100% found this document useful (10 votes)
10K views1,139 pages

Handbook of Lubrication and Tribology

Uploaded by

douglas
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
100% found this document useful (10 votes)
10K views1,139 pages

Handbook of Lubrication and Tribology

Uploaded by

douglas
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 1139

HANDBOOK of

LUBRICATION
and TRIBOLOGY
VOLUME II
Theory a n d D e s i g n
S E C O N D E D I T I O N

Edited by Robert W. Bruce

Sponsored by the SOCIETY OF TRIBOLOGISTS AND LUBRICATION ENGINEERS


HANDBOOK of
LUBRICATION
and TRIBOLOGY
VOLUME II
T h e o r y a nd Design
S E C O N D E D I T I O N
HANDBOOK of
LUBRICATION
and TRIBOLOGY
VOLUME II
Theory an d D e s i g n
S E C O N D E D I T I O N

Edited by Robert W. Bruce

Boca Raton London New York

CRC Press is an imprint of the


Taylor & Francis Group, an informa business
MATLAB® is a trademark of The MathWorks, Inc. and is used with permission. The MathWorks does not warrant the accuracy
of the text or exercises in this book. This book’s use or discussion of MATLAB® software or related products does not consti-
tute endorsement or sponsorship by The MathWorks of a particular pedagogical approach or particular use of the MATLAB®
software.

CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742

© 2012 by Taylor & Francis Group, LLC


CRC Press is an imprint of Taylor & Francis Group, an Informa business

No claim to original U.S. Government works


Version Date: 20120418

International Standard Book Number-13: 978-1-4200-6909-9 (eBook - PDF)

This book contains information obtained from authentic and highly regarded sources. Reasonable efforts have been made to
publish reliable data and information, but the author and publisher cannot assume responsibility for the validity of all materials
or the consequences of their use. The authors and publishers have attempted to trace the copyright holders of all material repro-
duced in this publication and apologize to copyright holders if permission to publish in this form has not been obtained. If any
copyright material has not been acknowledged please write and let us know so we may rectify in any future reprint.

Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmitted, or utilized in any
form by any electronic, mechanical, or other means, now known or hereafter invented, including photocopying, microfilming,
and recording, or in any information storage or retrieval system, without written permission from the publishers.

For permission to photocopy or use material electronically from this work, please access www.copyright.com (http://www.copy-
right.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA 01923, 978-750-8400.
CCC is a not-for-profit organization that provides licenses and registration for a variety of users. For organizations that have been
granted a photocopy license by the CCC, a separate system of payment has been arranged.

Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used only for identifica-
tion and explanation without intent to infringe.
Visit the Taylor & Francis Web site at
http://www.taylorandfrancis.com

and the CRC Press Web site at


http://www.crcpress.com
Contents

Preface....................................................................................................................... xi
Advisory Committee.............................................................................................. xiii
Review Board.. .......................................................................................................... xv
Contributors...........................................................................................................xvii

SECTION I  Theory and Practice of Lubrication and Tribology


1 Tribology......................................................................................................... 1-1
Kenneth C Ludema
2 Surface Analysis..............................................................................................2-1
Irwin L. Singer
3 Roughness.. ......................................................................................................3-1
Christopher A. Brown
4 Contact Mechanics.. ........................................................................................4-1
George G. Adams
5 Friction............................................................................................................5-1
Peter J. Blau
6 Wear: A Synoptic View...................................................................................6-1
Shyam Bahadur
7 Adhesive Wear.. ............................................................................................... 7-1
Kyriakos Komvopoulos
8 Abrasive Wear.................................................................................................8-1
Koji Kato
9 Rolling Contact Fatigue Wear........................................................................9-1
Michael N. Kotzalas
10 Fretting.......................................................................................................... 10-1
Thomas N. Farris and N. Sundaram

v
vi Contents

11 Erosion.. ......................................................................................................... 11-1


Awatef A. Hamed
12 Oxidative Wear.. ............................................................................................ 12-1
John R. Nicholls and Richard G. Wellman
13 Wear Models.. ................................................................................................ 13-1
Kenneth Holmberg and Anssi Laukkanen
14 Lubrication.................................................................................................... 14-1
Robert L. Jackson
15 Hydrostatic Lubrication................................................................................ 15-1
Gregory J. Kostrzewsky
16 Hydrodynamic Lubrication.......................................................................... 16-1
John A. Tichy
17 Compressible Gas Film Lubrication............................................................. 17-1
Luis San Andrés
18 Elastohydrodynamic Lubrication................................................................. 18-1
Farshid Sadeghi and Trevor Slack
19 Mixed Lubrication.. ....................................................................................... 19-1
Martin N. Webster
20 Boundary Lubrication and Boundary Lubricating Films............................20-1
Hugh Spikes
21 Additive Technology..................................................................................... 21-1
Elaine S. Yamaguchi, Gaurav Bhalla, and Vincent J. Gatto

SECTION II  Lubricants


22 Lubricants.. ....................................................................................................22-1
Robert W. Bruce
23 Incompressible Fluids...................................................................................23-1
Ronald A. Reich and James R. Anglin
24 Base Oils........................................................................................................24-1
Joseph M. Perez and Kimberly Wain Fick
25 Additives for Lubricants...............................................................................25-1
Leslie R. Rudnick
26 Rheology.. ......................................................................................................26-1
Scott S. Bair
27 Lubricant Application................................................................................... 27-1
Paul W. Hetherington and Evan S. Zabawski
28 Lubricating Grease........................................................................................28-1
Paul A. Bessette
Contents vii

29 Solid Lubricants............................................................................................29-1
Robert W. Bruce
30 Metalworking Lubricants.. ............................................................................30-1
Gregory J. Foltz
31 Hydraulic Fluids.. .......................................................................................... 31-1
James B. Hannon
32 Fluid Maintenance........................................................................................ 32-1
Allison M. Toms and George J.W. Staniewski

SECTION III  Wear Materials


33 Wear Materials.............................................................................................. 33-1
William A. Glaeser and Robert W. Bruce
34 Friction and Wear of Polymer Materials......................................................34-1
Thierry A. Blanchet
35 Metals............................................................................................................35-1
Thomas W. Scharf
36 Wear and Lubrication of Ceramics.. .............................................................36-1
Said Jahanmir
37 Composite Materials..................................................................................... 37-1
Li Chang and Klaus Friedrich
38 Coatings and Surface Treatments.................................................................38-1
Arup Gangopadhyay
39 Low Friction Coatings.. ................................................................................. 39-1
Thomas W. Scharf
40 Wear Coating and Treatments......................................................................40-1
Gary L. Doll and Allan Matthews
41 Coatings and Surface Treatments: Interactions with Lubricants................ 41-1
Staffan Jacobson

SECTION IV  Design for Lubrication and Tribology


42 Design for Lubrication and Tribology.. ........................................................42-1
Robert W. Bruce
43 Fluid Film (Hydrodynamic) Lubrication.. ....................................................43-1
Andras Z. Szeri
44 Journal Bearings.. ..........................................................................................44-1
John C. Nicholas
45 Thrust Bearings.. ...........................................................................................45-1
Scan M. DeCamillo and Bruce R. Fabijonas
viii Contents

46 Hydrodynamic Step and Wedge Bearings....................................................46-1


Theo G. Keith, Sorin Cioc, and L. Moraru
47 Compliant Foil Bearing Technology: An Overview.. ................................... 47-1
Hooshang Heshmat
48 Components with Nonconforming Contacts...............................................48-1
Andrew V. Olver
49 Lubrication of Rolling Element Bearings.....................................................49-1
E. Ioannides and Guillermo E. Morales-Espejel
50 Gear Lubrication...........................................................................................50-1
Robert F. Handschuh
51 Cams.............................................................................................................. 51-1
Andrew V. Olver
52 Lubrication Oil Systems................................................................................ 52-1
Jan Ploszaj, Hooshang Heshmat, and George J.W. Staniewski
53 Surface Texturing.. ........................................................................................ 53-1
Izhak Etsion
54 Sliding Bearings............................................................................................54-1
Timothy Alan Parsons and Jianpeng Feng
55 Magnetic Bearings.. .......................................................................................55-1
Alan B. Palazzolo, Zhiyang Wang, Jung Gu Lee, Albert F. Kascak,
and Andrew J. Provenza
56 Face Seals.. .....................................................................................................56-1
Tom W. Lai
57 Lip Seals.. ....................................................................................................... 57-1
Robert K. Flitney
58 Brake and Clutch.. .........................................................................................58-1
Roberto C. Dante, Carlo Navire, and Bruno Tron
59 Automotive Tribology................................................................................... 59-1
Edward P. Becker
60 Turbomachinery Tribology...........................................................................60-1
William D. Marscher
61 Natural and Artificial Human Joints........................................................... 61-1
Francis E. Kennedy and Douglas W. Van Citters
62 Nuclear Reactor Power Station Lubrication................................................. 62-1
Ken J. Brown and Steven Lemberger
63 Space Mechanism Lubrication......................................................................63-1
Stuart Loewenthal
64 Magnetic Storage.. .........................................................................................64-1
Nan Liu and David B. Bogy
Contents ix

65 Diagnostics....................................................................................................65-1
Richard S. Cowan
66 Tribology Testing..........................................................................................66-1
Terry L. Merriman
Preface

The term “tribology” has its origins in the Greek word “tribos,” which means rubbing.
The importance of tribology—friction, lubrication, and wear—has been expressed by the Jost
Report in 1966 as a potential savings quantified to 1.3%–1.6% of the gross national product (GNP).
Subsequently, similar studies in Canada by the National Research Council Canada (NRC; in 1986),
BMFT Germany (in 1976), and by ASME in the United States (in 1981) and in China (in 1986) esti-
mated the potential annual savings in those countries as 1%–2.5% of the GNP. The proper use of
generally accepted tribological practices may thus enable annual savings of over $140 billion in the
United States alone, or $470 per capita.
Since the publication of the first edition of Handbook of Lubrication and Tribology, Volume II: Theory
and Design in 1984, the increase in the price of energy and the environmental cost of energy have both
increased the significance of tribology. This second edition is meant to cover the field with few excep-
tions, leading to a highly condensed treatment of the most relevant information. It summarizes estab-
lished knowledge and practices and provides references for detailed study.
Section I, Theory and Practice of Lubrication and Tribology, reviews the basic principles of tribology
as currently used and describes wear mechanisms and modes of lubrication.
Section II, Lubricants, covers the full range of lubricants, or, as these are often called, coolants, includ-
ing mineral oil, synthetic and water-based fluids.
In Section III, Wear Materials, a wide range of wear- and friction-reducing materials and treatments
is reviewed. Today, this is the fastest growing area of tribology, with announcements of new coatings,
better performance, and new vendors being made every month.
Section IV, Design for Lubrication and Tribology, covers components and equipment commonly
found in tribological systems, followed by coverage of some specific industrial areas and their processes.
With contributions by a large number of authors and coauthors—some of the foremost experts in the
areas covered—this handbook provides the user with the most relevant information. The work of these
contributors and the support of their management are gratefully acknowledged.
I would like to acknowledge the Society of Tribologists and Lubrication Engineers for sponsoring the
development of this second edition. I would also like to express my gratitude to all the reviewers.
It is hoped that this handbook will enable the users to solve problems, make improvements, and claim
some of the savings projected by Jost, NRC, and the ASME.

Robert W. Bruce

xi
xii Preface

Bibliography
ASME, Strategy for energy conservation through tribology, NU, 1977 and 1981.
BMFT, German Federal Ministry for Research and Technology, Tribologie, Reibung-Verschleiss,
Schmierung. BMFT—Forschungsbericht T76-38, Bonn, Germany, 1976.
BMFT, German Federal Ministry for Research and Technology, Damit Rost und Verschleiss nicht
Milliarden fressen, Bonn, Germany, 1984.
Dudgeon, E.H., National Research Council of Canada, Associate Committee on Tribology: A strategy for
tribology in Canada, 1986, NRC Number: 26556.
Jost, J.P., Lubrication (tribology) education and research, Jost report, Department of Education and
Science, HMSO, London, U.K., 1966.
Tribology Institute, Chinese Mech. Eng. Soc., An investigation on the application of tribology in China,
Beijing, China, 1986.

MATLAB® is a registered trademark of The MathWorks, Inc. For product information, please contact:
The MathWorks, Inc.
3 Apple Hill Drive
Natick, MA, 01760-2098 USA
Tel: 508-647-7000
Fax: 508-647-7001
E-mail: [email protected]
Web: www.mathworks.com
Advisory Committee

E. Richard Booser Carlton Rowe Ward Winer


General Electric Mobil School of Mechanical
Vero Beach, Florida University Park, Florida Engineering
Georgia Institute of Technology
William A. Glaeser
Ed Salek Atlanta, Georgia
Battelle
Society of Tribologists and
Columbus, Ohio
Lubrication Engineers
Kenneth C Ludema Park Ridge, Illinois
Mechanical Engineering
Department
University of Michigan
Ann Arbor, Michigan

xiii
Review Board

Gary C. Barber Lois J. Gschwender Timothy C. Ovaert


Automotive Tribology Center University of Dayton Research University of Notre Dame
Oakland University Institute Notre Dame, Indiana
Rochester, Michigan Dayton, Ohio
Stephen H. Roby
Doug P. Hunsicker Chevron Energy Technology
Robert W. Bruce
Caterpillar Company
GE Aviation
Peoria, Illinois Richmond, California
Cincinnati, Ohio
Tom Karis Richard F. Salant
Jerry P. Byers Hitachi Global Storage Georgia Institute of Technology
CIMCOOL Industrial Technologies Atlanta, Georgia
Products, LLC San Jose, California
Cincinnati, Ohio
Gregory W. Sawyer
R. Gordon Kirk
University of Florida
Virginia Polytechnic Institute
Marc Carpino Gainesville, Florida
and State University
Consultant
Blacksburg, Virginia
State College, Pennsylvania Carl E. Snyder
Alan O. Lebeck University of Dayton Research
Thomas F. Conry Mechanical Seal Technology, Institute
University of Illinois Inc. Dayton, Ohio
Urbana, Illinois Albuquerque, New Mexico
Evan S. Zabawski
Piet M. Lugt Fluid Life
Ben M. DeKoven Edmonton, Alberta, Canada
SKF Engineering and Research
X-Lubes, Inc.
Centre
San Jose, California
Nieuwegein, the Netherlands Erwin V. Zaretsky
Glenn Research Center
Itzhak Green Karen M. Marvich National Aeronautics and Space
Georgia Institute of Technology GE Aviation Administration
Atlanta, Georgia Cincinnati, Ohio Cleveland, Ohio

xv
Contributors

George G. Adams Peter J. Blau Richard S. Cowan


Department of Mechanical and Oak Ridge National Laboratory Manufacturing Research
Industrial Engineering Oak Ridge, Tennessee Center
Northeastern University Georgia Institute of
Boston, Massachusetts Technology
David B. Bogy
Atlanta, Georgia
James R. Anglin Department of Mechanical
Alcoa Inc. Engineering
Roberto C. Dante
Pittsburgh, Pennsylvania University of California,
Universidad Nacional
Berkeley
Autónoma de México
Shyam Bahadur Berkeley, California
Mexico City, Mexico
Department of Mechanical
Engineering
Christopher A. Brown Scan M. DeCamillo
Iowa State University
Worcester Polytechnic Institute Kingsbury, Inc.
Ames, Iowa
Worcester, Massachusetts Philadelphia, Pennsylvania
Scott S. Bair
Georgia Institute of Technology Gary L. Doll
Ken J. Brown
Atlanta, Georgia University of Akron
Eco Fluid Center Ltd.
Akron, Ohio
Toronto, Ontario, Canada
Edward P. Becker
General Motors Company Izhak Etsion
Pontiac, Michigan Robert W. Bruce Department of Mechanical
GE Aviation Engineering
Paul A. Bessette Cincinnati, Ohio Technion—Israel Institute of
TriboScience & Engineering, Technology
Inc. Haifa, Israel
Dartmouth, Massachusetts Li Chang
Aerospace, Mechanical and
Bruce R. Fabijonas
Gaurav Bhalla Mechatronic Engineering
Kingsbury, Inc.
Chevron Oronite Company University of Sydney
Philadelphia, Pennsylvania
LLC Sydney, New South Wales,
Richmond, California Australia
Thomas N. Farris
Thierry A. Blanchet Rutgers University
Department of Mechanical, Sorin Cioc Piscataway, New Jersey
Aerospace, & Nuclear Department of Mechanical
Engineering Engineering Jianpeng Feng
Rensselaer Polytechnic Institute University of Toledo Oiles America Corporation
Troy, New York Toledo, Ohio Concord, North Carolina

xvii
xviii Contributors

Kimberly Wain Fick Hooshang Heshmat Francis E. Kennedy


Department of Chemical Mohawk Innovative Dartmouth College
Engineering Technology, Inc. Thayer School of Engineering
Pennsylvania State University Albany, New York Hanover, New Hampshire
University Park, Pennsylvania
Paul W. Hetherington Kyriakos Komvopoulos
Robert K. Flitney Fluid Life Department of Mechanical
Sealing Technology Edmonton, Alberta, Canada Engineering
Consultant University of California,
Rowton Villa, Craven Arms, Berkeley
Kenneth Holmberg
United Kingdom Berkeley, California
VTT Technical Research Centre
of Finland
Gregory J. Foltz
Helsinki, Finland Gregory J. Kostrzewsky
CIMCOOL Industrial Products,
Cummins Inc.
LLC
E. Ioannides Columbus, Indiana
Cincinnati, Ohio
S Ioannides Tribology and
Klaus Friedrich Engineering Consultants Ltd Michael N. Kotzalas
Institute for Composite Materials London, United Kingdom The Timken Company
IVW GmbH Canton, Ohio
Technical University Robert L. Jackson
Kaiserslautern Auburn University Tom W. Lai
Kaiserslautern, Germany Auburn, Alabama John Crane Inc.
Morton Grove, Illinois
Arup Gangopadhyay Staffan Jacobson
Ford Motor Company Uppsala University Anssi Laukkanen
Dearborn, Michigan Uppsala, Sweden VTT Technical Research Centre
of Finland
Vincent J. Gatto Helsinki, Finland
Said Jahanmir
Albemarle Corporation
Mohawk Innovative
Baton Rouge, Louisiana
Technology, Inc. Jung Gu Lee
Albany, New York Texas A&M University
William A. Glaeser
College Station, Texas
Battelle
Columbus, Ohio Albert F. Kascak
US Army Research Laboratory Steven Lemberger
Awatef A. Hamed Glenn Research Center Lemberger Consulting Services
School of Aerospace Systems Cleveland, Ohio LLC
University of Cincinnati Chicago, Illinois
Cincinnati, Ohio Koji Kato
Department of Mechanical Nan Liu
Robert F. Handschuh Engineering Department of Mechanical
Glenn Research Center Nihon University Engineering,
National Aeronautics and Space Koriyama, Japan University of California,
Administration Berkeley
Cleveland, Ohio Theo G. Keith Berkeley, California
Department of Mechanical
James B. Hannon Engineering Stuart Loewenthal
ExxonMobil Corporation University of Toledo S. Loewenthal & Associates
Allentown, New Jersey Toledo, Ohio San Jose, California
Contributors xix

Kenneth C Ludema Timothy Alan Parsons Trevor Slack


Mechanical Engineering Oiles America Corporation School of Mechanical
Department Concord, North Carolina Engineering
University of Michigan Purdue University
Ann Arbor, Michigan Joseph M. Perez West Lafayette, Indiana
Department of Chemical
William D. Marscher Engineering
Mechanical Solutions, Inc. Pennsylvania State University Hugh Spikes
Whippany, New Jersey University Park, Pennsylvania Imperial College London
London, United Kingdom
Allan Matthews Jan Ploszaj
University of Sheffield Howard Marten Company Ltd. George J.W. Staniewski
Sheffield, United Kingdom Pickering, Ontario, Canada Ontario Power Generation
Terry L. Merriman Pickering, Ontario, Canada
Andrew J. Provenza
Battelle Glenn Research Center
Columbus, Ohio National Aeronautics and Space N. Sundaram
Administration Purdue University
Guillermo E. Morales-Espejel
Cleveland, Ohio West Lafayette, Indiana
SKF Engineering and Research
Centre
Ronald A. Reich
Nieuwegein, the Netherlands Andras Z. Szeri
Alcoa Inc.
Department of Mechanical
L. Moraru Pittsburgh, Pennsylvania
Engineering
Department of Aerospace University of Delaware
Sciences Leslie R. Rudnick
Newark, Delaware
Politehnica University of Ultrachem, Inc.
Bucharest New Castle, Delaware
Bucharest, Romania John A. Tichy
Farshid Sadeghi Rensselaer Polytechnic
Carlo Navire Emeritus Professor of Institute
Isibond S.a.s Mechanical Engineering Troy, New York
Moncalieri, Italy School of Mechanical
Engineering
John C. Nicholas Purdue University Allison M. Toms
Lufkin—Rotating Machinery West Lafayette, Indiana GasTOPS Inc.
Technology Pensacola, Florida
Wellsville, New York Luis San Andrés
Texas A&M University Bruno Tron
John R. Nicholls
College Station, Texas NVH Advisor
Cranfield University
Barge, Italy
Bedfordshire, United Kingdom
Thomas W. Scharf
Andrew V. Olver Department of Materials
Douglas W. Van Citters
Imperial College London Science and Engineering
Thayer School of Engineering
London, United Kingdom University of North Texas
Dartmouth College
Denton, Texas
Hanover, New Hampshire
Alan B. Palazzolo
Department of Mechanical Irwin L. Singer
Engineering US Naval Research Laboratory Zhiyang Wang
Texas A&M University Chemistry Division Texas A&M University
College Station, Texas Washington, DC College Station, Texas
xx Contributors

Martin N. Webster Elaine S. Yamaguchi Evan S. Zabawski


ExxonMobil Research and Chevron Oronite Company Fluid Life
Engineering LLC Edmonton, Alberta, Canada
Clinton, New Jersey Richmond, California

Richard G. Wellman
Cranfield University
Bedfordshire, United Kingdom
I
Theory and
Practice of
Lubrication
and Tribology

1 Tribology  Kenneth C Ludema.........................................................................................1-1


Introduction  •  Friction  •  Lubrication  •  Wear  •  Bibliography
2 Surface Analysis  Irwin L. Singer.....................................................................................2-1
Introduction  •  Lubrication and Surface Wetting  •  Methods of Surface
Analysis  •  Characterization of Surfaces and Subsurfaces  •  Mechanical
Testing  •  Summary  •  Acknowledgments  •  References
3 Roughness  Christopher A. Brown...................................................................................3-1
Introduction  •  Measurement  •  Analysis and Characterization  •  Concluding
Remarks  •  Acknowledgments  •  References
4 Contact Mechanics  George G. Adams........................................................................... 4-1
Introduction  •  Basic Principles  •  Elastic Contact of Nonconformal Bodies  •  Spherical
Bodies  •  Cylindrical Bodies  •  Smooth Nonconforming Bodies  •  Contact of Conformal
Bodies  •  Effect of Friction  •  Elastic–Plastic and Fully Plastic Contacts  •  Scale-
Dependent Plasticity  •  Effect of Adhesion  •  Elastic–Plastic Contact with
Adhesion  •  Contact of Rough Surfaces  •  References
5 Friction  Peter J. Blau.........................................................................................................5-1
Introduction  •  Historical Studies of Friction  •  Static and Kinetic Friction
Coefficients  •  Frictional Transients and Instabilities  •  Friction and Energy
Dissipation  •  Frictional Heating  •  Friction and Interfacial Shear Strength  •  Localized
Material Displacement and Its Effect on Friction Modeling  •  Material-Specific
Friction  •  Friction Testing  •  Summary  •  References
6 Wear: A Synoptic View  Shyam Bahadur....................................................................... 6-1
Introduction  •  Adhesive Wear  •  Abrasive Wear  •  Fretting Wear  •  Erosive
Wear  •  Electrical Sliding Contact Wear  •  Arc Erosion Wear  •  Complexity of Wear
Situations  •  Investigating Tribological Failures   •  References

I-1
I-2 Theory and Practice of Lubrication and Tribology

7 Adhesive Wear  Kyriakos Komvopoulos..........................................................................7-1


Introduction  •  Phenomenological Aspects of Adhesive Wear  •  Factors Affecting
Adhesive Wear  •  Simple Theory of Adhesive Wear  •  Wear Coefficient  •  Alternative
Forms of the Adhesive Wear Equation  •  Formation of Loose Wear Particles  •  Evolution
of Surface Roughness and Minimum Clearance in Sliding Systems  •  Closing
Remarks  •  References

8 Abrasive Wear  Koji Kato................................................................................................. 8-1


Specific Wear Rate ws and Wear Coefficient K  •  Abrasive Wear Mode and Degree
of Penetration Dp  •  Friction Coefficient and Abrasive Wear Mode as Functions
of Dp  •  Abrasive Wear Mode Transition in Repeated Sliding  •  Abrasive Wear
Mode and Degree of Wear β  •  Hardness Increase on Groove Surface by Repeated
Rubbing  •  Concluding Remarks   •  References

9 Rolling Contact Fatigue Wear  Michael N. Kotzalas.....................................................9-1


Origin of Rolling Contact Fatigue  •  Distribution of Rolling Contact Fatigue  •  Predicting
RCF  •  Material Considerations  •  Other Considerations  •  Conclusion  •  References

10 Fretting  Thomas N. Farris and N. Sundaram............................................................ 10-1


Introduction  •  Mechanics of Fretting  •  Fretting Experiments  •  Design for
Fretting  •  Fretting in Practice  •  Summary  •  References

11 Erosion  Awatef A. Hamed..............................................................................................11-1


Introduction  •  Experimental Facilities Used in Erosion Research  •  Erosion Test
Results  •  Effect of Particle Impact Velocity  •  Effect of Particle Impact Angle  •  Effect
of Temperature  •  Effect of Particle Size and Loading  •  Effect of Particle Hardness,
Sharpness, and Mineral Composition  •  Volcanic Ash  •  Roughness of Eroded
Surfaces  •  References

12 Oxidative Wear  John R. Nicholls and Richard G. Wellman...................................... 12-1


Introduction and Background  •  Oxidation in the Absence of Wear  •  Transition to
Mild Oxidative Wear  •  Development of Wear Protective Oxide Layers  •  Oxidation
Effects in Fretting Wear  •  Modeling Erosion–Oxidation Processes  •  Basic
Erosion Mechanisms  •  Erosion–Oxidation  •  Modeling and Mapping
Erosion–Oxidation  •  Summary  •  References

13 Wear Models  Kenneth Holmberg and Anssi Laukkanen............................................ 13-1


Nomenclature  •  Introduction  •  Wear Equations  •  Wear Maps  •  Computerized Wear
Models  •  Simulation of Surface Stresses and Deformations  •  Parametric Analysis of
a Wear Contact  •  Component Wear Models  •  Conclusions  •  Acknowledgments  • 
References

14 Lubrication  Robert L. Jackson...................................................................................... 14-1


Introduction  •  Brief History of Lubrication  •  Types and Regimes of
Lubrication  •  Squeeze Film Lubrication  •  Rough Surface Considerations  •  Surface
Texturing  •  Nanoparticle Laden Lubricants  •  Chapter Overview  •  References

15 Hydrostatic Lubrication  Gregory J. Kostrzewsky....................................................... 15-1


Nomenclature  •  Introduction  •  History  •  Principles of Operation  •  Fundamental
Relationships  •  Flow Control  •  Design Procedure: Multi-Recess Journal
Bearing  •  Hybrid Bearings  •  Optimization  •  Summary  •  References

16 Hydrodynamic Lubrication  John A. Tichy.................................................................. 16-1


Introduction  •  Governing Equations of Fluid Mechanics for Lubrication
Flow  •  Fundamental Simplifications of Lubrication Theory  •  Reynolds Equation
of Hydrodynamic Lubrication: The One-Dimensional Bearing Case  •  Two-
Dimensional Versions of Reynolds Equation  •  Various Modifications of Reynolds
Equation  •  References
Theory and Practice of Lubrication and Tribology I-3

17 Compressible Gas Film Lubrication  Luis San Andrés................................................17-1


Nomenclature  •  Subscripts  •  Acronyms  •  Introduction  •  Fundamentals of
Gas Film Lubrication Analysis  •  Performance of One-Dimensional Slider Gas
Bearings  •  Plain Cylindrical Gas Journal Bearings  •  Introduction to Flexure Pivot
Bearings  •  Introduction to Foil Bearings  •  Recommendations for Oil-Free Rotating
Machinery  •  References
18 Elastohydrodynamic Lubrication  Farshid Sadeghi and Trevor Slack..................... 18-1
Nomenclature  •  Introduction  •  Regimes of Lubrication  •  EHL Models  •  EHL
Minimum Film Thickness Equations  •  Thermal and Non-Newtonian
Effects  •  Application of EHL Theory to Machine Components  •  References
19 Mixed Lubrication  Martin N. Webster........................................................................ 19-1
Introduction  •  Running-In (Break-In)  •  Experimental Methods and Tests  •  Wear
and Durability under Mixed Lubrication Conditions  •  Modeling Mixed Lubrication
Problems  •  Mixed Lubrication: The Future  •  References
20 Boundary Lubrication and Boundary Lubricating Films  Hugh Spikes.................. 20-1
Origins of Boundary Lubrication  •  Stribeck Curves and Boundary Lubrication  •  Solid-
Like and Viscous-Like Boundary Films  •  Detection and Measurement of Boundary
Lubrication  •  Film Replenishment  •  Types of Boundary Film  •  Friction Control by
Boundary Films  •  Wear Control by Boundary Films  •  Correlation between Friction
and Wear Properties of Boundary Films  •  Strength of Boundary Films  •  Prediction of
Boundary Lubrication Performance  •  References
21 Additive Technology  Elaine S. Yamaguchi, Gaurav Bhalla,
and Vincent J. Gatto...................................................................................... 21-1
Mechanism of Oxidation and Antioxidants  •  Mechanism of Contamination and Deposit
Control  •  Mechanism of Sludge and Varnish Control  •  Mechanism of Wear Control
by Antiwear and EP Additives  •  Mechanism of Corrosion Control  •  Mechanism of
Friction Modifiers  •  Mechanism of Viscosity Index Improver  •  Mechanism of Foam
Inhibitors  •  Conclusions and Future Needs  •  References
1
Tribology
1.1 Introduction....................................................................................... 1-1
Atomic Bonding, the Common Root of the Tribology Triad
1.2 Friction................................................................................................ 1-2
Surface and Contact Mechanics  •  Area of Contact  •  Adhesion as
the Cause of Friction  •  Measurement of Friction Force  •  Static
versus Sliding Friction  •  Break-In Friction Force  •  Frictional
Vibrations, Stick Slip, Stiction, Etc.
1.3 Lubrication.......................................................................................... 1-4
Kenneth C Ludema 1.4 Wear..................................................................................................... 1-5
University of Michigan Bibliography................................................................................................... 1-6

1.1  Introduction
Tribology is the science of substances in the state of tribulation or, more specifically, rubbing. It includes
the subfields of friction, wear, and lubrication. It involves all materials, organic and inorganic, all
machinery with moving parts, and even the human body.
Tribology is studied in several very different academic disciplines. The research aspects of friction
had been the domain of physicists in early years and remains so today. Friction problems are now mostly
the domain of product engineers. Lubrication became the major topic in tribology when the age of
machinery began. One new emphasis in tribology is the development of wear-resistant coatings, which
is a topic mostly in materials engineering. Another is in the bioengineering field, dealing with artificial
bone joints, artificial hearts, and many more issues.
This chapter is made up of three sections, one each on friction, wear, and lubrication. In these sec-
tions, atomic bonding or adhesion is taken as the fundamental cause of friction and wear, though not in
the simplest interpretation of adhesion. One problem with this concept is that adhesion is usually taken
as a force normal to a surface whereas friction is a force that is parallel to a surface. Further, the adhesion
theory of friction leaves the question of how atomic bonds can be so readily released as to allow sliding
with only reasonable friction force and wear.
The answer lies in the fact that the adhesion theory of friction is usually discussed in the context
of contact between two bare crystalline materials. In reality, whenever reasonable friction and wear
occurs, there is a bridge of van der Waals bonds, lamellar substances, and extremely disordered crystal-
line materials between the solid, load-carrying bodies.

1.1.1  Atomic Bonding, the Common Root of the Tribology Triad


Many of the observations in tribology are manifestations of the bonding forces between atoms. We
use the hard-ball model for atoms, even though we know that the localization of atoms is not that
simple. Further, we use the Lennard-Jones potential model of the bonding forces between atoms, which
assumes that widely separated atoms are drawn toward each other and nearby atoms repel each other.

1-1
1-2 Theory and Practice of Lubrication and Tribology

The ­“ balance point” varies with the material: for plain carbon austenite, the average spacing is a little
larger than the minimum atomic spacing, which is about 0.254 nm. The size of the atom is defined by the
location of electrons around the nucleus, which is ever indefinite.
The summation of the maximum forces of the attraction of many atoms is the force to fracture a bar
of the material in tension.
When two pieces of metal are placed in a vacuum chamber where the pressure is ≈10−12 atm, cleaned of
“contaminants,” and moved into contact with each other, they will become bonded or welded together.
No heat, pressure, or sliding is required. Until the parts are joined, their individual surfaces existed in
a high state of (negative) energy. If a very precise and instrumented specimen manipulator was used to
bring the bodies toward each other, the manipulator would sense a negative force and then a snapping
together of the bodies beginning when they had come within just a few nanometers of each other. The
two bodies become one if the atomic arrays match. The bonded bodies cannot be readily separated from
each other and neither can one be slid over the other.
Ceramic materials, with their covalent or ionic bonds, behave the same way but most ceramics are
more prone to fracture than are most metals when strains are imposed. Polymers, with their combina-
tion of covalent and van der Waals bonds, are much more flexible than are metallic bonds.
It seems that contact and bonding as achieved in a high vacuum might be the basic state of solids. But
in our common experience, we consider it normal that most pairs of bodies will slide over each other
readily. The answer may be seen when air is admitted into the vacuum chamber. Within 10−8 s, the oxy-
gen and water vapor in the air cover the specimens, forming two or three layers of oxide (for many met-
als) and a film of adsorbed water. These layers keep two contacting metal bodies separated far enough
from each other to prevent the very high bonding forces to operate. The two bodies will slide over each
other with an ease or difficulty as measured in a friction test.
In this model, friction is seen as a limited adhesion. Sliding resistance is determined mostly by the
physical properties of the substances covering the two contacting bodies. Many substances besides
oxides do form on clean surfaces and many different gases condense on the surfaces. These condensates
usually have an ordered structure (somewhat like a solid) with a viscosity that is orders of 10 higher than
that of bulk liquid.
Since we cannot determine how much or what kind of adhesion-interrupting substances are formed
on the surfaces around us, it is simply not possible to predict the coefficient of friction for any sliding pair.
Neither can we expect that friction force would be the same from 1 day to the next or at different sliding
speeds. And since we cannot expect the interrupting substances to be uniform in any way over a surface
and certainly not when smeared by sliding, we cannot expect that friction forces will be smooth or steady.
But we might expect that some regions on sliding surfaces could have extremely thin films upon
them, allowing sufficient shear stress transfer during sliding to remove small bits of material as wear
particles. (This explains only a few mechanisms of wear.)
The definitions of friction and wear seem clear enough, but the term “lubrication” is less so.
Conventional lubrication amounts to inserting a fluid film or special solid substances between two sur-
faces to keep them separated beyond the most intense force fields emanating from passing solid surfaces.
Adsorbed gas and a layer of the products of chemical conversion (e.g., oxide) do the same but are not
called lubricants even though they produce the effect of lubrication. Perhaps, the difference in definition
is whether a lubricant was applied by the hand of man rather than by “nature.”

1.2  Friction
1.2.1  Surface and Contact Mechanics
The resistance to the sliding and/or rolling of one solid body along another is usually called friction. We
simply say that the bodies are in contact, but contact is not easy to define.
Tribology 1-3

Contact as envisioned in solid mechanics occurs when the planar boundaries of two semi-infinite
solids merge into one plane. Two bodies are in contact with each other over an apparent area of contact.
Practical surfaces are not perfectly smooth nor even atomically smooth. Rather they are uneven or
rough, having undulations of several scales. The larger scale is called waviness and the smaller are called
asperities. Often asperities are depicted as steep mountains but they are actually more like gentle hills.
These can all be characterized by the several available surface roughness–measuring devices. Two solids
in contact rest on each other’s asperities.

1.2.2  Area of Contact


Asperities are usually deformed plastically when pressed against a counter surface (which also contains
asperities). As load increases more asperities carry a load, but the stress in each is not much increased.
It is useful for some purposes to determine the real area of contact between two bodies, but very dif-
ficult to measure. Since the asperities are in the plastic state of strain, the stress in each must be near the
yield strength, Y, of the material. So Areal ≈ W/Y, where W is the applied load.

1.2.3  Adhesion as the Cause of Friction


Scientists had speculated for centuries on the cause(s) of sliding friction. The earliest notion was that
the asperities on the opposing surfaces could not readily pass by each other but rather “interlocked”
with each other. Later scientists noted that when metal specimens slid with high applied load, metal was
found to have “adhered” and transferred from one to the other. The adhesion theory of friction was sup-
ported strongly by the finding that a film of fatty acid, of thickness far less than the height of asperities,
reduced the friction force very considerably whereas the friction force was not strongly influenced by
surface roughness of metal.

1.2.4  Measurement of Friction Force


Ever since friction forces were first measured, scientists have expressed frictional behavior in terms of
the friction coefficient, μ = F/W, where F is the force to slide or roll and W is the load (or force) between
bodies. Almost every book covering friction, mechanics, and machine design has tables of the coeffi-
cient of friction of many substances. These tables are not often helpful in that large differences may be
found from one table to another for the same material. One example is mild steel. Handbooks usually
show single numbers and possibly small ranges of value, in the region of 0.2, or perhaps 0.4. The author
of these different values likely did measure them but the values are put into tables without explanation.
This author has measured μ for mild steel on mild steel in the range from 0.1 to 1.1, covering a fairly wide
range of several variables. Often the counter surface against the listed material is not mentioned. Neither
are the conditions of sliding, for example, sliding speed, applied load, specimen shape, and several more
variables. This is a major dilemma in engineering. In the absence of good data, the obvious solution is
to measure friction force oneself. Because measuring friction force is bothersome to do, many problems
are solved by using a handbook value and then assuming that friction does not vary whatever the condi-
tions of sliding. This is essentially the approach attributed to Coulomb.

1.2.5  Static versus Sliding Friction


For many materials, the friction force to begin sliding is often greater than the friction force to sustain
sliding. The starting friction force is defined as the static friction whereas that associated with sustained
sliding is called the dynamic, kinetic, or sliding friction.
1-4 Theory and Practice of Lubrication and Tribology

No value of friction force is particularly predictable but the static friction force is the least predictable.
It often varies with time of standing still before sliding begins or resumes. This may be due to squeezing
out of some viscous-like substances from between load-carrying asperities.

1.2.6  Break-In Friction Force


Single-pass friction force is often different from repeat-pass sliding. At least three changes may occur
during sliding to cause that. For metals, there is usually some plastic flow and work hardening extending
down into the substrate. Sliding now occurs on a harder material. In metals and some ceramics, there
are new chemical species formed, such as oxides, which over time can become thicker or could flake off.
This changes friction. And in some polymers, the surface molecules become oriented in the direction of
sliding in single-pass sliding but become randomly arranged in reciprocating sliding.

1.2.7  Frictional Vibrations, Stick Slip, Stiction, Etc.


If you set a block of wood on a table and pull it with a rubber band, the rubber band stretches as the
prime mover (hand) moves. At some point, the block begins to move and it accelerates to the point of
moving faster than prime mover. The force in the rubber band becomes less than what is required to
sustain sliding so the block stops—it “sticks.” But the prime mover continues to stretch the rubber band
to the point where the force is enough to get the block moving again—it “slips.”
But under some conditions of block mass and stiffness of rubber band, the block never sticks. This
cannot be called stick slip but rather “frictional vibrations.”
The earlier discussion shows how frictional vibration can be caused by a static friction force that
exceeds sliding friction force. Frictional vibration can also be induced by three other phenomena,
namely, by a friction force that diminishes as speed increases, by ratcheting of atomic bonds, and by the
natural variations in friction force with sliding distance.
The more important technological cause of frictional vibration is the natural variations in friction
force during sliding. This author knows of no instance where unlubricated friction force is found to be
absolutely smooth. Prominent examples of frictional vibrations are in brakes of cars, trucks, airplanes,
and bicycles. The sliding members produce irregular friction force, which can be taken as an oscillation
of force upon a steady state of friction force. These oscillations incite the vibration of nearby mechanical
components, yielding a noisy system. These noises can be reduced by placing damping at various loca-
tion of the overall system.
A third but less understood cause for frictional vibration may be the ratcheting of atomic bonds, not
very practical for most people yet. Where a single atom moves from one well to another on a flat array of
atoms, the force to begin the transition is maximum at start. The force decreases to zero in mid-travel,
then becomes negative when settling into the second well—approximately. No energy is lost in this
cycle. Should sliding friction force then be zero or very small? With zero normal load on the atom, static
coefficient of friction could be infinite, and with a lifting force on the atom, the coefficient of friction
would be negative.

1.3  Lubrication
Little squirts of oil now and then are enough to keep household devices functioning. But when powered
ships and railroad locomotives were developed, their bearings needed more reliable and sustained lubri-
cation. Shafts in bearings are the ideal geometry for lubrication since the rotating motion of the shaft
draws lubricant into what would be the contact area if the bearings were dry. Several liquid lubricants
were put to use and close attention was devoted to designing the geometry of bearings and sliders of
the high-speed machinery. This was aided by the development of many equations for sizing bearing
Tribology 1-5

parts and for selecting lubricants. The analytical methods used were those of fluid mechanics, and when
applied to fluid film lubrication becomes hydrodynamics. The intent of hydrodynamic lubrication was
to develop and maintain a thick fluid film within the bearings so that the bearing components hardly
ever contacted each other, except at start-up and stopping. In the limit, the fluid film need only be
thicker than the combined heights of the tallest asperities on the two opposing surfaces. Though many
equations are available for calculating the fluid film thickness for different applications, it is still not
possible to predict the friction or drag forces in hydrodynamic systems.
The chief property of the lubricant in early days was its viscosity and the amount that the viscosity
decreased as its temperature increased. Later, it was found beneficial for the viscosity of a lubricant to
increase with applied pressure. But though two oils might be identical in all obvious aspects, one was
sometime found to lubricate steel better than the other. The better performing lubricant was found in
the 1930s to have been less completely refined. Not all of the sulfur compounds had been removed. The
chemical nature of lubricant was thus first recognized. Over the next decades, many rather compli-
cated oil formulations were developed. Most formulations were developed for automobiles since they
use about half of all the oil consumed. Several following chapters describe oil formulations and their
function in more detail.
Very specific chemically active additives in oil react with specific metals and specific ceramic bearing
material to form new compounds, often referred to as a boundary films. This compound has a lower
shear strength than do the bearing parts. There are several opinions on the composition of the film. The
compound gathers on the forward slopes of the larger undulations (waviness features) of surfaces. It is
actually not a film, but islands. This mode of lubrication is referred to as boundary lubrication.
When a fluid film becomes too thin to carry a design load, perhaps due to overload or high tem-
perature of the bearing system, or due to misalignment or other geometric problems, a form of dam-
age known as scuffing (some call it scoring) takes place. A very thin fluid film shearing at a high rate
can transmit a very high shear stress, at least up to the shear limiting properties of the fluid. The high
shear stress region on the shaft or bearing surface causes plastic shearing by thermal-plastic instability,
leaving minor surface roughness. Often these regions “heal” and smooth over again when overload is
relieved. But if overload continues, the surfaces may become too rough for continued function and in
the extreme the parts may seize together. Usually, there is no measurable loss of material in scuffing.
A humbler example of scuff healing is seen in an inadequately lubricated door hinge in a house.
Sometime such a hinge begins to turn with increasing difficulty. But if before seizure occurs, a drop or
two of light oil is touched to the hinge joints, it can be worked back to smooth and easy operation again.
There is one area of tribological hope that has yet to be realized and that is in MEMS​(microelectro-
mechanical systems). These are micro-nanoscale machines that have great potential for medicine and
other applications. They will not survive without lubrication but they cannot yet overcome the surface
(wetting) forces that liquid lubricants (e.g., condensed vapor) bring.

1.4  Wear
Very often, hardness reduces wear rate, but not always. Of the various forms of wear
1. Erosion by hard particles impinging at low angles is reduced by higher hardness but not much by
impingement at high angles. Because of the repeated impact of hard particles on a solid at high
angle, it is the high cycle fatigue properties that resist loss by erosion.
2. Abrasion is almost always reduced by higher hardness.
3. Wear by fatigue mechanisms, as in rolling contact, is reduced by higher hardness but up to a
point, beyond which wear rate increases again. Thus, for example, cam shafts are hardened to ≈50
Rc rather than to the maximum ±64 Rc. At lower hardness, abrasive particles wear the cam shaft;
at high hardness, the shaft lobes are vulnerable to fatigue failure.
1-6 Theory and Practice of Lubrication and Tribology

4. Fretting, the oscillatory sliding of very low amplitude, is reduced by higher hardness, usually. But
it is very complicated.
5. Wear by corrosion and oxidation is usually increased by higher hardness because chemical reac-
tivity is usually greater for the harder state of each metal.
Exceptions abound, and whereas tables of values of the coefficient of friction are not very reliable, tables
of values of wear rate are even less so.
Many events are ponderous to record. For example, if you measure the wear rate in simple reciprocat-
ing sliding of the polyethylene used in artificial hip joints and then switch to lateral reciprocation, the
wear rate increases several fold. If you stop a wear test on steel and later resume testing, the wear rate will
be significantly higher than before because of oxidation during the rest period. If a sand slurry contains
an electrolytic solution, the wear rate of a metal specimen will be greater than the sum of wear by the
sand and the rate of corrosion. The reason is that the abrasive action removes the passive layer on the
metal allowing faster corrosion.
There is hardly a way to tabulate data to cover all material and all circumstances of wear. There are
thousands of them. Each year hundreds of new coatings are developed. Several new polymers and copo-
lymers each year and many tens of new polymer composites are introduced. New alloys of metals con-
tinue to be offered, for example, the multicomponent alloys for artificial hip joints. Ceramic materials
are also under constant development with two goals: one is to make ever harder ceramics and the other
to make more ductile ceramics.
Though there are many new materials developed each year, there are few new mechanisms of wear.
Small particles can be removed from a larger body by only a few mechanisms: brittle failure (single
cycle) or ductile failure, low cycle fatigue and high cycle fatigue, chemical dissolution, ablation, ion
bombardment, and several more.
It should be possible to determine the mechanism by which material is lost from a surface and then to
determine what properties of material or operating variables should prevent or inhibit that mechanism
of wear. Over 100 equations may be seen in the literature, which were attempts to do the aforemen-
tioned. It is a formidable task. Most authors had scaled their task down by selecting the most influential
variables, but unfortunately too few authors select the same variables, thereby losing the possible benefit
of synergy. Unfortunately also, developing models are validated against data from tests, which often
ends up writing equations for lab devices rather than for real parts.
The full story of tribology is in the following chapters.

Bibliography
Armstrong-Helouvry, B. Control of Machines with Friction, Boston, MA: Kluwer Academic Publishers,
1991. ISBN 0-7923-9133-0.
Blau, P.J. Friction Science and Technology, 2nd edn., Boca Raton, FL: CRC Press, 2009. ISBN 978-1-4200-5404-0.
Budinski, K.G. and M.K. Budinski. Engineering Materials, Properties and Selection, 9th edn., Upper Saddle
River, NJ: Prentice-Hall, 2009. ISBN 0-13-904715-8.
Holmberg, K. and A. Matthews. Coatings Tribology, 2nd edn., Amsterdam, the Netherlands: Elsevier, 2009.
ISBN 978-0-444-52750-9.
Stachowiak, G.W. and A.W. Batchelor. Engineering Tribology, 3rd edn., Amsterdam, the Netherlands:
Elsevier, 2005. ISBN 978-0-7506-7836-0.
Vizintin, J., M.J. Kalin, K. Dohda, and S. Jahanmir, Eds. Tribology of Mechanical Systems, New York: ASME
Press, 2004. ISBN 0-7918-0209-4.
2
Surface Analysis
2.1 Introduction....................................................................................... 2-1
2.2 Lubrication and Surface Wetting.................................................... 2-2
2.3 Methods of Surface Analysis........................................................... 2-3
2.4 Characterization of Surfaces and Subsurfaces.............................. 2-5
Microscopy: Imaging to the Nanoscale  •  Structure: Texture and
Atomic Arrangements  •  Composition: Mostly Vacuum-Based
Techniques  •  Composition: Optical Techniques
2.5 Mechanical Testing......................................................................... 2-15
2.6 Summary........................................................................................... 2-16
Irwin L. Singer
U.S. Naval Research
Acknowledgments....................................................................................... 2-17
Laboratory References..................................................................................................... 2-17

2.1  Introduction
Bearings are expected to provide relative motion between the components, for example, balls and race,
with little or no service for the life of the machine. The components must be strong enough to carry the
load without permanent deformation and smooth enough to allow a lubricant to cover the asperities in
the contact zone. Ideally, the lubricant would always separate the two surfaces, to prevent any contact
between the mating surfaces. But that is not possible when the relative speed is zero, that is, at start-up
and slow down, so bearings are often designed to run in the mixed asperity contact/lubrication condi-
tion. Ideally, the surfaces would be durable enough to withstand contact. But even when the utmost care
is taken, from initial processing of raw materials to final finish, there is no guarantee that the compo-
nent is free of damage-inducing surface or subsurface defects. After the bearing is assembled, tested,
and put into service, what type of damage might occur? Damage can range from mild scoring to gross
spallation, and the causes can range from difficult to avoid contamination like dust (e.g., silica particles)
to mechanical instability of the assembly that houses the bearing. Why? Because real bearing materials
can deform, transform, form oxides on their surfaces, and accumulate superficial films—for good or for
bad—as illustrated by the schematic (a) and real1 (b) cross sections of worn-bearing surfaces in Figure
2.1. Bearing engineers need to be aware that even the best prepared components can be damaged by the
various mechanisms detailed in the following chapters. Here we focus on how one uses surface charac-
terization methods to discover and identify surface and subsurface damage.
This chapter describes the methods available to engineers for analyzing surfaces. What surface con-
ditions are conducive for the lubricant to lubricate? What techniques can be used to determine the
topography, structure chemistry, and mechanical properties of surfaces? The answers are in the ways
that modern scientists are able to “see” surfaces. “Seeing is believing” can be said to underlie all experi-
mental science. Until the late nineteenth century, what scientists could see was usually dictated by the
optics in a light microscope. With Maxwell’s equations, scientists recognized that visible light was only
a small region of the entire spectrum and began to “see” objects with light outside the visible spectrum.

2-1
2-2 Theory and Practice of Lubrication and Tribology

0.1 – 20 nm Tribofilm

0.3 – 5 nm Oxide

0.1 – 100 µm Deformation layer

Bulk material

20 µm

(a) (b)

FIGURE 2.1  (a) Schematic cross section of a worn-bearing surface (nonlinear depth scale for emphasis).
(b) Optical micrograph of a worn 316L stainless steel surface. (From Rainforth, W.M. et al., Philos. Mag., 66, 621,
1992. With permission.)

The discovery of electron beams and ion beams, combined with an understanding of beam physics,
allowed scientists to explore electron and ion beam interactions with surfaces. Today, there are literally
hundreds of “beam-in, beam-out” techniques, in which a beam is aimed at an area of the surface and
the interaction generates secondary particles or rays that can then be analyzed along with the reflected
beam. Scientists use these to see not only the topography of surfaces—down to the atomic scale—but
also the microstructures and compositions of surfaces from the nanoscale to the mesoscale, from the
outermost atomic layer to depths over 1–100 μm below the surface. The methods of surface analysis
are described in Section 2.3 and examples of surface analytical techniques are given in Sections 2.4.1
through 2.4.4. Finally, Section 2.5 covers mechanical testing of surfaces. However, before entering the
world of surface characterization, it will be valuable to understand how liquids interact with surfaces.

2.2  Lubrication and Surface Wetting


Liquids must be able to wet the surface to be effective lubricants. Wetting is determined by the interaction
between a liquid, a solid, and the surrounding gas. Consider a droplet on a flat surface (Figure 2.2). There
are three interfaces: the solid–liquid interface, the liquid–vapor interface, and the solid–vapor interface.
Each of these interfaces has an associated surface tension, γ, which represents the energy required to
create a unit area of that particular interface, or equivalently, the force per unit length at the interface.
By force arguments, the angle q between a liquid drop and a solid surface, the contact angle, is given by
or
vap

Receding
id–

Vapor
liqu

θ
Liquid
γ

γ solid–vapor γ solid–liquid Solid θ


θ Advancing

Solid

(a) (b)

FIGURE 2.2  Contact angle θ of a droplet on a flat surface indicating (a) three interfacial forces, γ, and (b) contact
angle hysteresis.
Surface Analysis 2-3

gsolid − vapor − gsolid − liquid


cos q = (2.1)
gliquid − vapor

where the “gammas” are surface tensions of the three interfaces. According to Equation 2.1, when
γ liquid–vapor < γsolid–vapor, cos θ increases and the area of contact between the liquid and the solid interface
increases. As the drop spreads and wets the surface, the contact angle approaches zero, hence cos θ
approaches 1. Using a contact angle goniometer, Zisman et al. discovered a universal behavior in plots
of cos θ vs. the surface tension of (nonpolar) liquid: cos θ increases linearly as the surface energy
decreases.2 They defined the point at which the plot crossed the cos θ = 1 axis as the critical surface ten-
sion, an important parameter because it is a characteristic of only the solid and therefore can be used to
predict the wettability of the surface. Early studies of wetting were done with model liquids on polished
and smooth substrates. Nonideal surfaces can be rough and have a chemically inhomogeneous surface.
Contact angle measurements on such surfaces often exhibit hysteresis. The hysteresis is characterized by
dynamic contact angle measurements, obtained by subtracting advancing (θa) and receding (θr) contact
angles. The physics behind wetting, surface energy, and contact angle can be found elsewhere.3,4
How do liquid lubricants maintain contact with the surface? As stated earlier the lower the sur-
face energy, the more easily liquids will wet (spread) on the surface. To increase wetting and reduce
­“ beading,” lubricants are doped with surfactants (surface active agents), compounds that lower the sur-
face energy of a liquid. Since metals generally have a high surface energy, due to their strong interatomic
bonds, one would expect liquids to wet them easily. However, in ambient environments, metal atoms
are not present as metals at the top surface (except for noble metals). Instead, the surface consists of
many layers, the topmost being a contaminant film consisting of hydrocarbons and condensed water
molecules, then an oxide layer, each of which lowers the surface energy.5 To achieve robust, wettable
surfaces, proper cleaning procedures must be followed after grinding, abrading, and polishing. Often, a
final solvent cleaning is sufficient. Occasionally, stronger cleaning processes are necessary, like hot alka-
line detergent cleaning, electrocleaning, and acid etching; in other cases, plasma (ion bombardment)
cleaning is required to remove the last layer of impurities. Similar cleaning procedures are necessary
before coating a surface. Sometimes, however, the surface energy of the metal is much higher than the
surface tension of the lubricant, causing the lubricant to spread too thin to provide lubrication. Here, a
stable, durable film of low surface energy—a barrier film—can be deposited on both sides of the contact
zone to prevent out-migration of fluids.6
Liquid lubricants, in turn, carry additives that form protective tribofilms on bearings. The films are
most important when sliding or rolling speeds are too low for hydrodynamic lubrication. These films
are usually too thin to be seen by the naked eye and, typically are neither smooth nor homogeneous;
nonetheless, they can protect the surfaces from wear. They have been and still are prime candidates for
investigation using modern surface analytical techniques.7

2.3  Methods of Surface Analysis


All “beam-in, beam-out” techniques operate on the same principle: send a beam of photons, electrons,
or ions at a surface and collect photons, electrons, or ions from the surface. These techniques developed
with advances in beam generators combined with basic understanding of beam–surface interactions.8
Perhaps, the earliest and best example of this collaboration is x-ray (Bragg) diffraction. In 1913, William
Lawrence Bragg and his father, William Henry Bragg, discovered that x-rays reflected off crystalline
solids produced patterns representative of a lattice structure. For this discovery, they were awarded the
Nobel Prize in physics in 1915, the only father–son team to win. Less than 50 years later, the technique
unraveled the mysteries of the double helix shape of DNA.9
The many “beam-in, beam-out” techniques described in this chapter can be understood with the help
of the cartoon in Figure 2.3. A beam of photon or electron particles hits a solid layer. The beam can be
2-4 Theory and Practice of Lubrication and Tribology

Auger electron Backscattered


Electrons Secondary electron electron

Photons
Photon

Collector

Elastically scattered
electron
Transmitted photon Transmitted electron

FIGURE 2.3  Cartoon of the “beam-in, beam-out” method. Electron or photon beams interact with a thin solid
layer. They can reflect or pass through the layer, or be inelastically scattered, producing secondary particle beams.
The collector determines what particles can be detected and the incident and exiting beam angles control the depth
sensitivity.

reflected (specularly) from the surface, can be transmitted, or more likely, will enter the layer but then
scatter from atoms below the surface. The scattering can be elastic, changing the direction but not the
energy of the beam, or it can be inelastic, depositing some energy into the atoms. The energy deposited
usually produces secondary particles, which can also be scattered elastically or inelastically, dictated by
the physics of interactions of the secondary particles with atoms. The “beam out” can be the scattered
primary beam, or a beam of secondary particles, depending on how the collector, shown on the right, is
tuned. Figure 2.3 illustrates several ways in which “beam-in, beam-out” techniques are used to analyze
surfaces. For example, an electron beam (in—from the left) can either reflect specularly from the sur-
face or pass into the layer. The electrons that scatter specularly from the surface can be identified by a
detector that collects and counts electrons that have the same energy as the primary beam. The electrons
that enter the layer will be scattered inelastically by atoms and decay within a tear drop–shaped volume
beneath the surface.10 Inelastic scattering will excite atoms, and their subsequent decay will produce
either x-rays or Auger electrons. The latter can be identified by electron detectors tuned to Auger elec-
tron energies and x-ray detectors tuned to capture the fluorescent x-rays. The depth sensitivity of the
technique depends on both the inelastic mean free path (the distance that the primary and secondary
beams travel before losing their energy) and the incident and exiting beam angles. Similarly, an incom-
ing photon beam can interact with solids and produce electrons, known as “photoelectrons.”
During the past 100 years or so, microscopies and spectroscopies evolved using the “beam-in, beam-
out” approach, with beams of electrons, ions, and photons from microwaves to x-rays and even gamma
rays. Today, there are hundreds of materials analysis techniques;11 many of them are surface-specific
and most are available commercially. Figure 2.4 displays a chart with the most widely used, commer-
cially available surface analytical techniques;12 all are referred to by acronyms13 that will be spelled out
and described in Section 2.4. The vertical axes give the composition sensitivities and the horizontal
axis gives the diameter size of the spot that can be detected. Each technique is represented by an area
(bubble) on the chart. The boundary of the bubble gives the range of compositions and lateral spot size
detectible. The techniques can determine compositions from a few atom percentage (AES, XPS, EDS) to
parts per trillion (dynamic SIMS), structures depicting arrangements of atoms with their nearest neigh-
bors (STEM/EELS) to defects in centimeter-thick steel plates (RTX), and surface sensitivities that span
Surface Analysis 2-5

SPM Imaging
TEM/STEM techniques
FIB
SEM
RTX
5E22 100 at%
STEM/
XRR
EELS AES SEM/ Raman XPS/
1E22 STEM/ XRD 10 at%
EDS EDS ESCA
1E21 1 at%
FTIR
XRF RBS
1E20 0.1 at%
Ph
Detection range

Lexes 100 ppm


ys
1E19
(Atoms/cm3)

Ph

ica
ys

l li
Ph

ica

TOF-SIMS
m
ys

1E18 10 ppm
l li

it
ica

fo

LA-ICPMS TXRF
l li

it

r0
m

fo

.3

1 ppm
it

1E17
nm mp
3n
fo
r3

sa
0n

m
sa

1E16 100 ppb


pl
m

in
sa

lin

g d th
m

ep
pl

1E15 10 ppb
ep
in

th
gd
ep

Dynamic SIMS
th

1E14 1 ppb

1E13 100 ppt

1E12 Copyright©2010 evans analytical group LLC 10 ppt


0.1 nm 1 nm 10 nm 100 nm 1 µm 10 µm 100 µm 1 mm 1 cm
Analytical spot size

FIGURE 2.4  Chart of surface analytical techniques. (Chart Courtesy of Evans Analytical Group®, http://www.
eaglabs.com/techniques/analytical_techniques/.)

depths from 1 nm (STEM) to 1 cm (XRR). So, if one needed to identify, for example, a thin boundary
film in a wear scar 100 μm wide, AES could do the job but XRF or RBS could not.

2.4  Characterization of Surfaces and Subsurfaces


There are many books on surface analysis14–17 and several that specialize in applications of surface analy-
sis to tribology.18–20 Herein is a survey of many techniques that are used by tribologists today and intro-
duce a few others that might be useful in the future.

2.4.1  Microscopy: Imaging to the Nanoscale


The unaided human eye is capable of resolving features down to tens of micrometers in size and seeing
colors over the visible region of the electromagnetic spectrum, roughly from 400 to 700 nm. An optical
microscope magnifies features as small as a micrometer in diameter. White light is used to illuminate
the feature and the scattered (reflected or transmitted) light is collected by lenses to form an image.
Alternatively, an image can be formed by scanning/rastering a narrow beam of light over the surface,
and the scattered light is collected and displayed as a function of the position of the incident beam on the
surface. In both cases, the spatial resolution of the image or spectrum is determined by the diffraction
limit of the microscope,21 which is proportional to the size of optical objective and inversely propor-
tional to the wavelength of the light. The light can also be analyzed spectroscopically, qualitatively with
a prism, or quantitatively with a spectrometer.
2-6 Theory and Practice of Lubrication and Tribology

2.0 mm 60.0 µm
(a) (b) (c)

FIGURE 2.5  (a) Damaged cone and rollers of taper roller bearing, due to heavy load and inadequate lubrication;
(b) severe plastic deformation (galling) of Nitronic 60™ oscillating against a flat of the same material at 485°C; and
(c) damage to AISI 4340 steel surface after impacting 100 times on a platen covered by a layer of coarse, but mobile,
alumina abrasives. (Photo (a) from Bearing failures and their causes, SKF Publication PI 401 E, Figure 55; photos
(b) and (c) are courtesy of P. Blau, ORNL, Oak Ridge, TN.)

Many of the advances in science over the last five centuries can be attributed to developments in
optical instruments operating with visible light. In the last 50 years or so, dozens of new “white light”
and ultraviolet–visible–infrared (UV/VIS/IR) light microscopies have been developed. Furthermore,
during the last 25 years, nanotechnologies have allowed us to break through the barrier of the “far-field”
diffraction limit, which has limited the resolution of light microscopes to a micrometer. Today, with
ever-increasing understanding of the physics of “near-field” techniques, numerous microscopes have
been developed that allow us to see nanoscale features with IR and VIS light.22
Digital cameras and VIS light microscopes are indispensible for examining worn surfaces at magni-
fications from 1× to 100×. Figure 2.5 gives three examples of visible light images of wear features. Photo
(a) shows surface fatigue of both the rollers and race in the tapered roller bearing,* photo (b) depicts
galling of a high temperature alloy steel, and photo (c) reveals both impressions and particle detachment
caused by impacting an abrasive-covered steel. Components can be inspected for defects and debris with
a low-magnification (1× to 50×) stereo microscope or a high-magnification (50× to 1000×) reflecting
microscope, the smallest size being about 1 μm, limited only by the diffraction of VIS light. Lighting
can be used in many ways. Off-axis (oblique) lighting can bring out vertical features on flat surfaces.
Transmitted light can illuminate buried structures and subsurface damage in transparent ceramics or
plastics and semitransparent materials, like fiber-reinforced epoxies. Polarized transmitted light of pho-
toelastic materials can depict stress fields in materials. Two examples of photos in transmitted light are
shown in Figure 2.6. The photo on the left displays both the bulk structure and worn surface of a G10
composited burned by molten Al; the photo24 on the right reveals the stress contours produced by a cyl-
inder pressed against a photoelastic flat. Surface topography can be enhanced using Nomarski objectives,
dark field, oblique, or back-lit illumination. Confocal microscopes create in-focus images by computer-
aided image reconstruction of 3D features that would be blurred in a standard microscope. Even simpler,
3D features can be captured as a stereoscopic image by photographing an object twice, at an angle about
4° apart; an example of a stereo image taken in a scanning electron microscope (SEM) will be given later.
Optical profilometers provide the most detailed images on reflective surfaces, capturing both
topography and height variations, by a variety of techniques: interferometry, confocal microscopy,
chromatic aberration, laser triangulation, and others. Depending on the technique used, these non-
contacting profilometers can measure and image 3D features over lateral scales from micrometers to
meters and vertical scales from below 1 nm to many centimeters. Contact (stylus) profilometers can
also produce images of 3D surface features, but are more invasive (scratch surfaces, move around
debris, etc.) and usually take longer to acquire images than optical profilometers. On the positive
side, their spatial resolution is limited mainly by the size (radius) of the stylus’ tip, and they are more

* Photo (a) is Figure 55 from Reference 23.


Surface Analysis 2-7

(a) (b)

FIGURE 2.6  (a) G10 epoxy, burned by molten Al, is illuminated by back lighting, showing fiber weave within and
on top of translucent epoxy. (b) Stresses produced by a cylinder pressed against a photoelastic flat, with both normal
and tangential loading, are made visible by polarized light. Main images are magnified areas from insets in upper
right. (Photo (a) from author; photo (b) is a public domain image from http://en.wikipedia.org/wiki/File:Kontakt_
Spannungsoptik.JPG.)

reliable than optical profilometers when the reflectivity is poor or the surfaces are transparent or have
semitransparent layers. Optical profilometers can produce significant height-quantification artifacts
on poorly reflective or transparent surface, both of which can be avoided by depositing a thin reflective
coating on the surface.
Scanning probe microscopy (SPM), in particular atomic force microscopy (AFM), can operate in a
contact mode like a stylus profilometer, imaging surfaces with nanometer resolution laterally and verti-
cally. The stylus probe, with a tip of radius 20 nm or less, is attached near the end of a cantilever. The tip,
loaded to around 1 nN, is guided across a surface by a piezoelectric positioning element. The location
of the tip, both lateral and vertical coordinates, is displayed either as a true 3D or pseudo-3D map, the
latter being a 2D image with the z-coordinate color coded. SPMs can also be used in more gentle modes,
by either cruising above the surface while sensing the attractive van der Waals forces between the tip
and the surface or tapping the surface. In both cases, the tip is set oscillating at or near the resonant
frequency of the cantilever. In the former, noncontact mode, the tip-to-surface distance is changed to
maintain a constant resonance frequency, providing a topographic image of the surface. In the tapping
mode, the initial vibration amplitude is increased to several hundred nanometers to allow the tip to
intermittently touch the surface. A servo motor is then used to adjust the tip height to maintain a con-
stant vibration amplitude, and this height profile becomes a force map of the surface.25
The SEM magnifies up to 100,000 times, showing features with dimensions at the nanometer scale.
A beam of high-energy electrons (from 0.5 to 30 keV) is rastered across the sample and either secondary
or backscattered electrons are collected. Image contrast can often be improved by operating at voltages
below 5 kV, but with a loss of lateral resolution. For samples that are nonconducting or have nonconduc-
tive features (oxide debris or insulating second phases), thin conductive films of carbon or gold reduce
“charging” artifacts. Figure 2.7 shows a stereo image of a Vickers indent in a ceramic.26 Stereo images of
3D features bring out the shape of indent impressions, surface bulges, and detached chips. An SEM must
be housed in a vacuum chamber to operate the electron gun. However, an environmental SEM (ESEM),
with differentially pumped sections, can operate at relatively high pressure (0.001 atm), allowing a range
of investigations, from in situ SEM analysis of atmospheric effects on friction on ceramics27 to analysis
of articular cartilage.28
The transmission electron microscope (TEM) can achieve even higher magnification, as small as
a single row of atoms. The lateral resolution is improved over SEM because the beam is transmitted
2-8 Theory and Practice of Lubrication and Tribology

6 µm 6 µm

FIGURE 2.7  SEM stereo image of a Vickers indent (at 4.9 N) in sintered SiC after heating in vacuum to 900°C.
Note that the ceramic deformed plastically inside the indent but chipped at the tensile-stressed edges. (From Singer,
I.L., Surf. Coat. Technol., 33, 487, 1987. With permission.)

10 nm

s
s

FIGURE 2.8  Cross-sectional HRTEM micrograph of a wear track on an amorphous Pb-Mo-S after 1000 cycles
(sliding direction parallel to image plane). Sliding created and lifted sheets of basal planes MoS2 (area marked by
downward arrows). Also, subsurface patches of basal MoS2 sheets are marked by upward arrows and the letter
“s.” (From Dunn, D.N. et al., MRS Proceedings Volume 522, Fundamentals of Nanoindentation and Nanotribology,
Moody, N.R. et al., Eds., MRS, Pittsburgh, PA, 1998, pp. 451–456; Wahl, K.J. et al., Wear, 230, 175, 1999. With
permission.)

through an ultrathin specimen, eliminating most of the secondary and backscattered electrons that
broaden the electron-illuminated region. When a finely focused beam is rastered across a thin section,
the technique is called scanning TEM (STEM) or high-resolution TEM (HRTEM). In the past, the prep-
aration of thin sections—as much an art as a skill—was done by mechanical grinding and chemical and/
or ion etching in several time-consuming steps. Figure 2.8 shows a HRTEM image of a cross-­sectional,
tribomechanically crystallized layer of an amorphous Pb-Mo-S coating.29 The image gave direct evi-
dence that single layers of ordered MoS2 (area marked by downward arrows) were created by sliding on
the surface. Also seen are subsurface patches of basal MoS2 sheets, marked by upward arrows and the
letter “s.” Today, thin sections can be made in one step by focused ion beam (FIB) etching in a SEM.
Areas to be thinned are identified in the microscope, then thin rectangular sections are etched with the
FIB and removed with a manipulator for TEM analysis.30,31
Microscopy can also be used to examine subsurface features of materials. One can access the near
subsurface top down, to depths from 1 to 500 nm, by chemical or electrochemical etching, by ion etch-
ing, or by mechanical or chemomechanical polishing (recognizing that each of the etching process men-
tioned earlier can leave artifact in the surface). Figure 2.9 shows both pseudo-3D (gray scale as height)
and 3D interferometric profilometry images of a polished 304 stainless steel surface that was ion-etch
100 nm deep. The etched surface reveals large austenitic grains; the height difference from grain to
grain is a result of the different sputtering rates of grains in different orientations.32 A more common
way to examine the subsurface is by cross-sectioning, that is, cutting perpendicular to the surface fol-
lowed by polishing, to expose the entire subsurface, for example, the cross-sectional image of a worn
surface on the right in Figure 2.1. A variant of cross-sectioning that magnifies the near subsurface is
Surface Analysis 2-9

Z Range: 188.8 nm

205
Z (µm) 40

0.4
0.2
0.0 0
0.2
50.0

103
µm
200.0
100.0 150.0 –40
50.0 100.0
0.0 50.0
Y (µ

0.0 )
–50.0 –50.0 m –80

m)

–100.0 –100.0 X 50 µm
–150.0
–150.0 –200.0 –120

0
0 138 276
(a) (b) µm

FIGURE 2.9  (a) Pseudo-3D (grayscale as height) and (b) 3D interferometric profilometry images of a polished
304 stainless steel surface that was ion-etch 100 nm deep. The etched surface revealed large austenitic grains; the
height difference from grain to grain is a result of the different sputtering rates of grains in different orientations.
Spikes on left image are noise artifacts. (Interferometric profiles shown here were taken on samples investigated in
Singer, I.L., J. Vac. Sci. Technol., 18, 175, 1981.)

called taper sectioning. This is performed by polishing the substrate at a shallow angle, θ, nearly parallel
to the surface (typically 1°–5°), which magnifies the cross-sectional view by sin−1 (θ).

2.4.2  Structure: Texture and Atomic Arrangements


X-ray diffraction (XRD) is the most commonly used technique to identify the structure of materials.
Specularly reflected x-rays produced intense peaks at angles that depend on the lattice spacing (Bragg’s
law), allowing identification of crystal structure, crystallite (grain) size, lattice strain, and preferred ori-
entations of grains. Although x-rays can penetrate tens of micrometers into solids, the technique can be
made very surface-sensitive by aiming the incoming x-ray beam at a small incident angle (0.3°–3°) to the
surface. Below a critical angle, the incident x-ray is confined to a thin layer near the surface, ­t ypically,
10–100 nm, and the reflected x-ray conveys the layer’s structure. The technique is known as grazing
incidence x-ray diffraction (GIXD or GID).
Electrons possess the same wavelike behavior as x-rays, and a variety of electron diffraction tech-
niques are used to study structures of single atomic layers at and below the surface. The TEM/STEM,
mentioned earlier, is able to image structures as well as produce electron diffraction patterns (TED) at
the nanometer scale. Besides identifying structures, these electron techniques are capable of determin-
ing the location and extent of plastic deformation down to the 10 nm scale. Low-energy electron diffrac-
tion (LEED), performed in ultrahigh vacuum chambers, impinges low-energy electrons (20–200 eV)
onto well-ordered crystalline surface. Spot patterns of backscattered diffracted electrons depict the
symmetry of the atomic surface and adsorbed species. Reflection high-energy electron diffraction
(RHEED) is a similar surface structural technique, but uses a more focused, higher-energy electron
beam (10–30 keV). The incident beam hits the surface at a shallow angle; the scattered electrons project a
diffraction pattern onto a screen. Electron backscatter diffraction (EBSD) operates like SEM, but is able
to produce “Kikuchi” diffraction patterns that identify texture or preferred orientation of crystalline or
polycrystalline material.
Finally, two very powerful, but specialized, techniques are electron energy loss spectroscopy (EELS)
and x-ray absorption near-edge structure (XANES), also called near-edge x-ray absorption fine structure
(NEXAFS). Both exploit the energy lost by inelastically scattered electrons to measure element-specific,
pair-distance distribution functions, that is, interatomic distances. By calculation-intense analysis, both
techniques are capable of determining the precise location of nearest neighbors, the ultimate atomic
2-10 Theory and Practice of Lubrication and Tribology

structure technique. High-resolution EELS (HREELS) is an even more precise EELS technique, but it
must be performed in an ultrahigh vacuum chamber.

2.4.3  Composition: Mostly Vacuum-Based Techniques


Before the advent of modern “dry” surface analytical techniques, metallurgists relied on wet chemical
methods to determine the microstructure of bearing material and detect impurities on the surface.33
These methods, while not quantitative nor low spot size, are easy to use and can be performed in the
field. Spot test kits, with reactive reagents, can detect specific elements found in stainless steels, tool
steels, high alloy steels, and many others. A drop of solution is placed on the sample, and the color spot
is used to identify the alloy. The tests are sensitive to a microgram.
Modern surface analytical techniques, as mentioned earlier, depend on beam-in, beam-out methods.
The three most commonly used surface analytical techniques are (1) energy-dispersive spectroscopy
(EDS), also known as energy-dispersive x-ray analysis (EDX), (2) Auger electron spectroscopy (AES),
and (3) x-ray photoelectron spectroscopy (XPS), also known as electron spectroscopy for chemical anal-
ysis (ESCA). When an atom is bombarded by photons (x-ray) or electrons with energies in the range of
2–50 keV, a core state electron can be ejected leaving behind a hole, an unstable electronic structure.
The hole can be filled in two ways. An electron from a higher orbit can fill the hole and the difference
in energy released as a photon (x-ray), the basis for x-ray fluorescence (XRF) analysis. Or, the core hole
can be filled by an outer shell electron and a second electron is ejected with kinetic energy equal to the
energy difference between the outer shell and the core; this is a three electron process, known as an
Auger process, named after one of its discoverers, Pierre Auger. In both cases, the energy of the emitted
x-ray or electron is characteristic of the atom and thus serves as a unique fingerprint of the material.
Which process is more likely to occur during electron excitation? Typically, Auger yields are higher at
low excitation energies (0.5–10 keV) and for lower atomic number elements, while XRF dominates at
high electron energies and higher atomic number elements.
Compositions of surface layers from 1 to 5 monolayers thick can be identified using electron and
x-ray sources in ultrahigh vacuum chambers. AES uses an electron beam in the 1–10 keV range to
excited atoms, which decay preferentially by Auger electron emission.14–17 Auger electrons have energies
from 10 eV to several thousands of electrovolts. At these energies, the electron mean free path is small,
0.3–3 nm. Therefore, Auger spectra represent the atomic compositions of atoms in the near-surface layer,
making AES a very surface-sensitive technique. All atoms, except for H and He, emit Auger electrons.
Furthermore, the kinetic energy of the Auger electron is sensitive to how the atom bonds with its neigh-
bors, providing chemical “fingerprinting” as well as elemental identification. Although XPS relies on
x-rays to excite atoms, it is very much like AES. X-ray-induced photoelectrons are emitted in the same
energy range as the electron-induced Auger electron, and therefore, XPS and AES have similar depth
sensitivities. The chemical information obtained from XPS is directly related to the binding energy of
the photoelectron and, therefore, the chemical state of the atom is more easily identified than that from
the three electron Auger process. The main advantage of AES is that its electron beam is much smaller
(hundreds of nanometers) than the x-ray beam (5–10 μm at the present time), so the spatial resolution
of scanning Auger microscopy (SAM) is much better than that of XPS. AES and XPS can be used in
conjunction with ion beam sputtering (ion beam erosion) to obtain compositions vs. depth information.
Figure 2.10 shows an example of Auger sputter depth profiling, with chemical fingerprinting of both
Ti and C as a function of depth below the surface of a Ti+-implanted bearing steel.34 Sputter depth pro-
filing, however, can be misinterpreted if the following caveats are not taken into account: first, prefer-
ential sputtering—a very common process—often makes it impossible to quantify compositions below
the surface.35 As an example, S sputters up to four times faster than Mo in a MoS2 matrix, giving S/Mo
ratios of 0.5 instead of the bulk ratio of 2.36 Second, sputtering breaks bonds, reducing the compounds
being interrogated at the surface, for example, sulfates (SO4)2− are reduced to sulfites (SO3)2− and even
to sulfides (S)2−.37,38
Surface Analysis 2-11

Depth (nm)
0 5 50 100 150 200

Auger peak-to-peak intensities (arb.unit)

O515 Ti
420
380
Fe650

C270
C
270

O
Ti
420

0 10 20 30 40
Ion milling time (min)

FIGURE 2.10  Auger sputter depth profile of Ti-implanted 52100 steel. Oxide layer (left) is shown with a more
expanded depth scale than the subsurface (right). The chemical states of both Ti and C can be inferred from the
line shapes of Ti(LMM) spectra (360–430 eV) and C(KLL) spectra (250–280 eV): Ti oxide on surface, Ti carbide at
oxide–steel interface, metallic Ti as C concentration falls, and an Fe carbide in the bulk of the steel. (From Singer,
I.L. et al., Nucl. Instrum. Methods, 182/183, 923, 1981. With permission.)

Energy-dispersive x-ray spectroscopy (EDS) is the analytical arm of the SEM. EDS can detect ele-
ments from B up (due to detector limitations), and unlike AES and XPS, EDS does not require ultrahigh
vacuum; SEM/EDS can be performed on large and not pristinely clean samples. EDS takes advantage of
fast, solid-state detectors to energy resolve the XRF emitted by excited atoms. Collection times are fast
(minutes) and energy resolution is sufficient to quantify elemental compositions to several percent. The
depth sensitivity of EDS in an SEM depends on the depth over which the electrons decay yet are suf-
ficiently energetic to excite atoms, typically 0.5–5 μm, depending on electron beam energy and atomic
mass of the near-surface atoms. While nowhere as depth sensitive as AES, it can be much faster and
simpler for qualitatively identifying local compositions than AES depth profiling. The two combined,
however, are an unbeatable tool for investigating the wear of coated surfaces: the SEM locates features;
EDS assesses the local compositions and thereby identifies candidate spots for AES depth profiling.
Figure 2.11 shows an example of how several surface analytical techniques were combined to study
the wear behavior of a MoS2 coating.39 In this study, EDS was used to located features in the wear track
according to composition, for example, oxygen-rich and oxygen-depleted areas. Then, AES depth pro-
files were taken to identify the layered compositions that gave rise to the features. The three depth pro-
files identify thick MoS2 debris at the edge of the track; a very thin, hence worn layer of MoS2 on the
track; and a nearly bare patch of track, with oxidized Fe and a very thin layer of MoS2.
When EDS is used with thin sections in the TEM/STEM, the depth sensitivity is controlled by the
section’s thickness, which can be as thin as 10 nm. Elemental maps generated by EDS complement
secondary and backscattered electron maps. While EDS is a fast way to identify the composition of a
surface, it cannot easily resolve overlapping emission energies, for example, the severe overlap of the
Mo(Lα) and S(Kα) spectra at 2300 eV. Wavelength-dispersive x-ray spectroscopy (WDS) uses diffraction
gratings to detect the precise wavelength of emitted x-rays and is therefore much better at quantifying
2-12 Theory and Practice of Lubrication and Tribology

Legend
60
S Point 2 Fe S
48
Mo
EDS 36 Mo
O
24
Line Fe
12 C O
C
0
60

Normalized intensity (%)


Point 1
48
1 36 S Mo
3 5 O
4 6 24 O
2 Mo
12 C C
Fe Fe
0
60
Point 3
48 Fe
S O
EDS intensity

O
36 Fe
24 Mo
12 S
Mo + S C
0
Position across track 0 3 6 9 12 15 18 21 24
Sputtering time (min)

FIGURE 2.11  (Top left) SEM image of a heavily worn track on MoS2-coated steel. (Bottom left) EDS line scans
depict the relative concentrations (thickness) of Fe(Kα), O(Kα), and Mo(Lα) + S(Kα) collected along line indicated
on SEM image, indicating compositions and thicknesses. (Right) Auger sputter depth profiles at three of the six
circles indicated in the SEM image. Circles indicate spots where Auger sputter depth profiles were taken. Point 1
was taken on a debris flake along the edge of the track; the debris is a mixture of Mo-S-Fe-O; point 2, an area of the
track with a MoS2 layer thinner than the original coating; point 3, mainly the oxidized wear track (Fe oxide), with a
thin cap of MoS2. (From Ehni, P.D. and Singer, I.L., Appl. Surf. Sci., 59, 45, 1992. With permission.)

composition and avoiding overlapping/interfering peaks. However, WDS is much slower than EDS.
Low-energy x-ray emission spectrometry (LEXES) is a variant of WDS. The electron probe microana-
lyzer (EPMA) is an SEM/EDS/WDS machine designed for more quantitative analysis. Electrons can
also stimulate light emission, known as cathodoluminescence (CL), in many nonmetallic and semicon-
ducting materials. The light, covering the UV/VIS/NIR range, can be analyzed spectroscopically and
mapped to complement SEM and EDX maps.
Finally, there are several fluorescence techniques for measuring wear that do not require vacuum
chambers for either beam in or beam out. The first is the aforementioned XRF. It relies on the same
detection scheme as EDS, but uses x-rays instead of electrons to generate the fluorescence. Because
x-rays penetrate 10 times deeper than electrons, the composition is obtained to depths from 10 to 50 μm.
And, since electron beams are not used, the technique does not require a vacuum chamber. In fact,
handheld XRF devices are available. Two other XRF techniques are ion-induced x-ray analysis (IIX)
and particle (alpha or proton)-induced x-ray spectroscopy (PIXE). As the names imply, they rely on
ions and protons, respectively, to generate XRF. Both are available commercially and used in real-time
monitoring of engine wear. PIXE can even be performed with a small radioactive source, which, while
not as efficient as more intense electron, x-ray, or particle beam sources, is very compact and requires
low power. A device built on this technique, the alpha proton x-ray spectrometer (APXS), was used by
the Sojourner rover to analyze rocks during the 1997 Pathfinder mission on Mars.
The second is thin-layer activation (TLA), a related technique that relies on gamma-ray fluorescence
and, therefore, also does not need a vacuum chamber.40 The surface to be studied is “labeled” with a
radioactive isotope marker, and then a gamma-ray detector monitors the loss of material from the com-
ponent or the increase in gamma-ray activity in the engine oil. TLA has been used to monitor wear (and
corrosion) of engines and other machinery. It can be performed under engine-operating conditions in
real time, is noncontacting, and can be made sensitive to submicrometer levels.
Surface Analysis 2-13

Secondary ion mass spectrometry (SIMS) uses an ion beam to sputter remove surface layers, then col-
lects and identifies the removed ion species with a mass spectrometer. While more difficult to quantify
than AES or XPS, SIMS can detect impurities at parts per trillion. SIMS can be used for composition
vs. depth profiling (dynamic SIMS) or simply to determine the composition of surfaces (static SIMS)
with minimal damage. Techniques that detect the backscattered or forward scattered primary ions are
called ion-scattering spectroscopies (ISS). Low-energy ion-scattering spectroscopy (LEIS) is sensitive
to the outermost atomic layer and can give both composition and structure. Medium-energy ion scat-
tering (MEIS) and high-energy ion scattering (HEIS), known in practice as Rutherford backscattering
spectroscopy (RBS), use higher-energy beams, from 100 keV to over 1 MeV, to probe compositions and
structure of surface and subsurface layers.
In glow discharge optical spectroscopy (GDOS) and glow discharge mass spectrometry (GDMS),
surfaces are bombarded by inert gas atoms in a glow discharge, producing sputtered atoms in either
an excited atomic state or an ionized state, respectively. In GDOS, the excited atoms decay by emitting
radiation, whose spectrum can be detected with an optical emission spectrometer (OES). Emission
peaks in the spectrum occur at wavelengths characteristic of the sputtered atoms; peak intensities
help determine atomic concentrations. In GDMS, the ionized atoms are collected in a mass spec-
trometer, and concentrations for most elements can be found down to the subparts per billion
range. Furthermore, in GDMS, unlike the related SIMS technique, quantification is nearly matrix
independent.
RBS, which relies on the energy loss of an ion passing through a material, is one of the few nonde-
structive techniques capable of composition profiling to depths of a micrometer or more. It is also highly
quantitative, thanks to decades of precision measurements of nuclear stopping powers and scattering
cross sections. RBS can also be used in the “channeling mode” to quantify defects in nearly perfect
crystals. Two related nuclear scattering techniques are particularly valuable for detecting hydrogen in
surfaces and films. Nuclear reaction analysis (NRA) relies on a nuclear reaction between the incoming
ion and a targeted nucleus, for example, the proton in hydrogen. The ion resonantly excites the nucleus,
which decays and emits ionizing radiation. A reaction used to profile hydrogen is

15
N + 1H →12 C + a + g ( 4.965 MeV )

with a resonance at 6.385 MeV. The concentration profiles of H vs. depth is obtained by detecting the
energy of the emitted gamma ray as a function of the 15N incident beam energy and computing the
depth from the known energy loss of 15N in the material. Elastic recoil detection (ERD), also referred to
as forward recoil scattering, uses a relatively low-energy (2 MeV) 4He beam to depth profile hydrogen,
but requires a foil sample from 1 to 10 μm thick.
Mössbauer spectroscopy is a spectroscopic technique based on the recoil-free, resonant absorption
and emission shifts of gamma rays emitted by atoms in solids. It is particularly useful for identifying
compositions of iron-containing specimens and, as with PIXE, data collection of Mössbauer spectra has
also been carried out on Mars.41 Gamma rays have very long path lengths in solids, which excludes their
use for surface-sensitive experiments. However, the same resonant shifts are detectable in emitted x-rays
and electrons, whose mean free paths in solids are tens of micrometers and 1–10 nm, respectively. While
not applicable for studying most elements, the latter two techniques, conversion x-ray and conversion
electron Mössbauer spectroscopy (CXMS and CEMS), have been exploited to study tribological coatings
and wear.42

2.4.4  Composition: Optical Techniques


Raman and IR spectroscopies are used to investigate mainly nonmetallic materials, including
organic and inorganic coatings and surface films, plastic and ceramic bearings, and lubricants. Both
2-14 Theory and Practice of Lubrication and Tribology

techniques can identify compounds based on vibrational and rotational modes of molecules. In IR
spectroscopy, the light directly excites molecules at frequencies characteristic of their structure. It
can be done by sweeping the frequency/wavelength of a monochromatic beam vs. time, or by using a
broadband light source and an interferometer to measure all wavelengths at once, then performing
a Fourier transform on the IR spectrum (FTIR). FTIR has been a standard practice in industry and
the military for condition monitoring of lubricants.43 In Raman spectroscopy, the electric field of the
laser (monochromatic) light is inelastically scattered by molecular, vibronic or electronic excitations,
resulting in small shifts, positive and negative, in the wavelength of the exiting light. The measured
shift between the incoming and outgoing photon energy, typically reported in units of wavelength or
frequency, corresponds to the excited mode.
Laser Raman spectroscopy is an excellent tool for analyzing wear scars and small features. It is com-
monly combined with an optical microscope, so locating features is easy. Its spatial resolution is about
1 μm in diameter by several micrometers in depth, as determined by the collection optics. It has also
been combined with a tribometer that allows in situ, real-time optical and Raman analysis of slid-
ing contacts.44 Raman is useful for detecting many inorganic compounds45 used as solid lubricants,
like MoS2 and diamond-like carbon (DLC). Moreover, it can be very sensitive, due to resonant Raman
effects,46 and detect films as thin as 10 nm.47 Sensitivity to selected compounds can also be enhanced
by selecting appropriate laser wavelengths. Raman can be made many orders of magnitude more sensi-
tive by enhancing the electric field in the neighborhood of the Raman interaction, as occurs when the
light excites surface plasmons. Surface-enhanced Raman spectroscopy (SERS) has been achieved by
­adding colloidal nanoparticles of Ag or Au to surfaces. Another SERS approach is to take Raman spec-
tra in the vicinity of a metallic (usually silver- or gold-coated) AFM or STM tip. The latter approach is
known as tip-enhanced Raman spectroscopy (TERS) and has been shown to have sensitivity down to
the single molecule level. TERS is one of many near-field scanning optical microscopy (NSOM/SNOM)
techniques that breaks the far-field resolution (diffraction) limit to investigate nanostructures with
UV/VIS/IR light.
An IR microscope allows one to locate areas on a sample, then to acquire IR spectra. Infrared micro-
scopes operate in either transmission or reflection mode. The spatial resolution is typically diffraction
limited to 20 μm or larger, determined by the wavelength of the blackbody radiation and the objective
apertures. Spectra of samples must be normalized to that of a “control” sample to eliminate the influ-
ence of the instrument and gases. In the lab, IR spectra of lubricants can be obtained by spreading the
lubricant on an IR-transmitting salt crystal and acquiring spectra in the transmission mode. Attenuated
total reflectance (ATR) IR is an alternative technique for studying thin lubricant films. ATR is based on a
special behavior of light in matter known as total internal reflection: When a ray of light enters a trans-
parent medium at or below a critical angle with respect to its surface, the light cannot pass through the
surface and, instead, reflects and becomes totally internally reflected. A film is deposited onto the ATR
crystal. A beam of infrared light is passed through a bevel edge of the crystal, then undergoes internal
reflection, producing an evanescent wave that penetrates typically 0.5–2 μm into the film. The technique
allows for in situ studies of tribofilm formation during boundary lubrication, as shown recently by Piras
et al.48 for zinc dialkyldithiophosphate (ZnDTP) against Fe and Sasaki et al.49 for tricresyl phosphate
against Fe.
Several instruments have been developed around the AFM and IR light sources that allowed IR spec-
troscopy to break through the diffraction limit (tens of microns) and measure IR absorption at the
micro- and nanoscale.50,51 In photothermal microspectroscopy (PTMS), IR light is beamed onto a sur-
face. If the radiation is absorbed by the sample, heat is generated locally and the surface expands and
is detected by the tip of an AFM. Topographical AFM maps are acquired at discrete IR frequencies or
by sweeping the IR frequencies, giving IR absorption maps with the spatial resolution of the AFM. A
related technique is scanning thermal microscopy (SThM), in which the AFM tip is replaced by a sub-
miniature temperature sensor like a thermocouple or a resistance thermometer.
Surface Analysis 2-15

2.5  Mechanical Testing


Understanding the mechanical properties of surfaces is crucial to bearing design, from run-in to steady-
state performance. For a given bearing contact geometry (see Chapters 4 and 13), the loading conditions
are conservatively estimated from bulk elastoplastic properties of the counterfaces. But these properties
do not usually account for changes in surface mechanical properties that arise during rolling or slid-
ing contact, for example, changes in surface hardness or in fracture toughness. Nor are bulk properties
sufficient to account for behaviors of coated or surface-engineered bearings. For these, it is necessary to
characterize the deformation and fracture resistance of thin layers.
The most common techniques for characterizing mechanical properties of surfaces are indentation
and scratch testing. In both techniques, a hard tip is pushed into and along a surface, respectively, while
the deformation response is recorded by sensors; later, the scratch tracks are examined with a micro-
scope. Indentation testing is often performed quasi-statically, that is, a load is applied to the surface,
held for a period of seconds, then removed; this method will be described in detail later. Alternatively,
dynamic indentation testing is performed by impacting the tip quickly, either once or repetitively, and
monitoring the rebound response. This technique is referred to as rebound “hardness” testing, but
that is a misnomer since it mainly measures elastic response whereas hardness usually refers to plastic
response. One such device is a scleroscope, in which a diamond-tipped or hardened steel tip is allowed to
fall from a known height, and the higher the rebound, the higher the elastic “hardness” of the material.
A more modern, electronic rebound tester, the Leeb rebound tester, measures the rebound velocities and
determines the rebound “hardness” from the energy lost during impact.
In scratch testing, a sharp tip is loaded against the surface as the sample moves transverse to the nor-
mal load. The earliest scratch tester was a sclerometer, an instrument used by mineralogists to measure
the width of a scratch made by a diamond under fixed load (not to be confused with a scleroscope).
Modern scratch testers are instrumented to record normal and tangential forces, and often have an
acoustic sensor to detect brittle fracture events.52 After testing, the scratched tracks are examined in a
microscope, then the wear patterns are associated with different mechanisms of failure.53 Scratch test-
ing is most valuable for investigating the limits of decohesion of the coating–substrate interface, what
is often referred to imprecisely as the “adhesion” of the coating. The combined normal and tangential
force of the tip applies increasing stresses to the interface, causing cracking, crack growth, delamina-
tion, and eventually spallation. Scratch tests can be used for quality control of coated bearings by, for
example, assigning a cutoff critical load for delamination. In addition, scratch testing has provided a
deeper understanding of coating–substrate mechanics.54 By correlating the behavior of the force curves,
acoustic emissions, and scratch patterns, numerous failure modes have been postulated, investigated,
and modeled. Scratch testing has also been used to characterize ceramics, polymers, ceramic-coated
polymers, and boundary-lubricant films. For thin (<1 μm) coatings, scratch tests are performed by scan-
ning force microscopy (SFM), which uses AFM-like devices as nanoscratch testers. They can be oper-
ated at high loads, simultaneously scratching the surface and measuring the scratch track depth during
deformation, then retracted and rerun at much lower loads to profile the depth of the permanently
deformed track.
Quasistatic indentation testing is categorized in terms of the load range applied or depth penetrated.
Macroindentation tests are usually performed with commercial machines designed for selected hard-
ness ranges of metals and alloys, for example, the Rockwell scales HRA-HRG; the procedures are well
documented in ASTM or ISO55 standards. The tests are performed with either a sharp diamond tip
(Rockwell, Vickers, and Knoop) or a hard spherical tip (Rockwell, Brinell); loads are in the kilogram-
force range and penetration depths around 0.1–1 mm. Most tests have been designed for engineering
metals; the Shore durometer test is used on polymers and elastomers.
Microindentation tests are carried out at lower loads, typically 0.01–10 N (1–1000 gf) and require a
sharp-tipped diamond indenter (Vickers, Knoop, or Berkovich) that penetrates only 10–100 μm into
2-16 Theory and Practice of Lubrication and Tribology

the surface. The geometry of the tip determines the depth of penetration at a given load. Vickers and
Berkovich have symmetric pyramidal geometries with the same projected area, but the Berkovich is a
three-sided pyramid and the Vickers, a four-sided pyramid. Knoop is an asymmetric (length to width
ratio = 7:1) four-sided pyramid, flatter than the Vickers, so it penetrates less than the Vickers at the
same projected area. Traditionally, microindentation tests were used to find the hardness of the surface,
where the hardness is computed as the ratio of the maximum load to the projected impression area.
A skilled operator measured the projected area of the impression with a light microscope. Modern
microindentation (and nanoindentation) machines are instrumented to collect force vs. time vs. depth
(FTD) data, a method known as depth-sensing indentation (DSI). The DSI approach is used to compute
parameters associated with the material’s deformation response—like hardness, elastic modulus, and
viscoelasticity—from FTD data.56 But it, like the traditional impression imaging approach, cannot take
into account the “pileup” of material along the rim of the impression. Pileup, however, can be measured
using 3D profilometry to compute the actual volume of material displaced by the indenter, above and
below the original surface. Finally, the indentation technique has many other uses besides hardness
testing.56 It can be used to evaluate the fracture toughness of ceramics and hard coatings, known as
indentation fracture toughness; stereo micrographs, like that in Figure 2.7, were used to assess the
toughness of sintered SiC subjected to several surface engineering treatments. It can be performed
with spherical indenters on soft coatings or with sharp diamond tips, like cube-corner indenters, on
hard coatings. It is also used to investigate fracture toughness of coatings and the durability of the
­coating–substrate interface.57 DSI can also be used to obtain the viscoelastic properties of polymers and
elastomers.58
Nanoindentation testing is also performed by DSI, but at loads from nN to tens of mN and is thus
able to obtain mechanical properties of nanometer-thin films and structures as well as image surface
mechanical properties like stiffness and hardness. 59 Such testers are built with very sensitive dis-
placement (nm) and load (μN) sensors, and require isolation from vibrations, thermal fluctuations,
and airflow. In addition to acquiring FTD data, they are capable of outputting 3D and pseudo-3D
(2D images with the third dimension color coded) spatial maps of topography, as well as normal
and lateral (friction) forces. The most sensitive ones are based on SPMs or AFMs, which can sense
individual surface atoms, measure the “spring constants” of molecules, and image areas at a magni-
fication 1000× greater than light microscopes. SPMs can be used to obtain mechanical properties of
tribofilms, transfer films, and liquid layers. Researchers have also exploited the nanometer sensitiv-
ity of SPMs as a near-field detector to avoid the diffraction limit on photon optics and extended the
property domain of SPMs to include magnetic, chemical, and thermal properties of surfaces, 60 as
mentioned earlier.
Finally, while indentation testing has been used extensively to extract the elastic modulus of coatings,
other techniques can often do a better job. For example, the elastic modulus of a surface layer can be
obtained with good accuracy using surface acoustic waves (SAWs). And since the depth sensitivity of
the wave is the acoustic wavelength in the surface, that is, inversely proportional to the frequency, sub-
micron films can be probed with frequencies in the 1–50 GHz range. Surface Brillouin scattering (SBS)
from SAWs in this range of frequencies is capable of determining elastic moduli of hard coatings,61 some
thinner than 10 nm.62

2.6  Summary
As tribology continues to look toward new materials and new bearing designs to meet the needs of
the engineering community, tribologists will have to rely more on surface analytical techniques to
solve problems. These approaches to understanding surface and subsurface damage were pioneered
by surface scientists who possessed equipment like AES, XPS, and SIMS to analyze thin solid and
liquid films and by materials specialists who had access to XRD, SEMs, and TEMs. Today, all these
techniques and more are available (at a price) to all who need to obtain direct evidence about materials
Surface Analysis 2-17

successes as well as failures. And, with continued development in both analysis and technology at
the nanoscale, tribologists will have to become more versed in the alphabet soup of techniques, even
as they continue to rely on their optical microscopes to “see” what is happening in their moving
mechanical assemblies.

Acknowledgments
The author wishes to thank NRL and ONR for support, and the following people and organizations
for contributing figures: The Evans Analytical Group; W.M. Rainford, Sheffield University; SKF
Corporation; P. Blau, ORNL and DOE.

References
1. W.M. Rainforth, R. Stevens, and J. Nutting, Deformation structures induced by sliding contact,
Philos. Mag. 66 (1992), 621–641.
2. W.A. Zisman, Relation of the equilibrium contact angle to liquid and solid constitution, in Contact
Angle, Wettability, and Adhesion, F.M. Fowkes (Ed.), Vol. 43, American Chemical Society, Washington,
DC, 1964, pp. 1–51.
3. A.W. Adamson, Physical Chemistry of Surfaces, 5th edn., Wiley, New York, 1990.
4. J.N. Israelachvili, Intermolecular and Surface Forces, Academic Press Inc., San Diego, CA, 1992.
5. M.K. Bernett and W.A. Zisman, Effect of adsorbed water on the critical surface tension of wetting on
metal surfaces, J. Colloid Interface Sci. 28 (1968), 243–249.
6. B.J. Kinzig and H. Ravner, Factors contributing to the properties of fluoropolymer barrier films,
Tribol. Trans. 21 (1978), 291–298.
7. D.H. Buckley, Surface Effects in Adhesion, Friction, Wear and Lubrication, Elsevier, Amsterdam, the
Netherlands, 1981.
8. JVST Special Issue: 50 Years of science, technology, and the AVS (1953–2003), P.A. Redhead and H.F.
Dylla (Eds.), J. Vac. Sci. Technol. A 21 (2003).
9. J.D. Watson and F.H.C. Crick, A structure for deoxyribose nucleic acid, Nature 171 (1953), 737–738.
10. J.I. Goldstein, D.E. Newbury, P. Echlin, D.C. Joy, C. Fiori, and E. Lifshin, Scanning Electron Microscopy
and X-Ray Microanalysis, Plenum Press, New York, 1981 (2nd edn., 1992).
11. A comprehensive list of material analysis techniques can be found at http://en.wikipedia.org/wiki/
Acronyms_in_microscopy
12. Chart Courtesy of Evans Analytical Group®, http://www.eaglabs.com/techniques/analytical_​
techniques/
13. Most techniques discussed in this chapter are summarized in Wikipedia articles of the same name
or acronym.
14. P.P. Kane and G.B. Larrabee (Eds.), Characterization of Solid Surfaces, Plenum Press, New York, 1974.
15. A.W. Czanderna (Ed.), Methods of Surface Analysis, Elsevier, New York, 1975.
16. J.C. Vickerman and I. Gilmore (Eds.), Surface Analysis: The Principal Techniques, John Wiley & Sons
Ltd., Chichester, U.K., 2009.
17. J.M. Walls (Ed.), Methods of Surface Analysis, Cambridge University Press, Cambridge, U.K., 1989.
18. K. Miyoshi and Y.W. Chung (Eds.), Surface Diagnostics in Tribology, World Scientific, Singapore,
1993.
19. W.A. Glaeser (Ed.), Characterization of Tribological Materials, Butterworth-Heinmann, Boston, MA,
1993.
20. G. Stachowiak and A.W. Batchelor, Experimental Methods in Tribology, Elsevier Science, Amsterdam,
the Netherlands, 2004.
21. M. Born and E. Wolf, Principles of Optics. Cambridge University Press, Cambridge, U.K., 1997. ISBN
0521639212; see also http://en.wikipedia.org/wiki/Diffraction-limited_system
2-18 Theory and Practice of Lubrication and Tribology

22. H.M. Pollock and D.A. Smith, The use of near-field probes for vibrational spectroscopy and photo-
thermal imaging, in Handbook of Vibrational Spectroscopy, J.M. Chalmers and P.R. Griffiths (Eds.),
John Wiley & Sons Ltd., Chichester, U.K., 2002.
23. Bearing failures and their causes, SKF Publication PI 401 E, online at http://www.skf.com/portal/skf/
home/aptitudexchange?contentId=0.237932.237933.237935.237962.238715
24. Public domain image from http://en.wikipedia.org/wiki/File:Kontakt_Spannungsoptik.JPG
(accessed March 2011)
25. F.J. Giessibl, Advances in atomic force microscopy, Rev. Mod. Phys. 75 (2003), 949–983.
26. I.L. Singer, Surface chemistry and mechanical behavior of silicon carbide and silicon nitride
implanted with titanium to high fluences and high temperatures, Surf. Coat. Technol. 33 (1987),
487–499.
27. M.N. Gardos and S.A. Gabelich, Atmospheric effects of friction, friction noise and wear with silicon
and diamond. Part I. Test methodology, Tribol. Lett. 6 (1999), 79–86.
28. R. Crockett, S. Roos, P. Rossbach, C. Dora, W. Born, and H. Troxler, Imaging of the surface of human
and bovine articular cartilage with ESEM and AFM, Tribol. Lett. 19 (2005), 311–317.
29. D.N. Dunn, K.J. Wahl, and I.L. Singer, Nanostructural aspects of wear in ion-beam deposited Pb-Mo-S
films, in MRS Proceedings Volume 522, Fundamentals of Nanoindentation and Nanotribology, N.R.
Moody, W.W. Gerberich, S.P. Baker, and N. Burnham (Eds.), MRS, Pittsburgh, PA, 1998, pp. 451–456;
K.J. Wahl, D.N. Dunn, and I.L. Singer, Wear behavior of Pb-Mo-S solid lubricating coatings, Wear
230 (1999), 175–183.
30. J.-H. Wu, D.A. Rigney, M.L. Falk, J.H. Sanders, A.A. Voevodin, and J.S. Zabinski, Tribological behav-
ior of WC/DLC/WS2 nanocomposite coatings, Surf. Coat. Technol. 188–189 (2004), 605–611.
31. T.W. Scharf, P.G. Kotula, and S.V. Prasad, Friction and wear mechanisms in MoS2/Sb2O3/Au nano-
composite coatings, Acta Mater. 58 (2010), 4100–4109.
32. I.L. Singer, Surface morphologies produced by ion milling on ion-implanted 18Cr8Ni steels, J. Vac.
Sci. Technol. 18 (1981), 175–178.
33. T.R. Dulski, Trace Elemental Analysis of Metals—Methods and Techniques, Marcel Dekker Inc.,
New York, 1999.
34. I.L. Singer, C.A. Carosella, and J.R. Reed, Friction behavior of 52100 steel modified by ion implanted
Ti, Nucl. Instrum. Methods 182/183 (1981), 923–932.
35. R. Behrish (Ed.), Sputtering by Particle Bombardment, 1st edn., Springer, Berlin, Germany, 1981.
36. H.C. Feng and J.M. Chen, Effects of low-energy argon-ion bombardment on MoS2, J. Phys. C (Solid
State Phys.) 7 (1974), L75.
37. G. Betz and G.K. Wehner, Sputtering of multicomponent materials, in Sputtering by Particle
Bombardment II, Topics in Applied Physics, R. Behrisch (Ed.), Vol. 52, Springer, Berlin, Germany,
1983, 11–90.
38. R.J. Colton, M.M. Ross, and D.A. Kidwell, Secondary ion mass spectrometry: Polyatomic and molec-
ular ion emission, Nucl. Instrum. Methods Phys. Res. Sect. B 13 (1986), 259–277.
39. P.D. Ehni and I.L. Singer, Electron-beam microprobe analysis of the wear behavior of sputter-­
deposited MoS2 coatings, Appl. Surf. Sci. 59 (1992), 45–53.
40. C. Blatchley, Radionuclide methods, in Metals Handbook: Friction, Lubrication and Wear Technology,
P. Blau (Ed.), Vol. 18, ASM International, Materials Park, OH, October 1992, pp. 319–330.
41. G. Klingelhofer, B. Bernhardt, J. Foh, U. Bonnes, D. Rodionov, P.A. DeSouza, Ch. Schroder,
R. Gellert, S. Kane, P. Gutlich, and E. Kankeleit, The miniaturized Mössbauer spectrometer MIMOS
II for extraterrestrial and outdoor terrestrial applications: A status report, Hyperfine Interact. 144/145
(2002), 371–379.
42. J.A. Sawicki and B.D. Sawicka, Conversion electron Mössbauer spectroscopy: A new tool for
plasma-induced surface modification studies, Nucl. Instrum. Methods B 23 (1987), 482–486;
D.L. Williamson, F.M. Kustas, D.F. Fobare, and M.S. Misra, Mössbauer study of Ti-implanted 52100
steel, J. Appl. Phys. 60 (1986), 1493–1500.
Surface Analysis 2-19

43. F.R. Van De Voort, J. Sedman, R.A. Cocciardi, and D. Pinchuk, FTIR condition monitoring of
in-service lubricants: Ongoing developments and future perspectives, Trib. Trans. 49 (2006),
410–418.
44. I.L. Singer, S.D. Dvorak, K.J. Wahl, and T.W. Scharf, Role of third bodies in friction and wear of pro-
tective coatings, J. Vac. Sci. Technol. A 21 (2003), S232–S240.
45. J.S. Zabinski and N.T. McDevitt, Raman spectra of inorganic compounds related to solid state
­tribochemical studies, Wright Laboratory Technical Report-96-4034, 1996, http://www.dtic.mil/cgi-
bin/GetTRDoc?AD=ADA310647&Location=U2&doc=GetTRDoc.pdf (accessed April 2011)
46. A.C. Ferrari and J. Robertson, Resonant Raman spectroscopy of disordered, amorphous, and dia-
mond like carbon, Phys. Rev. B 64 (2001), 75414.
47. T.W. Scharf and I.L. Singer, Thickness of diamond-like carbon coatings quantified with Raman spec-
troscopy thin solid films, Thin Solid Films 440 (2003), 138–144.
48. F.M. Piras, A. Rossi, and N.D. Spencer, Combined in situ (ATR FT-IR) and ex situ (XPS) study of the
ZnDTP-iron surface interaction, Tribol. Lett. 15 (2003), 181.
49. K. Sasaki, N. Inayoshi, and K. Tashiro, Friction-induced dynamic chemical changes of tricresyl phos-
phate as lubricant additive observed under boundary lubrication with 2D fast imaging FTIR-ATR
spectrometer, Wear 268 (2010), 911–916; K. Sasaki, N. Inayoshi, and K. Tashiro, Development of new
in situ observation system for dynamic study of lubricant molecules on metal friction surfaces by
two-dimensional fast-imaging Fourier-transform infrared-attenuated total reflection spectrometer,
Rev. Sci. Instrum. 79 (2008), 123702.
50. A. Hammiche, H.M. Pollock, M. Reading, M. Claybourn, P.H. Turner, and K. Jewkes, Photothermal
FT-IR spectroscopy: A step towards FT-IR microscopy at a resolution better than the diffraction
limit, Appl. Spectrosc. 53 (1999), 810.
51. H.M. Pollock and D.A. Smith, The use of near-field probes for vibrational spectroscopy and photo-
thermal imaging, in Handbook of Vibrational Spectroscopy, J.M. Chalmers and P.R. Griffiths (Eds.),
John Wiley & Sons Ltd., Chichester, U.K., 2002.
52. S. Jacobsson, M. Olsson, P. Hedenqvist, and O. Vingsbo, Scratch testing, in Friction Lubrication and
Wear Technology, Vol. 18, ASM International, Materials Park, OH, 1992, pp. 430–437.
53. S.J. Bull, Failure modes in scratch adhesion testing, Surf. Coat. Technol. 50 (1991), 25–32.
54. S.J. Bull and E.G. Berasetegui, An overview of the potential for quantitative coating adhesion mea-
surement by scratch testing, Tribol. Int. 39 (2006), 99–114.
55. Standards are given at http://www.iso.org/iso/iso_catalogue/catalogue_tc/catalogue_tc_browse.
htm?commid=53558 (accessed April 2011)
56. A.C. Fischer-Cripps, Nanoindentation, 2nd edn., Springer, New York, 2004.
57. J. Chen and S.J. Bull, Approaches to investigate delamination and interfacial toughness in coated
systems: An overview, J. Phys. D: Appl. Phys. 44 (2011), 034001.
58. A.C. Fischer-Cripps, in Nanoindentation, 2nd edn., Springer, New York, 2004, Chapter 7.
59. S.A. Syed Asif, K.J. Wahl, R.J. Colton, and O.L. Warren, Quantitative imaging of nanoscale mechanical
properties using hybrid nanoindentation and force modulation, J. Appl. Phys. 90 (2001), 1192–1200.
60. H.M. Pollock and D.A. Smith, The use of near-field probes for vibrational spectroscopy and photo-
thermal imaging, in Handbook of Vibrational Spectroscopy, J.M. Chalmers and P.R. Griffiths (Eds.),
John Wiley & Sons Ltd., Chichester, U.K., 2002, p. 1472.
61. F. Vaza, S. Carvalho, L. Rebout, M.Z. Silva, A. Paúl, and D. Schneider, Young’s modulus of (Ti,Si)
N films by surface acoustic waves and indentation techniques, Thin Solid Films 408 (2002), 160–168.
62. M.G. Beghi, C.E. Bottani, A. Li Bassi, P.M. Ossi, B.K. Tanner, A.C. Ferrari, and J. Robertson,
Measurement of the elastic constants of nanometer-thick films, Mater. Sci. Eng. C 19 (2002), 201–204.
3
Roughness
3.1 Introduction....................................................................................... 3-1
Importance of Scale  •  Fundamental Nature of the Geometry
of Topographies  •  Objectives in Measurement, Analysis, and
Characterization
3.2 Measurement...................................................................................... 3-5
3.3 Analysis and Characterization........................................................3-8
Height Parameters  •  Material Ratio Curve  •  Spatial and Hybrid
Christopher A. Parameters  •  Multiscale Analysis and Characterization
Brown 3.4 Concluding Remarks....................................................................... 3-12
Worcester Polytechnic Acknowledgments....................................................................................... 3-12
Institute References..................................................................................................... 3-12

3.1  Introduction
The objective of this chapter is to discuss the relation between tribology and surface roughness, and more
broadly, surface metrology. Surface metrology is the measurement and analysis of surface topographies.
Handbook of Surface and Nanometrology (Whitehouse 2011) is recommended to readers wishing
more detail on this subject. In his weighty tome, David Whitehouse does an excellent job of reviewing
the state of the art in surface metrology, including surface roughness and its relation to friction. Here,
in this chapter, emphasis is placed concept of scale as the key underlying concept in surface metrology.
Scale is emerging as a relatively simple and effective concept for aiding in the understanding of rough-
ness and its interaction with tribological phenomena. The concept of scale applied to measurement and
analysis appears fundamental to the developing science of surface metrology. The concept of scale is
applied to the interactions that control tribological phenomena, the measurement of rough surfaces,
and to the analysis of rough surfaces.
Conventionally, the term roughness refers to the finest scales of the geometry of the surface of an
object. This surface geometry is often called the topography. The largest scales of an object generally
describe its form. Form can comprise things like spheres, cylinders, and rectangular prisms. Waviness
is the term applied to the intermediate scales. The measurement and analysis of topography are called
surface metrology.
The roughness of rubbing surfaces influences the friction, wear, and the retention and distribution of
lubrication. In turn, the wear influences the roughness. Much research has been done on these relation-
ships, but there is still much to learn. Despite decades of research, it can be said that the predictive power
about such relations, which would provide the possibility of controlling tribological phenomena with
certain kinds of designed and manufactured roughness, is lacking. The design of surface roughness for
tribological phenomena is based on experience, rather than a derivation from first principles. There is no
simple set of fundamental rules for the design of rough surfaces for tribological interactions that could
be applied to a broad range of design problems.

3-1
3-2 Theory and Practice of Lubrication and Tribology

Second First
scale scale
Tribologic behavior Manufacture or wear
Topography

Measurement third scale

Measured
surface

Analysis fourth scale


Manufacture or
Behavior wear
characterization Characterization characterization

Discrimination and
correlation

FIGURE 3.1  Propagation of scale from surface manufacture of wear interacting at first scale (to create a surface
and tribological phenomena interacting at a second scale, which is measured at a third scale, then analyzed, and
characterized at a fourth scale).

Tribological phenomena involve many kinds of mechanisms that can interact with the surface fea-
tures at or over a variety of ranges of scales. Rough surfaces are challenging to characterize in a way that
captures the essence of the often complex geometries that make up the topographies at the fine scales
that influence tribological phenomena. Logically, the characterization should include the appropriate
geometric nature of the features, for example, inclinations, area, and radii, at the appropriate scale.
A good understanding of the basics of measurement and analysis of roughness is important to under-
standing and advancing documentation of the relations associated with tribological phenomena. New
methods in the measurement and characterization of roughness have the potential to enable new dis-
coveries about these relations. Measurement and characterization of roughness are discussed in the
following sections of this chapter.
This section attempts to help in understanding the definition of scales of the geometry of surfaces
that include roughness. The scales of interaction that control tribological phenomena logically should
be related to these roughness scales. The user also needs to be concerned with the scales that are inher-
ent to different kinds of measurement and different kinds of characterization of roughness. If the scales
of interaction involved in the tribological phenomena are not included in the scales of measurement
and in the analysis used for characterization, then correlations between tribological phenomena will be
difficult or impossible to establish (Figure 3.1). Similarly, it may not be possible to discriminate surfaces
that behave differently with respect to friction and wear.

3.1.1  Importance of Scale


Scale should serve as an essential context for understanding how to measure and characterize rough-
ness, which is crucial to understanding how roughness influences, and is influenced by, tribological
phenomena. The concept of scale in regard to surface features is important in understanding the inter-
action between roughness and tribological phenomena. Conventionally, the scale ranges used in defin-
ing roughness are broad. A more thorough approach to analyzing the decomposition of the topography
with respect to scale, than the method that has been used traditionally, can be important to developing
further insights into roughness and tribology. A finer resolution is needed, which can be achieved by
partitioning into smaller divisions the relatively large range of scales, which is used traditionally to
characterize roughness. This will assist in determining which scales are involved in the interactions that
Roughness 3-3

influence tribological phenomena. The knowledge of these critical scales will assist in the design and
manufacture of surfaces that control friction and in formulating more fundamental hypotheses about
the relations between roughness and tribology.
The reader should note that use of the terms relating to roughness can vary with the technical disci-
pline. In mechanical engineering, roughness, waviness, and lay make up the texture (ASME B46.1 2009).
The lay describes the directionality of the features on an anisotropic surface, such as the grinding direc-
tion. In civil engineering, the roughness of pavements refers to geometry on a scale that influences the
feeling of the ride in a vehicle. The texture and microtexture are finer scales, which would tend to influ-
ence the pavement-tire friction and wear (ASTM E867-06 2006). Material scientists use the term texture
to describe grain orientation, which might develop during rubbing or processing such as rolling and
extrusion. Food scientists use the word texture to refer to the way food feels, emphasizing the rheological
properties. In this chapter, the term texture will be used in the sense that mechanical engineers use the
term. Topography will also be used to mean the geometry of the surface region. The terms roughness will
be used the way it is defined in mechanical engineering (ASME B46.1 2009), to refer to the finest scales
in the topography of a surface. The next section undertakes description and definition of those scales.

3.1.2  Fundamental Nature of the Geometry of Topographies


Part of the difficulty in establishing relations between roughness and tribological phenomena is the chal-
lenge in appropriately describing the complex geometry of the topography. In this chapter, two approaches
to defining roughness in topographies are discussed: one, traditional, is based on standardized cutoffs in
wavelengths and another is based on the nature of the geometry of the topography. The first is important
because engineers often use the traditional approach. The second is important because it has a basis in
fundamental geometrical distinctions, and can be helpful because it distinguishes regions in the geo-
metrical continuity of scales used in product and process design on a fundamental geometric basis.
The definition of the cutoff wavelengths separating these three topographical regimes—form, wavi-
ness, and roughness—can be found in the standards (ASME B46.1 2009). The selection of the appropri-
ate cutoffs depends on the intended use of the surface. Traditionally, the most commonly used cutoffs
were appropriate for many applications in the auto industry during the mid to late twentieth century,
because it was the auto industry that participated most heavily in the development of the standards at
that time. In bearings, for example, waviness has been said to be the wavelengths in the topography that
would cause vibrations, whereas roughness influences the lubricant retention and the wear.
There is a fundamental difference between the way the form and the roughness are specified in engi-
neering designs, as well as between how they are verified. For the moment, waviness will be considered
to exist in a kind of vague interface region of scales between the two regimes of form and roughness.
Waviness is more similar to roughness, in its complexity, measurement, and characterization, than it is
to form. Form is generally defined at relatively gross dimensions using Euclidian geometry. The geomet-
ric product specifications used to describe form are generally applied to regular shapes like the diameter
of a sphere or cylinder, or the extent and inclination of a plane. Conventionally, form is often verified
with calipers, or a micrometer, or on a coordinate measuring machine. Form is usually measured using
relatively few measurement locations. Roughness is generally defined at fine scales, where the geometry
has important irregular components that are chaotic or fractal, as opposed to Euclidean. Roughness
specifications are generally statistical characterization parameters, and the measurements typically
include thousands to millions of elevations.
The key feature of Euclidean geometry that distinguishes it from fractal or chaotic geometries is that,
when magnified sufficiently, the Euclidean shapes are smooth. When the form is defined in a design as a
Euclidean shape, the positions of all the points on that form are defined precisely by the equations that
define the shapes. The fact that tolerances are placed on the form does not diminish the precision with
which the ideal intended form is described. When a sufficient number of finite locations on a design
of a Euclidean form are specified in a design, then all the locations on that designed form are specified
3-4 Theory and Practice of Lubrication and Tribology

exactly. The same cannot be said for irregular geometries that have important fractal or chaotic compo-
nents. No matter how many locations are specified, no other location is specified exactly.
In a design, when the roughness is properly specified with the cutoff wavelength (ASME Y14.5), then
it could be said that the value of larger scales, where the geometry is Euclidean and deterministic, is also
specified. Specifying the roughness defines the intended crossover scale between larger geometries at
scales that are Euclidean and deterministic and smooth, and the finer scales that have geometries that
are chaotic and rough and nondeterministic.
The specification of the scale of roughness can be due to practical considerations of the limits of
manufacturing. The methods of manufacturing parts or molds have scales of interaction that influ-
ence the crossover between form and roughness. Conventional machining by chip removal and abrasive
processes demonstrates limitations to the control of geometry, depending on the material, tools, and
process variables (Brown et al. 1996, Zang et al. 2002).
For some time there has been a tendency to design with Euclidean geometry to finer scales, even
into the nanometer range. There are numerous examples of these kinds of surfaces, often called struc-
tured surfaces, in the literature. The manufacture of structured surfaces is reviewed well by Whitehouse
(2011). These kinds of structured, or engineered, surfaces apply Euclidean, or deterministic, geometries
to structures at micro- and nanoscales. Often the structured surfaces are used on tribological surfaces
to control friction, lubrication, and wear. Because these structures are Euclidean, they fall outside the
definition of roughness based on geometric kinds, although because they are at fine scales, they could
fall into a definition of roughness based on scale alone. In this chapter, the discussion of roughness will
be confined to surface topographies that have strong fractal, or chaotic, components. MEMS, or micro-
electromechanical devices, are examples of structures that are engineered at scales that are in a range
that could be considered to be roughness by some traditional measures, based on conventional cutoffs.
That is not the case here.
In this chapter, a distinction is drawn between scales to define roughness and form that is based on
the surface of interest. The topography at the finer scales, where the topography of the surface is mostly
irregular or chaotic, will be considered rough. The geometry at larger scales, where the topography of
the surface is regular and deterministic, and best described with elements of Euclidean geometry, will
be considered to be form.
Deterministic, regular surfaces can provide better control over tribological phenomena. This means
that the form extends into the fine scales that have traditionally been rough. Logical design of effective
forms at the fine scales to control tribological interactions requires that these interactions with the
surface are understood. These structured or engineered surfaces can be expensive to manufacture and
to maintain, although not necessarily so costly (Whitehouse 2011, chapter 6). Chaotic surface topog-
raphies are often favored by nature. Some of the most common manufacturing methods for surfaces
rely on processes that are not highly deterministic at fine scales. Abrasive processes, for example, rarely
control the position or shape of the abrasive grains precisely. Processes with components that tend to
produce chaotic topographies are often used in manufacturing, perhaps because it was believed that it
might be too expensive to control the geometry to finer scales.
There is an important distinction between chaos and randomness with respect to the geometry of
rough surfaces. The sequence of heights or elevations on a rough surface is chaotic, not random. The
heights of consecutive locations on a rough surface are correlated to some extent. The surfaces consid-
ered here are continuous. One height cannot differ too much from the adjacent heights. If the sequence
were random, then there would be no correlation, and adjacent heights might be close or far away. The
correlation length can be used to characterize a rough surface; this is the length where the value of the
autocorrelation function falls to some small value. If the sequence of heights along a profile of a rough
surface were random, there would be no correlation, and the surface would not be continuous. A chaotic
surface can be characterized in a number of ways, which are discussed later. Chaotic geometries can be
usefully modeled and characterized as fractal. Fractals have the interesting property of being continu-
ous but nowhere differentiable. More specifically, the slopes, on inclinations, on a fractal surface cannot
Roughness 3-5

be determined at one point, that is, fractal curves do not have tangents in the sense that Euclidean
curves do.
Many geometric properties of rough surfaces change with respect to the scale of observation or cal-
culation. These scale-sensitive properties have been shown to include the length of profiles, areas of
surfaces, slopes or inclinations, radii of features, and the available volume below a reference surface or
between two rough surfaces. This variation of basic geometric properties with scale is an important con-
cept to bear in mind when modeling the effects of geometry for interactions like heat transfer and con-
tact mechanics, both important in tribology. This is also important to consider when trying to select or
develop useful characterization parameters for describing rough surfaces. This means that the sequence
of the heights is important in characterizing roughness, not just the distribution of heights.
Scale should be considered when developing the understanding of the interaction between rough
surfaces and tribological phenomena. Tribological phenomena have scales of interaction with the sur-
face roughness. These scales of tribological interaction can depend on many things. In some cases, the
scales of interaction may not be known, and it may not be obvious how to determine them theoretically.
In these cases, there are experimental approaches, which will be discussed later, that are like tuning
in a radio to find the wavelength for the emission of interest. Sometimes, the Euclidean design in a
structured or engineered surface may be fine enough to be included in the scales of interaction for the
tribological phenomena. Indeed, this is often the intent of the design.

3.1.3  Objectives in Measurement, Analysis, and Characterization


The objectives of measurement and analysis of surface roughness, that is, surface metrology, are to sup-
port product and process design, as well as the design of quality assurance systems. The measurement
and analysis of roughness can also be used to advance the understanding of roughness-related phe-
nomena. There are many phenomena that are suspected to be influenced by roughness, where the exact
nature of that influence is not well understood. Finding fundamental relationships between roughness
and interactions like contact mechanics and heat transfer will advance the science of tribology. There is
much still to be learned.
Correlation and discrimination are two important results of surface metrology. Discrimination is
the ability to distinguish the topographies of two surfaces, which are known to be different because of
processing or behavior, by using measurement and analysis of the topographies. Because the roughness
can vary from location to location on the surface, adequate sampling is necessary. The ability to dis-
criminate should be demonstrated statistically. Correlation is based on a regression analysis between a
processing or performance parameter and some characterization of the roughness. The regression coef-
ficient indicates the extent to which the regression line explains the way in which the relation between
the changes in the processing influences the change in the roughness, or the way in which the changes
in the roughness influences the change in the performance.
The discovery of reasonably strong performance and process correlations with roughness can be used
to support product and process design and to assist in process control. The ability to make such discrim-
inations between the roughnesses of surfaces is essential for the design of quality assurance systems.

3.2  Measurement
Roughness is traditionally measured with stylus profilers that trace profiles. A measured profile (z =
z(x)) consists of elevations, or heights, z, measured, or sampled, as a function of tracing position in x.
Areal measurements of textures are becoming more common. A measured surface (z = z(x, y)) consists
of heights measured, or sampled, as a function of spatial position in x and y. The profiles typically con-
tain thousands of heights, sampled at regular intervals along the trace. A measured surface can contain
millions of heights spaced on a grid at regular intervals. The intervals between the height measurements
are called the sampling intervals. The verification of roughness is generally confirmed by some kind of
3-6 Theory and Practice of Lubrication and Tribology

statistical evaluation of the measured heights. Roughness characterization parameters are discussed in
Section 3.3.
There is a range of scales included in measured profiles and textures. The largest scale is the length of
the trace (x) or the size of the region over which the texture is measured. The finest scale is influenced
by two factors. The sampling interval is the lower limit on the finest scale included in the measurement.
The size of the region over which an individual height is sampled is usually small, but it can be larger
than the sampling interval.
The finest scale in a measured profile or texture could be considered the resolution. Currently, there
is no US or ISO standard method for determining the resolution of instruments used in surface metrol-
ogy. The resolution in the spatial or horizontal directions (x, y) is usually different from that found in
the vertical or height direction (z).
The resolution depends on the sensor type and the way in which it interacts with the surface. The
interaction with the surface is influenced by the material properties of the surface as well as by the local
topography. Therefore, the resolution of a height sensor can vary from location to location on the same
surface. Furthermore, one height measurement is generally not sufficient to resolve a feature. The num-
ber of height measurements required to resolve a feature depends on what needs to be known about the
feature in order to resolve it.
The height sensors used in surface metrology are for the most part either tactile or optical. Atomic
probe microscopes can have a variety of sensor types, but that will not be covered in any detail here.
The finest scale on measurements is a function of the geometry of the sensor and the way in which it
interacts with the surface.
Tactile sensors have a large range of scales. Conventional stylus sensors generally have spherical tips
with radii between 1 and 10 μm, and are thereby conic. Atomic force microscopes are essentially small,
sensitive stylus instruments with tip radii in the order of a nanometer.
Tip shape provides an obvious geometrical limitation on the ability of stylus to resolve surface fea-
tures. The size and shape of the tip keep it from measuring steep valleys or the sides of steep cliffs. If
the radius of a mountain or ridge feature is smaller than the radius of the tip, the trace will show more
of the radius of the tip than the radius of the feature. The condition of the tip of a stylus can be checked
by tracing over the edge of a razor blade. It is possible to deconvolve slightly more true estimates of the
geometry from the trace by knowing the radius of the tip.
The reaction of the material on the surface to the tip is also an issue. A conventional stylus can scratch
a surface. While a smaller tip might be favored geometrically, a finite contact force is required, and the
contact stresses are sufficient to alter the surface (Brown and Savary 1991). The sensitivity of the height
registration and the sampling intervals in the stylus instruments can be much finer than the ability of
the stylus to resolve surface features.
Optical instruments were reviewed by Hocken et al. (2005) and are well covered by Whitehouse (2011),
so a brief review of commonly used types is given here. The height resolution of the optical instruments
can be in the order of nanometers, substantially smaller than the wavelength of the light they use. The
spatial resolution is limited by the wavelength. Optical instruments are not limited by surface hardness.
However, the transparency and reflectivity of the surface can be an issue. Generally, light is projected
onto the surface. If sufficient light is not projected from the surface back to the sensor in the instrument,
then the surface cannot be detected. Steep slopes can be difficult to measure. Highly reflective surfaces
and sufficiently dark surfaces may not return enough light to the sensor to allow a measurement of the
surface. In the case of poor detection, instruments may report aberrant heights measurements, which
look like spikes sticking up from, or poking down into, the surface.
Optical profilers are instruments that replace the contact stylus with a light stylus. Generally, they
consist of a metrology frame, a horizontal scanning system, a vertical adjustment device, and a height
sensor. Triangulation and focus variation are two common height sensors, both of which rely on a spot
of light striking the surface. These are horizontal scanning instruments, which are generally slower
than vertical scanning instruments. Either the surface to be measured or the sensor must be scanned
Roughness 3-7

over the spatial measurement range. It is interesting to note that the spot size is not necessarily a lower
limit on the resolution. It is not clear how the sensor treats height variations within the spot. There are
advantages to these systems. They can be adaptable, have good stand-off distances and large ranges, and
they can have good flexibility for adjusting the field of view.
Confocal and interferometric microscopes are vertical scanning microscopes. Some can include a
horizontal scanning component; however, it is only a light beam that is scanning, which is a relatively
fast process, compared to optical profilers. There are several types of interferometric microscopes,
which detect the interference of light from the surface with a reference beam. Confocal microscopes use
a narrow depth of field to determine the height by detecting when a location is in best focus during the
vertical scan. Figure 3.2 is made with a confocal microscope.
Structured light measurement systems, also called fringe projection systems, project patterns onto
the surface from one direction and capture images from another direction. The measured surface is
reconstructed from the distortion of the patterns due to the surface features. These systems are relatively
efficient, although they rely on the projection of several patterns to better cover the surface. They can
also cover large components. The resolution is by the pixel density and the ability of the image analysis
system to detect the pattern.
Stedman (1987) proposed a diagram for comparing the resolutions of measurement systems. The
diagram, which has been named for her, plots the range of vertical and spatial scales for an instrument
on the vertical and horizontal axes, respectively. Practically, it can be difficult to establish the fine-scale
resolution of instruments, and it can be more difficult to predict what that resolution will be on an
untried surface. This is particularly true of surfaces with steep slopes and diverse optical properties,
abrasive media for example. It can be necessary to test the measurement on the surface of interest to
establish how well it can be measured.
All of these measurement systems require some time to make the measurement and will therefore do
better with more stable thermal, mechanical, and electric–magnetic environments. Repeatability of the
heights at each location between consecutive measurements without changing the location is one way of
determining the noise that reduces the repeatability in the system. The heights at each location can be

µm
2.25

1.75

1.5
0
10

2.3 1.25
20
30

0
0
40

10 1
µm 0 50

20
30
6

40
0.75
70

50
60
80

70
90

µm 80 0.5
0

90
10

100
0
11

110 0.25
0

120
12

FIGURE 3.2  Image of a ground and lapped surface on a hardened steel ball used for sliding and sealing in a valve,
rendered from a measurement made by a laser scanning confocal microscope.
3-8 Theory and Practice of Lubrication and Tribology

regressed between the two consecutive measurements. The strength of the regression is an indication of
the amount of noise. Sources of noise can be manipulated, as by letting the air out of an air table, to note
the influence on the regression coefficient.
Reproducibility between different measurement systems can be difficult to achieve. Comparisons can
be poor between measurements on the same measurement system with different measurement param-
eters, such as different lenses on the optical microscopic systems. Filters have been developed for opti-
cal systems that improve the comparisons with contact stylus systems. Standards for determining the
uncertainty in surface metrology systems require more development.

3.3  Analysis and Characterization


The measured surfaces and profiles discussed earlier contain thousands to millions of elevations. These
can be used to generate images that may give some insight to roughness-related tribological phenomena.
Finally, it is necessary to characterize these measurements with a reasonably small collection of num-
bers that characterize the measurements in some useful way. If the measurement is of sufficient quality
and at appropriate scales, then the ability to discriminate and to find correlations depends on the analy-
sis and characterization. Two aspects of the characterization appear to be particularly important in the
characterization of roughness; these are the nature of the geometry that is characterized, for example,
slope, area, radius, volume, and the scale of the characterization.
There are several useful objectives that can be assigned to the characterization of measured profiles
and textures. One is to reach a consensus between measurement systems on the state of roughness of
the surface of a part. This can be within one company or between a supplier and customer. Another is
to support product and product design, by finding correlations between roughness and performance,
and between processing and roughness. Yet another objective is to support quality assurance systems by
discriminating between acceptable and unacceptable surfaces. An additional objective is to advance the
scientific understanding of interactions with rough surfaces. The worth of a measurement and charac-
terization system depends on the capability of that system to fulfill these objectives.
The statistical relevance and stability is important in establishing the worth of analysis and character-
ization systems. Using a one-way ANOVA, Wang and coworkers (2011) were able to identify parameters
that are statistically stable. Such parameters have more promise in characterizations that can discover
correlations with performance functions.
It is not the intent here to give a comprehensive description of all the characterization parameters
that appear in the literature and in the standards. Whitehouse (2011) does an excellent job of summariz-
ing conventional texture characterization. Useful definitions can be found easily on the Internet with
common search engines. The most common characterization parameters and parameter types will be
discussed here and integrated into the discussion of context of scale. This context provides a useful and
common basis for understanding the development and selection of characterization parameters with
measurement instruments and tribological phenomena that is not treated extensively elsewhere.
There are so many characterization parameters that Whitehouse was already inspired in 1982 to write
a paper about the parameter rash (Whitehouse 1982). The problem is not so much the large number of
characterization parameters; rather, it is the paucity of functional correlations between roughness and
performance and the necessity for a broader understanding of basic principles upon which a science of
surface metrology could be based. The use of scale decompositions, that is, a multiscale approach, is
discussed here to improve the ability to discriminate surfaces and to find useful correlations between
roughness and performance.
Conventional roughness parameters are calculated after filtering form and waviness from the mea-
sured profiles and textures. The measured surface or profile is leveled before filtering. A mean line is
established as part of the filtering, which divides the profile such that the sum of the heights would be
zero. The designations start with R for profile parameters and S for areal surface parameters, according
to ISO and ASME standards (ASME B46.1 2009). The importance of the selection of the waviness scale is
Roughness 3-9

often overlooked, but it should not be, since it can change the value of roughness parameters by as much
as 30%. For most of the common conventional parameters, the only way to introduce sensitivity to scale
is the selection of the waviness cutoff length. The cutoff values are generally predetermined, based on the
length or size of the measurement, rather than on the geometrical properties of the surface.

3.3.1  Height Parameters


Height parameters are the most frequently used of all the roughness characterization parameters. These
parameters rely only on the heights and not on their relative positions, which means that they cannot
identify feature shapes on a surface.
The most common is the arithmetic average roughness, Ra, or S a for a surface, the sum of the absolute
values of the heights:


Ra = ∑z
Valleys, or features below the mean line, or plane, have the same influence on the average roughness as
do peaks, or features above the mean line, or plane. Ra can be an indication that something has changed
on a surface. It should not be expected that Ra will be a good predictor of tribological performance,
except in some special cases where the surfaces are limited to specific types, as can happen when a nar-
row range of manufacturing processes and variables have been used.
The root mean square (rms) roughness, Rq, or Sq for a surface, is also commonly used, although more
often in the scientific and technical literature than in industry:

1
⎛1 ⎞

Rq = ⎜
⎝n ∑ Z2⎟

2

Rq is often referred to as the rms roughness or as σ, because it is the standard deviation of the heights
about the mean line or plane. Rq is the first statistical moment of the heights.
The second and third moments, skew Rsk and kurtosis R ku (S sk or Sku for surfaces), are also used, but
less frequently. A negative skew indicates that a profile is full, with smaller peaks than valleys, perhaps
plateaued. Surfaces with negative skew are used for sliding applications. The valleys tend to retain lubri-
cant and keep debris out of the high pressure interface. Figure 3.2 is an example of a surface for use in a
ball valve that was first ground then lapped to produce a bearing surface with a negative skew.
Koshy and Tovey (2011) find that the skew and kurtosis are good indicators of friction reduction on
the rake face of cutting tools when the roughness is modified by electric discharge machining. The sur-
faces in their study all had positive skews.
The peak-to-valley roughness, indicated by Rt or St, is the vertical distance from the lowest valley to
the highest peak. Rp and Rv (Sp and Sv) represent height peak and valley depth from the mean line (or
plane), respectively. The sum of Rp and Rv equals Rt. Because Rt relies entirely on two height samples,
it is subject to measurement anomalies, especially the kinds of spikes discussed earlier. Before relying
on these parameters, the measured profiles and surfaces should be checked for spike anomalies. For
surfaces, this can be done by looking at an image rendered from the measurement at a grazing angle, so
that the spikes can be seen protruding from the surface. R z or Sz are the averages of peak-to-valley cal-
culations over several sections. The number of sections used depends on the standard being used. This
averaging makes R z and Sz less susceptible to anomalies.
The ratio of the peak height to the peak-to-valley roughness can be another indicator of the fullness,
or a profile, along with the skew. If the Rp:R z ratio is greater than 0.5, then the profile would tend to be
full, and the skew would be positive.
3-10 Theory and Practice of Lubrication and Tribology

Height parameters tend to be highly correlated with each other (Nowicki 1985). This means that for
many surfaces, the height parameters report similar knowledge about the surface. It also means that the
height parameters will be difficult or impossible to adjust independently when designing manufacturing
processes.
Bigerelle et al. (2011) have recently introduced scale sensitivity into analysis of height parameters in
a novel way, by using the extreme amplitudes versus the observation length. They also employ a kind of
fractal analysis. With these methods, they are able to develop three stages of a grit-size effect in abrasion.
Stemp et al. (2008, 2009) have also studied the influence of the measurement length on height param-
eters, although in a less-complex way than Bigerelle, and on stone tools. Stemp found height parameters
to be unstable with respect to evaluation length, especially when compared to relative length that is
determined by length-scale fractal analysis (ASME B46.1 2009). The differences in the utility of the
height measurements could be due to the greater sophistication of Bigerelle’s approach, or to the nature
of the kinds of wear in the study.

3.3.2  Material Ratio Curve


The material ratio curve, also known as the Abbott–Firestone curve and bearing-area curve, is an
interesting way to look at the distribution of heights on a surface. The height on a scale from peak
to valley is plotted on the vertical axis. The percentage of surfaces at each height is plotted on the
horizontal axis. The curve shows what percentage of the material in a rough surface is present at
what height.
The curve is generally interpreted as having a core over its mid-40%. Properties of the total surface
in the peaks and valleys can be determined from the upper and lower extremities of the curve. These
are height parameters and should correlate somewhat with the skew. Parameters from this curve are
used to assess sliding surfaces, such as cylinder liners. Like the skew, although more nuanced, these
parameters can be used to determine if there might peaks that could protrude through a lubricant layer
and damage a counter face during sliding, and if there are sufficient valleys for retaining lubricant and
collecting debris.
The material ratio curve can be used to determine the relative volume of the peaks and valleys on a
surface. It does not show how the material is distributed among the peaks or the valleys. In other words,
the material ratio curve cannot discriminate between surfaces with a few large peaks or valleys or sur-
faces with many small peaks or valleys. This distribution could easily be supposed to have an influence
on tribological phenomena.

3.3.3  Spatial and Hybrid Parameters


Spatial, spacing, or length parameters characterize the distance between features. The mean spacing
of profile irregularities RSm used on profiles is one such spacing parameter (ASME B46.1 2009). A pro-
file irregularity is defined as a feature whose height has exceeded a selected value, after the profile has
descended below another selected value. The default is ±5% of R z about the mean line. The advantage of
spatial parameters is that they do not correlate well with the height parameters (Nowicki 1985). These
parameters report different knowledge about the profile and can be adjusted independently from the
height parameters.
Hybrid parameters depend on both vertical and lateral components of the measurement. The devel-
oped area, slope, and curvature are the most used of the hybrid parameters, although the use is small
compared to Ra, and is principally in research. These parameters are interesting because they depend on
the scale at which they are determined. The scale of their determination is dependent on the sampling
interval. Users may not be aware of the sampling interval or its influence on these parameters. If the
sampling interval is altered, as by changing the magnification of lens on the interferometric or confocal
microscope, these parameters can change dramatically.
Roughness 3-11

The ISO standards report the formula for the calculations in integral form, as does Whitehouse (2011).
This obscures the discrete nature of the actual calculations. All the measurements that these parameters
are calculated from are discrete.

3.3.4  Multiscale Analysis and Characterization


Multiscale analyses of measured surfaces are those that calculate some geometric properties over a
range of scales. These have also been referred to as scale-sensitive analyses. The characterization method
must capture the results of these analyses in a way that can be used for correlation and discrimination.
The power spectral density (PSD and APSD, for profiles and surfaces, respectively) is a method of
analysis and characterization that appears in the standards (ASME B46.1 2009). The PSD is a Fourier
decomposition of the measured profile or surface into spatial frequencies, which represent the scale. The
characterization for a profile is usually a plot of power versus frequency, which can be related to a linear
scale. The surface representation can show the power as gray scale intensity in the x- and y-directions.
Vorburger et al. (1993), in an interesting combination of measurement and analysis methods, show that
the angular distribution of scattered light is closely related to the PSD. PSD can be useful for visualizing
anisotropy, but it is difficult to translate succinctly in characterization that can be used quantitatively
for discrimination (Jordan and Brown 2006).
Wavelets can be used to filter measurements in a sequential manner according to the scale
(Muralikrishnan and Raja 2009). Zahouani et al. (2008) have used wavelets to understand the topog-
raphy of surfaces made by abrasive finishing over a range scales, or wavelengths. This is a multiscale
approach. It decomposes the topography using a continuous wavelet transform in two dimensions. The
researchers determined a multiscale transfer function to characterize surfaces created by abrasion over
a wide range of scales.
Area-scale analysis determines the apparent or calculated area of a measured surface as a function
of scale (ASME B46.1 2009). It uses a series of triangular tiling exercises, progressively using smaller
triangles, and covers a range of scales from the size of the measurement region to the sampling interval.
The results are depicted in a log–log, area–scale plot of the relative area versus the scale. The scale is the
area of the triangular tile used in the tiling exercise. The relative area is the ratio of the area calculated
from the tiling exercise to the nominal or (x–y) area, calculated as if the surfaces were perfectly smooth.
This is a kind of fractal analysis. The fractal dimension can be determined from the slope of the log–log,
area–scale plot. It is noted that on manufactured surfaces and on many other surfaces of engineering
interest the fractal dimension changes with scale. The relative area can be related to the inclinations, or
slopes, on the measured surface by a weighted average of one divided by the cosine of the inclination.
Area and inclination are two interesting geometric properties, because they can be related to transfer
phenomena and to plowing.
Jordan and Brown (2006) were able to use area-scale analysis to discriminate polyethylene ski bases,
which were ground with slightly different process variables in order to improve their sliding. At each
scale that was used for the tiling, a modified F-test was performed to determine mean square ratio,
which indicates the level of confidence in the difference between the relative areas of two grinds. The
plots of mean square ratio versus scale clearly show that there are certain scale ranges where the rela-
tive area can be used to discriminate the surfaces. They were also able to use the profile equivalent,
length-scale analysis to distinguish the grinding direction. A similar approach has been used on dental
microwear to discriminate diets (Scott et al. 2005).
Using area-scale analysis, Berglund et al. (2010a) were able to discover strong correlations between
die roughness and the friction coefficients of steel sheets against the dies. This was done by performing
regression analysis of the friction coefficient against the relative area at each of the tiling scales. The plot
of the regression coefficient versus the scale shows a maximum R 2 of about 0.9 at about 10 μm. Only two
conventional parameters were shown to have R 2 values about 0.7: the developed area ratio (Sdr), which
is essentially the relative area, but calculated only at the sampling interval, and the rms of the gradient,
3-12 Theory and Practice of Lubrication and Tribology

or slope, on the surface (Sdq), which is also evaluated at the sampling interval and related to the relative
area. In further analysis of the same experiment, Berglund et al. (2010b) systematically applied a sliding
band pass filter to the measured surfaces, covering the full range of scales in the measurement, and then
applied conventional analyses to the filtered surfaces. They found strong correlations, R 2 values exceed-
ing 0.9, with the narrowest bands for six characterization parameters, including S a and Sq.

3.4  Concluding Remarks


The utility of measurement and analysis systems needs to be tested on the surfaces of interest to see how
well they might work in that situation for the desired tasks.
Multiscale analyses applied over large-scale ranges on measured surfaces increase the probability that
strong correlations can be found or discriminations made between surfaces that are known to have been
made or to perform differently.
The ability to find correlations and to discriminate surfaces based on their roughness is improved
when certain scales, that is, wavelengths or spatial frequencies, are analyzed. This is similar to tuning a
radio to the desired frequency for a certain emission.

Acknowledgments
The author would like to gratefully acknowledge the support of Olympus Industrial Microscopes for
the LEXT OLS4000 ® Confocal Laser Microscope that was used to make the measurement rendered in
Figure 3.2; Digital Surf for the use of Mountains to render the image shown in Figure 3.2; and Elaine
Kristant, the WPI graduate who made the measurement.

References
ASME B46.1 2009, Surface Texture (Surface Roughness, Waviness, and Lay), ASME, New York, 2009.
ASTM E867-06, Standard Terminology Relating to Vehicle-Pavement Systems, ASTM International, West
Conshohocken, PA, 2006.
Berglund, J., C. Agunwamba, B. Powers, C.A. Brown, B.-G. Rosén, On discovering relevant scales in sur-
face roughness measurement—An evaluation of a band-pass method, Scanning 32 (2010b): 244–249.
Berglund, J., C.A. Brown, B.-G. Rosen, N. Bay, Milled die steel surface roughness correlation with steel
sheet friction, CIRP Annals Manufacturing Technology 59(1) (2010a): 577–580.
Bigerelle, M., S. Giljean, T.G. Mathia, Multi-scale characteristic lengths of abraded surfaces: Three stages
of the grit-size effect, Tribology International 44 (2011): 63–80.
Brown, C.A., W.A. Johnsen, R.M. Butland, Scale-sensitive fractal analysis of turned surfaces, Annals of the
CIRP 45(1) (1996): 515–518.
Brown, C.A., G. Savary, Describing ground surface texture using contact profilometry and fractal analysis,
Wear 141 (1991): 211–226.
Jordan, S.E., C.A. Brown, Comparing texture characterization parameters on their ability to differentiate
ground polyethylene ski bases, Wear 261 (2006): 398–409.
Koshy, P., J. Tovey, Performance of electrical discharge textured cutting tools, CIRP Annals—Manufacturing
Technology 60 (2011): 153–156.
Muralikrishnan, B., J. Raja, Computational Surface and Roundness Metrology, 2009.
Nowicki, B., Multiparameter representation of surface roughness, Wear 102 (1985): 161–176.
Scott, R.S., P.S. Ungar, T.S. Bergstrom, C.A. Brown, F.E. Grine, M.F. Teaford, A. Walker, Dental microwear
texture analysis within-species diet variability in fossil hominins, Nature 436(4) (2005): 693–695.
Stedman, M., Mapping the performance of surface-measuring instruments, Proceedings of SPIE 803
(1987): 138–142.
Roughness 3-13

Stemp, W.J., B.E. Childs, S. Vionnet, C.A. Brown, The quantification of microwear on chipped stone tools:
Assessing the effectiveness of root mean square roughness (Rq), Lithic Technology 33(2) (2008):
173–189.
Stemp, W.J., B.E. Childs, S. Vionnet, C.A. Brown, Quantification and discrimination of lithic use-wear:
Surface profile measurements and length-scale fractal analysis, Archaeometry 51(3) (2009): 366–382.
Vorburger, T.V., E. Marx, T.R. Lettieri, Regimes of surface roughness measurable with light scattering,
Applied Optics 32(19) (1993): 3401–3408.
Whitehouse, D.J., The parameter rash—Is there a cure? Wear 83 (1982): 75–78.
Whitehouse, D.J., Surface metrology, Measurement Science and Technology 8 (1997): 955–972.
Whitehouse, D.J., Handbook of Surface and Nanometrology, CRC Press, New York, 2011.
Zahouani, H., S. Mezghani, R. Vargiolu, M. Dursapt, Identification of manufacturing signature by 2D
wavelet decomposition, Wear 264 (2008): 480–485.
Zang, B., X. Liu, C.A. Brown, T.S. Bergstrom, Microgrinding of nanostructured material coatings, Annals
of the CIRP 51(1) (2002): 251–254.
4
Contact Mechanics
4.1 Introduction....................................................................................... 4-1
4.2 Basic Principles.................................................................................. 4-1
4.3 Elastic Contact of Nonconformal Bodies...................................... 4-2
4.4 Spherical Bodies................................................................................. 4-3
4.5 Cylindrical Bodies............................................................................. 4-5
4.6 Smooth Nonconforming Bodies..................................................... 4-5
4.7 Contact of Conformal Bodies.......................................................... 4-7
4.8 Effect of Friction................................................................................ 4-7
4.9 Elastic–Plastic and Fully Plastic Contacts..................................... 4-7
4.10 Scale-Dependent Plasticity............................................................. 4-10
4.11 Effect of Adhesion............................................................................ 4-10
4.12 Elastic–Plastic Contact with Adhesion........................................ 4-15
George G. Adams 4.13 Contact of Rough Surfaces............................................................. 4-16
Northeastern University References..................................................................................................... 4-19

4.1  Introduction
Contact mechanics is relevant to a wide variety of problems in classical and modern applications.
Examples from traditional areas include the contacts occurring in railroad wheels on rails, roller
bearings, gears, brakes, clutches, seals, electrical contacts, hard disk drives, and in friction modeling.
Although contact mechanics has a base in these traditional areas of tribology, there is growing interest
in contact mechanics in emerging areas such as in MEMS/NEMS and in biomechanics. Because of the
scaling effect, contact forces become relatively more important as the size of a body decreases. Thus,
MEMS/NEMS devices are prone to failure either by sticking together during fabrication or later while in
use. This “stiction” is perhaps the largest impediment to commercialization of these devices. In biology,
the contact and adhesion between cells affect the growth of organs and the spreading of cancer. Hip and
knee joints are subject to wear, and the resulting arthritis often leads to joint replacement. Of course,
contact stresses and the generation of wear debris is a major problem in the longevity of these prosthetic
joints. For a thorough treatment of contact mechanics theory, the reader is referred to the classical book
by Johnson [1].

4.2  Basic Principles


A nonconformal contact is defined as one in which the contact region, in the absence of deformation,
will be either a point or a line. Furthermore, when the effect of deformation is included, the dimensions
of the contact area are small compared to the local radii of curvature and the overall size of the bod-
ies. Thus the contact stresses are localized to the immediate vicinity of the contact region. This stress
localization allows the geometries of the contacting bodies to be described by the local geometries of

4-1
4-2 Theory and Practice of Lubrication and Tribology

the contacting bodies. As a first approximation, each surface can be represented by a pair of principal
curvatures in the vicinity of the contact, that is,

x2 y2
z(x , y ) = + (4.1)
2 Rx 2 R y

where
The common tangent plane is defined by z = 0
The x- and y-directions are in the directions of principal curvature
R x and Ry are the two principal radii of curvature at the point of initial contact

It is noted that the directions of the principal curvatures of the two bodies might not align, in which case
the x- and y-directions of each body will differ. Also note that Equation 4.1 is an approximation of an
ellipsoid for |x/R x|, |y/Ry | ≪ 1. Of course a special case is when these two radii of curvature are equal, in
which case the body is locally spherical.
In contrast, conformal bodies first make contact over a finite area and/or the contact dimensions are
a significant fraction of at least one of the local radii of curvatures. It is the nonconformal contacts that
tend to be technologically the most important and will be the focus of this chapter.
When two bodies are brought into contact, both normal and shear stresses may act across this inter-
face. The shearing stresses may be sufficient to prevent relative tangential displacement, that is, a stick
condition. In a slip region of the contact, relative tangential displacement of the two bodies occurs. Full
stick is said to occur if the entire contact is in stick, partial slip corresponds to regions of stick as well as
regions of slip, and complete slip implies that the entire contact region slips. In slip, there is a friction
law that typically relates the shearing stress to the normal stress. Sliding corresponds to complete slip
whereas rolling is accompanied by partial slip. Meanwhile the resultant force transmitted across the
contact may be in the normal direction, or it may have both normal and tangential components. Note
that although the force transmitted may be in the normal direction, the stresses may be a combination of
normal and shear, in which case the resultant of the shearing stresses vanishes. In a frictionless contact,
the shear stresses vanish everywhere.
The material behaviors that will be considered here are elastic and plastic. More specifically that
material behavior may be linear elastic, perfectly plastic, or plastic with strain hardening. In an elastic–
plastic or fully plastic contact, the stresses in the contact region include a plastic zone.
All real surfaces are rough to a certain extent. One way to model this roughness is to construct
a multi-asperity model of the surface, that is, replace the actual rough surface with a collection of
asperities that sit on the ideal smooth geometry. With each of these asperities described locally as a
micro-/nano-ellipsoid, the contact problem of the rough body reduces to the contact of a multitude
of these ellipsoids. These multi-asperity models will be discussed in detail toward the end of this
chapter.

4.3  Elastic Contact of Nonconformal Bodies


Elastic contact of nonconforming bodies is described by Hertz contact theory [2]. This theory was devel-
oped by Hertz while he was a graduate student during his Christmas vacation in 1880. Hertz contact [1]
assumes that (1) the dimensions of the contact area are much less than the radii of curvature of the unde-
formed contacting bodies, (2) the strains are sufficiently small for linear elasticity to be valid, (3) each
body can be described as an elastic half-space, and (4) the effects of friction and adhesion are negligible
so that only normal (compressive) tractions are transmitted. Assumption (3) may be viewed as valid for
two contacting bodies in which the contact radii are such that assumption (1) holds.
Contact Mechanics 4-3

Just as importantly Hertz theory can be applied to the contact of rough surfaces in which each indi-
vidual asperity contact with its mating surface is treated as a Hertz contact. In that case, the dimensions
of the individual asperity contacts must be much less than the radii of curvature of the asperity. These
points will be discussed in detail later in Section 4.13.

4.4  Spherical Bodies


Although Hertz contact pertains to the contact of locally ellipsoidal bodies, for simplicity the case of two
locally spherical elastic bodies will be treated first (Figure 4.1). The solution can be obtained from the
elasticity result that the axisymmetric pressure distribution given by

⎧⎪ p0 (1 − r 2 / a 2 )1/2 , r≤a
p(r ) = ⎨ (4.2)
⎩⎪0, r>a

acting on an elastic half-space produces a surface displacement in the normal direction [1] of

1 − n2 pp0
u(r ) = (2a2 − r 2 ), r ≤ a (4.3)
E 4a

The significance of Equation 4.3 is that the pressure distribution given by Equation 4.2 produces a
parabolic deformation of the elastic half-space in the loaded region. Hence for small displacements,
the curvature is constant and consistent with the deformation produced by a spherical indenter with
radius R, provided

d 2u (1 − n2 ) pp0 1
2
= − =− (4.4)
dr E 2a R

Thus, for two elastic bodies of different radii in contact, each of which is locally spherical at its peak, the
change in curvature of the two bodies add together resulting in the maximum contact pressure, which
is given by

2aEC
p0 = (4.5)
pR

R2
E2 , ν2 r
a

R1
E1, ν1

FIGURE 4.1  Hertz contact of two spherical bodies. The contact radius is given by “a.”
4-4 Theory and Practice of Lubrication and Tribology

where

1 1 1 1 1 − v12 1 − v22
= + , = + (4.6)
R R1 R2 EC E1 E2

where
R1 and R 2 are the radii of curvatures of the two spherical bodies
E1 and E2 are the elastic Young’s moduli
v1 and v2 are the Poisson’s ratios of the contacting bodies

Furthermore R and EC are, respectively, called the composite radius of curvature and the composite
(sometimes referred to as the reduced) elastic Young’s modulus
A simple integration of the pressure over the contact region gives the applied force P as

a
2

P = 2p rp(r )dr = p p0a 2
3
(4.7)
0

and by combining Equations 4.5 and 4.7 the contact radius is

1/ 3
⎛ 3PR ⎞
a=⎜ (4.8)
⎝ 4 EC ⎟⎠

The relative approach of distant points (δ) can be found by summing the displacements at r = 0 for the
two bodies, resulting in

pp0a
d= (4.9)
2EC

which, along with Equations 4.5 and 4.7, gives

1/ 3
a2 ⎛ 9P 2 ⎞
d= and d = ⎜ (4.10)
R ⎝ 16RE 2 ⎟⎠ C

Furthermore, the maximum contact pressure can be written as

1/ 3
3P ⎛ 6E 2 P ⎞
p0 = = ⎜ 3C 2 ⎟ (4.11)
2pa 2
⎝p R ⎠

The maximum von Mises stress, which is equal to twice the maximum shear stress, occurs subsurface
on the axis of symmetry. For v = 0.3, these maximums are at a depth of 0.481a and the corresponding
maximum von Mises stress is 0.620p0.
In summary, for a given applied force, Hertz contact gives the contact radius from Equation 4.8, the
approach from Equation 4.10, and the maximum contact pressure from Equation 4.11, with the compos-
ite radius and reduced elastic modulus defined by Equation 4.6. It is noted that the maximum contact
pressure is equal to three-halves of the average contact pressure.
Contact Mechanics 4-5

4.5  Cylindrical Bodies


For cylindrical bodies in contact, the contact half-width is given by

1/ 2 1/ 2
⎛ 4PR ⎞ 2P ⎛ EC P ⎞
a=⎜ , p0 = =⎜ ⎟ (4.12)
⎝ pEC ⎟⎠ pa ⎝ pR ⎠

where P is the load per unit length. Because the plane strain solution for the displacement far away from
the contact does not converge, it is not possible to define the approach. The maximum shear stress and
the maximum von Mises stress both occur subsurface below the middle of the contact. For v = 0.3 this
plane strain case gives a maximum von Mises stress of 0.558p0 at a depth of 0.704, whereas the maxi-
mum shear stress is 0.300p0 at a depth of 0.786a.

4.6  Smooth Nonconforming Bodies


For the more general case of smooth nonconforming bodies, the local geometry can be approximated
as ellipsoidal in the vicinity of the point of first contact, that is, the origin of the coordinate systems. At
point contact, the separation can be written as [1]

x2 y2
z= + (4.13)
2Rʹ 2Rʹʹ

where R′ and R″ are the principal relative radii of curvature, both of which are positive. These radii can
be found from

1 1 1 1 1 1
+ = + + +
Rʹ Rʹʹ R1ʹ R2ʹ R1ʹʹ R2ʹʹ
2 2 1/ 2
1 1 ⎡⎛ 1 1⎞ ⎛ 1 1⎞ ⎛ 1 1 ⎞⎛ 1 1⎞ ⎤
− = ⎢⎜ − ⎟ + ⎜ − ⎟ + 2⎜ − ⎟ ⎜ − ⎟ cos 2a ⎥ (4.14)
Rʹʹ Rʹ ⎢⎣⎝ R1ʹ R1ʹʹ⎠ ⎝ R2ʹ R2ʹʹ⎠ ⎝ R1ʹ R1ʹʹ⎠ ⎝ R2ʹ R2ʹʹ⎠ ⎥⎦

where
R1�and R1��are the principal radii of curvatures of body 1 whereas R2�and R2��are the principal radii
of curvature of body 2
α is the angle between the axes of principal curvatures, that is, between the directions of R1�and R2�

It is noted that R′ has been taken to be greater than R″ by choosing a positive sign before the square-root
term on the right side of Equation 4.14. Finally, the geometric mean of the principal relative radii of
curvature is defined by

Re = (RʹRʹʹ)1/2 (4.15)

The contact region will be elliptical of semi-major axis a and semi-minor axis b. The contact pressure is [1]

1/ 2
⎛ x2 y2 ⎞
p(x , y ) = p0 ⎜ 1 − 2 − 2 ⎟ (4.16)
⎝ a b ⎠

4-6 Theory and Practice of Lubrication and Tribology

The parameter e is defined by

1/ 2
⎛ b2 ⎞
e = ⎜1 − 2 ⎟
⎝ a ⎠ (4.17)

and its value is determined by the solution of

Rʹ (a/b)2 E(e) − K (e)


= (4.18)
Rʹʹ K (e) − E(e)

It is noted that the ratio a/b is determined entirely from the definition of e and is independent of the load
or material properties.
Now the function

1/ 6
1/ 3 ⎧⎪ ⎡⎛ a ⎞ 2 ⎫⎪
⎛ 4 ⎞ ⎤
G(e) = ⎜ 2 ⎟
⎝ pe ⎠ ⎨ ⎢⎜⎝ ⎟⎠ E(e) − K (e)⎥ [ K (e) − E(e)]⎬ (4.19)
b
⎩⎪ ⎢⎣ ⎥⎦ ⎭⎪

is defined, which is slightly different from that given in [1]. The complete elliptic integrals of the first
kind (K(e)) and of the second kind (E(e)) are considered standard functions, which are defined by

p /2 p /2
dj
K (e) =
∫ 1 − e 2 sin2 j
, E(e) =
∫ 1 − e 2 sin2 j dj (4.20)
0 0

However, these integrals are defined somewhat differently in some software packages, where they are
defined by K(e1/2) and E(e1/2), respectively.
The relations between the semi-major axis of the contact area, the approach, the force, and the maxi-
mum contact pressure are given by

1/ 3 1/ 3
⎛ 3PRe ⎞ ⎛ 9P 2 ⎞ K (e) 3P
a=⎜ G(e), d = ⎜ 3 2 ⎟ , p0 = (4.21)
⎝ 4 EC ⎟⎠ ⎝ 2p EC Re ⎠ G(e) 2pab

The maximum shear stress occurs below the surface and its values vary with the eccentricity of the
ellipse from 0.300p0 to a maximum of 0.325p0 [3].
Because of the complicated form of Equations 4.18 through 4.20, some simplified equations have been
developed. Here, we present the result of Greenwood [4], which is valid with a maximum error of 6% for
R′/R″ < 100. An equivalent radius is defined according to

1/ 3
⎡ 2(RʹRʹʹ)2 ⎤
Rˆe = ⎢ ⎥ (4.22)
⎣ Rʹ + Rʹʹ ⎦

which can be used with Equation 4.8 to determine the contact radius. However, when computing the
approach from the contact radius (Equation 4.10), the equivalent radius must be modified according to

e = Re2 /3Rˆ 1/3


R (4.23)

Contact Mechanics 4-7

4.7  Contact of Conformal Bodies


Consider a circular cross-sectional flat-ended rigid punch of radius a pressed against an elastic half-
space by a resultant punch force P. The displacement of the punch is related to the force according to

(1 − ν2 )P (4.24)
d=
Ea

and the stress distribution is given by

P
tzz (r , 0) = − , r<a (4.25)
2p a 2 − r 2

It is noted that this stress distribution is singular around the circular boundary of the punch in much
the same manner as the stresses at the tip of a crack.
The plane strain solution for a flat-ended punch pushed against an elastic half-plane also gives singu-
lar stresses at its ends. However, it is not possible to relate the force to the displacements for this plane
strain contact problem.
In the contact of non-flat conforming bodies, a small increase in load can lead to a large increase in
the contact area. A detailed discussion of these contacts is given by Johnson [1], which includes a sum-
mary of the results of Stuermann [5] for axisymmetric conforming contacts.

4.8  Effect of Friction


Consider now the contact of a spherical Hertz geometry in which both a normal force P and a tangen-
tial force T act. Our attention is restricted to the case in which there is negligible coupling between
normal/shear stresses and relative tangential/normal displacements. This uncoupling occurs exactly for
identical bodies, if one body is rigid and the other incompressible (v = 1/2), or for certain other cases.
In other cases, this coupling is generally small and will be neglected in the following. This problem has
been studied independently by Catteneo [6] and Mindlin [7]. As the tangential load increases, the cen-
tral stick region is surrounded by an annular slip zone, which begins to grow inward. As the tangential
force approaches the coefficient of friction times the normal load, the stick zone disappears leading to
pure sliding.
Before sliding occurs, the tangential displacement is related to the tangential and normal forces by

2/3
3mP ⎛ 2 − v1 2 − v2 ⎞ ⎡ ⎛ T ⎞ ⎤
dx = + ⎢1 − 1 − ⎥ (4.26)
16a ⎜⎝ G1 G2 ⎟⎠ ⎢ ⎜⎝ mP ⎟⎠ ⎥
⎣ ⎦

where
G1 and G 2 are the shear moduli of the contacting bodies
μ is the coefficient of friction
a is the contact radius computed from normal contact alone

4.9  Elastic–Plastic and Fully Plastic Contacts


It is important to understand the limit of elastic contact. As the applied load increases, the onset of plas-
tic yielding occurs below the surface of the body with the lower yield strength. If either the Tresca or von
Mises yield criterion is used for a spherical contact, then for Poisson’s ratio v = 0.3, yielding initiates [1]
4-8 Theory and Practice of Lubrication and Tribology

along the axis of symmetry at a depth of 0.481a below the surface when the maximum contact pressure
p0 = 1.61Y. This value of the maximum contact pressure for elastic behavior corresponds to a yield load
(PY) given by

p 3 p03R2 Y 3R2
PY = 2
≅ 21.6 2 (4.27)
6 EC EC

where Y is the yield strength of the lower yield strength material.


For arbitrary values of Poisson’s ratio, Chang et al. [8] determined the onset of plastic deformation
according to

p 3K 3H 3R2 K 3Y 3R2
PY = 2
≅ 113 (4.28)
6 EC EC2

where K = 0.454 + 0.41ν, the hardness H = 2.8Y, and

2
⎛ 1.4pKY ⎞ 4 E a3
dY = ⎜ ⎟ R, aY = RdY , PY = C Y (4.29)
⎝ EC ⎠ 3R

In Equation 4.27, δY and aY are, respectively, the values of the approach and contact radius at which
yielding initiates.
In an elastic contact it is only the composite radius of curvature, not the individual radii, which
affects the contact behavior. Such is not the case when the effect of plasticity is included. Thus, two cases
need to be treated separately. Consider first the indentation of a deformable body by a rigid sphere.
As the load continues to increase above PY, the size of the plastically deformed subsurface region also
grows, but it is surrounded by a reinforcing elastic region during this elastic–plastic phase. The plastic
zone then breaks through to the surface. Eventually the contact is said to be fully plastic, with the mean
contact pressure thought to be almost constant in the contact region and not changing with a further
increase in load. This average contact pressure is called the hardness.
This transition between elastic and fully plastic behavior was investigated by many investigators
using simplified analytical models. Tabor [9] claimed that for ductile metals, the mean pressure during
full plasticity (i.e., the hardness H) is related to the yield stress in simple compression according to

H = CY (4.30)

where C was thought to be a constant that is often taken to be in the range 2.8–3.0. However, the use of
a constant value of C has recently been called into question. Kogut and Komvopoulos [10] used finite
elements to show that the ratio of hardness (for spherical indentation) to yield strength depends on the
ratio of the composite elastic modulus to the yield strength, that is,

H ⎛E ⎞
= 0.201 ln ⎜ C ⎟ + 1.685 (4.31)
Y ⎝Y ⎠

It is readily shown from Equation 4.31 that if EC/Y = 700 then H/Y = 3.0 and when EC/Y = 250 then
H/Y = 2.8. However for EC/Y = 60 then H/Y = 2.5. Furthermore, the approach at the inception of fully
plastic deformation was given by

d 1
= , aʹ = (2R − d)d (4.32)
aʹ 1 + 0.037(EC / Y )
Contact Mechanics 4-9

where a′ is the truncated contact radius (i.e., the radius of the intersection of the sphere and the unde-
formed top surface of the half-space). Note that for δ ≪ R (which is frequently the case) these expres-
sions simplify.
The elastic–plastic region was also investigated in [10] for loading and unloading. The relation-
ships among the force, approach, and contact area (A = πa2) during loading were curve fit in the range
{1.78/(EC/Y) ≤ δ/a′ ≤ [1 + 0.037(E/Y)]−1} by the following empirical relations given by

Aʹ ⎡⎛ E ⎞0.394 ⎛ d ⎞0.419 ⎤
= 2.193 − ln ⎢⎜ C ⎟ ⎜ aʹ ⎟ ⎥ ,
A ⎢⎣⎝ Y ⎠ ⎝ ⎠ ⎥⎦
0.656

( ) (d aʹ) ⎤
0.651
0.839 + ln ⎢ EC ⎥
P Y
= ⎣ ⎦ (4.33)
AʹY 0.394

( ) (d aʹ)
0.419 ⎤
2.193 − ln ⎢ EC ⎥
Y
⎣ ⎦

where A′ = π(a′)2 is the truncated contact area.


Now consider the contact between a deformable hemisphere and a rigid flat surface. A finite ele-
ment analysis of the elastic–plastic contact for this configuration was performed by Kogut and Etsion
[11]. Such a configuration is likely more applicable to the modeling of an asperity contact than is the
indentation of a half-space by a rigid sphere. The contact stresses and the growth of the plastic zone
from elastic to elastic–plastic and to fully plastic loading were determined. It was found that the
plastic core is surrounded completely by elastic regions for 1 ≤ δ/δY ≤ 6 so that the entire contact area
is elastic. For 6 ≤ δ/δY ≤ 68 the contact area is elastic–plastic whereas for δ/δY > 68 the entire contact
area is plastic. When δ/δY ≥ 110, the average contact pressure is equal to the hardness H = 2.8Y. The
relations among the force, the approach, and the contact radius were each fit by empirical relations
given by

1.425 1.136
P ⎛ d⎞ A ⎛ d⎞ d
= 1.03 ⎜ ⎟ , = 0.93 ⎜ ⎟ , 1≤ ≤6
PY ⎝ dY ⎠ AY ⎝ dY ⎠ dY
1.263 1.146

P ⎛ d⎞ A ⎛ d⎞ d
= 1.40 ⎜ ⎟ , = 0.94 ⎜ ⎟ , 6≤ ≤ 110 (4.34)
PY ⎝ dY ⎠ AY ⎝ dY ⎠ dY

Jackson and Green [12] performed an analysis similar to [11] but extended that calculation to larger val-
ues of the approach. They also found that the hardness depends on the yield strain that is qualitatively
similar to [11].
After the unloading of a plastically loaded contact, there will be a residual plastic deformation.
A curve fit to the numerical results of Etsion et al. [13] is given by

dRes ⎛ 1 ⎞⎛ 1 ⎞
= 1− 1− (4.35)
dMax ⎜⎝ (dMax /dY )0.28 ⎟⎠ ⎜⎝ (dMax /dY )0.69 ⎟⎠

where
δ Max is the maximum approach during loading
δ Res is the residual (permanent) deformation

During subsequent load–unload cycles, most, if not all, of the deformation will be elastic.
4-10 Theory and Practice of Lubrication and Tribology

4.10  Scale-Dependent Plasticity


With modern application at micro- and nanoscales, there is considerable interest in depth-dependent
hardness. It has been observed experimentally that indentations and smaller depths often give sig-
nificantly greater hardness than the same material at the macroscale. The phenomenological based
strain-gradient plasticity theory has been used to explain this behavior at length scales down to
hundreds of nanometers [14]. Although the mathematical theory can be quite complicated, a simple
result comes from the mechanism-based strain-gradient plasticity theory as

h∗
H = H0 1 + (4.36)
h

where
H0 is the bulk hardness
h* is a material length scale

4.11  Effect of Adhesion


When a new surface is created in a solid, energy is required to break atomic bonds as well as to reorga-
nize the atoms near the surface. The work needed to create reversibly and isothermally a unit surface
area is called the surface energy γ. The surface energies of metals and covalent solids are typically
1–3 J/m2, for ionic solids 0.1–0.5 J/m2, and for molecular solids <0.1 J/m2. Also because atoms near the
surface of the body are acted upon by a different set of forces than are the atoms than in the bulk, a
tensile force is created parallel to the surface. This force per unit length is called the surface tension and
is numerically equal to the surface energy.
The work per unit area needed to separate two solids reversibly and isothermally is given by

w = g1 + g2 − g12 (4.37)

where
γ1 and γ2 are the surface energies of bodies 1 and 2, respectively
γ12 is the interfacial energy
w is the Dupré energy of adhesion (work of adhesion)

For two identical bodies, the interfacial energy vanishes and in that case w = 2γ.
For two half-spaces separated by a distance Z, the potential energy due to adhesion varies as a func-
tion of Z. When the separation distance is equal to the equilibrium separation of the surfaces (Z0),
which typically corresponds to a few Angstroms, the potential energy is a minimum. The well-known
Lennard-Jones potential gives the potential energy (U) per unit area as a function of the separation
distance Z according to

A ⎡ ⎛ Z0 ⎞ 2 1 ⎛ Z0 ⎞ 8 ⎤
U (Z ) = − ⎢⎜ ⎟ − ⎜ ⎟ ⎥ (4.38)
12pZ02 ⎢⎣⎝ Z ⎠ 4 ⎝ Z ⎠ ⎥⎦

This potential can be used for many solids with the exception of clean metal contacts as discussed later
in this section.
Contact Mechanics 4-11

1.2 Maugis model


1
0.8
0.6
Lennard-Jones
0.4
σ/σ0

0.2
0
–0.2
–0.4
–0.6
–0.8
0 0.5 1 1.5 2 2.5 3 3.5 4
Z/Z0

FIGURE 4.2  Variation of the normalized adhesive stress with the normalized separation distance of the two half-
spaces for the Lennard-Jones potential and for the Maugis model.

The derivative of the potential with respect to the separation distance gives the adhesive stress (σ) as
a function of separation (Z) for parallel surfaces by

A ⎡ ⎛ Z0 ⎞ 3 ⎛ Z0 ⎞ 9 ⎤
s (Z ) = ⎢⎜ ⎟ − ⎜ ⎟ ⎥ (4.39)
6pZ03 ⎢⎣⎝ Z ⎠ ⎝ Z ⎠ ⎥⎦

as shown in Figure 4.2. The constant A appearing in Equations 4.38 and 4.39 is called the Hamaker
constant and is related to the work of adhesion according to

A = 16pZ02w (4.40)

The maximum value of the adhesive stress is often called the theoretical stress σ0 and is related to the
Hamaker constant and the work of adhesion according to

A 16w
s0 = = (4.41)
9p 3 Z03 9 3 Z0

The work of adhesion is equal to the area under the stress vs. separation curve between the equilibrium
separation and infinity. Equivalently it is also equal to the negative of the potential evaluated at the
equilibrium position.
For clean metals the use of the Lennard-Jones potential, along with a Hamaker constant appropriate
for metals, leads to an underestimation of the surface energy. The strong adhesion is due to short range
forces that can be approximated by the interaction potential given by [15]

⎛ Z − Z0 ⎞
U (Z ) = −2g(1 + b Z ∗)exp(− b Z ∗), Z ∗ = ⎜
⎝ l ⎟⎠

Eg
s = s 0 bZ ∗ exp(1 − bZ ∗), s 0 ≅ 0.5 (4.42)
Z0

where λ is the Thomas–Fermi screening length and β = 0.9.


4-12 Theory and Practice of Lubrication and Tribology

Bradley [16] determined the adhesive force between two rigid spheres by integrating the intermolecu-
lar forces acting between the two bodies. The force required to separate the spheres, that is, the pull-off
force (P), was determined to be

P = −2pwR (4.43)

where the negative sign denotes a tensile force.


The effect of the elastic deformation of the contacting bodies is included in the Johnson, Kendall, and
Roberts (JKR [17]) model of adhesion of spheres. Their derivation uses an energy approach that accounts
for the change in elastic energy, the change in surface energy, and the work done by the applied load. The
effect of adhesion outside of the contact area is neglected. The stress distribution is tensile and infinite
at the boundary of the contact region. The attractive adhesive forces cause deformation beyond that
predicted by Hertz theory, resulting in an increase of the contact area. The relations among the contact
radius (a), the approach (δ), and applied load (P) are given by

3P1R a2 2paw
a3 = , d= − , P1 = P + 3pwR + 6pwRP + (3pwR)2 (4.44)
4 EC R EC

in which P1 is the effective Hertz load, that is, the applied load necessary to produce the same contact
radius in the absence of adhesion. Equation 4.44 reduces to Hertz contact when the work of adhesion
vanishes.
During loading in the JKR model, there is no interaction between the bodies until contact is first
made under a zero load. At that point, the contact region instantly becomes finite due to the sudden
inclusion of adhesion. The contact radius and approach under zero applied load are

1/ 3 1/ 3
⎛ 9pwR2 ⎞ ⎛ 3p 2w 2 R ⎞
a0 = ⎜ , d0 = ⎜ (4.45)
⎝ 2E ⎟⎠
C ⎝ 4 E 2 ⎟⎠
C

After continuing to load in compression and subsequent unloading, a tensile force is necessary to sepa-
rate the spheres. This separation occurs at a finite contact radius where

1/ 3 1/ 3
3 ⎛ 9pwR2 ⎞ a2 ⎛ 3p 2w 2 R ⎞
Pmin = − pwR, amin = ⎜ ≅ 0.630a0 , dmin = − = −⎜ (4.46)
2 ⎝ 8EC ⎟⎠ 3R ⎝ 64 EC2 ⎟⎠

The magnitude of the pull-off force is three-quarters of that predicted by Bradley for rigid spheres. It is
also observed that the JKR model produces hysteresis in the loading–unloading behavior as shown in
Figures 4.3 and 4.4.
The elastic adhesion model of Derjaguin, Muller, and Toporov (DMT [18,19]) is based on the Hertz
contact model but also takes into account the adhesive attraction in the region outside of the contact
area. Thus the effect of adhesion is equivalent to an increase in the contact force by an amount equal to
the adhesion force, that is,

4 EC a3 a2
= P + 2pwR, d = (4.47)
3R R

The results for the DMT theory are shown in Figures 4.3 and 4.4. As the two spheres approach each
other, the DMT theory predicts a pull-on force equal to that obtained by Bradley. Continued loading,
Contact Mechanics 4-13

1.5

1 Hertz

0.5

0 JKR
P/(πwR)

λ=1
–0.5
λ = 0.5 λ = 0.25
–1 λ = 0.1

–1.5 DMT

–2
–1 –0.5 0 0.5 1 1.5 2
δ/(π 2 w 2R/K 2)1/3

FIGURE 4.3  Dimensionless normal force vs. dimensionless approach for Hertz, JKR, DMT, and Maugis model.
Results for the Maugis model are given for various values of the adhesion parameter λ. In this figure, K = (4/3)EC .

2.5
0.5
1
λ=2
2
JKR
a/(πwR 2/K )1/3

1.5
0.25
0.1
1 DMT Hertz

0.5

0
–3 –2 –1 0 1 2 3 4
P/(πwR)

FIGURE 4.4  Dimensionless contact radius vs. dimensionless approach for Hertz, JKR, DMT, and Maugis model.
Results for the Maugis model are given for various values of the adhesion parameter λ. In this figure, K = (4/3)EC .

followed by subsequent unloading, produces separation at zero contact radius at a pull-off force equal
to the pull-on force, that is, there is no hysteresis. Thus the pull-off force is given by the same expression
(Equation 4.43) as that determined by Bradley.
It is noted that for the Bradley, JKR, and DMT theories, the pull-off force decreases with the radius
of the contacting bodies. However, this linear decrease in pull-off force is more than offset for by, for
example, the cubic decrease of force due to the weight of an object. Consider a scaling in which the char-
acteristic dimensions change from millimeters to micrometers. Although the pull-off force decreases by
a factor of 103, the weight decreases by a factor of 109 and the relative importance of adhesion increases
by a factor of 106.
It is noted that the JKR and DMT theories predict different values of the pull-off force. The JKR/DMT
models are most suitable when the range of surface forces is small/large compared to the elastic defor-
mations. The Tabor parameter (μ) is a measure of this ratio and is given by
4-14 Theory and Practice of Lubrication and Tribology

1/ 3
⎛ Rw 2 ⎞
m = ⎜ 2 3⎟ (4.48)
⎝E Z ⎠
C 0

The JKR theory is valid for large μ (greater than about 3), whereas the DMT theory is applicable for small
μ (less than about 0.1).
An analytical model was introduced by Maugis [20] and describes a continuous transition between
the JKR and DMT models. In this model the adhesive stress, which is usually taken equal to the theoreti-
cal stress (σ 0), is assumed constant in that portion of the separation region where the local separation is
less than some prescribed value h0. If the separation is greater than h0, then the adhesive stress vanishes.
Thus the work of adhesion for the Maugis model becomes equal to the product σ 0h0 and it is equated to
the work of adhesion from the Lennard-Jones potential (Figure 4.2). Thus

w
h0 = ≅ 0.97 Z0 (4.49)
s0

and the adhesive stress acts in an annular region (a < r < c) immediately outside of the contact region.
The inner radius of the annulus is the contact radius and the outer radius is determined such that the
separation at that radius is equal to h0.
The Maugis model gives nondimensional relations among the applied load (PM), contact radius (aM),
and approach (δM) as

laM
2
⎡ m2 − 1 + (m2 − 2)tan −1 m2 − 1 ⎤ + 4l aM ⎡ m2 − 1 tan −1 m2 − 1 − (m − 1)⎤ = 1
2

2 ⎣ ⎦ 3 ⎣ ⎦
2 ⎡ 4
3
PM = aM − laM m2 − 1 + m2 tan −1 m2 − 1 ⎤ , dM = aM
2
− aM l m2 − 1 (4.50)
⎣ ⎦ 3

where

1/ 3
⎛ 4 EC ⎞
aM = a ⎜
⎝ 3pwR2 ⎟⎠
1/ 3
⎛ 9R ⎞
l = 2s 0 ⎜ ≅ 1.16 m
⎝ 16pwEC2 ⎟⎠
c
m=
a
P
PM =
pwR
1/ 3
⎛ 16 E 2 ⎞
dM = d ⎜ 2 C2 ⎟ (4.51)
⎝ 9p w R ⎠

The results are plotted in Figures 4.3 and 4.4, along with the results of the Hertz, JKR, and DMT
theories. Because the Maugis formulation is analogous to the Dugdale model of a crack [21], in which
the plastic yield zone at the crack tip has a constant tensile stress, it is sometimes referred to as the
Maugis–Dugdale model.
Contact Mechanics 4-15

104

Hertz .05
103
P =0
P a/

Load P= P /πwR
JKR
102
δ1 M–D
/h
=0

δa/h0 = 0.05
1
0 DMT
10 .05

δa/h0 = 20
100 Bradley
(rigid)
10–1
10–3 10–2 10–1 100 101 102
Elasticity parameter λ = 1.16 µ

FIGURE 4.5  An adhesion map. (Reprinted from Johnson, K.L. and Greenwood, J.A., J. Colloid Interface Sci., 192,
326, Figure 5, 1997. With permission.)

A very useful adhesion map for the contact of elastic spheres is provided by Johnson and Greenwood
[22]. In a dimensionless load vs. Tabor parameter plane, the regions in which the Hertz, Bradley, DMT,
Maugis, and JKR models are applicable are identified (Figure 4.5).

4.12  Elastic–Plastic Contact with Adhesion


The adhesive contact of elastic–plastic spheres is complicated by the interaction between adhesion and
plasticity. Finite element analyses of the loading and unloading of two spheres [23] and of a hemisphere
with a rigid flat surface [24] both include the effects of adhesion and plasticity. These investigations
reveal that the unloading behavior goes from brittle to ductile depending on a loading parameter and
on a nondimensional parameter S defined by

w
S= (4.52)
Z0 H

in [25] and with the hardness replaced by the mean contact pressure in [23].
The characteristics of these separation modes are summarized by the following characteristics. In a
ductile separation the contact radius decreases slowly and stepwise before a sudden separation, typi-
cally at a significant fraction of the maximum contact radius. In a brittle separation, the contact radius
decreases steadily to a small value before separation. For a ductile separation, the hemispherical bump
is stretched significantly during separation due to plastic deformation; it is this behavior that best char-
acterizes ductile separation. The adherence force for a ductile separation is larger than for a brittle sepa-
ration. Finally a plastic neck is sometimes formed during a ductile separation, but not during a brittle
separation. Note that w/Z0 ≅ σ 0 so that Equation 4.52 represents the ratio of the theoretical stress to the
hardness. With this interpretation, it becomes apparent that the parameter S should play a central role
in governing the onset of ductile vs. brittle separation. One could envision a “tensile indentation” during
separation where, for S greater than about unity, the theoretical stress exceeds the hardness and leads
to a ductile separation.
It is noted that an earlier approximate analytical model [26], more recent molecular dynamics simula-
tions of nanocontacts (e.g., [27]), and experimental evidence (e.g., [28]) also reveal qualitatively similar
phenomena, that is, the existence of ductile and brittle separation modes.
During subsequent load–unload cycles without adhesion, most of the plastic deformation only
occurs on the first load cycle. In the elastoplastic/plastic deformation models with adhesion, there
may be significant plastic deformation on subsequent load cycles. Currently, no models or simulations
4-16 Theory and Practice of Lubrication and Tribology

of contacts under repeated loading–unloading of asperities in the presence of adhesion under gen-
eral situations exist. There are potential difficulties in predicting the deformed asperity radius after
a load–unload cycle. If the separation of the contacts is ductile, there may be material transfer as the
contacts separate.
For brittle separations in the DMT region, the adhesion force has been determined for elastic–plastic
deformation and curve fits given by Kogut and Etsion [29]. Simulations for this case also include unload-
ing and repeated loading [30,31].

4.13  Contact of Rough Surfaces


The contact of rough surfaces is often modeled using multi-asperity contact models. In these models,
a surface is replaced by a set of asperities; it is only the contact of these asperities with the contacting
surface that is considered. The total area that makes contact at the tips of the asperities is called the real
contact area and it is typically a very small fraction of the nominal contact area. In most such models,
the effect of each individual asperity is local and considered separately from the other asperities; the
cumulative effect is the sum of the actions of individual asperities. On the other hand, fully coupled con-
tact problems of rough surfaces are more complicated mathematically because the equations of elasticity
must be solved for the entire body simultaneously. Solution procedures typically lead to mixed bound-
ary value problems that can be solved analytically only for simple configurations. The emphasis in this
section is on multi-asperity contact models of rough surfaces.
Most of the approaches to multi-asperity contact modeling are related to the pioneering work of
Greenwood and Williamson (GW) [32]. In the GW model, which neglects adhesion, it is assumed that in
the contact between one rough and one smooth surface: (1) the rough surface is isotropic; (2) asperities
are spherical near their summits; (3) all asperity summits have the same radius of curvature while their
heights vary randomly; (4) there is no interaction between neighboring asperities; and (5) there is no
bulk deformation. With these assumptions, the Hertz contact theory is valid for each spherical asperity
on the surface of an otherwise flat elastic body.
The rough surface is replaced by a set of asperities (Figure 4.6), each with radius of curvature R, and
with a statistical distribution of heights given by the probability density function φ(z) defined by


prob(z > d ) = j (z )dz (4.53)
d

Thus, φ(z) is the probability that the height of an asperity, with tip height z above a reference plane
(taken to be the mean of the asperity heights), will be greater than d. If the rough surface is separated
from the rigid flat surface by d, then the number of asperities of height greater than d that make
contact (n) is

Flat surface of upper body


z

d
x

Mean of asperity heights


Mean of surface heights

FIGURE 4.6  Contact of a rough surface (lower body) with a flat rigid surface (upper body).
Contact Mechanics 4-17


n = N j(z )dz (4.54)
d

where N is the total number of asperities on the surface.


In the Hertz equations, the approach of a given asperity is δ = z − d and therefore the real contact
area is


A = N pR (z − d )j (z )dz (4.55)
d

and the force is


4
P=
3 ∫
NEC R1/2 (z − d )3/2 j (z )dz (4.56)
d

which are valid for arbitrary height distributions.


One such asperity height distribution is Gaussian, where

1 ⎛ −z 2 ⎞
j (z ) = exp ⎜ 2 ⎟ (4.57)
s 2p ⎝ 2s ⎠

and σ is the standard deviation of asperity heights. This distribution is realistic for many surfaces and
can be used to generate results numerically. However, a simpler distribution is given by

1 ⎛ −z ⎞
j (z ) = exp ⎜ ⎟ (4.58)
s ⎝s ⎠

This exponential distribution, while not as generally applicable as the Gaussian distribution, is a rea-
sonable approximation to the 25% highest asperities that are likely to be where most contacts occur.
Furthermore, the simple form of Equation 4.58 allows for analytical solutions for the number of con-
tacting asperities, the contact area, and the contact force, that is [32],

⎛ −d ⎞ ⎛ −d ⎞
n = N exp ⎜ ⎟ , A = N pRs exp ⎜ ⎟ ,
⎝ s ⎠ ⎝ s ⎠
1/ 2
⎛ −d ⎞ ⎛ s ⎞
P = N p 1/2 R1/2s 3/2 EC exp ⎜ ⎟ , ⇒ P = ⎜ ⎟ EC A (4.59)
⎝ s ⎠ ⎝ pR ⎠

It is noted that in the elastic contact of a single sphere with an elastic half-space, the contact area varies
as the two-thirds power of the applied force. However when the asperities have an exponential distribu-
tion of heights, the real area of contact varies linearly with the applied force. For a Gaussian distribution,
the force is nearly proportional to the real contact area over several orders of magnitude variation of the
applied load. This near constant elastic contact pressure can be thought of an “elastic hardness” in which
P/A does not change with increasing approach.
4-18 Theory and Practice of Lubrication and Tribology

Greenwood and Williamson also define a plasticity index ψ given by

EC s
y = (4.60)
H R

Interestingly it is the plasticity index, and not the load, which dominates the behavior of the contact.
When ψ is less than about 0.6 (smooth topography and high yield strain), the contact remains elastic for
a wide range of realistic forces. This behavior arises because as the load increases, more asperities come
into contact to support the increasing load. However, for ψ greater than 1.0, almost any load is sufficient
to cause some plastic deformation. It is noted that, aside from a constant factor of π1/2, ψ is the ratio of
the elastic contact pressure to the hardness.
Greenwood and Tripp [33] showed that the contact of two rough surfaces can be replaced with the
contact of a single rough body with a smooth surface. The elastic moduli and radii of curvature combine
as in the Hertz model and the equivalent standard deviation of asperity heights (σ) is given by

1 1 1 (4.61)
= +
s s1 s 2

where σ1 and σ 2 are the standard deviation of asperity heights of each contacting body.
The contact of a rough elastic sphere with a rigid flat was investigated by Greenwood and Tripp [34].
The dimensionless load is given by

2F (4.62)
F=
s sRʹ EC

-
For low loads (F less than about 2), the pressure distribution is broader than it is for a single Hertz con-
tact. For higher loads, the results get closer to those of a perfectly smooth Hertz contact.
The first three assumptions of the GW theory were relaxed by McCool [35] who treated anisotropic
rough surfaces with ellipsoidal asperities. His results were in good agreement with the simpler GW
model. McCool [36] also extended the GW model by using an asymmetric distribution of asperity
heights and by including the effect of a surface coating.
The effect of plasticity has been included in a multi-asperity contact model by Kogut and Etsion [37].
Their solution parallels the GW model except that Hertz contact is replaced by the elastic–plastic finite
element model of a single asperity contacting a rigid surface. The solution in [37] shows that an “elastic–
plastic hardness” (Hep) can be defined by

P = H ep A, H ep = 0.41/y H (4.63)

where the plasticity index ψ is defined slightly differently than in Equation 4.60. However, the two defi-
nitions are equivalent when Poisson’s ratio is 0.3.
The effect of adhesion in multi-asperity contacts can be included by replacing the Hertz contacts of
the GW model with the elastic adhesive contacts from the JKR model [38]. The results depend on an
adhesive parameter that represents the ratio of the standard deviation of asperity heights to the elastic
displacement at which an asperity breaks away from the adhering surface. A modified adhesion param-
eter is defined by

ECs 3/2
q= (4.64)
R1/2w
Contact Mechanics 4-19

When this parameter is small the effect of adhesion dominates whereas when it is large the effect of
roughness dominates. If the asperity contacts are in the DMT region of adhesion then the Maugis
multi-asperity model [39] is applicable, whereas if the asperities are in the Maugis region then either the
Adams et al. [40] or the Morrow et al. [41] models can be used. These two models span the range of small
to large values of the Tabor parameter.

References
1. K.L. Johnson, Contact Mechanics, Cambridge University Press, Cambridge, U.K., 1985.
2. H. Hertz, Über die Berührung fester elastischer Körper (On the contact of elastic solids), Journal
für die reine und angewandte Mathematik, 92, 156–171, 1882.
3. H.R. Thomas and V.A. Hoersch, Stresses due to the pressure of one elastic solid upon another,
Bulletin of the University of Illinois Engineering Experimental Station, 27, Bulletin no. 212, 1930.
4. J.A. Greenwood, Formulas for moderately elliptical Hertzian contacts, Journal of Tribology, 107,
501–504, 1985.
5. E. Stuermann (Shtaerman, I., Ya), On Hertz theory of local deformation of compressed bodies,
Comptes Rendus (Doklady) de l’Académie des Sciences de l’URSS, 25, 359–361, 1939.
6. C. Catteneo, Sul Contatto di due Corpi Elastici: Distribuzione Locale Delgi Sforzi, Rendiconti
dell’Academia Nazionale dei Lincei, 27, 342–348, 434–436, 474–478, 1938.
7. R.D. Mindlin, Compliance of elastic bodies in contact, Journal of Applied Mechanics, 16, 259–268,
1949.
8. W.R. Chang, I. Etsion, and D.B. Bogy, An elastic–plastic model for the contact of rough surfaces,
Journal of Tribology, 109, 257–263, 1987.
9. D. Tabor, Hardness of Metals, Clarendon Press, Oxford, U.K., 1951.
10. L. Kogut and K. Komvopoulos, Analysis of the spherical indentation cycle for elastic-perfectly plastic
solids, Journal of Materials Research, 19, 3641–3653, 2004.
11. L. Kogut and I. Etsion, Elastic–plastic contact analysis of a sphere and a rigid flat, ASME Journal of
Applied Mechanics, 69, 657–662, 2002.
12. R.L. Jackson and I. Green, A finite element study of elasto-plastic hemispherical contact against a
rigid flat, Journal of Tribology, 127, 343–354, 2005.
13. I. Etsion, Y. Kligerman, and Y. Kadin, Unloading of an elastic–plastic loaded spherical contact,
International Journal of Solids and Structures, 42, 3716–3729, 2005.
14. J.W. Hutchinson, Plasticity at the micron scale, International Journal of Solids and Structures, 37,
225–238, 2000.
15. D. Maugis, Contact, Adhesion and Rupture of Elastic Solids, Springer-Verlag, Berlin, Germany, 2000.
16. R.S. Bradley, The cohesive force between solid surfaces and the surface energy of solids, Philosophical
Magazine, 13, 853–862, 1932.
17. K.L. Johnson, K. Kendall, and A.D. Roberts, Surface energy and the contact of elastic solids,
Proceedings of the Royal Society of London, A324, 301–313, 1971.
18. B.V. Derjaguin, V.M. Muller, and Y.P. Toporov, Effect of contact deformations on the adhesion of
particles, Journal of Colloid and Interface Science, 53, 314–326, 1975.
19. V.M. Muller, B.V. Derjaguin, and Y.P. Toporov, On two methods of calculation of the force of sticking
of an elastic sphere to a rigid plane, Colloids and Surfaces, 7, 251–259, 1983.
20. D. Maugis, Adhesion of spheres: The JKR–DMT transition using a Dugdale model, Journal of Colloid
and Interface Science, 150, 243–269, 1992.
21. D.S. Dugdale, Yielding in steel sheets containing slits, Journal of the Mechanics and Physics of Solids,
8, 100–104, 1960.
22. K.L. Johnson and J.A. Greenwood, An adhesion map for the contact of elastic spheres, Journal of
Colloid and Interface Science, 192, 326–333, 1997.
4-20 Theory and Practice of Lubrication and Tribology

23. S.D. Mesarovic and K.L. Johnson, Adhesive contact of elastic–plastic spheres, Journal of the Mechanics
and Physics of Solids, 48, 2009–2033, 2000.
24. Y. Du, L. Chen, N.E. McGruer, G.G. Adams, and I. Etsion, A finite element model of loading and
unloading of an asperity contact with adhesion and plasticity, Journal of Colloid Interface Science,
312, 522–528, 2007.
25. Y. Du, G.G. Adams, N.E. McGruer, and I. Etsion, A parameter study of separation modes in adhering
microcontacts, Journal of Applied Physics, 103, 044902, 2008.
26. D. Maugis and H.M. Pollock, Surface forces, deformation and adherence at metal microcontacts,
Acta Metallurgica, 32, 1323–1334, 1984.
27. J. Song and D. Srolovitz, Adhesion effects in material transfer in mechanical contacts, Acta Materialia,
54, 5305–5312, 2006.
28. L. Chen, N.E. McGruer, G.G. Adams, and Y. Du, Separation modes in microcontacts identified by the
rate-dependence of the pull-off force, Applied Physics Letters, 93, 053503, 2008.
29. L. Kogut and I. Etsion, Adhesion in elastic–plastic spherical microcontact, Journal of Colloid and
Interface Science, 261, 372–378, 2003.
30. Y. Kadin, Y. Kligerman, and I. Etsion, Loading–unloading of an elastic–plastic adhesive spherical
microcontact, Journal of Colloid and Interface Science, 321, 242–250, 2008.
31. Y. Kadin, Y. Kligerman, and I. Etsion, Cyclic loading of an adhesive contact, Journal of Applied Physics,
104, 073522, 2008.
32. J.A. Greenwood and J.B.P. Williamson, Contact of nominally flat surfaces, Proceedings of the Royal
Society of London, A295, 300–319, 1966.
33. J.A. Greenwood and J.H. Tripp, The contact of two nominally flat rough surfaces, Proceedings of the
Institution of Mechanical Engineers, 185, 625–633, 1970.
34. J.A. Greenwood and J.H. Tripp, The elastic contact of rough spheres, Journal of Applied Mechanics,
34, 153–159, 1967.
35. J.I. McCool, Limits of applicability of elastic contact models of rough surfaces, Wear, 86, 105–118,
1983.
36. J.I. McCool, Extending the capability of the Greenwood Williamson microcontact model, Journal of
Tribology, 122, 496–502, 2000.
37. L. Kogut and I. Etsion, A finite element based elastic-plastic model for the contact of rough surfaces,
Tribology Transactions, 46, 383–390, 2003.
38. K.N.G. Fuller and D. Tabor, The effect of surface roughness on the adhesion of elastic solids,
Proceedings of the Royal Society of London, A345, 327–342, 1975.
39. D. Maugis, On the contact and adhesion of rough surfaces, Journal of Adhesion Science and
Technology, 10, 161–175, 1996.
40. G.G. Adams, S. Müftü, and N. Mohd Azhar, A scale-dependent model for multi-asperity model for
contact and friction, Journal of Tribology, 125, 700–708, 2003.
41. C. Morrow, M. Lovell, and X. Ning, A JKR–DMT transition solution for adhesive rough surface con-
tact, Journal of Physics D: Applied Physics, 36, 534–540, 2003.
5
Friction
5.1 Introduction....................................................................................... 5-1
5.2 Historical Studies of Friction........................................................... 5-2
5.3 Static and Kinetic Friction Coefficients......................................... 5-2
5.4 Frictional Transients and Instabilities........................................... 5-3
5.5 Friction and Energy Dissipation.....................................................5-4
5.6 Frictional Heating.............................................................................. 5-5
5.7 Friction and Interfacial Shear Strength......................................... 5-7
5.8 Localized Material Displacement and Its Effect on Friction
Modeling.............................................................................................5-8
Nanoscale Solid Friction  •  Nanoscale Liquid Friction  •  Microscale
Friction  •  Application of Multi-Scale Friction Models in
Engineering
5.9 Material-Specific Friction............................................................... 5-10
Peter J. Blau 5.10 Friction Testing................................................................................ 5-11
Oak Ridge National 5.11 Summary........................................................................................... 5-13
Laboratory References..................................................................................................... 5-13

5.1  Introduction
Friction, the resistance to relative motion between a solid body and another solid body or substance,
is so common an everyday experience that it has been the source of philosophical contemplation and
experimental investigation throughout history. Friction broadly encompasses not only fluid friction in
pipes, aerodynamic friction, and the friction between dry or wet sliding surfaces in mechanical parts
like seals, brakes, bearings, gears, but also friction in natural phenomena, ranging from human joints
to tectonic plates. Because friction takes so many forms, no universal, overarching theory or model suc-
cessfully explains it.
A common quantity used to describe the magnitude of friction between sliding solid bodies is called
the friction coefficient (alternatively, the “coefficient of friction”). It is the dimensionless ratio of two
mutually perpendicular forces: a force that resists relative motion and a normal force that acts per-
pendicular to the interface. A popular notion persists that the friction coefficient is a single-valued,
fundamental property of a given material pair. That presumption fails in the face of experimental
evidence. The degree to which variables like normal force, relative velocity, directional changes in rela-
tive motion, contact surface roughness, temperature, humidity, magnetic fields, and material proper-
ties affect frictional response is tribosystem dependent. This chapter considers the causes of friction,
the concepts that researchers have used to explain friction, and the control of friction in engineering
systems.

5-1
5-2 Theory and Practice of Lubrication and Tribology

5.2  Historical Studies of Friction


Studies of friction, and the use of lubrication to solve friction-related problems, date back thousands
of years. In fact, the invention of rollers and wheels was in effect an engineering solutions to friction
problems. Dowson eloquently points out that the history of transportation and industrial growth are
strongly tied to developments in tribology [1]. Historical reviews of friction research [1,2] include the
work of notables like daVinci, Amontons, Morin, Coulomb, Euler, Martins, Stribeck, Hertz, and many
others. Before advanced microscopes and profiling systems were invented, modelers envisioned solid
surfaces as arrays of spheres or rigid, interpenetrating features that climbed up and over one another.
Work by Desaguliers, Tomlinson, Deraygin, and others considered friction to be an atomic level–based
phenomenon that was related to adhesion.
In the 1900s, new instruments were developed to allow surfaces to be imaged, measured, and mapped
down to the nanometer scale. First-principles friction models became more complicated as the effects of
more variables and the subtleties of surface structure were included. For example, the following affected
the evolution of friction models:
• The notion that asperities are tiny high spots on surfaces
• The scale and geometry (roughness) of surface features
• The true contact area, vis-à-vis the apparent (nominal) contact area
• The concept of junction growth when high spots are increasingly loaded
• The role of adhesive forces and surface tension on friction
• The role of atomic-level electron interactions and lattice phonons
• Whether deformation and/or wear occurs along with friction
• The effects of cleanliness and adsorbed species
• The role of interfacial films, particles, liquids, or other materials
• The effects of applied load (contact pressure) and sliding velocity
• The role of frictional heating, vibrations, and system dynamics
No known friction model explicitly handles the effects of all of the aforementioned factors. Rather, mod-
elers apply simplifying assumptions and select variables that they believe relate to the friction problem
they have defined.

5.3  Static and Kinetic Friction Coefficients


Two dimensionless ratios, called the static (μs) and kinetic (dynamic) friction coefficient (μ), are used
to connote the magnitude of friction in machines and in natural systems. They are defined as follows:

Fs (5.1)
ms =
P

F (5.2)
m=
P

where
Fs is the (static) resisting force that must be overcome to initiate relative motion
F is the resisting force that must be overcome to sustain relative motion
P is the normal force, perpendicular to the interface in which relative motion occurs

In this chapter, if there is no subscript given, μ will refer to kinetic friction, and if it contains the
subscript “s,” it refers to static friction. Other symbols and abbreviations used in the literature for the
friction coefficient include “CoF,” “FC,” and “ f.”
Friction 5-3

Liquid films

Hydrodynamic Boundary-lubricated Non-Lubricated


thick film Mixed film high adhesion, plowing
µ ~ 0.001 mild contact

µ ~ 0.05

µ ~ 0.10

µ ~ 0.50

µ ~ 1.0

µ > 2.0
Ultralow friction films

Ice, slightly below 0°C

Sputtered, low friction coatings

Graphite-lubricated metals

Teflon on metals

Oxidized metals, low load

Self-mated aluminum

Clean metals in high vacuum


Steel on steel

Brass on steel
various crystallographic
Diamond on diamond–
orientations

FIGURE 5.1  General attributes of time-dependent frictional behavior.

Tables of friction coefficients for various material combinations and states of lubrication (such as
“greasy steel”) abound. Such tabulations are limited [3], but they are still helpful for appreciating the
range of friction coefficients. At the risk of oversimplifying, a schematic representation of typical ranges
of kinetic friction coefficients is shown in Figure 5.1. The shaded box at the top of the figure depicts lubri-
cation conditions. These range from liquid films to clean, dry contacts. Approximate kinetic friction
coefficients corresponding to those situations are represented in the center box. The box at the bottom
of the diagram includes a few examples of kinetic friction coefficients for nonliquid-lubricated materials
(the inclusion of ice friction is an exception). Typically (but not exclusively), the following magnitudes
of the sliding friction coefficient are observed:
• For well-lubricated rolling element bearings, μ ∼ 10−3
• For boundary-lubricated surfaces, μ ∼ 0.07–0.12
• For oxidized metal surfaces, μ ∼ 0.15–0.4
• For dry metal surfaces, μ ∼ 0.4–0.9
• For aluminum alloys, and very clean metals in a vacuum, μ > 1.0
Precise values of μ cannot be given because variables like contact pressure, relative humidity, sliding
velocity, and sliding distance can affect the friction coefficient for a given material pair (see also, the
discussion by Ludema [4]). Friction coefficients for dry sliding solids are typically not reported to more
than two decimal places in recognition of their variability. In fact, as the following section discusses,
associating a characteristic range of friction coefficients with a tribosystem may be more useful than
specifying a single value.

5.4  Frictional Transients and Instabilities


Maintaining a specified level or range of friction—neither too high, nor too low—is an important con-
cern in certain types of machine and component design. Consider, for example, the friction of brakes
on public buses: too low, and the bus will not stop quickly enough to avoid a collision; too high, and
the passengers will be thrown forward and possibly injured. In internal combustion engines, friction
is generally kept as low as possible to reduce mechanical losses and to improve fuel economy. In figure
5-4 Theory and Practice of Lubrication and Tribology

skating, ice hockey, and speed skating, the temperature of the ice rink is adjusted to provide users with
the desired surface characteristics.
Real machines and their internal load-bearing surfaces display a range of friction coefficients as they
run-in, age, and wear out. In design, it may be as important to understand the variations in friction
experienced by an operating machine or component as it is to know the nominal friction coefficient.
Wear particles can collect in systems. Lubricants can age, changing composition, molecular weight,
and viscosity over time. In applications like friction brakes and clutches, it is important to understand
the characteristics and causes of time-dependent frictional transients. Practical engineering friction
models for designers should be capable of predicting both steady state (if any), as well as time-dependent
changes in interfacial processes.
The diverse potential causes for frictional instabilities range from low mechanical stiffness to exter-
nal factors like the incidental ingestion of particles from other parts of the machine. In non-lubricated
systems, instabilities can arise from adhesive transfer or wear-related phenomena like the formation
and detachment of built-up edges on one of the sliding partners. Records of frictional behavior reflect
phenomena like running-in, coating failure, and end-of-life (wear out) [2,5–7]. The attributes of fric-
tion versus time or sliding duration can be classified in three ways: (1) a general trend (curve shape), (2)
time to reach prominent features in the profile, and (3) an instantaneous variability level. The details of
such “friction signatures” are highly repeatable in some tribosystems [2,8]. Complex changes in friction
reflect the net effect of several conjoint processes that do not necessarily require the same amount of
time to achieve steady state. It has also been pointed out that different combinations of processes can
produce similar-appearing friction versus time responses [2,5]. Therefore, a tribosystem-specific analy-
sis is required to identify the mechanistic causes of complex friction–time behavior.
A special case of frictional instability, in which surfaces periodically adhere and release, is known
as stick–slip. This “relaxation–oscillation phenomenon” involves the buildup of elastic energy (as in a
spring) as the tangential force is applied. The tangential force increases until it exceeds the interfa-
cial resisting force, at which instant slip occurs. The velocity of relative motion rises rapidly and then
decelerates until relative motion ceases and the process repeats. One criterion for stick–slip in a given
tribosystem is therefore that the friction coefficient in that tribosystem decreases as the sliding velocity
increases. Classical examples of stick–slip are the squeal of chalk on a slate board, the squeal of brakes
during a stop, the sound of a fork scratching a dinner plate, and earthquakes arising from transform
faults (as opposed to harmonic tremors from volcanism).
Other chapters in this book focus on lubricants and their properties; however, frictional interfaces in
practical systems are rarely perfectly clean and dry. Therefore, a discussion of friction should consider
not only idealized, dry sliding but also the influence of solid and liquid films. Solid interfacial films can
consist of unintended contaminants, oxides (on metals), interposed low shear strength materials (such
as graphite, molybdenum disulfide, and other solid lubricants), and multiphase layers of “tribomateri-
als” that are formed in situ as a result of wear. There is a common notion that static friction (sometimes
called stiction) is always greater than kinetic friction. That is not the case when oily or contaminated
surfaces are involved. However, before considering such details, a general consideration of energy gen-
eration and dissipation during friction is helpful.

5.5  Friction and Energy Dissipation


Frictional phenomena can be described in terms of the transformation and dissipation of energy by
nonconservative processes. The manner of that dissipation is considered in models that span a range of
size scales. Simply stated, however, the non-recoverable frictional work (Wf) comprises the product of
the friction force (F) and the sliding distance (d) over which it acts. In terms of the kinetic friction coef-
ficient μ and normal force P,

W f = F ⋅ d = mPd (5.3)
Friction 5-5

TABLE 5.1  PV Limits for Some Polymeric Bearing Materialsa


PV Limit (v = 0.05 m/s) PV Limit (v = 0.51 m/s) PV Limit (v = 5.1 m/s)
Material (kN-m/m2-s) (kN-m/m2-s) (kN-m/m2-s)
Acrylonitrile butadiene styrene (ABS) 631 140 70
Acetal 140 123 88
Nylon 6/6 (polyimide) 105 88 88
Nylon 6/6, 30% glass, 15% PTFE 614 701 526
Polycarbonate, 30% glass, 15% PTFE 701 1227 526
Polytetrafluoroethylene, glass fiber filled 351 351 351
a Converted to metric units from tables in Reference 9.

The product of the nominal contact pressure (p) (that is, a normal load P acting on an area A), and the
sliding velocity (v) has been used as a measure of the severity of sliding conditions. The frictional energy
dissipation over a given bearing area (A) is therefore proportional to the product of friction force and
the velocity; thus,

mPd
Ef = = mPv (5.4)
t

where t represents total sliding time. The so-called “pv limits” (sometimes capitalized PV) have been
used in selecting bushing materials for specific applications. They are purported to represent the maxi-
mum allowable operating conditions for an engineering material (usually polymer-based materials).
Table 5.1 presents some typical PV limits from [9]. As the data show, it cannot be assumed that the PV
limit for polymers and composites always decreases with sliding speed. In fact, a more physically sen-
sible way to use PV is not to think of it in terms of the nominal bearing pressure p, but rather in terms
of the friction force that acts along the surface to shear the material and to generate heat and wear.
Therefore, the instantaneous PV can change during machine operation because the friction coefficient
changes with time, speed, and temperature. Data that simulate specific operating conditions are needed
for material selection, but in their absence, tables of PV data are a convenient place to start.
Once it is produced, energy from frictional work must be dissipated. Heat generation is one of the
ways this occurs.

5.6  Frictional Heating


The bulk surface temperature that arises from frictional heating during metal sliding is in general
related to Fv. In fact, the majority of models for frictional heating somewhere contain the Fv product
(or μPv), but they vary in the manner in which the thermophysical properties of the materials and the
contact geometry are introduced.
Frictional heating models commonly assume that all of the frictional work is converted to heat,
but strictly speaking, that is not correct. Some energy translates into sound and structural vibrations,
some deforms the contacting materials, some creates new surfaces (by wear or fracture), and some
is stored in the material as sliding-induced defects (e.g., work hardening of metals and alloys). The
proportion of frictional work that expresses itself as heat depends on the rate of energy production
and the available dissipation mechanisms. That is simply a consequence of the well-known energy
conservation law:

Input = Output + Accumulation (5.5)



The reason why certain pairs of materials have similar friction coefficients, but much different rates
of wear is likely related to differences in their manner of energy dissipation [10]. In well-lubricated
5-6 Theory and Practice of Lubrication and Tribology

tribosystems, not only is there less energy available to do work (μ is small), but the lubricant itself can
help to carry away the heat that is generated.
Three types of temperatures are associated with frictional heating. The ambient temperature is that
of the surroundings. The bulk temperature is the average temperature that is developed in the friction-
ally heated zone as a result of the spikes in temperature that occur at individual asperity contacts. The
much higher temperature that occurs at those localized asperity contacts is called the flash temperature,
a concept attributed to the work of Harmen Blok in the late 1930s [11]. A more complete discussion of
frictional heating calculations may be found in References 2 (pp. 171–178) and 12 (pp. 39–44). In most
frictional heating models, like those described in [12], it is assumed that the heat is generated within the
boundaries of a specific interfacial area. However, physically, the transformation must take within a vol-
ume of material, and not a 2-D area. This approach was explored in a 1977 paper by Malkin and Marmur
[13]. Furthermore, the scale of tribosystems has implications for frictional heat dissipation. In nanoscale
models, energy is completely dissipated within the first few layers of interatomic atomic bonds (by lat-
tice phonons and electronic interactions), but in micro- or mesoscale friction models, much larger scale
processes like plowing and cutting (fracture) play a role in dissipation.
As discussed elsewhere [2, p. 176], there are at least six ways in which frictional heating models differ:
1. The assumed shape of the heat source on the surface
2. The contact geometry (sphere-on-plane, flat-on-flat, etc.)
3. The manner in which the heat flow is partitioned
4. The form in which thermophysical properties (like thermal conductivity and thermal diffusivity)
are incorporated
5. Whether the heat source is planar or in a volume (as in [13])
6. The use of velocity criteria (e.g., the Peclet number [12]) to determine which mathematical form of
the model should be used
Frictional heating is important in the performance of friction brakes on automobiles, trucks, railroad
cars, and aircraft. “Fade” is a term that describes how the braking effectiveness falls precipitously when
brake surfaces overheat. Brake linings (known as friction materials) that are designed for severe use,
like those on buses, large trucks, and vehicles that operate on hilly terrain, are specially formulated to
resist fade. An experimentally determined quantity, called the friction heating parameter (Θ), has been
used [2, p. 351] to reflect the efficiency by which frictional heat in a sliding material pair is converted
into a surface temperature rise under otherwise similar conditions. Table 5.2 lists values measured for a

TABLE 5.2  Frictional Heating Parameter for Various Sliding


Combinations in a Subscale Brake Materials Testing System
Θ, Relative to
Disk Material Slider Material Θ(°C/J) Combination 1
1 Cast iron Jurid 539 (TM)a 1.18 × 10−3 1.00
2 Cast iron Armada AR4 (TM)a 1.04 × 10−3 0.88
3 Cast iron Carbon-metallic 1.23 × 10−3 1.04
commercial lining
4 SiC-ceramic Carbon-metallic 3.59 × 10−3 3.04
composite commercial lining
5 Thermally sprayed Highly metallic 6.3 × 10−3 5.34
Ti-6Al-4V proprietary lining
6 Titanium alloy Jurid 539 (TM) 8.33 × 10−3 7.06
Ti-6Al-4V
Source: Data from Blau, P.J., Friction Science and Technology—From Concepts to
Applications, Taylor & Francis/CRC Press, Boca Raton, FL, 2008, p. 351.
a A commercial, phenolic-based semimetallic brake pad material.
Friction 5-7

series of sliding experiments on common and experimental brake materials in a laboratory-scale, brake
materials testing apparatus. The higher the value of Θ, the more frictional energy has been converted
into heating the surface of the disk specimen. Values for typical truck brake material combinations are
about 1.1 × 10−3°C/J, but when the disk specimen was composed of a titanium alloy with a much lower
thermal conductivity than cast iron, Θ rose to over 8.0 × 10−3°C/J.

5.7  Friction and Interfacial Shear Strength


One can address sliding friction in terms of an interfacial contact area (Ao), a resisting force parallel to
the surface (F), and the shear strength (τ) of the material that must be overcome in order for relative
motion to occur. Thus,

F = tAo (5.6)
From Equation 5.2, the friction coefficient can be approximated as the ratio of the shear stress to the
contact pressure (p):

Ao t
m=t = (5.7)
P p

However, τ is not always independent of contact pressure and can be written more generally as

t = to + a p (5.8)

where α is the pressure coefficient. Table 5.3 provides measurements of α and τo that were measured for
pure metals tested as possible lubricating films [14]. Combining Equations 5.7 and 5.8, and recognizing
that when plastic flow occurs in a solid, p = (P/A) = hardness (H) by definition,

to + ap to
m= = +a (5.9)
p H

This relationship, discussed by Briscoe and Stolarski [15], accounts for the observed pressure depen-
dence of the friction coefficient.
The observed friction of a buried interface is the net effect of the shear strength(s) of the substances
that comprise the region in which the difference in relative velocity between one body and another is
accommodated. In studies of pure, clean metals in vacuum, friction coefficients higher than 20 have
been reported because the shear strength of the original interface, in which adhesion takes place,
becomes higher than the shear strength of the weaker member of the sliding pair. Thus, μ reflects the
shear strength of a bulk material rather than the strength of an interfacial bond.

TABLE 5.3  Pressure Coefficients for


Solid Lubricating Materials
Metal α τo (MPa) μ (p = 1 MPa)
Cu 0.049 107.61 0.160
Ag 0.036 109.04 0.144
Au 0.029 91.82 0.123
Al 0.035 47.56 0.082
Sn 0.012 12.30 0.024
In 0.006 5.65 0.011
Source: Bednar, M.S. et al., Lubr. Engr., 49,
741, 1993.
5-8 Theory and Practice of Lubrication and Tribology

If sliding friction is seen as a consequence of the interfacial shear strength (τ), and if the friction force
must exceed the product (τA) for relative motion to take place, then some permanent displacement of
material must occur. Depending on the situation, displacement can be subtle, on the micro-asperity
level, or it can be gross, and easily visible to the naked eye. If lubricant films are present, displacement
may occur entirely within them, and there will be no trace of damage on the bearing surfaces. The next
section describes the role of material displacement in sliding friction, and how the localization of mate-
rial displacement has affected the development of friction models.

5.8 Localized Material Displacement and


Its Effect on Friction Modeling
The relative motion between two bodies occurs within a certain volume that can be called the local
material displacement zone (LMDZ). Understanding the LMDZ can help when selecting lubricants and
surface treatment approaches for friction control. The material being displaced can be a solid, liquid, or
gas. Dissipation of frictional energy can take place both within the LMDZ and beyond it, through the
transmission of heat and vibrations to the surroundings. Dissipative processes, rather than the displace-
ment of material, were described at a conference that focused on scale effects in friction [16]. Another
notable approach, advanced by Bethier et al. [17], articulates as many as 20 mechanistic possibilities by
which materials in or adjacent to an interface between sliding solids can accommodate relative displace-
ment. Physically, the region in which displacement occurs can range in size from the subnanometer
gap between atomically clean surfaces to centimeter-thick volumes of material, like the cataclastic flow
zones of highly sheared geological formations [18]. It can even be as large as an earthquake zone that is
kilometers in extent.
Modelers tend to focus on different size scales of frictional phenomena. For example, typical lubri-
cant films in rolling or sliding bearings can range from a few nanometers to a few micrometers thick.
Asperities in classical friction models, applied to engineering surfaces, are approximately 0.02–10 μm
tall. Employing the concept of a LMDZ recognizes that different models are needed to describe fric-
tional phenomena that occur at different size scales.

5.8.1  Nanoscale Solid Friction


When extremely smooth and defect-free crystalline solid surfaces slide, then the tangential force can
take on a periodic nature; especially when the contact point is only a few atoms across and can detect
such nanoscale force variations. Experiments on atomic-scale friction force periodicity were first
reported in the late 1980s [19]. Work by McClelland and Glosli [20] and others renewed interest in fric-
tion within the physics community. Studies of the forces between minute rafts of molecules perched
on oscillating substrates took friction studies in a new direction [21], and a 1998 book by Persson [22]
further stimulated the tribophysics community to conduct additional research on electronic effects,
magnetic effects, and phonon effects on nanoscale interfaces.
The definition of the friction coefficient as the ratio between tangential and normal forces between
bodies can be applied to pristine surfaces on the atomic scale, but one cannot confidently assume that
friction coefficients measured at such fine scales, and under highly controlled experimental conditions,
are applicable to macroscale engineering situations.

5.8.2  Nanoscale Liquid Friction


When smooth-surfaced moving solids are separated by monolayers of “fluid” films, their frictional
characteristics can be quite different than when thicker films are involved. Confined 1 or 2 nm “fluid”
films can assume a time- and shear rate–dependent epitaxial relationship with the atoms on a solid sur-
face, sometimes behaving like liquids and other times like solids. The phenomenon of shear thinning
Friction 5-9

[23] describes a rise in shear strength when thin fluid films begin to behave as a quasi-solid. Likewise,
the orientation (steric factors) of molecular layers can be designed to control friction in smooth, narrow
interfaces [24]. Materials in monolayer form have different properties than those in bulk form, so the
frictional behavior in such cases is not well predicted by thick-film models. Fine-scale LMDZ phenom-
ena are of current interest to those developing lubricants for the computer hard disk industry and for
microelectromechanical devices, and a large body of specialized literature, well beyond the scope of this
chapter, has evolved in those areas (e.g., [25,26]).

5.8.3  Microscale Friction


Increasing the scale of the LMDZ to the micrometer level brings us to the realm of classical friction stud-
ies conducted in the mid-1900s. The ground-breaking work of F. P. Bowden and D. Tabor of Cambridge
University led friction researchers to consider two basic contributions to metallic friction: (i) growth and
shearing of metallic junctions and (ii) plowing in which hard asperities dig through softer surfaces. This
classical view of solid friction incorporates the both molecular-scale interactions (adhesion) and asperity
deformation on the scale of micrometers. Interposed liquids can reduce the adhesive contribution, and
likewise, increasing the hardness and roughness of one of the bodies relative to the other can increase
its plowing tendency. These kinds of arguments make physical sense, and widen the applicability of that
relatively simple, two-contribution approach. A discussion of these mechanisms is contained in a 1950
book that many investigators still consider the “bible” of friction and boundary lubrication [27].
Researchers such as Buckley and Johnson [28], Steijn [29], and Feng and Field [30] explored the effects
of grain boundaries and crystallographic orientation of metals, ceramics, and diamond on frictional
variations. Rigney and Hirth [31] elucidated the formation of frictionally induced, highly deformed
defect layers in metals. Friction was seen to be a consequence of changes below a contact surface not just
an effect of microgeometry.
Wear that results from frictional contact results in the displacement and detachment of material.
The formation and movements of loose wear particles and their agglomerates are called third-body
effects. Depending on interfacial conditions, third bodies can smear and adhere to counterfaces to form
transfer films. Such films are a key factor in the performance of friction materials in vehicle brakes
[2 (p. 352),32]. In the mid-1900s, ground-breaking work by Rabinowicz and Tabor [33] applied radioac-
tive tracers to study transfer film formation, verifying their role in friction. A body of work in the former
Soviet Union (e.g., [34]) attempted to design surfaces in order to optimize the tendency of some materi-
als to form transfer layers. Collectively, that approach was called “selective transfer.”
The third-body effect has been articulated by French researchers such as Godet [35,36]. The pre-
viously noted concept of velocity accommodation, proposed in the mid-1980s [17], described the
manner in which bulk materials and interfacial particles between them can share the task of local-
ized material displacement and contribute to the measured resistance to sliding. In fact, the flow
of triboparticulates (third bodies) has been the subject of substantial theoretical and experimental
work that led to a new understanding of the use of powders in lubrication. For example, Heshmat
developed a quasi-hydrodynamic theory of powder lubrication [37]. Fillot et al. [38] modeled the role
of adhesion in particle flows of various thicknesses. Understanding the physical processes of transfer
has been facilitated by fascinating in situ observations like those described by Scharf and Singer [39].
From the perspective of tribology transitions, the role of transfer processes in defining the shapes and
variability of time-dependent friction transitions has been discussed at some length [10, pp. 271–295].

5.8.4  Application of Multi-Scale Friction Models in Engineering


The advent of new instrumentation, molecular dynamics modeling, and high-speed computing has
enabled remarkable insights into nano- and microscale friction, but they have also led to a growing
complexity and compartmentalization of friction research. Therefore, those seeking guidance for
5-10 Theory and Practice of Lubrication and Tribology

Smooth layers Single asperity Oscillating raft Thin fluid film

΄΄Clean΄΄ asperities Boundary films Roughness Microstructure

Parallel Multiple slip within


Thick fluid films ΄΄Third bodies΄΄ a large volume
displacements

FIGURE 5.2  Schematic depiction of the various approaches to friction modeling, based on the degree of localized
material displacement during relative movement.

low-friction-bearing materials and lubricant selection are confronted with many choices. In the absence
of experimental data for the precise application of interest, the applications engineer may seek a viable
friction model, conduct experiments, rely on product vendors, or contact a consultant. The first step in
material selection is to establish the scale of the problem and to identify candidate materials and lubri-
cants based on a set of application-specific criteria [16].
Figure 5.2 summarizes 12 of the LMDZ concepts described in prior sections in the context of friction
modeling. They range from idealized arrays of atoms, at the upper left, to delocalized, multiple slip loca-
tions in a heterogeneous fault zone, at the lower right. Like the specialized tools in a machinist’s toolbox,
friction models suitable for some situations may be inappropriate for others. In addition, variables like
normal force (contact pressure), sliding speed, surface roughness, surface texture, and environment are
not independent. For example, if the normal force is increased, the contact points on the surface will
expand either elastically or elastoplastically, and wear may occur to roughen the surface. Likewise, if
the sliding speed increases, the interfacial temperature rises. It is difficult to change only one friction-
affecting variable without changing others as well.

5.9  Material-Specific Friction


Having defined the tribosystem, material properties, lubricant properties, surface finish, and the regime
of lubrication come into play. Metallically bonded metals, ionically bonded ceramics, and covalently
bonded materials behave differently when slid self-mated or against a dissimilar counterface. They have
different surface states and different deformation and fracture characteristics. Polymeric materials, like
rubber whose deformation is highly rate dependent, will react differently to frictional conditions than
more rigid materials. Likewise, the friction of some materials is humidity dependent. This is especially
true for static friction at light loads in which capillarity forces and adhesion can change with time
between loading and the onset of relative motion [40]. Therefore, attention to surface states and cleanli-
ness is important for applications like electrical contacts and computer disk drives.
Material-specific discussions and reviews of friction are listed in Table 5.4. When selecting materi-
als, the regime of lubrication and the corresponding film thickness-to-composite-roughness ratio are
Friction 5-11

TABLE 5.4  Models and Investigations of Material-Specific Friction Processes


[References/Page
Material Frictional Attributes Numbers, If Applicable]
Metals and alloys Adhesion, elastic and plastic deformation, metallic contact (true and [27]
apparent contact area), junction growth, plowing, and shear. Role of [2, pp. 184–188]
oxides and films [41, pp. 31–49]
Ceramics At low loads, friction tends to be controlled by adsorbed species. Low [42, pp. 2–49]
humidity or elevated temperatures that cause desorption of films tend to
promote higher friction and wear. At high temperatures, debris or glassy
layers can reduce friction to levels comparable to those at room
temperature
Ceramic Affected by the relative percentage of lubricating phases, ability of [42, pp. 163–197,
composites lubricant to spread (“butter”) the surface, and the ability of the matrix to 199–223]
wear so to supply new lubricating phases
Polymers Models involve adhesion, elasticity, and hysteresis; modeled by [43, pp. 67–201]
arrangements of springs and dashpots [2, pp. 192–199]
Elastomers Like polymers, elastomer friction models involve adhesion, elasticity, and [44, pp. 61–85]
hysteresis, with arrangements of springs and dashpots. Draping over [15, pp. 46–48]
asperities is a common argument to explain instabilities and velocity
effects
Granular masses, Depend on both internal interparticle contacts and particle–wall [37]
powders interactions. Principle of dilatancy applies during shear. Strongly affected [45]
by pressure and shear rate. Can behave as a viscous solid under some
circumstances
Ice Strongly affected by temperature. Controlled by a varying combination of [2, pp. 204–209]
factors including frictional heating, presence of quasi-fluid films (water),
and pressure-induced melting
Solid lubricants Layer-lattice species, like graphite, boric acid, and molybdenum disulfide, [46]
depend on properly oriented planes or textures with low shear stress to [47]
facilitate friction reduction [48, pp. 269–290]
Diamond-based Can exhibit the lowest friction of any solid, but that depends on [49, pp. 667–675]
materials morphology, surface chemistry (type of atomic bonding), and the
surrounding atmosphere
Wood Strong effects of texture and fiber orientation; also humidity, and hardness [44, pp. 101–103]
Yarns and fibers Strong effects of texture and tightness of weave, humidity, state of wear, [44, pp. 103–104]
and directionality of motion
Rocks, minerals, Brittle nature, development of frictional structures, effects of pressure [18, pp. 44–96]
and geomaterials

important. Material properties are more important in dry, non-lubricated, or boundary-lubricated con-
ditions than in thick-film conditions in which the lubricant properties play a dominant role. In fact,
it may not matter what material is used if the surfaces retain their shape and remain separated by the
lubricant throughout their operating lives. However, bearing components may have to start and stop,
and at such times thicker films can break down to allow solid contact, then the selection of bearing
materials becomes important.

5.10  Friction Testing


Historically, the motivation for most friction testing is to select material combinations and/or lubri-
cants to alleviate practical problems. The study of friction for its own sake, using specialized devices
that isolate materials from complicating influences, is a subset of friction studies, but a growing one in
5-12 Theory and Practice of Lubrication and Tribology

academia. Unfortunately, not all friction measurement devices are designed correctly, and the descrip-
tions of the methods used to obtain data are incomplete, limiting their usefulness. For example, few
investigators who report friction coefficients also quantify the stiffness and damping capacity of their
apparatus [50,51]. Most researchers also fail to report the relative humidity of the laboratory where
friction tests are performed. This omission can be key, in light of considerable research on the effects of
humidity on friction (e.g., [2, pp. 302–304]).
Friction measurement methods are described in a number of reviews (e.g., [2,4,52–54]. Like model-
ing, such testing also involves issues of scale. German standard DIN 50322 [55] divides wear testing into
six levels, and the following can also be applied to applied friction testing:
1. Machinery field tests (e.g., a truck in normal use)
2. Machinery bench tests (e.g., a full-sized truck on a test stand)
3. System-level bench tests (e.g., the truck’s transmission on a test stand)
4. Component-level bench tests (e.g., a subscale transmission on a test stand)
5. Model tests (e.g., gears on a test stand)
6. Laboratory tests (e.g., small coupon tests to simulate contact geometry)
The degree to which the responses of laboratory tribosystems, schematically shown in Figure 5.3, trans-
lates into the behavior of engineering applications determines whether laboratory data can effectively be
used to select materials or lubricants for specific applications. Each of the three laboratory-scale tribo-
systems in Figure 5.3 has its own set of characteristics, whether the apparatus happens to be a hardness
tester, a viscometer, or a tribometer. The degree to which a given material or lubricant responds to those
systems needs to be sufficiently similar to what happens in an engineering tribosystem for laboratory
data to be of use in applied research. The challenge confronting any applied investigation therefore is to
determine what criteria can be used to determine if the screening tests and applications are “sufficiently
similar.”

Laboratory-scale tribosystems

Tribosystem A: Primary friction test


Motion, contact geometry and alignment, stiffness,
contact pressure, environment, cleanliness, third-
body effects, lubrication regime, duration,
transients, material condition, temperature, heat
flow (DIN 50 322: test category 1–5)

Engineering tribosystem
Tribosystem B: Laboratory test(s) Motion, contact geometry and alignment, stiffness,
for candidate lubricants contact pressure, environment, cleanliness, third-
body effects, lubrication regime, operating
Lubricant screening test: e.g., viscosity, pressure–
spectrum, transients, material condition,
viscosity coefficient, load-carrying capacity, four-
temperature, heat flow
ball, etc.
(DIN 50 322: Test category 6)

Tribosystem C: supporting
laboratory test(s) for the materials
Property measurements: hardness, scratch
hardness, elastic modulus, shear strength, etc.

FIGURE 5.3  Effective correlation of tribosystem characteristics in key when extending laboratory results to
material and lubricant selection in engineering systems.
Friction 5-13

5.11  Summary
Great strides have been made in understanding and controlling friction, but the field is not mature. The
level of fundamental understanding of frictional behavior and the accuracy of models that attempt to
describe it vary widely on a case-by-case basis. In specific cases that involve well-known materials and
lubricants rubbing under controlled conditions, a relatively narrow range of friction coefficients can
be confidently predicted. However, as the complexity of the tribosystem increases, the confidence level
of such predictions diminishes. Thus, mechanical designers of friction-critical machinery, and those
who select materials and lubricants to control friction, tend to rely on experiments or direct operat-
ing experience more than first-principles theories. Synergistic and transient effects, such as friction in
the presence of corrosion and oxidation, friction in the presence of multiple wear modes, and friction
in machines whose operating conditions vary with time, are not easy to incorporate into predictive
models. Lacking a fundamental understanding of such complex situations, statistical approaches are
sometimes used. But that approach is hardly new, considering the development of asperity models for
wear more than a half-century ago by Archard and others.
Some types of machine components, like plain bearings, rolling element bearings, pistons in engines,
and friction brakes, have been widely analyzed and tested. Commercial catalogs provide detailed tech-
nical information and data. Empirical friction models and standardized tests have been developed.
Likewise, friction models have been developed to explain the results of carefully controlled laboratory
experiments conducted over a limited set of test conditions and with pristine materials. As computer
models and surface analysis techniques evolve, more complex frictional conditions are being treated,
and, as the cost of research and development increases, attempts to integrate computer models into the
material selection and design process, in lieu of full-scale experiments, will continue. Yet, in light of the
needs for precise friction control in diverse engineering components, the use of friction testing is likely
to continue as well.

References
1. Dowson, D. 1998. History of Tribology, 2nd edn., John Wiley, New York.
2. Blau, P. J. 2008. Friction Science and Technology—From Concepts to Applications, Taylor & Francis/
CRC Press, Boca Raton, FL.
3. Blau, P. J. 2001. The significance and use of the friction coefficient, Tribol. Int., 34(9), 585–591.
4. Ludema, K. C. 1996. Friction, Wear, and Lubrication—A Textbook in Tribology, CRC Press, Boca
Raton, FL, p. 98.
5. Blau, P. J. 2005. The nature of running-in, Tribol. Int., 38(11–12), 1007–1012.
6. Blau, P. J. 1989. Friction and Wear Transitions in Materials—Break-in, Wear-in, Run-in, Noyes
Publications, Park Ridge, NJ.
7. Dowson, D. et al. (eds.). 1981. The running in process in tribology, Proceedings of Eighth Leeds-Lyon
Conference on Tribology, Butterworths, Surrey, U.K.
8. Blau, P. J. 2009. Embedding wear models into friction models, Tribol. Lett., 34(1), 75–79.
9. Bayer, R. G. 2002. Wear Analysis for Engineers, HNB Publishing, New York.
10. Blau, P. J. 1998. Four great challenges confronting our understanding and modeling of sliding fric-
tion, in Tribology for Energy Conservation, Proceedings of 24th Leeds-Lyon Symposium on Tribology,
D. Dowson et al. (eds.), Elsevier, London, U.K., pp. 117–128.
11. Blok, H. 1937. Measurement of temperature flashes on gear teeth under extreme pressure condi-
tions, in Proceedings of the General Discussion on Lubrication and Lubricants, Vol. 2, Institute of
Mechanical Engineers, London, U.K., pp. 14–20.
12. Cowan, R. S. and Winer, W. O. 1992. Frictional heating calculations, in Friction, Lubrication, and Wear
Technology, Metals Handbook, P. J. Blau (ed.), Vol. 18, ASM International, Materials Park, OH, pp. 39–44.
5-14 Theory and Practice of Lubrication and Tribology

13. Malkin, S. and Marmur, A. 1977. Temperatures in sliding and machining processes with distributed
heat sources in the subsurface, Wear, 42(2), 333–340.
14. Bednar, M. S., Cai, B. C., and Kuhlmann-Wilsdorf, D. 1993. Pressure and structure dependence of
solid lubrication, Lubr. Engr., 49(10), 741–749.
15. Briscoe, B. J. and Stolarski, T. A. 1993. Friction, in Characterization of Tribological Materials,
W. A. Glaeser (ed.), Butterworth-Heinemann Publications, Boston, MA, pp. 30–64, Chapter 3.
16. Blau, P. J. 1992. Friction mechanisms and modeling on the macroscale, in Fundamentals of Tribology
and Bridging the Gap between the Macro and Micro/Nanoscales, B. Bhushan (ed.), Kluwer Academic
Publications, Dordrecht, the Netherlands, pp. 241–260.
17. Bethier, Y., Godet, M., and Brendle, M. 1989. Velocity accommodation in friction, Tribol. Trans.,
32(4), 490–496.
18. Scholz, C. H. 1990. The Mechanics of Earthquakes and Faulting, Cambridge University Press,
Cambridge, U.K., pp. 37–38.
19. Mate, C. M., McCelland, G. M., Erlandsson, R., and Chaing, S. 1987. Atomic scale friction of a tung-
sten tip on a graphite surface, Phys. Rev. Lett., 59, 1942–1945.
20. McClelland, G. M. and Glosli, J. N. 1992. Friction at the atomic scale, in Fundamentals of Friction:
Macroscopic and Microscopic Processes, I. L. Singer and H. M. Pollock (eds.), Kluwer Academic
Publications, Dordrecht, the Netherlands, pp. 405–436.
21. Daly, C. and Krim, J. 1996. Sliding friction of solid xenon monolayers and bilayers on Ag(111), Phys.
Rev. Lett., 76, 803–806.
22. Persson, B. N. J. 1998. Sliding Friction, Springer-Verlag Publications, Berlin, Germany, 462pp.
23. Grannick, S. 1992. Molecular tribology of films, in Fundamentals of Friction: Macroscopic and
Microscopic Processes, I. L. Singer and H. M. Pollock (eds.), Kluwer Academic Publications, Dordrecht,
the Netherlands, pp. 387–401.
24. Israelachvili, J. N. 1992. Adhesion, friction, and lubrication of molecularly smooth surfaces, in
Fundamentals of Friction: Macroscopic and Microscopic Processes, I. L. Singer and H. M. Pollock
(eds.), Kluwer Academic Publications, Dordrecht, the Netherlands, pp. 351–385.
25. Bhushan, B. 1990. Magnetic media tribology: State of the art and future challenges, Wear, 36(1),
169–197.
26. Hsu, S. M. and Ying, Z. C. (eds.). 2003. Nanotribology—Critical Assessment and Research Needs,
Kluwer Academic Publications, Boston, MA, 442pp.
27. Bowden, F. P. and Tabor, D. 1986. The Friction and Lubrication of Solids, Clarendon Press, Oxford,
U.K. (original printing 1950), 374pp.
28. Buckley, D. H. and Johnson, R. L. 1966. Friction and wear of hexagonal metals and alloys as related
to crystal structure and lattice parameters in vacuum, Tribol. Trans., 9(2), 121–135.
29. Steijn, R. P. 1964. Friction and wear of single crystals, Wear, 7(1), 48–66.
30. Feng, Z. and Field, J. E. 1992. The friction and wear of diamond sliding on diamond, J. Phys. D Appl.
Phys., 25, A33.
31. Rigney, D. A. and Hirth J. P. 1979. Plastic deformation and sliding wear of metals, Wear, 53(2),
345–370.
32. Blau, P. J. 2001. Compositions, Functions, and Testing of Friction Brake Materials and Their Additives,
Oak Ridge National Lab. Tech. report, ORNL TM-2001/64, 38pp.
33. Rabinowicz, E. and Tabor, D. 1951. Metallic transfer between sliding metals: An autoradiographic
study, Proc. R. Soc. A, 208, 455–475.
34. Litvinov, V. N., Mikhin, N. M., and Myshkin, N. K. 1979. The Physico-Chemical Mechanics of Selective
Transfer during Friction (book in Russian), Izdatel’stvo Nauka, Moscow, Russia, 188pp.
35. Godet, M. 1984. The third-body approach: A mechanical view of wear, Wear, 100, 437–452.
36. Godet, M. 1990. Third-bodies in tribology, Wear, 136, 29–45.
37. Heshmat, H. 1995. The quasi-hydrodynamic mechanism of powder lubrication—Part III: On theory
and rheology of triboparticulates, Tribol. Trans., 38(2), 269–276.
Friction 5-15

38. Fillot, N., Iordanoff, I., and Berthier, Y. 2007. Modelling third body flows with a discrete element
method—A tool for understanding wear with adhesive particles, Tribol. Int., 40(6), 973–981.
39. Scharf, T. W. and Singer, I. L. 2009. Role of the transfer film on the friction and wear of metal carbide
reinforced amorphous carbon coatings during run-in, Tribol. Lett., 36(1), 43–53.
40. Ahkmatov, A. S. 1939. Some items in the investigation of the external friction of solids, as ref. by I. V.
Kragheski, Friction and Wear, Butterworths, London, U.K., p. 159.
41. Ludema, K. C. 1984. Friction, in Handbook of Lubrication—Theory and Practice of Tribology, E. R.
Booser (ed.), Vol. II, CRC Press, Boca Raton, FL.
42. Jahanmir, S. (ed.). 1994. Friction and Wear of Ceramics, Marcel Dekker, New York.
43. Bartenev, G. M. and Lavrentev, V. V. 1981. Friction and Wear of Polymers, Elsevier Publications,
Oxford, U.K.
44. Moore, D. F. 1975. Principles and Applications of Tribology, Pergamon Press, Oxford, U.K.
45. Adams, M. J. 1992. Friction of granular non-metals, in Fundamentals of Friction: Macroscopic and
Microscopic Process, I. L. Singer and H. M. Pollock (eds.), Kluwer Publications, Dordrecht, the
Netherlands, pp. 183–207.
46. Clauss, F. J. 1972. Solid Lubricants and Self Lubricating Solids, Academic Press, New York.
47. Miyoshi, K. 2007. Solid Lubricants and Coatings for Extreme Environments: State-of-the-Art Survey,
NASA/TM2007-214668, NTIS, Springfield, VA, 22pp.
48. Lancaster, J. K. 1988. Solid lubricants, in Handbook of Lubrication, E. R. Booser (ed.), Vol. II, CRC
Press, Boca Raton, FL, pp. 269–290.
49. Erdemir, A., Halter, M., Fenske, G. R., Zuiker, C., Csencsits, R., Krauss, A. R., and Gruen, D. R. 1997.
Friction and wear mechanisms of smooth diamond films during sliding in air and dry nitrogen,
Tribol. Trans., 40(4), 667–675.
50. Czichos, H. and Mølgaard, J. 1977. Toward a general theory of tribological systems, Wear, 44(2),
247–264.
51. Aronov, V., D’Souza, A. F., Kalpakjian, S., and Shareef, I. 1984. Interactions among friction, wear, and
system stiffness—Part 1: Effect of normal load and system stiffness, J. Tribol., 106(1), 54–58.
52. Budinski, K. G. 1992. Laboratory testing methods for solid friction, in ASM Handbook, Vol. 18,
Friction, Wear, and Lubrication Technology, P. J. Blau (ed.), ASM International, Materials Park, OH,
pp. 45–58.
53. ASTM G115-04. Standard Guide for Measuring and Reporting Friction Coefficients, ASTM Annual
Book of Standards, Vol. 03.02, ASTM International, West Conshohocken, PA.
54. Lampman, S. (ed.). 1997. Friction and Wear Testing—Source Book, ASM International, Materials
Park, OH, 185pp.
55. Standard DIN 50322. Kategorien der Verschleiß Prüfung (Categories of Wear Testing), BAM, Berlin,
Germany, 1986.
6
Wear: A Synoptic View
6.1 Introduction....................................................................................... 6-1
6.2 Adhesive Wear................................................................................... 6-2
6.3 Abrasive Wear....................................................................................6-4
6.4 Fretting Wear.....................................................................................6-6
6.5 Erosive Wear....................................................................................... 6-7
Solid Particle Erosion  •  Solid Particle Erosion at High
Temperatures  •  Slurry Erosion and Abrasion  •  Cavitation
Erosion  •  Liquid Impingement Erosion
6.6 Electrical Sliding Contact Wear.................................................... 6-14
6.7 Arc Erosion Wear............................................................................ 6-14
6.8 Complexity of Wear Situations...................................................... 6-15
Shyam Bahadur 6.9 Investigating Tribological Failures............................................... 6-16
Iowa State University References..................................................................................................... 6-17

6.1  Introduction
The significance of friction and wear has been recognized through centuries and the first great tribo-
logical innovations were the wheel and the sled mechanisms, which helped to ease transportation [1].
The desire to increase the longevity of systems led to the recognition of wear. The early generations
tried to reduce friction and wear by using vegetable oils and animal fats such as tallow and lard in
bearings with wooden shafts sliding against wooden or stone bearing blocks. In those days, wooden
plows had stones embedded in them so as to reduce wear. The scientific understanding of wear in
terms of the complex phenomena originated in the early twentieth century and new materials such as
tungsten carbide and carboloy were developed in 1926 and 1942, respectively, to combat wear. Later
in the century, with the development of tools for surface examination and analysis combined with the
ability to treat the contact problem by mechanical analysis, a good understanding of the physics of
the friction and wear processes was obtained. Based on this understanding, the wear problem today
is solved by sound design principles, proper selection of materials and operating parameters, and
­effective lubrication.
Wear is a complex phenomenon. Complication arises because of the many factors that affect the wear
processes coupled with the changing conditions at the interface during service and the formation of
reaction products. Sometimes there are multiple wear mechanisms in play. These problems have made it
difficult to develop mathematical relationships that could be applied in designing for wear. The practi-
cal approach involves using proper materials, controlling the environment, and changing the operating
conditions so as to reduce the severity of anticipated wear.
There are many types of wear that are described in the following sections.

6-1
6-2 Theory and Practice of Lubrication and Tribology

6.2  Adhesive Wear


This form of wear occurs because of the adhesion between asperities on mating surfaces when they
are in intimate contact during sliding. Adhesion is a natural phenomenon that occurs because of the
interatomic attraction between the contacting surfaces. The cleaner the surfaces, the more intimate
is the contact and so the greater is the adhesion force. The latter is also enhanced by the increase in
load and/or temperature. Because of adhesion, adhesive bonds are formed at asperity contacts and
this gives rise to interfacial junctions. The latter are ruptured in sliding and the force needed to
rupture these junctions contributes to the friction force. The properties of the junction determine
where shear will occur during sliding such as at the junction interface or within one of the contacting
bodies involved. Adhesion often contributes to the transfer of material from one of the contacting
bodies to another body and this gives rise to adhesive wear. Thus, the key to reducing adhesive wear
is to reduce adhesion, which may be done by lubricating the surfaces, reducing the loading between
surfaces, suppressing the temperature rise during sliding, and using contacting materials with low
interatomic diffusion capability. For the latter, the materials should be noncompatible or with low
mutual solubilities. Figure 6.1 gives a chart showing the relative mutual solubilities of pairs of pure
metals [2].
There are four possibilities related to the formation and shearing of junctions:
1. If the shear strength of the junction is weaker than that of both contacting materials, shearing
during sliding occurs at the interface. In this case, wear is very small. This is essentially the case
under liquid lubrication.
2. If the junction is stronger than one of the contacting materials but weaker than the other, shearing
occurs at a short distance from the interface within the weaker material.

W Mo Cr Co Ni Fe Nb Pt Zr Ti Cu Au Ag Al Zn Mg Cd Sn Pb In
∆ ∆ ∆ ∆
In
∆ ∆ ∆
Pb
∆ ∆ ∆ ∆ ∆ ∆ ∆
Sn
∆ ∆ ∆ ∆
Cd
∆ ∆ ∆
Mg
∆ ∆
Zn
∆ ∆ ∆ ∆
AI
∆ ∆
Ag
∆ ∆ ∆
Au
∆ ∆
Cu

Ti
∆ ∆ ∆ ∆
Zr
Pt
Two liquid phases
Nb
Fe One liquid phase, solid
solubility below 0.1%
Ni ∆
Solid solubility between 1% and 0.1%
Co
Cr Solid solubility above 1%
Mo Identical metals
W

FIGURE 6.1  Chart showing the relative mutual solubilities of pairs of pure metals. (From Rabinowicz, E., Wear
coefficients­­—metals, in Wear Control Handbook, M.B. Peterson and W.O. Winer (eds.), ASME, 475, 1980.)
Wear 6-3

3. If the junction is stronger than the softer contacting material and occasionally stronger than the
weaker spots (such as due to inclusions) in the harder mating material, most of the shearing will
occur in the weaker material but some may also occur in the harder material.
4. The junction is stronger than both contacting materials, shearing then occurs within one or the
other material and the wear is severe. This is the case with similar metals in contact because the
junctions are strengthened by work hardening due to deformation.
There are some equations that have been proposed for adhesive wear, but none applies in general. The
Archard’s equation is the one most commonly referred by the researchers and used by the designers.
According to this equation, the wear per unit sliding distance

V W (6.1)
=K⋅
L H

where
V is the volume of wear
L is the sliding distance
W is the load between the surfaces in contact
H is the hardness of the wearing material

K is a dimensionless factor introduced to match the calculated values of wear with the experimental
values for a particular material combination. The factor K is known as the wear coefficient and its values
for a variety of material combinations are given in the literature [3]. The values of K vary by a factor
of 105. Some representative values of K are given in Table 6.1.
Archard’s wear equation shows a direct proportionality between wear rate and load and inverse pro-
portionality between wear rate and hardness. For many material combinations, these proportionalities
do not apply. This is particularly true for polymers sliding against a metal surface where the increase in
load may produce thermal softening resulting in the exponential increase in wear with sliding distance.
Similarly, the wear rate may not depend upon the hardness of the polymer but instead upon the ability
of polymer to form transfer film upon the metal counterface. The work by Archard [3] provides a good
coverage of the fundamental aspects of wear.
The surface features of adhesive wear are shown in Figure 6.2 [4]. Here, a laser-melted surface of
Ti-6Al-4V alloy was in dry sliding contact with a tool steel ring in a block-on-ring setup. There is evidence

TABLE 6.1  Wear Coefficient K for Dry Sliding


at 1.8 m/s between the Material in the Table and
the Hardened Tool Steel, Except Where Stated
Otherwise
Material K
Mild steel on mild steel 7.0 × 10−3
60/40 leaded brass on steel 6.0 × 10−4
Ferritic stainless steel on mild steel 1.7 × 10−5
Stellite on steel 5.5 × 10−5
Tungsten carbide on tungsten carbide 1.0 × 10−6
PTFE on steel 2.5 × 10−5
Polyethylene on steel 1.3 × 10−7
Source: From Archard, J.F., Wear, interdisciplinary
approach to friction and wear, in Proceedings of NASA-
Sponsored Symposium, P. M. Ku (ed.), San Antonio, TX,
November 28–30, 1967, pp. 267–333.
Note:  Pin-on-disk arrangement. Load 400 g.
6-4 Theory and Practice of Lubrication and Tribology

0004 15 KV 1.0 µm WD 14

FIGURE 6.2  Adhesive wear features on Ti-6Al-4V laser-melted surface (hardness 590 DPH) in sliding con-
tact with a tool steel ring (hardness 547 DPH). Sliding velocity 4 m/s, load 9.81 N. (From Yerramareddy, S. and
Bahadur, S., The effect of laser surface treatments on the tribological behavior of Ti-6Al-4V, in Wear of Materials,
K. C. Ludema and R. G. Bayer (eds.), 531, 1991. With permission.)

of the formation of strong adhesive junctions, which, when pulled apart, resulted in material removal. This
is consistent with the observation in Figure 6.1, which shows high solubility between steel and titanium.
Adhesive wear is prevented in practice by lubricating the contact surfaces. When the surfaces are
deprived of adequate lubrication because of high contact pressure or the failure of lubricant supply to the
contact location, high friction resulting in high temperature occurs. In this case, the surface appearance
indicates seizure and cratering, which is evidenced by rough and torn topography. This kind of failure
is known as scuffing or scoring.
Galling is another kind of adhesive wear that occurs in low-speed sliding when the contact zone is
very heavily loaded so that large plastic deformation and flow of material occur. It is characterized by
severely roughened surfaces accompanied by the transfer or displacement of large fragments of material.

6.3  Abrasive Wear


This form of wear occurs when hard protuberances or hard particles are forced against a surface in
sliding motion. Abrasion is a serious problem in agricultural, mining, and earthmoving equipment
that handle abrasive materials such as soil, rock, and minerals. It also occurs due to atmospheric dust
that gets in the contact region, as in rotating and sliding bearings, seals, etc. The loss from abrasion is
estimated to be 1%–4% of the gross national product (GDP) of an industrialized country. In contrast to
this, abrasion is also used to advantage in polishing operations and abrasive machining.
There are two kinds of abrasion: two-body abrasion and three-body abrasion. In two-body abrasion, the
loss of material from a surface occurs when it slides against another harder surface with protuberances, such
as a file or sandpaper. In three-body abrasion, wear occurs when an abrasive material is entrapped between
the sliding surfaces. The wear in two-body abrasion is generally 10–1000 times higher than that in three-
body abrasion for the same conditions of sliding. In both kinds of abrasion processes, the damage is caused
by the mechanisms such as cutting and plowing followed by repeated plastic deformation, which leads to
subsurface crack nucleation and propagation, resulting finally in the separation of material from the sur-
face. The distinguishing surface feature in abrasion is the formation of long grooves as shown in Figure 6.3.
Abrasion is also classified as low stress and high stress. In low-stress abrasion, the stress on the abra-
sive material is small so that it does not lose its structure, as in the sanding of wood with sandpaper.
In high-stress abrasion, the abrasive particles are broken up during abrasion as in ball-mill grinding.
Sometimes, the term gouging abrasion is also used to describe high-stress abrasion by large lumps of
hard abrasive particles as in the rock crushing machinery.
Considering an abrasive particle, idealized as a conical asperity, dragged under load W on a softer
surface of hardness H and assuming that the material displaced in the grooving process by plastic flow
Wear 6-5

15 KV 10 µm

FIGURE 6.3  SEM micrograph of the abraded surface of laser-melted Ti-6Al-4V surface in dry sand–rubber wheel
abrasion using AFS 50-70 sand. Load 133.5 N, sand flow rate 360–370 g/min, rubber wheel speed 200 rpm. (From
Yerramareddy, S. and Bahadur, S., The effect of laser surface treatments on the tribological behavior of Ti-6Al-4V,
in Wear of Materials, K. C. Ludema and R. G. Bayer (eds.), 1991, pp. 531–540. With permission.)

is removed, a simple equation for abrasive wear has been derived. In the presence of a large number of
such abrasive particles on surface, the equation for abrasive wear is simplified to the following form:

V W
= Ka ⋅ (6.2)
L H

where
V is the volume of material removed in a sliding distance L
Ka is the abrasive wear coefficient
W and H have the same meaning as in Equation 6.1

This equation is of the same form as the Archard’s equation (6.1) for adhesive wear. The values of Ka for
two-body abrasion of metals are in between 5 × 10−3 and 5 × 10−2 and for three-body abrasion in between
5 × 10−4 and 5 × 10−3, approximately. Notice that Ka is about one order of magnitude lower for three-body
abrasion than for two-body abrasion. Furthermore, the wear coefficients for abrasive wear are much
higher than those for adhesive wear (Table 6.1).
The works of Krushchov [5] and Mulhearn and Samuels [6] have shown that for metals the abrasive
wear is inversely proportional to the hardness of the surface being abraded. The abrasive wear equation
also shows the same relationship. This would indicate that for reduced abrasive wear the hardness of
the surface being abraded should be increased, which is commonly practiced in industry. The experi-
ments reveal that for effective abrasion the abrading surface should be 50% harder than the surface
being abraded [7]. For this reason, polishing operations are performed using the abrading powders of
diamond, SiC, Al2O3, and Garnet, which have high hardness values as given in Table 6.2.

TABLE 6.2  Abrading Materials and


Their Hardness
Abrading Materials Hardness (kg/mm2)
Diamond 8000
SiC 2500
Al2O3 2100
Garnet 1350
SiO2 (silica) 800
6-6 Theory and Practice of Lubrication and Tribology

The discussion of abrasion in further details is provided in [8,9] and in a subsequent chapter in this
book. These chapters discuss the abrasive wear mechanisms, provide the experimental abrasion data on
various materials, and provide the remedial measures if abrasion needs to be controlled.

6.4  Fretting Wear


This form of wear results from very small-amplitude oscillatory displacements (100 μm or less)
between the contacting surfaces that are nominally at rest. Such displacements are caused generally
by ­v ibrations. When such displacements occur in atmosphere, oxide debris is generated between the
contact surfaces. It interacts with the rubbing surfaces and produces wear. The wear in this case is
known as fretting corrosion or fretting wear. If the small-amplitude displacements occur because of
a cyclic loading that results in the initiation of fatigue cracks, the damage then is known as fretting
fatigue. Some of the examples where fretting failure occurs are shrink-fit and press-fit joints, bearing
housings on loaded rotating shafts, flexible couplings and splines, blade and shaft dovetails in turbines,
steel rope wires in cranes and elevators, tube supports in heat exchangers, automotive leaf springs, ball
and race in ball bearings, riveted and fastener joints in aircraft, and prosthetic joints in humans and
animals.
In terms of the progression of events, the following three stages can be identified in a fretting process:
1. Initial adhesion that results in welding at asperity junctions in the case of metallic contact. The
breakage of welded junctions results in roughening of the surface accompanied by high friction
and low contact electrical resistance.
2. Formation of oxidized debris and the beds of compacted oxide. The contact resistance in this
process oscillates between high and low values.
3. Steady-state situation in which the friction and contact resistance remain fairly stable with
­occasional dips.
In the case of steels, fretting in air results in the formation of black iron oxide Fe3O4 in the fretting
­location. It changes in the peripheral areas to ferric oxide (α-Fe2O3), which is reddish brown in color and
is highly abrasive. In high-humidity situations, the oxide formed is hydrated ferric oxide (α-Fe2O3·H2O).
The shape of this oxide debris is platelike.
In the case of cyclic loading, fatigue cracks are observed in fretted regions. These cracks can be initi-
ated at very low stresses ∼30 ksi, which are well below the fatigue limit of the material, but the stresses
at asperity junctions are very high. The direction of cracks is perpendicular to the maximum principal
stress in the fretting area. The surface fatigue produces local pitting and/or flaking, which are the char-
acteristic signs of fretting.
The factors or processes that affect fretting are as follows:
1. Slip amplitude—When the amplitude is too small (≤10 μm), fretting does not occur because the
deformation in asperity contacts is basically elastic. When the amplitude is large (≥100 μm), the
debris from the contact region is able to escape and so the situation corresponds to sliding and
fretting does not occur.
2. Normal load—Fretting wear increases with the increase in load.
3. Frequency of vibration—In the partial-slip situation, no significant effect of the increase in fre-
quency on fretting wear has been reported [10]. At higher frequencies the increase in the inter-
facial strain rate results in increased fatigue damage and increased corrosion, both of which
contribute to greater wear.
4. Surface roughness—Fretting damage has been reported to be less with rougher surfaces [11].
5. Atmosphere—With the decrease in atmospheric air pressure, fretting wear damage in steels
decreases because of the reduced ability of oxidation [12]. Humidity affects fretting wear also
because it affects oxidation. Goto and Buckley [13] reported that in pure metals the fretting wear
Wear 6-7

damage was maximum between 10% and 15% relative humidity (RH) for silver, copper, titanium,
and iron. At 50% RH, all the metals showed a decrease in fretting wear. In ceramics such as sili-
con carbide, alumina, and zirconia, the wear rate is greatly reduced at 50% RH compared to that
at 5% RH.
6. Solubility of materials—In the case of dissimilar materials, the mutual solubility determines the
fretting damage. For example, nickel, chromium, and iron are severely damaged when fretted
against iron or chromium because of high mutual solubility but the damage is much less when
fretted against copper or silver because there is no mutual solubility. In the case of materials with
no solubility, such as polymer and steel, fretting damage is caused by iron oxide that is formed on
the steel surface [14].
7. Adhesion may be prevented by lubrication or separation of the metallic surfaces by a layer of poly-
meric material and this would lead to reduced fretting failure.
8. Processes such as surface rolling, shot peening, and case hardening that retard fatigue crack
growth due to compressive stress near surface reduce fretting damage.
9. Fretting can be eliminated by tightening, which would reduce the amplitude of oscillatory
displacement.
A detailed discussion of fretting is provided in a subsequent chapter in this book. For general review,
readers are encouraged to consult [15–17].

6.5  Erosive Wear


The progressive loss of material from repeated impacts of solid or liquid particles is called erosive wear.
The key to erosion is the impact of particles traveling free while in abrasion the particles slide across
a surface under the action of an externally applied force. There are many kinds of erosion, which are
described hereafter.

6.5.1  Solid Particle Erosion


In this kind of erosion, solid particles traveling at high speeds impact a solid surface. The force associ-
ated with the impacting particles erodes the surface at the points of impact and the particles bounce
off the surface at an angle that depends upon the angle of impact and the target surface properties.
The examples where erosive wear occurs are pneumatic transport of solid particles as in coal handling
plants and mines, ash removal systems, airfoils in aircraft turbines and compressors, waste disposal
plants, and fluidized bed combustion systems. As opposed to the erosion damage in these cases, the
solid particle erosion process is also used to advantage in the case of sand blasting and abrasive machin-
ing processes.
The cumulative erosion damage caused by a stream of impacting hard particles depends upon the
particle velocity, angle of impact, particle shape and size, and the target material. Of all the variables
that affect erosion most significantly are the impact velocity and the angle of impact. The variation of
erosion rate € (defined as the surface mass loss per unit mass of impacting particles) with the impact
velocity of particles v has been reported to be of the form € α vn where n varies from 2 to 3 depending
upon the size of particles [18–21]. The schematic variation of erosion rate with the angle of impact of
particles is shown in Figure 6.4. The erosion is the highest at 20°–30° for ductile metals and at 90° for
brittle materials [22].
For ductile materials, an increase in particle size increases the erosive damage until a saturation level
is reached after which no further change is indicated. The shape of the particles has been indicated to
affect erosion significantly, with particles having sharp contours providing higher erosion rates [23,24].
As for the mechanisms of erosion, round particles deform the ductile target surface on impact by
plowing, which involves the displacement of material to the side and in front of the particle. Subsequent
6-8 Theory and Practice of Lubrication and Tribology

Ductile
Erosion, ε

Brittle

0 10 20 30 40 50 60 70 80 90
Angle of impact, α°

FIGURE 6.4  Schematic variation of erosion rate with impingement angle of impacting particles.

10 µm

FIGURE 6.5  SEM of Ti-6Al-4V surface, solution treated at 1065°C and aged at 480°C for 4 h, eroded with 120 grit
SiC particles at 55 m/s velocity and 30° angle of impact. (From Yerramareddy, S. and Bahadur, S., Wear, 157, 245,
1992. With permission.)

impacts lead to the formation of platelets, cracking of the edges of these platelets, and finally the detach-
ment of cracked rims of the crater. The surface features of an eroded ductile target material are shown
in Figure 6.5. Here the platelets can be clearly seen. In the case of angular particles, some cutting occurs
also along with the deformation described earlier.
In the case of brittle materials, the impact of particles produces erosion by the formation and intersec-
tion of cracks along with some of the aforementioned effects. More rounded or softer particles tend to
cause purely elastic deformation and conical Hertzian factures.
When solid particle erosion occurs in corrosive media, it is known as erosion–corrosion. Such situa-
tions involve exposure to high temperature or liquid media, as in coal gasification and liquefaction, gas
turbines, and fluidized bed systems. In this case, there is a synergistic interaction between solid particle
Wear 6-9

erosion and corrosion. The subject matter is discussed separately under the following headings, namely,
high-temperature erosion, slurry erosion, liquid impingement erosion, and cavitation erosion.

6.5.2  Solid Particle Erosion at High Temperatures


It is relevant basically to metals that form surface films such as oxides when exposed to high tem-
peratures. Erosion at high temperatures is thus an erosion–corrosion phenomenon. In this case, the
interacting effects of erosion and oxide layer need to be considered. Here along with erosion the thick-
ness, morphology, and deformability of oxide layer and its adherence to the substrate become important
­factors. If the oxide scale is adherent and flexible, the wear is controlled by erosion because oxide pro-
vides a protective layer. In other words, the erosive wear in this case will be less because of oxidation. On
the other hand, if the oxide layer is brittle and poorly adherent, the oxide will be lost with the impact of
particles. In this case, oxide does not provide any beneficial effect.
The variation of erosion rate (also called wastage rate) as a function of temperature and velocity is
shown in Figure 6.6a for mild steel impacted by 80 mesh alumina particles [26]. At any temperature,
the increase in impingement velocity results in higher erosion rate. As for the temperature, erosion
rate increases initially with the increase in temperature up to 400°C but decreases at higher tempera-
tures. In many cases, a rapid increase in erosion rate followed by leveling is observed with the increase

5 1.76 m/s
Wastage rate (µm/h)

2.04 m/s
4 2.58 m/s

0
0 100 200 300 400 500 600
(a) Temperature/°C

16 2 – 1/4 Cr–1 Mo
80 14 9 Cr–1 Mo
9 Cr–1 Mo (mod.) 304 SS
30° 310 SS
30 m/s 12
Erosion rate, nm/g

60 30°
Erosion rate, µg/g

250 µm SiC 2.7 m/s


10
1 mm Al2O3
8
40
6

20 4
Room
2 temperature
0 0
300 500 700 900 700 900 1100 1300 1500
(b) Temperature, K Temperature, K

FIGURE 6.6  Erosion rate vs. temperature for (a) mild steel and (b) 9 Cr–1 Mo steel, 304 and 310 SS, and
Cr–Mo steels. (a: From Hutchings, I.M. et al., Proceedings Conference on Corrosion–Erosion–Wear of Materials at
Elevated Temperatures, A. V. Levy (ed.), Berkeley, CA, January 31–February 2, 1990, Paper 14. With permission; b:
Zhou, J. and Bahadur, S., Further investigations on the elevated temperature erosion–corrosion of stainless steels,
in Proceedings of Conference on Corrosion–Erosion–Wear of Materials at Elevated Temperatures, A. V. Levy (ed.),
Berkeley, CA, January 31–February 2, 1990, Paper 13.)
6-10 Theory and Practice of Lubrication and Tribology

Chipping and
Substrate deformation
craking Outgoing
Incoming and chipping
particle
particle
Particle

Oxide Oxide
Substrate Substrate

Type 1 Type 3

Spalling Digging

Oxide Oxide
Substrate Substrate

Type 2 Type 4

FIGURE 6.7  Schematic representation of the various possibilities of oxide scale removal from the surface under
multiple impacts at high temperatures. (From Zhou, J. and Bahadur, S., Further investigations on the elevated tem-
perature erosion–corrosion of stainless steels, in Proceedings of Conference on Corrosion–Erosion–Wear of Materials
at Elevated Temperatures, A. V. Levy (ed.), Berkeley, CA, January 31–February 2, 1990, Paper 13. With permission.)

in temperature as shown in Figure 6.6b [27]. The leveling in erosion occurs because of the protective
role of oxide.
The whole erosion–corrosion behavior may be divided into the following four categories:
1. Erosion dominant—This is the case when material wastage is mainly from erosion and much less
from corrosion. This would apply to noncorroding atmospheres such as inert gas or relatively low
temperatures where oxidation is negligible.
2. Corrosion dominant—This is the case when the ratio of corrosion damage to erosion damage is
high. This would be the case when corrosion is severe and erosion is small either because of low
particle impingement velocities or fine or rounded impacting particles.
3. Oxidation-affected erosion—This is the case where oxide scale is adherent and pliable. There are
two scenarios possible here. When the oxide scale is thick enough to be felt by the impacting par-
ticles but not so thick as to shield the substrate from erosion damage, the oxide and the substrate
are both deformed from particle impacts and the erosion rate is a function of the mechanical
properties of both the metal and the oxide. When the oxide scale is fairly thick, the damage from
impacting particles may be mostly confined within the oxide scale. Here the erosion rate will be a
function of the properties of oxide.
4. Oxidation-controlled erosion—This is the case when oxide scale is brittle and nonadherent. Here
oxide spalls under impact conditions and the process of oxidation and spalling are continuous.
Figure 6.7 provides the schematic representation of the various possibilities of oxide scale removal from
the surface under multiple impacts at high temperatures [27].

6.5.3  Slurry Erosion and Abrasion


Slurry erosion is another example of erosion–corrosion. Here the surface of material is exposed to a
high-velocity stream of slurry. This type of situation occurs in centrifugal pumps that are used to pump
Wear 6-11

slurry and in the transportation of minerals through pipelines at high pressures. Whereas erosion
occurs on the vanes and surrounding walls, abrasion occurs on pump valve and seat surfaces, piston
walls, cylinder liners, and plungers. Along with erosion and/or abrasion in these cases, corrosive wear
occurs also because of the presence of corrosive elements present in the liquid and the galvanic effect
if two metals are involved. The pumps, elbows, T-junctions, valves, and hydrocyclones are parts of the
slurry handling system subjected to erosive wear. In submerged valves that open and close to control the
flow of slurry, the parts are subjected to erosion and corrosion both.
When dealing with wear in such systems, slurry abrasivity is an important factor. The ASTM G75
standard [28] provides a test procedure to determine the slurry abrasivity, which is denoted in terms of
the Miller number. The higher the Miller number, the greater is the abrasivity of the slurry. This stan-
dard gives the Miller numbers for some slurries.
Based on the same procedure, another parameter known as the SAR number has been developed. It
provides a measure of the relative abrasion response of the material in the particular slurry. The higher
the SAR number, the greater is the damage to material from a slurry.
As distinguished from slurry abrasion, in slurry erosion the slurry particles impacting the target
material are not constrained by another surface. Instead the particles strike the target surface and
bounce off under the influence of their momentum in the liquid atmosphere of slurry. The erosion rate
in recycled slurry is smaller than in fresh slurry because particles tend to get rounded on impacts [29].
Lynn et al. [30] have shown that the wear rate is increased as the particle size in slurry increases, because
for the same velocity larger particles cause greater damage than smaller particles.
The wear rate is highly dependent upon the slurry velocity showing a relationship of the form € α vn
where n varies from 2 to 3 [31]. When the particle impacts are of glancing nature, the wear damage is
lower than when the angle of impact of particles increases because in the latter case the corrosive film is
removed and so the wear contribution from corrosion is increased [32].

6.5.4  Cavitation Erosion


In devices such as hydraulic pumps, turbines, valves, pipelines, hydrofoils, ship propellers, etc., cavi-
ties or gaseous bubbles are formed because of the pressure gradients in the liquid or the presence
of gaseous impurities. When subjected to compressive stresses, the cavities or bubbles collapse and
produce a shock wave into the surrounding liquid. Those cavities that are either in contact with or
very close to a solid surface form a microjet of liquid toward the solid. The impact from the clusters of
these cavities applies mechanical loads large enough to deform the surface locally. The repeated load-
ing eventually leads to the removal of material from the impacted surface and this process is known
as cavitation erosion.
There is initially an incubation period during which material deforms either elastically or plasti-
cally when subjected to loading from bubble collapses and there is no visible loss of material from the
surface in this stage. However, the roughening of surface by plastic deformation during the incuba-
tion period may make the component unsuitable for some applications. After the incubation period,
the loss of material from erosion is initially rapid. The shape of the erosion rate–time curve may
be composed of all or some of the following periods or stages: incubation, acceleration, maximum
rate, deceleration, terminal, and occasionally catastrophic. These stages are schematically shown in
Figure 6.8 from ASTM G 32-98 Standard [33]. A general treatment of cavitation erosion is provided
by Preece [34].
Since the deformation in cavitation erosion is localized normally within a grain or a part thereof, the
bulk deformation behavior principles do not apply. As such no correlation of the erosion resistance with
bulk mechanical properties has been shown. Studies have shown that the cobalt-based alloys and, to a
lesser extent, austenitic stainless steels are most effective for cavitation erosion resistance [35]. In order
to reduce cost, weld overlays of stainless steel or cobalt alloys are used in practice to provide resistance
to cavitation erosion.
6-12 Theory and Practice of Lubrication and Tribology

Terminal
Maximum rate
rate line
line

Cumulative erosion (material loss)

A Exposure time

“Maximum
Acceleration stage

erosion
Incubation stage
Erosion rate

rate”
Maximum rate

Terminal stage
Deceleration

(if exhibited)
stage

stage

Exposure time
Notes: 1–A, =nominal incubation time; tan B, = maximum erosion rate;
tan C,= terminal erosion rate; and D, = terminal line intercept.

FIGURE 6.8  Normalized erosion resistance of various metals and alloys relative to 316 austenitic stainless steel of
hardness 170 VHN. Erosion test data is not very consistent and so the information herein should be used only as a
rough guide. (From Heymann, F.J., Liquid Impingement Erosion, ASM Handbook, Vol. 18, Friction, Lubrication and
Wear Technology, 1992, pp. 221–232. With permission.)

Since cavitation erosion occurs in a liquid environment, corrosion is also a possibility. The corrosion
process is an electrochemical process, which involves cathodic and anodic reactions. In the cathodic
reaction, hydrogen gas is evolved. This gas cushions the collapse of cavities and thereby suppresses
­cavitation erosion.

6.5.5  Liquid Impingement Erosion


It is similar to cavitation erosion with the only difference that in liquid impingement erosion the liquid
drops or discontinuous jets impinge on the solid surface as opposed to the cavities or gaseous bubbles in
cavitation erosion. In both the cases, the progressive loss of material from the target surface occurs due
to repeated impacts, which produce impulsive mechanical loading causing localized plastic deformation
and finally the separation of material. This type of erosion is observed on low-pressure steam turbine
blades operating with wet steam, and the outer skin of aircrafts and missiles, and rotor blades of helicop-
ters from the impact of rain drops. The degradation of the optical properties of window materials from
the impingement of rain drops in fast-moving devices is also the result of this kind of erosion. The ero-
sion behavior in liquid impingement erosion is similar to that shown for cavitation erosion in Figure 6.8.
Wear 6-13

The presence of liquid raises the possibility of corrosion. If impulsive mechanical interactions are
not significant, erosion is insignificant. Since the flow of liquid inhibits the equilibrium in the corro-
sion process, corrosion continues. If the impingement velocity is very high, erosion outweighs the effect
of corrosion. Coulon [36] provided the following guidelines: corrosion, 0–10 m/s; corrosion–erosion,
10–50 m/s; erosion–corrosion, 50–200 m/s; and erosion, >200 m/s. The comprehensive review of liquid
impingement erosion may be found in References 37 and 38, ASTM Symposia volumes [39,40], and the
Proceedings of “Rain Erosion” and “Erosion by Liquid and Solid Impact” Conferences [41–43].
The variables in liquid impingement erosion are impingement velocity, impact angle, and droplet size.
The maximum erosion rate has been reported to be proportional to vn where n lies between 4 and 5 for
ductile materials [44] and between 6 and 9 for brittle materials [45]. The erosion is small at low impact
angles and is affected mainly by the normal component of the impact velocity. As for the droplet size,
the smaller the size the less is the erosion for the same total amount of liquid.
For both liquid impingement erosion and cavitation erosion, materials with high toughness and
work-hardening ability are found to be good for erosion resistance. Figure 6.9 gives the normalized

Hardness, Normalized erosion resistance


Material
HB or HV

Carbon steel 110 – 190


Cast iron 140 – 230
1¼ Cr – ½ Mo and 2¼ Cr – 1 Mo cast steel
Austenitic stainless steel (300 series) 140 – 230
Austenitic stainless steel weld overlay (301, 308, 309, 316)
Martensitic stainless steel (403, 410, 422) 200 – 400
17 – 4 PH, 17 – 2 PH stainless steel (603, 631) 320 – 460
12 Cr – 4 Ni cast stainless steel
Maraging steel 450 – 650
Ausformed steel 450 – 620
Tool steel 400 – 900
Titanium alloys 300 – 400
Stellite 6, 6B, 12 380 – 500
Stellite 6, 21 weld overlay
Aluminum 20 – 90
Aluminum alloys 100 – 200
Copper
Cupro-nickel 70 – 200
Brass 60 – 200
Manganese bronze 120 – 230
Aluminum and nickel–aluminum bronze 140 – 220
Nickel 70 – 90
Monel 120 – 360
Inconel 150 – 380
Inconel weld overlay
Nitronic–60 bar
IRECA austenitic cobalt stainless steels 220 – 340
EPRI/AMAX norem non-cobalt weld overlay
Iron aluminide
Nickel aluminide bar
Nickel aluminide weld

0.01 0.02 0.05 0.1 0.2 0.5 1 2 5 10 20 50 100


Normalized erosion resistance relative to type 316
austenitic stainless steel of hardness HV 170

FIGURE 6.9  Characteristic stages of erosion rate–time behavior and parameters for representation of the cumu-
lative erosion–time curve for cavitation erosion. (From Annual Book of ASTM Standards, Vol. 03.02, Wear and
Erosion; Metal Corrosion, ASTM International, West Conshohocken, PA. With permission.)
6-14 Theory and Practice of Lubrication and Tribology

erosion resistance for a variety of materials [46]. It should be noted that Stellite 6 has the highest erosion
resistance and so is often used as a shielding or cladding material. For protection from rain erosion,
elastomeric coatings are often used because they absorb impact loading thereby reducing the load on
the substrate.

6.6  Electrical Sliding Contact Wear


The common example of electrical sliding contact wear is in the electrical brushes that run on metal-
lic slip rings or commutators as in a motor. In these contacts, arcing is not a significant factor. The
tribological aspects of the electrical contact field are covered collectively in a volume of Wear [47],
which contains 25 papers. There is also another publication [48] that contains 75 papers on the
­subject. Since the contacts operate in air, an insulating film is developed on the contact surfaces. The
film tends to break down due to intermittent contacts that results in the formation of a conductive
junction thereby reducing the contact resistance. Due to the heating effect from both sliding and
potential difference, the process of the formation and breaking of contact is an ongoing phenomenon.
In switches and direct current vehicular systems, circuit breaking under current flow is the same
kind of phenomenon. Thus the wear in these systems is caused by relative motion between the mat-
ing parts as well as the electrical contact phenomenon that results in the transfer of material from
cathode to anode.
The electrical sliding research has mostly been directed to the studies of friction, wear, electrical
contact resistance, and material transfer in the sliding systems involving carbon brush and slip-ring
materials. The wear of carbon brushes has been reported to increase significantly when mated against
steel and high-zinc brass rings [49,50]. Shobert [51] showed that excluding high friction situations in
light-load running, the wear rates of graphite and metal graphite brushes siding against copper and cop-
per alloy commutators were usually higher with electrical current than without it. Konchits [52] studied
the influence of the magnitude and the direction of electrical current on the wear of carbon brushes. He
concluded that the basic cause of increased brush wear under electrical current was the joule heat energy
released in the friction zone. The general pattern observed was the rapid increase in wear intensity with
the increase in electrical current density.
Garshasb and Vook [53] studied the electrical contact system involving a copper brush sliding on
a silver slip ring in humidified CO2. They found that with the increase in electrical current the metal
transfer across the interface increased and so did the number of wear particles. At higher currents, the
magnitude of metal transfer varied linearly with the square of the current. Reichner [54] studied the
sliding friction and wear of metal fiber brushes against silver-plated copper slip rings in humidified
CO2 atmosphere. He observed very low wear rate for fiber brushes at high current densities. Burton and
Burton [55] studied the friction, wear, and resistivity of the specimens made of copper fibers in vitre-
ous or glassy carbon matrix sliding against 303 stainless steel. The resistivity and friction observed in
this case were comparable to those reported for typical copper graphite specimens and the wear was
exceptionally low.

6.7  Arc Erosion Wear


This type of wear damage occurs in electrical contacts that are of two types: make-and-break contact
and sliding contact. In the make-and-break contact, a contact member moves perpendicularly to and
from another contact surface as in an electrical switch. In the sliding contact, the tangential motion
between the contact surfaces is involved as in electrical transport systems. In both the types of con-
tacts, arcing is a significant factor in governing the life of contacts. Arcing in make-and-break contacts
occurs when the contacts are in the process of establishing or interrupting the current flow. In sliding
Wear 6-15

contacts, arcing occurs because of high voltage and unsteady contact that causes the contact members
to separate. Arcing involves the gaseous discharge involving electrons, metal vapors and ions, high
current density in the arc column, and the rise in temperature [56]. Because of the high temperature
and mass flow, the contact surface is severely corroded and eroded. Michal and Saeger [57] documented
that the contacts of AgCdO, AgNi, and AgSnO2 in DC load gained material at the cathode and lost at
the anode. There was thus transfer of material involved. The continuous arc duration consists of two
parts: the metallic phase for low current and the gaseous phase for high current [58]. Sone and Takagi
[59] studied the influence of metallic phase arc discharge on contact performance and showed that
the contact resistance increased when the arc duration in the metallic phase arc region grew and it
remained high and almost constant after the arc reached the transition border to the gaseous state.
They said that the material loss and transfer were proportional to the arc duration. Shen and Gould
[60] studied the contact surface, cross section, and debris in order to understand the erosion modes
of AgCdO and Ag(Sn,In)O materials made by the internal oxidation method. They found that the
internally oxidized Ag(Sn,In)O was more brittle than AgCdO made by the same method and that the
fabrication method of the contact material was one of the important factors influencing the erosion
mode. Donaldson et al. [61] studied the arc erosion behavior of Cu–Nb composite at high currents and
high-energy transient arcs and found that Cu–Nb had lower erosion than that of the commercially used
Cu–W composite.
Ping and Bahadur [62] studied the low-energy make-and-break contacts and the high-energy arc
gaps made of OFHC Cu, Cu-15% Nb, and Cu-15% Cr composites. They found that in the case of make-
and-break contacts the contact resistance increased with increasing number of cycles and this was
due to the development of oxide film on the contact surface as well as surface deterioration. The arc
erosion rates of the composite materials were higher than that of OFHC copper. For the same number
of cycles, the thickness of oxide film in the case of OFHC copper was the lowest and that of Cu-15%
Cr was higher than that of Cu-15% Nb. This was explained in terms of the relative reactivities of Cu,
Cr, and Nb.
In the case of high-energy arc gaps, erosion rates for all the three materials increased with increasing
electrical energy. The arc erosion of pure copper was the highest whereas the erosion rates of the com-
posites were about the same. The examination of the hemispherical tip surfaces showed bulk melting in
the case of Cu–Cr composite but that was not the case for Cu–Nb composite. Instead, in this case some
craters were observed because of localized erosion and melting. Thus melting was the major mode of
erosion damage.

6.8  Complexity of Wear Situations


Many real-life situations are confronted with more than one kind of wear. For example, a system may
encounter adhesive wear in the beginning but the oxidized wear particles or the particles in the matrix
of the material may contribute to abrasive wear as well. In the systems operating in corrosive envi-
ronment, there is a synergistic action between corrosion and wear. Some lubricating oils are reactive
and produce a reaction product that may accelerate wear. Surface coating of hard materials are used
to reduce abrasive wear. If they crack in service, they will lead to enhanced wear due to spalling and
abrasion from the coating material itself. In many situations, the cause of failure is wear combined
with other failures. For example, the failure of die casting dies involves a complex interaction between
various failure modes. Here the impingement of molten jet stream produces erosion on die surfaces.
The severity of erosion depends upon the jet stream velocity and the impingement angle. Since the
hot metal may produce corrosion, the wear process is basically erosion–corrosion. Mechanical and
thermal stress fluctuations due to process cycle produce heat checking from thermal fatigue. The sud-
den temperature changes produce thermal shock, which results in microcracking and eventually total
failure of the die.
6-16 Theory and Practice of Lubrication and Tribology

6.9  Investigating Tribological Failures


The investigation of a tribological failure involves a number of steps that are briefly described as follows:
1. Gather the factual data.
a. Get the information about the kinds of motions involved and operating conditions such as
the load, speed, and temperature; sliding, rolling, or impact; dry or lubricated sliding; system
open or closed to atmosphere; the type of contact, sliding direction, and amplitude based on
the scratches in the scar, etc.
b. Compositions of the materials of contacting surfaces and their heat-treat condition, the lubri-
cant specifications if relevant.
c. The cause of failure such as normal operation, accidental overloads, lubrication failure, tem-
perature rise, etc.
2. Inspect the surfaces of mating members and draw preliminary conclusions.
a. Carry out visual and microscopic examination of the worn contact surfaces for characteristic
features and draw probable conclusions based on the information gathered in step 1. These
features are better seen in scanning electron microscopy (SEM) than in optical microscopy.
SEM also provides the capability for energy-dispersive x-ray analysis (EDXA), which can
identify the material transfer, if any.
The following are some of the features that are indicative of the kind of wear given against each:
i. Adhesive junctions and material transfer from the mating surface—adhesive wear
ii. Seizure and cratering—scuffing and scoring in lubricated sliding
iii. Extensive plastic deformation accompanied by transfer or displacement of material—
galling when contact pressure and/or temperature are high
iv. Grooving/plowing—two-body abrasion when one of the contact surfaces is harder and
rougher than the other and the abrasive particles of other materials are not present
v. Grooving/plowing with polishing/grinding effect—three-body abrasion when abrasive
particles are either generated during sliding or admitted into the contact area
vi. Abraded surface and pits with rounded edges in small-amplitude cyclic loading—­
fretting wear
vii. Flaky surface appearance produced by the impact of solid particles—ductile erosion
viii. Cracking and loss of material produced by the impact of solid particles—brittle erosion
ix. Surface pitting produced by the collapse of bubbles in liquid environment—cavitation
erosion
x. Loss of material from sliding surface when exposed to corrosive atmosphere—corrosive
wear
b. Examine the wear debris and see if it agrees with your analysis in 2(a).
3. Make computations for the following:
a. Wear coefficient from the wear data and operating conditions. See if the calculated value of the
wear coefficient falls within the range of the wear coefficient for the particular type of wear.
b. Temperature rise and see how it would have affected the wear process. A suitable value of the
coefficient of friction may be assumed from the literature data for the material pair and oper-
ating conditions.
4. Based on the information in items 1–3 and the consultation of literature, reach at the consensus
on the mode of wear.
5. Replicate the test in laboratory.
a. Develop or use a laboratory wear tester for the type of wear determined in item 4.
b. Decide the parameter values to be used in the laboratory test. The contact conditions should
match those of the failed part with respect to the temperature, contact pressure, surface rough-
ness, type of motion, and environment. For fast results, increase the frequency or velocity.
Wear 6-17

c. Determine the following: wear, coefficient of friction, temperature rise, wear particle shape
and size.
d. Examine the worn surfaces and see if they match the surfaces of the failed system.
6. Based on the microscopic examination, computations, and laboratory data arrive at the final
result.

References
1. Dowson, D., History of Tribology, 2nd edn., Professional Engineering Publishers, London, U.K., 1998.
2. Rabinowiz, E., Wear coefficients—metals, Wear Control Handbook, M.B. Peterson and W.O. Winer
(eds.), ASME, 1980, pp. 475–506.
3. Archard, J. F., Wear, interdisciplinary approach to friction and wear, Proceedings of NASA-Sponsored
Symposium, P. M. Ku (ed.), San Antonio, TX, November 28–30, 1967, pp. 267–333.
4. Yerramareddy, S. and Bahadur, S., The effect of laser surface treatments on the tribological behav-
ior of Ti-6Al-4V, Wear of Materials, K. C. Ludema and R. G. Bayer (eds.), American Society of
Mechanical Engineers, New York, 1991, pp. 531–540.
5. Krushchov, M. M., Principles of abrasive wear, Wear, 28, 1974, 69–88.
6. Mulhearn, T. O. and Samuels, L. E., The abrasion of metals: A model of the process, Wear, 5, 1962,
478–498.
7. Richardson, R. C. D., The wear of metals by hard abrasives, Wear, 10, 1969, 291–309.
8. Tylczak, J. H., Abrasive Wear, Friction, Lubrication and Wear Technology, ASM Handbook, Vol. 18,
ASM International, pp. 184–190.
9. Samuels, L. E., Polishing Wear, Friction, Lubrication and Wear Technology, ASM Handbook, Vol. 18,
ASM International, pp. 191–198.
10. Soderberg, S., Bryggman, U., and McCullough, T., Frequency effects in fretting wear, Wear, 110, 1986,
19–34.
11. Calhoun, S. P., Effect of metal hardness and surface finish upon fretting corrosion, Technical Report
No. 63-1835, Rock Island Arsenal, IL, May 28, 1963.
12. Iwabuchi, A., Kayaba, T., and Kata, K., Effect of atmospheric pressure on friction and wear of 0.45%
C steel in fretting, Wear, 91, 1983, 289–305.
13. Goto, J. and Buckley, D. H., The influence of water vapor on the friction behavior of pure metals dur-
ing fretting, Tribology International, 18, 1985, 237–245.
14. Higham, P. A., Stott, F. H., and Bethune, B., Mechanism of wear of the metal surface during fretting
corrosion of steel and polymers, Corrosion Science, 18, 1978, 3–13.
15. Hurricks, P. L., The mechanism of fretting—A review, Wear, 15, 1970, 389–409.
16. Waterhouse, R. B., Fretting, Treatise on Materials Science and Technology, D. Scott (ed.), Academic
Press, New York, 1978, pp. 259–286.
17. Waterhouse, R. B., Fretting Wear, Vol. 18, Friction, Lubrication, and Wear Technology, ASM
International, Materials Park, OH, 1992, pp. 242–256.
18. Finnie, I., Erosion of metals by solid particles, Wear, 3, 1960, 87–103.
19. Grant, G. and Tabakoff, W., Erosion prediction in turbo machinery resulting from environmental
solid particles, Journal of Aircraft, 12(5), 1975, 471–478.
20. Goodwin, J. E., Sage, W., and Tilly, G. P., Study of erosion by solid particles, Proceedings of the
Institution of Mechanical Engineers, 184, 1969–1970, 279–292.
21. Finnie, I. and McFadden, D. H., On the velocity dependence of the erosion of ductile metals by solid
particles at low angles of incidence, Wear, 48, 1978, 181–190.
22. Finnie, I., Wolak, J., and Kabil, Y., Erosion of metals by solid particles, Journal of Materials, 2(3), 1967,
682–700.
23. Goodwin, J. E., Sage, W., and Tilly, G. P., Study of erosion by solid particles, Transactions of the ASME,
Journal of Basic Engineering, 92, 1970, 619–626.
6-18 Theory and Practice of Lubrication and Tribology

24. Winter, R. E. and Hutchings, I. M., Solid particle erosion studies using single angular particles, Wear,
29, 1974, 181–194.
25. Yerramareddy, S. and Bahadur, S., The effect of laser surface treatments on the tribological behavior
of Ti-6Al-4V, Wear, 157, 1992, 245–262.
26. Hutchings, I. M., Little, J. A., and Ninham, A. J., Low viscosity erosion–corrosion of steels in a fluid-
ized bed, Proceedings Conference on Corrosion–Erosion–Wear of Materials at Elevated Temperatures,
A. V. Levy (ed.), Berkeley, CA, January 31–February 2, 1990, Paper 14, NACE International 1991.
27. Zhou, J. and Bahadur, S., Further investigations on the elevated temperature erosion–corrosion
of stainless steels, Proceedings of Conference on Corrosion–Erosion–Wear of Materials at Elevated
Temperatures, A. V. Levy (ed.), Berkeley, CA, January 31–February 2, 1990, Paper 13 NACE
International 1991.
28. Annual Book of ASTM Standards, Vol. 03.02, Wear and Erosion; Metal Corrosion, ASTM Standard
G 75-01: Standard test method for determination of Slurry Abrasivity (Miller Number) and Slurry
Abrasion Response of Materials (SAR Number), Copyright ASTM International, 100 Barr Harbor
Drive, West Conshohoken, Pennsylvania 19428.
29. Madsen, B. W. and Blickensderfer, R., A new flow-through slurry erosion wear test, Slurry Erosion:
Uses, Applications, and Test Methods, STP 946, J. E. Miller and F. E. Schmidt, Jr. (eds.), ASTM
International, West Conshohocken, PA, 1987, pp. 169–184.
30. Lynn, R. S., Wong, K. K., and Clark, H. McI., On the particle size in slurry erosion, Wear of Materials,
K. C. Ludema (ed.), ASME, Orlando, FL, 1991, pp. 77–82.
31. Clark, H. McI., Slurry erosion, Proceedings of Conference on Corrosion–Erosion–Wear of Materials at
Elevated Temperatures, A. V. Levy (ed.), Berkeley, CA, January 31–February 2, 1990, Paper 10.
32. Postlethwaite, J., The control of erosion–corrosion in slurry pipelines, Materials Performance, 26,
1987, pp. 41–45.
33. Annual Book of ASTM Standards, Vol. 03.02, Wear and Erosion; Metal Corrosion, ASTM Standard
G32-98: Standard test method for cavitation erosion using vibratory apparatus, Copyright ASTM
International, 100 Barr Harbor Drive, West Conshohoken, Pennsylvania 19428.
34. Preece, C. M., Cavitation Erosion, Treatise on Materials Science and Technology, Vol. 16—Erosion,
Academic Press, New York, 1979, pp. 249–308.
35. Simoneau, R. et al., Cavitation erosion and deformation mechanisms of Ni and Co austenitic stain-
less steels, 7th International Conference on Erosion by Liquid and Solid Impact, Cavendish Laboratory,
University of Cambridge, Cambridge, U.K., 1978.
36. Coulon, P. A., Erosion–corrosion in steam turbines. Part II: A problem largely resolved, Lubrication
Engineering, 42(6), 1968, 357–362.
37. Adler, W. F., The mechanics of liquid impact, Treatise on Materials Science and Technology, Vol. 16,
Erosion, C. M. Preece (ed.), Academic Press, New York, 1979, pp. 127–183.
38. Brunton, J. H. and Rochester, M. C., Erosion of solid surfaces by the impact of liquid drops, Treatise
on Materials Science and Technology, Vol. 16, Erosion, C. M. Preece (ed.), Academic Press, New York,
1979, pp. 185–248.
39. Symposium on erosion and cavitation, ASTM STP 307, ASTM International, West Conshohocken,
PA, 1962.
40. Symposium on erosion by cavitation and impingement, ASTM STP 408, ASTM International, West
Conshohocken, PA, 1967.
41. Fyall, A. A. and King, R. B. (eds.), Proceedings of the Third and Fourth International Conferences on
Rain Erosion and Allied Phenomena, Royal Aircraft Establishment, Bedford, England, 1970 and 1974.
42. Field, J. E. and Corney, N. S. (eds.), Proceedings of the Sixth International Conference on Erosion by
Liquid and Solid Impact, Cavendish Laboratory, University of Cambridge, Cambridge, England, 1983.
43. Field, J. E. and Dear, J. P. (eds.), Proceedings of the Seventh International Conference on Erosion by
Liquid and Solid Impact, Cavendish Laboratory, University of Cambridge, Cambridge, England,
1987.
Wear 6-19

44. Heymann, F. J., Toward Quantitative Prediction of Low Impact Erosion, ASTM STP 474, ASTM
International, West Conshohocken, PA, 1970, pp. 212–248.
45. Schmitt, G. F., Erosion Rate-Velocity Dependence for Materials at Supersonic Speeds, STP 474, ASTM
International, West Conshohocken, PA, 1970, pp. 323–352.
46. Heymann, F. J., Liquid Impingement Erosion, ASM Handbook, Vol. 18, Friction, Lubrication and Wear
Technology, ASM International, 1992, pp. 221–232.
47. Wear, 78 (1–2), 1982, 1–262.
48. Electrical Contacts and Electromechanical Components, J.-G. Zhang (ed.), International Academic
Publishers, Pergamon Press, Oxford, U.K., 1989.
49. Johnson, J. L., Sliding monolithic brush systems for large currents, Electrical Contacts-1986,
Proceedings 32nd IEEE Holm Conference on Electrical Contacts, October 1986, pp. 3–17.
50. Rabinowicz, E. and Chan, P., Wear of silver-graphite brushes against various ring materials at high-
current densities, IEEE Transactions, CHMT-3(2), 1980, 288–291.
51. Shobert II, E. I., Carbon Brushes: The Physics and Chemistry of Sliding Contacts, Chemical Publishing
Co., New York, 1965, pp. 11–12.
52. Konchits, V. V., Wear in the sliding contact of electrical machines, I, Soviet Journal of Friction and
Wear, 7(1), 1986, 90–96.
53. Garshasb, M. and Vook, R. W., Fundamental analysis of Cu brush-Ag slip ring sliding electrical con-
tacts, IEEE Transactions, CHMT-9(1), 1986, 23–29.
54. Reichner, P., High current test of metal fiber brushes, I, IEEE Transactions, CHMT-4(4), 1981, 1–4.
55. Burton, R. A. and Burton, R. G., Vitreous carbon matrix for low wear carbon/metal current collec-
tors, Electrical Contacts-1988, Proceedings 34th Holm Conference on Electrical Contacts, September
1988, pp. 223–228.
56. Pitney, K. E., Ney Contact Manual, Electrical Contacts for Low Energy Uses, The J. M. Ney Company,
Bloomfield, CT, 1973, p. 14.
57. Michal, R. and Saeger, K. E., Application of silver-based contact materials in air-break switching
devices for power engineering, Electrical Contacts-1988, Proceedings 34th IEEE Holm Conference on
Electrical Contacts, 1988, pp. 121–127.
58. Gray, E. W., Voltage fluctuations in low current atmospheric arcs, Journal of Applied Physics, 43(11),
1972, 4573–4575.
59. Sone, H. and Takagi, T., Role of the metallic phase arc discharge on arc erosion in Ag contacts,
Electrical Contacts-1989, Proceedings 34th IEEE Holm Conference on Electrical Contacts, 1989,
pp. 157–161.
60. Shen, Y. S. and Gould, L. J., Erosion modes of internally oxidized Ag-CdO and Ag-(Sn,In) O material,
Electrical Contacts-1987, Proceedings of 33rd IEEE Holm Conference on Electrical Contacts, 1987,
pp. 157–161.
61. Donaldson, A. L., Kristiansen, M., Watson, A., Zinsmeyer, K., and Kristiansen, E., Electrode erosion
in high current, high energy transient arcs, IEEE Transactions on Magnetics, 22(6), 1986, 1441–1447.
62. Liu, P., Bahadur, S., Verhoeven, J. D., Gibson, E. D., and Kristiansen, M., Arc erosion behavior of
Cu–15% Nb and Cu–15% Cr in-situ composites, Wear, 203–204, 1997, 36–45.
7
Adhesive Wear
7.1 Introduction........................................................................................7-1
7.2 Phenomenological Aspects of Adhesive Wear.............................. 7-2
7.3 Factors Affecting Adhesive Wear.................................................... 7-3
7.4 Simple Theory of Adhesive Wear.................................................... 7-4
7.5 Wear Coefficient................................................................................. 7-8
7.6 Alternative Forms of the Adhesive Wear Equation................... 7-12
7.7 Formation of Loose Wear Particles...............................................7-14
Kyriakos 7.8 Evolution of Surface Roughness and Minimum Clearance
Komvopoulos in Sliding Systems.............................................................................7-18
University of 7.9 Closing Remarks.............................................................................. 7-20
California, Berkeley References..................................................................................................... 7-20

7.1  Introduction
Mechanical components cease to be useful due to three main reasons: (1) obsolescence due to changes
in design, material shortage, and various economic or safety factors, (2) irreversible gross deformation
and fracture due to stresses exceeding the bulk strength of the material, and (3) surface degradation due
to the loss of material as a result of wear and corrosion. Although there is not much to say about obso-
lescence, and bulk deformation and fracture are relatively well understood, loss of material due to wear
and corrosion continues to be an enormously expensive problem to society. While both wear and corro-
sion involve complex phenomena, understanding of corrosion is significantly more advanced compared
to wear. A concise definition of wear is the removal of material from solid surfaces due to the effect of
mechanical action [1]. In the past, there was less interest in wear because worn parts were discarded and
replaced by new ones. This notion has changed in more recent years because of the dramatic decrease
in tolerances of mechanical systems, advances in precision manufacturing, and development of in situ
diagnostic methods. In today’s industrialized world, wear is a limiting factor in most machines, equip-
ment, and manufacturing processes. Reducing wear (and friction) has also important implications in
energy savings and environmental pollution.
The most common type of wear is adhesive wear. This process of material loss may be encountered
between contacting metallic, ceramic, and polymeric materials, or combinations of these, and its sever-
ity depends on the presence of surface films, such as oxide, contaminant, and boundary lubricant films,
as well as the mutual affinity of one material surface for the other. Traditional studies of adhesive wear
have been mainly based on empirical methods that rely on weight loss measurements and qualitative
evaluation of surface damage by various microscopic techniques. While such methods might have
practical implications, they provide limited information about the factors responsible for the loss of
material. Important insight into adhesive wear was obtained with the development of scanning elec-
tron microscopy (SEM) and various microanalysis techniques, such as x-ray diffraction, Auger electron
spectroscopy, and x-ray photoelectron spectroscopy. These high-resolution techniques have provided

7-1
7-2 Theory and Practice of Lubrication and Tribology

invaluable information about changes in the morphology, microstructure, and composition of sliding
surfaces induced by the rubbing process, the nature and size of loose wear particles, and the structure
and composition of material adhered to the counter surface. These techniques have played a crucial role
in the understanding of the origins of adhesive wear and quantification of the loss of material in terms
of various factors controlling this type of wear.
It is generally accepted that the formation and rupture of asperity junctions at sliding interfaces are
precursors of material loss in the form of relatively small particles. These wear particles can transfer
back and forth between the rubbing surfaces before they become loose particles, which may be trapped
at the sliding interface or pushed out of the sliding track. Wear particle formation by adhesive wear
results from plastic shearing of asperity junctions. The fraction of asperity junctions that shear below
the contact interface depends on the work of adhesion of the sliding system, including environmental
factors controlling the formation of interfacial solid and liquid films. The severity of adhesive wear
depends on the cleanness of the surfaces, the magnitude of the load transferred across the contact inter-
face, the surface resistance to plastic deformation, and the ambient conditions of the sliding system.
It is not uncommon for the wear rate to change sharply due to variation in any of the former factors.
Adhesive wear is observed whenever two solid bodies slide against each other or undergo repetitive
normal contact. Therefore, basic understanding of its origins and identification of the important fac-
tors affecting the loss of material by this process are of paramount importance in the design of durable
mechanical systems. The objective of this chapter is to provide a basic background on adhesive wear and
illustrate practical use of the presented information through simple examples.

7.2  Phenomenological Aspects of Adhesive Wear


Numerous studies of adhesive wear suggest that material is removed in the form of lumps or fragments
from asperity microcontacts undergoing internal plastic shearing in the presence of strong interfacial
adhesive forces. Attractive forces due to chemical (e.g., covalent, ionic, metallic, and hydrogen bonds)
and physical (e.g., van der Waals forces) interactions lead to the establishment of adhesive asperity junc-
tions. In the case of weak asperity bonding (e.g., contaminated or well-lubricated surfaces), shearing
occurs along the contact interface of asperity junctions and asperity separation does not involve the loss
of material (path A, Figure 7.1a). However, if the interfacial bond strength exceeds the cohesive strength

Sliding direction
∆L
∆F Path A
(1)
(2)
(a) Interfacial shearing No wear

Sliding direction
∆L
∆F Worn asperity
(1) Path B

(2)
(b) Plastic shearing Material transfer-wear

FIGURE 7.1  Schematic illustration of asperity-scale surface interaction during sliding contact. An asperity of
surface (1) slides against another asperity of the opposed surface (2). The two possible scenarios are (a) shearing
along the contact interface (path A) and (b) shearing within one of the two asperities (path B), most likely the
weaker asperity, resulting in material transfer to asperity (1) and, thus, material loss (wear) from surface (2).
Adhesive Wear 7-3

10 µm 10 µm
(a) (b)

FIGURE 7.2  SEM micrographs of wear tracks on slid steel surfaces showing typical features of (a) mild and
(b) severe adhesive wear.

of one of the two asperities, shearing occurs below the contact interface of the particular asperity, result-
ing in the transfer of a fragment to the other asperity (path B, Figure 7.1b). Loss and transfer of mate-
rial to the opposed surface during sliding have been observed in several radiographic and microscopic
studies [1–4].
Material loss from interacting surfaces due to the existence of strong adhesive forces leads to the for-
mation of surface morphologies characterized by small cavities, resulting from the extraction of mate-
rial that was strongly adhered to the opposed surface. The SEM micrographs shown in Figure 7.2 reveal
typical wear futures on steel surfaces that exhibited material loss due to adhesive wear. Under conditions
of mild adhesive wear, the worn surface shows mild roughening due to the nonuniform pullout of mate-
rial and the formation of small wear particles (Figure 7.2a). Alternatively, severe adhesive wear leads
to significant surface roughening due to the formation of large craters (Figure 7.2b). A transition from
relatively mild to severe adhesive wear is usually observed with increasing normal load and interfacial
adhesive strength. The latter may be encountered upon the removal of contaminant and native oxide
films, desorption of lubricious substances (boundary lubricant films), or rupture of solid films possess-
ing low surface energy. Hence, surface morphologies containing craters and formation of lump-like
wear debris are intrinsic features of adhesive wear.

7.3  Factors Affecting Adhesive Wear


Even nominally flat surfaces demonstrate multi-scale roughness at high magnifications. Three-
dimensional topographies of macroscopically smooth surfaces comprise peaks and valleys randomly
distributed over the apparent area of contact (Figure 7.3a). When two such surfaces are pressed together,
actual contact occurs between a few of these asperities and the resulting morphology resembles island-
like distribution (Figure 7.3b). Increasing the global interference (load) leads to the formation of new
asperity junctions and simultaneous growth of established microcontacts. If the surfaces are then sepa-
rated (either in the normal or tangential direction), material may be pulled out from either surface and
transfer to the opposed surface in the presence of attractive (adhesive) forces at the contact interface. The
amount of pulled-out material depends on the strength of the contacting surfaces and the magnitude of
the interfacial adhesion forces.
Early models of adhesive wear evolved from a basic understanding of the effect of surface adhesion
forces on friction [5,6]. Holm [7] was the first to point out that the volume of material loss V due to
7-4 Theory and Practice of Lubrication and Tribology

10

4 8

z (µm) 2 6

y (µm)
0
–2 4
–4
10 2
7.5 10
7.5
y (µ 5 2.5 2.5
5 0
m) )
0 x (µm 0 2 4 6 8 10
(a) (b) x (µm)

FIGURE 7.3  (a) Three-dimensional surface topography showing multi-scale roughness and (b) formation of
microcontact areas due to the truncation of the rough surface by a rigid plane. Each microcontact area is approxi-
mated by a circle of equal area.

adhesive wear is a function of the applied normal load L, total distance of sliding S, and hardness of the
worn material H, that is,

LS
V =c (7.1)
H

where c is a dimensionless constant. Archard [8] is credited for formulating the concept of adhesive
wear, according to which most of the asperity junctions established between the interacting surfaces
shear along their common interface without resulting in material loss, with only a small fraction shear-
ing internally to produce wear particles. Archard argued that the constant c in Equation 7.1 represents
one-third of the probability k of forming a hemispherical wear particle from an asperity junction formed
over a distance equal to its diameter, and reformulated Equation 7.1 as

LS
V =k (7.2)
3H

The new factor k is usually referred to as the adhesive wear coefficient, and typically varies in the range
of 10−8 to 10−3. (From a probability perspective, the maximum value of k is 1; however, such high wear
coefficient cannot be obtained with any type of wear process.) In recent tabulations, the factor 3 in the
denominator of Equation 7.2 was combined with k to obtain a wear relationship similar to that given by
Equation 7.1.

7.4  Simple Theory of Adhesive Wear


When two surfaces are in sliding contact, one or both surfaces will ultimately undergo adhesive wear.
A simple theoretical treatment of adhesive wear based on the original model of Archard [8] is presented
in this section. The analysis accounts for the most important variables affecting adhesive wear, such as
applied normal load, total distance of sliding, and hardness of the worn material, and provides a means
of assessing the severity of adhesive wear in terms of the magnitude of the wear coefficient.
A fundamental concept in adhesive wear is the real area of contact. As mentioned earlier, even
nominally flat surfaces exhibit multi-scale roughness. Thus, when two real surfaces are brought into
contact, solid–solid contact is confined between surface asperities, hereafter referred to as asperity
microcontacts. The sum of all microcontact areas represents the real area of contact Ar, which is a
very small fraction of the apparent area of contact Aa . Typically, Ar/Aa varies in the wide range of
Adhesive Wear 7-5

10−8 to 10−5, depending on the topography and material properties of the interacting solid surfaces
and the applied normal load; however, higher values on the order of 10 −1 may be encountered under
special circumstances, such as in some heavy-load manufacturing processes (e.g., extrusion, forging,
and stamping).
Assuming approximately circular microcontact areas (Figure 7.3b), the real area of contact at any
instant of sliding under a given normal load can be written as

n
Ar = ∑a
i =1
i (7.3)

where
n is the number of instantaneous microcontacts
ai is the area of the ith microcontact

The fraction of the normal load ΔLi transmitted through a microcontact area ai can be expressed as

ΔLi = pm ai (7.4)

where pm is the mean contact pressure, bounded by the hardness of the softer surface H.
Considering that the total applied load L is the sum of the load fractions transmitted through all the
microcontacts and in view of Equations 7.3 and 7.4, it follows that

L = HAr (7.5)

Equation 7.5 gives a lower bound of the real area of contact, equal to the ratio of the applied load and
the material hardness, because all microcontacts are presumed to be plastically deformed (pm = H).
It will be shown later that, when other uncertainties are taken into consideration, Equation 7.5 is a
good approximation of the load–area relationship, especially if the majority of microcontacts deform
­plastically. Equation 7.5 can be further modified to include the effect of strain hardening on the hard-
ness of the sliding surface using the following relationship [9]:

H = 3C − m SY (7.6)

where
C is a strain constant (typically equal to 0.1 for steel and aluminum)
m is the strain hardening exponent
SY is the yield strength

The proportionality between real area of contact and normal load reflected by Equation 7.5 has been
proven experimentally, including results showing a linear variation of the electrical contact conduc-
tance, which is proportional to the real area of contact, with the applied normal load [2].
Assuming an average size of microcontact area of equivalent diameter dm, Equation 7.3 can be simpli-
fied to the following equation:

pdm2
Ar ⊕n (7.7)
4
7-6 Theory and Practice of Lubrication and Tribology

From Equations 7.5 and 7.7, it follows that

4L
n⊕ (7.8)
pHdm2

If each microcontact is assumed to remain in existence for a sliding distance equal to its diameter di,
after which it is replaced by another microcontact, and because n has been defined as the average num-
ber of microcontacts at any given instant of sliding, the total number of microcontacts N formed during
the total distance of sliding S can be written as

S
N =n (7.9)
dm

The size of a wear particle may be assumed to scale to the size of the microcontact area from which it
was produced. Thus, assuming wear particles of hemispherical shape, the volume of a wear particle of
diameter di is dVi = pdi3 /12. Consequently, the total wear volume can be expressed as

N N
pdi3 pdm3
V= ∑
i =1
dVi = ∑ 12
i =1
≈N
12
(7.10)

Substitution of Equations 7.6, 7.8, and 7.9 into Equation 7.10 yields

LS C m ⎛ LS ⎞
V= = (7.11)
3H 3 ⎜⎝ 3SY ⎟⎠

Equation 7.11 is in qualitative agreement with Holm’s relationship (Equation 7.1) and experimental
results showing that the volume of material loss due to adhesive wear is proportional to the applied
normal load, total distance of sliding, and reciprocal of the hardness of worn material [10,11]. However,
Equation 7.11 suggests that all microcontacts established during the sliding period contribute to the
total wear volume, which is unrealistic because many microcontacts may either deform elastically or
shear along their contact interfaces without producing wear particles (Figure 7.1a). Archard [8] intro-
duced the idea of the probability of a microcontact to generate a wear particle k and expressed the total
wear volume as

LS LS
V =k =K (7.12)
3H H

where K (=k/3) is also referred to as the adhesive wear coefficient, since factor 3 is often omitted in wear
coefficient calculations. It is noted that Equation 7.12 does not contain the average microcontact diam-
eter dm. In fact, Archard [8] showed that Equation 7.12 applies for a wide range of microcontact area
and wear particle size. The factor 3 in Equation 7.12 is due to the assumption of hemispherical wear
particles and, therefore, can be considered as a shape factor. Because the range of variation in shape
factor is very small compared to that of the wear coefficient [1,7,8,12–14], factor 3 is usually neglected
in the wear equation as secondary. This explains the frequent use of K in Equation 7.12 instead of k.
For simplicity, the second form of Equation 7.12 will be used hereafter, referring to K as the adhesive
wear coefficient.
Adhesive Wear 7-7

Quantification of adhesive wear in terms of the magnitude of K is useful for comparing the steady-
state wear behavior of different types of materials. The proportionality between wear volume and slid-
ing distance (or time, for constant sliding speed) shown by Equation 7.12 is usually observed after the
initial period of sliding, known as the run-in stage. During this transient period, the wear rate fluctuates
continuously until equilibrium conditions are established at the sliding interface. Abrupt transitions of
the wear rate (wear coefficient) may be encountered with load variations. Although such transitions may
appear to contradict Equation 7.12, they are actually due to the evolution of different sliding systems.
This is due to changes in the physicochemical properties of the surfaces during sliding, resulting in
surfaces possessing different tribological characteristics. For example, light-load sliding between metal
surfaces normally produces low K values because the tribological properties are controlled by native
oxide films present at the sliding interface, which can sustain the low tractions produced by a relatively
light load. Alternatively, heavy-load sliding of metal surfaces leads to rupture of the oxide films (and any
adsorbed layers) during the run-in period, and steady-state sliding occurs between the freshly exposed
metal surfaces. Because clean metal surfaces exhibit much higher affinity for each other compared to
oxidized surfaces, higher K values are obtained with higher loads. For fixed normal load, the magnitude
of K is controlled by the affinity of the interacting surfaces for each other, which depends on their com-
patibility (Figure 7.4). In general, the higher the solid solubility limit, the higher the wear coefficient.
For viscoelastic materials, such as elastomers, time-dependent deformation does not allow for the
indentation hardness to be uniquely defined. Therefore, to evaluate the wear performance of these
materials it is preferred to use the wear factor, defined as K' = V/LS or K' = (V/t)/Lv. The maximum
value of the product of the normal load and the sliding speed that a polymer surface can sustain before
undergoing softening, melting, or oxidation is referred to as the Lv limit and is an intrinsic property
of a polymeric material. In general, polymers characterized by a high Lv limit also demonstrate a high
wear resistance.

W Mo Cr Co Ni Fe Nb Pt Zr Ti Cu Au Ag Al Zn Mg Cd Sn Pb In
In
Pb
Sn
Cd
Mg
Zn
Al
Ag
Au
Cu
Ti
Zr
Pt Metallurgical Metallurgical Sliding Anticipated
Nb Symbol solubility compatibility compatibility wear

Fe 100% Identical Very poor Very high


Ni Above 1% Soluble Poor High
0.1 % — 1% Intermediate Intermediate Intermediate
Co soluble
Cr Below 0.1% Intermediate Intermediate Intermediate
insoluble or good or low
Mo Very low
Two liquid Insoluble Very good
W phases

(a) (b)

FIGURE 7.4  (a) Material compatibility chart based on the (b) mutual solubility of materials derived from binary
phase diagrams. (From Rabinowicz, E., ASLE Trans., 14, 198, 1971. With permission.)
7-8 Theory and Practice of Lubrication and Tribology

Example 7.1
A Cr coating of thickness h and hardness H is deposited onto a flat metal substrate. A ball-on-
flat tribometer operated in unidirectional rotary mode is used to examine the durability of the
overcoat. Two smooth ball bearings of identical roughness and diameter, a stainless steel and an
Al2O3 bearing, are to be used as counter-surface specimens. In all sliding cases, the normal load
L is fixed and the state of lubrication is classified as poor. Assuming insignificant changes in the
width w of the circular wear track due to the type of bearing used, determine which bearing will
wear out the Cr overcoat in fewer sliding cycles and what will happen if an Al2O3 overcoat of the
same thickness is used instead of the Cr overcoat. (Note: Assume gradual wear of the overcoat
by adhesive wear.)

Solution
The wear volume of overcoat material to be removed is V = 2πRwh, where R is the radius of
the circular wear track. The total sliding distance can be expressed as S = 2πRn, where n is the
number of sliding cycles to remove the overcoat from the sliding track. Substituting the former
expressions of V and S into Equation 7.12, we obtain

L(2pRn)
2pRwh = K
H

where K is the adhesive wear coefficient. From the previous equation, we find that

1 ⎛ whH ⎞
n= ⎜ ⎟
K⎝ L ⎠

The only varying parameter in the former equation is the wear coefficient. Therefore, we only
need to compare the wear coefficients of the following four systems: stainless steel/Cr (K1),
Al2O3/Cr (K2), stainless steel/Al2O3 (K3), and Al2O3/Al2O3 (K4), all under poor lubrication
conditions.
As mentioned earlier, K is controlled by the metallurgical affinity (solid solubility limit) of the
sliding surfaces (Figure 7.4). Stainless steel/Cr represents a metal–metal system. To determine
the compatibility of the sliding surfaces, we examine the mutual solubility of Cr in Fe, which
according to Figure 7.4 is >1%. Thus, the surfaces are classified as compatible, and from Figure
7.6 we obtain K1 = 10−4. Al2O3/Cr is a nonmetal–metal system of incompatible surfaces. Hence,
from Figure 7.6 it is found that K2 = 3 × 10−6. Stainless steel/Al2O3 is also a metal–nonmetal sys-
tem of incompatible materials; thus, K3 = 3 × 10−6. Finally, Al2O3/Al2O3 is a nonmetal–nonmetal
system of identical materials, for which Figure 7.6 gives K4 = 3 × 10−4. Thus, K4 > K1 > K2 = K3, and
since n ~ K−1, it follows that n4 < n1 < n2 = n3. Hence, the stainless steel bearing will wear out the
Cr overcoat faster than the Al2O3 bearing, whereas the opposite will occur if the Cr overcoat is
replaced by an Al2O3 overcoat of the same thickness.

7.5  Wear Coefficient


The voluminous literature of adhesive wear coefficients suggests that in addition to the load, sliding
distance, and material hardness, which are explicitly introduced in Equation 7.12, there are other
important factors affecting adhesive wear, such as the affinity of one surface for the other, which
depends on the metallurgical compatibility of the sliding surfaces and the cleanness of the contact
interface. The affinity of a material surface for another can be described by the work of adhesion W,
defined as the work needed to separate two contacting surfaces of these materials, given by [1]
Adhesive Wear 7-9

W12 = g1 + g2 − g12 ≈ cm cl (g1 + g2 ) (7.13)

where
γ1 and γ2 are the surface energies of interacting surfaces (1) and (2), respectively
γ12 is the interface energy
cm is a coefficient indicative of the metallurgical compatibility of the contacting surfaces
cl is a coefficient that depends on the state of lubrication or the environment of the sliding surfaces,
for example, vacuum, air, or lubricant

The product cmcl, known as the compatibility index, assumes values between 1.5 (clean surfaces of the
same metal sliding in vacuum) and 0.016 (dissimilar material surfaces, that is, two liquid phases, sliding
under excellent lubrication conditions). The wide ranges of compatibility index and surface energy sug-
gest that the work of adhesion can vary by 3–4 orders of magnitude, depending on the solid solubility
limit (metallurgical affinity effect) of the sliding surfaces and the presence (or absence) of lubricious
films at the sliding interface (environment effect).
Since the work of adhesion reflects the interfacial bond strength of two contacting surfaces, it should
correlate with the adhesive wear coefficient. Indeed, experimental evidence confirms that sliding sys-
tems exhibiting high work of adhesion also demonstrate high wear coefficient. Thus, the variation of the
adhesive wear coefficient in the wide range of 10−3 to 10−8 can be attributed to the effects of metallurgical
compatibility and state of lubrication, and is in line with the variation of the work of adhesion by several
orders of magnitude, depending on the compatibility index and surface energies of the interacting sur-
faces. In general, higher wear coefficients are produced for metals sliding against metals than nonmetals
sliding against nonmetals or metals [13–15]. Further classification in K values can be obtained in terms
of the metallurgical compatibility, determined from the equilibrium phase diagram of the materials of a
sliding system (solid solubility limit). Rabinowicz [15] considered five levels of mutual solubility of pure
metals and developed the compatibility chart shown in Figure 7.4. In order of decreasing mutual solu-
bility, the five classes of compatibility are identical metals (100% soluble), compatible (solubility > 1%),
intermediate compatible I (0.1% < solubility < 1%), intermediate compatible II ­(solubility < 0.1%), and
incompatible (solubility = 0%, two liquid phases). In general, sliding metal surfaces demonstrating high
compatibility are characterized by high K values.
Metallurgical compatibility is not the only factor affecting friction and wear of sliding surfaces. Since
the tribological behavior also depends on the environment where surface sliding occurs, changes in the
physicochemical properties of the sliding surfaces due to environmental factors could have a signifi-
cant effect on wear behavior. For example, metal sliding under atmospheric conditions may promote
continuous replenishment of adsorbent contaminant layers and rapid formation of oxide films upon
rupture, maintaining surfaces of low surface energy and, in turn, low wear. Figure 7.5 shows the effects
of metallurgical compatibility and environment on the compatibility index cmcl of two classes of slid-
ing systems—metal sliding against metal (Figure 7.5a) and nonmetal sliding against metal or nonmetal
(Figure 7.5b). The graphs were constructed using compatibility index data obtained by Rabinowicz. For
both types of material systems, cmcl tends to decrease with decreasing mutual solubility and improving
lubrication.
Since the two most important factors affecting adhesive wear are the affinity of the two surfaces
for each other (indicated by their metallurgical compatibility) and the degree of cleanness (or state of
lubrication) at the sliding interface, the variation of the wear coefficient in terms of these factors, shown
in Figure 7.6, is of particular importance. The wear coefficient data used to construct Figure 7.6 were
obtained from a statistical analysis of wear coefficient values of various sliding surfaces [13,14]. The
figure shows that mutual solubility and lubrication state can change the adhesive wear coefficient by sev-
eral orders of magnitude, and that compatibility and lubricity are continuous functions. Environmental
effects (e.g., oxidation, corrosion, and adsorption/desorption of contaminant monolayers) and sliding
7-10 Theory and Practice of Lubrication and Tribology

1.50
1.20
1.0 1.00 1.00
1.0
0.70
Compatibility index, cmcl

Compatibility index, cmcl


0.5 0.50 0.70 0.62
0.45
0.60
0.32 0.37 Ide Iden
0.27 nti tica
Co cal 0.5 0.40 l
0.2 0.20 0.21 mp 0.40 Sim 0.39
0.16 Int ati ilar
er ble 0.14 0.36
0.14
0.1
0.12 Int media 0.08
0.27
Dissim
erm te I 0.27
0.096 e 7 ilar
Inco diate 0.06 0.25
mpa II 3 0.20
0.05 0.063 t ible 0.04 0.050 0.2 0.19
6
0.03 0.036
2 0.15
0.028
0.02 0.022
0.016 0.1

In Clean Poor Fair Good In Clean Poor Fair Good


vacuum in air lubricant Iubricant Iubricant vacuum in air lubricant lubricant lubricant
(a) Environment (b) Environment

FIGURE 7.5  Dependence of compatibility index of (a) metal–metal and (b) nonmetal–metal or nonmetal sliding
systems on environment.

10–2
–3 Severe wear
10
Ide
ntic
al
Wear coefficient

–4
10 Com
pat Intermediate
Inte ible
–5 rme wear
10 Inte
rme
dia
te I
dia and
te II
–6
10 Inco
mp
atib
le
–7
10 Low wear
–8
10
Clean Poor Average Excellent
State of lubrication

FIGURE 7.6  Dependence of adhesive wear coefficient on material compatibility and state of lubrication.

conditions (e.g., load, sliding speed [shear rate], and frictional heating) affecting lubrication by bound-
ary films may lead to significant fluctuations in wear rate during the operation of mechanical systems
possessing sliding interfaces.
The results shown in Figure 7.6 can be used to estimate the wear coefficient of metal–metal and
nonmetal–metal or nonmetal sliding systems under different states of lubrication. Caution should be
exercised in using these data for that they are applicable only when adhesive wear is the dominant
process of material loss. The metal–metal category includes pure metals and alloys sliding against
other metals or alloys. The metallurgical compatibility of alloys is determined by the element with the
highest concentration in the alloy composition. However, if no element in the alloy has a concentration
above 60%, then that alloy sliding on itself should be classified as compatible rather than identical [13].
Composite materials consisting of metals and nonmetals can also be classified based on the material
with the highest concentration. In cases of multiple-phase materials consisting of two or more ele-
ments of comparable concentrations, it is best to determine a wear coefficient for each of the dominant
elements in the composition and then compute the wear coefficient as the geometric average of the
calculated values.
Regarding the state of lubrication, “clean” implies surfaces thoroughly cleaned by some combination
of mechanical (polishing) and chemical (solvent) processes, which are sliding against each other in air
Adhesive Wear 7-11

in the absence of any adsorbing species or boundary lubricant layer. Under “poor” lubrication condi-
tions, we include surfaces sliding in the presence of a liquid that does not have lubricating properties
(e.g., water and ethyl alcohol). In general, a lubricant layer is characterized as “good” if it can adsorb onto
the sliding surfaces to form a fairly continuous, self-replenishing layer that reduces direct solid–solid
contact over a large fraction of the real area of contact (e.g., 70%–80%); if this fraction is above 95%, then
lubrication is considered to be “excellent.” Since the efficacy of a boundary lubricant layer depends on
its strength of adsorption, lubricant molecules that react chemically (chemisorption) with the surfaces
usually produce better lubrication conditions than relatively inert lubricants (e.g., naphthenic hydrocar-
bons) attached to the surfaces by van der Waals forces (physisorption). Thus, “good” lubrication includes
typical petroleum-based liquids and organic synthetic lubricants (e.g., chlorofluorocarbons and poly-
alkylene glycols), whereas “excellent” lubrication includes petroleum-based oils enriched with special-
compound additives (e.g., molybdenum disulfide and zinc dialkyl dithiophosphate).
In addition to the strength of adsorption, other factors affecting the state of lubrication in a tribosys-
tem are surface traction (both normal and shear), sliding velocity, surface roughness, molecular chain
length, and polarity. Surface tractions producing interfacial stresses that exceed the adsorption strength
of lubricant molecules and sliding speeds resulting in frictional heating that raises the surface tem-
perature above the transition temperature of the lubricant (on average about 50°C, 150°C, and 220°C
for poor, good, and excellent lubricants, respectively) are detrimental to the lubrication efficacy. The
degradation of lubricity may be understood as a downward shift in the state of lubrication (Figure 7.6).
Long-chain molecules with polar groups that exhibit strong affinity for the sliding surfaces are essential
for the formation of closely packed, conformal boundary layers that can minimize solid–solid contact
between rough surfaces. Therefore, changes in the aforementioned factors may affect adhesive wear
significantly.
The severity of adhesive wear can be characterized by the range of the wear coefficient. As shown
in Figure 7.6, a sliding system can operate under conditions resulting in low (K < 10−6), intermediate
(10−6 < K < 10−4), or severe (K > 10−4) adhesive wear. It is noted that it is difficult to operate identical or
compatible material surfaces under low-wear conditions even when lubrication can be classified as
excellent. However, very low wear can be achieved with incompatible materials, even when the lubri-
cation efficacy is below average. For very low wear coefficients (10−7 to 10−8), material loss occurs at
extremely small scales, perhaps on the atomic scale. Because these surfaces appear as burnished, this
extremely low adhesive wear is also known as burnishing wear. This type of wear is critical to mirror
polishing of bearings, planarization of wafers, and lapping of ceramic heads used in hard-disk drives.
In the intermediate wear range, material is removed in the form of wear particles ranging from about
1 to 100 μm (Figure 7.1a), whereas in the severe wear range, the surface appearance is dominated by
large cavities (Figure 7.1b) from which material has been pulled out as a result of strong adhesive forces.
Severe adhesive wear, also known as scoring or scuffing, is normally encountered during the final stage
of the operation life of a sliding system, and is characterized by the formation of large wear particles,
squeaking, and excessive vibration.

Example 7.2
A manufacturer considers changing the microstructure of a ceramic composite used to fabricate
cutting tools in order to improve the toughness and thermal shock resistance. The proposed
solution is to cement the brittle carbide grains with Co. The initial composite consists of 80%
Al2O3 and 20% ZrO2, while the new material will be made of 65% Al2O3, 25% TiC, and 10% Co
(all percentages are vol%). If the tools are used to machine aluminum alloys, determine if the
new composite material will exhibit better tribological properties than the original material
from the work of adhesion standpoint. (Note: The surface energy of Al2O3, ZrO2, TiC, Co, and
Al is equal to 0.74, 0.53, 0.9, 1.8, and 0.9 J/m2, respectively.)
7-12 Theory and Practice of Lubrication and Tribology

Solution
The surface energy of the work material can be approximated by that of Al, which is the ele-
ment with the highest concentration in the workpiece microstructure. However, because the
tool material is a composite, its surface energy will be obtained from the rule of mixtures,
which gives the property X of a composite material in terms of the same property of each
constituent Xi:

X= ∑a ⋅ X
i i

where ai is the volume fraction of the ith element in the composite.


Substituting the given volume fractions and the known surface energies in the previ-
ous equation, the surface energies of the initial (Al 2O3 –ZrO2) and new (Al 2O3 –TiC–Co) tool
materials are

ginitial = ∑ a g = 0.8 × 0.74 + 0.2 × 0.53 = 0.698 J/m


i i
2

i =1

gnew = ∑ a g = 0.65 × 0.74 + 0.25 × 0.9 + 0.1 × 1.8 = 0.886 J/m


i i
2

i =1

The tool (ceramic composite) and the workpiece (Al alloy) constitute a nonmetal–metal system
of incompatible (dissimilar) surfaces. (The presence of 10% Co in the new tool is too low to alter
the chemical character of the ceramic tool.) For dissimilar materials and machining under fair
lubrication conditions, Figure 7.5b gives cmcl = 0.2. Thus, the work of adhesion corresponding to
the initial (Winitial) and new (Wnew) tool-workpiece system is

Winitial = 0.2(ginitial + gAl ) = 0.2(0.698 + 0.9) = 0.3196 J/m2


Wnew = 0.2(gnew + gAl ) = 0.2(0.886 + 0.9) = 0.3572 J/m 2


It is noted that the higher the work of adhesion, the less effective the cutting process. This is
because more energy is dissipated to combat adhesion (friction) at the tool–workpiece contact
interfaces (tool rake and wake faces). In addition, frictional heating intensifies with increas-
ing work of adhesion due to the greater propensity of the tool and workpiece surfaces to inter-
act with each other. Consequently, from a tribological perspective, the new composition is less
desirable than the original composition because Wnew > Winitial. However, the higher (~12%) work
of adhesion of the new tool material may be an acceptable compromise because the incorpora-
tion of Co will enhance the tool toughness and thermal shock resistance. Caution should be
exercised in determining the state of lubrication. The effect of the composition change on the
work of adhesion and, in turn, the cutting efficiency becomes more pronounced under poor
lubrication conditions. Therefore, the effect of the change in tool composition on the state of
lubrication should not be overlooked.

7.6  Alternative Forms of the Adhesive Wear Equation


Equation 7.12 can be used to determine the recess rate of a surface due to the gradual loss of material by
adhesive wear. This is of particular interest in applications where maintaining tight tolerances is critical
Adhesive Wear 7-13

for effective operation, such as journal bearings, piston–cylinder interface, and pump systems. If adhe-
sive wear occurs over a given apparent area of contact Aa, the wear volume can be expressed as V = Aah,
where h is the thickness of the material removed from the sliding surface of interest. Thus, for constant
sliding speed v, Equation 7.12 can be written as

⎛ L ⎞ ⎛ vt ⎞ ⎛ v⎞
h = K ⎜ ⎟ ⎜ ⎟ = Kpa ⎜ ⎟ t (7.14)
⎝ Aa ⎠ ⎝ H ⎠ ⎝H⎠

where
pa (=L/Aa) is the average pressure
t is the time of sliding

Equation 7.14 indicates that the thickness of the worn material is a linear function of time. Differentiation
of Equation 7.14 yields the average rate of surface recess due to the loss of material by adhesive wear,
dh/dt, that is,

dh Kvpa
h = = (7.15)
dt H

Using Equation 7.15 and in situ measurements of the surface recess rate, the wear coefficient of a sliding
system can be determined in terms of the known parameters H, v, and pa.
Equation 7.15 can also be used to obtain a relationship between parameters controlling the forma-
tion and removal of a tribochemical surface layer, such as a corrosive layer or a tribofilm produced
from a chemical reaction between lubricant additives and the sliding surfaces. Assuming that tri-
bofilm formation follows an Arrhenius relationship, the thickness of the growing tribofilm hf can be
expressed as

h f = De −Q /RT t n (7.16)

where
D is the diffusion coefficient
Q is the activation energy
R is the gas constant
T is the temperature (=sum of the ambient (operation) temperature and the temperature rise due to
frictional heating)
n is typically equal to 0.5 or 1

Thus, for tribofilm removal due to adhesive wear (Equation 7.15) and tribofilm formation following an
Arrhenius type of law (Equation 7.16) for n = 1, the following relationship is obtained for the balance
between the removal rate and the replenishment rate of the tribofilm:

Kvpa
= De −Q / RT (7.17)
H

Equation 7.17 can be used to tune the design parameters (e.g., v, pa, and T) such that to maintain a bal-
ance between the removal and the replenishment rates of a lubricious or protective tribofilm.
7-14 Theory and Practice of Lubrication and Tribology

Another case of interest is adhesive wear under heavy-load conditions, such as those encountered in
metal forming, extrusion, and stamping. Under these conditions, the real area of contact approaches the
apparent area of contact (Ar/Aa ≈ 0.75 is considered to be an upper bound) and, in view of Equation 7.5,
Equation 7.12 can be modified as

⎛ L⎞ ⎛ HAr ⎞
Aa h = K ⎜ ⎟ S = K ⎜ (vt )
⎝H⎠ ⎝ H ⎟⎠

or

⎛A ⎞
h = K ⎜ r ⎟ v ≈ Kv (7.18)
⎝ Aa ⎠

Equation 7.18 indicates that the rate of material transfer from the workpiece to the die surface due
to adhesive wear in heavy-load manufacturing processes is proportional to the wear coefficient and
the processing speed. Considering the factors affecting the magnitude of K (Figure 7.6), it is clear that
reducing the material compatibility and improving the lubrication efficiency are critical for preventing
material transfer to the die and roller surfaces in metal working.

7.7  Formation of Loose Wear Particles


In addition to material transfer, the formation of loose wear particles is also of high importance
because it affects the operation of systems with small clearances and tight tolerances, the design of
filters, and the integrity of sliding surfaces because the agglomeration and entrapment of wear par-
ticles at the sliding interface accelerate surface deterioration. Since material loss due to adhesive wear
is a consequence of the strong bonding between the asperities on the sliding surfaces, the formation
of loose wear particles may appear to be an unlikely event. However, experimental evidence suggests
that sliding surfaces demonstrating loss of material by adhesive wear also show formation of loose
wear debris. Thus, besides material transfer in the presence of strong adhesive forces (Figure 7.1b)
another mechanism controls the production of loose wear debris. In general, the formation of loose
wear particles is a consequence of simultaneous chemical and mechanical effects. Surface rubbing
changes the chemical behavior of the surfaces and facilitates the transfer of material to the opposed
surface. In the case of metal sliding, for example, rapid oxidation of the adhered fragments occurs due
to their large surface-to-volume ratio. Because oxidation is generally accompanied by a volumetric
change, a residual stress arises at the interface of the adhered fragments and the surface. An adhered
fragment will become a loose wear particle if the produced residual stress exceeds the bond strength
of the adhesive fragment to the surface.
A more general mechanism of loose wear particle formation is based on the energy change in a
fragment at the instant of its transfer to the opposed surface. A simple energy-based analysis can
be carried out for the process depicted in Figure 7.1b. Without loss of generality, assume homoge-
neous, isotropic, elastic-perfectly plastic material behavior and hemispherical fragment of diameter
d. During plastic shearing, the candidate fragment is under a compressive stress σ. This stress causes
the material to expand in the other two orthogonal directions (Poisson effect), producing a strain of
νσ/E in the horizontal direction, where E and ν are the elastic modulus and Poisson’s ratio, respec-
tively. At the instant that the fragment forms and transfers to the upper asperity (1) (Figure 7.1b), the
normal stress σ reduces to zero; however, the horizontal strains remain because interfacial adhesion
prevents the fragment from contracting, that is, elastic recovery is restricted. This gives rise to a resid-
ual stress in the horizontal direction equal to νσ. Since we are dealing with plastic deformation, the
normal stress σ may be set equal to the yield strength of the material SY, which is the largest stress that
Adhesive Wear 7-15

an elastic-perfectly plastic material can sustain. Using this simple model and Equation 7.13, we can
obtain the following relationships for the elastic strain energy Ee stored in the fragment when it is still
attached to asperity (2) and the adhesion energy Ea that keeps the fragment attached to asperity (1):

⎛ vS ⎞ ⎛ pd ⎞ pv Sgd
3 2 2 3
1 1
Ee = seU = (vSY ) ⎜ Y ⎟ ⎜ = (7.19)
2 2 ⎝ E ⎠ ⎝ 12 ⎟⎠ 24 E

⎛ pd 2 ⎞
Ea = AW12 = ⎜ cm cl (g1 + g2 ) (7.20)
⎝ 4 ⎟⎠

where
U is the volume of the fragment
A is the contact area between the fragment and asperity (1)

The transferred fragment will become a loose wear particle if Ee ≥ Ea. Using this condition in conjunc-
tion with Equations 7.19 and 7.20, we obtain

⎛ E ⎞
d ≥ 6 ⎜ 2 2 ⎟ cm cl (g1 + g2 ) (7.21)
⎝ v SY ⎠

Substitution of Equation 7.6 into Equation 7.21 gives

54 E
d≥ cmcl (g1 + g2 ) (7.22)
C 2mv 2 H 2

The aforementioned relationship provides a lower limit of the size of loose wear particles generated by
a sliding system in terms of mechanical properties, compatibility index, and surface energies of the
interacting surfaces, and is in qualitative agreement with experimental results [1]. As expected, larger
loose wear particles are predicted for compliant/low-strength surfaces exhibiting high affinity for the
counter-surface material and in the absence of a good lubricant. It is interesting to note that the size of
loose wear particles is intrinsic to the particular sliding system, that is, it does not depend on design
parameters, such as the load, sliding speed, and duration of sliding. Equation 7.22 can be used to make
preliminary calculations of design clearances, specify the pore size of filters, and select appropriate over-
coats for improving the surface mechanical properties and lowering the compatibility index to mini-
mize material loss and formation of loose wear particles due to adhesive wear.

Example 7.3
A metal–ceramic composite consists of Cu matrix and aluminum oxide (Al 2O3) particles in
volume fractions of 5%, 10%, 20%, 30%, and 40%.
(a) Calculate the surface energy and hardness of each composite material.
(b) For each composite material sliding against titanium carbide (TiC) in clean air, find the
work of adhesion and estimate the minimum size of loose wear particles. (Hint: Assume
that pullout of the Al2O3 grains does not occur and that the softer material is worn off
faster.)
(c) What will be the effect of coating one or both surfaces with a teflon layer on the work of
adhesion? (Hint: Consider adhesive wear, back and forth transfer of teflon, and clean air
environment.)
7-16 Theory and Practice of Lubrication and Tribology

The surface energy and hardness of each material are given as follows.

Properties
Material γ (J/m2) H (GPa)
Al2O3 0.74 21.08
Cu 1.10 0.784
TiC 0.90 23.53
Teflon 0.02 0.039

Solution
(a) The surface energy and hardness of each composite material, obtained from the rule of
mixtures, are given in the following table.

Properties
Composite Material γ (J/m2) H (GPa)
95% Cu + 5% Al2O3 1.082 1.80
90% Cu + 10% Al2O3 1.064 2.81
80% Cu + 20% Al2O3 1.028 4.84
70% Cu + 30% Al2O3 0.992 6.87
60% Cu + 40% Al2O3 0.956 8.90

(b) Since TiC is much harder than any of the composite materials, it can be assumed that only
the softer surface (composite) wears out, without this affecting the composition of the com-
posite materials. To proceed with the solution, we need to use Equations 7.13 and 7.22, that is,

54 E
W12 = cmcl (g1 + g2 ), dmin = cmcl (g 1 +g2 )
C 2m ν2 H 2

where subscripts 1 and 2 denote composite and ceramic surfaces, respectively. Since the
metal phase is the dominant element in the composite (Cu/Al2O3) and the counter surface is
a ceramic (TiC), the sliding system consists of dissimilar surfaces. For metal–nonmetal sur-
faces sliding in clean air, the compatibility index is cmcl = 0.36 (Figure 7.5b). Thus, the work
of adhesion of each composite–ceramic pair can be obtained by substituting γ TiC = 0.9 J/m2,
cmcl = 0.36, and the surface energies of the composite materials listed in the table of answer
(a) in the equation of W12 given earlier. These results are listed in the following table.
Wear particles will be mainly produced from the softer surface, that is, the composite
material (see table in answer (a)). Since it is stated that the Al2O3 grains are not pulled out
during sliding and that the softer material (i.e., Cu) wears out faster, we can assume that
the wear particles will mainly consist of Cu (HCu = 0.784 GPa). For C2m ≈ 1, ECu = 120 GPa,
and νCu = 0.3, the size of loose wear particles d estimated from the previous equation is
given in the following table.

Sliding System W12 (J/m2) dmin (μm)


95% Cu + 5% Al2O3 vs. TiC 0.714 83.6
90% Cu + 10% Al2O3 vs. TiC 0.707 82.8
80% Cu + 20% Al2O3 vs. TiC 0.694 81.3
70% Cu + 30% Al2O3 vs. TiC 0.681 79.8
60% Cu + 40% Al2O3 vs. TiC 0.668 78.3
Adhesive Wear 7-17

It can be seen that increasing the volume fraction of the Al2O3 particles not only increases
the hardness, but also reduces the work of adhesion, suggesting an improvement of the
adhesion characteristics for sliding against TiC. However, despite this beneficial effect, a
relatively high volume fraction of hard particles (e.g., ≥30%) could be detrimental to other
properties of the material (e.g., fracture toughness). In this situation, failure by other wear
mechanisms, such as surface and/or subsurface contact fatigue, may commence due to the
reduced ductility of the composite material with a high volume fraction of Al2O3 particles.
(c) To fully answer question (c), we need to consider the following three possible scenarios:
(i) Both surfaces are coated with teflon. This case represents a system with identical non-
metal surfaces sliding in clean air. Hence, from Figure 7.5b, we find that cmcl = 1. Since
both surfaces are covered by the teflon layer, interfacial adhesion is solely controlled
by the teflon properties. Thus, regardless of the composite composition, the work of
adhesion is

W12 = cmcl (gTeflon + gTeflon ) = 1(0.02 + 0.02) = 0.040 J/m2


The aforementioned value is lower than the work of adhesion values obtained in
(b) by two orders of magnitude, implying dramatically improved adhesion char-
acteristics and, in turn, tribological properties by coating both surfaces with a
teflon layer.
(ii) Only the composite surface is coated with teflon. This case represents a nonmetal–
nonmetal system of dissimilar surfaces, for which cmcl = 0.36 (Figure 7.5b). Hence, for
all composite compositions,

W12 = cmcl (gTeflon + gTiC ) = 0.36(0.02 + 0.9) = 0.331 J/m2


(iii) Only the TiC surface is coated with teflon. This is again a nonmetal–nonmetal
system involving dissimilar surfaces, for which cmcl = 0.36 (Figure 7.5b). However,
in this case the surface energy of one of the surfaces (composite) is not fixed but
varies with composition. Following a similar procedure, we find the results given in
the following table.

Sliding System W12 (J/m2)


95% Cu + 5% Al2O3 vs. TiC/Teflon 0.397
90% Cu + 10% Al2O3 vs. TiC/Teflon 0.390
80% Cu + 20% Al2O3 vs. TiC/Teflon 0.377
70% Cu + 30% Al2O3 vs. TiC/Teflon 0.364
60% Cu + 40% Al2O3 vs. TiC/Teflon 0.351

It is clear that the most advantageous scenario is the first one, that is, the case of both
surfaces coated with teflon, which has the lowest work of adhesion and, therefore,
should produce the best tribological properties. Based on the obtained results (i.e.,
W12(i ) < W12(ii ) < W12(iii )), scenarios (i)–(iii) can be ranked in the following order (from best
to worse): (i), (ii), and (iii). For the general case that the teflon coating is partially worn off
from one surface and/or partially covers the opposed sliding surface, the work of adhe-
sion results corresponding to scenarios (ii) and (iii) may be considered as upper bounds
during the transient stage that teflon transfers back and forth to the sliding surfaces. In
the ideal case of all contacting asperities being coated with teflon, the work of adhesion
corresponding to scenario (i), that is, 0.04 J/m2, is a lower bound of the work of adhesion
during the transition stage of scenario (iii).
7-18 Theory and Practice of Lubrication and Tribology

7.8 Evolution of Surface Roughness and Minimum


Clearance in Sliding Systems
Surface roughness is one of the most redundant aspects in mechanical design. Although surface finish
is normally of secondary importance in noncontacting mechanical components, with the exception
of those subjected to dynamic loads, it may affect the performance and operation life of contacting
mechanical elements, such as ball bearings, gears, and guideways, in many profound ways. Despite
this well-known fact, the selection of the initial surface roughness of articulating machine elements
is still based on empirical knowledge. This lack of insight is probably a consequence of the continuous
evolution of the surface topography (texture) during articulation with the opposed surface, until an
equilibrium condition is achieved at the contact interface. During this initial stage (run-in period), the
surface texture changes gradually because of various factors, including material transfer, formation of
loose wear particles, asperity deformation (fracture), and plowing. Although the contribution of each
of these processes to the final (equilibrium) surface texture depends on the type of sliding materials,
magnitude of surface tractions, and loading rate, the final surface texture does not show a dependence
on the initial surface state. The evolution of the surface texture due to adhesive wear may be interpreted
by considering the effect of material loss on the peak-to-valley roughness Rpv. At any given time, the
tallest asperities are the most likely to produce wear particles. After a particle is formed from one of
those asperities, this spot on the surface is likely to be the lowest within the neighborhood of the failed
asperity (Figure 7.1b). Another nearby asperity of slightly smaller height will then take the place of the
failed asperity. Thus, to a first order of approximation, it may be assumed that Rpv scales with the size of
the produced wear particles, that is, Rpv ~ d. Since the root-mean-square roughness Rq is the commonly
used statistical measurement of surface roughness and Rq is a fraction of Rpv, typically, Rq ≈ Rpv/6 [16], it
follows that Rq ~ d/6. Hence, using Equation 7.22, the following expression of the equilibrium roughness
can be obtained:

9E
Rq ≈ cm cl (g1 + g2 ) (7.23)
C 2mv 2 H 2

Equation 7.23 indicates that the equilibrium roughness depends on the elastic–plastic properties, sur-
face energy, and compatibility index of the sliding surfaces. This relationship can be used to determine
if the initial texture of a surface designed for a certain application is to be considered smooth or rough.
For example, a surface may be presumed to be rough if its Rq is about two to three times higher than that
calculated from Equation 7.23.
Seizure of a sliding system may occur when the available clearance is smaller than the largest wear
particles produced during operation. Entrapment of the larger wear particles at the sliding interface may
also lead to rapid damage of the opposed surfaces. However, Equation 7.22 gives a lower limit of the wear
particle size. While it is not straightforward to presume a maximum wear particle size dmax, a statistical
analysis shows that the size distribution of wear particles produced by adhesive wear typically shows
a spread of three times the average particle size [1]. Thus, it may be assumed that dmax is approximately
six times larger than the size of the smallest wear particles dmin (Equation 7.22). Hence, the minimum
clearance of a sliding system cmin can be expressed as follows:

324 E
cmin ≈ cm cl (g1 + g2 ) (7.24)
C 2mv 2 H 2

From Equations 7.22 and 7.24, it follows that cmin ≥ 6dmin.


Adhesive Wear 7-19

Example 7.4
A sliding system operated in sea water (poor lubricant) consists of a horizontal stainless steel
(74% Fe, 18% Cr, and 8% Ni) shaft of diameter D = 0.25 in. rotating in a copper–lead (75% Cu and
25% Pb) bushing of diameter slightly larger than 0.25 in. The hardness of the bushing and the
shaft are 0.4 and 3 GPa, respectively, and their lengths are l = 0.5 ft.
If the system’s performance becomes unsatisfactory when it can no longer rotate due to the
entrapment of wear particles or when the diametral clearance becomes equal to 0.055 in., esti-
mate the life of the system when the applied vertical load is L = 8 lb and the rotational speed of
the shaft is ω = 500 rpm, for an initial clearance of (a) 0.04, (b) 0.02, (c) 0.003, (d) 0.001, (e) 0.0003,
(f) 0.0001, and (g) 0.00003 in. (Given: K = 8 × 10−5, bushing properties: C2m ≈ 1, Eb = 100 GPa, and
νb = 0.3)

Solution
The surface energy of each mechanical component can be calculated by the rule of mixtures,
that is,

gb = aCu × gCu + aPb × gPb = 0.75 × 1.1 + 0.25 × 0.450 = 0.9375 J/m2

gs = aFe × gFe + aCr × gCr + aNi × gNi = 0.74 × 1.7 + 0.18 × 1.4 + 0.08 × 1.7 = 1.646 J/m2

The dominant elements (highest concentration) in the microstructure of the bushing and the
shaft are Cu and Fe, respectively, which are metals of intermediate compatibility I (Figure 7.4).
For this type of metallurgical compatibility and poor lubricant, Figure 7.5a gives cmcl = 0.14.
It is reasonable to assume that the clearance will increase with time mainly due to adhesive
wear of the softer bushing, neglecting the wear of the steel shaft. Then, the minimum clearance
(Equation 7.24) for this system can be obtained as

324 E 324 × (100 × 109 )


cmin = 2m 2
c c (gb + gs ) =
2 m l
× 0.14(0.9375 + 1.646) = 8.13 × 10 −4 m = 0.032in.
C v Hb 1 × 0.32 (0.4 × 109 )2

This result indicates that clearances (b)–(g) will eventually lead to the cessation of the system’s
operation because they cannot accommodate the largest wear particles generated from the
wearing bushing surface. Thus, for clearances (b)–(g), system failure will occur due to particle
embedment between the shaft and the bushing surfaces. To determine the time of failure for
these clearances, we need to find the time needed to generate a wear volume equal to that of a
hemispherical wear particle of diameter dmin, dav = 3dmin, or dmax = 6dmin, noting that the estimates
obtained for dmin and dmax represent lower and upper bounds, respectively, of the system’s life
to failure.
From Equation 7.22, we find that dmin ≈ 5.33 × 10−3 in.; hence, dav = 1.6 × 10−2 in. and
dmax = 3.2 × 10−2 in. Then, the time needed to produce a hemispherical particle of diameter dmin,
dav, and dmax can be obtained from Equation 7.12 as follows:

pd 3 LS L(vt ) L(wpD)t
=K =K =K
12 Hb Hb Hb

H bd 3 0.4 × 109 × (0.0254d )3


t= = ≈ 6.04 × 10 4 d 3 [min]
12KLw D 12 × (8 × 10 ) × (4.448 × 8) × 500 × (00.0254 × 0.25)
−5
7-20 Theory and Practice of Lubrication and Tribology

Substitution of d = 5.33 × 10−3, 1.6 × 10−2, and 3.2 × 10−2 in. into the previous equation gives
t ≈ 9.15 × 10−3, 0.247, and 1.98 min. These short times suggest that clearances (b)–(g) will result
in premature system failure.
Since clearance (a) is larger than dmax, we need to determine the time for the diametral
clearance to reach 0.055 in. Assuming uniform adhesive wear along the length of the bushing,
the amount of material to be removed for the diametral clearance to reach 0.055 in. can be
obtained as follows:

Δc (0.055 − 0.04) × 0.0254


V = p Dl = p × (0.0254 × 0.25) × (0.5 × 0.3048) × = 5.8 × 10 −7 [m3 ]
2 2

Thus, using Equation 7.12, the system operation life for an initial clearance of 0.04 in. is esti-
mated to be

VH b 5.8 × 10 −7 × (0.4 × 109 )


t= = ≈ 136.2 [h]
KLwpD 8 × 10 × (4.448 × 8) × 500 × p × (0.0254 × 0.25)
−5

7.9  Closing Remarks


Some fundamental aspects of adhesive wear, the most common process of material loss due to surface
rubbing, were elucidated in this chapter. Emphasis was given on quantitative analysis of wear rate, wear
coefficient, surface compatibility, size of loose wear particles, and equilibrium roughness. The use of
the presented relationships in preliminary analysis was demonstrated by solving illustrative examples.
Although scale dependent physical and chemical effects were discussed, the overall treatment was based
on a continuum description. Recent advances in nano/microprobe-based tribometers and high-resolu-
tion imaging techniques have begun to shed light into adhesive wear commencing at atomic and molec-
ular scales. However, further studies must be carried out at such small scales before the development of
quantitative models capable of yielding estimates of material loss at the nanoscale due to the effect of
strong adhesive forces.

References
1. Rabinowicz, E., Friction and Wear of Materials, John Wiley & Sons, New York, 1965.
2. Bowden, F. P., The physics of rubbing surfaces, Proc. R. Soc. NSW 78, 187–219, 1945.
3. Kerridge, M., Metal transfer and the wear process, Proc. Phys. Soc. B 68, 400–407, 1955.
4. Kerridge, M. and Lancaster, J. K., The stages in a process of severe metallic wear, Proc. R. Soc. London,
Ser. A 236, 250–264, 1956.
5. Bowden, F. P. and Tabor, D., The Friction and Lubrication of Solids, Oxford University Press, London,
U.K., 1950.
6. Kragelskii, I. V., Friction and Wear, Butterworths, London, U.K., 1965.
7. Holm, R., Electrical Contacts, Almquist & Wiksells, Stockholm, Sweden, 1946.
8. Archard, J. J., Contact and rubbing of flat surfaces, J. Appl. Phys. 24, 981–988, 1953.
9. Cahoon, J. R., Broughton, W. H., and Kutzak, A. R., The determination of the yield strength from
hardness measurements, Metall. Trans. 2, 1979–1983, 1971.
10. Burwell, J. T. and Strang, C. D., On the empirical law of adhesive wear, J. Appl. Phys. 23, 18–28, 1952.
11. Merchant, M. E., Friction and adhesion, in Interdisciplinary Approach in Friction and Wear, P.M. Ku,
ed., NASA SP-181, Washington, D.C., 1968, pp. 181–265.
Adhesive Wear 7-21

12. Hirst, W., Wear of unlubricated metals, in Proceedings of the Conference on Lubrication and Wear,
Institute of Mechanical Engineers, London, U.K., 1957, pp. 674–681.
13. Rabinowicz, E., Wear coefficients—Metals, in Wear Control Handbook, ASME, New York, 1980,
pp. 475–506.
14. Rabinowicz, E., The wear coefficient—Magnitude, scatter, uses, ASME J. Lubr. Technol. 103,
188–194, 1981.
15. Rabinowicz, E., Determination of compatibility of metals through static friction tests, ASLE Trans.
14, 198–205, 1971.
16. Moore, D. F., The Friction and Lubrication of Polymers, Pergamon, New York, 1972.
8
Abrasive Wear
8.1 Specific Wear Rate ws and Wear Coefficient K.............................. 8-1
8.2 Abrasive Wear Mode and Degree of Penetration Dp................... 8-3
8.3 Friction Coefficient and Abrasive Wear Mode as Functions
of Dp..................................................................................................... 8-5
8.4 Abrasive Wear Mode Transition in Repeated Sliding.................. 8-7
8.5 Abrasive Wear Mode and Degree of Wear β................................. 8-9
8.6 Hardness Increase on Groove Surface by Repeated
Rubbing........................................................................................ 8-10
Koji Kato 8.7 Concluding Remarks....................................................................... 8-11
Nihon University References..................................................................................................... 8-12

8.1  Specific Wear Rate ws and Wear Coefficient K


Abrasive wear is defined as loss of volume or mass from a surface caused by sliding of a relatively harder
abrasive or asperity, which leaves a groove on the wearing surface.
In standard abrasive wear tests of metals with abrasive grids, abrasive papers, or files, the following
relationship is commonly observed experimentally [1,2]:

V = w sWL (8.1)

where
V is the wear volume
W is the normal load
L is the sliding distance
ws is the experimental constant

The proportional constant ws is called specific wear rate, specific wear volume, or specific wear
amount and generally ranges from 10−3 to 10−1 mm3/N m, depending on the contact materials and con-
tact conditions in abrasive friction.
It is commonly observed in abrasive wear of metals that the value of ws becomes smaller as the hard-
ness, Hv, of wearing material becomes larger [1,3–5]. Figure 8.1 shows an example of the experimental
relationship [5] observed between 1/ws and Hv in abrasive wear tests of 20 kinds of metals and alloys,
which is described approximately by

1 H
= v (8.2)
ws K

where K is the experimental constant.

8-1
8-2 Theory and Practice of Lubrication and Tribology

6
SUS304

Wear resistance ws–1, ×1010 N/m2


5
Mo

Fe
4
S45C
Ni

3 Fe
Phosphor Ni
bronze
2
Cu Unlubricated
Brass
Zn Hokkirigawa, kato
Cn Ag
1 Zu Sasada, oike
Sn
Al
Al
Sn
Pb
0
100 200 300 400 500 600 700
Hardness Hv, kgf/mm2

FIGURE 8.1  Relationship between wear resistance defined by 1/ws and hardness, Hv, of wearing material in
unlubricated abrasive wear of 20 different metals and alloys. (From Hokkirigawa, K. and Kato, K., Theoretical
estimation of abrasive wear resistance based on microscopic wear mechanism, in ASME, Proc. Int. Conf. on Wear
of Metals, Denver, CO, Vol. 1, 1989, pp. 1–8. With permission.)

By having ws = K/Hv from Equation 8.2 and introducing it into Equation 8.1,

K ⋅ WL
V= (8.3)
Hv

where K is called wear coefficient and ranges from 10−3 to 10−1 for metals in general [1]. It is a coefficient
that is related to the area ratio of a wear particle area against a micro-abrasive contact area and the ratio
of real sliding contact distance of an abrasive against its nominal sliding distance.
In Equation 8.3, W/Hv is approximately equal to real contact area, Ar, in plastic contact [6], and
Equation 8.3 is described with Ar as follows:

V = K ⋅ Ar ⋅ L (8.4)

where Ar is the sum of all micro-contact area at all abrasive contacts and is given by

Ar = ∑ ΔA ri (8.5)
i =1

where
ΔAri is the contact area at ith micro-abrasive contact
N is the total number of micro-abrasive contacts

The value of wear coefficient, K, can be different at each abrasive contact. Equation 8.4 is described for
the case as follows:
N

V= ∑ K ΔA L i ri (8.6)
i =1
Abrasive Wear 8-3

It is clear from Equation 8.6 that understanding abrasive wear mechanism is to know the value of Ki
at the ith contact area, ΔAri, and their distributions in the nominal contact area and nominal sliding
distance, L.
With this thought, abrasive wear modes and degree of wear at one abrasive contact are explained in
the following.

8.2  Abrasive Wear Mode and Degree of Penetration Dp


Three abrasive wear modes are known by the in situ observations in the SEM as shown in Figure 8.2
[7]. In Figure 8.2a, a long ribbon-like wear particle is generated, and this wear mode is called cutting
mode. A deep and long groove is formed distinctly on the wearing surface and relatively high wear rate
is generated in this wear mode.
In Figure 8.2b, a wedge is formed ahead of the pin and sticks to it. Once this wedge is formed, sliding
takes place at the interface between the wedge bottom and the flat surface in the following process. This
wear mode is called wedge-forming mode, which is observed when adhesion between a pin and a flat is
strong. The wear rate under this wear mode is much smaller than that of cutting mode, and the groove
surface is generally rough, with transverse cracks.
In Figure 8.2c, a wear particle is not clearly observed at the interface and only a shallow groove is
formed plastically on a flat surface. This wear mode is called ploughing mode, which is observed when
adhesion at the interface is weak and the penetration by an asperity is relatively small. Although a shal-
low groove is formed on a flat surface as a result of plastic deformation, real volume removal from the
surface does not take place in a single pass of sliding. Repeated sliding on the same groove is necessary
in this mode for generating wear in the form of plastic flow wear and/or low cycle fatigue wear.
In order to describe the severity of contact of an abrasive that causes different modes of wear shown in
Figure 8.2, a model in Figure 8.3 is introduced for the contact between a hard spherical pin and a soft flat
specimen. As an index of severity of contact in the model, the degree of penetration, Dp, is introduced
by the following definition [7]:

h
Dp = (8.7)
a

where
a is the contact radius
h is the penetration depth

n
ctio
g dire
Slidin

50 µm 20 µm 50 µm

(a) (b) (c)

FIGURE 8.2  Three abrasive wear modes confirmed by in situ SEM observations: (a) cutting mode, steel pin on
brass plate; (b) wedge-forming mode, steel pin on stainless plate; and (c) ploughing mode, steel pin on brass plate.
(From Hokkirigawa, K. and Kato, K., Tribol. Int., 21(1), 51, 1988. With permission.)
8-4 Theory and Practice of Lubrication and Tribology

Contact area

R
h

FIGURE 8.3  Model of contact between a hard spherical pin and a soft flat specimen: h, depth of groove; a, radius
of contact area; and R, radius of pin tip.

The geometrical condition of the model gives the following equation between a, h, and R:
12


(
h = R − R2 − a2 ) (8.8)

where
R is the pin tip radius

By introducing Equation 8.8 into Equation 8.7,

12
R ⎛ R2 ⎞
Dp = − − 1⎟ (8.9)
a ⎜⎝ a 2 ⎠

where Dp = 1 at a = R is the maximum value in the model. If the mean contact pressure is assumed equal
to the hardness, Hv, of the soft flat specimen, the semicircular contact area, πa 2/2, is expressed as follows:

pa2 W
= (8.10)
2 Hv

By introducing Equation 8.10 into Equation 8.9, the degree of penetration, Dp, is expressed as a function
of the pin tip radius, R, the normal load, W, and the hardness, Hv, as follows:
12 12
⎛ pH v ⎞ ⎛ pH v 2 ⎞
Dp = R ⎜ −⎜ R − 1⎟ (8.11)
⎝ 2W ⎟⎠ ⎝ 2W ⎠

For the case of h/2R << 1, Equation 8.8 gives the following equation as an approximation:

a2
h= (8.12)
2R

By introducing Equation 8.12 into Equation 8.7, Dp is given by the following equation as an approximation:

a
Dp = (8.13)
2R
Abrasive Wear 8-5

By introducing Equation 8.10 into Equation 8.13, Dp is simply described with R, W, and Hv as follows:

1 W
Dp = (8.14)
R 2pH v

The relationship between the introduced Dp and observed wear modes will be explained in the
following.

8.3 Friction Coefficient and Abrasive Wear


Mode as Functions of Dp
The friction coefficient, μ, changes depending on the value of Dp. Figure 8.4 shows examples observed
in sliding of a hard steel pin against a relatively soft S45C disk with and without grease lubricant [7]. All
tests with different pin tip radius and normal load are carried in the SEM and wear modes are confirmed
by the in situ observation together with the measurement of friction force during sliding.
Although the observed friction coefficient, μ, is complicatedly related to Dp, it can be explained
theoretically if Dp is translated into an attack angle θ in a 2-D abrasive model proposed by Challen and
Oxley [8].
Their theoretical analysis of 2-D models with slip line field theory gives the following relationships
for the cutting mode

⎛ p 1 ⎞
m = tan ⎜ q − + cos −1 f⎟ (8.15)
⎝ 4 2 ⎠

2.5
P W C
Dry

2.0
Coefficient of friction µ

1.5

1.0

Flat specimen: S45C


Dry
0.5
Lubricated
P W C
Lubricated

0 0.1 0.2 0.3 0.4


Degree of penetration Dp

FIGURE 8.4  Relationship between friction coefficient, μ, and degree of penetration, Dp, for quenched steel
pins of 27 or 62 μm tip radius sliding against S45C disks, lubricated and unlubricated: P, ploughing mode; W,
wedge-forming mode; and C, cutting mode. (From Hokkirigawa, K. and Kato, K., Tribol. Int., 21(1), 51, 1988. With
permission.)
8-6 Theory and Practice of Lubrication and Tribology

the wedge-forming mode

m=
1 − 2 sin g + (1 − f ) sinq + f cosq
2

(8.16)

1 − 2 sin g + (1 − f ) cosq + f sinq
2

and the ploughing mode

m=
(
A sin q + cos cos −1 f − q ) (8.17)
A cos q + sin ( cos −1
f − q)

⎧⎪ p ⎛ sin q ⎞ ⎫⎪
A = ⎨1 + + cos −1 f − 2q − 2 sin −1 ⎜ ⎟⎬
2 ⎝ 1 − f ⎠ ⎭⎪
⎩⎪

where
θ is the attack angle of a model abrasive
f is the shear strength at the contact interface normalized by the shear strength of the wearing
material
γ is the angle determined by the stress discontinuity

Equations 8.15 through 8.17 explain very well the observed relationship between μ and Dp in Figure 8.4
when Dp and θ are related geometrically in Figure 8.5 by the following equation:

q
D p = a tan (8.18)
2

where α is a shape factor to be determined for reducing differences between theoretical values with
2-D models and experimental values with 3-D models. The most suitable value of α for the spherical
pin tip is found as 0.8. With Equations 8.15 through 8.18, the value of f can be determined for the
experimentally designed Dp and the observed resultant value of μ in each wear mode.
Figure 8.6 shows abrasive wear regions of cutting, wedge-forming, and ploughing modes observed on
flat surfaces of brass, S45C steel, and SUS304 stainless steel as a function of degree of penetration, Dp,
and relative shear strength, f, at the contact interface, under sliding of quenched steel pins of tip radius
27 or 62 μm [5]. Solid lines in the figure are theoretical boundaries between two abrasive wear modes
obtained by the criteria of wear mode transition by Challen and Oxley [8].

R θ

a
h
θ

FIGURE 8.5  Definition of attack angle, θ, for a spherical pin tip of radius, R, grooving with the depth, h, and the
contact radius, a.
Abrasive Wear 8-7

0.6
R
a
D p = h/a
0.5 h

0.4

Degree of penetration Dp
Cutting
0.3

0.2
Wedge

0.1
Ploughing

0 0.2 0.4 0.6 0.8 1.0


Relative shearing strength
at the contact interface f

FIGURE 8.6  Regions of abrasive wear modes of cutting, wedge forming, and ploughing as a function of degree
of penetration, Dp, and relative shear strength, f, at the contact interface. (From Hokkirigawa, K. and Kato, K.,
Theoretical estimation of abrasive wear resistance based on microscopic wear mechanism, in ASME, Proc. Int. Conf.
on Wear of Metals, Denver, CO, Vol. 1, 1989, pp. 1–8. With permission.)

8.4  Abrasive Wear Mode Transition in Repeated Sliding


When sliding of an abrasive is repeated on the same wear groove on a flat surface, the abrasive wear
mode transits from cutting to wedge forming or from wedge forming to ploughing.
It is shown in Figure 8.7 for the case of a spherical WC pin repeatedly sliding on a flat SUS304 steel
disk, where the abrasive wear mode transitions from cutting to wedge forming and then to ploughing
are observed under a larger load and that from wedge forming to ploughing under medium loads, and
ploughing mode only under smaller loads in the repeated sliding up to 1000 cycles [9].

Cutting WC pin on SUS 304 disk


1.6

1.4
1.2
Load W, N

1.0 Wedge
Ploughing
0.8

0.6

0.4

0.2

0
1 5 10 50 100 500 1000
Number of cycles n, cycle

FIGURE 8.7  Abrasive wear mode transition caused by repeated sliding cycle, n, of a WC pin of 30 μm tip radius
on a SUS304 steel disk under a constant load without lubricant. (From Kitsunai, H. et al., Wear, 151, 279, 1991. With
permission.)
8-8 Theory and Practice of Lubrication and Tribology

R
hn

hn–1
hn
an

FIGURE 8.8  Model of abrasive contact at nth sliding cycle in repeated sliding of a spherical pin against a flat
surface: an, contact radius at nth cycle; hn, groove depth at nth cycle; and hn−1, groove depth at (n−1)th cycle.

The contact model in the process of repeated sliding is schematically shown in Figure 8.8 where hn−1
and hn show the groove depth at the (n − 1)th and nth sliding, and an the contact length at the nth sliding.
If the degree of penetration, Dpn, at the nth sliding is defined for n ≥ 2 by

Δhn
D pn = (8.19)
an

where Δhn = hn − hn−1, Equation 8.19 is described as follows by considering the geometrical relationship
between Δhn, an, and R in Figure 8.8:

Δhn
D pn = (8.20)
2
R − ( R − Δhn )
2

The values of Δhn observed for the first 10 cycles in Figure 8.7 are introduced into Equation 8.20 and
Dpn values are calculated and shown in Figure 8.9 in relation to sliding cycle n [10]. In the figure, the

Theory
WC pin on SUS 304 disk
Wear mode
Load Cutting Wedge Ploughing
0.07N C
0.2 N
0.5 N
0.3 1.0 N
1.5 N
0.27
Degree of penetration, Dpn

0.2
W

0.1
0.09

0
1 2 3 4 5 6 7 8 9 10
(a) Number of cycles n, cycle (b)

FIGURE 8.9  Abrasive wear mode transition in repeated sliding and its prediction: (a) abrasive wear modes and
degree of penetration, Dpn, at the nth cycle as a function of sliding cycle, n, and (b) cutting mode, C, wedge-forming
mode, W, and ploughing mode, P, at f = 0.85 in abrasive wear mode diagram of Figure 8.6. (From Kitsunai, H. et al.,
Wear, 135, 237, 1990. With permission.)
Abrasive Wear 8-9

observed ranges of Dp1 for cutting, wedge-forming, and ploughing modes are shown by the symbols of
C, W, and, P, which agree well with the ranges of Dp at f = 0.85 in Figure 8.6.
It is very clear from the figure that Dpn becomes smaller as the sliding cycle increases, and abrasive
wear mode transits from cutting to wedge forming or from wedge forming to ploughing as a result. The
critical value of Dpn for each transition is almost similar to the value of Dp given by Equation 8.11, which
is equal to Dp1, for a single pass of sliding.

8.5  Abrasive Wear Mode and Degree of Wear β


The apparent area of a groove formed in a single pass of an abrasive does not totally give the wear vol-
ume. Figure 8.10 shows cross-sectional profiles of grooves formed by a single pass of pin on a flat surface
[11]. The groove profile in Figure 8.10a generated by ploughing mode is mostly occupied by side ridges
and that in Figure 8.10b generated by wedge-forming mode is more occupied by the part of valley. The
groove profile in Figure 8.10c by cutting mode has very few parts of side ridges and shows mostly the
part of valley.
Such wear grooves are schematically shown in Figure 8.11, where ΔA′ means the cross-sectional area
of a valley and ΔA″ is the cross-sectional area of side ridges. (ΔA′ − ΔA″) should correspond to the real
wear volume. If degree of wear, β, at one groove is defined by the following equation:

ΔAʹ − ΔAʹʹ
b= (8.21)
ΔAʹ

β becomes zero when there exists no volume loss from the groove but only plastic deformation, and β
becomes unity when ΔA″ is zero. The wear volume, ΔV, from the groove after sliding distance, L, is given
by the following equation:

ΔV = b ΔAʹ L (8.22)

Figure 8.12 shows β as a function of Dp in the case of a steel pin sliding against a steel disk of different
hardness, Hv, prepared by heat treatment. In the figure, the wear modes of cutting, wedge forming, and
1 µm

20 µm
(a) (b) (c)

FIGURE 8.10  Cross-sectional profiles of abrasive grooves: (a) ploughing mode, (b) wedge-forming mode, and (c)
cutting mode. (From Hokkirigawa, H. et al., Wear, 123, 241, 1988. With permission.)

Degree of wear ∆ A˝

∆ A΄ – ∆ A˝
β
∆ A΄
∆ A΄

FIGURE 8.11  Definition of the cross-sectional area of a valley, ΔA′, and that of side ridges, ΔA″.
8-10 Theory and Practice of Lubrication and Tribology

1.0
Cutting
0.8 Wedge

Degree of wear β
Ploughing
0.6 Specimen E (Hv = 750)
Specimen D (Hv = 650)
0.4
Specimen C (Hv = 560) Specimen B (Hv = 380)

0.2 Specimen A (Hv = 250)

0 0.1 0.2 0.3 0.4 0.5


Degree of penetration D p

FIGURE 8.12  Degree of wear, β, as a function of degree of penetration, Dp, observed with heat-treated steel whose
hardness, Hv, ranges from 250 to 750 kgf/mm2. (From Hokkirigawa, H. et al., Wear, 123, 241, 1988. With permission.)

ploughing are shown by different symbols [11]. It is understood from these results that β varies between
0.8 and 1.0 in the cutting mode, between 0.1 and 0.8 in the wedge-forming mode, and is almost zero
in the ploughing mode. It is also shown in the figure that the cutting mode of large β is generated at
smaller Dp values as the hardness of material increases and the wedge-forming mode is generated in a
wide range of Dp values as the hardness of materials decreases.
If we introduce contact area, ΔAr, for an abrasive that forms a groove of ΔA′ and ΔA″ shown in
Figure 8.11, then suppose the following relationship:

ΔAʹ = g ΔAr (8.23)


where γ is a constant. ΔV is described as follows by introducing Equation 8.23 into Equation 8.22:

ΔV = bg ΔAr L (8.24)

When the total number of abrasive contacts is N for the real contact area, Ar, as described by Equation
8.5, the values of ΔAr, β, and γ at each abrasive contact can be different. Therefore, the wear volume, V,
described by Equation 8.6 can be expressed in a different way as follows:

V= ∑ b g ΔA L i i ri (8.25)
i =1

By comparing Equations 8.6 and 8.25, the character of Ki can be understood to some extent by the
following equation:

K i = bi gi (8.26)

where βi strongly depends on Dp as shown in Figure 8.12. It is also important to notice that ΔAri is
proportional to the inverse of Hv on the groove under constant load.

8.6  Hardness Increase on Groove Surface by Repeated Rubbing


The hardness of metal surfaces generally increases by having repeated rubbing. Figure 8.13 shows the
increase in hardness, Hv, on a wear groove of a SUS304 steel flat surface generated by repeated sliding of
a WC pin of tip radius 30 μm under the load of 0.2 N [12]. It is very clear in Figure 8.13 that the surface
Abrasive Wear 8-11

1200 WC pin on SUS304 disk


load 0.2 N

Hardness Hv
800

400

0
200 400 600 800 1000
Number of cycles n, cycle

FIGURE 8.13  Increase in hardness, Hv (kgf/mm2), on a wear groove of a SUS304 steel flat surface generated by
repeated sliding of a WC pin of 30 μm tip radius under 0.2 N load. (From Kitsunai, H., The study of generation
mechanism and controlling technology of nano/micro particles in production process of electric devices, PhD dis-
sertation, Tohoku University, Sendai, Japan, pp. 29–30, 2007. With permission.)

hardness is below 400 kgf/mm2 before rubbing by a pin but it is above 800 kgf/mm2 on a groove after
having 400 sliding cycles, which means that the groove surface hardness can be two or three times larger
than the initial hardness in general after having repeated rubbing by an abrasive.
Therefore, the real-time hardness value on a groove has to be used for Equation 8.11 or 8.14 in deter-
mining the value of degree of penetration, Dp, in practice, and the prediction of degree of wear, β, in
Figure 8.12 has to be made along the β–Dp curve of proper hardness.

8.7  Concluding Remarks


It is well observed in macroscopic wear tests that abrasive wear volume, V, is approximately proportional
to the inverse of hardness, Hv, of wearing material as shown in Figure 8.1 for 20 kinds of metals and
alloys. Observed values of wear coefficient, K, range from 10−3 to 10−1 in general. For explaining such
macroscopic abrasive wear properties, fundamental mechanisms of abrasive wear at one abrasive con-
tact have to be understood.
The in situ SEM observations of abrasive wear process in microscopic abrasive tests show that three
wear modes of cutting, wedge forming, and ploughing are generated in abrasive wear.
The regions of those three wear modes can be described on the abrasive wear map with parameters of
degree of penetration, Dp, and relative shear strength, f, as shown in Figure 8.6.
The degree of wear, β, at one abrasive groove is a function of Dp as shown in Figure 8.12 and is largest
in cutting mode, medium in wedge-forming mode, and smallest in ploughing mode.
By introducing the ratio, γ, of valley area A′ against contact area ΔAr of an abrasive, K at one abrasive
contact is given by β·γ.
Further understanding of abrasive wear property can be made by considering the increase in hard-
ness on the groove surface and the decrease in the degree of penetration generated by the repetition of
sliding cycle as shown in Figures 8.9 and 8.13.
In practical situations of multiple abrasive contacts, distributions of ΔAr, Dp, f, β, and γ have to be
considered, and certain models for them become necessary for wear prediction. It is not explained in
this chapter, but some models are shown in Reference 13.
The abrasive wear explained in this chapter belongs to the category of two-body abrasion where abra-
sives are firmly fixed to the sliding surface of one body. There exists another category of abrasive wear,
which is called three-body abrasion. Abrasives are free there and roll between two solid surfaces in rela-
tive motion, and wear is generated on surfaces by repeated indentations of abrasives.
Abrasive wear volume, V, by rolling abrasives is again proportional to the inverse of hardness, Hv, of
the wearing material, although specific wear rate, ws, is much smaller than that in two-body abrasion.
8-12 Theory and Practice of Lubrication and Tribology

A critical condition for the transition from three-body to two-body abrasion is predictable by introduc-
ing a criterion of severity of contact for the two bodies and abrasives in slurry. They are well explained
in References 14 and 15.

References
1. M. B. Peterson and W. O. Winer (Eds.), Wear Control Handbook, ASME, New York (1980).
2. J. M. Challen and P. L. B. Oxley, Prediction of Archard’s wear coefficient for metallic sliding friction
assuming a low cycle fatigue wear mechanism, Wear, 111 (1986), 275–288.
3. M. M. Khrushov, Resistance of metals to wear by abrasion-as related to hardness, Proceedings of
International Conference on Lubrication and Wear, Institute of Mechanical Engineering, London
U.K. (1957), pp. 655–659.
4. D. C. Evans and J. K. Lancaster, in D. Scott (Ed.), Treatise on Materials Science and Technology, Vol. 13,
Academic Press, London, U.K. (1979), pp. 85–139.
5. K. Hokkirigawa and K. Kato, Theoretical estimation of abrasive wear resistance based on micro-
scopic wear mechanism, ASME, Proceedings of International Conference on Wear of Metals, Denver,
CO, Vol.1 (1989), pp. 1–8.
6. F. P. Bowden and D. Tabor, Friction and Lubrication of Solids I, Clarendon Press, Oxford, U.K. (1950).
7. K. Hokkirigawa and K. Kato, An experimental and theoretical investigation of ploughing, cutting
and wedge formation during abrasive wear, Tribology International, 21(1) (1988), 51–57.
8. J. M. Challen and P. L. B. Oxley, An explanation of the different regimes of friction and wear using
asperity deformation models, Wear, 53 (1979), 229–243.
9. H. Kitsunai, K. Hokkirigawa, N. Tsumaki, and K. Kato, Transitions of microscopic wear mechanisms
for Cr2O3 ceramic coatings during repeated sliding observed in a scanning electron microscope tri-
bosystem, Wear, 151 (1991), 279–289.
10. H. Kitsunai, K. Kato, K. Hokkirigawa, and H. Inoue, The transition between microscopic wear modes
during repeated sliding friction observed by SEM-tribosystem, Wear, 135 (1990), 237–249.
11. H. Hokkirigawa, K. Kato, and Z. Z. Li, The effect of hardness on the transition of the abrasive wear
mechanism of steels, Wear, 123 (1988), 241–251.
12. H. Kitsunai, The study of generation mechanism and controlling technology of nano/micro particles
in production process of electric devices, PhD dissertation, Tohoku University, Sendai, Japan, 2007,
pp. 29–30.
13. K. Kato, Abrasive wear, Characterization of Tribological Materials, W. A. Glaeser (Ed.), Butterworth-
Heinemann, Boston, MA, 1993, pp. 80–97, Chapter 5.
14. K. Adachi and I. M. Hutchings, Sensitivity of wear rates in the micro-scale abrasion test to test condi-
tions and material hardness, Wear, 258 (2005) 318–321.
15. K. Adachi and I. M. Hutchings, Wear-mode mapping for the micro-scale abrasion test, Wear, 255
(2003), 23–29.
9
Rolling Contact
Fatigue Wear
9.1 Origin of Rolling Contact Fatigue.................................................. 9-1
9.2 Distribution of Rolling Contact Fatigue........................................ 9-3
9.3 Predicting RCF................................................................................... 9-5
9.4 Material Considerations...................................................................9-6
9.5 Other Considerations........................................................................9-8
Michael N. Kotzalas 9.6 Conclusion........................................................................................ 9-10
The Timken Company References..................................................................................................... 9-10

9.1  Origin of Rolling Contact Fatigue


Rolling contact fatigue (RCF) is classified within as a wear mode. This mode removes material from the
contact surface of a machine element primarily due to rolling motion. The most common machine ele-
ments associated with this wear mode are rolling element bearings; however, gears, cams, clutches, and
other mechanisms are also prone to this wear mode depending on their design style. At a high level, the
mechanism is associated with the concentrated contact causing stresses in the material that change, or
cycle, as the contact moves over the surface. The cyclical nature of the subsurface stress may result in
crack initiation and propagation from a weak point in the material and removal of material from the
surface. The result of this wear is the formation of a pit on the contact surface, which is often referred
to as a spall.
To prevent RCF, steel used for manufacturing the machine components is typically hardened to
obtain high strength. The hardening process yields a microstructure that is primarily martensite with
some retained austenite and metallic carbides interspersed in a crystalline matrix (see Figure 9.1).
When viewed under higher magnifications, an occasional inhomogeneity becomes evident. These
inhomogeneities are usually phases not intended to be part of the steel chemistry. Some alloying ele-
ments such as aluminum react with oxygen to form aluminum oxide. Most of the aluminum oxides
formed will be removed during steelmaking processing, when the steel is still liquid. However, a
fraction of them become trapped during solidification and remain in the steel as solid inclusions.
These inclusions have different material properties than the surrounding steel. When the material is
stressed, the differences in material properties cause the localized stresses to drastically increase and
the localized martensite transforms to ferrite. The result is a white etch region commonly referred
to as butterflies (see Figure 9.2). Over time the effect of the cyclic stresses may result in crack initia-
tion, coalescence, and propagation until a spall is formed on the contacting surface (see Figure 9.3a).
When RCF is initiated in the subsurface, it is often termed an inclusion fatigue. Inclusion RCF spalls
are typically deep in origin because the crack initiates beneath the surface and typically has an ellip-
soidal shape.

9-1
9-2 Theory and Practice of Lubrication and Tribology

20 µm 20 µm

(a) (b)

FIGURE 9.1  Typical microstructures of hardened machine elements manufactured from (a) high carbon steel
(e.g., AISI 52100) and (b) low carbon steel (e.g., AISI 8620) after carburizing.

FIGURE 9.2  Butterfly caused by a high stress region around an inclusion in a primarily martensitic steel
microstructure.

Over the last 50 years, steelmaking practices have improved to the point where inclusion RCF is no
longer the dominant mode of fatigue. Today, other conditions lead to RCF. One of these mechanisms
is associated with any feature on the contacting body. These features can increase or concentrate stress
in a localized area.
The most common example occurs when the edges of the mating bodies are not relieved through
special profiling. The loading of the body edges yields high surface pressure, which translates to high
subsurface stresses. When RCF is caused by this mechanism, it is often termed geometric stress concen-
tration (GSC).
Figure 9.3b shows a typical GSC spall. The spall is concentrated around a geometric feature, such as
the edge of a raceway on a roller bearing, and is often found in conjunction with areas exhibiting visual
plastic deformation.
The second most common form of RCF is similar to GSC in that it is associated with stress concen-
tration, but instead occurs at microscale features on the rolling contact surfaces. After a manufactured
surface is finished, it will contain small peaks or asperities. If the magnitude of these peaks or asperities
is large relative to the lubricant film thickness, this is commonly due to insufficient lubrication and true
Rolling Contact Fatigue Wear 9-3

(a) (b)

(c)

FIGURE 9.3  RCF spall morphology as related to the mode of origin: (a) inclusion spall that is deep and has a
well-formed ellipsoidal shape, (b) geometric stress concentration spalls that are located at an edge of a roller bear-
ing raceway, and (c) point surface origin spall started at the ridge of an asperity from the finishing process with an
arrow head shape.

solid-to-solid contact, and sliding occurs. This results in high stresses at or near the contact surface.
Failure caused by this mechanism is termed a point surface origin (PSO) (see Figure 9.3c). A PSO spall
has the shape of an arrowhead pointing to the initiation site and is shallow at its origin. After initiation,
stress cycles accumulate and the spall will propagate and widen, eventually encompassing the entire
contact area. As the spall widens, a smaller area must carry the same load since it is only the undam-
aged raceway in contact with the rolling element. This diminishing area increases the contact stress and
the spall often widens rapidly. The increase in contact stress, and deeper subsurface shear stresses, also
deepens the spall. Additionally, when foreign particles are present in the lubricant, they tend to damage
the surface; this can also lead to PSO RCF.

9.2  Distribution of Rolling Contact Fatigue


RCF begins at a weak point in the material. Repeated cycling of stresses initiate a crack, which subse-
quently propagates to create a spall. As such, one would not expect every component to have a weak
point in the same location. Therefore, each bearing has a unique fatigue life, even when seemingly
9-4 Theory and Practice of Lubrication and Tribology

β = 0.5

Probability distribution function


β=3

f (Lx)

β = 1.5

β=1

Lx

FIGURE 9.4  Weibull probability distributions, f(Lx), for various shape parameter values, β.

identical. When a population of mechanisms, all failing due to rolling contact, is collectively studied,
the fatigue life follows a distribution. This distribution of fatigue has proven to be large and almost ran-
dom in nature. For example, the time where 50% of a population of seemingly identical rolling element
bearings have fatigued is typically 3.4 times larger than the time where 10% has fatigued, and the point
where 90% of the population experiences fatigue is 7.5 times larger. The statistical probability of fatigue
failure was first studied by Weibull [1,2], whose name is now associated with the family of probability
distributions most commonly used when studying the fatigue of contact mechanics. The distribution is
very flexible and can describe populations with highly skewed or even Gaussian failure rates depending
on the value of the shape parameter (see Figure 9.4). This shape parameter is often termed the Weibull
slope. From Harris and Kotzalas [3], typical values of the shape parameter from RCF testing of rolling
bearings range from 0.7 to 3.5. As evident in Figure 9.4, varying the shape parameter results in signifi-
cant variation in the expected failure rates.
More recent data indicate that the three-parameter form of the Weibull distribution is superior in
correlating the fatigue lives to modern manufacturers’ materials [4,5]. The three-parameter RCF life is
given as a function of the Weibull distribution parameters:

1
b
⎡ ⎛ 100 ⎞ ⎤
Lx = h ⎢ log e ⎜ +g (9.1)
⎣ ⎝ R ⎟⎠ ⎥⎦

where
Lx is the expected life of the rolling contact at the given population reliability, R
R is the desired reliability level
γ is the location parameter
η is the scale parameter
β is the shape parameter

The best Weibull fit for rolling element bearing shape parameter β and the location parameter γ are,
respectively, 1.5 and 0.216 [4,5].
Table 9.1 shows the highly stochastic nature of RCF. In Table 9.1, the life is given relative to the life at
90% reliability as this is the most typical value for which machine elements are designed. Table 9.1 also
shows the median (R = 50%) and average (R = 36.6%) values are almost 3 and 3.75 times larger than the
typical design target (R = 90%), thus the previous statement that the fatigue life distribution is almost
Rolling Contact Fatigue Wear 9-5

TABLE 9.1  Rolling Contact


Fatigue Reliability Distribution
for β = 1.5 and γ = 0.216
R (%) Lx/L10
100 0.216
99.9 0.251
99.5 0.319
99 0.380
98 0.477
97 0.559
96 0.633
95 0.701
90 1.000
75 1.748
50 2.969
36.6a 3.742
25 4.586
1 9.944
a Average value of distribution.

random. Therefore, when RCF testing is conducted a large number of samples are required to ensure the
statistical validity of the estimated Weibull parameters.

9.3  Predicting RCF


The leading theory on predicting the occurrence of RCF is through the method of Ioannides and Harris
[6] (IH) (Equation 9.2):

1
⎡ A (s − s lim )c dV ⎤⎥ b × 106
( )∫
Lx = ⎢ (9.2)
⎢⎣ Ln 1 R zh ⎥⎦

where
R is the reliability (typically R = 90%)
A is the material constant based on strength
Lx is the number of stress cycles, or fatigue life, at (100 − x)% reliability
β is the Weibull shape parameter (Weibull slope)
σ is the stress at the point in the material (typically von Mises)
σ lim is the fatigue limit stress for the material
c is the stress life exponent, which is a material constant
z is the depth of the point in the material
h is the depth exponent, which is also a material constant
V is the stressed volume over which the function is integrated

In this model, the local material stress can be compared to the material strength in a way to best predict
failure for all of the aforementioned RCF modes, hence its popularity.
The current available data for typical high-quality heat-treated steel diverge from the log–log data
when compared with the previous generation of steels. A good example can be seen in the data of
9-6 Theory and Practice of Lubrication and Tribology

10,000
Maximum hertz stress (MPa)

No failures out of
20 test samples

1,000
1.E + 00 1.E + 01 1.E + 02 1.E + 03 1.E + 04 1.E + 05 1.E + 06

106 stress cycles at 90% reliability

FIGURE 9.5  Three ball-on-rod RCF test data for through hardened 52100 steel. (From McKenzie, M., Endurance
strength at 1 × 109 cycles for 52100, in Proceedings of STLE Annual Meeting, Calgary, Alberta, Canada, May 7–11,
2006. With permission.)

McKenzie [7] as shown in Figure 9.5. When good steel and heat-treat practices are utilized, the RCF life
can dramatically increase under very lightly stressed conditions. Looking at the IH model (Equation 9.2),
when the subsurface stresses are concentrated and increased, as would be the case near an inclusion or
other microstructure defect, the model predicts fatigue based on the relevant material properties of A,
c, and h, but most importantly σ lim. If the material includes a significant population of inclusions, car-
bides, or dislocations, then the local material stresses are higher than the material strength and fatigue
under lower surface pressures or applied loads are more likely. From a statistical standpoint, this yields
a lower fatigue limit stress (σ lim) in the IH RCF model. With cleaner steels and advanced heat-treatment
capabilities, the fatigue limit stresses found through testing are increased and the IH model is able to
determine the point at which subsurface plasticity risks are minor. This results in a substantial increase
in fatigue life. This is the fatigue limit concept shown in McKenzie’s [7] RCF data. Since the IH model is
based on the material stress state, it can also predict other modes of fatigue that occur at highly stressed
points in the material, such as under an asperity or under a point on the surface where there is a geo-
metric concentration of the contact pressure. Heat treatment, which can be used to instill compressive
residual stresses, mounting press fits, and/or high-speed operation effects, can also be accounted for in
the IH method.

9.4  Material Considerations


Typical high-quality heat-treated steels used in machine components subject to RCF have material prop-
erties per Harris and Kotzalas [3], Medlin et al. [8], and Dominik [9] as summarized in Table 9.2. The
listed values are typical of carbon vacuum degassed (CVD) ladle-refined steels heat treated to a mini-
mum of 58 HRc, with a fine grain structure and minimal carbides. The material constants listed in Table
9.2 should be confirmed through testing for the specific material and heat-treatment process used in the
manufacture of machine components; however, the values stated within can be used as initial design
guidelines.
Rolling Contact Fatigue Wear 9-7

TABLE 9.2  Typical Fatigue Model Material Properties for Rolling Contact Materials Hardened to
Minimum 58 on the Rockwell C Scale
Parameter AISI 52100 CVD (Through Hardened) SAE 8620/SAE 4320 (Case Carburized)
A (material constant) 552 N9.3/mm1.2 552 N7.56/mm0.89
β (Weibull slope) 10/9 3/2
c (stress exponent) 31/3 34/3
h (depth exponent) 7/3 7/3
σlim (fatigue limit stress) 684 MPa 590 MPa
σr (residual stress) 0 MPa −165 MPa

The most common problem for manufacturing machine components subject to RCF is achieving
the minimum 58 HRc when post heat-treat machineability is required. Reduced hardness indicates
a reduced material yield stress. Reduced hardness decreases the material strength properties A and
σ lim. This is because fatigue is based on plasticity, allowing dislocations in the material to agglomer-
ate and cause fatigue crack formation. Equation 9.3 gives an empirical relationship correlating the
rolling element bearing material constant parameter, and its influence on fatigue life, to material
hardness [3]:

3. 6
⎛ HRc ⎞
Aʹ = A ⎜ (9.3)
⎝ 58 ⎟⎠

Note that Equation 9.3 only adjusts the material constant based on strength and does not adjust the
fatigue limit stress, which is also expected to change. A conservative estimate in this condition would be
to set the fatigue limit to zero; however, this would not be expected in actual operation.
Another aspect that affects performance is the quality of the material. The most common method to
describe quality is by describing the manufacturing process, such as vacuum arc remelting (VAR) or
vacuum induction melting–vacuum arc remelting (VIM-VAR). Both of these processes remelt the steel
in a vacuum to further remove oxide inclusions and enhance the fatigue performance. The problem with
using the steel manufacturing processes to define fatigue performance is that not all manufacturing
plants and equipment are equal. This is shown in Eckel et al. [10] where enhanced air-melted, vacuum
ladle–refined, and bottom-poured steel showed RCF performance equivalent to VAR. In other words,
just because a specific process is used to manufacture the steel, it does not always guarantee the desired
results. It is heavily influenced on the quality processes and standards employed during the manufac-
turing process. It is good practice to ensure the process yields the desired results through RCF testing.
Table 9.3, taken from Eckel et al. [10] and Zaretsky [11], offers initial design guidelines and lists typical
material strength values for various manufacturing processes. It should be noted that the fatigue limit
has not been adjusted and tabulated as a function of steel quality as would be expected with decreased

TABLE 9.3  Steel Manufacturing Process Effect on RCF Material Constant Based on
Strength (A)
Process Effect on Material Constant (A)
Carbon vacuum degassing (CVD) A′ ≈ 1 × A
Vacuum arc remelting (VAR) A′ ≈ 2 × A
Vacuum induction melting–vacuum arc remelting (VIM–VAR) A′ ≈ 3 × A
Sources: Eckel, J. et al., Clean engineered steels—Progress at the end of the twentieth century,
Tech Pub. 1361, ASTM, Conshohocken, PA, 1999; Zaretsky, E., ed., STLE Life Factors for Rolling
Bearings, STLE, Park Ridge, IL, 1992. With permission.
9-8 Theory and Practice of Lubrication and Tribology

inclusion content. This is due to light load testing being cost and time prohibitive for most researchers.
As such, the conservative approach is to use the standard air-melted, CVD fatigue limit stress (Table 9.2)
in the absence of additional data.

9.5  Other Considerations


Although the IH method is used to estimate RCF performance of tribological systems, there are many
different aspects that cannot easily be quantified in a handbook or design guide. Some considerations
include the overall system stress conditions, such as surface texture and third body contamination.
Others are less well known and interact with material strength, for example, water contamination and
lubricant additives.
The debris effect on fatigue has been quantified in the most recent bearing fatigue standard pub-
lished by the International Organization for Standardization (ISO) [12]; however, this standard assumes
all contaminants are equally detrimental to the RCF life. This is not the case as shown by Nixon and
Cogdell [13], where the type of debris in the system causes different amounts of damage to the contact
surface and, therefore, different RCF performance. The best way to include the effect of debris is to ana-
lyze surfaces from the application at different stages in the life cycle. Unfortunately, this is not typically
feasible at the time of machine design.
The effect of surface texture on fatigue life has also been taken into consideration. To date, most of
the published data on surface fatigue are based on empirical data generated on a fixed texture, such
as ground, hard turned, or chemical–mechanical polished, with varying amplitude and lubricant film
thickness [14]. This works well to understand and design machinery when the manufacturing pro-
cesses do not change. It does not allow for quantification of the stress states in the material due to the
texture and net result on RCF performance. Although surface texture is important in RCF, a rough
design guideline is given in Figure 9.6 from Bamberger [15]. This figure is often adequate if the surface

3.5

3.0

2.5

2.0
L΄/L

1.5

1.0

0.5

0
0.6 0.8 1 2 4 6 8 10
hc/RA

FIGURE 9.6  RCF life adjustment for composite surface roughness (R A) and lubricant film thickness (hc). (From
Bamberger, E., ed., Life Adjustment Factors for Ball and Roller Bearings: An Engineering Design Guide, ASME Press,
New York, 1971. With permission.)
Rolling Contact Fatigue Wear 9-9

10
Cantley (18)-SAE 20 + R&O
Cantley (18)-SAE 5 + EP
Barnsby (19)
L΄/L

0.1
10 100 1000
Water concentration in lubricant (PPM)

FIGURE 9.7  Effect of water contamination of the lubricant on RCF life. (From Cantley, R., ASLE Trans., 20, 244,
1976; Barnsby, R., ed., Life Ratings for Modern Rolling Bearings, A Design Guide for the Application of International
Standard ISO 281/2, ASME Press, New York, 2003. With permission.)

roughness is under 0.6 μm. Research into this area, and its implication on the IH and contact mechanic
models, is ongoing [16,17].
Lubricant selection, and its reaction to environmental conditions, also influences RCF. When water
dissolves into a lubricant, the life of a component is decreased. Cantley [18] published data on the effect
of dissolved water, as did Barnsby [19]. Their combined data are given in Figure 9.7, which shows that
both sources are complementary until the point that typical lubricants saturate and free water would be
expected to exist in the system. Equation 9.4 summarizes Figure 9.7 and gives a good design guideline
for adjusting the estimated fatigue life to account for water concentration. This approach adjusts the
estimated life, not the material properties:

0. 6
⎛ 100 ⎞
Lʹ = L ⎜ (9.4)
⎝ PPM ⎟⎠

Unfortunately, the exact mechanism that causes RCF reduction in the presence of water is unknown.
Some researchers have postulated that the effective viscosity is reduced when water is present in the
lubricant and therefore the film thickness is reduced, which, as shown previously, has a negative effect
on life. In contrast, Benner et al. [20] have published data showing that the film thickness is not reduced
when water is dissolved in the lubricant. These data do not fit with the aforementioned theory and more
research is necessary before the mechanism is fully understood.
Finally, the effects of lubricant additives are not completely understood. Again, there are competing
theories. Some speculate the prevention of wear causes high stress states because the asperities do not
wear down during operation. Others speculate that there is a tribochemical attack on the surface, which
results in weakening of the steel microstructure and fatigue crack initiation [21]. In either case, Barnsby
[19] has published potential RCF ranges based on the typical additive package for a given application;
they are listed in Table 9.4. It should be noted that empirical validation of the specific oil and additive
package that will be used with the machine component subject to RCF should be conducted.
9-10 Theory and Practice of Lubrication and Tribology

TABLE 9.4  Potential Lubricant Additive Effect on RCF Life


Range of Actual-to-Predicted
Lubricant Class Fatigue Life (L′/L)
Industrial lubricants
Hydraulic oils 0.6–1.0
Roll bearing oils without antiwear 0.8–1.4
Roll bearing oils with antiwear 0.6–1.0
Turbine oils 0.6–1.0
Circulating oils without antiwear 0.8–1.4
Circulating oils with antiwear 0.6–1.0
Synthetic antiwear gear oils 0.8–1.7
Mineral antiwear gear oils 0.4–1.3
Automotive and aerospace lubricants
Gear lubricants 0.3–0.7
Automatic transmission fluids (ATF) 0.6–1.0
Aviation turbine oils 0.8–1.7
Source: Barnsby, R., ed., Life Ratings for Modern Rolling Bearings, A
Design Guide for the Application of International Standard ISO 281/2,
ASME Press, New York, 2003. With permission.

9.6  Conclusion
While RCF is a common design consideration for machine elements, it is not always understood or
properly considered. The known aspects can be included, such as material stresses and strength; how-
ever, many unknowns still exist. The content within should help guide the person in need of a better
understanding of the physical mechanisms related to RCF acquire some general design guidelines.

References
1. Weibull, W., 1939, A statistical theory of the strength of materials, Proc. R. Swedish Inst. Eng. Res.,
No. 151.
2. Weibull, W., 1949, A statistical representation of fatigue failure in solids, Acta Polytech. Mech. Eng.
Ser. 1, No. 9.
3. Harris, T. and Kotzalas, M., 2007, Rolling Bearing Analysis: Essential Concepts of Bearing Technology,
5th edn., CRC Press, Boca Raton, FL, pp. 208 and 221.
4. Takata, H., Suzuki, S., and Maeda, E., 1985, Experimental study of the life adjustment factor for reli-
ability of rolling element bearings, Proceedings of JSLE International Tribology Conference, Tokyo,
Japan, pp. 603–608.
5. Kotzalas, M., 2005, Statistical distribution of tapered roller bearing fatigue lives at high levels of reli-
ability, Trans. ASME J. Tribol., 127(4), 865–870.
6. Ioannides, E. and Harris, T., 1985, A new fatigue life model for rolling bearings, Trans. ASME J.
Tribol., 107, 367–378.
7. McKenzie, M., 2006, Endurance strength at 1 × 109 cycles for 52100, Proceedings of STLE Annual
Meeting, Calgary, Alberta, Canada, May 7–11.
8. Medlin, D., Cornelissen, B., Matlock, D., Krauss, G., and Filar, R., 1999, Effect of thermal treatments
and carbon potential on bending fatigue performance of SAE 4320 gear steel, SAE Tech. Paper
#1999-01-0603, SAE International, Warrendale, PA.
9. Dominik, W., 1984, Rating and life formulas for tapered roller bearings, SAE Tech. Paper #841121,
SAE International, Warrendale, PA.
Rolling Contact Fatigue Wear 9-11

10. Eckel, J., Glaws, P., Wolfe, J., and Zorc, B., 1999, Clean engineered steels—Progress at the end of the
twentieth century, Tech Pub. 1361, ASTM, Conshohocken, PA.
11. Zaretsky, E., ed., 1992, STLE Life Factors for Rolling Bearings, STLE, Park Ridge, IL.
12. ISO 281, 2007, Rolling Bearings—Dynamic Load Ratings and Rating Life, International Organization
for Standardization, Geneva, Switzerland.
13. Nixon, H. and Cogdell, J., 1998, Debris signature Analysis SM: A method for assessing the detrimen-
tal effect of specific debris contaminated lubrication environments, SAE Tech. Paper #981478, SAE
International, Warrendale, PA.
14. Hashimoto, F., Melkote, S., Singh, R., and Kalil, R., 2009, Effect of finishing methods on surface char-
acteristics and performance of precision components in rolling/sliding contact, Int. J. Machining
Machinability Mater., 6(1–2), 3–15.
15. Bamberger, E., ed., 1971, Life Adjustment Factors for Ball and Roller Bearings: An Engineering Design
Guide, ASME Press, New York.
16. Zhou, R. and Nixon, H., 1992, A contact stress model for predicting rolling contact fatigue, SAE Tech.
Paper #921720, SAE International, Warrendale, PA.
17. Morales-Espejel, G.E., Gabelli, A., and Ioannides, E., 2010, Micro-geometry lubrication and life
ratings of rolling bearings, Proc. IMechE, J. Mech. Eng. Sci., 224(12), 2559–2567.
18. Cantley, R., 1976, The effect of water in lubricating oil on bearing fatigue life, ASLE Trans., 20(3),
244–248.
19. Barnsby, R., ed., 2003, Life Ratings for Modern Rolling Bearings. A Design Guide for the Application
of International Standard ISO 281/2, ASME Press, New York.
20. Benner, J., Sadeghi, F., Hoeprich, M., and Frank, M., 2006, Lubricating properties of water in oil
emulsions, J. Tribol., 128(2), 296–311.
21. Evans, R., Nixon, H., Darragh, C., Howe, J., and Coffey, D., 2007, Effects of extreme pressure additive
chemistry on rolling element bearing surface durability, Tribol. Int., 40(10–12), 1649–1656.
10
Fretting
10.1 Introduction..................................................................................... 10-1
10.2 Mechanics of Fretting..................................................................... 10-2
Contacts and Partial Slip  •  Fretting Tractions  •  Fretting
Stress  •  Fretting Crack Formation
10.3 Fretting Experiments......................................................................10-6
Fretting Test Rig  •  Friction Evolution
10.4 Design for Fretting..........................................................................10-8
Advanced Fretting Contact Analysis  •  Life Estimation
Thomas N. Farris 10.5 Fretting in Practice........................................................................ 10-10
Rutgers University
Recognizing Fretting  •  Methods to Alleviate Fretting
N. Sundaram 10.6 Summary......................................................................................... 10-11
Purdue University References................................................................................................... 10-12

10.1  Introduction
Fretting may be defined as a friction-driven contact phenomenon in which damage occurs in the vicin-
ity of the contact of two nominally clamped surfaces. In the presence of cyclic tangential loads, the two
contacting bodies undergo small-scale, oscillatory, relative tangential motions known as “microslip.”
This slipping motion may be localized so that the contacting surfaces do not exhibit global relative
motion. Typical microslip amplitudes are on the order of 10–100 μm. The physical mechanism of the
damage involves a combination of wear, corrosion, and fatigue. These, in turn, are driven by high stress
gradients near the contact and microslip. The resulting nucleated fretting crack may grow in the pres-
ence of an external cyclic stress field, ultimately resulting in failure of the component. While fretting
damage is very localized, it can have a huge impact on the fatigue life of an engineering component,
reducing it by as much as 40%–60% (Waterhouse, 1972).
Fretting was reported as early as 1911 by Eden et al. (1911), but Tomlinson (1927) is credited with the
first experimental investigation using a machine to generate small relative rotations between two rings.
The fact that fretting exacerbates fatigue and reduces life was noted by McDowell (1953). Early evolution
of the friction coefficient was observed by Nishioka and Hirakawa (1969) and Endo and Goto (1976)
and details of the fretting damage process and microwelding were described by Waterhouse and Taylor
(1971) and Waterhouse (1972). Classical elastic contact mechanics began with the work of Hertz (1882),
who modeled frictionless spherical contacts. Subsequently, important work done by Cattaneo (1938)
and Mindlin (1949) provided the ability to analyze contacts in the presence of tangential forces. Elastic
contact mechanics is a key tool to analyze fretting.
Fretting is now widely recognized as the initiator of cracks in numerous engineering components,
many related to aerospace structures. Among the notable ones are blade-disk dovetail joints in gas-
turbine engines (Kalb, 1991) and riveted lap joints in aircraft structures (Iyer et al., 1995; Muller,
1995). Fretting has been observed in numerous other instances such as the landing gear of aircraft,

10-1
10-2 Theory and Practice of Lubrication and Tribology

High-frequency
aeroelastic drivers
and system vibrations

Rotating disk
Contact (bulk inertial

∆Q
surface stress)

P Blade
root

FIGURE 10.1  Schematic of loads that cause microslip and fretting in a blade-disk dovetail joint.

engine seals, rocket propulsion fuel tanks, and even artificial hip joints (Waterhouse, 1981). The sche-
matic of a blade-disk dovetail joint in Figure 10.1 shows the kinds of oscillatory loads that drive fret-
ting fatigue.

10.2  Mechanics of Fretting


Understanding the mechanics of the fretting contact is crucial to understand fretting. Methods devel-
oped for use in classical elastic contact mechanics have been applied successfully to fretting contacts.
A very brief introduction is provided here; Gladwell (1980), Hills et al. (1993), and Hills and Nowell
(1994) provide an extensive treatment of this subject.

10.2.1  Contacts and Partial Slip


Contact mechanics is a branch of mechanics concerned with the determination of contact extents, con-
tact pressure, and shear tractions and stresses caused by a fretting contact. The concept of partial slip
arises naturally when contact involves deformable bodies in the presence of friction. The contacting
bodies experience no relative motion in a central region called the stick zone of the contact. In the
peripheral slip zones, microslip occurs and damage accumulates.
Given the geometry and profiles of the contacting bodies, their elastic/plastic material properties,
interface properties such as the coefficient of friction and applied load history, it is possible to determine
the contact state and contact stresses in principle. In practice, one makes numerous assumptions about
each of these aspects to simplify the description of the fretting contact. The most common assumptions
are as follows:
1. The material behavior is elastic for the purposes of contact analysis.
2. The contact size is small in comparison to the size of the bodies, typically <0.1 times some
­characteristic dimension such as the radius of curvature.
3. Inertial effects may be ignored.
4. The influence of asperities may be ignored, that is, the contact may be modeled using the nominal
(smooth) profile.
5. The presence of a fretting crack does not influence the contact tractions.
6. Three-dimensional effects are significant only close to free edges, so plane-strain analysis is valid.
These assumptions, while seemingly restrictive, are backed by both theory and experimental evidence
to provide a good first-order description of the fretting contact. They allow one to treat the contact-
ing bodies as a pair of semi-infinite half-planes. The governing equations for the pressure and shear
in such bodies are then a pair of singular integral equations. These equations allow determination
Fretting 10-3

of the contact tractions, contact extents, and so on in closed form for a subset of profile geometries
and loads.

10.2.2  Fretting Tractions


A commonly occurring fretting contact configuration involves a pair of cylinders (or one cylinder and a
flat). In these cases, the pressure traction p(x) is described very well by an elliptical approximation. This
description is exact if the materials are elastically similar:

2
⎛ x⎞
p (x ) = p0 1 − ⎜ ⎟ (10.1)
⎝ a⎠

where the peak pressure, p0, and the contact half-length, a, are given by

PE ∗ 4PR ∗
p0 = a= (10.2)
pR ∗ pE ∗

P is the applied normal load per unit length (dimensions of N/mm). R* and E* are, respectively, the
composite radius of curvature and composite material stiffness given by

1 1 1 1 1 − n12 1 − n22
= + = + (10.3)
R∗ R1 R2 E∗ E1 E2

Under the assumption that the tangential load per unit length, Q, does not exceed μP in magnitude,
the partial slip condition described earlier prevails. Here, μ is the coefficient of friction. For |Q| > μP,
bulk sliding occurs. The shear traction q(x) is obtained by superposing two different elliptic tractions,
q(x) = q′(x) + q′(x) and q′(x) = sign(Q)μp(x), and q′(x) is described by different expressions in the stick
region and the slip zones:

mp0 2
q "(x ) = −sign(Q) c − x2 | x |< c (10.4)
a

q "(x ) = 0 c <| x |< a (10.5)


Note that the signs of q(x) will be reversed if the sign of the applied tangential load Q is reversed. If these
shear tractions are used in cyclic tangential loading calculations, please note that they are valid only at
the extremes of the cycling. Intermediate positions in the cycling have shear tractions that are partially
reversed. Closed-form expressions for such tractions are given in Szolwinski and Farris (1996). The ratio
of the stick-zone size to the contact size, c/a, is given by

c Q
= 1− (10.6)
a mP

Typical fretting tractions for partial slip contact of cylinders, including partially reversed shear trac-
tions, are shown in Figure 10.2.
10-4 Theory and Practice of Lubrication and Tribology

0.8

0.6

0.4
Q = 0.38 P

µ p(x)/p0, q(x)/p0
0.2
Q=0
0
(Partially
reversed)
–0.2
Q = –0.38 P
–0.4

–0.6

–0.8
–1 –0.5 0 0.5 1
x/a

FIGURE 10.2  Typical partial slip shear tractions. Note the shear tractions on partial reversal of the shear load.

10.2.3  Fretting Stress


In the absence of remote stresses, the peak trailing-edge stress on the surface is given by

mQ
s xx |max =2 p0 (10.7)
P

The following approximation has been found to model the edge-of-contact peak stress very well for
remote stresses in phase with the shear load Q:

mQ
s xx |total = 2 p0 + s0 (10.8)
P

By using representative values for p0, Q/P, slip zone μ, and bulk stress, it is possible to estimate the stress
concentration (notch factor) max(σxx)/σ 0 provided by the fretting contact.
Typical values are between 2.50 and 3.00.

10.2.4  Fretting Crack Formation


The damage associated with the fretting contact has been described as a “synergistic competition” between
wear-, corrosion-, and fatigue-related phenomena and is a three-stage process in metals (Szolwinski
and Farris, 1996). In the first stage, the thin oxide layer is removed. This only takes a few fretting cycles.
Subsequently, the underlying bare metal adheres forming “microwelds.” This adhesive process and initial
gathering of wear debris is associated with a rise in the friction coefficient at the beginning of a fretting test.
Additional cycles of fretting cause near-surface plastic deformation and finally a series of microcracks com-
parable to the grain size. Some of these cracks may be worn away in subsequent cycles, but some propagate a
few hundred microns into the surface. Characteristically, the influence of contact stresses dies away rapidly
at these depths. Subsequent propagation/growth of this crack is driven entirely by far-field bulk stresses.
Using crossed cylinders (a first-order correct spherical contact) and imposing tangential dis-
placement amplitudes from 1 to 35 μm on the contact, Vingsbo and Soderberg (1988) showed that
Fretting 10-5

Mixed stick Gross slip


and slip
Reci-
procat.
10–14 sliding 107

Fatigue life (number of cycles)


Stick
Wear (m3/Nm)

10–15 106

10–16 105
1 3 10 30 100 300 1000
∆ (µm)

FIGURE 10.3  Schematic of the dependence of fretting-fatigue life and wear rate on displacement amplitude Δ.
(Adapted from Vingsbo, O. and Soderberg, S., Wear, 126, 131, 1988.)

it is possible to define three fretting regimes, each exhibiting a different type of surface damage
(Figure 10.3):
1. Near-stick regime corresponding to relatively small amplitudes (≤2 μm) exhibits limited fretting
damage and no crack formation up to a million cycles.
2. Partial slip regime (called “mixed stick–slip” by Vingsbo and Soderberg) corresponding to inter-
mediate amplitudes, limited wear, and extensive fretting crack formation associated with fretting
fatigue.
3. Gross slip regime corresponding to large amplitudes and severe wear-driven surface damage,
with limited crack formation.
By plotting the (post-test) micrographically observed fretting regime as a function of two indepen-
dent variables, such as displacement amplitude and normal load, the method of Vingsbo and Soderberg
(1988) allows construction of fretting maps showing the expected type of damage that occurs under
particular loading conditions and transitions between such regimes. If the tangential displacement
amplitude is not independently controllable or known accurately, a fretting map generated by varying
tangential loads may be used instead. The reduced propensity for crack formation at large amplitudes is
due to the grinding away of cracks at an early stage of nucleation (Hills and Nowell, 1994).
The starting point for fretting crack nucleation models is the Smith–Watson–Topper (SWT) equa-
tion (Smith et al., 1970), which also considers mean stress effects in contrast to conventional strain-life
approaches:

⎛ Δe⎞ 1
Γ = s max ⎜ ⎟ = (s ʹf )2 (2N f )2b + s ʹf eʹf (2N f )b + c (10.9)
⎝ 2 ⎠ E

Plotting contours of the multiaxial parameter Γ for fretting contacts shows that the critical crack
nucleation location is the trailing edge of the contact; the critical plane at this location is oriented
10-6 Theory and Practice of Lubrication and Tribology

perpendicular to the contact surface, corresponding with experimental observations. Contours of


the multiaxial Γ parameter under typical fretting contact conditions may be found in Szolwinski
and Farris (1996).

10.3  Fretting Experiments


Fretting is a complex, multifactorial phenomenon and involves the interaction of many different aspects.
These may be broadly grouped as follows (Murthy, 2004):
1. Manufacturing aspects—surface profiles, manufacturing tolerance, and heat treatments
2. Materials properties—the fatigue and fracture behavior
3. Tribological aspects—friction coefficient evolution, wear, and wear-driven profile changes
4. Mechanics—fretting loads acting on the contact
At one time, the number of variables was considered too high to allow meaningful general purpose
fretting experiments, with estimates as high as 50 individual factors affecting fretting (Dobromirski,
1992). Historically, most studies were application specific and the results could not be transferred to a
different setting. However, modern fretting research has shown that it is possible to reduce this number
significantly and yet retain essential characteristics of fretting by the use of a “bridge-type” fretting test
rig described by Hills et al. (1988).

10.3.1  Fretting Test Rig


A schematic of the fretting test rig is shown in Figure 10.4. The test uses a standard dogbone-type speci-
men clamped at either end into a conventional servohydraulic fatigue test machine. The key feature is
the presence of a fretting chassis, which is an added structure that allows for the generation of cyclic
tangential loads, while keeping the normal clamping load approximately constant. These tangential
loads are in phase with the bulk stress σ 0. The normal load is applied to the specimen via a pair of pads
machined to the desired nominal profile radius. The pads are pushed against the specimen by hydraulic
actuators and fixed tightly to the chassis by using pad-tops. The pad holding platforms are mounted on
thin steel diaphragms, which have a very low resistance to pressure loading but resist tangential loading.

Cross-head and upper load cell


σ

Pad-top
Pressure load cell
X
Hydraulic actuator
P Y P
Q Q

Z
Stiff beam
H
Diaphragm Top platform R

F0
σ0
(a) Actuator and lower load cell (b)

FIGURE 10.4  (a) Schematic and (b) mechanics of fretting.


Fretting 10-7

This essentially allows transmission of most of the applied pressure to the contact, while providing suf-
ficient stiffness to generate a large tangential force.
A pair of pressure rods with pressure load cells allows equal normal loads to be applied to both con-
tacts. The width of the specimen (15.24 mm) is chosen so that the two contacts do not interact with each
other. This value is chosen so that w ≥ 10a. More details of the dimensions and design of the specimen
and pad may be found in Murthy (2004).
A simplified model of mechanics of this rig may be obtained by treating the specimen and diaphragms
as linear springs. This model shows that by altering the vertical location at which the pad meets the
specimen, one may, in effect, control the Q/P ratio, which is a crucial experimental parameter. The lower
part l1 of the specimen is treated as a spring of stiffness k1 = AsEs/l1, the part above the contact a spring of
stiffness k3 = AsEs/l3, and each diaphragm spring of stiffness k2 = AdEd/ld. If the force applied at the lower
actuator is F0 and its displacement δ 0, and the displacement of the contact is δc, it is possible to show that
the forces and displacements are related as

F0 = k1(d − dc ) F = k3dc Q = k2dc (10.10)


For vertical equilibrium, one must have F0 = F + 2Q

F0 F0
F= Q= (10.11)
1 + 2k2 /k3 2 + k3 /k 2

For the same applied bulk load F0, it is possible to obtain different shear loads Q by changing k3 (k2 is a
constant), which is accomplished by clamping the pad at different position to change l3. Typical experi-
mental procedures in fretting are quite involved, with several checks to ensure correct alignment, the
formation of a good contact, and so on (Gean, 2008). The bulk of the fretting test consists of applying a
cyclic axial load at about 10 Hz or so to the aligned, clamped specimen while a constant normal load is
applied to the contact. Tests are typically run until the specimen fails or are stopped at a predetermined
“run-out” point, such as million cycles.

10.3.1.1  High-Temperature Fretting Tests


In several fretting applications such as a gas-turbine high-pressure stage, the fretting contact is subjected
to high temperatures that significantly affect fretting life. The fretting chassis described earlier may be
modified to perform such testing. In this case, components of the rig closer to the high-­temperature
zone are built using Ti6Al4V, while 4140 steel is used for component further away. To avoid thermal
expansion mismatches, the high-temperature webs/platform are machined from a single Ti6Al4V block.
Local heating of the contact is provided by using a pair of igniters, one on each side of the contact and
in close proximity to it. The igniter surface heats up to about 1550°C. Radiative and convective trans-
fer heats the contact. A thermocouple is used to measure the contact temperature and used to provide
on–off control of the igniters. A ceramic block and sheets are used to minimize fluctuations due to
external heat transfer and provide a furnace-like environment around the contact.
The system provides a target temperature of 610°C ± 5°C.
To protect the surface of the diaphragms and pad holder from the high temperatures, they are covered
with ceramic sheets. The pad holders are also cooled with circulating water through specially machined
channels. More details of high-temperature fretting tests may be found in Murthy (2004).

10.3.2  Friction Evolution


As mentioned, a contact is in a state of partial slip if |Q| < μP and slides if the magnitude of Q exceeds
this value. However, in a fretting test, the friction coefficient is not a constant, but evolves temporally
10-8 Theory and Practice of Lubrication and Tribology

TABLE 10.1  Friction Coefficient


Evolution in Inco718/Ti6Al4V
Fretting Contact
Nfret μaverage μslip
2,500 0.496 0.587
10,000 0.537 0.630
20,000 0.521 0.600
50,000 0.597 0.791
80,000 0.590 0.726

and spatially as the test progresses. The mechanism is somewhat complicated and is only briefly dis-
cussed here. In the first few cycles, the friction coefficient is low and bulk slipping occurs. The surface
oxide layer is removed in the first few cycles and the value of μ continues to increase. At a certain stage,
μ becomes large enough for partial slip to set in. Once this occurs, the friction coefficient ceases to evolve
in the central stick zone, but continues growing in the peripheral slip regions. The rate of friction coef-
ficient increase in the slip zones slows as the cycling proceeds. Spatially, the edges of the contact have
the highest value of μ.
A separate experiment known as a friction test is required to characterize an average coefficient of
friction. In this test, a normal fretting test is stopped after a specified number of cycles. Without chang-
ing the pad or specimen, an increasing amplitude sinusoidal waveform is applied to the specimen. This
causes the tangential force to increase during each cycle; ultimately, this value of Q is sufficient to cause
bulk sliding and this event is typically characterized by a slight drop in the measured value of Q/P and a
notable acoustic event, both of which allow one to mark the onset of sliding and stop the test. The initial
coefficient of friction, μ 0, the average coefficient of friction in the friction test, μ, and the slip-zone fric-
tion coefficient, μs, are then related by the following expression (Harish, 2000):

ms
2 ⎛ Q Q Q ⎞
ms = m0 +
p ∫ ⎜⎝ m P
m0
s
1−
ms P
+ sin −1 1 −
ms P ⎟⎠
d ms (10.12)

This equation relies on several assumptions, the most important one being that in a given cycle μs is
spatially constant. To give an estimate of the numerical values involved, μ 0 is approximately 0.10–0.15,
the stable μs about 0.45–0.65, and μ about 0.30–0.55 in Al–Al. These values depend strongly on the
contact material pair. Table 10.1 shows the evolution of average and slip-zone friction coefficients in an
Inco718/Ti6Al4V contact pair (Rajeev, 2001). Please note that more sophisticated, alternate methods of
calculating μs exist (Murthy, 2004). It is recommended to use μs in design/life calculations because it
provides estimates that are more conservative.

10.4  Design for Fretting


10.4.1  Advanced Fretting Contact Analysis
If the contact profiles are complicated, or the effect of the remote stress on the tractions or material
elastic dissimilarity have to be modeled exactly, semianalytical methods may be used to solve the gov-
erning equations of fretting contacts (Hills and Nowell, 1994). Semianalytical procedures have been
developed for elastically similar contacts (Murthy et al. 2004), dissimilar or anisotropic contacts (Rajeev
and Farris, 2002, Rajeev et al. 2004), and multiple contacts (Sundaram and Farris, 2009).
Fretting 10-9

Alternately, one may use the finite element method to solve for fretting contacts. Modern FEA
solvers such as ABAQUS provide good-quality contact analysis capabilities. Even so, obtaining
high-quality contact FEA results is nontrivial; the challenges in contact FEA may be summarized
as follows:

1. The small size of the contact patch compared to the characteristic dimensions of the contacting
bodies (sometimes can be as many as three orders of magnitude).
2. Inability of standard mesh makers to generate optimized contact meshes.
3. FEA solvers may need matched meshes along the line of contact, placing considerable restrictions
on the mesh geometry in the vicinity of the contact.
4. Contact FEA analysis is inherently nonlinear and analyses times can be as many as two orders of
magnitude larger than linear, static analyses.

Item (1) implies that some scheme is needed that allows elements of quite disparate sizes in the same
mesh and a suitable method to transition between them. Automatic meshing tools that accompany
standard analysis software do not usually yield satisfactory results in such cases. Further, naive element
selection by mesh makers may generate triangular elements along contact regions, which leads to volu-
metric locking and analysis failure. Apart from the additional contact degrees of freedom, the stiffness
matrix in contact procedures has a large wave front, which can drastically increase analysis time. Every
effort must be made to optimize the contact mesh.

10.4.2  Life Estimation


Once the contact stresses for an application have been determined either by using an analytical descrip-
tion or semianalytical methods as described in earlier sections, cyclic stresses and strain fields may be
calculated. FEA may also be used for this purpose, but it is computationally more expensive than the
other options.
The objective is to use a fatigue life model such as that described by Equation 10.9 to predict fret-
ting crack nucleation life. The smallest life (2Nf) corresponds to the maximum value of Γ = σmaxΔ/2.
The process is time consuming, as it requires stepping through each point in the shear-load cycle,
with stress calculations performed at a grid of subsurface physical locations at each increment and
further examination of planes of different orientation at each grid point. Elastic stress transforma-
tion laws must be used for the latter and the plane-strain Hooke’s law then provides the strain. By
stepping through an entire loading cycle, the values of εmax, εmin, and σmax are known at each plane at
each point. The stress σmax corresponds to the maximum stress normal to the plane experienced by it
during the load cycle. With the maximum Γ known at each spatial location, a final global maximum
damage parameter Γmax may be calculated (Szolwinski and Farris, 1996). Exhaustive spatial study
is not needed in every case and it may suffice to confine attention to the trailing edge of contact
(Szolwinski, 1998).
The total life is considered to consist of two separate regimes: a nucleation life, which is the cycles to
nucleate a crack of roughly between a and 2a in size, and propagation life to cause failure of the speci-
men. The rationale behind using a depth between a and 2a (order of magnitude 1 mm deep) is related
to the fact that the effect of the contact is negligible at these depths. The propagation phase is thus
independent of the contact and considered to propagate in Mode I under the exterior σ 0 stress only. The
nucleated crack has been shown to be a semielliptical surface crack, for which a K I stress intensity factor
may be computed, and a Paris-type law used to estimate propagation life.
For fretting experiments, an experimental nucleation life may be computed by defining Nnucleation,​
expmt = Nfailure,expmt − Npropagation and it may be compared with predicted Nnucleation using the
SWT parameter.
10-10 Theory and Practice of Lubrication and Tribology

350
50

300

σmax (ksl) = 2po (µQ/P)1/2+ σo


40
250

σmax (MPa)
30 200

150
20
2024-T351 strain-life curve 100
Hills and Nowell (1990)
10 Current experimental data
3× upper bound 50
3× lower bound

0
105 106 107
Cycles to nucleation

FIGURE 10.5  Plot of maximum σxx stress amplitude at the trailing edge of contact plotted against the experi-
mental crack nucleation life. The solid line is the predicted lifetime using the multiaxial fatigue parameter. (From
Szolwinski, M.P., The mechanics and tribology of fretting fatigue with application to riveted lap joints, PhD thesis,
School of Aeronautics and Astronautics, Purdue University, West Lafayette, IN, August 1998. With permission.)

Investigations reveal good agreement between nucleation life as predicted using the multiaxial
approach and experimental nucleation life as defined earlier, for different metals such as the Al-2024-T351
alloy (Figure 10.5). Observations show that the nucleation life is typically 85%–90% of total life, so more
complicated propagation models do not add much by way of predictive capability. From a design per-
spective, given the contact conditions and contact stress analysis, the total life may thus be predicted as
Nfailure,predicted = Nnucleation,predicted + Npropagation.
Finally, it must be mentioned that several alternate models to the multiaxial fatigue approach exist
such as the equivalent stress parameter approach used by Murthy et al. (2001).

10.5  Fretting in Practice


10.5.1  Recognizing Fretting
Fretting is recognized by the characteristic oxidized wear debris found in the vicinity of the contact. In
fact, this was among the earliest observations of fretting by Eden et al. (1911). In addition, fretting cracks
emanate from the edge of contact, and post-failure examination reveals a characteristic wear scar. In
particular, a mostly undamaged region of stick surrounded by a significantly worn slip region may be
seen. In fretting-fatigue tests with dogbone specimens, the wear scar is an oval-shaped, “race-track”
pattern (Gean, 2008).

10.5.2  Methods to Alleviate Fretting


The following methods have been used to alleviate fretting:
1. Shot peening—Shot peening induces subsurface compressive residual stress by cold working.
Subsurface residual stresses are typically induced by high-velocity bombardment of the compo-
nent with metallic shots. Fretting tests in Ti have shown that shot peening increases life by as
Fretting 10-11

much as 100% at room temperature. Due to thermal relaxation of residual stresses, the efficacy of
shot peening is reduced at higher temperatures (Gean, 2008).
2. Laser shock peening—In laser shock peening, residual compressive stresses are imparted to
the surface by using a high-power pulsed laser. Room temperature fretting-fatigue tests on
shock-peened Ti6Al4V show life improvements of up to 10× (Srinivasan et al., 2009). As a
residual stress–based technique, the efficacy of LSP at high temperatures is expected to be
lower.
3. Coatings—Coatings such as CuNiIn and AlBr have been used to reduce fretting in Ti17 and
Ti64Al contacts. However, fretting experiments show (Gean and Farris, 2009; Murthy et al., 2009)
that such coatings are removed early in the fretting process and provide negligible improvement
in fretting-fatigue life.
4. Lubrication—Lubricants like MoS2 and graphite provide little improvement in life (Gean and
Farris, 2009).
5. Contact elimination—By using monolithic “blisks,” the number of blade-disk dovetail joints (and
hence, fretting contacts) in gas-turbine engines may be reduced.
6. Undercuts—This is another dovetail-specific design change that separates stress concentrations
from the edge of contact and the minimum neck-radius stress concentration. Fretting tests on
R88 specimens with a simulated undercut show life improvements up to 30% (Gean, 2008).

10.6  Summary
Critical learning from fretting-fatigue tests and analyses is as follows:
1. Bridge-type fretting tests allow controlled, effective, and reproducible studies of fretting.
2. Fretting cracks nucleate at the trailing edge of contact and are propagated to failure by the remote
stress field.
3. Accurate modeling of edge-of-contact stresses is essential for fretting design.
4. The friction coefficient evolves spatially and temporally during testing.
5. Slip-zone friction coefficients may be estimated effectively.
6. Conventional life-prediction techniques may be used to estimate fretting life.
7. Nucleation life occupies the dominant share of total life.
8. Coatings and lubricants are not effective in reducing fretting fatigue.
9. Residual stress based alleviation techniques are less effective at elevated temperatures.
10. Higher-order effects may usually be ignored in fretting-fatigue tests.
A large body of the fretting literature pertains to fretting behavior/life observed using different types of
material contact pairs, profiles, temperatures, or surface treatments. Table 10.2 provides a selection of
primary references.

TABLE 10.2  Summary of Fretting Studies


Material (Specimen) Temperature Surface Treatment Reference
Al-2024 Room temperature None Szolwinski and Farris (1998)
Ti6Al4V Room temperature None Murthy et al. (2001)
Ti6Al4V Room temperature Various Murthy et al. (2009)
Ti6Al4V Room temperature LSP Srinivasan et al. (2009)
Ti17 High temperature Various Gean and Farris (2009)
Super alloys High temperature None Gean et al. (2009)
SCN High temperature None Matlik et al. (2009)
10-12 Theory and Practice of Lubrication and Tribology

References
Cattaneo, C., Sul contatto di due corpi elastici: Distribuzione locale degli sforzi, Rendiconti dell’ academica
nazionale die lincei, 1938.
Dobromirski, J.M., Variables of fretting process: Are there 50 of them? in Standardization of Fretting
Fatigue: Test Methods and Equipment, eds. M.H. Attia and R.B. Waterhouse, ASTM STP 1159,
American Society of Testing and Materials, Philadelphia, PA, 1992, pp. 60–66.
Eden, E.M., W.N. Rose, and F.L. Cunningham, The endurance of metals, Proc. Inst. Mech. Eng., Parts 3–4
(1911), 831–880.
Endo, K. and H. Goto, Initiation and propagation of fretting fatigue cracks, Wear, 38 (1976), 311–324.
Gean, M.C., Elevated temperature fretting fatigue of nickel based alloys, PhD thesis, School of Aeronautics
and Astronautics, Purdue University, West Lafayette, IN, August 2008.
Gean, M.C. and T.N. Farris, Elevated temperature fretting fatigue of Ti-17 with surface treatments, Tribology
Int., 42(9) (September 2009), Fifth International Symposium on Fretting Fatigue, pp. 1340–1345.
Gean, M.C., N. Tate, and T.N. Farris, Fretting fatigue of nickel based superalloys at elevated temperature, in
50th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference, Palm
Springs, CA, 2009, pp. 2009–2626.
Gladwell, G.M.L., Contact problems in the classical theory of elasticity, in Monographs and Textbooks
on Mechanics of Solids and Fluids, Sijthoff & Noordhoff, Adphen aan den Rijn, the Netherlands,
1980.
Harish, G., Coupled thermoelastic analysis of fretting contacts, PhD thesis, School of Aeronautics and
Astronautics, Purdue University, West Lafayette, IN, December 2000.
Hertz, H., Ueber die Berührung fester elastischer Körper, Journal für die reine und angewandte
Mathematik, 92 (1882), 156–171.
Hills, D. and D. Nowell, Mechanics of Fretting Fatigue, Vol. 30 of Solid Mechanics and Its Applications,
Kluwer Academic Publishers, Dordrecht, the Netherlands, 1994.
Hills, D.A., D. Nowell, and J.J. O’Connor, On the mechanics of fretting fatigue, Wear, 125 (1988), 129–146.
Hills, D.A. and D. Nowell, Crack initation criteria in fretling fatigue, Wear, 136 (1990), 329–343.
Hills, D., D. Nowell, and A. Sackfield, Mechanics of Elastic Contacts, Butterworth Heinemann, Oxford, U.K.,
1993.
Iyer, K., G.T. Hahn, P.C. Bastias, and C.A. Rubin, Analysis of fretting conditions in pinned connections,
Wear, 181–183 (1995), 524–530.
Kalb, B.J., Friction stresses between blade and disk dovetail possible case of numerous dovetail problems,
in Proceedings of USAF Structural Integrity Conference, San Antonio, TX, 1991.
Matlik, J.F., T.N. Farris, F.K. Haake, G.R. Swanson, and G.C. Duke, High-frequency, high-temperature fret-
ting-fatigue experiments, Wear, 261(11–12) (2006), 1367–1382.
Matlik, J.F., T.N. Farris, J. Haynes, G.R. Swanson, and G. Ham-Battista, Prediction of fretting crack location
and orientation in a single crystal nickel alloy, Mech. Mater., 41(10) (2009), 1133–1151.
McDowell, J.R., Fretting corrosion tendencies of several combinations of materials, ASTM STP 144:
Symposium on Fretting Corrosion, Philadelphia, PA, 1953, pp. 24–39.
Mindlin, R.D., Compliance of elastic bodies in contact, J. Appl. Mech., 16 (1949), 259–268.
Muller, R.P.G., An experimental and analytical investigation on the fatigue behavior of fuselage riveted lap
joints, PhD thesis, Delft University of Technology, Delft, the Netherlands, 1995.
Murthy, H., Fretting fatigue of anisotropic materials at elevated temperatures, PhD thesis, School of
Aeronautics and Astronautics, Purdue University, West Lafayette, IN, December 2004.
Murthy, H., G. Harish, and T.N. Farris, Efficient modeling of fretting of blade/disk contacts including load
history effects, ASME J. Tribology, 126 (2004), 56–64.
Murthy, H., G. Mseis, and T.N. Farris, Life estimation of Ti-6Al-4V specimens subjected to fretting fatigue
and effect of surface treatments, Tribology Int., 42(9) (2009), Fifth International Symposium on
Fretting Fatigue, pp. 1304–1315.
Fretting 10-13

Murthy, H., P.T. Rajeev, T.N. Farris, and D.C. Slavik, Fretting fatigue of Ti-6Al-4V subjected to blade/disk
contact loading, in Developments in Fracture Mechanics for the New Century, 50th Anniversary of
Japan Society of Materials Science, Osaka, Japan, 2001, pp. 41–48.
Nishioka, K. and K. Hirakawa, Fundamental investigations of fretting fatigue, part 3, Bull. JSME, 12(51)
(1969), 397–407.
Rajeev, P.T., Fretting contact of dissimilar isotropic/anisotropic materials, PhD thesis, School of Aeronautics
and Astronautics, Purdue University, West Lafayette, IN, December 2001.
Rajeev, P.T. and T.N. Farris, Numerical analysis of fretting contacts of dissimilar isotropic and anisotropic
materials, J. Strain Anal. Eng. Des., 37(6) (2002), 503–517.
Rajeev, P.T., H. Murthy, and T.N. Farris, Load history effects on fretting contacts of isotropic materials,
ASME J. Eng. Gas Turbines Power, 126 (2004), 385–390.
Smith, K.N., P. Watson, and T.H. Topper, A stress–strain function for the fatigue of metals, J. Mater., 5(4)
(1970), 767–778.
Srinivasan, S., D.B. Garcia, M.C. Gean, H. Murthy, and T.N. Farris, Fretting fatigue of laser shock
peened Ti-6Al-4V, Tribology Int., 42(9) (2009), Fifth International Symposium on Fretting Fatigue,
pp. 1324–1329.
Sundaram, N. and T.N. Farris, Multiple contacts of similar elastic materials, ASME J. Tribology, 131 (2009),
021405.
Szolwinski, M.P., The mechanics and tribology of fretting fatigue with application to riveted lap joints,
PhD thesis, School of Aeronautics and Astronautics, Purdue University, West Lafayette, IN, August
1998.
Szolwinski, M.P. and T.N. Farris, Mechanics of fretting fatigue crack formation, Wear, 198 (1996), 93–107.
Szolwinski, M.P. and T.N. Farris, Observation, analysis and prediction of fretting fatigue in 2024-T351
aluminum alloy, Wear, 221(1) (1998), 24–36.
Tomlinson, G.A., The rusting of steel surfaces in contact, Proc. R. Soc. Ser. A, 115 (1927), 472–483.
Vingsbo, O. and S. Soderberg, On fretting maps, Wear, 126 (1988), 131–147.
Waterhouse, R.B., Fretting Corrosion, Pergamon Press, Oxford, U.K., 1972.
Waterhouse, R.B., Fretting fatigue in aqueous electrolytes, in Fretting Fatigue, Applied Science, London,
U.K., 1981, pp. 159–176.
Waterhouse, R.B. and D.E. Taylor, The initiation of fatigue cracks in a 0.7% carbon steel by fretting, Wear,
17 (1971), 139–147.
11
Erosion
11.1 Introduction..................................................................................... 11-1
11.2 Experimental Facilities Used in Erosion Research..................... 11-1
11.3 Erosion Test Results......................................................................... 11-3
11.4 Effect of Particle Impact Velocity.................................................. 11-4
11.5 Effect of Particle Impact Angle...................................................... 11-4
11.6 Effect of Temperature...................................................................... 11-4
11.7 Effect of Particle Size and Loading............................................... 11-5
11.8 Effect of Particle Hardness, Sharpness, and Mineral
Composition..................................................................................... 11-6
11.9 Volcanic Ash..................................................................................... 11-6
Awatef A. Hamed 11.10 Roughness of Eroded Surfaces...................................................... 11-6
University of Cincinnati References..................................................................................................... 11-7

11.1  Introduction
Erosive particles can be found in many applications such as pneumatic conveying, food processing,
surface blasting, fluidized beds, and gas turbines. Whether their presence is induced or unavoid-
able, particles impact these systems’ surfaces and cause erosion damage. The resulting changes in the
impacted surface geometry and finish adversely affect the performance and life of the equipment.
Extensive research efforts have been directed at developing materials and coatings that are resistant to
erosion and at eliminating or reducing the frequency and intensity of particle surface impacts. Erosion
rate is known to be affected by both impact velocities and impingement angles of the erosive particles.
It is also affected by the surrounding temperature, material properties of the abrasive particles and the
impacted surfaces, and the angularity or roundness of particles.

11.2  Experimental Facilities Used in Erosion Research


Various types of facilities have been used in experimental erosion research. They include jet blasting,
whirl arm, centrifugal acceleration, and erosion wind tunnels. In general, the test range of particle
velocity depends on the facility design parameters. Velocities as high as 350 m/s were reported in cen-
trifugal accelerators, and 800 m/s in whirling arm facilities (Klies and Kulu 2008). However, the control
of particle impact conditions varies considerably depending on the type of erosion test facility.
Jet blasting stationary specimens with particles entrained in a small gas jet is the most commonly used
method in erosion testing. According to Klies and Kulu (2008), pneumatic test devices equipped with
a nozzle already existed in 1904. Erosion test results using such facility were reported by Finnie (1960),
Sheldon (1970), Tilly (1973), and Hutchings (1983). Gauge pressure is used to control particle impact
velocities in these facilities, up to 130 m/s, mostly at room temperatures. However, because of the spread-
ing and mixing of the jet flow, the erosion data obtained at a given test condition actually encompasses

11-1
11-2 Theory and Practice of Lubrication and Tribology

a range of particle impingement velocities and impact angles. The deviation of actual particle impact
velocities from the mean depends on the size and shape of the particles, the distance between the speci-
men and nozzle, and, to a lesser extent, particle flux and flow turbulence (Dosanjh and Humphry 1985).
In addition, the greatest damage occurs at the nozzle axis of jet blasting rig because of particle dispersion
and their divergence from the geometrical impact angle. Photos of samples tested at different inclination
angles using this type of blast facility by Oka et al. (2001) are shown in Figure 11.1 (Hamed et al. 2006).
Large variations in the depth and roughness of the tested surface can be seen in the figure.
In centrifugal accelerator facilities, particles are hurled through the radial channels of a rotor against
test specimens at the desired impact angle. It is possible to test several specimens simultaneously in
these facilities. However, they also produce a fairly wide scatter in ejection angles and particles outlet
velocities with up to 16% deviation from the mean value (Kleis and Kulu 2008). Both jet blasting and
centrifugal accelerator erosion test facilities are hence best used for determination of relative erosion
resistance of different materials because they fail to provide test results with a high degree of accuracy.
In whirling arm rigs that operate in vacuum, the elimination of aerodynamic effects enables high par-
ticle velocities (up to 548.64 m/s) with a low-power drive (Goodwin et al. 1969).
Erosion wind tunnels provide uniform particle impact velocities and impingement angles as well
as uniform particle distribution in the test section (Tabakoff et al. 1974). As shown schematically in

30°

60°

90°

5 mm

FIGURE 11.1  Aluminum samples tested in jet blasting facility at 104 m/s by SiO2 (326 μm) particles.
Erosion 11-3

Propane igniter
Fuel
(C)
Combustor

Particle (E) (D) Particle preheater


injector
Steam

F
Acceleration
tunnel
Exhaust
gases
Steam
jacket

Secondary
air

(H) Exhaust
Test tank
G
section
(A)

Cooling
water
Particle
feeder

FIGURE 11.2  Schematic of the high-temperature erosion tunnel.

Figure 11.2 (Hamed et al. 2006), hot erosion wind tunnels also provide uniform high test section tem-
peratures up to 704°C (Tabakoff and Wakeman 1979). Erosion wind tunnels have been used for testing
not only material sample erosion rates but also compressor cascades (Balan and Tabakoff 1984), turbine
vanes, probes (Hamed et al. 2005), and other configurations in environments that are representatives of
the system’s operating conditions (Kline and Simpson 2004).

11.3  Erosion Test Results


A large experimental data base exists on erosion caused by solid particle impact for a wide range of
metallic alloys, plastic, and rubber, as well as for ceramic and composite materials. Erosion test results
for compressor and turbine blades cooling were also reported by Swar et al. (2009) and Tabakoff and
Simpson (2002). Erosion tests are conducted by placing a target sample at predetermined angle and
measuring the change in the sample weight before and after erosion. Erosion test results are usually
expressed in terms of the erosion mass parameter (milligrams of specimen weight loss per gram of
11-4 Theory and Practice of Lubrication and Tribology

impacting particles), or they are converted to an erosion volume parameter (cubic centimeters loss per
gram of impacting particles).
Particle velocity at surface impact has the strongest influence on erosion rate, and it was earlier
proposed that the erosion loss is proportional to the kinetic energy of the erodent particles. However,
velocity exponents both higher and lower than 2 have been reported in the erosion rate results by
various investigators.

11.4  Effect of Particle Impact Velocity


Particle velocity at surface impact has the strongest influence on erosion rate, and it was earlier proposed
that the erosion loss is proportional to the kinetic energy of the erodent particles. Consequently, the
erosion rate (ε) was first believed to be proportional to the square of the impacting particle velocities (v).
However, experimental results indicated that the velocity exponent (n) in the erosion rate equation,
shown later, can vary widely depending on the target material and that it also depends on the impinge-
ment angle and temperature:

e = vn

According to Klies and Kulu (2008), the velocity exponent n also depends on the particle impingement
angles, and values as high as 4 and as low as 0.68 have been reported for different materials.

11.5  Effect of Particle Impact Angle


The variation of erosion rate with impingement angle is characteristically different for ductile and
brittle materials. This is attributed to the predominantly different mechanisms of cutting and brittle
fracture. Brittle-material erosion rate continuously increases with the impingement angle up to its
maximum at normal impingement. However, the erosion rate of ductile materials initially increases
until it reaches a maximum at an impingement angle between 20° and 35°, depending on the material.
It then decreases with further increase in the impingement angle to a residual value at normal impact.
Polymer matrix reinforced composite materials’ erosion behavior is quasi-ductile, with maximum ero-
sion rate occurring at less than 90° but at higher impingement angles than purely ductile materials
(Drensky et al. 2011 and Harsha et al. 2007).
Following the general consideration that the erosion damage is associated with the gradual removal
of material caused by deformation and cutting actions, Oka et al. (1997) proposed a trigonometric func-
tion to represent the dependence of the normalized erosion rate on impact angle and material hard-
ness. The erosion rate was normalized by that at normal impact for five metallic materials, a plastic and
a ceramic material, and the formulation was based on the observed independence of the normalized
erosion rate from the impact velocity. Their semi-theoretical equation consisted of the product of two
terms, the first related to the normal component of the impact energy, and approximates repeated plastic
deformation, while the second approximates the tangential component related to the cutting action.

11.6  Effect of Temperature


Experimental evidence indicates that the erosion rate at 650°C can be double that at room tempera-
ture (Wakeman and Tabakoff 1982). Typical variations of erosion rates with velocity and temperature
as measured are shown in Figure 11.3 (Hamed et al. 2006). It is therefore important that the erosion
experiments at high temperatures be conducted in a carefully controlled environment and that the
target is heated to a steady temperature in the hot stream before introducing the heated particles in the
erosion tests. Both SiC and SiO2 particles produced marked increase in the erosion wear of boiler tube
Erosion 11-5

100 100
INCO 718 B1 = 25° INCO 718 B1 = 90°

16°C, 60°F 16°C, 60°F


40 427°C, 800°F 40 427°C, 800°F
704°C, 1300°F 704°C, 1300°F

20 20

10 10
Erosion rate (cm3/GM) ×104

Erosion rate (cm3/GM) ×104


6.0 6.0

4.0 4.0

2.0 2.0

1.0 1.0

0.6 0.6

0.4 0.4

0.2 0.2

0.1 0.1
1 2 4 6 8 10 1 2 4 6 8 10
Particle velocity (ft/s) × 10–2 Particle velocity (ft/s) × 10–2
34.8 69.6 1.39 2.09 3.48 34.8 69.6 1.39 2.09 3.48
(m/s) (m/s)

FIGURE 11.3  Erosion test results showing effects of temperature and impact velocity.

material above 300°C–400°C, while fly ash particles only produced a slight increase in erosion rate
with temperature. The increased erosion rate is a reflection of the change in target material mechani-
cal properties (Suckling and Allen 1997). The erosion test results of Drensky et al. (2011) for polymer
matrix composites indicated that peak and overall erosion rate increased with temperature; however,
the temperature effect was found to reverse at normal impact conditions.

11.7  Effect of Particle Size and Loading


Experimental evidence indicates negligible erosion rates by smaller particles below 10 μm (Duke 1968).
Erosion test results of ductile materials measured by Grant and Tabakoff (1975) using aluminum oxide
particles and by Kotwal and Tabakoff (1981) using alumina and silica particles of different sizes indicate
that larger particles produced higher erosion rates, but that the effect of particle size on erosion rate
diminished as impact velocity decreased. Experimentally measured erosion rates of ductile materials
by quartz particles sieved into size ranges indicate that increasing particle size causes an increase in the
erosion rate, but that there is a critical particle velocity above which erosion is not influenced by size,
and this value appear to increase linearly with impact velocity (Goodwin et al. 1969). Very high particle
11-6 Theory and Practice of Lubrication and Tribology

50, 3X 25KV WD : 39MM S : 00000


(c) 1MM
(a) (b)

FIGURE 11.4  Electron micrographs of (a) ∼100 μm fly ash, (b) ∼200 μm silica sand, and (c) ∼250 μm aluminum
oxide particles.

fluxes produce a shielding effect by the rebounding particles that interfere with the incoming particles.
Hence, the erosion rate was found to increase with a decrease in particle loading, up to a point where
further decrease in particle flux had little effect on the erosion rate (Suckling and Allen 1997).

11.8 Effect of Particle Hardness, Sharpness,


and Mineral Composition
Erosion rate is also affected by particle composition and shape. Figure 11.4 shows magnified scanning
electron micrographs of fly ash, silica sand, and aluminum oxide particles (Hamed et al. 2006). The lat-
ter are most erosive because of their angular shapes and very sharp corners. Quartz is the most common
dust constituent, but the composition and size distribution of erosive dust can vary with location. Since
quartz is one of the hardest common constituents, an increase in quartz content results in a significant
increase in erosion rate. Furthermore, general abrasion between particles on the ground results in pref-
erential reduction in the size of the softer constituents. Goodwin et al. (1969) observed that the hardest
dust particles tended to also have the sharpest profiles.

11.9  Volcanic Ash


Several airplanes have been nearly lost and many damaged by volcanic ash encounters (Miller and
Casadevall 1999). The level of damage to aircraft surfaces and windshields is influenced by the length of
time spent in the volcanic ash clouds, which can be transported as far as 1000 miles from eruption loca-
tions. In addition, volcanic ash damages the electronics, power generators, navigation instruments, and
infiltrates the engine lubrication system (Grindle and Burcham 2003). However, the loss of jet engine
power is the most serious occurrence in these encounters. Since jet engine combustor exit temperatures
are well above the 1000°C ash melting temperatures, the critical engine damage occurs in the hot section
where the molten ash solidifies inside the turbine cooling passages and coats the combustor fuel nozzles
and turbine airfoils as confirmed experimentally by Dunn et al. (1996), and through detailed examina-
tion of the engines of NASA DC-8 airplane that flew inadvertently through the volcanic ash cloud of
Mt. Hekla in 2000 (Grindle and Burcham 2003).

11.10  Roughness of Eroded Surfaces


Surface roughness characteristics were measured after erosion tests in some investigations and were
found to correlate closely with the erosion rate in terms of variation with the impact angle, velocity, and
particle size (Balan and Tabakoff 1984; Kline and Simpson 2004; Hamed et al. 2005). The eroded surface
Erosion 11-7

FIGURE 11.5  Vane surface damage of metal alloy due to erosion by ∼1500 μm crushed quartz at 300 ft/s com-
pared to the protected original surface.

roughness did not change beyond a certain limit even with additional mass removal by erosion (Hamed
et al. 2005). Figure 11.5 clearly shows the difference between the roughness of the exposed and protected
vane surfaces after being tested in the erosion tunnel (Hamed et al. 2006).

References
Balan, C. and Tabakoff, W. (1984) A method for predicting the performance deterioration of a compressor
cascade due to sand erosion, AIAA Paper 84-1208.
Dosanjh, S. and Humphry, J. A. (1985) The influence of turbulence on erosion by a particle-laden fluid jet,
Wear, 102, 309–329.
Drensky, G., Hamed, A., Tabakoff, W., and Abot, J. (2011) Experimental investigation of polymer matrix
reinforced composite erosion characteristics, Wear, 270(3–4), 146–151.
Duke, G. A. (1968) Erosion tests on modified Rover Gas Turbine, Australian Defense Scientific Service,
Research Laboratories, Note No. ARL/ME279.
Dunn, M. G., Baran, A. J., and Miatech, J. (1996) Operation of gas turbine engine in volcanic ash clouds,
Journal of Engineering for Gas Turbine and Power, 118, 724–731.
Finnie, I. (1960) Erosion of surfaces by solid particles, Wear, 3, 87–103.
Goodwin, J. E., Sage, W., and Tilly, G. P. (1969) Study of erosion by solid particles, Proceedings of
Instrumentation Mechanical Engineers, 184, 279–289.
Grant, G. and Tabakoff, W. (1975) Erosion prediction in turbomachinery resulting from environmental
particles, Journal of Aircraft, 12(5), 471–478.
Grindle, T. J. and Burcham, F. W., Jr. (2003) Engine damage to a NASA DC-8-72 airplane from a
high-altitude encounter with diffuse volcanic ash cloud, NASA/TM-2003-212030.
Hamed, A., Tabakoff, W., Rivir, R. B., Das, K., and Arora, P. (2005) Turbine blade surface deterioration by
erosion, Journal of Turbomachinery, 127(3), 445–452.
Hamed, A., Tabakoff, W., and Wenglarz, R. (2006) Erosion and deposition in turbomachinery, Journal of
Propulsion and Power, 22(2), 250–260.
Harsha, A. P. and Thakre, A. A. (2007) Investigation on solid particle erosion behavior of polytherimide
and its composites, Wear, 206, 807–818.
Hutchings, W. H. (1983) A model for the erosion of metals by spherical particles at normal incidence,
Wear, 34, 269–281.
11-8 Theory and Practice of Lubrication and Tribology

Kleis, I. and Kulu, P. (2008) Solid Particle Erosion Occurrences, Prediction and Control, Springer-Verlag,
London, U.K.
Kline, M. and Simpson, G. (2004) The development of innovative methods for erosion testing a Russian
coating on GE T64 gas turbine engine compressor blades, ASME Paper GT2004-54336.
Kotwal, R. and Tabakoff, W. (1981) A new approach for erosion prediction due to fly ash, Journal of
Engineering for Power, 103, 265–227.
Miller, T. P. and Casadevall, T. J. (1999) Volcanic Ash Hazards to Aviation, Encyclopedia of Volcanoes,
Academic Press, San Diego, CA, pp. 915–930.
Oka, Y. I., Nishimura, M., Nagahashi, K., and Matsumura, M. (2001) Control and evaluation of particle
impact conditions in a sand erosion test facility, Wear, 250, 736–774.
Oka, Y. I., Ohnogi, H., Kosowaka, T., and Matsumura, M. (1997) The impact angle dependence of erosion
damage caused by solid particle impact, Wear, 204, 573–579.
Sheldon, G. L. (1970) Similarities and differences in the erosion behavior of materials, Journal of Basic
Engineering, 89, 619–625.
Sucking, M. and Allen, C. (1997) Critical variables in high temperature erosive wear, Wear, 203–204,
528–536.
Swar, R., Hamed, A., and Tabakoff, W. (2009) Deterioration of Thermal Barrier Coated Turbine Blades,
ISABE 2009-1248, Montreal, Quebec, Canada.
Tabakoff, W., Grant, G., and Ball, R. (1974) An experimental investigation of certain aerodynamic effects
on erosion, AIAA Paper 74-639.
Tabakoff, W. and Simpson, G. (2002) Experimental study of deterioration and retention on coated and
uncoated compressor and turbine blades, 40th AIAA Aerospace Sciences Meeting & Exhibit, January
14–17, Reno, NV.
Tabakoff, W. and Wakeman, T. (1979) Test facility for material erosion at high temperature, ASTM Special
Publication 664, pp. 123–135.
Tilly, G. P. (1973) A two stage mechanism of ductile erosion, Wear, 23, 87–93.
Wakeman, T. and Tabakoff, W. (1982) Measured particle rebound characteristics useful for erosion
prediction, ASME Paper 82-GT-170.
12
Oxidative Wear
12.1 Introduction and Background....................................................... 12-1
12.2 Oxidation in the Absence of Wear................................................ 12-2
Oxide Growth and Oxide Growth Models  •  Damage to the Oxide
Scale Resulting from Mechanical Stress  •  Oxidative Wear under
Sliding Wear Conditions
12.3 Transition to Mild Oxidative Wear..............................................12-8
12.4 Development of Wear Protective Oxide Layers........................ 12-10
12.5 Oxidation Effects in Fretting Wear............................................. 12-12
12.6 Modeling Erosion–Oxidation Processes.................................... 12-13
12.7 Basic Erosion Mechanisms........................................................... 12-13
John R. Nicholls 12.8 Erosion–Oxidation........................................................................ 12-14
Cranfield University
12.9 Modeling and Mapping Erosion–Oxidation............................. 12-17
Richard G. Wellman 12.10 Summary......................................................................................... 12-19
Cranfield University References................................................................................................... 12-19

12.1  Introduction and Background


Many wear processes can be significantly influenced by heat, especially in sliding wear, fretting, and
erosion where the higher temperatures can facilitate the oxidation of the contacting metal surfaces
[1–18]. Such heating processes are often externally applied/generated, associated with plant and equip-
ment operation, but may also be generated at the contacting surface through frictional forces, impact,
and material deformation. This is particularly true for dry sliding wear conditions.
Thus, the aim of this chapter is to provide an overview and review of the synergistic interactions of
oxidation/corrosion processes and wear.
There are many examples where environmental effects exacerbate wear [1,3,4,7,8,12–14,16,17] and
others where the formation of an oxide glaze provides protection and minimizes wear damage [2,5,6,10–
12,14,19,20]. It is well known that oxides can reduce wear rates by reducing or eliminating metal–metal
contact [1,6,10,11,19,20]. However, under fretting conditions, reciprocating micromovement of contact-
ing surfaces, the severity of damage during fretting may increase due to oxide debris accumulating at
the contacting interface [7,21]. If these oxide particles are harder than the adjacent metal, then abrasive
action begins exacerbating the wear process. This influence of the environment on wear is well recog-
nized in industry and with correct material design may be used to minimize wear. In diesel engines,
oxidation-enhanced wear is used to control diesel valve seat recession through the formation of protec-
tive oxide “glazes” [20,22,23]. These “glazes” are compact nanocrystalline oxides (originally thought to
be amorphous—hence the term “glaze”), which can provide high-temperature wear protection to slid-
ing contact surfaces [14]. With the “glaze” layer present, severe metal–metal interactions (local welding)
cannot occur and therefore wear is greatly reduced. From these examples, it is evident that oxidative
wear damage may be catastrophic or beneficial and this depends on whether a protective oxide layer can
form that acts as an environmental barrier.

12-1
12-2 Theory and Practice of Lubrication and Tribology

From the foregoing, it is apparent that “oxidative wear” is a subset of Tribo-corrosion. The lat-
ter defined as “the complex degradation process affecting metals and alloys, which results from the
combined affect of mechanical loading and the environment, when two materials are in mechanical
contact.” Tribo-corrosion process may include sliding wear, abrasive wear, erosion by solid particle or
liquid impact, cavitation or fretting fatigue, while the environmental interaction may be classified as
“dry”—for example, high-temperature oxidation—or “wet,” which involves condensed, liquid phases.
Aqueous corrosion at room temperature falls into this latter class [16]. Tribo-corrosion is observed in
a diverse range of industries from transport (automotive, trucks, aeronautics, and ships), through the
offshore industry and energy production to dentistry and health care (implants). Tribo-corrosion has
been well reviewed recently [16,24–26]. In this chapter, only the influence of “dry” corrosion conditions,
associated with oxidation at increasing metal temperatures, will be reviewed.

12.2  Oxidation in the Absence of Wear


Thermodynamically, all metals and alloys, with the possible exception of gold, are unstable in air, or
oxygen-containing environments, and react with oxygen to form their native oxide. In the absence of
wear or sliding mechanical contact, they rely on the formation of a slow-growing, protective oxide scale
to limit metal loss through oxidation. Thus, many of our accepted construction materials are designed
to be self-protecting, forming either stable chromia (Cr2O3) or alumina (Al2O3) surface oxides. For
example, many stainless steels have sufficient chromium content for Cr2O3 protective scales, which
can resist oxidation processes well, up to 900°C [27]. However, for specialized steels designed to work
at temperatures up to 1300°C, both chromium and aluminum are added, along with minor amounts
of reactive elements (Y, Zr, Hf, Ce, etc.), and these specialist steels develop a protective alumina scale
[28–30]. Moreover, once a protective scale is formed, the rate of oxidation is generally controlled by the
diffusion of reactants through the oxide scale. Such diffusion-controlled oxidation processes were first
modeled by Wagner in 1933 [31,32] and since have been well reviewed in the many text books available
on high-temperature oxidation and corrosion [33–36].
Wagner’s model and the current understanding of oxidation would suggest that for diffusion-­
controlled growth the oxide scale will thicken parabolically with time and that the oxidation rate will
increase exponentially with temperature, according to an Arrhenius dependence:

x 2 = k pt (12.1)

⎛ −Q ⎞
k p = A exp ⎜ (12.2)
⎝ RT ⎟⎠

where
x is the oxide thickness
t is the time
kp is the oxidation rate constant
A, Q, and R are constants with Q known as the activation energy
T is the absolute temperature

Such ideal behavior is often not found in practice. At low oxidation temperatures, the scaling rate may
no longer be parabolic and may slow to a limiting oxide thickness condition as has been reported by
Cabrera and Mott [35,36]. Similarly, at very high temperatures, structural changes, for example, grain
growth, may limit available diffusion paths and again give rise to oxidation rates slower than predicted
by Wagner, known as sub-parabolic oxidation [37,38]. Clearly though, whatever the exact oxidation
mechanism, environmental protection and minimal metal loss are associated with the formation of an
Oxidative Wear 12-3

adherent, slow-growing, oxide layer that acts as a diffusion barrier limiting the ingress of oxidant or cor-
rosive species. If this was under aqueous corrosion conditions, this state of stable scale formation would
be known as “passivation.”
Wear processes, impact, abrasion, and sliding contact, damage this “passivation” and thus may
enhance the oxidation processes. Reoxidation and new scale growth will continually try to repair this
damage; thus, there is a synergy between the extent and severity of the wear events and the potential
repair of protection possible by new scale formation. To better understand this synergy, it is necessary to
understand how oxide scales grow, how they can repair themselves, how oxide scale mechanical proper-
ties change with scale thickness, and how these properties change and developed oxidation morpholo-
gies influence scale fracture and failure.

12.2.1  Oxide Growth and Oxide Growth Models


From an understanding of the interaction of oxidation and wear processes, there are in practice three
types of oxide growth that may occur during the oxidation of metals and alloys in the absence of
wear. Each will later be shown to play a role in determining the friction and wear behavior; thus, all
three are important and often they may overlap and interact together. As noted by Stott and Wood
[5], the type of oxide growth determines the amount of oxide present on the alloy surface prior to
sliding (Stott and Wood were researching sliding wear) as well as that produced between the surfaces
during sliding contact. It would also determine the chemistry, composition, and rate of formation
of any protective oxide “glazes” that may form the tribo-oxidation process as part of oxidative wear
[5,10,11].
Thus, some important features of the oxidation process and growth of oxide scales are described
in brief before considering the implications of the dynamics of such oxide formation on various wear
processes. It will be shown further that the mechanical properties of the oxide scales are a critical factor
determining whether the oxide cracks, spalls, and thus mechanically fails or is plastically deformed and
compacted to form an oxide glaze.
Stott and Wood [5] in their study of “glaze” formation considered the influence of oxides on friction
and wear from a corrosion engineer’s perspective. Prior to this, many researchers had researched oxi-
dative wear from a tribologist’s perspective [1–3,6,8] and other recent papers build on the conceptual
models derived by Quinn and coworkers [2,4,6,8]. It is interesting to contrast and compare oxidative
wear from these two perspectives. As identified by Stott and coworkers [5,19], the following aspects of
oxidation–corrosion are important when considering oxidative wear.

12.2.1.1  Early Oxide Nucleation and Growth


How the oxide forms and grows is particularly important for “fresh,” chemically clean metals in low
oxygen-containing gaseous environments when exposed at relatively low temperatures. The oxidizing
gas must first be absorbed on to the clean metal surface, whereupon small nuclei of oxides of most
alloying elements in the alloy are formed at preferred sites. These nuclei grow, slowly, extend laterally,
and with time coalesce to give a complete oxide layer. Figure 12.1 illustrates the thickness of oxide films
formed on a clean surface of various metals at room temperature (20°C–25°C). This figure was redrawn
using data that was originally published in Kubaschewski and Hopkins in 1953 [39], from early research
papers on the oxidation of Al, Cu, and Fe.
It can be seen that very rapidly, within the first minutes of exposure of a fresh clean surface, an oxide
of some 1.5–2 nm is formed. This thickens over the next 2–10 h, depending on the metal, and after that
grows very slowly. For aluminum, an amorphous Al 2O3 scale first forms, which thickens to 2.5 nm
after 50 h. For iron and copper, crystalline oxides form, with thicknesses of 3.4 and 4.4 nm after 50 h
exposure. Why some oxides first form as amorphous layers and others as crystalline material depends
on local surface energies, the local crystal structure of the metal and oxide, and whether any epitaxy
may exist.
12-4 Theory and Practice of Lubrication and Tribology

Oxide thickness (nm)


3

1
Cu
Fe
AI
0
0 10 20 30 40 50 60
Time (h)

FIGURE 12.1  Thickness of oxide films formed on various metals at room temperature.

This region of film formation, where many oxides form and grow to a limiting thickness, may
extend to higher temperatures with the maximum temperature at which this early nucleation and
growth process controls film thickness varying from metal to metal. With a further increase in tem-
perature, there is another region where neither asymptotic oxidation nor high temperature parabolic
oxidation (see Section 12.2.1.3) applies. In this intermediate region, a number of oxidation rate law
models have been applied, including a logarithmic law, a cubic law, and sometimes a parabolic law
but with a rate constant different to that of steady state. An explanation for this sub-parabolic or near
asymptotic oxidation behavior was first proposed by Cabrera and Mott [36] based on electric field
gradients that modify the diffusion transport of ions across the oxide scale. Thus when ionic defect
concentrations within the oxide do not change with thickness one has a parabolic law, when these
ionic defect concentrations in this film decrease as 1/thickness one has a cubic law and for other
distributions the logarithmic law approximates. This theory of Cabrera and Mott [36] therefore cov-
ers a number of mathematical relationships for oxide growth on various alloys up to several tens of
nanometers.

12.2.1.2  Transient Oxidation


At higher temperatures for a freshly exposed metal surface in an oxygen-rich environment, small nuclei
of single component oxides impinge and grow thus more complex oxides quickly develop. These first
nuclei rapidly coalesce to form a complete oxide layer. An important observation is that all elements in
the alloy oxidize during this transient stage and the amount of these various oxides are roughly in pro-
portion to the concentrations of the elements in the alloy. This transient oxidation stage is observed at
higher temperatures than the low temperature, slow nucleation, and growth stage, where this complexed
mixed oxide scale rapidly forms and grows. This transient stage of oxide formation is probably most
important in determining wear behavior under oxidative wear. For the corrosion engineer, it is critical
[40] determining the spatial distribution, composition, and structure of the first formed oxide. In turn,
this influences the development of the final oxide scale structure and its thickness at which steady-state
oxidation is achieved.

12.2.1.3  Steady-State Oxidation


Following transient oxidation, a steady state is achieved eventually, whereupon thermodynamically
more favored oxides grow in preference to those that are less thermodynamically favored, to produce
Oxidative Wear 12-5

a complete slower growing layer at the base of the transient oxide layer. For example, as cited by Stott
and Wood [5], a Ni–20%Cr alloy will initially oxidize to form NiO and Cr2O3 nuclei. The Cr2O3 nuclei
will develop below the transient formed NiO to form a complete Cr2O3 layer. In turn, this layer forms an
environmental barrier and limits the rate of growth of the oxide on this alloy. Similarly, for an FeCrAl +
RE (RE = reactive element), the most stable slow-growing oxide is alumina; thus, this alloy would form
a steady-state protective alumina layer on exposure to an oxidizing environment [29,30,37,38]. These
protective oxide layers are not always formed, especially when the oxide scales are repeatedly dam-
aged and are required to regrow a new surface oxide. This would progressively deplete the chromium
(or aluminum, depending on which oxide is formed) present in the alloy. When reduced, such that the
scale-forming element concentration is below some threshold level, then it is no longer possible to form
a healing layer be it chromia or alumina, instead internal oxidation of these elements may occur within
the alloy and an external base metal oxide scale forms. Thus the oxide that forms under steady-state
conditions depends on many factors including the alloy composition, diffusion coefficients within the
alloy and the oxide (both of which depend on temperature), oxygen solubility, oxide growth rates, the
mechanical properties of the oxide scale, and the local oxidation conditions.
Growth of these protective oxide scales is controlled by the diffusion of ionic species across the grow-
ing oxide, either metal cations, for example, the formation of chromia scales on the nickel–chrome alloy
reported earlier, or oxygen anions in the case of the growth of an alumina scale on a FeCrAl + RE alloy.
In these cases, provided equilibrium has been established at the gas–oxide and oxide–metal phase
boundaries, the diffusion of either the cations of the oxide-forming elements or oxygen anions will con-
trol film growth. This behavior has been modeled by Wagner [31,32], see also [33–36]; the Wagner oxide
growth model predicts a parabolic oxidation rate. In simplistic terms, the oxide growth rate is found to
be inversely proportional to the oxide thickness, with the proportionality constant (kp)—the parabolic
rate constant. Thus, in differential form, Equation 12.1 becomes

dx k p
= (12.3)
dt 2x

demonstrating that the oxide growth rate is inversely proportional to thickness.


This on integrating gives

x 2 = k pt + c (12.4)

Only when oxidation is assumed to commence on a clean metal surface, on which no initial oxide
film has formed, the constant of integration is zero, giving the classical parabolic rate equation (see
Equation 12.1). If an initial oxide film is present on the surface, then a constant of integration must be
included to account for this initial oxide film thickness.

12.2.2  Damage to the Oxide Scale Resulting from Mechanical Stress


Various forms of damage are possible, from direct mechanical loading, from thermal loads, or as a
result of any prevailing wear conditions, which may alter the stress situations established in the oxide
scale–metal composite system. Thus, if the wear damage results in plastic damage to the substrate-
forming ridges and hillocks, then tensile fracture of the oxide scale may occur, cracking through the
oxide thickness, over such ridges and hillocks leading to loss of oxide from the asperities. Equally,
fracture at the scale–metal interface may occur by shear, introduced as a result of friction forces.
Whereas, buckling stress can lead to separations at the scale–metal interface, which on relaxation
from compressive loads can lead to disbonding at the oxide–scale interface and the loss of oxide due
to spallation. The cause of such fracture processes and the associated oxide mechanical properties
12-6 Theory and Practice of Lubrication and Tribology

have been researched by Hancock and coworkers [41–44] and Schütze and coworkers [45–50] and
have been well reviewed in a recent book published by Schütze [34]. It can be assumed that there
are always defects in the oxide scales (pores, microcracks, and different oxide phases). Fracture will
occur when the energy released by the growth of existing defects in the scale exceeds the energy
to create two new surfaces. Thus, the critical stress (σc) calculated on the basis of this criterion is
given by [34]

12
⎛ E ∗Gc ⎞
sc = ⎜ (12.5)
⎝ pc ⎟⎠

where
E* is the effective elastic modulus, which can be calculated for the case of a phase boundary (in this
example, the boundary between the metal and the oxide scale) from Poisson’s ratio and the shear
moduli of the adjacent phases [51]
Gc is the critical surface energy (energy for enlarging the crack)—values can be taken from [52]
c denotes the geometric dimensions of the composite defect size [41]

Experimentally, it has been shown that σc varies with oxide thickness and temperature, decreasing as
the oxide thickness increases, but increasing as the surface temperature is increased [41]. Figure 12.2
illustrates such behavior for a chromia (Cr2O3) scale formed on Nimonic 75 (Ni–20%Cr), where a plot
of fracture stress against scale thickness is plotted from experimentally determined data. Given that

320

280

1000°C
240
Fracture stress (MN/m2)

200

160 950°C

120 900°C

80 800°C

40 700°C

0
0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0
Thickness (µm)

Cracked Uncracked Temp, °C KIC’ MN m–3/2


700 2.0
800 2.7
900 3.7
950 4.5
1000 6.2

FIGURE 12.2  Variation of the fracture stress for a chromia scale as a function of oxide thickness and ­temperature.
(From Stephenson, D.J. and Nicholls, J.R., Mater. Sci. Technol., 6, 96, 1990. With permission.)
Oxidative Wear 12-7

Kc = (E*Gc)1/2 and from measurements of the composite defect size (see [34] or [41]) within the scale
at different temperatures, predicted fracture stresses at temperatures between 700°C and 1000°C are
superimposed [53].
Table 12.1 presents surface fracture energies, elastic moduli, and calculated fracture toughnesses for
a range of oxide scales that may be encountered in oxidative wear behavior. This room temperature data
for Fe3O4 and Cr2O3 is superimposed with high temperature fracture data of Hancock and Nicholls [41]
in Figure 12.3, illustrating the significant increase in fracture toughness of these two oxide systems at
surface temperatures (contact temperatures in the case of wear) above 600°C.

TABLE 12.1  Surface Fracture Energies, Elastic


Moduli, and Calculated Mode I Fracture
Toughnesses for Some Oxide Scales
Oxide γo [J/m2] Eox [GPa] KIC [MPa (m)1/2]
CoO 3.0 156 1.0
NiO 3.6 191 1.2
Al2O3 7.7 419 2.5
Cr2O3 5.8 283 1.8
Fe2O3 6.0 219 1.6
Fe3O4 4.5 208 1.4
FeCr2O4 5.0 233 1.5
SiO2 4.4 86 0.9
Source: Data taken from Robinson, J. and Manning,
M.I., Mater. Sci. Technol., 6, 81, 1990.

16

14

FeO/Fe3O4 on
12 iron

10
KIC MNm–3/2

Cr2O3 on
2 nimonic 75

0
0 100 200 300 400 500 600 700 800 900 1000 1100
T (°C)

FIGURE 12.3  Influence of temperature on the critical stress intensity factor (KIC) for iron oxide and chromium
oxide scales. (Data from Table 12.1 and Hancock, P. and Nicholls, J.R., Mater. Sci. Technol., 4, 398, 1998.)
12-8 Theory and Practice of Lubrication and Tribology

Hence the observation of a critical oxide thickness in the oxidative wear model of Quinn and
Sullivan [54] reflects a critical strain energy sufficient to propagate an oxide crack–interface delamina-
tion for an oxide scale and its interface with the metal that becomes more defective as the oxide grows
thicker.

12.2.3  Oxidative Wear under Sliding Wear Conditions


Oxides have long been known to influence wear rates of metallic surfaces, recognized as early as 1930
[55] when addressing mild wear of steels. The manner by which oxides influence the wear rates has
been well reviewed by Stott and coworkers [5,10,11,14,56]. It is well known that oxides can reduce metal
loss rates by reducing metal–metal contact [1], giving a transition from severe to mild wear. Stott and
coworkers have noted that oxides formed in wear may be present as fine wear debris particles, as a
smooth burnish top layer or as a “glaze.” Such a “glaze” can form on the surface of compacted particles,
leading to reduced friction and very low wear rates [10,14]. It is clear that there is an influence of the
oxidation rate, thus for instance for steels one sees a transition from severe wear to mild wear as the
temperature of the contacting surface increase [2–6,8,12,16] and so does the rate of oxide formation.
Factors that influence the oxide growth rate will influence oxidative wear, including temperature at the
wear (rubbing) surface, a function of contact pressure and sliding speed, the oxygen partial pressure,
and other species in the environment that may exacerbate corrosion rates. Thus, as reported by Stott
[56], during the early stages of sliding, frictional forces lead to local heating, mechanical deformation,
and oxidation, particularly at high ambient temperatures. The wear debris that may form consists of a
mixture of oxide particles and partially oxidized metallic particles. They further showed that three lim-
iting cases of oxide generation occur during slide wear, that is attrition and oxidation of metallic debris
(1 μm diameter particles are oxidized in a fraction of a second at 20°C [56]); oxidation of metal asperities
on the sliding surface while in contact; and in some situations, especially at high ambient temperatures
the oxide may develop on surfaces prior to the sliding action. As the sliding surfaces traverse each other,
the oxides may be removed completely as proposed in the models of Quinn and coworker [2,54,58,59]
for mild oxidative wear, may be partially removed, or may be thickened and compacted into a protective
“glaze.” Thus, oxide properties such as oxide strength, tenacity (degree of adhesion), elastic properties,
fracture toughness, and the strain to crack the oxide must also play a role, but this is an area where there
is insufficient data for most of the systems that have been studied under oxidative wear conditions. Thus,
from a tribological perspective, the formation of “oxide glazes” can be thought to prevent wear; however,
this requires that the oxide film remains protective and does not fail mechanically during the wear con-
tact. If the oxide does fail, then the oxide particles that may form can become abrasive exacerbating the
wear, as was recognized by Archard and Hirst [1] in 1956, when studying the wear of hardened steels.

12.3  Transition to Mild Oxidative Wear


Consider, as proposed by Quinn [2] in 1962 and Welsh [3] in 1965, that the transition between mild
and severe wear is controlled by a balance of competing tribological and oxidation processes. Thus,
the rate of exposure of new surface (a result of oxide removal during the wear process) will depend on
tribological controlled factors such as applied load, rubbing speed, friction force, and temperature of
the contacting surfaces. Figure 12.4 illustrates a wear mechanism map for steels illustrating the contact
conditions for “mild oxidative wear” [57]. Whereas, the rate of coverage with new oxide will depend on
the alloy composition, environmental factors such as oxygen partial pressure, again the local surface
temperature, and whether a sufficiently resilient oxide has time to form (that is, will the oxide crack on
a subsequent loading–wear cycle). Parameters controlling the rate of oxidation and the conditions that
may lead to oxide failure have been described in Sections 12.2.1 and 12.2.2.
From this type of argument, Quinn and Sullivan [54] in 1977, following a review of oxidative wear,
proposed a model to predict wear rates under mild oxidative wear conditions, as follows:
Oxidative Wear 12-9

Sliding velocity v (m/s)


–4 –2 2
10 10 1 10
10
Steel
Wear–mechanism map
Pin–on–disk configuration
Seizure

–3
10 –4 –3 –4
10
Melt (10 )
–2
–1 10 (10 )
Normalized pressure F

–5
10
10 (10 ) –6
–3
–4 (10 ) Wear 10
–5
10 –5 –7 10
–6
–6
10
–4 10 (10 ) 10
–7
Mild– 10
oxidational
wear –6 –8
Martensite –8
10 (10 ) formation 10
–5
10
–3
Severe-
10 Delamination oxidational
wear –7 wear
10 10–9
–6 –8
10 (10 ) Mild to severe
wear transition
–8
10 –10
10
Ultra- –9
mild 10 –11
10
–9 10
–5 wear
10
–2 1 2 4
10 10 10
Normalized velocity v

FIGURE 12.4  Wear mechanism map for steels. (From Lim, S.C. and Ashby, M.F., Acta Metall., 35, 1, 1987. With
permission.)

⎛ Q ⎞
exp ⎜ −
⎛a⎞ ⎝ RTC ⎟⎠
g = A⎜ ⎟ (12.6)
⎝U ⎠ xc2

where
γ is the wear rate
A is the real area of contact over which the wear process occurs
TC is the temperature in the contact zone
(a /U) is an estimate of the time available for oxidation, derived from α = the mean radius of asperity
contact and U = the sliding speed

The oxidation parameters are as follows: Q is the activation energy for oxide formation and xc is a critical
thickness of oxide at which oxide mechanical failure and spallation may occur.
This oxidation model of Quinn and coworker [2,4,8,54], in its early form, involved the oxidation of
metal asperities while in contact. The extent of oxidation will depend primarily upon the temperature
developed at these asperity contacts. In each of these asperity contact areas, it is assumed that the local
oxide thickness increases to some critical value, whereupon the oxide is lost from the point of contact
and the process repeats. The critical oxide thickness is such that after a number of asperity contacts,
oxidation, and removal, the wear rate observed matches that predicted by the Archard wear equation.
Some encouraging correlations between the model and experimental data have been achieved
[58,59]. From a study of the mild oxidative wear of a high chromium ferritic steel pin against a
stainless steel plate at 400°C, it is shown that the activation for tribo-corrosion may be half that for
static oxidation, with this activation energy varying with applied load. It is obvious that this model
12-10 Theory and Practice of Lubrication and Tribology

is solely based on the rate of buildup of oxide than its loss. No account is taken of how the oxide is
lost or whether the rate of loss varies over the wear surface, which could depend on the number of
asperities in contact. The degree of asperity contact may account for the observed activation energy
dependence on load. Clearly the model has deficiencies, as it takes no account of the oxide mechani-
cal properties, except through the proposed critical oxide thickness (xc). It does, however, recog-
nize that oxide growth follows parabolic kinetics (as per Equation 12.3) and that the oxide growth
dependence on temperature follows an Arrhenius rate law (as per Equation 12.1). However, why
the tribo-corrosion activation energy should be half that for static oxidation is an area for further
research. Stott has suggested a much lower value [56]. Does this reflect a size effect, that is, fine debris
oxidize more easily or oxidize due to the local stored strain energy, or does it just reflect the rate of
transient oxidation?
One must further ask, “What is meant by ‘the critical oxide thickness’ as proposed in Quinn and
Sullivan’s model?” It can be considered as a fit factor, in which case Equation 12.6 would be used as an
empirical model to fit experimental data. On the other hand, the “critical oxide thickness” may be an
important material parameter that defines the local mechanical properties of the metal–oxide system,
when fracture of the oxide occurs under the load conditions imposed by the wear process.

12.4  Development of Wear Protective Oxide Layers


Stott and coworkers [11,14,19,56,60,62], in their studies of sliding wear, have shown that above 150°C the
wear of the pin, in a pin on disk study, is significantly reduced compared to lower temperatures. Further,
at higher temperatures, there is a sudden, significant, drop in the coefficient of friction after some given
sliding time. They observed that the time to this drop in measured friction was a reproducible func-
tion of the sliding test temperature, varying from several hours at 250°C to a few seconds at 800°C, as
observed, for like-on-like wear of Nimonic 75 and Nimonic 80, both are, chromia-forming nickel-based
alloys [5,10,56,60]. Figure 12.5 illustrates this observed change in the coefficient of friction at tempera-
tures up to 800°C for the like-on-like sliding of Nimonic 75 and Nimonic 80 under reciprocating sliding
wear conditions. This figure has been drawn from data in [56,60] for the sliding wear of Nimonic 75,

1.6
Coefficient of friction

1.3

250°C
0.8
400°C

0.4 800°C

0
1 10 100 1,000 10,000 100,000
Time, s

FIGURE 12.5  Change in friction coefficient due to the formation of a compact oxide “glaze” during reciprocating
sliding wear of Nimonic alloys. (Drawn from data for Nimonic 75, published in Stott, F.H., Tribology Int., 35, 489,
2002; Lin, D.S. et al., ASLE Trans., 17, 251, 1973; data at 250°C is for Nimonic 80 from Stott, F.H., Tribology Int., 35,
489, 2002.)
Oxidative Wear 12-11

with a load of 15 N and a displacement of 2–5 mm at 8.3 Hz in air [60] at test temperatures of 400°C and
800°C and Nimonic 80 at a test temperature of 250°C.
Stott and coworkers further showed from detailed studies of the wear scars that this change in friction
was associated with the development of compacted layers of oxide and partially oxidized alloy particles
on the sliding surfaces. This compacted layer results from the accumulation of oxidized wear debris.
During the sliding wear process, some debris is lost, but most is retained and undergoes further defor-
mation, fragmentation and attrition, progressively being broken down into smaller and smaller particles.
As sliding continues these fine particles become compacted due to thermoelastic stress formed between
the contacting surfaces, ultimately becoming load-bearing surfaces as the surrounding regions of metal-
lic asperity contact are worn down. Under these thermoelastic sliding conditions, the fine oxide particles
sinter together forming more dense layers, an oxide “glaze.” As noted by Stott [56], sintering of nanosized
particles can occur to a significant degree even at temperatures just above room temperature [61].
Once the compacted oxide layer—the “glaze”—has been established, two competing processes occur
during further sliding: the fracture and breakdown of these surface oxide layers resulting in the forma-
tion of further debris, and consolidation and compaction of this debris to repair these areas of damage.
Thus, at low surface temperatures the rate of surface damage may exceed the possibility for repair, with
loose oxide particles being lost resulting in the breakdown of any protective “glaze.” While at higher
temperatures the more rapid oxidation of any debris formed, plus the oxidation of freshly exposed metal
surfaces, ensures sufficient micro-fine oxide particles are available for compaction to repair any dam-
age. This competition would account for the observed sharp transition from high friction—involving
oxide breakdown and metal–metal contact—to low friction when a compacted “glaze” has the capability
to form.
From this understanding, Stott and coworkers [62] proposed a model to account for the development
of compacted oxide, wear debris layers based on the criteria that must be met for the loss or retention
of wear debris particles, during reciprocating sliding wear. A key feature in this model is the need for
newly formed, loose debris to be able to move, rotate, and thus attrite before becoming entrapped in
the surface. Thus, the model calculates a wear volume, from a knowledge of the applied load, the size of
asperity contact, measured values of friction forces for the region of asperity contact, and predicts com-
pact debris formation and the proportion of debris that has converted to a glaze, where the formation of
oxide debris and its conversion to compacted oxide vary with time and temperature. The model would
appear to use Monte Carlo methods to estimate when oxide debris will form and the conditions under
which this debris compacts and converts to a “glaze”. The starting condition for this computer-based
model is the assumption that one of the contacting surfaces is rough with hemispherical asperities, with
heights that follow a Gaussian distribution; the counter surface is assumed smooth [5,56,60,62]. The
model is shown to predict experimental trends in a reciprocating sliding wear test [5,60,62], predicting
wear volumes both for the severe wear and oxide “glaze” (high temperature) conditions. The model also
demonstrates that increasing temperatures facilitates “glaze” formation and the formation of wear pro-
tective layers. Experimental values and model predictions for the activation energy for the wear process
of like-on-like reciprocating sliding were similar, circa 12.6 kJ/mol. This value is much lower than that
expected for the static oxidation of Nimonic 75 to form a chromia scale (268 kJ/mol). As noted by Stott
[56], even though tribo-oxidation processes differ from static oxidation, one should expect the activa-
tion energy barrier within the oxide layers that controls the oxidation process to be similar. Such an
observation concurs with that of Quinn and coworkers [63,64]; thus on this basis Stott [56] concludes
that the formation of an oxide glaze is not controlled by the underlying oxidation process but by physical
adsorption interactions between contacting particle surfaces as the glaze sinters.
One limit of this model is that it fails to predict the temperature for transition from severe to mild
wear, underestimating the wear observed at temperatures over the range 150°C–400°C by a factor of
3–5 times. This would suggest that the model incorrectly predicts the rate of oxide debris formation,
prior to its compaction into a protective glaze; possibly an alternative mechanism, similar to that pro-
posed by Quinn et al. [63–65] operates in this low temperature regime.
12-12 Theory and Practice of Lubrication and Tribology

12.5  Oxidation Effects in Fretting Wear


Fretting wear results when two contacting surfaces experience small oscillatory movement under load.
Studies of the friction and wear behavior in dry environments under fretting load conditions of a variety
of materials, including mild steels, stainless steels, and titanium alloys [21,66–72], have shown that as
the temperature of the contact zone is increased there is an abrupt decrease in wear rate, above some
threshold value. For mild steels [66,67] and stainless steels [69,72] this threshold is around 200°C, which
Hurricks [66,67] associated with the formation of a compact oxide debris layer between the contacting
surfaces. For titanium alloys, such oxidation debris effects may be observed at room temperature [21,70].
Thus according to Hurricks [67], fretting wear and oxidative effects on the fretting behavior pro-
ceeds in three phases. First during the early cycles, up to a few thousand cycles, metal–metal contact
prevails at the fretting surfaces and the major mechanism of damage is metal transfer due to adhesion
by local welding of asperities, which then fracture leading to a roughened surface. Phase two sees oxi-
dation of this debris and the accumulation of this oxidized or partially oxidized material within the
contact zone. In this second phase, oxidation–surface interactions are complex, oxidation of existing
wear particles may occur, abrasive wear by hard oxide particles may occur if the oxide particles are
harder than the mechanically deformed sub-rate metal, and oxidation of the newly exposed surface
metal may also result. Ultimately, phase two transforms to the final phase where the abrasive action of
the wear particles is decreased and a compact bed of entrapped oxide debris results, separating the fret-
ting surfaces. The generation of this trapped debris is of critical importance in fretting wear as it pro-
vides a load-carrying plateau. In this stage, fatigue interactions on damage processes begin to control
fretting failures. In this respect, the early material wear behavior under oxidative fretting conditions
is similar to that reported by Stott and coworkers [10,11,14,56,60] in their studies of friction and wear
of like-on-like reciprocating sliding in air at elevated temperatures (see Section 12.4). They observed a
decrease in friction and wear as the temperature increased, see Figure 12.5, which was the direct result
of the formation of an oxide “glaze,” an oxidized, or partially oxidized, compact, load-bearing layer on
the wear surfaces.
In analyzing their observed fretting wear of various aerospace stainless steels, Rybiak et al. [72] intro-
duced a novel “accumulated energy”-based analysis to predict the transition from severe to mild fretting
wear. They measured the tangential force and displacement of the contacting surfaces for various applied
loads and at various test temperatures between 23°C and 400°C. From the hysteresis loop of these two
measured parameters, they were able to calculate the energy dissipated per cycle and by summation,
the total energy dissipated over a large number of repeat cycles. Taking the test at 200°C (later shown
to be the condition just before a stable compact oxide layer is formed) as a reference, it was shown that
the wear volume was a linear function of the energy accumulated in the contact zone (see Figure 12.6).
Further, by plotting the slope of these energy–wear rate curves against temperature, it was shown that a
sharp drop in wear rate was observed at 220°C, with the wear rate per joule of energy dissipated in the
surface dropping by an order of magnitude. This condition corresponded to the formation of a compact,
“glaze-like” oxide layer on the contact surface, which for the stainless steel studied was shown to be a
mixture of γ-Fe, Fe2O3, and/or Cr2O3, that is, partially oxidized wear debris.
For the case of oxidative fretting wear of various stainless steels, this dependence of wear volume on
the accumulated energy density proves an interesting concept. It has been used to successfully model the
effects of variable temperature cycles, variable number of cycles at each temperature condition, and the
total exposure, with a precision better than ±10% for tests cycled between 200°C and 400°C.
Given the close similarity between the oxide glazes under reciprocating wear observed by Stott
and coworkers [10,11,14,56,60], one must ask “would such an energy density criteria also predict
the onset of glaze formation under high temperature sliding wear conditions?” One caveat is that
under reciprocating slide wear, the sliding amplitudes are much greater and this would aid particle
release and thus improve debris ejection from the contacting surfaces. This could modify the local
Oxidative Wear 12-13

3.5
α200°C =1.2
3.0 2
R = 0.97

Normalized total wear volume:


2.5 α400°C = 0.14
R2 = 0.95
2.0
V = V/V ref.
α200°C 200°C
1.5 400°C

1.0

0.5

α400°C
0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Normalized accumulated dissipated energy:
ΣEd = ΣEd/ΣEdref.

FIGURE 12.6  Normalized total wear volume as a function of the accumulated dissipated energy at 200°C and
400°C for an aerospace stainless steel. (From Rybiak, R. et al., Wear, 268, 413, 2010. With permission.)

surface energy balance and may result in some observed nonlinearity of wear volume on accumu-
lated energy density.

12.6  Modeling Erosion–Oxidation Processes


In many high-temperature systems, such as gas turbines, steam turbines, fluidized bed combustor, gas-
ifiers, etc., solid particle erosion and erosion–oxidation synergistic processes are a major cause of wear.
In such systems, the survival of high temperature alloys depends on the formation and maintenance
of a protective surface oxide scale (see Section 12.2). Solid particle erosion is one mechanism by which
this protective oxide barrier can fail. The continual breakdown of this surface oxide allows contami-
nants direct access to the substrate, increasing alloy degradation through enhanced rates of corrosion
[13,73,74]. Under conditions where the particle loading is very high, erosion may be the dominant deg-
radation mechanisms; here, only transient surface scales form [74–76]. In this case, the degradation pro-
cess is controlled primarily by the substrate alloy properties and ductile mechanisms can be expected
to predominate.
Worst case scenarios are normally found where there is a combination of both erosion and ­oxidation/
corrosion, where the erosive component is continuously able to damage to the oxide scale, negating its
ability to act as an environmental barrier. If the surface oxide layer does not crack or spall, then oxida-
tion may reduce material loss due to the formation of stable oxide glazes in a similar manner to that
described by Stott and coworkers [10,11,14] in the mitigation of sliding wear. More often than not the
combination of oxidation/corrosion and erosion results in excessive wear rates.

12.7  Basic Erosion Mechanisms


Erosion mechanisms can be broadly divided into two main categories: ductile erosion and brittle ero-
sion. The major difference between the two erosion mechanisms becomes obvious when comparing
the erosion rate at different angles of impact. In the erosion of ductile materials the maximum erosion
12-14 Theory and Practice of Lubrication and Tribology

rate is usually found to occur at an impact angle close to 30°, while in brittle materials this will occur at
approximately 90° impact.
The erosion of soft, tough materials is predominantly ductile and material loss can occur in two dif-
ferent ways: cutting wear due to impact at low angles [77] and extrusion at high angles. Fragmentation
of the impacting particles can occur at normal or near-normal impact angles, giving rise to second-
ary erosion [78]. These two mechanisms are not mutually exclusive, and erosion of ductile materials is
essentially a combination of the two processes, with one being dominant over the other depending on
impact angle, material properties, and particle properties and shape. There are many papers available
in the literature, which deal with calculating and modeling the wear rates of ductile materials; classical
papers are [77,79,80].
Material loss in solid particle erosion of brittle materials occurs predominantly through the forma-
tion and interaction of a subsurface microcrack network. In order for these cracks to develop, the sur-
face stresses must reach a critical value to initiate microcracking. When these cracks propagate and
intersect with the surface, material is lost [77,81–84]. Hence, for a brittle material, the erosion resistance
is a function of its fracture toughness and its resistance to crack initiation.
The modes of deformation and fracture depend on the particle velocity, shape, and its mechanical
properties relative to those of the target material [77,85–87]. Blunt particles traveling at low velocities set
up Hertzian stress fields in the target, which initiate cone cracking, while angular particles traveling at
high velocities produce inelastic deformation zones and initiate median and lateral cracking [88].

12.8  Erosion–Oxidation
The big question is what happens when both erosion and oxidation are combined. It all depends on
the oxidation kinetics of the system being investigated, that is, fast growth (e.g., NiO) or slow growth
(e.g., Cr2O3) oxide, temperature, whether the oxide is protective, its mechanical properties, and the ero-
sion conditions. First, one should consider the fundamentals of oxide growth from an erosion perspec-
tive. As explained in Section 12.2, oxidation is a thermochemical process, and this manifests itself, for
most structural materials, in the growth of a surface oxide scale. In alloy systems, there is competition
between the various alloy constituents, which gives rise to selective oxidation, the formation of multi-
layered scales, internal oxidation, and other more complex bioxidant oxidation reactions. One initially
observes the formation of a “transient oxide” on any exposed metallic alloy surface. The transient oxide
usually consists of stable oxide species of all the oxidizable elements in the alloy. Competition between
the growth of these species usually leads to the formation of some protective oxide, be it alumina, chro-
mia, or an iron–chromium spinel, depending on the alloy being oxidized. This is the desirable service
condition, with components protected by a stable, slow-growing surface oxide. Problems are encoun-
tered when this oxide scale is damaged, be it due to ballistic impaction, erosion, mechanical stressing,
or thermal cycling, whereupon non-protective oxidation may ensue leading to thicker oxide scales and
faster oxidation rates.
The onset and extent of such mechanically induced failures depends primarily on interplay between
three factors:
1. Defects, such as voids, pores, and microcracks, that develop in the scale as it grows [34,41,89,90].
2. Stresses that act on the scale. These may result from impact events, but are supplemented by those
generated due to oxide growth and due to thermally induced transients within the system under
evaluation. The influences of such stresses will be locally concentrated by the defects in the scale.
3. Stress relief, within the oxide scale–substrate system can limit the extent of such mechanical fail-
ures of the oxide scale. This occurs primarily by plastic deformation of the near surface region,
while at temperature. From an erosion–oxidation perspective, this stress relief has a significant
influence on the boundary between “oxide-modified” and “substrate-dominated” erosion, using
the terminology of Stephenson and Nicholls [76].
Oxidative Wear 12-15

The interaction between these three factors, and how they depend on the impact event, will determine
when scale failure occurs. The mechanically damaged scale can be healed by forming a new oxide—this
involves both the transient oxidation and protective growth stages—and the relative times between this
oxide loss, healing, and regrowth process dictate the observed erosion–oxidation regime.
One final point, continued impact, scale spallation, and regrowth will lead to depletion of oxide-
forming elements from within the near surface region of the alloy. This will have an influence on the
type of oxide formed, its mechanical properties, and the mechanical properties of the underlying alloy.
These time-dependent phenomena further complicate the modeling/prediction of erosion–oxidation.
Over the last 30 years, there has been considerable research on high-temperature erosion [91–109],
with most of the work concentrating on stainless steels, nickel-based superalloys, and titanium alloys,
although some work was also conducted on ceramics. These studies have researched various test meth-
ods, as reviewed by Nicholls [110], and have considered alternative modeling [111–115] and mapping
techniques [16,17,26,116,117].
The oxidation properties of the base metal combined with the flux of the erodent will determine the
thickness of the oxide that is formed; the velocity and size of the particles will also have an influence.
Thus, Barkalow and Pettit [74] defined erosion–oxidation in terms of the kinetic energy of the impact-
ing particles and scale growth rate identifying four regimes, which are with increased particle kinetic
energy and/or particle flux:

1. Oxidation
2. Erosion-modified oxidation
3. Oxidation-modified erosion
4. Metallic erosion

The erosion properties of the oxide layer become important when the impact energy and the particle flux
are both sufficiently low that the oxide grows to a sufficient thickness that the impact event is totally con-
tained within the oxide layer. Conversely, at very high particle fluxes and kinetic energies only substrate
properties are important. Under intermediate oxidation rates and particle fluxes, erosion and oxidation
processes synergistically interact.
Most researchers have divided the erosion–oxidation process into four or six regimes, ranging from
pure erosion on one side to pure oxidation on the other. For most cases there is a close similarity in
approach, with four regimes the most popular as summarized in Table 12.2 [17]. For the proposed
six-step regimes [121,122], the erosion–oxidation interactions may be further subdivided as illustrated
in Figure 12.7, where three modes of erosion-enhanced corrosion are introduced.

TABLE 12.2  Various Proposed Erosion–Oxidation Regimes, Divided in Three or Four Steps
Roy et al. [118] Metal erosion Oxidation-affected Oxidation-controlled Oxide erosion
erosion erosion
Stephenson and Substrate dominated Oxide modified Oxide dominated Oxidation kinetics
Nicholls [76]
Barklow and Pettit [74] Metallic erosion Oxidation-modified Erosion-modified Oxidation
erosion oxidation
Kang et al. [119] Pure erosion Oxidation-affected Erosion-enhanced Erosion of oxide
erosion oxidation only
Stack and Stott [115] Erosion dominated Erosion–corrosion Corrosion dominated 1 Corrosion
dominated dominated 2
Wright et al. [120] Erosion dominated Erosion–oxidation Oxidation dominated
interaction
Sources: Hogmark, S. et al., Proceedings of Sixth International Conference on Erosion by Solid and Liquid Impact, 1983,
p. 37; Rishel, D.M. et al., Corrosion Sci., 35, 1007, 1993.
Note: Six-step regimes would further subdivide the erosion–oxidation interactions.
12-16 Theory and Practice of Lubrication and Tribology

Oxide Erosion of
C oxide only
Metal
E o
r r
o r
s o
i s Type II
o Oxide
i
n o
r n
Erosion-
a r enhanced
t a corrosion
e t Oxide
e
Metal Type I

Spalls

Type III
Metal

Composite layer
Oxidation

Affected
Metal

No Erosion of
Metal metal only
Corr

FIGURE 12.7  Schematic of the different erosion–oxidation regimes proposed by Rishel et al. (From Rishel, D.M.
et al., Corrosion Sci., 35, 1007, 1993. With permission.)

Laboratory tests indicate that a distinct erosion–corrosion response with respect to temperature
exists, with wastage rates first increasing with an increase in temperature to some peak temperature
before starting to decrease with a further increase in temperature [95,106]. This behavior differed from
in situ tests within a fluidized bed combustor where wastage due to erosion–corrosion was found to
decrease with an increase in temperature, above a critical temperature [123]. Such behavior is similar to
that observed for oxidative wear under reciprocating sliding wear conditions. This difference in oxida-
tive wear under erosion–oxidation conditions was attributed to the internal cooling introduced with in
situ tests on heat-exchanger tubes. The internal cooling of a component is thought to cause the isother-
mal erosion–corrosion curve to broaden, with a shift of maximum wastage rates to lower temperatures,
resulting in wear damage similar to that observed for oxidative-sliding wear.
Rishel et al.’s [122] description of erosion–oxidation is very similar to that of Stephenson and Nicholls
[76], but includes three different erosion-enhanced corrosion regimes, as illustrated in Figure 12.7. Type
III, based on oxide spallation, is the same as the oxide-modified regime proposed by Stephenson and
Nicholls. However, in type I and type II the thickness of the oxide is reduced due to erosion and as a
result the rate of oxidation increases compared to pure oxidation, also cracks caused by the impacting
particles lead to short circuit diffusion for oxygen, further increasing the local oxidation rate.
Oxidative Wear 12-17

12.9  Modeling and Mapping Erosion–Oxidation


Erosion–oxidation mapping (E-O maps) and more recently erosion–corrosion (aqueous) mapping
(E-C maps) have become a popular way to represent how systems will behave under different ­erosion–
oxidation (and erosion–corrosion) conditions [16,17,26,104,111]. Stephenson and Nicholls [111] plot-
ted particle velocity vs. oxide thickness for a specific particle size and temperature, while Stack and
Pena [104] plotted velocity vs. temperature. In both cases contours indicate the different regimes of
­erosion–oxidation wastage, which in turn translates into regions of high, medium, and low mass loss.
Figure 12.8 illustrates such an erosion–oxidation map for IN738, following 90° impact at 700° [111]. In
producing this map, Stephenson and Nicholls took the concept of maps one step further by incorpo-
rating wastage rate contours onto an erosion–oxidation map of impact velocity vs. particle size for a
set temperature. The wastage rate contours were calculated by Monte Carlo modeling of the erosion–­
oxidation process, which allowed differentiation between the different erosion–oxidation regimes for
each impact event [111,124].
The major drawback of such maps is that in order to fully characterize a system one needs a whole
family of maps to cover the main variables, namely, velocity, temperature, particle size, and flux.
Such compendiums have been produced by Stack and Pena [116], who derived a family of wear
maps of velocity vs. particle size for three different compositions and four different temperatures.
These were then used to map materials performance from the low wastage regimes identified in each
erosion–oxidation map.
Erosion–oxidation modeling has followed one or two main development routes. The first involves an
empirical “equation” approach, where the oxidation kinetics of a system are combined with simple mass
(or volume) loss equations to generate a single complex equation to predict mass loss [114,115]. However,
this type of model has a limited applicability and is normally only valid for one erosion–oxidation
regime.
Stack and Stott [115] modeled erosion–oxidation in four regimes. It was based on the assumption
that the oxidation process followed parabolic kinetics between impacts, while oxidation that occurred
during impact was negligible. It was also assumed that a critical scale thickness existed, which was

1000
1 mm/1000 h

100
Impact velocity (m/s)

Substrate dominated
100 µm/1000 h

Scale + substrate

10
10 µm/1000 h

Scale modified

1
1 10 100 1000
Particle size µm

FIGURE 12.8  Erosion–oxidation map for IN738 impacted at 90° by silica beads at 700°C. (From Stephenson, D.J.
and Nicholls, J.R., Wear, 186–187, 284, 1995. With permission.)
12-18 Theory and Practice of Lubrication and Tribology

used to limit the oxide thickness that could be removed, until above some critical temperature. By
defining weight gain per unit area as a function of time and temperature and mass loss due to erosion
as a function of time per unit area, Stack and coworkers were able to construct erosion–oxidation
maps of velocity vs. temperature, from which critical factors that change the transition zones could be
identified.
An alternative, but more complex, approach, which was taken by Stephenson and Nicholls [124,125],
was to use Monte Carlo statistical methods to calculate the damage caused by each individual impacting
particle. This approach enables the model to assess each particle–oxide (or substrate) interaction and
hence ascertain the erosion–oxidation regime and determine the amount of material removed as a result
of the local impact conditions. It is this area of substrate–oxide–particle interaction that is impossible to
solve analytically, especially as it strongly depends on the individual particle’s size and kinetic energy.
Unfortunately, not all particles are the same size. Under these complexed particle–surface interaction
conditions, the Monte Carlo method offers a solution that can integrate the damage across a range
of particle sizes, involving one or more erosion–oxidation mechanisms. The Monte Carlo modeling
method has been described in detail in a recent review of erosion–oxidation mechanisms, maps, and
models [17] by the current authors.
Figure 12.9 illustrates the model’s ability to predict metal losses across a range of erosion mecha-
nisms. In Figure 12.9, predicted metal loss rates for IN738LC impacted at 90° impact angle, in a particle
flux of 0.1 kg/m2/h at 700°C, are presented. At high impact velocities, the erosion rate is dominated by
the substrate behavior. As the impact velocity falls, oxide scales have a chance to form, such that scale
removal becomes the controlling mechanism. Under these conditions, predicted erosion rates drop by a
factor of 1000, as a protective oxide glaze is formed.

10,000
n = 2.1
Accumulated
strain
1,000

100
Metal loss (µm/1000 h)

Scale removal n=6


n = 0.8 Low cycle
4 µm fatigue
10

1
1000 µm

0.1

Scale modified Substrate dominated


0.01
1 10 100 1000
Impact velocity (m/s)

Particle size
4 µm 10 µm 100 µm 1000 µm

FIGURE 12.9  Predicted metal loss rates for IN738LC impacted at 90° impact angle, in a particle flux of 0.1 kg/m2/h
at 700°C. (From Nicholls, J.R. and Stephenson, D.J., Wear, 186–187, 64, 1995. With permission.)
Oxidative Wear 12-19

Monte Carlo methods have been used to model the erosion–oxidation behavior of a number of high
temperature alloys and coatings proposed for use within a gas turbine as part of a coal-fired combined
cycle power plant [124]. The materials studied included IN738, various coatings on IN738, a stainless
steel (AISI 310), and IN800H. A comparison between measured and predicted erosion rates in this
study [124] has shown that Monte Carlo methods can be successfully used to simulate the high tem-
perature erosion conditions expected in such a coal-fired power plant. Agreement between experiment
and prediction was very good, within a factor of 3, with predicted erosion rates marginally above those
measured, thus providing a conservative estimate of component life.

12.10  Summary
This review has illustrated the complex nature of oxidative wear processes. The review has examined
mechanisms and models for oxidative-sliding wear, fretting-oxidation interactions, and the mecha-
nisms and models developed to predict erosion–oxidation behavior. The numerous factors that affect
system performance has shown that oxidative wear is a complex process involving thermochemical
processes, ductile and brittle material behavior, fracture mechanics, and thermally induced material
aging mechanisms. Various different mechanisms and how the system variables affect the transitions
between the different oxidative wear regimes have been discussed. Mapping offers a useful tool to iden-
tify the regime in which a system will be operated. This has proved useful in the study of oxidative wear
and erosion–oxidation processes to determine the degree of material wastage that can be expected for
a given system. Finally, different concepts of modeling have been discussed by example, from analyti-
cal models of oxidative wear, through finite element methods to Monte Carlo statistics in modeling the
erosion–oxidation behavior of different systems.

References
1. J. F. Archard, W. Hirst, The wear of metals under unlubricated conditions, Proc. R. Soc. A236 (1956),
397–410.
2. T. F. J. Quinn, The role of oxidation in the mild wear of steel, Br. J. Appl. Phys. 13 (1962), 33–37.
3. N. C. Welsh, The dry sliding of steels, Phil. Trans. R. Soc. A257 (1965), 31–50.
4. T. F. J. Quinn, The effect of ‘hot spot’ temperatures on the unlubricated wear of steel, ASLE Trans. 10
(1967), 158–165.
5. F. H. Stott, G. C. Wood, The influence of oxides on the friction and wear of alloys, Tribol. Int. 11
(1978), 211–218.
6. T. F. J. Quinn, D. M. Rowson, J. L. Sullivan, Application of the oxidational theory of mild wear to the
sliding wear of low alloy steel, Wear 65 (1980), 1–20.
7. R. B. Waterhouse, P. E. Taylor, High temperature fretting and wear of like metallic contacts, Rev. High
Temp. Mater. 4 (1980), 259–298.
8. T. F. J. Quinn, Review of oxidational wear, Tribol. Int. 15 (1983), 257–315.
9. J. P. Archard, R. A. Rowntree, The temperature of rubbing bodies. Part 2. The distribution of tempera-
tures, Wear 128(1) (1988), 1–17.
10. F. H. Stott, J. Glascott, G. C. Wood, Factors affecting the progressive development of wear-protective
oxides on iron-base alloys during sliding at elevated temperatures, Wear 97 (1984), 93–106.
11. F. H. Stott, M. P. Jordan, The effects of load and substrate hardness on the development and mainte-
nance of wear-protective layers during sliding at elevated temperatures, Wear 250 (2001), 391–400.
12. J. L. Sullivan, S. S. Athwal, The role of oxidation in the wear of alloys, Tribology Int. 16 (1983), 123–131.
13. N. Birks, F. S. Pettit, D. M. Rishel, Erosion–corrosion and wear, J. Phys. IV 4 (1993), 667–678.
14. F. H. Stott, The role of oxidation in the wear of alloys, Tribol. Int. 31 (1998), 61–71.
15. H. A. Abdel-Aal, On the influence of thermal properties on wear resistance of rubbing metals at
elevated temperatures, J. Tribol. 122 (2000), 657–660.
12-20 Theory and Practice of Lubrication and Tribology

16. M. M. Stack, Mapping tribo-corrosion processes in dry and in aqueous conditions, Tribol. Int. 35
(2002), 681–689.
17. R. G. Wellman, J. R. Nicholls, High temperature erosion–oxidation mechanisms, maps and models,
Wear 256 (2004), 907–917.
18. P. N. Bogdanovich, V. M. Belov, D. V. Tkachuk, Dimensions and kinetics of local heat sources in
rubbing solid contact, Int. J. Appl. Mech. Eng. 11 (2006), 5–13.
19. J. E. Wilson, F. H. Stott, G. C. Wood, The development of wear protective oxides and their influence
on sliding friction, Proc. R. Soc. 369A (1980), 557–574.
20. I. A. Inman, S. Datta, H. L. Du, J. S. Burnell-Gray, Q. Luo, Microscopy of glazed layers formed during
high temperature sliding wear at 750°C, Wear 254 (2003), 461–467.
21. J. Ding, S. B. Leen, E. J. Williams, P. H. Shipway, A multi-scale model for fretting wear with oxidation-
debris effects, J. Eng. Tribology 223 (2009), 1019–1032.
22. J. R. Nicholls, Coatings and hardfacing alloys for corrosion and wear resistance in diesel engines,
Mater. Sci. Tech. 10 (1994), 1002–1012.
23. S. Atamert, H. K. D. H. Bhadeshia, Comparison of the microstructure and abrasive wear properties
of stellite hardfacing alloys deposited by arc welding and laser cladding, Metals Technol. 20 (1989),
1037–1054.
24. S. Michler, Tribo-electrochemical techniques and interpretation methods in tribo corrosion: A com-
parative evaluation, Tribol. Int. 41 (2008), 573–583.
25. R. J. K. Wood, Tribo-corrosion of coatings: A review, J. Phys. D 40 (2007), 5502–5521.
26. M. M. Stack, Bridging the gap between tribology and corrosion: From wear maps to Pourbaix
diagrams, Int. Mater. Rev. 50 (2005), 1–17.
27. U. R. Evans, Metallic Corrosion, Passivity and Protection, 2nd edn., Edward Arnold and Co-publishers,
London, U.K. (1946), pp. 233–239.
28. D. R. Sigler, Aluminum oxide adherence on Fe-Cr-Al alloys modified with group IIIB, IVB, VB, and
VIB elements, Oxid. Met. 32 (1989), 337–355.
29. W. J. Quadakkers, M. J. Benett, Oxidation induced lifetime limits of thin walled iron based ODS alloy
components, Mater. Sci. Technol. 10 (1994), 126–131.
30. B. A. Pint, Experimental observations in support of the dynamic-segregation theory to explain the
reactive-element effect, Oxid. Met. 45 (1996), 1–37.
31. C. Wagner, Beitrag zur theorie des anlaufvorgangs, Z. Phys. Chem. B21 (1933), 25–41.
32. C. Wagner, Theoretical analysis of the diffusion processes determining the oxidation rate of alloys,
J. Electrochem. Soc. 99 (1952), 369–380.
33. P. Kofstad, High Temperature Corrosion, Elsevier Applied Science, London, U.K. (1988).
34. M. Schütze, Protective Oxide Scales and Their Breakdown, Wiley, Chichester, U.K. (1997).
35. N. F. Mott, Atomic physics and the strength of metals, J. Inst. Met. 72 (1946), 367–380.
36. N. Cabrera, N. F. Mott, Theory of the oxidation of metals, Rep. Prog. Phys. 12 (1949), 163–184.
37. W. J. Quadakkers, Growth mechanisms of oxide scales on ODS alloys in the temperature range
1000–1100°C, Werkst. Korros. 41 (1990), 659–668.
38. W. J. Quadakkers, D. Naumenko, E. Wessel, V. Kochubey, L. Singheiser, Growth rates of alumina
scales on Fe-Cr-Al alloys, Oxid. Met. 61 (2004), 17–38.
39. O. Kubaschewski, B. E. Hopkins, Oxidation of Metals and Alloys, Butterworths Scientific Publications,
London, U.K. (1953), p. 131.
40. B. Chattopadhyay, G. C. Wood, The transient oxidation of alloys, Oxid. Met. 2 (1970), 373–399.
41. P. Hancock, J. R. Nicholls, Application of fracture mechanics to the failure of surface oxide scales,
Mater. Sci. Technol. 4 (1988), 398–406.
42. D. Bruce, P. Hancock, Influence of mechanical properties of surface oxide films on oxidation mech-
anisms I. A vibration technique to study the nature and growth of thermally formed oxide films
metals, J. Inst. Met. 97 (1969), 140–147.
Oxidative Wear 12-21

43. D. Bruce, P. Hancock, II Mechanical properties and adhesion of surface oxide films on iron and
nickel measured during growth, J. Inst. Met. 97 (1969), 148.
44. R. C. Hurst, P. Hancock, Measurement of the mechanical properties of growing surface films in oxi-
dising environments, Werkst. Korros. 23 (1972), 773–776.
45. M. Schütze, A. Rahmel, Influence of constant strain rates on growth and cracking behavior of oxide
scales and on internal corrosion of a 18Cr-0. 8Al-1. 5Si Steel, In International Corrosion Conference
Series High Temperature Corrosion (R. A. Rapp, ed.), NACE, Houston, TX (1983), 421–429.
46. M. Schütze, Deformation and cracking behavior of protective oxide scales on heat-resistant steels
under tensile strain, Oxid. Met. 24 (1985), 199–232.
47. M. Schütze, Stress and decohesion of oxide scales, Mater. Sci. Technol. 4 (1988), 407.
48. W. Christi, A. Rahmel, M. Schütze, Behavior of oxide scales on 2.25Cr-1Mo steel during thermal
cycling. I. Scales formed in oxygen and air, Oxid. Met. 31 (1989), 1–34.
49. W. Christl, A. Rahmel, M. Schütze, Behavior of oxide scales on 2.25Cr-1Mo steel during thermal
cycling. II. Scales grown in water vapor, Oxid. Met. 31 (1989), 35–69.
50. M. Schütze, Plasticity of protective oxide scales, Mater. Sci. Technol. 6 (1990), 32–38.
51. T. Suga, I. Kvernes, G. Elssner, Fracture energy measurements of ceramic thermal barrier coatings.
Mater. Sci. Eng. Technol. 15 (1984), 371–377.
52. J. Robinson, M. I. Manning, Limited to the adherence of oxide scales, Mater. Sci. Technol. 6 (1990),
81–92.
53. D. J. Stephenson, J. R. Nicholls, Role of surface oxides in modifying solid particle impact damage.
Mater. Sci. Technol. 6 (1990), 96–99.
54. T. F. J. Quinn, J. L. Sullivan, A review of oxidational wear, in Proceedings of International Conference
on Wear of Materials, St. Louis, MO, ASME, New York (1977), p. 110.
55. M. Fink, Wear oxidation—A new component of wear, Trans. Am. Soc. Steel Treat. 18 (1930),
1026–1034.
56. F. H. Stott, High temperature sliding wear of metals, Tribology Int. 35 (2002), 489–495.
57. S. C. Lim, M. F. Ashby, Wear-mechanism maps, Acta Metall. 35 (1987), 1–24.
58. T. F. J. Quinn, Oxidational wear modelling: I. Wear 153 (1992), 179–200.
59. T. F. J. Quinn, Oxidation wear modelling: Part II. The general theory of oxidational wear, Wear 175
(1994), 199–208.
60. D. S. Lin, F. H. Stott, G. C. Wood, Effects of elevated ambient temperatures on the friction and wear
behaviour of some commercial nickel based alloys, ASLE Trans. 17 (1974), 251–262.
61. Y. H. Zhou, M. Harmelin, J. Bigot, Sintering behaviour of ultra-fine Fe, Ni and Fe-25wt%Ni powders,
Scr. Metall. 23 (1989), 1391–1396.
62. J. Jiang, F. H. Stott, M. M. Stack, A mathematical model for sliding wear of metals at elevated tempera-
tures, Wear 181–183 (1995), 20–31.
63. T. F. J. Quinn, J. L. Sullivan, D. M. Rowson, Origins and development of oxidational wear at low
ambient temperatures, Wear 94 (1984), 175–191.
64. T. F. J. Quinn, Oxidational wear modelling part III. The effects of speed and elevated temperatures,
Wear 216 (1998), 262–275.
65. T. F. J. Quinn, Review of oxidational wear. Part II: Recent developments and future trends in oxida-
tional wear research, Tribol. Int. 16 (1983), 305–315.
66. P. L. Hurricks, The fretting wear of mild steel from room temperature to 200°C, Wear 19 (1972),
207–229.
67. P. L. Hurricks, The mechanisms of fretting—A review, Wear 15 (1970), 389–409.
68. A. Iwabuchi, The role of oxides particles in the fretting wear of mild steel, Wear 151 (1991), 301–311.
69. T. Kayaba, A. Iwabuchi, The fretting wear of 0.45%C steel and austenitic stainless steel from 20 to
650°C in air, Wear 74 (1981), 229–245.
70. R. B. Waterhouse, Fretting Corrosion, Pergamon Press, Oxford, U.K. (1972).
12-22 Theory and Practice of Lubrication and Tribology

71. R. B. Waterhouse, D. E. Taylor, Fretting debris and the delamination theory of wear, Wear 29 (1974),
337–344.
72. R. Rybiak, S. Fouvry, B. Bonnet, Fretting wear of stainless steels under variable temperature condi-
tions: Introduction of a composite wear law, Wear 268, (2010), 413–523.
73. D. J. Stephenson, J. R. Nicholls, P. Hancock, The interaction between corrosion and erosion during
simulated sea salt compressor shedding in marine gas turbines. Corros. Sci. 26 (1986), 757–767.
74. R. H. Barkalow, F. S. Pettit, Corrosion/Erosion of Coal Conversion System Materials, Berkeley, CA
(1979).
75. J. E. Restall, D. J. Stephenson, High temperature erosion of coated superalloys for gas turbines. Mater.
Sci. Eng. 88 (1987), 273–282.
76. D. J. Stephenson, J. R. Nicholls, Modelling the influence of surface oxidation on high temperature
erosion, Wear 186–187 (1995), 284–290.
77. I. Finnie, Erosion of surfaces by solid particles, Wear 3 (1960), 87–103.
78. G. P. Tilly, A two stage mechanism of ductile erosion, Wear 23 (1973), 87–96.
79. A. V. Levy, The platelet mechanism of erosion of ductile metals, Wear 108 (1986), 1–21.
80. G. P. Tilly, Erosion caused by impact of solid particles. Treatise Mater. Sci. Technol. 13 (1979), 287–319.
81. B. R. Lawn, A. G. Evans, D. B. Marshall, Elastic/plastic indentation damage in ceramics: The median/
radial crack system, J. Am. Ceram. Soc. 63 (1980), 574–581.
82. A. G. Evans, M. E. Gulden, M. Rosenblatt, Impact damage in brittle materials in the elastic-plastic
response regime, Proc. R. Soc. 361 (1978), 343–365.
83. B. Lawn, R. Wilshaw, Indentation fracture: Principles and applications, J. Mater. Sci. 10 (1975),
1049–1081.
84. C. M. Perrott, Elastic-plastic indentation: Hardness & fracture, Wear 45 (1977), 293–309.
85. S. Wada as quoted by S. Srinivasan and R. O. Scattergood, Effect of erodent hardness on erosion of
brittle materials, Wear 128 (1988), 139.
86. S. Srinivasan, R. O. Scattergood, Effect of erodent hardness on erosion of brittle material, Wear 128
(1988), 139–152.
87. I. M. Hutchings, Transitions, threshold effects and erosion maps, in Erosion of Ceramic Materials
(J. E. Ritter, ed.), Trans Tech Publications, Enfield, NH (1992), pp. 75–92.
88. S. M. Wiederhorn, B. J. Hockey, Effect of material parameters on the erosion resistance of brittle
materials, J. Mater. Sci. Technol. 18 (1983), 766–780.
89. M. M. Nagal, W. T. Evans, The mechanical failure of oxided scales under tensile or compressive load,
J. Mater. Sci. 28 (1993), 6247–6260.
90. P. Hancock, J. R. Nicholls, Failure of oxide scales, Mater. High Temp. 12 (1994), 209–218.
91. W. Tabakoff, Erosion resistance of superalloys and different coatings exposed to particulate flows at
high temperature, Surf. Coat. Technol. 120–121 (1999), 542–547.
92. P. M. Rogers, I. M. Hutchings, J. A. Little, Coatings and surface treatments for protection against
low-velocity erosion-corrosion in fluidized beds, Wear 186–187 (1995), 238–246.
93. P. M. Rogers, T. E. Howes, I. M. Hutchings, J. A. Little, Wastage of low alloy steels in a fluidized bed
environment, Wear 186–187 (1995), 306–315.
94. T. E. Howes, P. M. Rogers, J. A. Little, I. M. Hutchings, Erosion-corrosion of mild steel in a tempera-
ture gradient, Wear 186–187 (1995), 316–324.
95. V. K. Sethi, R. G. Corey, High temperature erosin of alloys in oxidising environments, in Proceedings
of the Seventh International Conference on Erosion by Liquid and Solid Impact, Cambridge, U.K.
(1987).
96. S. Chinnadurai, S. Bahadur, High-temperature erosion of Haynes and Waspaloy: Effect of tempera-
ture and erosion mechanisms, Wear 186–187 (1995), 299–305.
97. P. Hancock, J. R. Nicholls, D. J. Stephenson, The mechanism of high temperature erosion of coated
superalloys. Surf. Coat. Technol. 32 (1987), 285–304.
Oxidative Wear 12-23

98. A. Levy, Y. F. Man, Surface degradation of ductile metals in elevated temperature gas-particle streams,
Wear 111 (1986), 173–186.
99. A. V. Levy, J. Yan, J. Patterson, Elevated temperature erosion of steels, Wear 108 (1986), 43–60.
100. N. Gat, W. Tabakoff, Some effects of temperature on the erosion of metals, Wear 50 (1978), 85–94.
101. D. J. Stephenson, J. R. Nicholls, P. Hancock, Particle–surface interactions during the erosion of a gas
turbine material (MarM002) by pyrolytic carbon particles, Wear 111 (1986), 15–29.
102. D. J. Stephenson, J. R. Nicholls, P. Hancock, Particle-surface interactions during the erosion of
aluminide-coated MarM002, Wear 111 (1986), 31–39.
103. M. Suckling, C. Allen, Critical variables in high temperature erosive wear, Wear 203–204 (1997),
528–536.
104. M. M. Stack, D. Peña, Particle size effects on the elevated temperature erosion behaviour of
Ni-Cr/WC MMC-based coatings, Surf. Coat. Technol. 113 (1999), 5–12.
105. M. M. Stack, Q. Song-Roehrle, F. H. Stott, G. C. Wood, Computer simulation of erosion-corrosion
interactions at elevated temperatures, Wear 181–183 (1995), 516–523.
106. M. M. Stack, J. Chacon-Nava, F. H. Stott, Relationship between the effects of velocity and alloy cor-
rosion resistance in erosion-corrosion environments at elevated temperatures, Wear 180 (1995),
91–99.
107. R. L. Howard, A. Ball, The effect of test temperature on the particle erosion performance of titanium
aluminide alloys, Wear 205 (1997), 11–14.
108. J. R. Zhou, S. Bahadur, Erosion characteristics of alumina ceramics at high temperatures, Wear
181–183 (1995), 178–188.
109. A. L. Ham, J. A. Yeomans, J. F. Watts, Effect of temperature and particle velocity on the erosion of a
silicon carbide continuous fibre reinforced calcium aluminosilicate glass-ceramic matrix composite,
Wear 233–235 (1999), 237–245.
110. J. R. Nicholls, Corrosion in advanced power plants, in Proceedings of Second International Workshop
on Corrosion in Advanced Power Plants, Tampa, FL (1997), p. 219.
111. D. J. Stephenson, J. R. Nicholls, Modelling the influence of surface oxidation on high temperature
erosion, Wear 186–187 (1995), 284–290.
112. D. J. Stephenson, J. R. Nicholls, Modelling erosive wear, Corros. Sci. 35 (1993), 1015–1026.
113. A. J. Markworth, A stochastic model for the simultaneous occurrence of oxidation and erosion,
Mater. Sci. Eng. A 150 (1992), 37–42.
114. A. J. Markworth, V. Nagarajan, I. G. Wright, An approach to modelling simultaneous high tempera-
ture oxidation and erosion, Oxid. Met. 35 (1991), 89–106.
115. M. M. Stack, F. H. Stott, An approach to modelling erosion-corrosion of alloys using erosion-
corrosion maps, Corros. Sci. 35 (1993), 1027–1034.
116. M. M. Stack, D. Pẽna, Mapping erosion of Ni-Cr/WC-based composites at elevated temperatures:
Some recent advances, Wear 250–251 (2001), 1433–1443.
117. M. M. Stack, Corrosion in advanced power plants, in Proceedings of Second International Workshop
on Corrosion in Advanced Power Plants, Tampa, FL (1997), p. 243.
118. M. Roy, K. K. Ray, G. Sundararajan, Erosion-oxidation interaction in Ni and Ni-20Cr alloy, Metall.
Mater. Trans. 32A (2001), 1431–1451.
119. C. T. Kang, F. S. Pettit, N. Birks, Mechanisms in the simultaneous erosion-oxidation attack of nickel
and cobalt at high temperature. Metall. Trans. A 18 (1987), 1785–1803.
120. I. G. Wright, V. K. Sethi, A. J. Markworth, A generalised description of the simultaneous process of
scale growth by high-temperature oxidation and removal by erosive impact, Wear 186–187 (1995),
230–237.
121. S. Hogmark, A. Hammarsten, S. Soderberg, On the combined effects of corrosion and erosion, in
Proceedings of the Sixth International Conference on Erosion by Solid and Liquid Impact, Cambridge,
U.K. (1983), pp. 37–31.
12-24 Theory and Practice of Lubrication and Tribology

122. D. M. Rishel, F. S. Pettit, N. Birks, Spalling types and mechanisms in the erosion-corrosion of metals
at high temperature, Corros. Sci. 35 (1993), 1007–1013.
123. G. J. Holtzer, P. L. F. Rademakers, Studies on 90 MWth AKZO and 4 MWth TNO FBC show excel-
lent erosion-corrosion results, In Proceedings of the International Conference on Fluidized Bed
Combustion 2, ASME, New York (1991), pp. 743–753.
124. J. R. Nicholls, D. J. Stephenson, Monte Carlo modelling of erosion processes, Wear 186–187 (1995),
64–77.
125. D. J. Stephenson and J. R. Nicholls, Modelling high temperature erosion, in Plant Corrosion—
Prediction of Materials Performance (J. E. Strutt and J. R. Nicholls, eds.), Ellis Horwood Ltd,
Chichester, U.K. (1987), pp. 315–324.
13
Wear Models
Nomenclature.............................................................................................. 13-1
13.1 Introduction..................................................................................... 13-2
13.2 Wear Equations................................................................................13-4
Adhesive Wear  •  Abrasive Wear  •  Fatigue Wear  •  Impact Wear
13.3 Wear Maps........................................................................................ 13-7
13.4 Computerized Wear Models..........................................................13-9
Kenneth Holmberg 13.5 Simulation of Surface Stresses and Deformations.................... 13-11
VTT Technical Research 13.6 Parametric Analysis of a Wear Contact..................................... 13-15
Centre of Finland 13.7 Component Wear Models............................................................ 13-17
Anssi Laukkanen 13.8 Conclusions.................................................................................... 13-18
VTT Technical Research Acknowledgments..................................................................................... 13-19
Centre of Finland References................................................................................................... 13-19

Nomenclature
Aa Apparent contact area
C Constant or basic load capacity of rolling bearing
d Wear depth
E Young’s modulus, modulus of elasticity
E′ Composite modulus of elasticity, 1/E' = 0.5 {[(1 – ν12)/E1] + [(1 – ν22)/E2]}
H Hardness
h Coating thickness
Kc Fracture toughness
K′ Wear rate
k General proportionality constant or wear coefficient
L Load
L′ Equivalent load for rolling bearings
m Particle mass
N Number of cycles
N Number of asperities in contact
P Apparent normal pressure (p = L/Aa)
p′ Bearing-type-specific load-life exponent or normalized nominal contact pressure divided by
the contact hardness
R Composite mean curvature of surface asperities or radius
SWR Surface wear resistance
S Sliding distance

13-1
13-2 Theory and Practice of Lubrication and Tribology

T Time
T Temperature
V Wear volume
V′ Material volume lost per unit surface area and per unit sliding distance
V Velocity
v′ Sliding velocity divided by the velocity of heat flow
α Asperity shape factor
β Probability factor for material cuts
μ Coefficient of friction
ν Poisson’s ratio
σ Composite standard deviation of surface asperities
σo Yield strength

13.1  Introduction
Wear is commonly defined as the removal of material from solid surfaces as a result of one contacting
surface moving over another. The material is removed in the form of wear debris. The typical size of
wear debris in visible wear conditions is from about half a micrometer to tens or hundreds of microm-
eters. In severe mining wear conditions, single wear particles can be of centimeter size. In steady-state
low-wear-lubricated conditions in an engine, the size of wear particles in continuous wear may be of
nanometer size.
The detachment of material from the surface is a result of the loading conditions in the contact.
These are the moving normal load and the moving tangential load due to friction. The loading condi-
tions deform the surface material and result in stresses and strains that the material cannot anymore
withstand. The result is shear within the material, deformation, clustering of voids, crack initiation, and
crack propagation, resulting in the formation of loose material debris. Tribochemical reactions at the
top surface may degrade or improve the surface properties and thus have a considerable influence on
wear. The wear products may disappear from the surface, they may agglomerate and attach to the coun-
tersurface to form transfer layers, or they may stay as loose particles in the contact zone and influence
future contact and wear conditions.
Wear is not a material property but a contact system property reflecting the dynamic conditions when
one surface moves against another. One contact system may differ very much from another, resulting in
contact conditions where the variables that influence wear are very different. Similar contact conditions
with similar mechanisms of material removal from the surface have been classified as wear mecha-
nisms. Such wear mechanisms are adhesive, abrasive, rolling fatigue, fretting, erosion, and corrosive
wear. These have been described in detail in the previous chapters.
It is important to understand the influence of various variables on the wear process to be able to pre-
dict and control this process. These variables are numerous, and their influence and interactions have
long been studied but are today understood only on a very general level [1–7].
The basic variables influencing wear as well as friction in a tribological contact are energy, ­geometrical,
material, and environmental variables, as shown in Figure 13.1.
Energy variables are those that bring energy into the moving contact. They are normal load, tangen-
tial load, velocity, time, and temperature.
Geometrical variables are those that define the geometrical contours between solids and gases or
liquids and between different materials or phases within solids. They are typically macrogeometry (cur-
vatures, etc.), surface roughness, surface topography, particle size and shape, as well as microstructures
of composite materials.
Material variables define the properties of the materials involved. They define the elastic, plastic,
fracture, and viscous behavior of the materials. Here the materials include solids, liquids, and gases.
Wear Models 13-3

Variables Wear process Outcome

Energy Macromechanical
. Load mechanisms
. Velocity Micromechanical
. Time mechanisms
. Temperature Tribochemical
Geometry
. Macrogeometry mechanisms Wear
. Topography
. Loose particles Friction
.Material
Chemical composition
. Microstructure Changes in
. Hardness
. Elasticity geometry, material,
. Viscosity and environment
. Toughness
.Environment
Contamination
. Gases, liquids Nanophysical
. Radiation phenomena

FIGURE 13.1  In a tribological contact, there are a number of energy, geometrical, material, and environmental
variables influencing the wear process on macro-, micro-, and nanoscales, resulting in friction, wear, and changes
in the prevailing tribocontact variables influencing the future wear process.

The most common properties for solid materials are hardness, yield strength, elastic modulus, frac-
ture toughness, and Poisson’s ratio. For liquids and gases, the most common properties are, for exam-
ple, the chemical structure and viscosity, and material parameters such as pressure and temperature
dependence.
Environmental variables define the surrounding of the tribocontact. They are typically surrounding
gases, liquids, contaminations, radiation, etc.
These variables determine the wear process taking place and also the wear volume and wear rate.
The contact conditions change all the time during the wear process and, thus, some of the influencing
parameters will also change with time. Such changes are, for example, smoothening of the surfaces,
strain hardening, material transfer, ageing, and chemical reactions.
The changes in the contact that take place during the wear process can be observed and explained on
macro-, micro-, and nanolevels. Macroscale contact mechanisms are related to stress distribution and
deformations over the whole contact area and the total wear particle formation process and its dynam-
ics over time. These phenomena are typically of a size level of 1 𝜇m or more, up to milli- or centimeters.
Microscale contact mechanisms are related to stresses and strains at the asperity level, crack generation
and propagation, material liberation, and single particle formation. In typical engineering contacts,
these phenomena are of a size level of about 1 𝜇m or less. The chemical effects take place on a microscale
or less, while material transfer occurs both at macro- and microscales. Nanoscale contact phenomena
are related to the physics and chemistry of the material at atomic and molecular levels including effects
such as van der Waal’s forces and crystal structures. The nanoscale phenomena are the basis for wear
mechanisms on micro- and macrolevels [8].
In this chapter, we will focus on how the aforementioned complex wear processes can be simplified,
without the loss of any essential features, to such a level that they can be described by wear models or
wear equations. In the literature, there are numerous descriptions of wear mechanisms where the wear
model is left in a word form. We will concentrate on wear models, showing in quantitative terms the
interaction of various parameters and variables and how they influence wear in unlubricated condi-
tions. The models may be wear equations, wear maps showing the regions of validity of wear equations,
and computer programs like finite element methods (FEMs) linking different interactions to huge 3D
meshes representing the contact geometry.
13-4 Theory and Practice of Lubrication and Tribology

13.2  Wear Equations


Wear is a limitation to the proper functioning of machines and devices. Thus, there is a need to predict
wear. Over the years, many researchers have tried to formulate wear models and wear equations for wear
prediction. This has, however, turned out to be a very challenging task because of the complexity of the
wear process. To date we do not have any universally valid wear equation.
The large variety of different approaches to solve this problem is illustrated by a study carried out by
Meng and Ludema [9]. The study found that up to 1994 more than 300 wear models had been published and
of them 182 were formulated as wear equations, including more than 100 different variables and constants.
This shows the large number of possible variables influencing wear in different contact conditions.
The wear equations presented have typically been shown to be valid in some limited contact conditions,
but they all lack generality. They can be divided into three groups: (1) empirical equations formulated based
on experimental data, (2) contact-mechanics-based equations assuming some simple contact condition
relationships, and (3) material failure equations based on observations of the worn surfaces.
We shall present in the following only the most commonly used and most generic wear equations
related to common wear mechanisms developed mainly for metallic surfaces. Their derivation is
contact-mechanics based and can be found in textbooks such as Hutchings [4] and Bhushan [6]. In
real contact conditions, the prevailing wear mechanisms are very often a combination of the simplified
contact-mechanics-based wear mechanisms that are analyzed in the following.

13.2.1  Adhesive Wear


For contact conditions with one surface sliding unlubricated over another, the wear volume is com-
monly directly proportional to the normal load and the sliding distance and inversely proportional to
the hardness or shear strength of the softer surface. This relationship has been shown to be largely valid
especially for metallic surfaces with plastic-type contact conditions. It can be formulated as

V=
(k ⋅ L ⋅ s ) (13.1)
H

where
V is the wear volume
L is the applied load
s is the sliding distance
H is the hardness of the softer material of the mating surfaces

Even if shear strength is the more generic material property for plastic behavior, we use in the follow-
ing the closely related material property hardness, because it is in common use. The nondimensional
proportionality constant, k, also frequently called the wear coefficient, relates here to an adhesive plastic
type of sliding contact and can be experimentally determined for various sliding conditions. It would
typically include the effects of other variables that influence wear such as surface topography, contami-
nation, material microstructure, chemically formed surface layers, and probability of wear particle for-
mation. In mild wear contacts, k is typically in the range of 10−8 to 10−4, and in severe sliding conditions,
it is about 10−4 to 10−2.
The equation expresses the wear of the softer surface, but it can be used for the wear of the harder
surface if H is replaced by the hardness of the harder surface.
Of all wear equations, this is the most commonly used today and it is called the Archard’s equation.
It was first developed by Holm [10] for sliding electrical contacts and later Archard [11] presented the
theoretical basis for the expression, covering more generally typical adhesive sliding conditions. This
Wear Models 13-5

wear equation is in accordance with Amontons’ laws of friction as it indicates that wear is directly
proportional to load and to sliding distance but independent of apparent contact area and sliding
velocity.
In mild sliding conditions with materials of low modulus of elasticity or with very smooth surfaces,
the contact is dominated by elastic behavior and the wear equation can be formulated as

V=
(k ⋅ L ⋅ s ) (13.2)
⎡ Eʹ ⋅ s R 1 2 ⎤
⎣⎢
( ) ⎦⎥

where
V is the wear volume
L is the normal load
s is the sliding distance
E′ is the composite modulus of elasticity
σ and R are the composite standard deviation and the composite mean curvature of the surface
asperities, respectively

The nondimensional wear coefficient, k, relates now to an adhesive elastic type of sliding contact.
The wear depth is of interest from a design point of view as it influences tolerances and functional-
ity of tribological contacts. The wear depth rate can correspondingly be formulated for a plastic-type
contact as

d (k ⋅ p ⋅ v ) (13.3)
=
t H

where
d is the wear depth
t is the time
p is the apparent normal pressure
v is the sliding velocity
H is the hardness

For an elastic-type contact, the wear depth rate is


d
=
(k ⋅ p ⋅ v ) (13.4)
t ⎡Eʹ ⋅ s R 1 2 ⎤
⎢⎣ ( ) ⎥⎦

13.2.1.1  Wear Coefficient and Wear Rate


In this presentation, we use k as proportionality constant that would be different for different sliding
contact systems as it actually includes the effects of all the influencing variables, except those variables
considered to have major effects that have been separately written out in the equations. This constant has
also been used in another way. From Equation 13.1, the constant k can be written as wear volume mul-
tiplied by hardness and divided by load and sliding distance. In this form, k is called the wear coefficient
and is used as a universal quantitative parameter for wear to support design and material selection. This
is often more illustrative to use than just the wear volume of removed material because of the very dif-
ferent contact conditions. However, it should be noted that the wear coefficient corresponds specifically
to the surface to which the used hardness value is related.
13-6 Theory and Practice of Lubrication and Tribology

Wear rate K′ is a widely used practical and more general expression. K′ is the wear volume divided by
load and sliding distance and is normally given with the dimension (10−6 mm3/N m). It is recommended
that wear results from wear testing are given as the wear rate, whenever it is possible, since this is much
more general than to give it as wear volume or wear depth. An advantage to use wear rate instead of
coefficient of wear is that wear test results can be expressed as wear rate even if the hardness of the sur-
faces is not known. There is a clear physical argument to express wear in the form of wear rate because it
represents the quantity of material loss divided by the mechanical energy input into the contact.

13.2.2  Abrasive Wear


In abrasive contact conditions, hard countersurface asperities or hard particles deform the surface. The
wear rate is typically 2–3 orders of magnitude higher than for adhesive wear. The modification of the
surfaces and particles is a result of plastic deformation or material fracture. In the plastic deformation
case, the abrasive wear volume can be expressed as

V=
(a ⋅ b ⋅ L ⋅ s ) (13.5)
H

where
α is an asperity shape factor
β is a probability factor for material cuts that is related to the shear strength of the contact interface
and the mechanical properties of the wearing material

If we replace α · β with k, we end up in the Archard’s equation, which in general terms is valid in many
abrasive wear conditions. The wear coefficient for abrasive wear is typically in the range of 10−6 to 10−1.
This wear equation has experimentally turned out to be valid for a wide range of abrasive wear condi-
tions, both for two-body and three-body abrasive wear.
At higher loads and sharp topographies, brittle fracture of the hard asperities or particles occurs. In
sliding, where n asperities carry the load, the abrasive fracture related wear volume is

⎡ ⎛ E ⎞ 9 8⎤
⎢k ⋅ n ⎜⎝ H ⎟⎠ ⋅ L ⎥
V=⎣ ⎦ (13.6)
(12
Kc ⋅ H
58
)
where
k is a material-independent constant
Kc is the fracture toughness [12]

This relationship has been observed to be valid in experimental conditions, especially with ceramic
surfaces.

13.2.3  Fatigue Wear


The wear in rolling contacts is typically related to fatigue failures in the surface material. It is espe-
cially more common for rolling element bearings to express the wear in the form of fatigue life. This
is a reliability-based material life prediction method based on the Weibull distribution. The lifetime
L10 of rolling element bearings is that 90% of the bearing population can function without fatigue
failure:
Wear Models 13-7


(
L10 = k ⋅ C L ʹ ) (13.7)

where
C is the basic load capacity of the bearing
L′ is the equivalent bearing load
p′ is a bearing-type-specific load-life exponent

The basic load capacity C for a bearing is defined as the highest load at which 90% of the bearing
population can function without failure for 1 million inner-race revolutions under given running
conditions. For bearings, the proportionality constant k includes several design-related adjustment
factors such as bearing material, bearing processing, lubrication, speed effect, misalignment, and
contamination [13,14,51].

13.2.4  Impact Wear


Erosive wear is the most common form of impact wear, and it is a combination of abrasive and fatigue
wear. In erosive wear, the loading condition on the surface is because of continuous impacts from col-
liding particles that cause plastic deformation and changes in the surface strength. The wear is directly
proportional to the total mass, m, of the impacting particles, the square of the impacting velocity, v, and
inversely proportional to the surface hardness, H, as follows:

V=
(k ⋅ m ⋅ v )
2

(13.8)
(2 ⋅ H )

where k now is related to the impact angle, the shape and size of the particles, material ductility, fatigue
strength, and the proportion of the displaced material resulting in wear debris.

13.3  Wear Maps


A wear map, or a wear-mechanism map, is a diagram that shows the regimes of different wear mecha-
nisms, and, thus, it shows the limits for the validity of wear models describing each wear mechanism.
In a wear map, the regimes are shown as a function of two variables along the two coordinate axes. This
is, of course, a simplified and approximate way to present wear-mechanism regimes, because only two
variables are included and all the others are kept constant. If, however, the two variables are those that
have a dominating influence on wear in defined contact conditions, the wear map gives a good indica-
tion of how the wear mechanisms change and where the more uncontrolled transition regimes between
different wear mechanisms can be expected to appear.
Lim and Ashby [15] showed how wear maps can be developed for steel pairs using pin-on-disk con-
tact conditions. The map, which is shown in Figure 13.2, is constructed on the basis of a detailed
analysis of a large number of published wear test results combined with a theoretical analysis of the
wear mechanisms. It shows how the dominating wear mechanism changes with the two main energy
variables, load and speed, given as normalized pressure and normalized velocity, respectively. The
geometrical and environmental variables are considered constant as referred to a pin-on-disk device
in laboratory environment and the material variables are also considered constant as referred to steel
against steel contact. Levels for the normalized wear rates are given in the map as curves of constant
wear rates.
The wear mechanisms identified by Lim and Ashby in their wear map are as follows:
13-8 Theory and Practice of Lubrication and Tribology

Sliding velocity v (m/s)


–4 –2 2
10 10 1 10
10
Steel
Wear-mechanism map
Seizure pin-on-disk configuration

10–3
–1
–4
10–2(10 )
10–3 Melt 10–4
~

10
Normalized pressure F

10–3(10–5)
Wear –5
10–4(10–6)–7 10–6
–4 10–5(10 ) 10–6 10
10 –7
Mild- 10
oxidational Martensite –8
wear 10–6(10–8) formation 10
–5
10
Severe-
–3 Delamination
10 oxidational
wear –7
10 wear 10–9
–6 –8
10 (10 ) Mild to severe
wear transition
–8 –10
10 10
Ultra- –9
10 –11
mild –9 10
–5 wear 10
10 –2 2 4
10 1 10 10
~
Normalized velocity v

FIGURE 13.2  Wear-mechanism map for steel pairs sliding in pin-on-disk contact conditions and showing the
wear-mechanism regions as functions of normalized pressure and normalized sliding velocity. Wear values are
given as lines of constant normalized wear rate. (From Lim, S.C. and Ashby, M.F., Acta Metall., 35, 1, 1987. With
permission.)

• Seizure
• Melt-dominated wear
• Oxidation-dominated wear (mild and severe oxidational wear)
• Plasticity-dominated wear (including adhesion and delamination)
For each wear mechanism, they derive a normalized wear rate equation of the form

V ʹ = f ( pʹ, vʹ,T , material and geometrical variables ) (13.9)

where
V′ is the material volume lost per unit surface area and per unit sliding distance
p′ is the normalized nominal contact pressure divided by the contact hardness
v′ is the sliding velocity divided by the velocity of heat flow
T is temperature

The wear mechanisms Lim and Ashby used in their wear map are combinations of some of the basic
wear mechanisms described earlier. Seizure typically combines adhesion and abrasion, melt wear is
mainly adhesion, oxidational wear combines adhesion and chemical wear, while plasticity-dominated
wear combines adhesion and surface fatigue. The transition from one wear regime of mild wear to
another representing severe wear may be due to the load-dependent rupture of thin oxide surface films
or the velocity-dependent formation of hard martensitic surface layers. The transition from severe to
mild wear at the running in process may be due to distance-dependent martensite formation at higher
velocities, or a combination of work hardening, surface oxide films, and smoothening of original surface
roughness at lower velocities [16].
Wear Models 13-9

It is important to note that Lim and Ashby have limited their wear map to the rotational sliding pin-
on-disk contact conditions and to steel pairs. For this reason, this map is only valid in applications with
steel and with similar contact conditions. The development of similar wear maps for ceramics has been
reported by Hsu et al. [17], Adashi et al. [18], and Hsu and Shen [19]. Wear maps for alumina ceramics
have been developed by Wang et al. [20] and local yield maps of hard coatings by Diao et al. [21].
In wear maps for abrasive wear conditions with a sharp and hard steel tip sliding against a softer steel
surface, Hokkirigava and Kato [22] identify three wear mode regimes: microcutting, wedge formation,
and ploughing. They introduce as variables for the two coordinate axes the degree of penetration, repre-
sented by indentation depth per contact radius, and the shear strength at the contact interface.
Vingsbo et al. [23] developed wear maps for fretting wear to show the wear behavior in reciprocating
moving contacts. They used displacement amplitude and frequency or tangential force amplitude and
normal load as coordinate axes in their maps. Further developments of fretting wear maps were made
by Fouvry and Kapsa [24] and Zhou et al. [25]. Wear maps for uncoated high-speed steel-cutting tools
with feed rate and cutting speed as coordinate axes were published by Lim et al. [26].
In a review paper of wear-mechanism maps, Lim [27] covers the development of wear maps including
maps for metals, ceramics, metal–matrix composites, polymers, coatings, tools, fretting, erosion, and
maps for time-dependent wear transitions. He cites wear maps especially for coated surfaces related
to wear of abradable coatings on gas-turbine components and flank-wear as well as crater-wear maps
for TiN-coated high-speed steel inserts. Wear maps are useful as they enable implications of wear on
changes in design, material, and operating parameters to be assessed and allow sensible correlation to
be made between laboratory-based experimental investigations and observations in the field [28]. Wear
maps illustrate the validity regions and limits of models describing the related wear mechanisms.

13.4  Computerized Wear Models


We have discussed the numerous possible variables that influence wear and have seen that in the more
generic wear equations, only very few variables—typically the load, sliding distance, hardness, and elas-
tic modulus—have been included. All other possible variables have been excluded as separate variables
and are jointly included in the value k of the general proportionality constant. Still, we need a more accu-
rate description of the wear process to be able to predict wear. We need wear models showing the interac-
tion of all important variables in certain contact conditions and there are often some 5–10 of them. For
this we need to develop more precise wear equations, and four routes are considered in the following.
Empirical wear equations based on experimental data can be developed to become more precise. The
accuracy of test and measurement methods can be improved, the test conditions can be better controlled, the
number of experiments and test repetitions can be increased, and advanced statistical methods can be taken
into larger use. There is an impressive development going on in this field that gives us much useful informa-
tion about the wear process in the specific studied contact conditions. However, the new information and
understanding is very often lacking in generality and has not resulted in many new generic wear models.
Material failure equations based on observations from worn surfaces can be improved with the help
of the new and advanced surface microscopy and analysis methods available. However, it is very difficult
to construct accurate wear models based on observations only from the end result of the wear process,
and this is what the failed surface observations typically represent. The quantification of failure observa-
tions is a most difficult task.
Contact-mechanics-based equations can be further developed by analytical approaches, but this route
is slow due to the complexity of the contact conditions and the numerous variables involved. On the
other hand, the computerized-contact-mechanics-based approach is a very promising route. It repre-
sents the contact conditions that are the origin for the wear process and they can be modeled today very
accurately and quantified into equations by new computerized techniques [29].
There are three reasons why the computerized wear prediction methods can today handle the com-
plex and dynamic conditions in a tribological contact in a completely new way:
13-10 Theory and Practice of Lubrication and Tribology

1. There has been a rapid development in software technology, which today offers new mod-
eling and simulation tools that can very accurately describe even complex material
structures on several size and timescales and their response to dynamic loaded contact condi-
tions. Such approaches are, for example, advanced finite element method (FEM) techniques
including extended FEM and mesh-free techniques, discrete and continuum dislocation
dynamics (DDD and CDD), molecular dynamic simulation (MDS), and meso-mechanical
approaches.
2. There has been a very rapid development in computer hardware technology, which today offers
an increased computing capacity that can handle even 3D multiscale complex software programs
within a reasonable computing time frame.
3. New surface characterization methods make it possible to characterize more accurately the
surface properties and their changes at the surface and very close under the surface, down to
nanostructural level. These techniques are, for example, SEM and TEM (scanning and transmis-
sion electron microscopy), Raman spectroscopy, SIMS (secondary ion mass spectrometry), AES
(Auger electron spectroscopy), AFM (atomic force microscopy), and nanoindentation. These can
be used as good input values for computerized modeling and simulation.

The computerized modeling approach includes several steps combining tribocontact analysis, software
development, material characterization, and empirical testing. The methodology to develop an accurate
and generic wear equation can be characterized by the following steps, as shown in Figure 13.3:
1. Identify the relevant scale level for the tribocontact to be modeled.
2. Analyze the contact conditions and identify possible influencing variables and parameters.
3. Construct the contact and material model by, for example, FEM or MDS techniques.
4. Identify the relevant material properties of the material pair, for example, by macro-, micro-, or
nanolevel material characterization.
5. Carry out computer simulations for the chosen material pairs and contact conditions to show the
prevailing stresses, strains, and deformations.
6. Validate the model by comparing the simulations with results from empirical measurements in
similar contact conditions.

Identify relevant Contact analysis Model contact and


scale level material pair

Material Simulate deforma- Empirical model


characterization tions and stresses validation

Identify dominating
Parametric analysis
Calculate risk for variables and interact
gives the wear
crack initiation and ions
equation
crack growth

KE = √K 2I + K 2II + (1+v)K 2III


0.308 0.64
k R 0.047 σ0,su h
SWR =
0.9 E 0.470 E 0.86 µ1.05
Eco su ti

FIGURE 13.3  Route and steps for developing an SWR equation by computerized wear modeling and simulation
approach.
Wear Models 13-11

Wear

Adhesion + Abrasion + Fatigue +


material material material
detachment detachment detachment

Friction

Adhesion Ploughing Hysteresis

FIGURE 13.4  Basic mechanisms for wear are related to material detachment as a result of material breakage due
to adhesive pulling, abrasive deformation, and fatigue. The corresponding mechanisms resulting in friction are
adhesive, ploughing, and hysteresis resistance to motion.

7. Calculate the tendency for crack initiation and crack growth, resulting in material detachment
based on, for example, fracture mechanics for tough and brittle materials.
8. Identify the dominating variables with respect to critical local stress and strain peaks that may
exceed the material capacity.
9. Carry out parametric analysis of the dominating variables showing their weight and interaction
and formulate the surface wear resistance (SWR) equation, its limits of validity and accuracy.
The wear model shows how the material close to the surface responds to the loading conditions. This
response can be elastic, plastic, or fracture resulting in the formation of wear debris. The basic mecha-
nisms for material detachment from the surface are related to stretching the top material due to adhe-
sion to the countersurface, deformation by a hard countersurface, or fatigue failure due to repeated
loading, as shown in Figure 13.4. These basic mechanisms of material detachment are the same as those
resulting in the friction force.
Tribocorrosion is frequently listed as one basic wear mechanism, but it is on purpose not included
here. The reason is that chemical reactions taking place on the top surface certainly considerably influ-
ence wear, but their role is to cause deterioration or improvement of the material properties on the top
surface. After that, the detachment of material is actually caused by some of the three basic wear mecha-
nisms shown in Figure 13.4.
A wear resistance equation that has been developed according to the aforementioned methodology is
typically limited to some wear-mechanism conditions like severe abrasive wear, mild adhesive wear, impact
fatigue wear, etc. The equation covers a large range of material, geometrical, and energy variables. Simulations
can be done to explore the risk for surface failure and wear in various conditions. An equation like this
does not show the accumulated wear process nor predict the lifetime with regard to wear for components.
It quantifies the risk for initiation of surface failure resulting in wear and can be expressed as the surface
wear resistance (SWR). It is a tool that can be used by designers for improved surface material selection and
robust tribological design of machine components. In the following sections, we shall show as an example
how such an SWR equation was developed for a metallic surface covered with a thin hard coating, such as
the commonly used PVD (physical vapor deposition) or CVD (chemical vapor deposition) coated surfaces.

13.5  Simulation of Surface Stresses and Deformations


When a surface is loaded, it will respond by deforming elastically or plastically, or by breaking. The
deformation causes stresses and strain in the material and when these exceed the strength of the material
13-12 Theory and Practice of Lubrication and Tribology

it will break by cracking. The surface material response to loading was modeled by advanced FEM tech-
niques on a microscale level in a typical scratch test contact. A spherical diamond tip was made to slide
with increasing load over a flat steel surface coated with a few-micrometer-thick ceramic TiN coating,
as shown in Figure 13.5. Three effects are important to consider with regard to the failure behavior of
the surface. They are the ploughing of the spherical stylus, resulting in plastic deformation; the sliding
of the stylus on the coating, resulting in friction; and from this the pulling force on the coating behind
the contact, resulting in tensile stresses and fracture [30].
A detailed analysis of the contact conditions was needed for accurate contact condition modeling.
The complex stress field in the surface is a result of the following four effects (see Figure 13.5a):
1. Friction force. The friction force between the sliding stylus and the surface generates compressive
stresses in front of the stylus, originating from the pushing action, and tensile stresses behind the
tip originated from the pulling.
2. Geometrical changes. The elastic and plastic deformations are spherical indent, groove, and torus
shaped. They result in bending, stretching, and compressing of the coating. The stresses arising
are both compressive and tensile.

Residual stress Increasing


Bend, loading Bulk plasticity
torus Bend, concentration
groove Speed

Bend,
Bend,
Pull torus
indent
Bend,
torus Bend, Push
Plastic indent
deformation

Elastic + plastic
(a) deformation

R = 200 µm

(b) 1 µm 2 µm

Tip with r = 200 µm

Symmetry and
continuity
specified slice
different sizes
Linear-elastic
Elastic-plastic Slave
contact
1 2 3 region
12 mm

1 mm
(c) 0.1 mm

FIGURE 13.5  Stress field in the coated surface resulting from a sliding sphere is a result of four loading effects:
friction force, geometrical deformations, bulk plasticity concentration, and residual stresses. Illustration (a) shows
the loading effects with exaggerated dimensions and deformations and (b) with correct dimension interrelation-
ships. Illustration (c) is a schematic illustration of the 3D finite element mesh. The mesh sizing is in the range of
0.25–100 μm and the number of mesh nodes is about 190.000.
Wear Models 13-13

3. Bulk plasticity concentration. The spherical indentation causes the hard coating to deform elas-
tically in a circular wavelike shape and the substrate under the coating to deform plastically,
reaching its peak value at an angle of about 45° from the plane of symmetry in the plane of the
coating.
4. Residual stresses. It is very common for thin ceramic coatings, due to the deposition pro-
cess, to contain considerable compressive residual stresses. These are typically of the order of
0.5–4 GPa.
The FEM modeling can be carried out with a very complex finite element mesh; the more complex it
is, the more detailed is the information that can be achieved. The drawback with complex meshes is
that they result in very long computing times. So the optimal solution is always a balance between
the complexity of the mesh and minimizing of the computing time. The rapid development in com-
puting capacity and speed offers a possibility to run even relatively complex 3D FEM meshes in a
reasonable time. In line contacts, 2D FEM can be used successfully but, for example, in point con-
tacts where the stress maxima can be found outside the vertical plane of symmetry, 3D FEM analyses
are more beneficial.
The contact conditions of a 200 μm radius diamond tip sliding 10 mm with a load increasing from
5 to 50 N and indentation depth increasing from 0.5 to 3 μm, over a 2 μm thick TiN coating attached
with complete adhesion to a high-speed steel substrate were modeled by 3D FEM [30,31]. For the adhe-
sive component of the coefficient of friction, the empirically measured value 0.08 was used. The friction
originating from ploughing is integrated as plastic deformation in the FEM model. The stylus was mod-
eled as linear elastic but with the Young’s modulus of diamond it will behave practically as rigid. The
ceramic TiN coating was assumed not to deform plastically and thus modeled as linear elastic. The high-
speed steel substrate was modeled as elastic–plastic including the strain-hardening effect. A schematic
illustration of the FEM mesh is shown in Figure 13.5c.
Thin ceramic coated surfaces in tribocontacts typically fail by cracking and delamination. Thus, it
was of interest to study the fracture behavior of the hard coating. The deformations and stresses were
primarily analyzed and inferred with respect to first principal stress. This is defined as the largest stress
in tension in any direction and it is the major mechanism causing crack growth.
Topographical stress field maps showing the first principal stresses on the top of the coating and at
the subsurface symmetry plane in the direction of sliding are shown in Figure 13.6. Figure 13.6a shows
the stress pattern in the tip, in the coating, and in the steel substrate. Red color indicates high tensile
stresses and blue color high compressive stresses. The quantitative stress values are given in the scale
insert. Figure 13.6b shows more in detail the stress pattern in the coated surface at the location of the
trailing part of the contact in a case where there is an additional intermediate 500 nm hard ceramic bond
layer between the coating and the substrate. Figure 13.6c shows the stress pattern surrounding a vertical
crack going through the coating and ending at the substrate. The image is 2D along the symmetry plane.
The topographical stress field maps are very informative since they show both the location and the
quantitative level of stress peaks that possibly overload the material and result in fracture. A detailed
analysis of the topographical stress field and deformation maps resulted in the following generic conclu-
sions for steel surfaces with thin hard coatings [30–36]:
• The elastic modulus of the coating has a bigger effect on the stress peaks compared to the hardness
of the coated surface and the coating thickness.
• Tensile stress maxima were found at the groove edge typically at a distance of 1–2 times the con-
tact length behind the tail part of the contact and at this location the first cracks were observed
empirically; the stress pattern analysis could explain the crack generation mechanisms and both
the location and the direction of the angular cracks.
• The cracking of the coating most commonly starts from the top of the surface and develops verti-
cally down to the interface and then generally continues horizontally along the coating/substrate
interface, resulting in coating delamination and wear.
13-14 Theory and Practice of Lubrication and Tribology

S, max, principal
(Avg: 75%)
+3.769e-03
+2.500e-03
+1.958e-03
+1.417e-03
+8.750e-04
+3.333e-04
–2.083e-04
–7.500e-04
–1.292e-03
–1.833e-03
–2.375e-03
–2.917e-03
–3.458e-03
–4.000e-03
–6.807e-03

(a)

3000
2222.2
1444.4
666.67
111.11
888.89
1666.7
2444.4
y 3222.2
z 4000
x
(b)

FIGURE 13.6  Topographical stress-field maps showing the first principal stresses on the top of the coating and
at the symmetry plane intersection of a high-speed steel sample coated with a 2 μm thick TiN coating and loaded
by a sliding spherical diamond stylus. Sliding direction is from left to right; the sliding stylus is invisible in figures
(b) and (c). The values on the color scale are given in MPa. (a) Shows the stress field at 10 N. (From Holmberg, K.
et al., Wear, 264, 877, 2008. With permission.) (b) Is a close-up at the tail part of a similar contact but with a 500 nm
hard ceramic bond layer added. (From Holmberg, K. et al., Tribol. Int., 42, 137, 2009. With permission.) (c) Shows
the stress field around a preexisting 2 μm deep vertical crack going through the coating. (From Holmberg, K. et al.,
Int. J. Fracture, Submitted for publication, 2011. With permission.) See reference articles for colored pictures.
Wear Models 13-15

300
222.22
144.44
66.667
–11.111
–88.889
–166.67
–244.44
y –322.22
z x –400

(c)

FIGURE 13.6  (continued)

• Preexisting residual compressive stresses in the coating are relaxed in the material deformation
process taking place below the sliding contact, and they do not affect the coating cracking behind
the contact in and at the formed groove.
• The thickness of a bond layer between the coating and the substrate does not have much effect on
the stress peaks, but its elastic modulus is very important.

Always when a model has been developed, it is important to validate it empirically to be sure that it
accurately reflects the real conditions that have been modeled. The validation can be done on various
levels. The first level is to ensure that the material model deforms and behaves similarly as the real mate-
rial. This can typically be done by simple indentation or scratch testing. In the case described earlier,
the location and direction of the first cracks to appear were observed in a real scratch test and this was
compared with the stress field generated in the simulations with identical variable values. It showed that
there was a high tensile stress peak exactly at the location of the first crack and the stress filed analysis
could predict the direction of the crack. This was taken as a first validation of the model. A second level
of validation would be to compare the wear resistance of a model-optimized coated surface with the
wear results from some tribotest such as a pin-on-disk tester.

13.6  Parametric Analysis of a Wear Contact


At this stage, we have a model that has been validated to produce very accurate information about
local stresses and strains in the material in and in the vicinity of the studied tribocontact, down to the
microlevel and even close to the nanolevel. Based on this information, conventional fracture mechan-
ics is used to calculate the crack growth and fracture toughness of the coated surface. The fracture
analysis is carried out using a mixed-mode approach, where all stress intensity factor (SIF) components
and the mixed-mode state are presented using an equivalent SIF. The equivalent SIF components are
determined using a weight-function-based methodology, where the weight functions are determined
numerically on the basis of an interpolation to a 3D boundary element analysis of coating cracking.
The results of such analyses can then be utilized to describe the surface behavior in terms of fracture
mechanical performance. Thus, the methodology relies in the phenomenon measuring up against the
fracture mechanical crack driving force (or SIF), and system resistance to crack generation scales with
the increase and decrease of the computed fracture mechanical quantities.
13-16 Theory and Practice of Lubrication and Tribology

Based on the tribological contact analysis and previous experience from empirical measurements,
seven variables were identified as major parameters and chosen to be included in a parametric analysis.
These were the tip radius, elastic modulus and yield strength of the substrate, elastic modulus of the
coating and the tip, coating thickness, and coefficient of friction.
The parametric analysis was carried out by forming an SQLite database of all numerical computa-
tions that have been carried out with thin hard coatings. The database can be considered to be a repre-
sentation of an n-dimensional (in this case a 7D) surface. The various dependencies between variables
affecting the fracture mechanical response of the surface can now be investigated and queried. By add-
ing a nonlinear regression module, the dependencies can be represented in a quantified form as pre-
sented in the following. The database is complemented by new results as they become available, and new
dependencies are and can be investigated.
The parametric analysis computations for one single hard coating on an elastic–plastic substrate
including strain hardening, like for steel, show the influence of the chosen variables on tensile stress
peaks that may exceed the strength of the material and result in breakage. Here we introduce the SWR as
a parameter for the wear related property of the surface defined as the resistance of the surface to surface
crack growth, which is the initial process of wear. The SWR concept is derived by expressing the critical
condition (crack initiation) by use of the effective SIF as outlined above for characterizing the fracture
mechanical conditions of the coating-substrate system, and quantifying by means of FE results how this
state is affected by a variation in any or several of the parameters. As such SWR ~ fracture toughness
and a critical measure of stress for crack initiation. The SWR can be used to investigate how the critical
condition is affected by the various identified parameters. The computations were carried out and the
database resulted in a simplified equation for the SWR with seven variables (SWR7) in the form

⎡⎣k ⋅ R0.047 ⋅ s 00,.su


308
⋅ h0.64 ⎤⎦
SWR 7 = (13.10)
⎡⎣ Eco0.9 ⋅ Esu0.470 ⋅ Eti0.86 ⋅ m1.05 ⎤⎦

where
k is a proportionality constant for dimensional purposes
R is the tip radius
σ 0,su is the yield strength of the substrate
Eco, Esu, and Eti are the elastic modules of the coating, substrate, and tip, respectively
h is the coating thickness
μ is the coefficient of friction

The SWR7 can be understood as a cut of the numerical database with respect to studied parameters and
their respective influences. The expression can be presented in a dimensional form and the cross terms
dependencies evaluated, but for clarity and condensed presentation it is here given in absolute terms.
We can conclude from the exponent values in the equation that the coefficient of friction and the elas-
tic modulus of the coating have a major effect on the stresses generated under load in the coated system
and, thus, on the SWR. Coating thickness and contacting tip elastic modulus are also important, while
both the elastic properties and hardness of the substrate influence less than one traditionally would
expect. The influence of the tip radius reflecting the loading conditions is minor.
The SWR equation (13.10) can be used to compare the wear resistance of various coated surfaces when
looking for an optimal solution. As an example, the typical values for the variables

• R = 200 μm
•  σ 0,su = 2000 MPa
• Eco = 300 GPa
• Esu = 200 GPa
Wear Models 13-17

• Eti = 1000 GPa
• h = 2 μm
•  μ = 0.1
would result in SWR7 = 1.4 · 106 mm2/N2 · k. The unit for the SWR depends on how many and which vari-
ables are included in the equation. Commonly, one would apply it in relative terms and the coefficient
“k” is used to specify that this is the intended manner.
It is also important to reflect on the generality of Equation 13.10. As described earlier, this equation
can predict the SWR in a contact system

• With one hard elastic coating


• On an elasto-plastic substrate
• Loaded by a spherical elastic sliding countersurface
• Where the coating is thin and resulting in a material response from the coating/substrate surface
system

This contact system is typically represented by industrial steel components covered by a thin hard
ceramic coating in a severely loaded sliding contact, where surface and coating cracking turn out to be
detrimental.
The validity range of the equation is fairly large, covering most applications of the previously defined
contact system. The variable validity ranges are

• R = 10–1 0000 μm
•  σ 0,su = 200–3 000 MPa
• Eco = 40–500 GPa
• Esu = 100–500 GPa
• Eti = 100–1 000 GPa
• h = 0.5–10 μm
•  μ = 0–2

13.7  Component Wear Models


Wear models and wear equations have a real value when they can be applied to practical machine com-
ponents and help in wear prediction and robust design with respect to wear. In the generic wear models
that we have so far discussed are the contact conditions more or less simplified compared to many com-
ponents in use in products and machines.
There have been numerous attempts to formulate wear and lifetime equations for tribological com-
ponents such as bearings, gears, seals, bushings, joints, sliders, tools, piston rings, cam and tappets,
etc. Kragelsky et al. [37] analyzed the wear of several machine components both in sliding and abrasive
conditions and developed equations for wear prediction. Component-level wear analyses and models
are found in some textbooks [1,5,38]. The models have normally been shown to be in agreement with
empirical measurements in the specific conditions for which they have been developed, but they lack a
more general validity and are, thus, not in wider use.
There are at least three approaches that can be observed among these component wear models. One
is to develop a wear equation starting from the Archard’s equation, where the wear is proportional to
load and sliding distance and inversely proportional to hardness. Then some additional variables often
related to geometrical and functional parameters are added to the equation and these are tuned to cor-
respond with empirical measurements.
Another approach that is typical for high-speed contacts of adhesive type and often with softer mate-
rials, such as bushings and seals, is to correlate the wear with the contact pressure and sliding speed,
also called the pv-factor. This is a good approach when the surface temperature is important because the
13-18 Theory and Practice of Lubrication and Tribology

speed and pressure have a major influence on the generation of heat in the contact. The wear can then be
shown as curves or regions in pv-diagrams.
A third approach is to consider the energy balance in the contact and assume that the amount of wear
is proportional with the work done, which is expressed as friction force times sliding distance [39]. The
validity of this very relevant statement is sometimes limited by the complex character of component
contacts.
There is one brilliant exception where a fatigue wear lifetime equation has been developed for a
machine component and is widely used. It is the bearing life formula (Equation 13.7) for rolling ele-
ment bearings such as ball and roller bearings. The formula was originally developed by Lundberg and
Palmgren [40,41]. It has over the years been complemented according to the elasto-hydrodynamic lubri-
cation theory [42,43] and has been adjusted to more precisely fit additional functional and environmen-
tal parameters [13,14,51].
One new and interesting approach to modeling component wear was published by Hedgadekatte
et al. [44,45]. They modeled the sliding contact by FEM and used a global incremental wear model
(GIWM) that considered also the wear history. Global indicates that the wear model considers
macro-quantities like average contact pressure over the whole contact. Incremental indicates that
the model is continuously updating the contact pressure according to the new worn topography at
various intervals of sliding. The phenomenological wear model they use is the Archard model. The
GIWM approach has successfully been used for micro-machine applications like lobe pumps and
gears [46,47].
It is very challenging to also include the dynamics of wear debris in wear models. Especially in fret-
ting wear contacts, wear particles are typically entrapped in the contact zone; they influence wear by
forming a load-carrying platform and undergo a series of processes before they are eventually removed.
Ding et al. [48] have presented a finite-element-based approach to simulating the effects of debris on
fretting wear. The debris accumulated in the contact interface are modeled as a layer structure with
the mechanical properties being described by an anisotropic elastic–plastic material model. Migration
of wear particles within the surfaces determines the evolution of the thickness of the debris layer and
its dynamics. This results in redistribution of the contact pressure and slip over the contact region for
prediction of debris effect on wear damage.

13.8  Conclusions
It is a very challenging task to generate a wear model and especially to develop it quantitatively to a
wear equation that describes the wear process satisfactorily and can be used for wear prediction and
tribological design. This is due to the variety of wear mechanisms, the great number of potential param-
eters that influence wear, the dynamics of the wear process, and the various scales with different wear
mechanisms involved.
Some generic wear equations for adhesive, abrasive, erosive, and fatigue wear have been developed
and can be used within well-defined limits. Still, they lack in accuracy and general validity. Only for
rolling contact bearings there is a good wear model and a wear equation with well-developed and accu-
rate statistical basis that is in use in industrial design worldwide. Wear maps are very useful as they
illustrate the regions of validity of wear mechanisms and wear models.
The computerized wear modeling approach has great potential due to new possibilities offered by
the rapid development in hardware capacity, in software programming methods, and in micro- and
nanoscale surface material characterization techniques. Computer simulations of stresses, strains, and
deformations from macro- to nanoscale give very useful information about how the surface material
behaves under tribological loading, how it breaks into wear debris, and how the wear process develops.
Parametric analysis of results from tribological contact simulations is a new technique that can be used
for developing wear models and wear equations for defined contact loading conditions. It is a tool for
wear prediction, tribological surface optimization, and robust design.
Wear Models 13-19

Acknowledgments
The financial support of Tekes, the Finnish Technology Agency; Taiho Kogyo Tribology Research
Foundation, Japan; Savcor Coatings, Finland; and the VTT Technical Research Centre of Finland is
gratefully acknowledged. The research was started as part of the EC-supported COST532 Triboscience
and Tribotechnology Action. It is now continuing as a Finnish joint industrial consortium strategic
research action coordinated by FIMECC Ltd within the program on breakthrough materials called
DEMAPP in projects wear, friction, and energy. We gratefully acknowledge the financial support of
Tekes and the participating companies.

References
1. M.B. Peterson and W.O. Winer (eds.), Wear Control Handbook, ASME, New York, 1980.
2. I.V. Kragelskii and E.A. Marchenko, Wear of machine components, J. Lubr. Tech. Trans. ASME, 104,
1–8, 1982.
3. N.P. Suh, Tribophysics, Prentice-Hall, Englewood Cliffs, NJ, 1986.
4. I.M. Hutchings, Tribology: Friction and Wear of Engineering Materials, Butterworth Heinemann,
Oxford, U.K., 1992.
5. R.G. Bayer, Mechanical Wear Prediction and Prevention, Marcel Dekker, New York, 1994.
6. B. Bhushan, Principles and Applications of Tribology, John Wiley & Sons, New York, 1999.
7. G.W. Stachowiak, Wear: Materials, Mechanisms and Practice, John Wiley & Sons, Chichester, West
Sussex, U.K., 2005.
8. K. Holmberg and A. Matthews, Coatings Tribology: Properties, Mechanisms, Techniques and
Applications in Surface Engineering, Elsevier Tribology and Interface Engineering Series No. 56,
Elsevier, Amsterdam, the Netherlands, 2009.
9. H.C. Meng and K.C. Ludema, Wear models and predictive equations: Their form and content, Wear,
181–183, 443–457, 1995.
10. R. Holm, Electric Contacts, Hugo Gebers Förlag, Stockholm, Sweden, 1946.
11. J.F. Archard, Contact and rubbing of flat surfaces, J. Appl. Phys., 25, 981–988, 1953.
12. A.G. Evans and D.B. Marshall, Wear mechanisms in ceramics. In: D.A. Rigney (ed.), Fundamentals of
Friction and Wear of Materials, ASM, Metals Park, OH, 1981.
13. E.V. Zaretsky, A. Palmgren revisited : A basis for bearing life prediction, Lubr. Eng. J. STLE, February,
18–24, 1997.
14. E. Ioannides, G. Bergling, and A. Gabelli, An Analytical Formulation for the Life of Rolling Bearings,
Acta Polytechnica Scandinavia, Mechanical Engineering Series, The Finnish Academy of Technology,
Espoo, Finland, No. 137, 1999.
15. S.C. Lim and M.F. Ashby, Wear-mechanism maps, Acta Metall., 35, 1–24, 1987.
16. S.C. Lim, M.F. Ashby, and J.H. Brunton, Wear-rate transitions and their relationship to wear mecha-
nisms, Acta Metall., 35, 1343–1348, 1987.
17. S.M. Hsu, D.S. Lim, Y.S. Wang, and R.G. Munro, Ceramics wear maps: Concept and method develop-
ment, Lubr. Eng. J. STLE, 47, 49–54, 1991.
18. K. Adashi, K. Kato, and N. Chen, Wear maps of ceramics, Wear, 203–204, 1997.
19. S.M. Hsu and M.C. Shen, Wear mapping of materials. In: G.W. Stachowiak (ed.), Wear: Materials,
Mechanisms and Practice, John Wiley & Sons, Chichester, West Sussex, U.K., pp. 367–423, 2005.
20. Y.S. Wang, S.M. Hsu, and R.G. Murno, Ceramics wear maps: Alumina, Lubr. Eng. J. STLE, 47, 63–69, 1991.
21. D. Diao, K. Kato, and K. Hayashi, The local yield map of hard coating under sliding contact. In:
D. Dowson et al. (eds.), Thin Films in Tribology, Elsevier Science Publishers, Amsterdam, the
Netherlands, pp. 419–427, 1993.
22. K. Hokkirogawa and K. Kato, An experimental and theoretical investigation of ploughing, cutting
and wedge formation during abrasive wear, Tribol. Int., 21, 51–57, 1988.
13-20 Theory and Practice of Lubrication and Tribology

23. O. Vingsbo, M. Odfalk, and N. Shen, Fretting maps and fretting behaviour of some F.C.C. metal
alloys, Wear, 138, 153–167, 1990.
24. S. Fouvry and P. Kapsa, An energy description of hard coating wear mechanisms, Surf. Coat.
Technol., 138, 141–148, 2001.
25. Z.R. Zhou, K. Nakazawa, M.H. Zhu, N. Maruyama, P. Kapsa, and L. Vincent, Progress in fretting
maps, Tribol. Int., 39, 1068–1073, 2006.
26. S.C. Lim, S.H. Lee, Y.B. Liu, and K.H.W. Seah, Wear maps for uncoated high-speed steel cutting tools,
Wear, 170, 137–144, 1993.
27. S.C. Lim, Recent developments in wear-mechanism maps, Tribol. Int., 31, 87–97, 1998.
28. J. Williams, Wear and wear particles: Some fundamentals, Tribol. Int., 38, 863–870, 2005.
29. K. Holmberg, E. Turunen, H. Ronkainen, A. Laukkanen, and K. Wallin, Computer modelling and
simulation approach to developing wear resistant materials, Tribol. Finn. J. Tribol., 28, 38–43, 2009.
30. K. Holmberg, A. Laukkanen, H. Ronkainen, K. Wallin, and S. Varjus, A model for stresses, crack
generation and fracture toughness calculation in scratched TiN-coated steel surfaces, Wear, 254,
278–291, 2003.
31. K. Holmberg, A. Laukkanen, H. Ronkainen, K. Wallin, S. Varjus, and J. Koskinen, Tribological contact
analysis of a rigid ball sliding on a hard coated surface–Part I: Modelling stresses and strains, Surf.
Coat. Technol., 200, 3793–3809, 2006.
32. K. Holmberg, A. Laukkanen, H. Ronkainen, K. Wallin, S. Varjus, and J. Koskinen, Tribological con-
tact analysis of a rigid ball sliding on a hard coated surface–Part II: Material deformations, influence
of coating thickness and Young’s modulus, Surf. Coat. Technol., 200, 3810–3823, 2006.
33. K. Holmberg, A. Laukkanen, H. Ronkainen, and K. Wallin, Surface stresses in coated steel surfaces:
Influence of a bond layer on surface fracture, Tribol. Int., 42, 137–148, 2009.
34. K. Holmberg, H. Ronkainen, A. Laukkanen, K. Wallin, S. Hogmark, S. Jacobson, U. Wiklund,
R.M. Souza, and P. Ståhle, Residual stresses in TiN, DLC and MoS2 coated surfaces with regard to
their tribological fracture behaviour, Wear, 267, 2142–2156, 2009.
35. A. Laukkanen, K. Holmberg, J. Koskinen, H. Ronkainen, K. Wallin, and S. Varjus, Tribological con-
tact analysis of a rigid ball sliding on a hard coated surface–Part III: Fracture toughness calculation
and influence of residual stresses, Surf. Coat. Technol., 200, 3824–3844, 2006.
36. A. Laukkanen, K. Holmberg, H. Ronkainen, and K. Wallin, Cohesive zone modeling of initiation and
propagation of multiple cracks in hard thin surface coatings, J. ASTM Int., 8(1), 1–21, 2010.
37. I.V. Kragelsky, M.N. Dobychin, V.S. Kombalov, Friction and Wear Calculation Methods, Pergamon
Press, Oxford, U.K., 1982.
38. G.E. Totten (ed.), Handbook of Lubrication and Tribology, CRC Taylor & Francis, London, U.K.,
2006.
39. A. Ramalho and J.C. Miranda, The relation between wear and dissipated energy in sliding systems,
Wear, 260, 361–367, 2006.
40. G. Lundberg and A. Palmgren, Dynamic capacity of rolling bearings, Acta Polytechnica, Mechanical
Engineering Series, Royal Swedish Academy of Engineering Sciences, Stockholm, Sweden, 1(3), 7,
1947.
41. G. Lundberg and A. Palmgren, Dynamic capacity of roller bearings, Acta Polytechnica, Mechanical
Engineering Series, Royal Swedish Academy of Engineering Sciences, Stockholm, Sweden, 2(4), 96,
1952.
42. D. Dowson and G.R. Higginson, Elasto-Hydrodynamic Lubrication: The Fundamentals of Roller and
Gear Lubrication, Pergamon Press, Oxford, U.K., 1966.
43. B.J. Hamrock and D. Dowson, Ball Bearing Lubrication: The Elastohydrodynamics of Elliptical
Contacts, John Wiley & Sons, New York, 1981.
44. V. Hegadekatte, N. Huber, and O. Kraft, Modeling and simulation of wear in a pin on disc tribom-
eter, Tribol. Lett., 24, 51–60, 2006.
Wear Models 13-21

45. V. Hegadekatte, S. Kurzenhäuser, N. Huber, and O. Kraft, A predictive modeling scheme for wear in
tribometers, Tribol. Int., 41, 1020–1031, 2008.
46. V. Hegadekatte, J. Hilgert, O. Kraft, and N. Huber, Multi time scale simulations for wear prediction
in micro-gears, Wear, 268, 316–324, 2010.
47. Y.J. Chen and N. Huber, Transient simulation of wear in a lobe pump using the Wear Processor. In:
M. Hadfield, C.A. Brebbia, and J. Seabra (eds.), Tribology and Design, WIT Press, Southampton, U.K.,
pp. 66, 49–60, 2010.
48. J. Ding, I.R. McColl, S.B. Leen, and P.H. Shipway, A finite element based approach to simulating the
effects of debris on fretting wear, Wear, 263, 481–491, 2007.
49. K. Holmberg, H. Ronkainen, A. Laukkanen, K. Wallin, A. Erdemir, and O. Eryilmaz, Tribological
analysis of TiN and DLC coated contacts by 3D FEM modelling and stress simulation, Wear, 264,
877–884, 2008.
50. K. Holmberg, A. Laukkanen, H. Ronkainen, and K. Wallin, Simulation of coating adhesion failure in
a pre-existing crack field, Int. J. Fracture, Submitted for publication, 2011.
51. E.V. Zaretsky, Rolling bearing life prediction, theory, and application. In: G. K. Nikas (ed.) Recent
Developments in Wear Prevention, Friction and Lubrication, Research Signpost, Trivandrum, India,
pp. 46–136, 2010.
14
Lubrication
14.1 Introduction..................................................................................... 14-1
14.2 Brief History of Lubrication........................................................... 14-1
14.3 Types and Regimes of Lubrication................................................ 14-3
14.4 Squeeze Film Lubrication............................................................... 14-5
14.5 Rough Surface Considerations......................................................14-6
14.6 Surface Texturing............................................................................14-8
14.7 Nanoparticle Laden Lubricants................................................... 14-10
Robert L. Jackson 14.8 Chapter Overview.......................................................................... 14-11
Auburn University References................................................................................................... 14-11

14.1  Introduction
This section first introduces the concept of lubrication and some of the associated nomenclature. Then it
discusses the different regimes and types of lubrication, the effects of surface roughness on lubrication,
and finally some techniques for improving lubrication, including surface texturing. Readers should be
aware that this serves mostly as an introduction and many of the topics are covered in more detail in
other parts of the book or in the cited references.
Although lubrication is used in many different applications and under many different conditions, its
actual function is often not well understood. Lubrication is the usage of a third material between two
surfaces. These two surfaces could be in relative sliding motion or they could be stationary. Lubrication
is used to reduce or control friction and wear between these surfaces. The lubricant or third material
between the two surfaces is usually either a fluid, a solid, or a semisolid (grease, gel, paste) with a low
shear strength. Reducing the shear stress between two surfaces has the effect of reducing the friction and
wear. The thickness of the lubricant layer does not have to be substantial as even a single molecular layer
can be effective at reducing the friction and wear (may be classified as a type of boundary film lubrica-
tion). However, a thick layer of lubricant can be more reliable and resilient. If lubricant is removed from
between surfaces due to wear or squeezing, it can sometimes be replenished by a supply inlet or by being
imbedded in the bulk surface itself. Surface dimples and textures can also act as an effective reservoir
for lubricants and will be discussed in more detail in Section 14.6.

14.2  Brief History of Lubrication


Lubrication and lubricants have been used by mankind for many thousands of years with archeological
evidence of it dating back to the Egyptian Empire (Dowson 1998). The typically used lubricants were
mostly animal fats and oils derived from creatures such as whales up until the arrival of the industrial
revolution and the discovery of reservoirs of mineral oil. The beginning of the industrial usage of mined
or mineral oil is often marked by Edwin Drake’s discovery of oil in 1859 in Oil Creek, Pennsylvania

14-1
14-2 Theory and Practice of Lubrication and Tribology

FIGURE 14.1  Photograph of early oil derricks at Oil Creek, Pennsylvania.

(Wicks 2009) (Figure 14.1). George Bissell actually first postulated the value of oil when examining oil
naturally flowing into a creek in 1851. Bissell processed it into lamp fuel and lubricants for machines.
This discovery eventually led to the foundation of John Rockefeller’s Standard Oil Company, which
eventually was broken up by the federal government for antitrust violations into 34 smaller companies.
We now see remnants of Standard Oil with remaining offspring such as ExxonMobil, British Petroleum
(BP), and Chevron. Interestingly, mineral oils at first did not perform well because they did not possess
the natural lubricity of animal-derived lubricants, but later additives improved their performance.
Synthetic lubricants have also since taken their place alongside conventional mineral oils (LePera
2000). Synthetic lubricants or oils are defined as substances that are made by artificially creating or
chemically modifying hydrocarbons. Since most conventional mineral oil lubricants are modified
chemically in someway, the divide between the two is often not well defined. Originally, synthetic lubri-
cants were used in aircraft engine applications because they could perform at high temperatures.
Initially, it appears that most believed that lubrication was just acting as a barrier between surfaces
that reduced the friction and wear (i.e., boundary lubrication), and not generating significant load-
carrying pressure. However, Beauchamp Tower (1885), somewhat by accident, found that lubrication
under some conditions can generate a pressure large enough to actually carry the entire load pressing
two surfaces together and therefore separate them completely. Since the surfaces are no longer in con-
tact, the friction and wear are usually very low. This is defined as the hydrodynamic or full-film regime
of lubrication. This occurred while Tower was testing the application of oil to a journal bearing via radial
holes rather than just submersion of the open sides. When he tried to plug these holes, the corks would
be pushed out by the pressure building in the lubricant. He then inserted a gage to measure the pressures
generated by the oil in the journal bearing.
At nearly the same time, Nikolay Petrov (1883) made significant advancements of his own. He is most
well known for his equation for predicting the frictional torque of a lubricated hydrodynamic journal
bearing. However, Petrov also made contributions in the area of lubricant selection for particular appli-
cations based on its viscosity.
However, a complete analytical theory explaining the mechanics of the hydrodynamic bearings came
only a few years later when Osborne Reynolds (1885) (shown in Figure 14.2) simplified the Navier–
Stokes fluid dynamics equations to a single second-order differential equation by considering that a
lubricating film is very thin relative to the size of the bearing surface area. This equation is
Lubrication 14-3

FIGURE 14.2  Osborne Reynolds.

∂ ⎛ rh3 ∂p ⎞ ∂ ⎛ rh3 ∂p ⎞ ∂ ( rh ) ∂ ( rh )
+ = 6U + 12 (14.1)
∂x ⎜⎝ m ∂x ⎟⎠ ∂y ⎜⎝ m ∂y ⎟⎠ ∂x ∂t

where
x and y are the coordinate axes in the plane of the slide surfaces
h is the film height
p is the fluid pressure
μ is the viscosity
ρ is the fluid density
U is the relative velocity between the surfaces in the x-direction
t is time

Note that the left side of the equation is associated with the pressure-driven flow effects, the first term
on the right side is the velocity-driven effects that generate substantial load-carrying capacity in hydro-
dynamic bearings, and the second term on the right is the squeeze film term which considers changes in
film thickness with time or relative normal motion of the surfaces.
Equation 14.1, usually named for Reynolds, can also be analytically solved for several simple cases
and so provided an accessible analytical tool for engineers even before the arrival of the computer.
The arrival of the computer age accelerated the progress in this field. Raimondi and Boyd (1955) used
advanced numerical algorithms to solve the hydrodynamic journal bearing problem for cases which
could not be considered analytically. This also enabled the modeling and solution of the solid and fluid
coupling in elastohydrodynamic problems (Petrusevich 1951). Raimondi and Boyd (1957) also provided
a methodology for the design of hydrostatic bearings which separate surfaces using an externally sup-
plied pressurized lubricant. Today researcher’s employ the ever increasingly available computational
resources by modeling lubricants at the molecular scale using molecular dynamics methods.

14.3  Types and Regimes of Lubrication


Typically, lubrication is classified into three different regimes: boundary, mixed or partial, and full-
film lubrication. Hamrock (1994) lists four regimes, including elastohydrodynamic lubrication (EHL),
which some may label as a subcategory of full-film hydrodynamic lubrication. In the full-film lubrica-
tion regime, the film thickness is much larger than the surface roughness, and so they are completely
14-4 Theory and Practice of Lubrication and Tribology

separated which results in essentially no asperity contacts. Therefore, the wear and friction are low, only
resulting from the shearing of the lubricant. EHL occurs in several cases, but is defined by when the
fluid pressure deforms the surfaces significantly. When extremely high interface pressures develop in
hard concentrated contacts (such as rolling element bearings), causing the viscosity of the lubricant to
increase by several orders of magnitude and for the surfaces to deform significantly under these pres-
sures. Soft materials such as the rubber in lip seals and cartilage can deform significantly at much lower
pressures and also be in the EHL regime.
Boundary lubrication, which was first defined by Hardy and Doubleday (1922), occurs for high loads
when the hydrodynamic or hydrostatic film thickness cannot separate the surfaces except for a chemi-
cally small film residing from an applied lubricant. In some cases, this film may be only a monolayer of
molecules thick. In this regime, there is significant asperity contact and therefore the friction and wear
are usually relatively high. However, it should be noted that the friction and wear are usually lower than
without a lubricant between the two surfaces (i.e., dry contact).
Mixed or partial lubrication occurs when the film thickness is not sufficient enough to completely
separate the peaks or asperities of the two surfaces and marks the realm between boundary and full-film
lubrication. This typically results in an increase in the friction and wear. As the film thickness decreases
further, the friction will increase further until a saturation point is often reached. Once this occurs, the
region defined as boundary lubrication is reached.
The Stribeck curve (see Figure 14.3) is often used to graphically define these various regions of lubri-
cation (hydrodynamic, mixed, and boundary) and often attributed to Richard Stribeck (1901) or Mayo
Hersey (1914). A Stribeck curve is a plot of the coefficient of friction versus the product of the bearing
linear speed and viscosity normalized by the average bearing pressure. Using the traditional notation
for the Stribeck curve, the equation form of this relation is

⎛ ZN ⎞
f eff = f eff ⎜ (14.2)
⎝ P ⎟⎠

where
feff is the effective coefficient of friction
Z is the dynamic viscosity of fluid (N·s/m2)
N is the rotational speed (rev/s)
P is the average bearing pressure (N/m2)

Boundary
lubrication
Mixed
Coefficient of friction

0.1 (Severe lubrication


wear) (EHL)
(Moderate
wear)
Full-film lubrication
0.01 (negligible wear)

0.001
0
ZN/P

FIGURE 14.3  Stribeck plot with lubrication regimes marked. Typical values of the friction coefficients are shown.
Lubrication 14-5

U
h

Surface asperity contact

FIGURE 14.4  Depiction of boundary lubrication between two rough surfaces. U is the relative velocity between
the surfaces. (Please note the exaggerated roughness due to disproportional magnification in the direction perpen-
dicular to the surface.)

This curve is frequently used to qualitatively identify the transitions from boundary lubrication to
mixed lubrication, and to hydrodynamic lubrication. On the right side of the Stribeck curve, the rota-
tional speed and viscosity are high enough to generate enough hydrodynamic lift to overcome the aver-
age bearing pressure and separate the surfaces with a thin film of fluid. When a film separates the
surfaces, little or no contact occurs between the bumps or asperities on the surfaces. Thus, when a
bearing is operating on the right side of the Stribeck curve, the friction and wear are low. However, if
the bearing operates under heavier loads, with lower viscosity and/or lower speed, it will operate on the
left side of the Stribeck curve. Moving left on the Stribeck indicates a loss of hydrodynamic lift and so
eventually some of the applied load is carried by contact between the surfaces (see Figure 14.4). This
contact occurs at the peaks or asperities of the surface. As more asperities come into contact, the fric-
tion and wear will increase drastically. The surfaces are then in the boundary lubrication regime, where
the lubricant mainly benefits the contacts by providing a coating that reduces the adhesion and shear
strength of the asperity contacts.

14.4  Squeeze Film Lubrication


The Stribeck curve is generally used for surfaces in relative sliding motion (in a direction parallel or
tangent to the surfaces); however, lubricant film thicknesses and load-carrying capacity can also be gen-
erated by relative normal motion, or a compressive and tensile loading of the lubricant (see Figure 14.5).
This type of lubrication is known as a squeeze film. Actually, it may play a large role in generating the
load-carrying capacity of the cartilage in human and animal articular joints.
A significant amount of work has characterized the squeeze film effect (hydrodynamic pressure cre-
ated due to perpendicular motion between two surfaces), with most of the focus on squeeze film damping.

Normal velocity

Generated pressure

Lubricating film

Stationary surface

FIGURE 14.5  Schematic of squeeze film motion that results in bearing lift.
14-6 Theory and Practice of Lubrication and Tribology

Griffen et al. (1966) suggested that a squeeze film between two parallel plates provides viscous damping
action over a certain frequency range and provides an expression for cutoff frequency below which the gas
film acts as a damper. Blech (1983) studied the squeeze film damping and divided the squeeze film force into
damping and spring force components. According to him, the damping force is maximum approximately at
the cutoff frequency and excitation above or below the cutoff frequency causes a decrease in damping force.
Etsion (1980) analyzed squeeze effects in liquid-lubricated radial face seals and obtained squeeze damping
coefficients. Etsion (1980) and Green and Etsion (1983) calculated dynamic damping and stiffness coeffi-
cients of the fluid films in mechanical face seals considering squeeze film effects along with hydrostatic and
hydrodynamic effects. Research into the effect of squeeze film damping is also currently prominent in the
design of microelectromechanical systems (MEMS) and microstructures (Starr 1990; Pan et al. 1998; Bao
et al. 2003; Nayfeh and Younis 2004; Kim et al. 2011).
The largely underutilized squeeze film effect has a potential for use as a means to lubricate surfaces
and to create squeeze film bearings (Minikes and Bucher 2003; Mahajan et al. 2008). Applications of
such squeeze film bearings can be and are sometimes used in read/write heads in hard disk drives, man-
ufacturing processes, vibrating machinery, low-speed applications, and hydrodynamic bearing start-up
and shutdown. When the two bearing surfaces in relative motion approach each other, positive film
pressure is developed between them and viscous resistance to flow at the sides of the bearing increases
(Hamrock 1991). Thus, the fluid contained between the two bearing surfaces acts as a lubricating film.
Incompressible squeeze film bearings were discussed by Hamrock (1991) and also analytical expressions
for the pressure profiles were derived.

14.5  Rough Surface Considerations


In regions near contact where asperities between surfaces come in close proximity, the asperities can
influence the lubrication flow (see Figures 14.4 and 14.6). Here the lubrication regime is known as
boundary lubrication; in other words there is only a thin film of lubricant separating the surfaces, and
the microtopography of the surfaces greatly affects the flow of the lubricant. Even before contact occurs
between the asperity peaks, they can influence and obstruct the flow of lubricant (see Figure 14.6).
It should be noted that when a lubricant film thickness is of the same magnitude as the surface rough-
ness (average or RMS), the lubricant flow is generally considered obstructed and it is possible that there
is contact.
Patir and Cheng (1978, 1979) were the first to formulate these asperity flow effects between two three-
dimensional surfaces. These flow effects were taken into account in the form of flow factors which were
incorporated into a modified form of the Reynolds equation given by

Asperity flow
obstructions
Sliding
direction

FIGURE 14.6  Overhead depiction of a lubricant film flow being obstructed by surface asperities or roughness.
Lubrication 14-7

∂ ⎛ h3 ∂p ⎞ ∂ ⎛ h3 ∂p ⎞ U1 + U 2 ∂hT U1 − U 2 ∂fs ∂hT


fx + fy = + s + (14.3)
∂X ⎜⎝ 12 m ∂x ⎟⎠ ∂y ⎜⎝ 12 m ∂y ⎟⎠ 2 ∂x 2 ∂x ∂t

where
ϕx, ϕy, and ϕs are flow factors which describe the affect the asperities have on the lubricant flow in
different directions
h‾T is the average gap
h is the film height between the surfaces
p is the lubricant pressure
μ is the lubricant viscosity
U1 and U2 are the relative velocities of the two surfaces sliding against each other in the x-direction

Patir and Cheng (1978, 1979) give formulations for the flow factors as functions of surface roughness,
asperity orientation (longitudinal or transverse), and film height. Patir and Cheng’s formulations for the
flow factors when the roughness is isotropic are

fx = fy 1 − 0.90e −.56(h /s ) (14.4)


⎡⎛ s ⎞ 2 ⎛ s ⎞ 2 ⎤
fs = ⎢⎜ 1 ⎟ − ⎜ 2 ⎟ ⎥ Φs (14.5)
⎢⎣⎝ s ⎠ ⎝ s ⎠ ⎥⎦

where

Φs = 1.899 * (h/s )0.98 e −0.92h /s + 0.05h /s for h/s ≤ 5 (14.6)


Φs = 1.126 e −0.25h /s for h/s > 5 (14.7)


where
σ1 and σ 2 are the RMS roughness of each surface
s = s 12 + s 22

Also, h‾T is known as the average gap and is defined as

hT =
∫ (h + d)G(d)dd (14.8)
−h

where
δ is the combined roughness of the two surfaces given by δ = δ1 + δ 2
G is the Gaussian height distribution of the surface

Additional details of the model can be found in the original works.


In regions where the surface roughness is significantly less than the film thickness (h/σ > 6), the
lubrication model is ascribed to the full-film regime by the modified Reynolds equation. This is
because that as h/σ increases ϕx and ϕy approach the value of one and ϕs approaches zero so that
Equation 14.3 reduces to the unmodified version of the Reynolds equation (Equation 14.1). Equations
14-8 Theory and Practice of Lubrication and Tribology

14.3 through 14.8 are usually solved numerically using a technique such as the finite difference or
finite element method (Kazama and Yamaguchi 1993; Ruan et al. 1997; Jackson and Green 2006;
Wang et al. 2006).

14.6  Surface Texturing


Surface texturing is a methodology for improving the performance and load-carrying capacity of lubri-
cated surfaces by, in essence, controlling the texture or roughness of the surface. Surface texturing
appears to be gaining usage by industry in such areas as pistons and mechanical seals. An example laser-
textured surface is shown in Figure 14.7. This concept may have been arrived at by noticing that in some
cases rough surfaces can also generate load-carrying capacity by the small-scale features actually acting
as very small hydrodynamic bearings (Hamilton et al. 1966; Lebeck 1987). For instance, it has also been
noticed that flat-faced thrust washer bearings appear to be able generate load-carrying capacity, which
may be due to a combined effect of thermal deformation and surface roughness (Cameron and Wood
1958; Jackson and Green 2001, 2008; Yu and Sadeghi 2002).
Recently, extensive research (Etsion and Burstein 1996; Baroud et al. 2000; Ronen et al. 2001; Wang
et al. 2002; Brizmer et al. 2003; Hsu 2004) has been made to optimize these textured surfaces and to
characterize their performance. This led to many advances like microtextured surfaces (Wang et al.
2002) and laser surface texturing (LST) (Etsion and Burstein 1996; Ronen et al. 2001; Brizmer et al.
2003), which actually improve lubrication by expanding the hydrodynamic lubrication regime of the
Stribeck curve, especially to the left (see Figure 14.3). The two main mechanisms of lubrication enhance-
ment that surface textures provide are (1) they act as reservoirs of lubricant that can be tapped if the
external supply is interrupted and (2) the individual dimples act like hydrodynamic bearings. Surface
textures can be created using many different techniques including vibrorolling (Bulatov et al. 1997),
reactive ion etching, abrasive jet machining, lithography (Stephens et al. 2004), and LST (Etsion 2005).
The geometry of the textures or individual dimples, such as density, shape, and depth, must be opti-
mized for a given set of operating conditions. Many applications that possess nearly parallel sliding con-
tacts such as mechanical seals and pistons show improvement in friction when surface textures are used.
Recent work has also shown that the textures at the nanoscale and for very thin lubricating films may

FIGURE 14.7  A picture of laser surface texturing. (Courtesy of I. Etsion, SurfTech, Mississauga, Ontario,
Canada.)
Lubrication 14-9

Hydrodynamic load carrying pressure

Flexible membrane structures

Bulk material
(a)
Sample surface Covered grooves

(b)

FIGURE 14.8  Schematic (a) and top view picture (b) of a prototype of self-adapting surfaces using deformable
bridges to cause controlled deformation.

also produce load-carrying capacity and lower friction coefficients than larger-scale textures due to the
shear thinning effect and elimination of a nanoscale stick–slip effect (Capozza et al. 2008; Jackson 2010).
Recently, researchers have developed surface textures that may actually adapt for different conditions
by changing shape when the fluid pressure increases (Jackson 2005; Duvvuru et al. 2009; Fesanghary
and Khonsari 2010). This is a form of EHL, but used in adapting textures that are specifically designed
to behave in this manner. One concept is to have a textured geometry with deep grooves that are covered
by a deformable membrane (see Figure 14.8), but other techniques may also be possible. The concept of
an adapting surface may also be linked to the soft articular cartilage of synovial joints in humans and
animals (Tysoe and Spencer 2006; Bougherara et al. 2007; Singh et al. 2009). Self-adapting surfaces may
also already exist unintentionally in deformable lip seals or in the elastohydrodynamic contacts of mul-
tiple asperity contacts (Lebeck 1987; Shi and Salant 2000, 2001; Harp and Salant 2002; Jacobson 2002).
In addition, similar adaptive mechanisms have long been used in air-lubricated foil bearings and
analyzed numerically and experimentally (Heshmat et al. 1982; Peng and Carpino 1993; Arakere 1996;
Lez et al. 2007). However, these bearings use the mechanisms at a fairly large scale in comparison to the
microscale self-adapting surfaces. In addition, the foil bearings use gas as the lubricant, which is not sus-
ceptible to cavitation. Hardie and Ettles (1988) analyzed the performance of a foil slider bearing (instead
of the conventional journal bearing). One of the major findings was that the slider foil bearings could
provide a near-constant film thickness under dynamic situations (which is a similar finding to Jackson
(2005) for self-adapting step bearings).
Similar to foil bearings, San Andrés (2006) investigated the performance of hybrid tilted gas bear-
ings which used deflection to improve performance. Salant (1994) also researched a similar concept by
altering a typical mechanical seal to include a deformable or compliant surface (also similar to a foil
14-10 Theory and Practice of Lubrication and Tribology

bearing). Salant’s work found that the stiffness of the compliant surface mechanical seal could be sig-
nificantly better than conventional mechanical seals.

14.7  Nanoparticle Laden Lubricants


Advances in material fabrication and formulation techniques have initiated substantial interest in the
area of nanoparticle laden lubricants, sometimes referred to as nanofluids, but many in the field prefer
the term colloids. These particles are fabricated at the nanoscale, and materials such as silica, copper,
silver (see Figure 14.9), diamond, and graphite, among others, have been considered based on their
individual properties. The addition of nanoparticles to paraffinic oil can have a marked effect on the
lubricating properties (Tao et al. 1996; Hwang et al. 2006), and also the thermal conductivity and the
phase transitions. Research has suggested that there are a few different mechanisms for nanoparticles to
improve the performance of the oil, including (1) viscosity, (2) thermal properties and thermal stability,
(3) the particles could roll between the surfaces and act as “nano ball bearings,” and (4) the particles
could mend worn surfaces by adhering to them (see Figure 14.10). Copper particles, for instance, appear
to provide exceptional mending properties. In addition, paraffinic oils with nanoparticles could provide
more “resistance” to being completely squeezed out from between surfaces and still have similar bulk
properties (because the particles lodge themselves between the surfaces effectively creating lubricant
reservoirs similar to the surface textures).
A few researchers have investigated the effect of CuO particles on lubricating oils and have shown
promising results. Sajith et al. (2007) found that although CuO and Al 2O3 particles had only a slight
effect on the fluid viscosity, the flash temperature of the fluid decreased significantly, thus improving
the fluid stability without affecting its lubricating capabilities. Wu et al. (2007) found that the addition
of CuO nanoparticles reduced friction and improved the wear resistance. Copper particles covered
by an oxide layer have also been tested by Liu et al. (2004). Copper particles with organic shells have
been tested and shown to be effective at reducing friction and wear by bonding to the surfaces (Li et al.
2006). There are some promising results, but there is still significant room for further fundamental
characterization of the effect of the wide array of available nanoparticles, each with its own possible
advantages, in oil and other lubricants. One of the main problems of nanoparticle laden lubricants is
that the suspended particles do not remain in the lubricant forever and may eventually drop out, result-
ing in a short shelf life.

20

15
Volume (%)

10

20 nm 0
0 10 20 30 40 50 60 70 80 90 100
(a) (b) Diameter (nm)

FIGURE 14.9  TEM images of prepared silver nanoparticles with a mean diameter of approximately 20 nm (a) and
the histogram of the size distributions (b). (Excerpted with permission from Zhang, M. et al., STLE Tribol. Trans.,
52(2), 157, 2009.)
Lubrication 14-11

2.5 µm
–1 µm

45 nm
4.8 µm 30 nm

(a)

4 µm

2 µm 135 nm
m
0.4 µ (c)

(b)

FIGURE 14.10  Pictures showing a lubricated surface worn without nanoparticles (a), with nanoparticles (b),
and a close SEM picture of a mended surface covered by particles (c). (With kind permission from Springer
Science+Business Media: Tribol. Lett., Investigation of the mending effect and mechanism of copper nanoparticles
on a tribologically stressed surface, 17(4), 2004, 961–966, Liu, G. et al., Figure 5.)

14.8  Chapter Overview


This section has provided a brief history of the lubrication field, along with an introduction to its con-
cepts and a few insights to current developments. In addition, the remaining sections cover a wide
range of topics relating to lubrication and mostly hydrodynamic or full-film lubrication, but sections on
mixed and boundary lubrication are also provided. This includes the assumptions and derivation of the
aforementioned Reynolds equation. Then it is explained how to apply Reynolds equation to a number
of different geometries and motions such as cylindrical journal bearings, thrust bearings, and squeeze
film bearings. The different cases of liquid-lubricated and gas-lubricated bearings are also addressed.
Section 14.3 discussed the effect that the surface deformations induced by the fluid pressure may have
on bearing performance (i.e., EHL). Cases which are exceptions to the assumptions made when deriving
Reynolds equation, such as for inertial effects and turbulence, are also considered.

References
Arakere, N. K. (1996). Analysis of foil journal bearings with backing springs. Tribol. Trans. 39(1): 208–214.
Bao, M., H. Yang et al. (2003). Modified Reynolds equation and analytical analysis of squeeze-film air
damping of perforated structures. J. Micromech. Microeng. 13(6): 795–800.
Baroud, C., I. Busch-Vishniac et al. (2000). Induced micro-variations in hydrodynamic bearings. J. Tribol.
Trans. ASME 122(3): 585–589.
Blech, J. J. (1983). On isothermal squeeze films. J. Lubr. Technol. Trans. ASME 105: 615–620.
Bougherara, H., M. Bureau et al. (2007). Design of a biomimetic polymer–composite hip prosthesis.
J. Biomed. Mater. Res. Part A 82(1): 27–40.
Brizmer, V., Y. Kligerman et al. (2003). A laser surface textured parallel thrust bearing. Tribol. Trans. 46(3):
397–403.
14-12 Theory and Practice of Lubrication and Tribology

Bulatov, V. P., V. A. Krasny et al. (1997). Basics of machining methods to yield wear and fretting resistive
surfaces, having regular roughness patterns. Wear 208: 132–137.
Cameron, A. and W. L. Wood (1958). Parallel surface thrust bearing. ASLE Trans. 1: 254–258.
Capozza, R., A. Fasolino et al. (2008). Lubricated friction on nanopatterned surfaces via molecular dynam-
ics simulations. Phys. Rev. B 77: 235432.
Dowson, D. (1998). History of Tribology. New York: Wiley.
Duvvuru, R. S., R. L. Jackson et al. (2009). Self-adapting microscale surface grooves for hydrodynamic
lubrication. Tribol. Trans. 52(1): 1–11.
Etsion, I. (1980). Squeeze effects in radial face seals. J. Lubr. Technol. Trans. ASME 102(2): 145–152.
Etsion, I. (2005). State of the art in laser surface texturing. J. Tribol. Trans. ASME 127(1): 248.
Etsion, I. and L. Burstein (1996). A model for mechanical seals with regular microsurface structure. Tribol.
Trans. 39(3): 677–683.
Fesanghary, M. and M. M. Khonsari (2010). On self-adaptive surface grooves. Tribol. Trans. 53(6): 871–880.
Green, I. and I. Etsion (1983). Fluid film dynamic coefficients in mechanical face seals. J. Lubr. Technol. 1–5.
Griffin, W. S., H. H. Richardson et al. (1966). A study of fluid squeeze-film damping. ASME J. Basic Eng.
451–456.
Hamilton, D. B., J. A. Walowit et al. (1966). A theory of lubrication by microasperities. ASME J. Basic Eng.
88(1): 177–185.
Hamrock, B. J. (1991). Fundamentals of Fluid Film Lubrication. Washington, DC: NASA Office of
Management, Scientific and Technical Information Program [Supt. of Docs., U.S. G.P.O., distributor].
Hamrock, B. J. (1994). Fundamentals of Fluid Film Lubrication. New York: McGraw-Hill.
Hardie, C. E. and C. M. M. Ettles (1988). The analysis of a self-acting flexible foil slider bearing. ASME J.
Tribol. 110(1): 134–138.
Hardy, W. B. and I. Doubleday (1922). Boundary lubrication—The paraffin series. Proc. R. Soc. Lond. A
100: 25–39.
Harp, S. R. and R. F. Salant (2002). Inter-asperity cavitation and global cavitation in seals: An average flow
analysis. Tribol. Int. 35(2): 113–121.
Hersey, M. D. (1914). The laws of lubrication of horizontal journal bearings. J. Wash. Acad. Sci. 4: 542–552.
Heshmat, H., W. Shapiro et al. (1982). Development of foil journal bearings for high load capacity and high
speed whirl stability. J. Lubr. Technol. Trans. ASME 104(2): 149–156.
Hsu, S. M. (2004). Nano-lubrication: Concept and design. Tribol. Int. 37(7): 537.
Hwang, Y., H. S. Park et al. (2006). Thermal conductivity and lubrication characteristics of nanofluids. Curr.
Appl. Phys. 6(1): e67–e71.
Jackson, R. L. (2005). Self adapting mechanical step bearings for variations in load. Tribol. Lett. 20(1): 11–20.
Jackson, R. L. (2010). A scale dependent simulation of liquid lubricated textured surfaces. ASME J. Tribol.
132(2): 022001.
Jackson, R. L. and I. Green (2001). Study of the tribological behavior of a thrust washer bearing. Tribol.
Trans. 44(3): 504–508.
Jackson, R. L. and I. Green (2006). The behavior of thrust washer bearings considering mixed lubrication
and asperity contact. Tribol. Trans. 49(2): 233–247.
Jackson, R. L. and I. Green (2008). The thermoelastic behavior of thrust washer bearings considering
boundary lubrication, asperity contact and thermoviscous effects. Tribol. Trans. 51(1): 19–32.
Jacobson, B. (2002). Nano-meter film rheology and asperity lubrication. J. Tribol. Trans. ASME 124(3): 545.
Kazama, T. and A. Yamaguchi (1993). Application of a mixed lubrication model for hydrostatic thrust
bearings of hydraulic equipment. ASME J. Tribol. 115: 686–691.
Kim, S.-J., R. Dean et al. (2011). An investigation of the damping effects of various gas environments on a
vibratory MEMS device. Tribol. Int. 44(2): 125–133.
Lebeck, A. O. (1987). Parallel sliding load support in the mixed friction regime. Part 1—The experimental
data; Part 2—Evaluation of the mechanisms. J. Tribol. Trans. ASME 109(1): 189–205.
LePera, M. E. (2000). Synthetic automotive engine oils—A brief history. Lubr. World. 10(1).
Lubrication 14-13

Lez, L. S., M. Arghir et al. (2007). Static and dynamic characterization of a bump-type foil bearing struc-
ture. ASME J. Tribol. 129(1): 75–83.
Li, B., X. Wang et al. (2006). Tribochemistry and antiwear mechanism of organic–inorganic nanoparticles
as lubricant additives. Tribol. Lett. 22(1): 79–84.
Liu, G., X. Li et al. (2004). Investigation of the mending effect and mechanism of copper nanoparticles on
a tribologically stressed surface. Tribol. Lett. 17(4): 961–966.
Mahajan, M., R. L. Jackson et al. (2008). Experimental and analytical investigation of a dynamic gas squeeze
film bearing including asperity contact effects. Tribol. Trans. 51(1): 57–67.
Minikes, A. and I. Bucher (2003). Coupled dynamics of a squeeze-film levitated mass and a vibrating piezo-
electric disc: Numerical analysis and experimental study. Journal of Sound and Vibration. 263(2): 426.
Nayfeh, A. H. and M. I. Younis (2004). A new approach to the modelling and simulation of flexible micro-
structures under the effect of squeeze film damping. J. Micromech. Microeng. 14(2): 170–181.
Pan, F., J. Kubby et al. (1998). Squeeze film damping effect on the dynamic response of a MEMS torsion
mirror. J. Micromech. Microeng. 8(3): 20–208.
Patir, N. and H. S. Cheng (1978). An average flow model for determining effects of three-dimensional
roughness on partial hydrodynamic lubrication. ASME J. Tribol. 100: 12–17.
Patir, N. and H. S. Cheng (1979). Application of average flow model to lubrication between rough sliding
surfaces. ASME J. Tribol. 101: 220–230.
Peng, J.-P. and M. Carpino (1993). Calculation of stiffness and damping coefficients for elastically sup-
ported gas foil bearings. ASME J. Tribol. 115(1): 20–27.
Petrov, N. P. (1883). Friction in machines and the effect of the lubricant. Inzh. Zh. St. Petersburg 1–4:
71–140, 227–279, 377–436, 535–564.
Petrusevich, A. I. (1951). Fundamental conclusions from the contact-hydrodynamic theory of lubrication.
Izv. Akad. Nauk, SSSR, Otd. Tekh. Nauk. 2: 209–233.
Raimondi, A. A. and J. Boyd (1955). Applying bearing theory to the analysis and design of pad-type bear-
ings. ASME Trans. 77(3): 287–309.
Raimondi, A. A. and J. Boyd (1957). An analysis of orifice and capillary compensated hydrostatic journal
bearings. Lubr. Eng. 13(1): 28–37.
Reynolds, O. (1885). On the theory of lubrication and its application to Mr. Beauchamp tower’s experi-
ments, including an experimental determination of viscosity of olive oil. Philos. Trans. R. Soc. Lond.
Ser. A 177: 157–234.
Ronen, A., I. Etsion et al. (2001). Friction-reducing surface texturing in reciprocating automotive compo-
nents. Tribol. Trans. 44(3): 359–366.
Ruan, B., R. F. Salant et al. (1997). Mixed lubrication model of liquid/gas mechanical face seals. Tribol.
Trans. 40(4): 647–657.
Sajith, V., M. D. Mohiddeen et al. (2007). An investigation of the effect of addition of nanoparticles on the
properties of lubricating oils. Proceedings of the 2007 ASME–JSME Thermal Engineering Summer
Heat Transfer Conference. British Columbia, Canada: Vancouver.
Salant, R. F. (1994). Analysis of a hydrostatic gas seal with a compliant face. Proceedings of 14th International
Conference on Fluid Sealing, BHR Group Ltd., Firenze, Italy, pp. 385–395.
San Andrés, L. (2006). Hybrid flexure pivot-tilting pad gas bearings: Analysis and experimental validation.
ASME J. Tribol. 128(3): 551–558.
Shi, F. and R. F. Salant (2000). Mixed soft elastohydrodynamic lubrication model with interasperity cavita-
tion and surface shear deformation. J. Tribol. Trans. ASME 122(1): 308–316.
Shi, F. and R. F. Salant (2001). Numerical study of a rotary lip seal with a quasi-random sealing surface.
J. Tribol. Trans. ASME 123(3): 517–524.
Singh, R. A., E.-S. Yoon et al. (2009). Biomimetics: The science of imitating nature. Tribol. Lubr. Technol.
65(2): 40–47.
Starr, J. B. (1990). Squeeze-film damping in solid-state accelerometers. Technical Digest, Solid State Sensor
and Actuator Workshop, June 4–7, Hilton Head Island, South Caroline, pp. 44–47.
14-14 Theory and Practice of Lubrication and Tribology

Stephens, L. S., R. Siripuram et al. (2004). Deterministic micro asperities on bearings and seals using a
modified LIGA process. J. Eng. Gas Turbines Power 126(1): 147.
Stribeck, R. (1901). Kugellager für beliebige Belastungen (ball bearings for any stress). Z. Verein. Deutsch.
Ing. 45.
Tao, X., Z. Jiazheng et al. (1996). The ball-bearing effect of diamond nanoparticles as an oil additive. J. Phys.
D Appl. Phys. 29: 2932–2937.
Tower, B. (1885). Second report on friction experiments (experiments on the oil pressure in a bearing).
Proc. Instn. Mech. Engrs 58–70.
Tysoe, W. and N. Spencer (2006). Nature’s soft touch. Tribol. Lubr. Technol. 62(4): 56.
Wang, X., K. Kato et al. (2002). The lubrication effect of micro-pits on parallel sliding faces of SiC in water.
Tribol. Trans. 45(3): 294–301.
Wang, Y., Q. J. Wang et al. (2006). Development of a set of Stribeck curves for conformal contacts of rough
surfaces. Tribol. Trans. 49(4): 526–535.
Wicks, F. (2009). The oil age. Mech. Eng. 131(8): 42–45.
Wu, Y. Y., W. C. Tsui et al. (2007). Experimental analysis of tribological properties of lubricating oils with
nanoparticle additives. Wear 262: 819–825.
Yu, T. H. and F. Sadeghi (2002). Thermal effects in thrust washer lubrication. J. Tribol. Trans. ASME 124(1):
166–177.
Zhang, M., X. Wang et al. (2009). Investigation of electrical contact resistance of Ag nanoparticles as addi-
tives added to PEG 300. STLE Tribol. Trans. 52(2): 157–164.
15
Hydrostatic Lubrication
Nomenclature.............................................................................................. 15-1
15.1 Introduction..................................................................................... 15-2
15.2 History............................................................................................... 15-3
15.3 Principles of Operation................................................................... 15-3
15.4 Fundamental Relationships...........................................................15-4
Pressure Distribution  •  Flow Rate  •  Load Capacity  •  Frictional
Torque  •  Power Loss
15.5 Flow Control.....................................................................................15-9
Capillary Compensation  •  Orifice Compensation  •  Constant Flow
Compensation  •  Pressure Sensing Valve Compensation
15.6 Design Procedure: Multi-Recess Journal Bearing.................... 15-11
15.7 Hybrid Bearings............................................................................. 15-13
Gregory J. 15.8 Optimization.................................................................................. 15-16
Kostrzewsky 15.9 Summary......................................................................................... 15-18
Cummins Inc. References.................................................................................................. 15-18

Nomenclature
Aeff Effective bearing friction area, m2
Af Friction area for one pad, m2
Ao Bearing land area for one pad, m2
Ar Recess area for one pad, m2
a Axial flow land width, m
B′ Flow shape factor for thin land bearings
b Circumferential flow land width, m
Cd Orifice discharge coefficient
cp Specific heat capacity, J/N°C
c1, c2 Integration constants
D Bearing diameter, m
dc Capillary tube bore diameter, m
f Shear force on a fluid element, N
Hf Friction or viscous power loss, Pa·s (or N·m/s)
Hp Pumping power loss, Pa·s (or N·m/s)
Ht Total power loss, Pa·s (or N·m/s)
h Film thickness, m
ho Design bearing clearance, lubricant film thickness, m
hr Depth of bearing recess, m
h′min Minimum running clearance, m
K Dimensionless power ratio

15-1
15-2 Theory and Practice of Lubrication and Tribology

kc Capillary flow factor, m3


ko Orifice flow factor, m4/s·N1/2
k s Dimensionless stiffness factor for bearings with slots
L Bearing length, m
lc Length of capillary tube, m
N Shaft speed, rpm
n Number of bearing pads
pL Liftoff recess pressure, Pa (or N/m2)
pr Recess pressure, Pa (or N/m2)
ps Minimum supply pressure, Pa (or N/m2)
Q Total volumetric flow rate, m3/s
Qc Volumetric flow (laminar) in a capillary tube, m3/s
Qo Volumetric flow in an orifice, m3/s
q Volumetric flow rate per unit width, m2/s
r Radius, m
ri Inner radius; recess radius, m
ro Outer radius, m
T Bearing torque, N·m
t Frictional torque on a fluid element, N·m
ΔT Adiabatic temperature rise, °C
u Surface velocity, m/s
vc Average fluid velocity in the capillary, m/s
W Normal or applied bearing load, N
W′ Maximum applied bearing load, N
β Pressure ratio, β = pr/ps
γ Circumferential flow factor for journal bearings
η Dynamic or absolute viscosity, Pa·s (or N·s/m2)
θ Coordinate in cylindrical polar coordinates, pad angle
λ′ Dimensionless stiffness
λo Actual stiffness, N/m
ρ Mass density, kg/m3
ω Angular velocity, rad/s

15.1  Introduction
Hydrodynamic lubrication develops a fluid film due to shearing action in the lubricant caused by
relative motion between two surfaces. In contrast, hydrostatic lubrication achieves a fluid film
via the supply of lubricant under pressure from an external source. There is no physical contact
during start-up and shutdown. Hydrostatic bearings are typically used when an application requires
full film lubrication and is unable to develop an adequate film hydrodynamically. Since hydrostatic
lubrication does not require relative motion between the bearing surfaces to develop a fluid film
pressure to support a load, this makes hydrostatic bearings particularly well suited to very low
speed, reversing, or even zero speed operation. Other advantages of hydrostatic bearings over
other bearing types include low friction, high stiffness, long life, and high load capacity. The main
disadvantage of hydrostatic bearings is the cost and complexity of an external pressurized lubricant
supply system.
Successful applications include slow-moving or heavily loaded machines such as rolling mills, large
machine tools, rolling-road chassis dynamometers, telescopes, large radar antennas, as well as high
speed or lightly loaded machines such as gyroscopes or precision balances.
Hydrostatic Lubrication 15-3

15.2  History
The origin of hydrostatic bearings can be traced to a Frenchman, L.D. Girard, who in 1851 used high-
pressure water to feed a hydrostatic bearing in a railway propulsion system [1,2]. Since then, a num-
ber of publications and patents have been issued, with research interest increasing notably soon after
World War II.
Several authors deserve special mention due to their significant contributions. In 1947, Fuller [3] and,
in 1963, Rippel [4] each presented a series of articles addressing basic concepts and design equations for
a range of hydrostatic bearing applications.
A comprehensive review of hydrostatic bearing design was conducted in 1971 by Rowe and
O’Donoghue [5]. The authors cite 103 references considered to be landmarks of bearing design and
analysis. With a bias toward design information, the topics covered include pad coefficients, various
means of lubricant supply and control, optimization, dynamic effects, thermal effects, and consider-
ation of tolerances.
A sense of the maturity of the topic developed by 1974 as evidenced by a series of three articles by
Stout and Rowe [6–8] who address practical considerations on design for manufacture, including toler-
ance variation.
In 1992, Bassani and Piccigallo [9] authored a book on the theory and design of a wide variety of
hydrostatic bearings. The authors provide a number of worked examples along with an extensive list of
references.
With the basics of hydrostatic bearing design well documented, subsequent research reflects a move
from approximate solutions to numerical analyses as well as exploration of the many permutations
available given the wide range of bearing geometries and lubricant supply and control methods.

15.3  Principles of Operation


The basic principles used in all hydrostatic bearings are graphically illustrated in Figure 15.1. The main
design parameters for operation are lubricant flow rate, load capacity, and power loss. The main design
input parameters include bearing geometry, lubricant viscosity and density, supply pressure, applied
load, and minimum oil film thickness requirements.
The formation of a fluid film in a hydrostatic bearing system is illustrated in Figure 15.1 via six operat-
ing conditions. Prior to operation, Figure 15.1(a), the supply flow, supply pressure, and recess pressure
are zero. The bearing pad and the thrust runner are in contact. Shortly after the supply pump is turned
on, Figure 15.1(b), the supply pressure ps becomes > 0. Because the bearing has not “lifted off,” the flow
is zero. Thus the pressure drop in the restrictor is zero and the recess pressure pr equals the supply pres-
sure ps. In Figure 15.1(c), pressure continues to build in the manifold and the recess, reaching a value
pL = W/Ar where the load pressure equals the applied load W divided by the recess area Ar . This condition
enables “lift off” or separation of the bearing surfaces. Flow is still zero at this instant and the recess
pressure equals the supply pressure. After “lift off,” Figure 15.1(d), lubricant flows through the bearing.
The recess pressure pr decreases from the value pL because the applied load W is now carried by the
integral of the pressure in the bearing recess in conjunction with the pressure over the land region. Due
to viscous friction, the pressure in the land clearance space gradually reduces in the transverse or radial
direction from pr at the edge of the recess (the start of the land) to zero at the end of the land. Clearance
can be controlled by adjusting the flow rate. Since the flow rate is > 0, there is a pressure drop in the
restrictor and pr < ps. Figure 15.1(e) shows that when the applied load increases, W+∆W, the clearance
decreases, h – ∆h, resulting in reduced lubricant flow. The lower flow rate reduces the pressure drop in
the restrictor allowing the recess pressure to increase pr + ∆ pr. Conversely, Figure 15.1(f) shows that a
reduction in applied load, W – ∆W, leads to an increase in the bearing clearance, h + ∆h. Bearing flow
increases, which increases the loss in the restrictor, leading to a lower recess pressure pr – ∆ pr.
15-4 Theory and Practice of Lubrication and Tribology

Load, W W W
Bearing runner

Bearing pad

Recess pressure, p>0 p = pL


Bearing pr = 0 Q=0
Flow,
recess Restrictor
Supply pressure, Q=0 p>0 p = pL
ps = 0
Manifold

(a) Pump off (b) Pressure building up (c) Pressure × recess area = W

W W + ∆W W – ∆W

ho ho – ∆ho ho + ∆ho

Q p = pr Q p = pr + ∆pr Q p = p r – ∆ pr

p = ps p = ps p = ps

(d) Bearing operating (e) Increased load (f ) Decreased load

FIGURE 15.1  Formation of fluid film in a hydrostatic bearing system. (From Rippel, H.C., Machine Design, Parts
1–10, 1963. With permission.)

The maximum theoretical load capacity is Wmax = psAr. At this load, the bearing pad and the thrust
runner are in contact. The bearing reaches a practical maximum load capacity when the minimum
bearing clearance h is several times the machining deviation of the bearing pad and thrust surface
profiles.
A view of the wide range of possible bearing configurations and control principles is shown in Figure
15.2 [10]. Each bearing family is a form of a journal, thrust, or combined journal and thrust style and
may have single or multiple recesses. For all configurations, starting torque is effectively zero and load
capacity is independent of speed.

15.4  Fundamental Relationships


As in hydrodynamic lubrication, the Reynolds equation provides a basis to obtain a relationship among
the geometry of the bearing surfaces, applied load, relative sliding velocity, lubricant viscosity, and den-
sity, as well as flow rate, supply pressure, and power loss.
A standard form of the Reynolds equation as expressed in polar coordinates, r, θ, is

1 ∂ ⎛ r rh3 ∂p ⎞ 1 ∂ ⎛ rh3 ∂p ⎞ w ∂ ( rh ) (15.1)


⎜ ⎟+ ⎜ ⎟=
r ∂r ⎝ 12h ∂r ⎠ r ∂q ⎝ 12hr ∂q ⎠ 2 ∂q

This equation assumes one surface is rotating about its center with an angular velocity equal to ω.
Assuming that the density and viscosity are constant, this equation becomes

1 ∂ ⎛ 3 ∂p ⎞ 1 ∂ ⎛ 3 ∂p ⎞ ∂h
rh + h = 6hw (15.2)
r ∂r ⎜⎝ ∂r ⎟⎠ r 2 ∂q ⎜⎝ ∂q ⎟⎠ ∂q
Hydrostatic Lubrication 15-5

Externally
pressurized bearings

Liquid Gas

Journal Combined journal and thrust Thrust

Partial Conical Opposed pads Complex pad arrangement

Full Spherical Multiple recess

Single pad

Capillary Orifice Constant flow Pressure sensing Inherent stability


valve

FIGURE 15.2  Hierarchy of hydrostatic bearings. (From O’Donoghue, J.P. and Rowe, W.B., Tribology, 2, 25, 1969.
With permission.)

Examining a simple hydrostatic thrust bearing helps understand the basic theory and the relationships
among the bearing parameters. Figure 15.3 shows a flat circular thrust pad with a recess. The recess can
be machined in either surface. The bearing has an outer radius ro and an inner radius ri defining the
recess. The recess is deep enough to allow the pressure to be constant at pr over the entire recess. Constant
pressure is effectively achieved if the recess depth, hr, is on the order of 50 ho for liquid lubricants.

15.4.1  Pressure Distribution


A view of the bearing from above is shown in the upper part of Figure 15.4. The film thickness is the
same across the land region in both the radial and circumferential directions. The pressure is constant
in the θ direction. In the land region, the Reynolds equation, Equation 15.2, reduces to

∂ ⎛ ∂p ⎞
r =0 (15.3)
∂r ⎜⎝ ∂r ⎟⎠

Integrating gives

dp c1
=
dr r

and

p = c1 ln r + c2

For the bearing shown in Figure 15.4, the boundary conditions are

p = pr at r = ri

p = 0 at r = ro

15-6 Theory and Practice of Lubrication and Tribology

Load, W

ro

ri
Film
thickness, Thrust surface
ho
p=0
Depth, hr
pr

Bearing pad

Flow control device


pressurized supply, ps
flow, Q

FIGURE 15.3  Geometry of a circular step hydrostatic thrust bearing.

q

ri
dr
r
Land
region

ro

p = pr

ri

r
p=0
dr r
ro

FIGURE 15.4  Pressure distribution for a circular step hydrostatic thrust bearing.
Hydrostatic Lubrication 15-7

Applying the boundary conditions gives

ln ( r ro )
p = pr (15.4)
ln ( ri ro )

and

dp pr (15.5)
=
dr r ln ( ri ro )

15.4.2  Flow Rate


The upper half of Figure 15.4 shows a small volumetric element at radius r having a radial length dr and
an angular width dθ. The radial flow rate q across the volumetric element is

ho3 ⎛ dp ⎞
q= − (15.6)
12h ⎜⎝ dr ⎟⎠

There is a negative sign on the pressure term since the gradient is in the opposite direction to the flow.
The total volumetric flow rate Q must be constant from the inner to outer radius, giving

ho3 ⎛ dp ⎞
Q = 2prq = 2pr − (15.7)
12h ⎜⎝ dr ⎟⎠

Substituting Equation 15.5 gives

ppr ho3
Q= (15.8)
6h ln ( ro ri )

15.4.3  Load Capacity


The normal load W is carried by the integral of the pressure over the whole bearing surface as shown in
the lower half of Figure 15.4. The pressure in the recess is constant, so the load capacity is the sum of the
recess pressure pr times the recess area plus the integral of pressure over the land region:

2p ri 2p ro

W=
∫ ∫ p dA + ∫ ∫ pdA
0 0
r
0 ri
(15.9)

Using dA = rdθdr gives

2p ri 2p ro ro

W=
∫∫
0 0
pr rdqdr +
∫∫
0 ri

prdqdr = pri2 pr + 2p prdr
ri
(15.10)

Substituting Equation 15.4 for p and integrating by parts lead to an expression for load capacity that is
only a function of recess pressure and bearing geometry
15-8 Theory and Practice of Lubrication and Tribology

W=
(
ppr ro2 − ri2 ) (15.11)
2 ln (ro /ri )

15.4.4  Frictional Torque


The shear force is proportional to the area A and the shear strain rate

u
f = hA (15.12)
h

For the fluid element shown in Figure 15.4, the force is

⎛ rw ⎞
f = h ( rdqdr ) ⎜ ⎟ (15.13)
⎝ h ⎠

and the frictional torque is

t = fr (15.14)

The bearing torque T is obtained by integrating over the bearing surface. From Equations 15.13 and
15.14,

2p ri 2p ro
w w 3 phw ri4 phw 4 4
T=
∫∫ h r 3dqdr +
hr ∫∫ h
ho
r dqdr =
2hr
+
2ho
( )
ro − ri (15.15)
0 0 0 ri

The torque has two components. The first term is for the recess area and the second term is for the land
region. When hr ≫ ho, the first term, torque over the recess, becomes negligible.

15.4.5  Power Loss


There are two sources of heat generation within the bearing, the friction or viscous dissipation power Hf
and the pumping power Hp. The total power loss is the sum of the losses:

H t = H f + H p = Tw + prQ (15.16)

Since hr is typically much greater than ho, the second (land region) term in Equation 15.15 may be used
for torque T and Equation 15.8 used for flow Q:

phw 2 4 4 ppr2ho3
Ht =
2ho
(
ro − ri + )
6h ln ( ro ri )
(15.17)

If the assumptions are made that the total power loss is dissipated as heat and that all of the heat appears
in the fluid, then
Hydrostatic Lubrication 15-9

H t = Q rc p ΔT (15.18)

where
ρ is the density of the lubricant
cp is the specific heat capacity

The fluid temperature rise is then

Ht
ΔT = (15.19)
Q rc p

A similar analysis can be carried out for other bearing geometries including a flat rectangular step, flat
annular step, conical or spherical shape. The procedure for determining the pressure, flow, load capacity,
friction, and power loss is the same as for the flat circular step bearing. The difference lies in accounting
for the particular geometry.

15.5  Flow Control


Bearings having one recess, such as a circular step or simple thrust pad, may be fed directly from a single
pressurized fluid source. For a common source to supply more than one bearing or pad, some means of
restricting the flow to each pad is required. Controlling the pressure in each recess allows load to be car-
ried independently by each bearing pad. The device used to control the pressure is called a compensator
or restrictor. By limiting the flow to each recess, restrictors enable all of the bearing pads to become
active and contribute to carrying the applied load.
A second function of a restrictor is to control the variation of film thickness as a result of changes
in applied load. Bearing stiffness is similar to a spring constant whereby the load capacity increases as
the film thickness reduces. High stiffness is desirable, particularly in precision machine tools so that
additional loads may be supported with minimal corresponding displacement. Different flow control
devices produce different bearing stiffness characteristics due to the manner in which pressure drop
varies as a function of flow rate in the restrictor.
Four common approaches to controlling pressure are capillary tubes, sharp-edged orifices, con-
stant flow, and pressure sensing valves. These flow controls are listed in the order of increasing bearing
stiffness.

15.5.1  Capillary Compensation


A capillary is a relatively long, narrow diameter tube used to control the lubricant flow by creating a
viscous pressure drop between a main supply manifold and a bearing recess. The pressure drop is given
by the Hagen–Poiseuille equation:

128hlcQc
Δp = (15.20)
pdc4

Rearranging and introducing a constant kc,

kc Δp pdc4
Qc = where kc = (15.21)
h 128lc
15-10 Theory and Practice of Lubrication and Tribology

Thus, kc is a constant based on the geometry of the capillary tube and the capillary flow is linearly related
to the pressure drop Δp = ps − pr. A typical practice is to make Δp/pr ≈ 1 such that the pressure drop Δp is
about one-half of the supply pressure ps.
The capillary flow equation assumes laminar flow, so the Reynolds number should be <2000, where
Re = ρvcdc/η, where ρ = density, vc = average fluid velocity in the capillary, dc = capillary bore diameter, and
η = dynamic viscosity. Viscosity changes due to temperature and pressure are neglected.
The minimum recommended capillary length is greater than 20 times the inside diameter. For best
accuracy, the length-to-diameter ratio should be greater than 100 so that end effects are negligible.
Capillaries with inside diameters in the range of 0.08–0.18 cm are typical. Diameters less than about
0.06 cm are avoided due to the risk of blockage.
Capillary length and diameter are selected to provide a stiffness suitable to the design requirements.
Because the lubricant flow is independent of viscosity, the bearing’s stiffness is independent of tempera-
ture since the effect of viscosity on flow through the bearing is balanced by the effect of viscosity on flow
through the capillary.

15.5.2  Orifice Compensation


In contrast to capillary tubes, sharp-edged orifices are turbulent flow devices. The volumetric flow of
lubricant through an orifice is given by

Qo = Cd Ao 2Δp/r (15.22)

Rearranging and introducing a constant ko,

pdo2Cd
Qo = ko ps − pr where ko = (15.23)
8r

Thus, ko is a constant based on the geometry of the orifice and the flow is proportional to the square root
of the pressure drop Δp = ps − pr.
Orifices are generally more compact than capillary tubes. To avoid the risk of blockage, orifice diam-
eters greater than about 0.05 cm are recommended. The pipe diameter at the orifice should be at least
10 times the orifice diameter. The discharge coefficient Cd is a function of the Reynolds number and is
typically in the range of 0.6–0.65 when the pipe diameter condition is met.

15.5.3  Constant Flow Compensation


The previous two compensating elements use a constant pressure supply and restrict flow by generat-
ing pressure drops. These pressure drops increase the power loss in the system, thus making capil-
lary tubes and orifices more suitable for small machines or intermittent operation. When long-term
power loss is important, another solution is to achieve constant flow by using either a separate pump
to directly supply each recess or flow control valves to divide the flow from a single, larger pump.
Flow is independent of recess pressure because the flow is constant regardless of the pressure dif-
ference across the device. Common flow dividers include positive displacement, series progressive
piston designs, and rotary gear designs. Flow dividers may also be cascaded to supply a large number
of recesses.
Hydrostatic Lubrication 15-11

15.5.4  Pressure Sensing Valve Compensation


Pressure sensing valves represent the most sophisticated compensating devices. The two main groups
are spool valves and diaphragm valves. Both types vary the controlling restrictor as a function of load
on the bearing. The dynamic response for diaphragm valves is generally faster than for spool valves.
Further information about pressure sensing valves can be found in Bassani and Piccigallo [9] and
O’Donoghue and Rowe [10].

15.6  Design Procedure: Multi-Recess Journal Bearing


The majority of design procedures found in the literature are based on graphs and tables. More recently,
researchers have published numerical solutions of the Reynolds equation to study various hydrostatic bear-
ings. The use of graphs and tables requires access to specific references while numerical solutions are gener-
ally not available for others to use, thus requiring independent implementation and validation of solutions.
A third approach is based on a simple design procedure developed by O’Donoghue and Rowe [10]. The
method is applicable to thin land bearings (where a/L ≈ 0.1 to 0.3). The procedure may also be applied
to thick land bearings with the result that the flow and stiffness parameters are slightly overestimated.
The consequence is not serious and can be effectively addressed by providing an excess pump capacity
of ∼20% to allow for flow rate variations due to temperature effects.
An explanation of the design procedure [10] begins with a discussion of bearing geometry. A typical
multi-recess journal bearing having six pads is shown in Figure 15.5. The load orientation is on recess,
or between lands. Axial dividing slots can be added in the land region (Figure 15.6), for bearings oper-
ating at high speeds. This example shows a five pad design, with the load centered between lands. The
axial slots will increase lubricant flow rate, which will assist in heat dissipation. Due to manufacturing
simplicity, journal bearings without axial slots are more common.
General recommendations can be made for the bearing geometry:
• Bearing length-to-diameter ratio L/D = 1
• Axial land width a = L/4

3 5


2 6

1 h
h1,2 h1,6 a
AL

Ar
n
b

πD
πD
B΄= 6na γ = na(L–a)
πDb

FIGURE 15.5  Multi-recess hydrostatic journal bearing, without axial dividing slots. (From O’Donoghue, J.P. and
Rowe, W.B., Tribology, 2, 25, 1969. With permission.)
15-12 Theory and Practice of Lubrication and Tribology

3 4

2 5

h1,2 h1,5

c b

Ar
n a
AL

πD
1+γ a(L–a)
B΄= ( πD –b –c) ( ) γ=
πD
n 6a b( n –b–c)

FIGURE 15.6  Multi-recess hydrostatic journal bearing, with axial dividing slots. (From O’Donoghue, J.P. and
Rowe, W.B., Tribology, 2, 25, 1969. With permission.)

• Circumferential land width b = (πD/4n)


• Design land film thickness, ho, 5–10 times the machining tolerance
• Recess depth ≈ 20 times the design land film thickness (the range is normally 10–100)
• Pressure ratio β = pr /ps = 0.5
A guideline for bearing diameter is given in Table 15.4. This guideline will lead to a supply pressure
that is not significantly greater than 2.1 MPa (∼300 psi), allowing a low cost gear pump to be used.
Higher supply pressures will increase temperature rise but will allow a smaller bearing diameter or
length.
There are two paths by which lubricant can exit the bearing recesses: axially and circumferentially.
Lubricant that exits axially leaves the bearing while lubricant that exits circumferentially passes to a
neighboring recess (or, if present, an axial exhaust slot). A circumferential flow factor, γ, accounts for
the circumferential flow and is defined as a dimensionless ratio of axial flow resistance divided by the
circumferential flow resistance.
Bearing stiffness may be determined using Table 15.1 for designs without axial slots or Table 15.2
for bearings using axial slots. The dimensionless stiffness values are given as a function of the num-
ber of bearing pads and the type of flow compensation. A value for the circumferential flow factor, γ,
is determined during the design process and depends solely on the bearing geometry. Equations for
calculating γ are given later and in Figures 15.5 and 15.6; a typical range is 0.5 < γ < 2.0. The pressure
ratio, β, is assumed to be 0.5, which achieves maximum stiffness at the design load condition. There
are only a few exceptions where it might be advantageous to use a ­pressure ratio other than 0.5 [5].
Dimensionless stiffness values for pressure ratios other than 0.5 are available [10].
Hydrostatic Lubrication 15-13

TABLE 15.1  Dimensionless Stiffness λ′ for a


Journal Bearing with n Pads, without Axial Slots
and with a Pressure Ratio β = pr/ps = 0.5
Dimensionless Stiffness, λ′
Pads, n Capillary Orifice Constant Flow
0.27 0.36 0.54
3
1 + 0.75g 1+ g 1 + 1.5g
0.955 1.28 1.91
4
1 + 0.5g 1 + 0.67g 1+ g
1.03 1.37 2.13
5
1 + 0.35g 1 + 0.46g 1 + 0.69g
1.08 1.43 2.15
6
1 + 0.25g 1 + 0.33g 1 + 0.5g

TABLE 15.2  Dimensionless Stiffness λ′ for a


Journal Bearing with n Pads, with Axial Slots and a
Pressure Ratio β = pr/ps = 0.5
Dimensionless Stiffness, λ′
Pads, n Capillary Orifice Constant Flow
1.13ks 1.5ks 2.25ks
3
1 + 0.5g 1 + 0.67g 1+ g
1.5ks 2k s 3ks
4
1 + 0.5g 1 + 0.67g 1+ g
1.88ks 2.5ks 3.75ks
5
1 + 0.5g 1 + 0.67g 1+ g
2.25ks 3ks 4.5ks
6
1 + 0.5g 1 + 0.67g 1+ g

⎡ sin q ⎤
ks = sin q ⎢ + g cos q ⎥
⎣ q ⎦

The remainder of this section is an adaptation of O’Donoghue’s and Rowe’s [10] design procedure for a
multirecess journal bearing. The example is organized as a series of four tables. The design requirements
are listed in Table 15.3. The bearing geometry, Table 15.4, has been optimized for a length to diameter
ratio of one. To minimize the effect of viscosity variation with temperature, capillary compensation is
chosen over orifice compensation. Table 15.5 gives miscellaneous coefficients and the calculated bearing
performance parameters are summarized in Table 15.6.

15.7  Hybrid Bearings


Some bearings are designed to operate in a hybrid mode of operation, combining hydrodynamic and
hydrostatic lubrication. This hybrid mode can be further characterized as a function of shaft speed.
At low shaft speeds, the mode of operation is principally hydrostatic and the ratio of friction power
to pumping power is low. An example is a large industrial hot gas fan operating on a turning gear to
15-14 Theory and Practice of Lubrication and Tribology

TABLE 15.3  Design Requirements


Symbol Example, with Units Parameter Comments
N 1000 rpm Shaft speed Specified
W′ 8900 N Maximum bearing load Specified
— Capillary Type of flow Select capillary, orifice, or constant flow device
compensation
— Without axial slots Without axial slots Slots are usually not present due to
With axial exhaust slots manufacturing complexity. Slots are
advantageous at high speed for heat dissipation
n 4 Number of bearing Typical range is 3 ≤ n ≤ 6
pads Use n = 4 for most bearings
n = 6 for high precision applications
β 0.5 Pressure ratio β = pr/ps Use β = 0.5
The maximum stiffness at the design load is
obtained when β = 0.5. Values > 0.5 are to be
avoided because stiffness and load range are
both reduced

TABLE 15.4  Bearing Geometry


D 160 mm Bearing diameter Use D = 2.9W ʹ when D is mm and W is N
(use D = 0.02W ʹ when D is inches and
W is lbf )
L 160 mm Bearing length Use L = D
ho 50 μm Design bearing Use ho = 5–10 times the machining tolerance
clearance on ho. Typical values for ho are greater for
hydrostatic bearings than for the same size
hydrodynamic bearing. Usually on the
order of 50 μm (0.002 in.)
a 40 mm Axial flow land Use a = L/4 for optimum width
width
b 42 mm Circumferential Use b = πD/3n
flow land width
c — Axial slot width Use c = πD/8n
⎛ pD ⎞
Ar 100 cm2 Recess area for Without axial slots, Ar = ⎜ − b ⎟ (L − a)
one pad ⎝ n ⎠

⎛ pD ⎞
With axial slots, Ar = ⎜ − b − c ⎟ (L − a)
⎝ n ⎠
pDL 3Ar
Af 125 cm2 Friction area for Af = −
one pad n 4

avoid shaft bow as the fan cools to ambient temperature. This type of bearing is also known as oil lift
or lift-assisted. Small recesses, typically 2%–5% of the bearing’s projected area (bearing length times
diameter), are supplied with high-pressure oil during start-up and shutdown in order to reduce wear
and drive torque (friction). For thrust bearings, the pocket area is usually somewhat larger, about 5% of
the bearing’s thrust area.
At high shaft speeds, the mode of operation may be primarily hydrodynamic. The ratio of friction
power to pumping power is high. An example is a high-speed precision grinding head spindle. In order
to reduce the friction power loss, the hydrodynamic action can be reduced by increasing the hydrostatic
Hydrostatic Lubrication 15-15

TABLE 15.5  Miscellaneous Coefficients


γ 0.91 Circumferential flow factor for journal Carryover lubricant to adjacent recess or axial
bearings slot. Effect is least for small values of γ. Typical
range is 0.5 < γ < 2.0
γ = (axial flow resistance/circumferential na(L − a)
flow resistance) Without axial slots, use g =
pDb

With axial slots, use g =


na(L − a)
b(pD − nc − nb)
B′ 0.52 Flow shape factor for thin land bearings pD
Without axial slots, use Bʹ =
6an
(pD − nc − nb)(1 + g)
With axial slots, use Bʹ =
6an
θ — Pad angle Without axial slots, not applicable
p b+c
With axial slots, use q = −
n D
ks — Dimensionless stiffness factor for bearings Without axial slots, not applicable
with slots
⎡ sin q ⎤
With axial slots, use ks = sin q ⎢ + g cos q ⎥
⎣ q ⎦
0.955
Use lʹ =
1 + 0.5g
λ′ 0.66 Dimensionless stiffness Without axial slots, use Table 15.1
With axial slots, use Table 15.2

TABLE 15.6  Performance Parameters


3W ʹ
ps 2.11 MPa or Minimum supply Use ps =
lʹD(L − a)
N/m2 pressure
ps D(L − a)
λo 535 × 106 N/m Actual stiffness Use lo = lʹ
ho
If stiffness is too low, increase ps and/or D
h′min 33 μm Minimum running Use hmin
ʹ = ho − W ʹ/lo
clearance
pDN
u 8.38 m/s Sliding velocity Use u = If u ≈ 0, skip the next
60
step and use a convenient viscosity η,
perhaps 20 cP

ps ho2 bBʹ
ƞ 0.00285 Dynamic viscosity Use h =
u Af
Pa·s or N·s/m2 If η is too low, reduce land width or
(2.85 cP) increase ho and repeat the calculations

ps ho3
Q 96 cm3/s or Flow rate Use Q = nbBʹ
5.76 L/min h
ΔT 2.5°C Adiabatic Speed ΔT (°C) ΔT (°F)
temperature rise u ≠ 0 1.21 × 10−6 ps 0.015ps
u ≈ 0 0.61 × 10−6 ps 0.0075ps
15-16 Theory and Practice of Lubrication and Tribology

pad area. Recirculation within a recess increases bearing friction and the impact on total power con-
sumption and temperature rise needs to be considered.
A key difference between hydrostatic and hybrid bearings is the absence of large recesses in the hybrid
bearings. When hydrodynamic operation is desired, recesses reduce the hydrodynamic pressure in the
loaded portion of the bearing surface. Therefore, a small recess or supply groove of minimal area is used
in the loaded zone of a hybrid bearing so as to minimize the interruption of the hydrodynamic film.
Any slot or hole outlets should be located away from the loaded zone to allow for more efficient hydro-
dynamic operation.

15.8  Optimization
Various criteria may be applied to the optimal design of hydrostatic bearings. Desirable features include
low power loss, high stiffness, low temperature rise, and large clearance to allow large machining toler-
ances. Parameters such as bearing size, recess area, lubricant flow rate, and pressure can be varied to
maximize load capacity or stiffness for a given oil flow or minimum power loss. The general design task
is to minimize total power loss given a set of required operating conditions, typically applied load and
minimum film thickness.
For both hydrostatic and hybrid bearings, Rowe [11] defines a power ratio K as a function of the two
sources of heat generation within the bearing:

Hf
K= (15.24)
Hp

This ratio can be used as a measure of the proportion of hydrodynamic effects to hydrostatic effects. For
the case of no rotation, K ≈ 0 and hydrostatic effects dominate. As K increases, hydrodynamic effects
become increasingly significant.
Power loss for a specific geometry, a flat circular step thrust bearing, was examined in Section 15.4.5.
The following discussion considers the optimization of power loss for the same bearing, with rotation
and without any hydrodynamic effects.
The friction power is proportional to the fluid viscosity and the shear rate

u2
H f = hA f (15.25)
h

and the pumping power is the supply pressure ps times the flow rate Q

ps2ho3
H p = psQ = Bʹ (15.26)
h

where B′ is a dimensionless flow factor based on the geometry of the bearing. For the example of the
circular step thrust bearing, the volumetric flow rate Q from Equation 15.8 may be used with Equation
15.26 to find B′ = π/(6 ln(ro/ri)).
The friction area Af is comprised of two parts: area over the recess and area over the bearing lands:

A f = Ar + Ao (15.27)

Substituting into Equation 15.25 gives


Hydrostatic Lubrication 15-17

⎛ Ar A ⎞
H f = hu 2 ⎜ + o⎟ (15.28)
⎜ ( hr + ho ) ho ⎟
⎝ ⎠

There are two depths over the friction area: the depth over the land is ho and the depth over the recess
is (hr + ho).
Since the recess depth is much greater than the film thickness in the land region, the value of the
recess term will be small compared to that of the land term; thus,

⎛A ⎞
H f ≈ hu 2 ⎜ o ⎟ (15.29)
⎝ ho ⎠

The relationship for total power may then be written as the sum of Equations 15.29 and 15.26:

u 2 ps2ho3
H t = hAo + Bʹ (15.30)
ho h

Total power loss may be minimized by varying film thickness. It is readily seen that by taking the
derivative with respect to ho, the condition Hf = 3Hp results. In this case, K = 3, and the hydrostatic and
hydrodynamic effects on load are of the same order. If the total power loss is minimized with respect
to viscosity, then the condition Hf = Hp, or K = 1, results. Bearings operating with K ≪ 1 are considered
“low speed” and are considered “high speed” when K ≥ 1. Optimum values for bearing parameters thus
occur in the range Hp ≤ Hf ≤ 3Hp.
Rowe [11] also describes the ratio of the design recess pressure to supply pressure, β = pr/ps, as a design
parameter. Regardless of the compensation method employed, maximum stiffness is achieved when
β = 0.5. Furthermore, there is a preferred range for the power ratio K such that when 1 < K < 3, the total
power will not exceed the minimum value by >15%.
Due to recirculation in the bearing recesses, the recess friction is nonzero and may be accounted for
by modifying the friction area. For bearings with nonturbulent flow in the recesses, Ao may be replaced
by an effective friction area:

ho
Aeff = Ao + 4 Ar (15.31)
hr

The friction area is increased to some extent by considering the recess area scaled by the ratio of the film
thicknesses in the land and recess regions.
Using Equation 15.24 as Hf = KHp, the total power loss may be expressed as a function of the power
ratio K:

H t = H f + H p = KH p + H p = H p ( K + 1) (15.32)

Solving the adiabatic heat equation, Equation 15.18, for ΔT and from Equation 15.26 Hp = psQ

ps ( K + 1)
ΔT = (15.33)
rc p

The power ratio is now related to the maximum possible temperature rise. If the temperature rise is to
be minimized, then the supply pressure should be as low as practical subject to controlling the load.
15-18 Theory and Practice of Lubrication and Tribology

15.9  Summary
Unlike hydrodynamic bearings, the load-carrying capacity of hydrostatic bearings does not rely on the
relative motion between the bearing surfaces; rather, it depends on a supply of pressurized lubricant in
the loaded region. Thus, two surfaces may be separated by a film of lubricant at zero speed and wear is
effectively eliminated. Load capacity and stiffness are largely dependent on the geometry of the bearing
and the type of flow restrictors used. Hydrostatic bearing performance parameters may be manipulated
more easily to suit design requirements than those for hydrodynamic bearings because the number of
design variables is greater for hydrostatic bearings.
The effect of manufacturing tolerances on the design of externally pressurized journal bearings is not
widely covered in the literature. Stout and Rowe [6–8] address this topic in a series of three works. A
non-recessed bearing feeding two rows of circumferential slots (double-entry feeding arrangement) is
shown to produce the best overall performance when considering load capacity, flow rate, and sensitiv-
ity to manufacturing variation. In order to minimize the sensitivity to manufacturing errors, a design
pressure ratio β = 0.5 is recommended; the power ratio should be in the range 1 ≤ K ≤ 3; and the operat-
ing eccentricity should be conservatively designed, e = 0.5.
Comprehensive information about hydrostatic bearing design procedures can be found in Bassani
and Piccigallo [9]. The field of hydrostatic bearings has evolved to the point that detailed design and cal-
culation methods for hydrostatic journal bearings, both with and without drainage grooves, are docu-
mented and available in international standards [12–15].

References
1. Girard, L.D., Comptes Rendus, April 28 and October 6, 1851, February 23, 1852.
2. Girard, L.D., Nouveau systeme de locomotion sur chemin de fer, Bachelier, Paris, France, 1852, p. 40.
3. Fuller, D.D., Hydrostatic lubrication, Machine Design, June through December, Parts 1–5, 1947.
4. Rippel, H.C., Design of hydrostatic bearings, Machine Design, August 1–December 5, Parts 1–10,
1963.
5. Rowe, W.B. and O’Donoghue, J.P., A review of hydrostatic bearing design, Proceedings of Conference
on Externally Pressurized Bearings, IMechE/Institution of Production Engineers, London, C21/71,
1971, pp. 157–187.
6. Stout, S.J. and Rowe, W.B., Externally pressurized bearings—Design for manufacture—Part I—
Journal bearing selection, Tribol. Int., 7(3), 1974, 99–106.
7. Stout, S.J. and Rowe, W.B., Externally pressurized bearings—Design for manufacture—Part 2—
Design of gas bearings for manufacture including a tolerancing procedure, Tribol. Int., 7(4), 1974,
169–180.
8. Stout, S.J. and Rowe, W.B., Externally pressurized bearings—Design for manufacture—Part 3—
Design of liquid externally pressurized bearings for manufacture including tolerancing procedures,
Tribol. Int., 7(5), 1974, 195–214.
9. Bassani, R. and Piccigallo, B., Hydrostatic Lubrication: Theory and Practice, Tribology Series 22,
Elsevier Science Publishers, Amsterdam, the Netherlands, 1992.
10. O’Donoghue, J.P. and Rowe, W.B., Hydrostatic bearing design, Tribology, 2(1), 1969, 25–71.
11. Rowe, W.B., Advances in hydrostatic and hybrid bearing technology, Proc IMechE, Part C: Journal
of Mechanical Engineering Sciences, 203, 1989, 225–242.
12. ISO 12167-1:2001, Plain bearings—Hydrostatic plain journal bearings with drainage grooves under
steady-state conditions—Part 1: Calculation of oil-lubricated plain journal bearings with drainage
grooves.
Hydrostatic Lubrication 15-19

13. ISO 12167-2:2001, Plain bearings—Hydrostatic plain journal bearings with drainage grooves under
steady-state conditions—Part 2: Characteristic values for the calculation of oil-lubricated plain jour-
nal bearings with drainage grooves.
14. ISO 12168-1:2001, Plain bearings—Hydrostatic plain journal bearings without drainage grooves
under steady-state conditions—Part 1: Calculation of oil-lubricated plain journal bearings without
drainage grooves.
15. ISO 12168-2:2001, Plain bearings—Hydrostatic plain journal bearings without drainage grooves
under steady-state conditions—Part 2: Characteristic values for the calculation of oil-lubricated
plain journal bearings without drainage grooves.
16
Hydrodynamic
Lubrication
16.1 Introduction..................................................................................... 16-1
16.2 Governing Equations of Fluid Mechanics for Lubrication
Flow.................................................................................................... 16-2
16.3 Fundamental Simplifications of Lubrication Theory.................16-4
16.4 Reynolds Equation of Hydrodynamic Lubrication: The
One-Dimensional Bearing Case.................................................... 16-7
16.5 Two-Dimensional Versions of Reynolds Equation.....................16-8
Two-Dimensional Incompressible Reynolds Equation in Cartesian
Coordinates  •  Two-Dimensional Compressible Reynolds Equation
in Cartesian Coordinates  •  Two-Dimensional Incompressible
Reynolds Equation in Polar Coordinates  •  Two-Dimensional
Incompressible Reynolds Equation in Spherical Coordinates
16.6 Various Modifications of Reynolds Equation........................... 16-10
John A. Tichy Modified Reynolds Equation for Lubricant Inertia  •  Modified
Rensselaer Polytechnic Reynolds Equation for Turbulence  •  Some Additional Examples
Institute References................................................................................................... 16-12

16.1  Introduction
The purpose of lubrication is to separate two nearby surfaces in relative motion with a film of inter-
vening material. If no contact between the surfaces occurs, and if the intervening material is a fluid,
one could call conditions fluid film lubrication. If the forces separating the two surfaces and the fluid
flow are generated by the surface motion itself, the mode of lubrication is referred to as hydrodynamic
lubrication. If the separating forces and fluid flow are ultimately due to some external pumping action
conditions are those of hydrostatic lubrication, the subject of Chapter 15 of this handbook. The basic
physics and fluid mechanics occurring in hydrodynamic and hydrostatic lubrication are essentially the
same, although the technologies of many applications are different.
In the “traditional” tribology applications of hydrodynamic lubrication, such as journal and thrust
bearings, films are referred to as “thick” when surfaces are separated by greater than, say, 10 μm, a
dimension much greater than the molecular scale of nanometers. However, in the terminology of fluid
mechanics, thin film flow refers to flow in which the film thickness is much less than a characteristic
dimension in the direction of flow. A thin film thickness in this case could easily be in the millimeter
range, much greater than the thick film of tribology. Occasionally in the scientific literature, one sees
reference to “lubrication flow” or hydrodynamic lubrication to describe the fluid mechanics of a situa-
tion even though there is no desire to separate two surfaces.

16-1
16-2 Theory and Practice of Lubrication and Tribology

The other extreme of lubrication in which there is significant contact of the surfaces is referred to
as boundary lubrication. There are other intermediate conditions with partial surface contact (mixed
lubrication) or significant elastic deformation (elastohydrodynamic lubrication) which are discussed
elsewhere in other sections.
In the case of hydrodynamic lubrication, as to the characteristics of the lubricant, only the thermal
and mechanical properties play a role in the lubrication process. The fluid viscosity is by far the most
important lubricant property, with density playing an occasional secondary role. For “unusual” lubri-
cants such as, say, synovial fluid in animal joints, additional mechanical properties such as relaxation
time may also be significant. The chemical properties associated with lubricity, slipperiness, stickiness,
and other phenomena which depend on how the lubricant reacts with the surface are definitely not
important in hydrodynamic lubrication. Under typical conditions, water would have exactly the same
hydrodynamic lubrication properties as an oil of the same viscosity.
In a highly simplified and heuristic fashion, hydrodynamic lubrication can be explained as follows.
Viscous fluids resist a shearing motion by developing a shear stress. To balance the forces on a fluid ele-
ment, the stress generates a pressure field which tends to separate the surfaces. The shearing motion is
produced by either sliding or squeezing of the confining surfaces.
A 1-D hydrodynamic contact (the film shape described by one spatial coordinate) requires a 2-D fluid
flow (coordinates along and across the film). Likewise, a two-dimensional contact requires 3-D flow.
Thus, confusion can arise by way of terminology, if one simply refers to the “two-dimensional” case.

16.2 Governing Equations of Fluid Mechanics


for Lubrication Flow
The compressible continuity equation (16.1) and the generalized Newtonian viscous fluid model (16.2),
along with the linear momentum equation (16.3) and thermal energy equation for an incompressible
fluid (16.4) are as follows:

3
∂r ∂
∂t
+ ∑ ∂x m
(rvm ) = 0 (16.1)
m =1

∂v ∂v j
tij = hg ij , g ij = i + , pij = − pdij + tij (16.2)
∂x j ∂x i

3 3
⎛ ∂v ∂v i ⎞ ∂pmi
r⎜ i +
⎝ ∂t
∑ vm ⎟=
∂x m ⎠ ∑ ∂x m
+ rg i (16.3)
m =1 m =1

3 3 3 3
⎛ ∂T ∂T ⎞ ∂ ⎛ ∂T ⎞ ∂v n
rcv ⎜
⎝ ∂t
+ ∑ vm ⎟=
∂x m ⎠ ∑ k
∂xm ⎜⎝ ∂xm ⎟⎠
+ ∑∑t mn
∂x m
(16.4)
m =1 m =1 m =1 n =1

where
t denotes time
The symbol xi denotes the three coordinate directions (i = 1, 2, 3)
vi denotes the flow velocity components
T is the fluid temperature
p is the pressure
Hydrodynamic Lubrication 16-3

The symbols cv, k, ρ, and η are the fluid specific heat, thermal conductivity, and density, respectively,
while ρgi is a body force per unit volume (gravity in this case). The nomenclature of the section gener-
ally follows that of Bird et al. [1]. In Equation 16.2, τij is the extra or deviatoric stress (the portion of the
stress determined by the flow field), g ij is the rate of strain tensor, πij is the total stress (the physical force
per area on a surface of a fluid element), p represents the fluid pressure, and δij is the unit tensor (=1 for
i = j, = 0 otherwise).
The term generalized Newtonian (or purely viscous) means that the extra stress is only a function of
the rate of strain, as shown in Equation 16.2 with viscosity being the ratio of the two quantities. If vis-
cosity is constant, the flow is said to be Newtonian. In the shear-dominated flows of lubrication, terms
such as τ11 are small and the normal stress τ11 equals −p. Likewise, in shear flows, the off-diagonal extra
stress τ12 is the shear stress, and the strain rate g⋅12 is the shear rate. Equation 16.2 can be substituted in
Equation 16.3, and if the viscosity η is constant, one obtains the incompressible Navier–Stokes equations
(data not shown) which govern classical fluid mechanics.
These equations can be found in any standard fluid mechanics text such as Schlichting and Gersten
[2]. The incompressible form of the momentum equation will prove to be adequate even for compress-
ible gas bearings (see Section 16.5.2). The Navier–Stokes equations are the momentum or force balance
equations. The left-hand side describes fluid inertia or acceleration, while the first two right-hand side
terms represent the net pressure and viscous forces, respectively. The inertia includes an unsteady term
∂vi/∂t and nonlinear convective terms such as v1(∂v1/∂x1).
We begin with the full 3-D flow, of a fluid confined between a lower surface 1 and upper surface 2
at (x3)1 and (x3)2, respectively. The bearing geometry has three spatial coordinate directions with three
length scales (see Figure 16.1). The primary flow direction (usually the direction of sliding) is x = x1 of
reference or characteristic length B. Such a reference length might be the breadth of the bearing in the
case of a slider or thrust bearing, and the radius in the case of a journal bearing. The coordinate direc-
tion across the film, and normal to x is z = x3, and the third orthogonal direction is y = x2 in the general
plane of the bearing surfaces.
The film thickness h(x, y, t) is the local distance between the two surfaces, = z2 − z1, with reference film
thickness H. The reference film thickness could be the inlet, exit, or mean thickness. The characteristic
length in the y-direction is a bearing length L. There are three flow velocities: vx, vy, and vz, with char-
acteristic values Vx, Vy, and Vz . If Figure 16.1 represents a linear slider (thrust) bearing, the velocity Vx is
simply the sliding velocity and Vz is a velocity in the squeezing direction.
Consistent with the purposes of a handbook—to assist the practitioner as opposed to provide a defini-
tive scholarly exposition—in the analysis and presentation that follows, the 2-D incompressible flow
(1-D bearing) case is first presented. In this way, the basic physical and mathematical principles involved
can be more clearly illustrated, with fewer mathematical terms and expressions required. Later, the
3-D flow (2-D bearing) cases with the compressible effect are presented as natural extensions. Several

Vz

2
h(x,y,t)
z B
H Vy x
Vx
1
y

FIGURE 16.1  Schematic of 2-D film.


16-4 Theory and Practice of Lubrication and Tribology

excellent textbooks such as Szeri [3] or Hamrock [4] develop the hydrodynamic theory in careful detail,
and the interested reader is referred to these works.
The incompressible continuity equation, Equation 16.1 for ρ = constant, becomes

∂v x ∂v y ∂v z (16.5)
+ + =0
∂x ∂y ∂z

Now, consider an incompressible flow in a thin film driven by a sliding velocity Vx. The velocities in
the directions along the film have orders of magnitude vx ∼ Vx and vy ∼ Vx. The spatial derivatives of the
velocities have orders of magnitude ∂/∂x ∼ 1/B, ∂/∂y ∼ 1/L, and ∂/∂z  ∼ 1/H. In the case of an unsteady pro-
cess, the time derivatives have the order of magnitude ∂/∂t ∼ 1/tT, where tT is an imposed time scale such
as 1/ω where ω is an oscillation frequency [rad/s]. For a “long” bearing L >> B, and thus, ∂vy/∂y << ∂vx/∂x,
the continuity equation becomes

∂ v x ∂v z
+ =0 (16.6)
∂x ∂z

Thus, the problem is now 2-D, and the constitutive flow model (16.2) and momentum equations (16.3
and 16.4) are expressed as

∂v x ⎛ ∂v ∂v ⎞
txx = −tzz = 2h , txz = tzx = h ⎜ x + z ⎟ (16.7)
∂x ⎝ ∂z ∂x ⎠

⎛ ∂v ∂v ∂v ⎞ ∂p ∂txx ∂tzx
r ⎜ x + vx x + vz x ⎟ = − + + + rg x (16.8)
⎝ ∂t ∂x ∂z ⎠ ∂x ∂x ∂z

⎛ ∂v ∂v ∂v ⎞ ∂p ∂t ∂t
r ⎜ z + v x z + v z z ⎟ = − + xz + zz + rg z (16.9)
⎝ ∂t ∂x ∂z ⎠ ∂z ∂x ∂z

⎛ ∂T ∂T ∂T ⎞ ∂ ⎛ ∂T ⎞ ∂ ⎛ ∂ T ⎞ ∂v x ⎛ ∂v ∂v ⎞
rcv ⎜ + vx + vz ⎟= ⎜k ⎟ + ⎜k ⎟ + 2txx + txy ⎜ x + z ⎟ (16.10)
⎝ ∂t ∂x ∂z ⎠ ∂x ⎝ ∂x ⎠ ∂ z ⎝ ∂ z ⎠ ∂x ⎝ ∂z ∂x ⎠

The terms in parentheses on the left-hand side of Equations 16.8 and 16.9 represent the acceleration of
a fluid element in a local sense (velocity change at a spatial point) and convective sense (velocity change
following the motion). The first three right-hand side terms are the net pressure and viscous force on
the surfaces of a fluid element (tractions). The terms in parentheses on the left-hand side of Equation
16.10 represent the temperature (thermal energy) change of a fluid element in a local sense (temperature
change at a spatial point) and convective sense (temperature change following the motion). The first
three right-hand side terms represent the net conduction on the surfaces of a fluid element, while the last
two terms represent the viscous dissipation (work due to fluid friction).

16.3  Fundamental Simplifications of Lubrication Theory


Considering the earlier 2-D flow problem of Equations 16.6 through 16.9, the fundamental character-
istic of lubrication flow is now addressed: the fact that there are two small length scale ratios which
enable one to simplify these equations considerably. Figure 16.2 schematically portrays the 1-D contact
to be studied. The lower surface is called surface 1, and the upper surface is called surface 2. The lower
Hydrodynamic Lubrication 16-5

2 ∂h/∂t

Vz
V2T
z
Vx h(x,y,t) V2S

x
1
V1S
B

FIGURE 16.2  Schematic of 1-D film.

surface slides at velocity V1S in the x-direction. The upper surface slides at speed V2S and translates in
the x-direction at speed V2T. The normal or z-component of the translation is V2N = ∂h/∂t. Regarding
the sliding of the upper surface, one visualizes a ribbon which is always moving tangentially to its local
direction. In the case of a cylindrical upper surface, this motion is achieved by a simple rotation; oth-
erwise, a more complex mechanism would be required. In any case, for a small slope it is seen that
(=∂h/∂x « 1), the V2Sx ≈ V2S and V2Sz ≈ V2S (∂h/∂x).
The first length scale ratio is that the confined film is thin, i.e., H/B « 1. The two terms of Equation 16.6
must balance each other which requires that vz ∼ VxH/B « vx . This conditions reduces Equations 16.7
through 16.10 to

∂v x ∂v ∂v
txx = −tzz = 2h  txz = tzx = h x , g = x , h = h(T , g ) (16.11)
∂x ∂z ∂z

⎛ ∂v ∂v ∂v ⎞ ∂p ∂ ⎛ ∂v x ⎞
r ⎜ x + vx x + vz x ⎟ = − + ⎜h ⎟ + rg x (16.12)
⎝ ∂t ∂x ∂z ⎠ ∂x ∂z ⎝ ∂z ⎠

∂p
0=− + rg z (16.13)
∂z

2
⎛ ∂T ∂T ∂T ⎞ ∂ ⎛ ∂T ⎞ ⎛ ∂v x ⎞
rcv ⎜ + vx + vz ⎟= ⎜k ⎟ + h⎜⎝ ⎟ (16.14)
⎝ ∂t ∂x ∂z ⎠ ∂z ⎝ ∂z ⎠ ∂z ⎠

The body force terms are of order 103 N/m3. The middle term of the right-hand side of Equation 16.11
(shear stress gradient) is of the order ηVz/H2, and for η = 0.01 Pa·s, Vx = 1 m/s, and H = 0.1 mm, the shear
stress term is ∼107 N/m3. It is clear that the body force terms can be omitted without significant loss
of accuracy. One need not regard the omission of body forces terms as an ad hoc “assumption,” of the
theory.
The second length scale ratio can be expressed in a number of fashions, but one such statement is
that the entrance length Le into the lubrication contact is short, i.e., Le/B « 1. This implies that contrary
to most fluid mechanics, conditions upstream do not affect the flow at a given position—only local
conditions are of influence. If the derivative in the primary direction of the motion is ∂vx/∂x ∼ Vx/Le, the
entrance region can be seen as a competition between convective inertia force per volume ∼ρVx(Vx/Le)
and viscous resistance ∼ηVx/H2. Equating these two quantities and rearranging, one finds that

Le H rVx H
= Re∗ = Re , Re = (16.15)
B B h

16-6 Theory and Practice of Lubrication and Tribology

where
Re is the conventional Reynolds number of channel flow
Re* is the so-called modified or reduced Reynolds number

One could also say that the reduced Reynolds number represents the ratio of the inertia force (mass
times acceleration) to the net viscous force on a fluid material element. Using the same parameter values
as given earlier, and if density ρ is of order 1000 kg/m3, film thickness ratio H/B = 0.001 (a typical ratio
in most applications), contact breadth B = 10 cm, Reynolds number Re equals 1.0, and reduced Reynolds
number Re* = 10−3. Clearly, the inertia force is negligible, although in certain high-speed cases in which
the film thickness ratio is larger, an inertia effect can be important.
The relative orders of magnitude of the terms of the energy equation (16.14) can also be evaluated.
The energy of convection is ∼ρcvΔTVx/B (energy transport due to movement of lubricant along the film),
energy of conduction is ∼kΔT/H2 (energy transport to the bearing surfaces across the film), and viscous
dissipation is of order ηVx/H2. The ratio of convection to conduction is given by

rcv ΔTVx /B H rc V H hc p
= Pe∗ = Pe , Pe = x x = Re (16.16)
k ΔT /H 2 B k k

where
Pe is the Péclet number
Pe* is the modified Péclet number

For cv ≈ cp ∼ 1000 J/kg·K and k ∼ 0.1 W/m·s, Pe* = 10, meaning energy transport by conduction is about 10
times that of conduction. Based on this argument, most design models omit the conduction terms of the
energy equation, and regard the temperature as T(x, t).
Therefore, (for the 2-D flow case), the basic equations are

∂ v x ∂v z
+ =0 (16.17)
∂x ∂z

∂p ∂tzx ∂p ∂ ⎛ ∂v x ⎞ ∂v x
0=− + =− + ⎜h ⎟ , tzx = h (16.18)
∂x ∂z ∂x ∂z ⎝ ∂z ⎠ ∂z

∂p
0= (16.19)
∂z

2
⎛ ∂T ∂T ⎞ ⎛ ∂v x ⎞
rcv ⎜ + vx ⎟ = h⎜⎝ ⎟ (16.20)
⎝ ∂t ∂x ⎠ ∂z ⎠

One notes that the length scales present require that pressure not vary across the film (Equation 16.16),
and just as in the case of discarding body force terms, this condition on pressure does not have to be
invoked as an assumption of the theory. Apart from generalized Newtonian conditions in which viscosity
η is a function of the strain rate (to be discussed later), for constant viscosity, the problem is entirely linear.
An averaged version of the energy equation is presented in Equation 16.20, which is consis-
tent with T = T(x, t) and assumes no energy loss through the bearing surfaces (adiabatic). The set of
Equations 16.17 through 16.20 is strongly coupled through the viscosity variation. A more complete
resolution of the problem would require solving Equation 16.14 with Equations 16.17 through 16.19.
Hydrodynamic Lubrication 16-7

The boundary conditions for nonporous rigid surfaces are as follows:

z = 0 : v x = V1S , v z = 0

∂h ∂h
z = h(x , t ) : v x = V2S + V2T , v z = + V2S
∂t ∂x

x = 0, B : p = pa , x = 0 : T = Tin (16.21)

Several points should be noted. We measure the normal coordinate z with respect to the lower surface,
which may not be flat. Although proof is not presented here, for the linear case, there is no loss of gener-
ality in transforming a film space bounded by a lower curved surface to a flat boundary. Thus, the lower
surface is defined by z = 0, and the upper surface by z = h(x, t) where h is the film thickness. One assumes
that fluid particles adhere to the surface, and thus, the normal velocity vz is zero.
Another consequence of fluid elements adhering to the contiguous surface is that the fluid slid-
ing velocity is that of the surface: the “no-slip condition.” The validity of this hypothesis is generally
accepted at micrometer length scales and above. The normal velocity (the z-component) at the upper
surface is simply ∂h/∂t. The flow rate through an element of surface normal to the z-direction Ldx is
∂h/∂t plus a second term due to the sliding of the inclined surface V2Tx(−∂h/∂x).
The pressure is set equal to the ambient pressure Pa at the ends of the film. The lubricant has an inlet
temperature Tin. Because the inertia terms are suppressed, and no terms such as ∂vx/∂x are present in
the momentum equation, no inlet or exit boundary conditions on the flow are required as would nor-
mally be the case in a fluid flow problem. Likewise, because the conduction terms are suppressed, no
temperature boundary conditions at the surfaces are required. As a practical matter, these conditions
are extremely difficult to discern and would involve the geometry and thermal properties of the material
of the bearing structure. The adiabatic assumption used here is much simpler and conservative in the
sense that it predicts a greater temperature rise due to viscous heating, and thus a lower viscosity and
lower load capacity.
Although time appears in the formulation through the boundary conditions, the theory is funda-
mentally steady (quasi-steady) due to the suppression of the inertia terms of the momentum equation.
In a mathematical sense, time is a parameter, not an independent variable. The problem is entirely local
and instantaneous. The resulting fluid velocity will be completely in phase with an imposed surface
motion, and depends only on the local pressure gradient and film thickness.
One of the fundamental properties of lubrication is simply revealed by the scales, without actu-
ally solving the differential equations. Load carrying is proportional to pressure p and friction force
is proportional to shear stress τzx. If f is the coefficient of friction, from the scaling, it is observed that
f  ∼ τzx/p ∼ H/B, i.e., the friction coefficient is proportional to the thinness ratio of the contact.

16.4 Reynolds Equation of Hydrodynamic Lubrication:


The One-Dimensional Bearing Case
The goal of this section is to derive the 1-D form of the renowned Reynolds equation which governs
hydrodynamic lubrication. Although the present development is limited for illustrative purposes to
2-D flow in a 1-D confined film, the physical principles and the mathematics are essentially the same as
in the 3-D case. In the development to this point, and in what follows, the fewest possible assumptions
are made, and they are introduced only as needed.
Equation 16.15 is integrated twice with respect to z, and the two boundary conditions on vx applied.
Nonconstant viscosity is acceptable, but only η = η(x, t), i.e., the viscosity cannot vary across the film.
16-8 Theory and Practice of Lubrication and Tribology

In a practical sense, one could regard the viscosity used as a sort effective value, averaged across the film.
One thus obtains

z z 1 ∂p 2
v x = V1S ⎛ 1 − ⎞ + (V2S + V2T ) + (z − zh) (16.22)
⎝ h⎠ h 2h ∂x

Following the continuity equation (16.14), integrating Equation 16.18 for vx with respect to x and then
integrating with respect to z, using vz(Z = 0) = 0, the cross-film velocity vz is found:

V2S + V2T − V1S ⎞ ∂h z 2 ∂ ⎡ 1 ∂p ⎛ z 3 z 2h ⎞ ⎤


vz = ⎛ + ⎢ − − ⎥ (16.23)
⎝ 2 ⎠ ∂x h 2
∂x ⎣ h ∂x ⎜⎝ 6 4 ⎟⎠ ⎦

Applying the boundary condition (16.21) on vz at the upper surface z = h after some rearrangement pro-
duces Reynolds equation:

∂ ⎛ h 3 ∂p ⎞ 1 ∂h ∂h
= (V1S + V2S − V2T ) + (16.24)
∂x ⎜⎝ 12h ∂x ⎟⎠ 2 ∂x ∂t

originally derived by Reynolds in 1886 [5]. The right-hand side terms are geometric and kinematic,
describing the shape of the film and the surface velocities. Usually, they are loosely considered to be
the problem inputs. The left-hand-side term describes the pressure generated and must be found by
integration or equivalently solving the differential equation (16.20). The first term on the right-hand
side is the hydrodynamic wedge term. This term, when linked to the pressure term, describes how
fluid drawn into the gap by the sliding velocities generates pressure due to viscous flow resistance.
The second term on the right-hand side is the so-called squeeze film term, which, when linked to the
pressure term, describes how fluid confined within the gap generates pressure due to squeeze flow
resistance. Reynolds equation is linear for constant viscosity, thus doubling the speed doubles the
pressure, etc.
The overall load-carrying capacity (normal force) and friction (tangential force) are calculated from

B B

∫ ∫
Fz = L pdx , Fx = L tzx (z = 0)dx (16.25)
v 0

16.5  Two-Dimensional Versions of Reynolds Equation


There is a huge number of possible versions of Reynolds equation, some of which will be presented here.

16.5.1 Two-Dimensional Incompressible Reynolds


Equation in Cartesian Coordinates
A general 2-D version of Reynolds equation in Cartesian coordinates is shown as follows:

∂ ⎛ h 3 ∂ p ⎞ ∂ ⎛ h 3 ∂p ⎞ ∂ ⎡ ⎛ V1Sx + V2Sx ⎞ ⎤ ∂ ⎡ ⎛ V1Sy + V2Sy ⎞ ⎤ ∂h ∂h ∂h


+ = h + h ⎟⎠ ⎥ − V2Tx − V2Ty +
∂x ⎜⎝ 12h ∂x ⎟⎠ ∂y ⎜⎝ 12h ∂y ⎟⎠ ∂x ⎣⎢ ⎝ 2 ⎠ ⎥⎦ ∂y ⎢⎣ ⎜⎝ 2 ⎦ ∂ x ∂y ∂t

(16.26)
Hydrodynamic Lubrication 16-9

This version of Reynolds equation is found in Cameron [6], a venerable text, and a useful reference
source of lubrication equations. Sliding of either surface in either direction is accommodated, and the
sliding velocity can vary with position. As discussed earlier, the lower surface 1 is “flat” by definition
(z = 0). The translational velocities of the upper surface V2Tx and V2Ty could represent, for example, the
motion of a bump or protuberance in the upper surface.

16.5.2 Two-Dimensional Compressible Reynolds


Equation in Cartesian Coordinates
It should first be noted that the compressible lubrication approach given later (which follows conven-
tional practice) is not a true compressible theory as generally perceived in fluid dynamics. Wave phenom-
ena, the hallmark of compressible fluid dynamics, cannot be described in the compressible lubrication
approach. This is not a serious drawback because the magnitude of the flow velocities are typically far
below the speed of sound (low Mach number) where the peculiar features of compressible flow (shock
waves and the like) do not take place. The approach is to first solve for the incompressible flow field and
then substitute this result into the incompressible continuity equation (16.1).
The development is similar to that of the incompressible case and will not be repeated here. To best
illustrate the key features, the equation is limited to the case for which there is no translation of the
upper surface, and no motion of the surfaces in the y-direction:

∂ ⎛ rh3 ∂p ⎞ ∂ ⎛ rh3 ∂p ⎞ ∂ ⎡ ⎛ V1Sx + V2Sx ⎞ ⎤ ∂rh


+ = rh ⎜ ⎟⎠ ⎥ + (16.27)
∂x ⎜⎝ 12h ∂x ⎟⎠ ∂y ⎜⎝ 12h ∂y ⎟⎠ ∂x ⎢⎣ ⎝ 2 ⎦ ∂t

The extension of this equation to the more complex geometric and kinematic conditions of Equation
16.26 should be apparent. For a perfect gas ρ = p/(RT), where R is the perfect gas constant, under isother-
mal conditions, one obtains

∂ ⎛ ph3 ∂p ⎞ ∂ ⎛ ph3 ∂p ⎞ ∂ ⎡ ⎛ V1Sx + V2Sx ⎞ ⎤ ∂ph


⎜ ⎟ + ⎜ ⎟ = ph + (16.28)
∂x ⎝ 12h ∂x ⎠ ∂x ⎝ 12h ∂x ⎠ ∂x ⎢⎣ ⎝ 2 ⎠ ⎥⎦ ∂t

These equations are found in References 3, 4, and 6. Note that this version of Reynolds equation is
nonlinear.

16.5.3 Two-Dimensional Incompressible Reynolds


Equation in Polar Coordinates
Many possible cases and versions are possible. The mathematical details of the derivation are omitted
but the result is

1 ∂ ⎛ rh3 ∂p ⎞ 1 ∂ ⎛ h3 ∂p ⎞
+
r ∂x ⎜⎝ 12h ∂x ⎟⎠ r 2 ∂q ⎜⎝ 12h ∂q ⎟⎠
V2 x cos(q) + V2 y siin(q) ∂h −V2 x sin(q) + V2 y cos(q) 1 ∂h ∂h ∂h
= + +w + (16.29)
2 ∂r 2 r ∂q ∂q ∂t

where r = x2 + y2 and θ = arctan(y/x) are the radial and angular coordinates, respectively. This equation
applies to, say, a ball spinning on its axis at constant angular velocity ω while sliding at constant velocity
on a plane.
16-10 Theory and Practice of Lubrication and Tribology

16.5.4 Two-Dimensional Incompressible Reynolds


Equation in Spherical Coordinates
Again, the mathematics to obtain this equation is complex, but parallel to those of the Cartesian equa-
tion. Likewise, a virtual infinity of possible cases and generalizations are possible:

1 ∂ ⎛ h 3 ∂p ⎞ ∂ ⎛ 2 h 3 ∂p ⎞ ∂h 1 ⎛ ∂h ∂h ⎞
⎜ sin q ⎟ + ⎜ csc q = cos q + w x ⎜ sin f + cos f cot q ⎟ (16.30)
sin q ∂q ⎝ 12R h ∂q ⎠ ∂f ⎝
2
12R2h ∂f ⎟⎠ ∂t 2 ⎝ ∂q ∂f ⎠

In the earlier equation, from Tichy and Bou-Saïd [7], the film is described between two hemispher-
ical surfaces of radius R and R + c, c « R. The spheres are positioned above the x–y plane where Z is
the distance to the point on the surface from the plane. The angle of the point from the vertical is
q = arctan( x 2 + y 2 /Z ), and the circumferential angle is ϕ = arctan(y/x). The inner sphere rotates about
the x-axis at angular velocity ωx. This is essentially the geometry and kinematics of the human hip joint
during forward walking.

16.6  Various Modifications of Reynolds Equation


The tribology and lubrication literature is replete with modifications to Reynolds equation for a num-
ber of specialized cases and applications. The topics discussed in the following sections and the list of
topics that follow are by no means comprehensive, but should give the reader a feeling for the richness
of research on lubrication theory, despite its age of more than a century. Results in this section are not
presented in sufficient detail to be used without consulting the research literature on the various topics.
For simplicity, the steady (∂h/∂t = 0), Newtonian, 1-D bearing case with the lower sliding surface is used
to illustrate the basic principles.

16.6.1  Modified Reynolds Equation for Lubricant Inertia


The reduced Reynolds number Re* can become significant (order one and greater) in certain high-speed
applications, such as gas turbine engines, and in microsystems due to relatively large thickness ratios.
There are numerous approaches to this topic (see, for example, [8–10]), but perhaps the most common
is to expand dimensionless pressure (p* = (p − pa)H2/ηV1xB) and dimensionless velocity (v ∗x = v x /Vx) in
terms of reduced Reynolds number:

p∗ = p0∗ + Re∗ p1∗ + Re∗2 p2∗ + 


v ∗x = v ∗x 0 + Re∗v ∗x1 + Re∗2v2∗x +  (16.31)

The dimensionless momentum equation (having used x* = x/B and z* = z/H) is as follows:

⎛ ∂v ∗ ∂v ∗ ⎞ dp∗ ∂2v ∗
Re∗ ⎜ v x∗ x∗ + v z∗ x∗ ⎟ = − ∗ + ∗x2 (16.32)
⎝ ∂x ∂z ⎠ dx ∂z

The expressions of Equation 16.31 are substituted in Equation 16.32, and terms gathered in powers of
Re*, following essentially the same steps in the derivation of Reynolds equation and the formalities of
regular perturbation analysis. The zero-order term p0∗ (x ∗ ) is the nondimensionalized solution to the
Hydrodynamic Lubrication 16-11

Reynolds equation of the inertialess case (Equation 16.24). Regarding that expression as a known vari-
able, the modified Reynolds equation for inertia involves the first-order pressure term p1∗ (x ∗ ):

2 2
∂ ⎛ ∗3 ∂p1∗ ⎞ ∂ ⎡⎢ ∗7 ⎛ ∂p0∗ ⎞ ⎤⎥ ∗6 ∂h ⎛ ∂p0 ⎞
∗ ∗
∂ 2 p∗ ∂h∗ ∂p∗
∗ ⎜
h ∗ ⎟
= a1 h ⎜ ∗⎟ + a2 h ∗⎜ ∗ ⎟
+ a3h∗5 ∗20 + a4h∗4 ∗ 0∗
∂x ⎝ ∂x ⎠ ∂x ⎢ ⎝ ∂x ⎠ ⎥

∂x ⎝ ∂x ⎠ ∂x ∂x ∂x
⎣ ⎦
∂h∗ ∂h∗ ∂p∗
+ a5h∗2 + a6h∗4 ∗ 0∗ (16.33)
∂x ∗
∂x ∂x

The ai are known numerical constants whose values are not given here. Note that in both Reynolds equa-
tions, the right-hand side is known, from which the left-hand-side pressure is calculated. Rather than
just ∂h*/∂x*, the right-hand side of this inertia-modified Reynolds equation depends on the inertialess
pressure and various quadratic coupled terms. In theory, the procedure can be carried out to any order,
but with increasing complexity, and diminishing returns.

16.6.2  Modified Reynolds Equation for Turbulence


Turbulence is a fluctuating fluid motion in which flow properties vary rapidly with time and position,
which is described by various fluid mechanics models. While normally not occurring in most lubrica-
tion conditions, turbulence can play a significant role in important specialized applications. The most
common type of fluid flows observed in nature is turbulent, although turbulent flow is a relative rarity
in lubrication. The Reynolds number (Re as opposed to reduced Reynolds number Re*) determines the
presence and intensity of turbulence, and represents the relative strength of the inertia to viscous forces
on the scale of the film thickness.
Turbulent flow is one in which the velocity field varies irregularly in position and in time. Although
it is generally accepted that the Navier–Stokes equations describe turbulence in a completely determin-
istic fashion, the resulting complex motion is chaotic and random, meaning it cannot be predicted from
a set of initial and boundary conditions. In an instantaneous sense, the details of turbulent flow are
always three-dimensional. Turbulence exhibits scales-within-scales (eddies). The large-scale motion is
influenced by the geometry while small-scale motion is largely independent of the geometry and rather
universal. Relative to laminar flow, turbulent motion is strongly diffusive (meaning rapid mixing) and
strongly dissipative (meaning energy is converted to heat).
Most turbulent approaches to lubrication, for example [11–13], do not attempt to predict instan-
taneous values of pressure and velocity, and such values are of little practical importance, in any
case. The dependent variables of the flow are separated into mean (time-averaged) and fluctuating
components:

p = p + pʹ, v x = v x + vʹx , etc. (16.34)


After many assumptions and simplifications, the results are finally sorted out to the semi-empirical form:

∂ ⎛ h3 ∂p ⎞ Vx ∂h
= (16.35)
∂x ⎜⎝ hkx ∂x ⎟⎠ 2 ∂x

The k x is a curved-fitted turbulent flow factor determined from the analysis:

rVx h(x )
kx = 12 + 0.530(k 2Reh )0.725 , k = 0.125Re0h.125 , Reh = ,
h
16-12 Theory and Practice of Lubrication and Tribology

16.6.3  Some Additional Examples


A partial list of other research topics in hydrodynamic lubrication includes surface roughness [14,15],
Stokes roughness [16,17], non-Newtonian lubrication [18,19], thin film compressible lubrication (molec-
ular mean free path issues) [20,21], porous media lubrication [22,23], etc. Dozens, if not hundreds, of
works on these and related topics appear each year. In most, but not all, such cases, a modified Reynolds
number, similar in form to Equation 16.35, is obtained.

References
1. Bird, R.B., Stewart, W.E., and Lightfoot, E.N. (2002) Transport Phenomena, 2nd edn., Wiley & Sons,
New York.
2. Schlichting, H. and Gersten, K. (2000) Boundary Layer Theory, 8th edn., Springer, New York.
3. Szeri, A.Z. (1998) Fluid Film Lubrication, Cambridge Press, Cambridge, U.K.
4. Hamrock, B.J. (1994) Fundamentals of Fluid Film Lubrication, McGraw-Hill, New York.
5. Reynolds, O. (1886) On the theory of lubrication and its application to Mr. Beauchamp Tower’s
experiments, including an experimental determination of the viscosity of olive oil, Philos. Trans.
R. Soc., 177, 157–234.
6. Cameron, A. (1966) Principles of Lubrication, Wiley, New York.
7. Tichy, J. and Bou-Saïd, B. (2008) The Phan-Thien and Tanner model applied to thin film spherical
coordinates: Applications for the lubrication of hip joint replacement, ASME J. Biomech. Eng., 130(4),
021012.1–021012.6.
8. Reinhardt, E. and Lund, J.W. (1975) Influence of fluid inertia on the dynamic properties of journal
bearings, ASME J. Lubr. Tech., 97(2), 159–157.
9. Hill, D.L., Baskharone, E.A., and San Andres, L. (1994) Inertia effect in a hybrid bearing with a 45
degree entrance region, ASME J. Tribology, 117(3), 498–505.
10. Marrero, V., Borca-Tasciuc, D.-A., and Tichy, J. (2010) On squeeze film damping in microsystems,
ASME J. Tribol., 132(3), 031701–031707.
11. Elrod, H. and Ng, C.W. (1967) A theory for turbulent films and its application to bearings, ASME
J. Lubr. Tech., 89, 346–362.
12. Constantinescu, V.N., Smith, R.N., and Galetuse, S. (1980) Influence of inertia forces in film rupture
in laminar and turbulent lubrication, STLE Trans., 23(1), 61–69.
13. Brunetière, N. and Tournerie, B. (2009) Study of hydrostatic mechanical face seals operating in a
turbulent rough flow regime, ASME J. Tribol., 110(2), 314–321.
14. Patir, E. and Cheng, H. (1979) Application of average flow model to lubrication between rough sur-
faces, ASME J. Lubr. Tech., 101(2), 220–230.
15. Minet, C., Brunetière, N., Tournerie, B., and Fribourg, D. (2010) Analysis and modeling of the topog-
raphy of mechanical seal faces, STLE Tribol. Trans., 53(6), 799–815.
16. Sun, D.-C. and Chen, K.-K. (1977) First effects of stokes roughness on hydrodynamic lubrication,
ASME J. Lubr. Tech., 99(1), 2–9.
17. De Kraker, A., Van Ostayen, R.A.J., and Rixen, D.J. (2010) Development of a texture averaged
Reynolds equation, Tribol. Int., 40(1), 2100–2109.
18. Racou, A., Rajagapol, K.R., and Szeri, A. (1987) Flow of a fluid of the differential type in a journal
bearing, ASME J. Tribol., 109(1), 100–108.
19. Boucherit, H., Bou-Saïd, B., and Tichy, J. (2010) Comparison of non-Newtonian constitutive laws in
hydrodynamic lubrication, Tribol. Lett., 43(11), 49–57.
20. Zhang, W.-M., Meng, G., Zhou, J.-B., and Chen, J.-Y. (2010) Slip model for the ultra-thin gas-lubricated
slider bearings of an electrostatic micromotor in MEMS, Microsyst. Technol., 15(6), 953–961.
Hydrodynamic Lubrication 16-13

21. Chen, D. and Bogy, D.B. (2010) Comparison of slip corrected Reynolds lubrication equations for the
air bearing film in the head-disk interface of hard disk drives, Tribol. Lett., 37(2), 191–201.
22. Lin, J.R. (1996) Viscous shear effects on the squeeze film behavior in porous circular discs, Int. J.
Mech. Sci., 38(4), 373–384.
23. Nabhani, M., El Khlifi, M., and Bou-Saïd, B. (2010) A general model for porous medium flow in
squeezing film situations, Lubr. Sci., 22(2), 37–52.
17
Compressible Gas
Film Lubrication
Nomenclature.............................................................................................. 17-1
Subscripts..................................................................................................... 17-2
Acronyms..................................................................................................... 17-2
17.1 Introduction..................................................................................... 17-3
17.2 Fundamentals of Gas Film Lubrication Analysis....................... 17-5
17.3 Performance of One-Dimensional Slider Gas Bearings............ 17-8
17.4 Plain Cylindrical Gas Journal Bearings..................................... 17-12
Thin Film Flow Analysis for Plain Cylindrical
Bearings  •  Performance of a Plain Cylindrical Gas Journal
Bearing  •  Bearing Force Coefficients and Dynamic Stability
17.5 Introduction to Flexure Pivot Bearings..................................... 17-19
17.6 Introduction to Foil Bearings...................................................... 17-19
Luis San Andrés 17.7 Recommendations for Oil-Free Rotating Machinery.............. 17-20
Texas A&M University References................................................................................................... 17-21

Nomenclature
B Bearing width [m]
c Radial clearance in journal bearing [m]
C αβ Damping coefficients [N·s/m]; αβ = X,Y. C – αβ = Cαβ/C*
3
D⎛L⎞
C* m ⎜ ⎟ . Factor for damping coefficient in radial bearing
4⎝c⎠
– C
C 3 . Dimensionless damping coefficient (slider bearing)
⎛ L* ⎞
12 m B ⎜ h ⎟
⎝ *⎠
D Journal or rotor diameter [m]
eX, eY Components of journal eccentricity vector [m]. ε = e/c
FX, FY Components of bearing reaction force [N]
f Torque/cW. Drag friction coefficient in journal bearing
h Film thickness [m]. H = h/h* or h/c
Kαβ Damping coefficients [N·s/m]; αβ = X,Y. K –αβ = Kαβ/K*
K* C*Ω. Factor for stiffness coefficient in radial bearing
– K
K . Dimensionless stiffness coefficient (slider bearing)
BLpa h*
Kn (λ/h). Knudsen number. >15 for continuum flow
L Length of bearing [m]

17-1
17-2 Theory and Practice of Lubrication and Tribology

M Rigid rotor mass [kg]


Mc Critical mass of rigid rotor-bearing system [kg]
N Rotational speed [rev/s]
p Pressure [Pa]. P = p/pa
pa, pS Ambient and supply pressures [Pa]
R ½D. Journal radius
Re ρUh/μ. Shear flow Reynolds number
ℜg Gas constant [J/kg·K]
2
mN LD ⎛ R ⎞
S Sommerfeld number. S =
W ⎜⎝ c ⎟⎠
s s = λ±iϖ. Eigenvalue of characteristic equation
t Time [s]
T Temperature [K]
Torque Drag torque [Nm]
U Surface speed [m/s]. ΩR in journal bearing
W Load [N]. w = W/(BLpa) or W/(LDpa)
WFR (ϖ/Ω). Whirl frequency ratio
X,Y Inertial coordinate system for journal bearing analysis
x, y, z Coordinate system in plane of bearing
z {x(t),y(t)}T. Vector of journal center dynamic displacements [m]
Z Complex impedance [N/m]; Z = (K + iωC), i = −1
α (h1/h2). Ratio of inlet to outlet film thickness in slider bearing
γ (L2/L). Ratio of lengths in Rayleigh step and tapered-flat slider bearings
ϕ Attitude angle (between load vector and journal eccentricity vector) [degree]
τ ωt. Dimensionless time
ε (e/c). Journal eccentricity ratio
Θ x/R. Circumferential coordinate fixed to stationary
λ Gas molecular free path [m]
2
6 mUL* 6 mΩ ⎛ R ⎞
Λ Speed number. Λ = , Λ =
pah*2 pa ⎜⎝ c ⎟⎠
μ Gas viscosity [Pa·s]
ρ Gas density [kg/m3]
12mw L2* w
σ Frequency number. s = =Λ 1
pah*2 2Ω
ω Frequency of dynamic motions [rad/s]
ϖ Whirl frequency of unstable dynamic motions [rad/s]
Ω (2πN). Rotor or journal speed [rad/s]

Subscripts
* Characteristic value
u Ultimate limit at Λ → ∞

Acronyms
GFB Gas foil bearing
RBS Rotor-bearing system
Compressible Gas Film Lubrication 17-3

17.1  Introduction
In a gas bearing, a film of gas, hereby referred to as the lubricant, separates the rotating component
(e.g., a journal) from the stationary part (a housing or stator). Gas film bearings offer advantages of low
friction and reduced heat generation. These advantages enable their successful applications in air-cycle
units for airplane cabins, high-precision instruments, auxiliary power units, and high-speed microtur-
bomachinery* (MTM). Gas-bearing systems do not require costly, complex sealing, and lubricant cir-
culation systems, hence ensuring system compactness, low weight, and extreme temperature operation.
Furthermore, these bearings eliminate process fluid contamination and are environmental-friendly.
However, their excessive cost, protected technology, and lack of calibrated predictive tools have pre-
vented widespread use in mass-produced applications.
Gas bearings have a low load-carrying capacity and require a minute film thickness to accomplish their
intended function. Thus, their fabrication and installation tends to be expensive and time-consuming.
Another disadvantage is poor damping because of the inherently low viscosity of the gas.
The literature on the analyses of gas-bearing analyses is extensive, albeit experimental verification
and successful commercial implementations have not always been reported. Gross [1] covers the funda-
mentals of analysis that span the fast development of gas-bearing technology in the 1960s. Pan [2] gives
a serious description of the analysis and performance of (rigid surface type) gas bearings as of 1980.
Hamrock [3] details analyses for the static load performance of both thrust and radial gas bearings,
and Czolczynski [4] reviews the analyses for prediction of frequency-dependent force coefficients in gas
bearings.
The last decade (2000s) has seen a rebirth of gas bearings, in particular, gas foil bearings (GFBs) for
MTM [5] and aerostatic gas bearings for spindle machines [6]. San Andrés and coworkers [7–16] report
the results of a comprehensive research program, experimental and analytical, evaluating and develop-
ing simple and cost-effective gas bearings for MTM.
Bearings in rotating machinery are of two types: (1) radial bearings supporting lateral loads including
rotor weight and (2) thrust bearings carrying axial loads. See Figure 17.1 for a few relevant gas-bearing
configurations. These loads can be either static or dynamic or both. Gas bearings behave as mechani-
cal elements that provide stiffness, damping, and inertia force coefficients that, in conjunction with
the structural parameters of a rotor, determine the stability and dynamic behavior of the entire rotor-
bearing system (RBS).
In gas bearings, hydrodynamic shear action from the moving component (sliding or rotating) enables
the generation of the lubricant wedge where a hydrodynamic pressure evolves to produce the reac-
tion force opposing the externally applied load. Gas bearings operating under the hydrodynamic (self-
acting) principle are, in general, of simple construction although at times difficult to manufacture and
install because of the required minute film clearances. Other bearings employ external pressurization
supplied through restrictors (orifices, slots, or capillaries) to enable a hydrostatic action that separates
the surfaces thus inducing journal or rotor lift without rotation, for example. Hydrostatic bearings are
mechanically more complex than hydrodynamic bearings because of their additional supply ports;
albeit their major advantage is in applications without rotor spinning. This advantage must be weighed
against the extra cost plus the need of an external source of pressurized gas. More importantly, in a
hybrid-bearing configuration, that is, one where both hydrostatic and hydrodynamic operating prin-
ciples act jointly, the external supply pressure is typically used to promote early rotor lift off thus reduc-
ing temporary rubs, avoiding wear of surfaces, and extending bearing life.
There are (probably) as many types of gas-bearing configurations as there are applications; that is, a
gas bearing is selected to fulfill certain functions while keeping a cost low, including component fabrica-
tion and installation, and of course, operation. The archival literature features successful applications

* As per the International Gas Turbine Institute (IGTI), a microturbomachinery has power <500 kW.
17-4 Theory and Practice of Lubrication and Tribology

Flexure pivot tilting pad


Spiral grooved thrust and radial bearings bearing

Bump-type foil bearing Flexure pivot tilting pad


bearing with hydrostatic
pressurization

Metal mesh foil bearing Overleaf-type foil bearing

FIGURE 17.1  Typical commercial gas bearings for microturbomachinery.

of gas bearings; often failing to notice that, in contrast to liquid lubricated bearings, gas bearings have
inherent limitations that prevent their widespread usage as load support elements in (heavy) commer-
cial machinery.
Gases, although chemically more stable than liquids, have an inherent low viscosity—one or
two orders of magnitude lower than that of mineral oils, for example. Recall that the load-carrying
capacity (W) of a self-acting hydrodynamic film bearing is roughly proportional to mAU hmin ( 2
) [3],
where μ is the lubricant viscosity, U is the surface speed, A is the area of action, and hmin is the minimum
film thickness. Hence, in order to achieve a desired load capacity, a gas-lubricated bearing replacing
a similar size oil-lubricated bearing must operate at an exceedingly high surface speed (U) or with a
minute film thickness (h). That is, hydrodynamic gas bearings are not intended for supporting rotating
machinery that operates with relatively low surface speeds or if the film clearance or gap is too large.
Hence, the need for accurate manufacturing of parts which both increases cost and makes installation
complicated. Of course, externally pressurized (aerostatic) gas bearings can be used efficiently to carry
loads at low or even zero surface speeds. However, aerostatic bearings require a source of pressurized gas
which adds cost and complexity [1,6,10].
To enhance the hydrodynamic action, designers have produced a number of bearing configurations
that exploit geometrical features such as steps, grooves, pockets, and dimples, for example. Figure 17.1
Compressible Gas Film Lubrication 17-5

shows several typical commercial gas-bearing configurations. The bearing types with textured surfaces,
known as (spiral) grooved bearings and herringbone journal bearings, have been instrumental to the
operation of gyroscopes for aircraft and satellite navigation [2], enabled noncontacting gas face seal
technology [7], and more recently, allowed the revolution in digital storage hard-drive technology [17].
In these applications, static and dynamic loads are relatively low. Do note that, for optimum load per-
formance giving the maximum static (centering) stiffness, the depth of the machined steps or grooves
or pockets is just equal or a little larger than the operating film gap or clearance, as will be demonstrated
later. Until recently, these geometrical features were difficult to machine at low cost, except in certain
materials like silicon carbide for noncontacting face seals. However, current casting and manufactur-
ing processes allow the manufacturing of these bearings (or seals) at a relatively low cost and with near
identical performance in one or millions of pieces.
Other radial bearing configurations of interest, that is, undergoing close scrutiny and commercial
development, include bump-type foil bearings [5,15,18], flexure pivot tilting pad bearings [13], and (low
cost) metal mesh foil bearings [19].

17.2  Fundamentals of Gas Film Lubrication Analysis


The fluid flow in a hydrodynamic gas bearing or gas face seal is typically laminar and inertialess, that
is, the Reynolds numbers Re = ρUh/μ < 1, because of the smallness in film thickness (h) and the low
lubricant density (ρ). Gas annular seals, such as labyrinth and honeycomb types, are notable exceptions,
since in these applications large pressure drops, high surface speeds, and large clearances promote flow
turbulence accompanied by strong fluid compressibility effects [20].
Consider, as shown in Figure 17.2, the flow of an ideal gas in a region confined between two surfaces
separated by the small gap h. The top surface has velocity U along the x-direction. For an isothermal
p
process, the gas density (ρ) and pressure (p) are related by r = , with ℜg and T representing the
ℜg ⋅T
gas constant and operating temperature, respectively.
Table 17.1 shows a list of the physical properties of the most common gases used as lubri-
cants in thin film bearings. The gas viscosity (μ) increases with its absolute temperature (T) as
⎛⎛ T ⎞ ⎛ T ⎞⎞ T
m = mo ⎜ ⎜ 1 + * ⎟ ⎜ 1 + * ⎟ ⎟ where T* and To are reference temperatures and μo = μ(To) [22].
⎝ ⎝ To ⎠ ⎝ T ⎠ ⎠ To
Reynolds equation describes the generation of the film pressure within the flow region [2]:

 ⎛ h3 p  ⎞ U ∂
∇⋅⎜ ⋅ ∇ p⎟ = ( ph) + ∂∂t ( ph) (17.1)
⎝ 12m ⎠ 2 ∂x

Equation 17.1 represents an isoviscous gas and flow without fluid inertia effects. Furthermore, the
derivation of Equation 17.1 assumes that the gas satisfies the no-slip condition, that is, it adheres to

Y V = dh/dt + U dh/dx

Gas lubricant Lz
X
h(x,z,t)

Z Lx
Film thickness h << Lx, Lz

FIGURE 17.2  Geometry of a gas-lubricated thin film bearing.


17-6 Theory and Practice of Lubrication and Tribology

TABLE 17.1  Viscosity and Molecular Weight of Gases Used in


Thin Film Bearings
Molecular
Gas Formula Weight μo, μPa·s To, K T *, K
Acetylene C2H2 26.036 10.2 293 198
Air O2 + N 29.000 17.1 273 124
Ammonia NH3 17.034 9.82 293 626
Argon Ar 39.950 22.04 289 142
Carbon dioxide CO2 44.010 13.66 273 274
Carbon monoxide CO 28.010 16.65 273 101
Chlorine Cl2 70.900 12.94 289 351
Chloride HCl 36.458 13.32 273 360
Helium He 4.003 18.6 273 38
Hydrogen H2 2.016 8.5 273 83
Hydrogen sulfide H2S 34.086 12.51 290 331
Methane CH4 16.042 10.94 290 198
Neon Ne 20.180 29.73 273 56
Nitrogen N2 28.020 16.65 273 103
Nitric oxide NO 30.010 17.97 273 162
Nitrous oxide N2O 44.020 13.66 273 274
Oxygen O2 32.000 19.2 273 138
Steam H2O 18.016 12.55 372 673
Sulfur dioxide SO2 64.070 11.68 273 416
Xenon Xe 131.300 21.01 273 220
Source: http://periodic.lanl.gov/default.htm
Note: Gas constant Rg = (8.314 34 J/kg·K)/MW.

the surfaces.* As a boundary condition, the pressure is typically ambient (pa) on the boundary of the
domain.
The gas film Reynolds equation is nonlinear, and hence exact solutions exist for a handful of limiting
conditions [2,3]. The left-hand side of Equation 17.1 is elliptic in character, while the terms on the right-
hand side are known as the shear-induced flow and squeeze film flow terms.
It is convenient to normalize Equation 17.1 in terms of dimensionless variables and parameters. To
this end, let

x z h p
x= ; z = ; t = w t; H = ; P = (17.2)
L* L* h* pa

where
L* is a characteristic length of the bearing surfaces
h* is a characteristic film thickness, typically the minimum film thickness or the clearance (c) in a
radial bearing

Above ω denotes an excitation whirl frequency representative of unsteady or time transient effects. With
the definitions given, Reynolds equation becomes in dimensionless form:

* As the film thickness (h) decreases into the nanometer scale, its size approaches that of the gas molecular free path
(λ = 60 nm for air under standard conditions), and hence, slipping effects become significant. Magnetic recording and
digital hard drive applications fall within this category. The Knudsen number (Kn = λ/h) aids to distinguish the flow
regime of operation; Kn > 15 denotes molecular flow; 0.01< Kn < 15 represents slip flow; and Kn < 0.01 gives a continuum
flow, as in the applications discussed herein [21].
Compressible Gas Film Lubrication 17-7

∂ ⎛ 3 ∂P ⎞ ∂ ⎛ 3 ∂P ⎞ ∂ ∂
⎜ PH
∂x ⎝
⎟+
∂x ⎠ ∂z
⎜⎝ PH ⎟ =Λ
∂z ⎠ ∂x
( PH ) + s ∂t ( PH ) (17.3)

where

6 mUL* 12 mw L2*
Λ= and s = (17.4)
pah*2 pah*2

are known as the speed number and the frequency number, respectively [2]. Both parameters represent
the influence of fluid compressibility on the performance of the gas bearing. For Λ and σ small, typically
<1, the gas bearing operates as an incompressible fluid film bearing, as seen next.
For steady-state conditions (σ = 0), and operation at low speeds, Λ ≪ 1, an expansion of the dimen-

sionless pressure as P = 1 + Λ P gives the simplified Reynolds equation [2]:

∂ ⎛ 3 ∂P ⎞ ∂ ⎛ 3 ∂P ) ⎞ ∂
⎜H
∂x ⎝
⎟+
∂x ⎠ ∂z
⎜H ⎟=
∂z ⎠ ∂x
(H ) (17.5)

which is formally identical to the Reynolds equation for an incompressible lubricant. Hence, its solution
can be easily sought—analytically for either the short length or long journal bearings, or using numeri-
cal schemes for finite length bearings of any geometry.
Clearly, the assumed solution is strictly valid for Λ → 0. Hence, the pressure field cannot be much
higher than ambient pressure (pa), and consequently, the bearing load capacity is also small albeit pro-
portional to the speed number, that is, it increases linearly with surface speed (U), for example. Note
that the dimensionless pressure P = P − 1 = (
p − pa ) h*2
as is typical in mineral oil–lubricated bearings.
L 6mUL*
Analytical solutions to Equation 17.5 are available for either the short length or infinitely long cylin-
drical journal bearings, for example. Closed-form solutions are also available for simple 1-D slider or
Rayleigh step bearing geometries (see References 2 and 21).
On the other hand, for large speed numbers, in the limit as Λ → ∞, Equation 17.3 reduces to*

∂ h
( PH ) = 0 ⇒ p = pa b (17.6)
∂x h(x )

where hb is the film thickness at the boundary where the pressure is ambient. The limiting speed solution
(Equation 17.6) shows that the pressure within the film is bounded and independent of the surface speed
U. This result is in opposition to that in incompressible fluid bearings where the generated hydrody-
namic film pressure is proportional to the surface speed U. Since the pressure has a definite limit, it also
means that the bearing load capacity has also a limit, that is, an ultimate value. In this regard, gas film
bearings do show a significant difference with incompressible fluid (mineral oil–lubricated) bearings
whose (theoretical) load capacity increases with surface speed.
Closed-form solutions for finite speed numbers (Λ) are not readily available. Hence, predictions of
bearing film pressure and its force reaction supporting an applied load must rely on numerical analysis.
For low to moderate speed numbers, finite differences or finite element methods applicable to ellipti-
cal differential equations are quite adequate. See Reference 8 for an advancement resolving the issue of
pressure oscillations and numerical instability for operation at large speed numbers (Λ).

* The PH solution is an inner field which must be matched to an outer (boundary) solution satisfying the side pressure
condition (P = 1) [2]. For the purposes of this review, the PH solution is adequate.
17-8 Theory and Practice of Lubrication and Tribology

L
L = L1 + L2 L1 L2

Taper
W Step or ridge W
load Gas lubricant
load

h1 Land
h1 y h(x) Inlet h(x)
h2 h2
x
U Moving surface U

(a) (b)

L = L1 + L2 L1 L2
Width B>>L Taper
W
load
Exit
Flat
h1 h(x) h2

U Moving surface

(c)

FIGURE 17.3  Schematic view of (simple) one-dimensional slider bearings. (a) Tapered bearing; (b) Rayleish step
bearing; (c) taper-flat bearing.

17.3  Performance of One-Dimensional Slider Gas Bearings


Consider, as shown in Figure 17.3, three typical 1-D* slider-bearing configurations: tapered, Rayleigh
step, and tapered flat. In these configurations, the width (B) ≫ length (L), and thus the hydrodynamic
pressure does not vary along the z-axis. The bearing peak pressure and maximum load capacity are a
function of the ratio between the inlet film thickness (h1) and the exit film thickness (h2) and the extent
of the step or tapered length (L1).
Integration of the pressure field over the bearing surface gives the reaction force that opposes the
applied load (W):

L 1
W
W =B
∫ ( p − pa )dx or w =
B L pa
=
∫ ( P − 1)dx (17.7)
0 0

For small speed numbers Λ < 1 (incompressible fluid), Table 17.2 shows closed-form expressions for the
peak hydrodynamic pressure and the bearing load as a function of the film thickness ratio α = h1/h2 and the
land-to-length ratio γ = L2/L in a Rayleigh step bearing [22]. Simple calculations show that the maximum
load w– requires thickness ratios on the order of two, that is, α = 2.189 for a tapered bearing (w
– = 0.0267),
max

and α = 1.843 for a step bearing with γ = 0.30 (w = 0.034). Hence, the taper height difference or the step
height (h1–h2) is of similar size as the minimum film thickness (h2). In gas bearings, the smallness of the
film thickness required to support realistic loads also poses a difficulty in manufacturing mechanical fea-
tures such as ridges and steps. Furthermore, manufacturing processes must ensure a surface roughness
(RMS value) at least one order of magnitude (∼1/10) lower than the minimum film thickness [21].

* In this case, the bearing width (B) is much longer than its length (L), and hence, the film pressure is only a function of
∂P ∂P
the coordinate (x). The analysis calls for  .
∂z ∂x
Compressible Gas Film Lubrication 17-9

TABLE 17.2  Closed-Form Expressions for Peak Hydrodynamic


Pressure and Load in One-Dimensional Tapered Bearing and
Rayleigh Step Bearing
Tapered Bearing Step Bearing

Peak pressure
p=
( p − pa ) h
2
2 (a − 1) (a − 1) g
6mUL 4a (1 + a ) ga 3
1+
1− g

Load
w=
Wh22 1 ⎡ (1 − a ) ⎤⎥ g (a − 1)
⎢ ln(a ) + 2
6mUBL2 2
(1 + a ) ⎥⎦ 2 ga 3
(1 − a ) ⎢⎣ 1+
1− g

Source: Williams, J.A., Engineering Tribology, Chap. 7, Oxford Science


Publications, Oxford, Great Britain, 2000.
Note: Low-speed operation (incompressible fluid approximation)
h L
( )
Λ = 6 mUL pa h22 < 1, a = 1 h , g = 2 L .
2

An example of gas-bearing performance follows. Predictions are obtained for a film thickness ratio
α = h1/h2 = 2.2 and length ratio γ = L2/L = 0.30 for the Rayleigh step and tapered-flat bearings. The param-
eters used are close to those delivering a maximum reaction force (load capacity) in an incompressible
lubricant slider bearing.
Figure 17.4 depicts for increasing speed numbers (Λ) the evolution of the hydrodynamic pressure field
versus the coordinate (x/L). Note that the peak pressure displaces toward the minimum film location as
Λ increases. Most important is to realize that the peak pressure is not proportional to the speed, as is the
case in incompressible lubricant bearings. The largest peak pressure cannot exceed that of the limit at
high speeds, that is, pmax Λ→∞ =a  = h1/h2. This feature may entice designers to implement or promote high
aspect ratios for the film thicknesses, α ≫ 1. However, too large inlet/exit film ratios (α ≫ 1) will cause
the gas flow to choke at the bearing exit plane. This is an undesirable operating condition that produces
noise and shock wave instabilities and could cause severe mechanical damage [2].
Figure 17.5 depicts the (dimensionless) load (w = W/BLpa) versus speed number (Λ) for the three slider
bearings. Note that at low speeds, Λ < 5 typically, the load capacity is proportional to the speed number.
However, as Λ increases, the load reaches an asymptotic magnitude. It is important to note that knowl-

edge derived from incompressible lubrication theory does not extend to gas lubrication theory. For exam-
ple, the selected Rayleigh step configuration offers the largest load at small speed numbers, that is, in the
incompressible fluid flow region. However, as evidenced in the predictions, at the highest speed numbers
(Λ > 100), the Rayleigh step bearing produces the smallest load albeit it shows the largest hydrodynamic
peak pressure. Note that, see Figure 17.4, in the step bearing the region of pressure generation is confined
to the film land with small thickness (h2); while the rest of the bearing is basically at ambient pressure.
Fluid film bearings support both static and dynamic loads. Thus far, the analysis has focused on the
static load capacity. Consider a bearing that undergoes motions of small amplitude (Δy) and frequency
(ω) about an equilibrium condition with film thickness ho(x). Dynamic force coefficients, stiffness (K)
and damping (C), follow from small amplitude motions with frequency (ω.) about the equilibrium or
static conditions.* Unlike bearings lubricated with incompressible fluids, the stiffness and damping
force coefficients of gas bearings are strong functions of frequency [2,4,23]. In particular, for high speeds
and high-frequency operation (Λ → ∞, σ → ∞)C → 0, that is, damping is lost. Thus, gas bearings need to

* See Lund [23] for the original and most elegant description of the analytical perturbation method for calculation of
dynamic force coefficients in gas bearings.
17-10 Theory and Practice of Lubrication and Tribology

2.4 2.4 Λ = inf


Speed number Λ=inf Speed number
10 10 L2/L = 0.3
2.2 2.2 h2
20 h2 Λ=500 20 h1 Λ = 500
h1 100
100 U
Film pressure (P/Pa)

Film pressure (P/Pa)


2 500 U Λ=100 2 500
Λ = 100
Infinite Infinite

1.8 h1 1.8 h1
α= =2.18 α= =2.2 Λ = 20
h2 Λ = 20 h2
1.6 1.6
Λ = 10
1.4 1.4

1.2 1.2
Λ = 10
1 1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
(a) Coordinate x/L (b) Coordinate x/L

2.4
Speed number Λ = inf
2.2 10
L2/L=0.3 Λ = 500
20 h2
100 h1
Film pressure (P/Pa)

2 500 U Λ = 100
Infinite
Λ = 20
1.8 h1
α= =2.2
h2
1.6

1.4
Λ = 10
1.2

1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
(c) Coordinate x/L

FIGURE 17.4  Pressure field in one-dimensional slider bearings (tapered (a), Rayleigh step (b), and tapered flat (c))
for increasing speed numbers (α = 2.2, γ = 0.3).

0.7
Tapered bearing
Step bearing L2/L = 0.3
0.6 Tapered flat L2/L = 0.3
Normalized load capacity (–)

L2/L = 0.3
h2
h1 h1
0.5 α= =2.2
h2 U

0.4

0.3 L2/L = 0.3


h1 h2

0.2 U

0.1 h1
h2
U
0
1 10 100 1000
Speed number (Λ)

FIGURE 17.5  Load capacity (w) in one-dimensional slider gas bearings versus speed number (Λ) (α = 2.2, γ = 0.3).
Compressible Gas Film Lubrication 17-11


K =–
K
BLpa/h*
– 1.6
(K) Sp # = 1
Sp # = 5 Λ = 100
1.4 Sp # = 10
Sp # = 20
1.2 Sp # = 100

Stiffness coefficient
Sp # = 500
1
Λ = 20
0.8

0.6
Λ = 500
0.4 Λ = 10
L2/L = 0.3
Λ=5 h2
0.2 h1
Λ=1
U
0
1 10 100 1000
Frequency number (σ)

FIGURE 17.6  Stiffness coefficient for 1D tapered-flat gas bearing versus frequency number (σ) and increasing
speed numbers (Λ) (α = 2.2, γ = 0.3).

3.00E – 02
C Sp # = 1
(C) Λ=1 C= 3 Sp # = 5
2.50E – 02 12µ B L* Sp # = 10
h*
Sp # = 20
Damping coefficient

Sp # = 100
2.00E – 02 Λ = 5 Sp # = 500

1.50E – 02 L2/L = 0.3


Λ = 10 h1
h2

U
1.00E – 02 Λ = 20
Λ = 100
5.00E – 03
Λ = 500
0.00E + 00
1 10 100 1000
Frequency number (σ)

FIGURE 17.7  Damping coefficient for 1D tapered-flat gas bearing versus frequency number (σ) and increasing
speed numbers (Λ) (α = 2.2, γ = 0.3).

be used with great caution for applications that require mechanical energy dissipation characteristics to
ameliorate or reduce vibrations of the mechanical system.
For the tapered-flat slider with α = h1/h2 = 2.2 and length ratio γ = L2/L = 0.30, Figures 17.6 and 17.7
– –
depict the dimensionless stiffness (K) and damping (C) coefficients versus increasing frequency num-
bers (σ) and various speed parameters (Λ). Note two important dynamic force performance features:
(1) the bearing stiffness rises rapidly with frequency, a typical hardening effect of gas bearings and (2)
damping decreases quickly, as expected.* It is also important to realize that, at low frequencies (σ ∼ 1),
the (nearly static) stiffness reaches a maximum at a certain speed (Λ > 50), not increasing further with
sliding speed. This is also expected since, as shown in Figure 17.5, the load capacity also reaches its ulti-
mate limit for operation at Λ > 50.

* Negative damping coefficients are not unusual in stepped gas bearings such as in spiral grooved or herringbone grooved
configurations, see Reference 7.
17-12 Theory and Practice of Lubrication and Tribology

17.4  Plain Cylindrical Gas Journal Bearings


Cylindrical hydrodynamic bearings support radial (or lateral) loads in rotating machinery. Figure
17.1 depicts a few commercial gas-bearing configurations. Using gas as the lubricant in the fluid film
bearing offers distinct advantages such as lesser number of parts, avoidance of mineral oils with lesser
contamination, and most importantly, little drag friction (minute power losses) and the ability to oper-
ate at extreme conditions in temperature, high or low.
In hydrostatic or hybrid (hydrostatic/hydrodynamic) bearings, pressurized gas flowing through
restrictor ports creates a centering stiffness and nearly journal-centered operation. The external pres-
surization aids to promote early journal liftoff and reduces the possibility of “hard landings” or transient
rubs that cause wear of surfaces. Hydrostatic operation enables gas bearings with relatively large clear-
ances, hence reducing manufacturing costs and difficulties associated with installation [10].
Disadvantages in gas bearings stem from two types of instabilities [2]: pneumatic hammer con-
trolled by the flow versus pressure lag in the pressurized gas feeding system and hydrodynamic insta-
bility, a self-excited motion characterized by subsynchronous (forward) whirl motions. Proper design
of a hybrid-bearing system minimizes these two kinds of instabilities.* Gas-bearing design guidelines
available since 1967 [24] dictate that, to avoid or delay a pneumatic hammer instability, externally pres-
surized gas bearings have restrictors impinging directly into the film lands, that is, without any (deep)
pockets or recesses.
The analysis herein does not discuss textured or etched bearings, that is, ones with herringbone
grooves, for example. See Reference 7 for the appropriate analyses and predictions. The textured
bearings are still costly to manufacture, offer little improvements in load capacity, and have severe
limitations in rotordynamic stability [12].

17.4.1  Thin Film Flow Analysis for Plain Cylindrical Bearings [10]
In a plain cylindrical radial bearing (see Figure 17.8), the journal rotates at speed (Ω) and (eX,eY) denote
the journal displacements within the bearing clearance (c). The dimensionless film thickness is

h
H= =1 + eX cos Θ + eY sin Θ (17.8)
c

Θ
W, static load

θ Ω rotor speed

Bearing center
Bearing
Lubricant Y ey
Journal ex Journal
center

e, Journal
h, film thickness eccentricity
h = c+ eX cosΘ + ey sinΘ X φ, attitude angle

FIGURE 17.8  Geometry of plain cylindrical gas bearing, coordinate system, and nomenclature.

* A self-excited instability means that a change in the equilibrium or initial state (position and/or velocity) of the RBS
leads to a permanent departure with increasing amplitudes of motion at a certain frequency, usually a natural fre-
quency. A self-excited instability does not rely on external forces (load condition), including mass imbalance, for its
manifestation.
Compressible Gas Film Lubrication 17-13

Reynolds equation for the laminar flow of an ideal gas and under isothermal conditions governs the
generation of hydrodynamic pressure within the thin film region, that is [2],

∂ ⎛ 3 ∂P ⎞ ∂ ⎛ 3 ∂P ⎞ ∂ ∂
⎜ PH
∂Θ ⎝
⎟ +
∂Θ ⎠ ∂z
⎜⎝ PH ⎟ =Λ
∂z ⎠ ∂Θ
( PH ) + s ∂t ( PH ) (17.9)

where

2 2
6 mΩ ⎛ R ⎞ 12 mw ⎛ R ⎞ w
Λ= and s = =Λ 1 (17.10)
pa ⎜⎝ c ⎟⎠ pa ⎜⎝ c ⎟⎠ Ω
2

are the well-known speed and frequency numbers, respectively. An applied external static load (Wo)
determines the journal center to displace eccentrically to the equilibrium position (eX,eY)o with steady
pressure field po and film thickness ho. Radial bearings are regarded as two degrees of freedom mechani-
cal elements with lateral forces reacting to radial displacements (x, y). Bearing rotordynamic force
coefficients are, by definition, changes in reaction forces due to small amplitude motions about an equi-
librium position. The linearized force reaction model for a gas bearing is

⎧FX ⎫ ⎡ K XX K XY ⎤ ⎧ x ⎫ ⎡C XX C XY ⎤ ⎧ x ⎫
⎨F ⎬ = − ⎢ K − (17.11)
⎩ Y⎭ ⎣ YX K YY ⎥⎦ ⎨⎩ y ⎬⎭ ⎢⎣ CYX CYY ⎥⎦ ⎨⎩ y ⎬⎭

Gas bearings (rigid surfaces, tilting pads, and foil types) have frequency-dependent force coefficients
because of the fluid compressibility and the compliance of the bearing pad surfaces. The dependency on
frequency cannot be overlooked. See References 9 and 10 for details on the modeling of cylindrical and
tilting pad hybrid (hydrostatic /hydrodynamic) gas bearings.
Reynolds equation (17.9) is nonlinear because the gas density depends on the pressure. In the case
of aerostatic or pure hydrostatic bearings carrying only static loads, the equation becomes linear,
that is,

 ⎛ −h 3  ⎞
∇⋅⎜ ⋅ ∇p 2 ⎟ = 0 (17.12)
⎝ 24 m ⎠

This equation can be solved efficiently for (p2) as the independent variable with either central finite dif-
ferences or finite element methods.
At large speed numbers or frequency numbers (Λ ≫ 1, σ  ≫ 1), the first-order terms on the right-hand
side of Equation 17.9 dominate the generation of the hydrodynamic pressure in the gas film region. For

low rotational speeds (Ω) and low frequencies, (Λ, σ → 0), the expansion P ≃ 1 + Λ P gives the linear
Reynolds equation:

∂ ⎛ 3 ∂P ⎞ ∂ ⎛ 3 ∂P ⎞ ∂H w ∂H
⎜H ⎟+ ⎜⎝ H ⎟= + (17.13)
∂Θ ⎝ ∂Θ ⎠ ∂z ∂z ⎠ ∂Θ 12 Ω ∂t

This equation is formally identical to the Reynolds equation for a thin film with an incompressible fluid.
The numerical solution of the linear equation earlier can be easily performed using (central) finite dif-
ferences, for example. More importantly, any computational tool predicting pressure fields for bearings
17-14 Theory and Practice of Lubrication and Tribology

lubricated with incompressible fluids (oils) can also be used for gas films operating at low rotational
speeds and/or low whirl frequencies.
Note from Equation 17.9 that for operation at large rotor speeds (Ω → ∞) and/or large whirl frequen-
cies (ω → ∞), the character of the Reynolds equation changes from elliptical to parabolic. For operation
with large speeds, the infinite speed (Λ → ∞) pressure field becomes

∂ h
0=
∂Θ
( ph ) ⇒ p = pa a
h (Θ)
(17.14)

which* establishes a limit on the generation of hydrodynamic pressure in a radial bearing. Consequently,
the bearing reaction load will also reach a definite limit. The ultimate load capacity (wu) of the cylindri-
cal gas bearing is, as Λ → ∞ [3],

1/ 2
W (1 + e)
2p
cos q p (1 + e) ⎡1 − (1 − e2 ) ⎤
⎣ ⎦
wu = max = −
pa LD 2 ∫( 1 + ecos q
dq =
) e(1 − e2 )
1/ 2 (17.15)
0

with the journal static eccentricity (ε = e/c) is along the direction of the applied load. Figure 17.9 shows
that the ultimate load (wu) grows modestly with journal eccentricity. Gas bearings, unlike mineral
oil–lubricated journal bearings, are not able to support heavy loads, as those typical in large rotating
machinery.† The graph shows a recommended safe upper bound for load capacity selection at wu = 2
which renders an eccentricity (ε) ∼ 0.60. Note that operation at any finite rotational speed will produce
a higher eccentricity. Furthermore, safe operation should avoid very large journal eccentricities that can
provoke transient rubs and impacts that could quickly destroy the RBS.
Incidentally, for operation with infinite frequency (σ → ∞), and for simplicity not accounting for
shear flow effects (Λ = 0), the squeeze film pressure is just

∂ ha
0= ( ph ) ⇒ p ≈ pa h Θ, (17.16)

∂t ( t)
Thus, the pressure is in-phase with the film thickness, that is, solely determined by the displacements
(eX,eY) and not its time variations, that is, not a function of the velocity at which the film thickness

5.0
4.5 W/( paLD)
4.0
3.5
Load, w (–)

3.0 Recommended limit


2.5 for safe operation
Load, W
2.0
1.5
Y
1.0 Ω
(Λ ∞) 0.5 X
0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Eccentricity (e/c)

FIGURE 17.9  Ultimate load capacity (W/paLD) of plain cylindrical journal bearing. Infinite speed solution
(Λ → ∞).

* This solution is to be taken with caution since it does not satisfy all the boundary conditions, in particular at the bearing
axial edges, i.e., p = pa at z = ± L2 .

Specific load capacities ( W
pa L D ) in oil bearings easily exceed 20 bar or more, i.e., w > 20.
Compressible Gas Film Lubrication 17-15

changes. These operating conditions thus lead to a stiffening or hardening of the gas film and absence
of squeeze film-damping effects. Importantly enough, high-frequency motions of a squeeze gas film can
generate a mean pressure above ambient and, hence, the ability to carry a static load (albeit small). See
Reference 2 for details on this rectification phenomenon.

17.4.2  Performance of a Plain Cylindrical Gas Journal Bearing


Table 17.3 shows the geometry and operating conditions of a cylindrical journal bearing operating with
air at ambient condition. The bearing application is typical for a miniature high-speed spindle.
The performance of the bearing is calculated in terms of the dimensionless load parameter (w), speed
2
mNLD ⎛ R ⎞
number (Λ), and the Sommerfeld number S = with N as the rotational speed in rev/s. In
W ⎝c⎠
the design (and selection) of a gas bearing, S serves to size the bearing clearance.*
Figures 17.10 and 17.11 show the static (equilibrium) eccentricity (ε) and attitude angle (ϕ) versus
Sommerfeld number (S). ϕ is the angle between the load vector and the ensuing journal eccentricity vec-
tor. Each graph includes the (single) curve representative of the operation for the journal bearing with
an incompressible lubricant. With an incompressible lubricant, large Sommerfeld numbers S, denoted
by either a small load W, a high rotor speed Ω, or large lubricant viscosity μ, determine small operating
journal eccentricities or nearly a centered operation, that is, ε → 0 and ϕ → ½π (90°). That is, the journal
eccentricity vector e is nearly orthogonal or perpendicular to the applied load vector W.
A cylindrical (plain) gas bearing does not offer a unique performance curve; albeit the maximum
journal eccentricity is bounded by the solution for the incompressible lubricant† and accounting for
liquid cavitation. The specific loads in a gas bearing are, by necessity, rather small. That is, even a w = 1.50
(see Figure 17.10) determines large operating eccentricities, in particular, when the speed number (Λ) is
also low. As per the attitude angle, gas bearings show smaller ϕ than with incompressible lubricants, in
particular at high speeds.
Figure 17.12 shows the drag friction coefficient, f  = Torque/cW, is indistinguishable between incom-
pressible fluid and gas film journal bearings. This is so since the shear stress model is viscous in char-
acter, that is, not affected by fluid compressibility. The result does not mean a gas bearing has the same

TABLE 17.3  Geometry and Operating Conditions


of Plain Cylindrical Gas Bearing
Journal diameter, D 0.0285 m L/D = 1
Length, L 0.0285 m
Clearance, c 0.020 mm R/c = 712
Lubricant: Air at 26.7°C
Ambient pressure, pa 1.01 bar
Viscosity, μ 0.0185 cP
Density, ρ 1.16 kg/m3
Specific load, paLD 82 N
Journal speed 10–100 krpm Ω = RPM π/30
Load, W 10–100 N

* Even to this day, turbomachinery is designed (and built) with little attention to the needs of bearings and adequate lubri-
cation for cooling and load support, static and dynamic. That is, thermofluidic and aerodynamic considerations dictate
the speed and size of the rotating elements. Fixed diameter and length for a bearing and the lubricant to be used, as well
as the load to be supported, severely constrain the design space. The bearing designer has only the bearing clearance (c)
to play with.
† The solution for the incompressible fluid accounts for lubricant (gas) cavitation at ambient pressure.
17-16 Theory and Practice of Lubrication and Tribology

Λ = 1.17 – 11.7

1.0 w = 0.244

0.9 w = 0.488

0.8 w = 1.000

Eccentricity (e/c)
0.7 w = 1.500
0.6 Incompressible
f luid
0.5
0.4 Load, W
0.3 Y
0.2 Ω
0.1
X
0.0
0.01 0.10 1.00
2
Sommerfeld # S=
µN L D R
W c

FIGURE 17.10  Journal eccentricity (ε) versus Sommerfeld number for plain cylindrical gas journal bearing. Load
(w) increases.

w = 0.244
Λ = 1.17 – 11.7
90 w = 0.488
Load, W
80 w = 1.00
Y
Attitude angle (degree)

70 Ω w = 1.50
60 Incompressible
X f luid
50
40
30
20
10
0
0.01 0.10 1.00
2
µN L D R
Sommerfeld # S = W c

FIGURE 17.11  Journal attitude angle (ϕ) versus Sommerfeld number for plain cylindrical gas journal bearing.
Load (w) increases.

Λ = 1.17 – 11.7
100.0

f = Torque/(cW)

10.0
Drag friction
coefficient

w = 0.244
w = 0.488
1.0
w = 1.00
w = 1.50
Incompressible
0.1 f luid
0.01 0.10 1.00
2
µN L D R
Sommerfeld # S=
W c

FIGURE 17.12  Drag friction coefficient (f ) versus Sommerfeld number for plain cylindrical gas journal bearing.
Load (w) increases.
Compressible Gas Film Lubrication 17-17

drag torque (and power loss ℘ = Torque Ω) as a mineral oil bearing. The difference in viscosities causes
the gas bearing to have a much lower drag coefficient; f is quite small, two orders of magnitude at least.

17.4.3  Bearing Force Coefficients and Dynamic Stability


Figure 17.13 depicts the bearing stiffness and damping force coefficients evaluated at a frequency coin-
ciding with the journal rotational speed (ω = Ω). In the example, the dimensionless load w = 0.488
while the journal speed increases from 10 to 100 krpm. Hence, the bearing Speed number Λ = 1.17
to 11.7, and the Sommerfeld number S = 0.032 to 0.318. The dimensionless force coefficients are
3
D ⎛ L⎞
K = K (C*Ω ) and C = C C*, where C* = m ⎜ ⎟ . See Figure 17.10 for the relation between the journal
4 ⎝ c⎠
eccentricity and the Sommerfeld number. Note that the direct stiffnesses (K XX, K YY) and damping (CXX,
CYY) coefficients increase with the journal eccentricity (ε).
The stability of the RBS is of interest. In general, this is a elaborated procedure that requires the inte-
gration of the fluid film-bearing reaction forces into a rotordynamics model. Simple analyses consider
a point mass (M) rigid rotor supported on a gas bearing. The (linearized) equations of motion of the
system about an equilibrium condition (F = W) are

M z + K z + C z = Fe (17.17)

where z = {x(t),y(t)}T is the vector of dynamic displacements of the journal center. Previously, Fe = {{FX,FY}T
is the external dynamic force vector acting on the system, for example, due to rotor mass imbalance. The
stability of the system considers the homogeneous form of Equation 17.17 and assumes an initial state
(zi, żi) away from the equilibrium condition. With z = zo est, then

⎡(K + s 2 M ) + Cs ⎤ z o = 0 (17.18)
⎣ ⎦

The roots of the characteristic equation |(K + s2M) + Cs| = 0 are s1,2 = λ ± iϖ. If the real part λ < 0, then the
RBS is stable returning to its equilibrium position, that is, z → 0 as t → ∞. If, on the other hand, λ > 0,
then the RBS is unstable and z grows without bound.*

C/C*
10.0
Cxx
Cxx
Cyy
Cyy
Bearing damping

Cxy –Cxy

1.0
Cyx
Load, w

Y
Ω
w = 0.488 Cyx
X
0.1
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Eccentricity (e/c)

FIGURE 17.13  Synchronous frequency stiffness and damping force coefficients versus journal eccentricity for
plain cylindrical gas bearing. Load w = 0.488.

* It is a common misconception that the “no bound” statement implies system destruction. In actuality, the journal
will whirl with a large amplitude whirl orbit bounded by the bearing clearance. As the motion amplitude grows, the
bearing nonlinearity determines the size of the limit cycle. Of course, sustained operating under this condition is not
recommended.
17-18 Theory and Practice of Lubrication and Tribology

At the threshold of instability, when λ = 0,


. the system will perform self-excited motions with whirl
frequency, ϖ, that is, z = zo eιϖt. Hence, Equation 17.18 becomes

⎡⎣Z − v 2 M ⎤⎦ z o = 0 where Z = (K + iv C ) (17.19)


Solution of Equation 17.19 is straightforward for an incompressible fluid, rigid surface, journal bearings
since their force coefficients are frequency independent. The analysis leads to the estimation of the sys-
tem critical mass (Mc) and the whirl frequency ratio (WFR) [25]:

K XX CYY + K YY C XX − CYX K XY − C XY K YX
Mcw S2 = K eq =
C XX + CYY

⎛ v ⎞ ( K eq − K XX ) ( K eq − K YY ) − K XY K YX
2

WFR 2 = ⎜ s ⎟ = (17.20)
⎝Ω⎠ C XX CYY − C XY CYX

On the other hand, gas bearings have frequency-dependent force coefficients, K(ω) and C(ω). As an
example, for the particular operating conditions noted, Figure 17.14 depicts the dimensionless stiffness
(Kij)ij = X,Y and damping (Cij)ij = X,Y coefficients versus frequency ratio (ω/Ω) where Ω is the rotational speed;
ω/Ω = 1 denotes whirl frequency excitation synchronous with the rotational speed). Note that the direct
stiffnesses increase with whirl frequency, a typical hardening effect due to fluid compressibility. On the
other hand, the damping coefficients at high frequencies are zero, Cij → 0 as ω → ∞, also due to fluid
compressibility. An iterative method is required to solve for the characteristic Equation 17.19 |Z(ϖ) −
ϖ 2M| = 0. Lund [24] restated this equation as Z(ϖ) = ϖ 2M, and hence, the instability threshold occurs at
frequency ϖs where the imaginary part of the complex impedance Ze is zero while its real part must be
greater than zero. The equivalent impedance is

2
1 2
Ze(v ) =
2
( ) ⎣
(
Z XX(v ) + ZYY(v ) − ⎡⎢ 1 4 Z XX(v ) − ZYY(v ) ) +Z XY(v ) ZYX(v ) ⎤⎥

(17.21)

Kxx Kyy –Kxy Kyx Cxx Cyy –Cxy Cyx

6 3

–Cxy
Kxx
4 2 Cyx

Kyy Cxx
2 1
–Kxy
0 Cyy
0

Kyx
–1
–2
0.01 0.1 1 10 0.01 0.1 1 10
(a) Frequency ratio (b) Frequency ratio

FIGURE 17.14  Bearing stiffness and damping force coefficients versus whirl frequency ratio (ω/Ω). Plain cylin-
drical gas bearing. Load w = 0.488, speed = 50 krpm (Λ = 5.843), S = 0.158 (e/c = 0.485). (a) Stiffness coefficients, K;
(b) damping coefficients, C.
Compressible Gas Film Lubrication 17-19


( ) ( )
Im Ze(v s ) = 0 and Re Ze(v s ) > 0 (17.22)

The first statement earlier implies the effective damping is zero. For the data shown in Figure 17.14, the
RBS critical mass is just Mc = 0.968 kg and the WFR = 0.48. That is, for operation with journal rotation at
50 krpm (833 Hz), the RBS becomes unstable if its physical mass is greater than Mc. If the actual system
mass M > Mc, the RBS will begin self-excited motions at a frequency equaling 48% of the running speed,
that is, ϖs = 400 Hz. This whirl frequency is also the natural frequency of the RBS for the noted operat-
ing condition. For cylindrical gas bearings of various types, Czolczynski [4] lists tables of rotordynamic
force coefficients, critical mass, and whirl ratios.

17.5  Introduction to Flexure Pivot Bearings


For certain static load dispositions, tilting pad bearings can eliminate the typically harmful hydro-
dynamic instability by not generating cross-coupled stiffness coefficients. Critical turbomachinery
operating well above its critical speeds is customarily implemented with tilting pad bearings. The mul-
tiplicity of parameters associated with a tilting pad bearing demands complex analytical methods for
predictions of force coefficients and stability calculations [10]. Incidentally, conventional (commercial)
tilting pad bearings cannot be easily modified to add external pressurization (holes through pivots
and pads) without constraining severely the pads’ motion and adding sealing issues. A modern bear-
ing configuration suitable for high-speed rotating machinery is the flexure pivot—tilting pad bear-
ing (see Figure 17.1). This bearing type offers a marked improvement over the conventional design
since its wire discharge machined (EDM) construction renders an integral pads-bearing configura-
tion, thus eliminating pivot wear and stack up of tolerances on assembly [13]. Each pad connects to
the bearing through a thin flexural web, which provides a low rotational stiffness, thus ensuring small
cross-coupled stiffness coefficients and avoiding subsynchronous instabilities during very high-speed
operation. See [13,26,27] for details on rotordynamic measurements and predictive model validation.
The tests included reliability of the bearings under intermittent shocks and periodic forces simulat-
ing maneuver loads and uneven road conditions. Furthermore, the test data suggest the possibility
of controlling the supply pressure to move critical speeds and avoid the passage through resonances.
Reference [13] discusses and implements a simple and inexpensive control strategy that demonstrated
remarkable results. In brief, external pressurization is only needed at low rotor speeds, while at high
rotor speeds it can be safely dispensed with.

17.6  Introduction to Foil Bearings


Oil-free systems have a reduced part count, footprint, and weight and are environmentally friendly
with demonstrated savings in long-interval maintenance expenses. Until recently, gas bearings were
constructed with hard or rigid surfaces to reduce friction during start-up or shutdown events. However,
bearing types such as herringbone groove bearings require tight clearances (film thicknesses), and with
their hard surfaces offer few advantages for use in high-speed MTM.
GFBs have emerged as a most efficient alternative for load support in high-speed machinery. These
bearings are compliant surface hydrodynamic bearings using ambient air as the working fluid media.
Recall Figure 17.1 showing two typical GFB configurations: one is a multiple overleaf bearing and the
other is a corrugated bump bearing. Both bearing types are used in commercial rotating machinery,
yet the open literature presents more details on bump-GFBs, along with measurements and analy-
ses. The corrugated bump foil bearing is constructed from one or more layers of corrugated thin
metal strips and a top foil. In operation, a minute gas film wedge develops between the spinning rotor
17-20 Theory and Practice of Lubrication and Tribology

and top foil. The bump-strip layers are an elastic support with engineered stiffness and damping
characteristics [5,18].
GFBs offer distinct advantages over rolling elements bearings including no DN* value limit, reli-
able high-temperature operation, and large tolerance to debris and rotor motions, including tem-
porary rubbing and misalignment. Current commercial applications include auxiliary power units,
cryogenic turbo expanders, and micro-gas turbines. Envisioned or underdevelopment applications
include automotive turbocharger and aircraft gas turbine engines for regional jets and helicopter
rotorcraft systems [5]. Also, GFBs have demerits of excessive power losses and wear of protective
coatings during rotor startup and shutdown events. In addition, expensive developmental costs and,
until recently, inadequate predictive tools limited the widespread deployment of GFBs into midsize
gas turbines. In particular, at high-temperature conditions, reliable operation of GFB-supported
rotor systems depends on adequate engineered thermal management and proven solid lubricants
(coatings).
Successful implementation of GFBs in commercial rotating machinery involves a two-tier effort:
that of developing bearing structural components and solid lubricant coatings to increase the bearing
load capacity while reducing friction, and that of developing accurate performance prediction models
anchored to dependable (non commercial) test data. Chen et al. [18] and DellaCorte and coworkers
[5,28] publicize details on the design and construction of first-generation foil bearings, radial and thrust
types, aiming toward their wide adoption in industry.

17.7  Recommendations for Oil-Free Rotating Machinery


Until recently, GFB design was largely empirical, each foil bearing being a custom piece of hardware,
with resulting variability even in identical units, and limited scalability. At present, the advances
in radial GFB technology (design, construction, and predictability) permit OEMs and end users to
implement radial GFBs for deployment into novel MTM or to upgrade and improve outdated rotating
machinery. That is, there is enough published know-how on materials, guidelines for design and con-
struction of radial GFBS including engineered coatings for high-temperature applications, a reliable
data base of GFB forced performance (static and dynamic), and computational tools benchmarked to
test data.
Research on radial GFBs for lateral support of oil-free rotating machinery has steadily progressed
with comprehensive analyses accounting for most relevant physical aspects to accurately predict GFB
static and dynamic load performance, power loss, and the management of thermal energy in high-
temperature applications. Empirical research has gone beyond showing a few instances of acceptable
mechanical performance, to demonstrate GFB multiple-cycle repeatable performance in spite of per-
sistent large amplitude whirl motions at low frequencies, typically coinciding with the system natural
frequencies. Many developmental efforts have attempted to fix or suppress these undesirable motions.
One could hastily attribute the subharmonic whirl motions to a typical rotordynamic instability
induced by hydrodynamic effects of the gas film, that is, generation of too-large cross-coupled stiffness
coefficients that destabilize the RBS. However, as learned from the measurements [29], rotor imbalance
triggers and exacerbates the severity of subsynchronous motions. The subsynchronous behavior is a
forced nonlinearity due to the foil bearing strong nonlinear (hardening) stiffness characteristics, as is
demonstrated in Reference 30. The predictions and measurements validate the simple FB model, that
is, a minute gas film with effective infinite stiffness, with applicability to large amplitude rotordynamic
motions.
Challenges for GFBs include intermittent contact and wear at startup and shutdown, and potential
for large amplitude rotor whirl at high speeds. Subsynchronous motions are common in foil bearings

* DN, the product of journal diameter (mm) times rotational speed (RPM), is a limiting factor for operation of rolling
element bearings (e.g., DN = 2 million in specialized bearings with ceramic balls).
Compressible Gas Film Lubrication 17-21

due to their strong structural hardening nonlinearity. Incidentally, the ultimate load capacity of a GFB
depends mainly on its support structure. Hence, engineers must pay close attention to the bearing
structural components (design, fabrication, and assembly).

References
1. W.A. Gross, Gas Film Lubrication, John Wiley & Sons, New York, 1962.
2. C.H.T. Pan, Gas bearings, Tribology: Friction, Lubrication and Wear, Ed. A.Z. Szeri, Hemisphere
Publication Corporation, Washington, DC, 1980.
3. B.J. Hamrock, Fundamentals of Fluid Film Lubrication, Chaps. 16–17, McGraw-Hill, New York,
1994.
4. K. Czolczynski, Rotordynamics of Gas-Lubricated Journal Bearing Systems, Springer-Verlag,
New York, 1999.
5. C. DellaCorte, K.C. Radil, R.J. Bruckner, and S.A. Howard, Design, fabrication, and performance of
open source generation I and II compliant hydrodynamic gas foil bearings, Tribology Transactions,
51(3), 254–264, 2008.
6. G. Belforte, T. Raparelli, V. Viktorov, A. Trivella, and F. Colombo, An experimental study of high-
speed rotor supported by air bearings: Test rig and first experimental results, Tribology International,
39(8), 839–845, 2006.
7. N. Zirkelback and L. San Andrés, Effect of frequency excitation on the force coefficients of spiral
groove thrust bearings and face gas seals, Journal of Tribology, 121(4), 853–863, 1999.
8. M. Faria and L. San Andrés, On the numerical modeling of high speed hydrodynamic gas bearings,
Journal of Tribology, 122(1), 124–130, 2000.
9. L. San Andrés and D. Wilde, Finite element analysis of gas bearings for oil-free turbomachinery,
Revue Européenne des Eléments Finis, 10(6/7), 769–790, 2001.
10. L. San Andrés, Hybrid flexure pivot-tilting pad gas bearings: Analysis and experimental validation,
Journal of Tribology, 128(3), 551–558, 2006.
11. D.A. Osborne and L. San Andrés, Experimental response of simple gas hybrid bearings for oil-free
turbomachinery, Journal of Engineering for Gas Turbines and Power, 128(3), 626–633, 2006.
12. X. Zhu and L. San Andrés, Experimental response of a rotor supported on Rayleigh step gas bear-
ings, American Society of Mechanical Engineers Paper No. GT 2005-68296, ASME, New York, 2005.
13. L. San Andrés and K. Ryu, Hybrid gas bearings with controlled supply pressure to eliminate rotor
vibrations while crossing system critical speeds, Journal of Engineering for Gas Turbines and Power,
130(6), 062505, 2008.
14. T.H. Kim and L. San Andrés, Limits for high speed operation of gas foil bearings, Journal of Tribology,
128(3), 670–673, 2006.
15. T.H. Kim and L. San Andrés, Heavily loaded gas foil bearings: A model anchored to test data, Journal
of Engineering for Gas Turbines and Power, 130(1), 012504, 2008.
16. L. San Andrés and T.H. Kim, Analysis of gas foil bearings integrating FE top foil models, Tribology
International, 42(1), 111–120, 2009.
17. C. Carnes, Hard-driving lubrication, Tribology and Lubrication Technology, November, 32–38, 2004.
18. H.M. Chen, R. Howarth, B. Geren, J.C. Theilacker, and W.M. Soyars, Application of foil bearings
to helium turbocompressor, Proceedings of 30th Turbomachinery Symposium, Turbomachinery
Laboratory, Texas A&M University, Houston, TX, 2000, pp. 103–113.
19. L. San Andrés, T. Chirathadam, K. Ryu, and T. H. Kim, Measurements of drag torque, lift-off journal
speed and temperature in a metal mesh foil bearing, Journal of Engineering for Gas Turbines and
Power, 132(11), 112503, 2010.
20. D. Childs, Turbomachinery Rotordynamics, Chap. 5, John Wiley & Sons, New York, 1993.
21. A.Z. Szeri, Fluid Film Bearings: Theory & Design, Chap. 11, Cambridge University Press, Cambridge,
U.K., 1998.
17-22 Theory and Practice of Lubrication and Tribology

22. J.A. Williams, Engineering Tribology, Chap. 7, Oxford Science Publications, Oxford, Great Britain,
2000.
23. J.W. Lund, Calculation of stiffness and damping properties of gas bearings, Journal of Lubrication
Technology, 90, 793–803, 1968.
24. J.W. Lund, A theoretical analysis of whirl instability and pneumatic hammer for a rigid rotor in
pressurized gas journal bearings, Journal of Lubrication Technology, 89, 154–163, 1967.
25. J.W. Lund, The stability of an elastic rotor in journal bearings with flexible damped supports, Journal
of Applied Mechanics, 87, 911–920, 1965.
26. L. San Andrés and K. Ryu, Dynamic forced response of a rotor-hybrid gas bearing system due to
intermittent shocks, American Society of Mechanical Engineers Paper No. GT2009-59199, ASME,
New York, 2009.
27. L. San Andrés, K. Ryu, and Y. Niu, Dynamic response of a rotor-hybrid gas bearing system due to
base induced periodic motions, American Society of Mechanical Engineers Paper No. GT2010-
22277, ASME, New York, 2010.
28. B. Dykas, R. Bruckner, C. DellaCorte, B. Edmonds, and J. Prahl, Design, fabrication, and performance
of foil gas thrust bearings for microturbomachinery applications, Journal of Engineering for Gas
Turbines and Power, 131(1), 012301, 2009.
29. L. San Andrés, D. Rubio, and T.H. Kim, Rotordynamic performance of a rotor supported on bump
type foil gas bearings: Experiments and predictions, Journal of Engineering for Gas Turbines and
Power, 129(3), 850–857, 2007.
30. L. San Andrés and T.H. Kim, Forced nonlinear response of gas foil bearing supported rotors,
Tribology International, 41(8), 704–715, 2008.
18
Elastohydrodynamic
Lubrication
Nomenclature.............................................................................................. 18-1
18.1 Introduction..................................................................................... 18-2
18.2 Regimes of Lubrication................................................................... 18-2
Hydrodynamic Lubrication  •  Elastohydrodynamic
Lubrication  •  Partial or Mixed Lubrication  •  Boundary
Lubrication
18.3 EHL Models......................................................................................18-4
Line and Point Contacts
18.4 EHL Minimum Film Thickness Equations................................. 18-7
Farshid Sadeghi Thermal Correction Factor
Purdue University
18.5 Thermal and Non-Newtonian Effects..........................................18-9
Trevor Slack 18.6 Application of EHL Theory to Machine Components............. 18-10
Purdue University References................................................................................................... 18-16

Nomenclature
a Half width of Hertzian contact across rolling direction, m
b Half width of Hertzian contact along rolling direction, m
E′ Equivalent modulus of elasticity, 2((1 − n12 )/E1 + (1 − n22 )/E2 )−1 , Pa
E1, E2 Modulus of elasticity of contacting surfaces 1 and 2, Pa
G Dimensionless material parameter, αE′
H Dimensionless film thickness, hR x/b2
H0 Dimensionless film thickness constant
H Film thickness, m
hc Film thickness at center of contact, m
hmin Minimum film thickness, m
Λ Dimensionless film parameter
Mg Gear ratio
P Dimensionless pressure, p/PH
PH Maximum Hertzian pressure, Pa
p Pressure, Pa
R x Reduced radius of curvature in the x-direction, m
Ry Reduced radius of curvature in the y-direction, m
Rq Root-mean-square surface roughness
U Dimensionless speed parameter, um η 0/E′R x

18-1
18-2 Theory and Practice of Lubrication and Tribology

u1, u2 Velocities of contacting surfaces 1 and 2 in rolling direction, m/s


um Average surface velocity, (u1 + u2)/2, m/s
Um Dimensionless mean velocity of contacting surfaces in rolling direction, um/uM
W Dimensionless load parameter, w/E′ Rx2
w External normal load, N
w′ External normal load per unit width, N/m
X dimensionless location along rolling direction, x/b
x Coordinate in rolling direction, m
Y Dimensionless location across rolling direction, y/b
y Coordinate across rolling direction, m
z Roelands pressure–viscosity index of lubricant
α Barus pressure–viscosity coefficient, Pa−1
η Viscosity of lubricant, Pa·s
η 0 Absolute viscosity of the lubricant at p = 0 and constant temperature, Pa·s
h Dimensionless absolute viscosity, η/η 0
κ Ellipticity ratio, a/b
ν1, ν2 Poisson’s ratio for contacting surfaces 1 and 2
ρ Density of lubricant, kg/m3
ρ 0 Ambient density of lubricant, kg/m3
r Dimensionless density, ρ/ρ 0
ωg Angular velocity of gear

18.1  Introduction
Elastohydrodynamic (EHD) lubrication (EHL) is a form of lubrication occurring between rolling/sliding
nonconformal machine components. EHL consists of the hydrodynamic action of fluid entrainment at
pressures high enough to elastically deform the contacting surfaces under conditions where lubricant
viscosity variation with pressure is significant. The discovery of EHL has provided the explanation for
the effective lubrication and satisfactory operation of rolling/sliding machine components (e.g., rolling
element bearings, cam and its followers, gears, etc.).
The development of the EHL theory occurred throughout the twentieth century. The development
began with the work of Ertel (1939) who included the elastic deformation of the contacting solids and
the viscosity variation of the lubricant with pressure in analyzing the inlet zone of lubricated noncon-
formal contacts. His pioneering work showed significantly larger film thickness than those previously
predicted by the hydrodynamic theory. The early investigations primarily focused on the basic EHL
solution for isothermal lubrication using the Newtonian fluid model. These works led to the develop-
ment of formulas for minimum and central film thicknesses which continue to be widely used today.
Attention then shifted toward better quantifying the friction occurring at the contact. Improvements in
friction predictions were made by including thermal and/or non-Newtonian effects in EHL simulations.
The improvements in computational methods and computer hardware have allowed EHL investigators
to extend their efforts and develop models for mixed EHL.
Dowson and Ehret (1999) provide a comprehensive review of EHL. More detailed discussion of vari-
ous aspects of EHL problems is included in the following sections and additional references relevant to
EHL are listed at the end of this chapter.

18.2  Regimes of Lubrication


A lubricant is a substance which is introduced into the contact of loaded rolling/sliding bodies to
control friction and wear. Liquid lubricants are pressurized in a converging contact due to surface
drag; they can lower contact temperatures and remove contaminants. Additives in lubricants can
Elastohydrodynamic Lubrication 18-3

provide corrosion–oxidation and wear resistance. Four different lubrication regimes are identified
for lubricated contacts: (i) hydrodynamic, (ii) elastohydrodynamic, (iii) partial or mixed, and (iv)
boundary.

18.2.1  Hydrodynamic Lubrication


Hydrodynamic or full film lubrication is the lubrication condition where the load-carrying surfaces
are separated by a relatively thick film of lubricant. This regime of lubrication is stable and surface-to-
surface contact does not occur during the steady-state operation of the bearing. Pressure generated in
the contact is on the order of megapascals. The minimum film thickness separating the surfaces is typi-
cally a few micrometers and the coefficient of friction is on the order of 10−2. This lubrication regime is
typical of conformal contacts.

18.2.2  Elastohydrodynamic Lubrication


EHL is the lubrication condition where the load is large enough for the surfaces to elastically deform
during the hydrodynamic action. Typical pressures occurring in this regime can range from 0.5 to over
4 GPa. The minimum film thickness separating the surfaces is usually ≤1 μm and the coefficient of fric-
tion is on the order of 10−2 to 10−1. This regime of lubrication is typical of nonconformal contacts such as
ball and rolling element bearings.

18.2.3  Partial or Mixed Lubrication


Partial or mixed lubrication regime is the lubrication condition when the tallest asperities of the contact-
ing surfaces protrude through the lubricant film and occasionally come into contact. This occurs when
the speed is low, the load is high, or the temperature is sufficiently large to significantly reduce lubricant
viscosity. Partial EHL deals with simultaneously occurring solid-to-solid contact and EHD-lubricated
contact. It is generally believed that if the average film thickness is less than three times the composite
surface roughness, surface asperities will penetrate the lubricant film and force direct solid-to-solid
contact, resulting in mixed EHL. The average film thickness in partial lubrication is between 0.01 and
0.1 μm. Since part of the load is supported by a bulk lubricant film, the coefficient of friction of mixed
lubrication is usually <0.1.

18.2.4  Boundary Lubrication


Boundary lubrication is the condition when the fluid films are negligible and there is considerable asper-
ity contact. The physical and chemical properties of thin surface films are of significant importance
while the properties of the bulk fluid lubricant are insignificant. In boundary lubrication, the rolling and
sliding bodies are only separated by a few layers of molecules and the surface film varies in thickness
from 1 to a few tens of nanometers. Boundary lubrication occurs when a fluid film cannot be formed
or sustained due to heavy loads, low running speeds, high surface roughness, and/or a lack of lubricant
supply. The frictional characteristics of boundary lubrication depend on the rheological behavior of the
lubricant and the interaction (both physical and chemical) between the molecular surface film and the
solids. The coefficient of friction for sliding boundary lubrication ranges from 0.05 to 0.15 with an aver-
age of 0.1, which is larger than that occurring in the mixed EHL regime but still less than what occurs
for unlubricated contacts.
A schematic of the relative film thicknesses encountered in the different lubrication regimes is given
in Figure 18.1. To distinguish the lubrication regimes possible in rough surface lubrication, a dimen-
sionless film parameter Λ is defined as
18-4 Theory and Practice of Lubrication and Tribology

w w w

um um um

(a) (b) (c)

FIGURE 18.1  Fluid film thickness in the different lubrication regimes. (a) Full film (HD/EHL); (b) mixed;
(c) bounday.

hmin
L= (18.1)
Rq2,a + Rq2,b

where
hmin is the minimum lubricant film thickness
Rq,a and Rq,b are the root-mean-square surface finish of the contacting bodies

The lubrication regimes are characterized as

⎧ 5 ≤ L < 100 : hydrodynamic lubrication


⎪ 3 ≤ L < 10 : elastohydrodynamic lubrication


⎪1≤ L < 5 : partial or mixed lubrication
⎪⎩ L < 1 : boundary lubrication

The film parameter plays a significant role in the expected life of a lubricated contact. For the particular
case of rolling element bearings, a fatigue life adjustment factor has been developed in terms of the film
parameter (Tallian 1967; Bamberger et al. 1971).

18.3  EHL Models


18.3.1  Line and Point Contacts
Line contact occurs when two cylindrical bodies are brought in contact under load. In general, rolling
element bearings and uncrowned spur and helical gears are considered as line contacts. It is important
to note that even cylindrical rolling element bearings usually have a large crown. Most machine elements
that are nonconformal have elliptical contact and seldom circular contact however; they are commonly
referred to as point contacts. Note that the term point contact is sometimes reserved only for circular con-
tacts; however, in this chapter the term point contact is used to refer to both elliptical and circular contacts.
To obtain a solution to an EHL problem, the Reynolds and elasticity equations are simultaneously
solved subject to boundary conditions while allowing for viscosity and density variation with pressure.

18.3.1.1  Line EHL Contact


The dimensionless steady-state one-dimensional Reynolds equation is given as

∂ ⎛ rH 3 ∂P ⎞ ∂
−l ( rH ) = 0 (18.2)
∂X ⎜⎝ h ∂X ⎟⎠ ∂X

Elastohydrodynamic Lubrication 18-5

The dimensionless elasticity equation is given as

X2 1
H ( X ) = H0 + −
2 p ∫
P ( X ) ln X − X ʹ dX ʹ (18.3)
Ω

The dimensionless viscosity and density variations with pressure are


{ ⎣
z
h ( p ) = exp ⎡⎣ ln (h0 ) + 9.67 ⎤⎦ ⎡(1 + 5.1 × 10 −9 p ) − 1⎤
⎦ } (18.4)

⎛ 0.6 × 10 −9 p ⎞
r ( p) = ⎜ 1 + (18.5)
⎝ 1 + 1.7 × 10 −9 p ⎟⎠

The dimensionless boundary conditions are given by

⎧0, at X = Xinlet

P=⎨ dP (18.6)
⎪⎩P = dX = 0, at X = X exit

The dimensionless load balance equation is

p
∫ P(X)dX = 2 (18.7)
Ω

Figure 18.2 depicts the pressure and film thickness results for an EHL line contact. The results dem-
onstrate that as the speed decreases and/or the load increases the well-known pressure spike moves
toward the exit zone of the contact. Figure 18.2 also depicts the restriction in film thickness near the
exit location. The pressure spike and/or film thickness restrictions are typical of EHL contacts.

1.6 1.5
1
Nondimensional film thickness (hRx/b )
2

1.4
Nondimensional pressure (p/ph)

1.2
2
1
1
3
0.8 1

0.6
0.5
0.4
2
0.2 3

0 0
–1.5 –1 –0.5 0 0.5 1 1.5 –1.5 –1 –0.5 0 0.5 1 1.5
(a) Nondimensional rolling direction (x/b) (b) Nondimensional rolling direction (x/b)

FIGURE 18.2  (a) Pressure and (b) film thickness results for EHL line contact at various speeds (W = 1.1e−4,
G = 3500, (1) U = 5e−10, (2) U = 1e−10, (3) U = 5e−11).
18-6 Theory and Practice of Lubrication and Tribology

18.3.1.2  Point EHL (Elliptical and Circular) Contact


The steady-state dimensionless Reynolds equation is given as

∂ ⎛ rH 3 ∂P ⎞ 1 ∂ ⎛ rH 3 ∂P ⎞ ∂
+ −l (rH ) = 0 (18.8)
∂X ⎜⎝ 12h ∂X ⎟⎠ k 2 ∂Y ⎜⎝ 12h ∂Y ⎟⎠ ∂X

The dimensionless elasticity equation is given as

X 2 k 2 Y 2 2Rx PH κ Pl ( X ʹ,Y ʹ)dX ʹdY ʹ


H ( X ,Y ) = H 0 +
2
+
Ar 2
+
pE ʹb ∫∫ ( X − X ʹ)2 + (Y − Y ʹ)2
(18.9)
Ω

The dimensionless boundary conditions are

⎧0, at (Xinlet ,Y ), (X ,Yinlet ), (X ,Youtlet )



P=⎨ dP (18.10)
⎪⎩P = dX = 0, at X = X exit

The dimensionless load balance equation is given by

∫∫ P(X,Y )dX = W (18.11)


Ω

Figure 18.3 illustrates the pressure and film thickness for an EHL point (circular) contact. The results
demonstrate the usual pressure spike near the end of the contact and the film reduction in these areas. It
is believed that this pressure spike contributes significantly to rolling contact fatigue.
The line and point contact results were presented for smooth surfaces. To illustrate the effects that
surface roughness plays at lower values of the film parameter, results from a mixed EHL model are
now presented. Figure 18.4 depicts the pressure and film thickness distribution for a mixed EHL con-
tact at a load of W = 1.33e−5 and material parameter of G = 3853 at different nondimensional speeds.
The pressure and film thickness distribution along the center of contact in the rolling direction is also

2 1
0.8
1.5 0.6
0.4
–1
P 0.2
1
–1.2
H 0
Y

–1.4
-0.2
0.5
–1.6
-0.4
1 -0.6
0 0.5
0.5
1
0 -0.8
1 –0.5 –0.5 0
Y –1 –1
X
–1
Y0 –1 –0.5 0 0.5 1
–1 0 X
–1 X

(a) Analytical pressure (b) Analytical film thickness (c) Analytical film thickness
distribution contours

FIGURE 18.3  Analytical film thickness and pressure results at W = 1.0e−5, U = 5.0e−9, G = 3000.
Elastohydrodynamic Lubrication 18-7

5
5 5
4
4 4
3
3 3
P (p/PH)

P (p/PH)

P (p/P )
H
2
2 2
1 1 1
0 0 0
5 5 5 4
4 3 4 3 3
2 2 2 1
1 0 –1 1 0 –1 1
1 0 –1 1
Y (y/b) –2 –3 0 Y (y/b) –2 –3 0 Y (y/b)
–2 –3 0
–4 –5 –1 X (x/b) –4 –4 –5 –1 X (x/b)
–5 –1 X (x/b)

Pressure distribution

H (hR/b2)
H (hR/b )
H (hR/b2)

0 0

2
–0.5 –0.5 –0.5
–1 –1 –1
5 5 5
4 4 4
3 3 3
2 2 2
1 1 1
0 0 0
Y (y/b) –1 Y (y/b) –1 Y (y/b) –1
–2 –2 –2
–3 1 –3 1 –3 1
–4 0 –4 0 –4 0
–5 –1 X (x/b) –5 –1 X (x/b) –5 –1 X (x/b)

Film thickness distribution


1 5 1 5 1 5
Dimensionless film thickness (hRx/b )

Dimensionless film thickness (hR /b )

Dimensionless film thickness (hR /b )


2

2
0.8 4 0.8 4 0.8 4
Dimensionless pressure (P/PH)

Dimensionless pressure (P/P )

Dimensionless pressure (P/P )


x

x
H

H
0.6 3 0.6 3 0.6 3

0.4 2 0.4 2 0.4 2

0.2 1 0.2 1 0.2 1

0 0 0 0 0 0
–2 –1.5 –1 –0.5 0 0.5 1 1.5 –2 –1.5 –1 –0.5 0 0.5 1 1.5 –2 –1.5 –1 –0.5 0 0.5 1 1.5
X (x/b) X (x/b) X (x/b)
Pressure and film thickness distribution along the center line
0 0.3 0 0.6 0 0.3
0.2 0.4 0.2
–0.5 0.1 –0.5 –0.5 0.1
0.2
Y (y/b)

Y (y/b)
Y (y/b)

0 0
0
–1 –0.1 –1 –1 –0.1
–0.2 –0.2 –0.2
–1.5 –0.3 –1.5 –0.4 –1.5 –0.3
–1.5 –1 –0.5 0 0.5 1 1.5 –1.5 –1 –0.5 0 0.5 1 1.5 –1.5 –1 –0.5 0 0.5 1 1.5
X (x/b) X (x/b) X (x/b)

Internal shear stress distribution along the center line σxy/PH


U = 2e–13 U = 2e–12 U = 2e–11

FIGURE 18.4  Mixed EHL contacts (W = 1.33e−5, G = 3853, ellipticity ratio = 4.2).

shown. The results indicate that the pressure undergoes large fluctuations where there is surface asperity
contact. The orthogonal shear stress along the center line of contact in the rolling direction is shown to
demonstrate the effect of surface asperities on subsurface stresses.

18.4  EHL Minimum Film Thickness Equations


The minimum film thickness separating rolling and sliding surfaces is of significant importance for
tribological contacts. Its value is indicative of the ability of the lubricant film to support the load while
preventing surface-to-surface contact. The conditions that influence the minimum film thickness are
the material properties of the contacting bodies, the rheology of the lubricant, the applied load, and the
surface velocity of the contact.
Grubin and Vinogradova (1949) were the first to develop an elegant solution for the minimum film
thickness in line contacts. Dowson and Higginson (1959) developed a similar minimum film thickness
equation based on curve fits of numerically obtained EHL solutions.
Pan and Hamrock (1989) presented a curve-fit formula for line contact minimum and central film
thickness as
18-8 Theory and Practice of Lubrication and Tribology

0.694 −0.128
hmin ⎛u h ⎞ 0.568 ⎛ wʹ ⎞
= 1.714 ⎜ m 0 ⎟ (aE ʹ ) ⎜⎝ E ʹR ⎟⎠ (18.12)
Rx ⎝ E ʹRx ⎠
x

0.692 −0.166
hc ⎛u h ⎞ ⎛ w ⎞
= 2.922 ⎜ m 0 ⎟ (aE ʹ )0.470 ⎜⎝ E Rʹ (18.13)

Rx ⎝ E ʹRx ⎠ ʹ x ⎟⎠

Hamrock and Dowson (HD) (1977) developed the minimum and central film thickness for elliptical
contacts:

2
0.68 −0.073
⎛ ⎛ Ry ⎞ p ⎞
hmin ⎛u h ⎞ ⎛ w ⎞ ⎜ −0.68 ⎜
⎝ Rx ⎠ ⎟

Rx
= 3.63 ⎜ m 0 ⎟
⎝ E ʹRx ⎠
(aE ʹ )0.49 ⎜⎝ E R2 ⎟⎠ ⎜1 − e ⎟ (18.14)
ʹ x
⎜⎝ ⎟⎠

2
0.67 −0.067 ⎛ ⎛ Ry ⎞ p ⎞
hc ⎛u h ⎞ 0.53 ⎛ w ⎞ −0.73 ⎜
⎜ 1 − 0.61e ⎝ Rx ⎠ ⎟ ⎟
= 2.69 ⎜ m 0 ⎟ (aE ʹ ) ⎜⎝ E ʹR2 ⎟⎠ (18.15)
Rx ⎝ E ʹRx ⎠ x
⎜ ⎟
⎜⎝ ⎟⎠

Optical interferometry is used to measure film thickness in EHL contacts by bringing a highly
polished steel ball in contact under load with a transparent coated disk (usually glass) in order to
enhance and provide color interference fringes of the contact. Figure 18.5 illustrates the interfer-
ence photomicrographs of an EHL circular contact along with corresponding analytical results.
Koye and Winer (1981), Smeeth and Spikes (1997), and others have all used optical interferometry to
experimentally measure the minimum film thickness of point contacts. These measurements have
confirmed that the Hamrock and Dowson (1977) equation predicts the minimum film thickness for
most EHL applications to within the accuracy normally required in mechanical design calculations.
Koye and Winer (1981) showed good correlation between measured and predicted minimum film
thicknesses, except for at higher loads where the measured minimum film thickness falls below the
predicted values. Some recent studies indicate that at low speeds the Hamrock and Dowson (1977)
equation does not compare well with experimental results. Coy and Zaretsky (1981) found a fair

1.2
1
0.8
P 0.6
0.4
0.2
0
1

0
Y 1
0.5
–1 0
–0.5
–1 X

(a) (b) (c)

FIGURE 18.5  (a) Experimental interference photomicrograph and (b) corresponding analytical film thickness
and (c) pressure results at W = 4e−6, U = 1.94e−10, G = 4127.
Elastohydrodynamic Lubrication 18-9

agreement between the measured and theoretical film thicknesses for dimensionless speeds <10 −10.
They found that the differences between the film thicknesses predicted using different EHL theories
were less than the scatter in the experimentally measured data. At higher speeds, they found that the
film was smaller than the theoretical predictions which they attributed to inlet shear heating and
starvation effects.

18.4.1  Thermal Correction Factor


Gupta et al. (1991) presented a thermal correction factor formula for calculating the percentage of film
thickness reduction due to inlet heating given by

⎛P ⎞
1 − 13.2 ⎜ H ⎟ (L* )0.42
⎝ Eʹ ⎠
Ct = (18.16)
1 + 0.213(1 + 2.23AC 0.83 )(L* )0.64

where

⎛ ∂h ⎞ u 2
L* = ⎜ − (18.17)
⎝ ∂t m ⎟⎠ k f

Here
PH is the maximum Hertzian pressure
E′ is the effective modulus
η is the absolute viscosity
tm is lubricant temperature
ũ 2 is the mean surface velocity
kf is the lubricant thermal conductivity
AC is the slide-to-roll ratio
Ct is used to multiply the Hmin equation and obtain the effects of temperature on film thickness.

18.5  Thermal and Non-Newtonian Effects


Newtonian isothermal EHL analyses predict film thicknesses that closely match experimental results but
they fail to predict friction at the contact accurately. The need to include thermal and/or non-Newtonian
fluid effects has been the subject of much research since the early 1970s. To allow for thermal effects in
EHL contacts, the thermal Reynolds (Newtonian and/or non-Newtonian), elasticity, and energy equa-
tions are solved simultaneously subject to boundary conditions while allowing for viscosity and density
variation with pressure and temperature. Cheng (1970) investigated the effects of temperature on EHL
contacts. Sadeghi and Sui (1989) investigated thermal effects in EHL line contacts using both Newtonian
and non-Newtonian (Ree–Eyring) fluid models. They demonstrated that for higher loads and slide-to-
roll ratios up to 30% the minimum film thickness could be reduced up to 15%. They also showed that at
elevated pressures and high slide-to-roll ratios the temperature within the lubricant film is significant
and better representations of friction variation with slide-to-roll ratio can be achieved when thermal
effects are included. Kim and Sadeghi (1991) followed the work of Sadeghi and Sui (1989) and presented
results for thermal EHL of point contacts using both Newtonian and non-Newtonian (Ree–Eyring)
fluid models. They demonstrated that for point contacts film thickness is only slightly affected by tem-
perature, but better corroboration between experimental friction measurements and analytical results
is achieved when thermal effects are included in the analysis.
18-10 Theory and Practice of Lubrication and Tribology

The Newtonian fluid assumption overestimates the friction generated in EHL contacts. The
Ree–Eyring fluid model has been used extensively to incorporate the effects of non-Newtonian fluids
in EHL contacts. The Ree–Eyring fluid model is based on the activation energy concept and provides a
physical understanding for the decrease in viscosity with shear rate and the change in viscosity with tem-
perature. However, the Ree–Eyring fluid model does not capture the fact that the shear stress is bounded
by a critical value. Bair and Winer (1979) proposed a viscoelastic lubricant rheology model which limits
the shear stress as the shear rate increases. Lee and Hamrock (1990) used a circular non-Newtonian
fluid model to investigate thermal EHL of line contacts. The Ree–Eyring fluid rheology model was used
by Kim and Sadeghi (1991) to investigate EHL of point contacts. Their results indicate that the pressure
distribution maintains similar attributes to the predictions of the Newtonian fluid model; however, the
pressure spikes are reduced. The film thickness was found to be essentially the same as that predicted
using the Newtonian fluid model.

18.6  Application of EHL Theory to Machine Components


The effectiveness of lubrication in machine elements operating in the EHL regime is governed by calcu-
lating the oil film parameter lambda (λ) which is the ratio of the minimum film thickness to the com-
posite root-mean-square roughness of the surfaces in contact. In this section, examples of film thickness
and lambda ratio calculation for a deep groove ball bearing, a spur gear, and a cam and follower are
provided.
Deep groove ball bearings are the most commonly bearings used in industry. In order to determine
which lubrication regime the bearing operates in, one needs to determine the minimum film thickness
separating the surfaces (i.e., the ball with inner and outer races). Here an approach commonly used to
determine the minimum film thickness for a deep groove ball bearing is highlighted.
A radial load of 22,240 N is applied to the deep groove ball bearing shown in Figure 18.6. The bearing
has 10 balls. The inner race is turning at a rate of 10,000 rpm and the outer race is fixed on the ground.
The lubricant viscosity is 0.007 Pa·s and the pressure–viscosity coefficient is 1.54 × 10−8 Pa−1. The bear-
ing is assumed to have the standard properties for steel. The root-mean-square surface roughness of
the race and roller are 0.0447 and 0.01 μm, respectively. From the given information, several important
parameters can be calculated.
Equivalent modulus of elasticity:

−1 −1
⎛ 1 − n12 1 − n22 ⎞ ⎛ 1 − 0.32 1 − 0.32 ⎞
Eʹ = 2 ⎜ + = 2⎜ + ⎟ = 231 GPa (18.18)
⎝ E1 E2 ⎟⎠ ⎝ 210 GPa 210 GPa ⎠

R 7.9375 -
R 8.255
100.84

69.08

R 8.255

FIGURE 18.6  Dimensions for deep groove ball bearing (dimensions in mm).
Elastohydrodynamic Lubrication 18-11

Pitch diameter:

1
de =
2
(dir + dor ) = 12 (69.08 + 100.84) = 84.963 mm (18.19)

Surface velocities assuming pure rolling conditions:

2p
) = 10, 000 × 60 (84.963 )
2
(
w o − w i de2 − db2 − 15.8752
u= = 21, 466.7 mm/s (18.20)

4de ( 4)(84.963)
Equivalent radii in the rolling direction:

−1 −1
⎛ 1 1 ⎞ 1 1 ⎞
Ror −b, x = ⎜ + =⎛ + = 9.42 mm (18.21)
⎝ Ror , x Rb, x ⎟⎠ ⎝ −50.419 7.9375 ⎠

−1 −1
⎛ 1 1 ⎞ 1 1 ⎞
Ror −b, y = ⎜ + ⎟ =⎛ + = 206.4 mm (18.22)
⎝ Ror , y Rb, y ⎠ ⎝ −8.255 7.9375 ⎠

−1 −1
⎛ 1 1 ⎞ 1 1 ⎞
Rir −b, x = ⎜ + =⎛ + = 6.45 mm (18.23)
⎝ Rir , x Rb, x ⎟⎠ ⎝ 34.544 7.9375 ⎠

−1 −1
⎛ 1 1 ⎞ 1 1 ⎞
Rir −b, y = ⎜ + ⎟ =⎛ + = 206.4 mm (18.24)
⎝ Rir , y Rb, y ⎠ ⎝ −8.255 7.9375 ⎠

Here, d and ω are the diameter and angular velocity while the subscripts ir, or, and b denote inner race,
outer race, and ball, respectively. Using the Stribeck approximation, the force on the ball carrying the
maximum load can be obtained:

5 5
wr ,max = wr = × 22, 240 N = 11,120 N (18.25)
N ball 10

where
wr,max is the force on the ball carrying the maximum load
Nball is the number of balls in the bearing
wr is the radial load applied on the bearing

Following Hamrock and Brewe (1983), the ellipticity parameter can be estimated from the bearing
geometry as

2/p 2/p
⎛ Ror −b, y ⎞ 206.4 ⎞
k =⎜ =⎛ = 7.14 (18.26)
⎝ Ror −b, x ⎟⎠ ⎝ 9.42 ⎠

The Hamrock–Dowson minimum film thickness equation can now be used to determine the minimum
film thickness for the inner and outer race contacts.
18-12 Theory and Practice of Lubrication and Tribology

Outer race ball minimum film thickness:

uh0 21.47 m/s × 0.007 Pa ⋅ s


U= = = 6.91 × 10 −11 (18.27)
E ʹRor −b, x 231 GPa × 9.42 mm

G = aE ʹ = 1.54 × 10 −8 Pa −1 231 GPa = 3554 (18.28)


wr ,max 11,120 N
W= 2
= 2 = 5.42 × 10 −4 (18.29)

E ʹRor −b, x 231 GPa × 9.42 mm( )
H min = 3.63U 0.68G 0.49W −0.073 (1 − e −0.68k ) = 4.22 × 10 −5 (18.30)

hmin = Ror −b, x H min = 9.42 mm × 4.22 × 10 −5 = 0.398 mm (18.31)


hmin 0.398 mm
L= 2 2
= 2 2
= 8. 7 (18.32)
Rq + Rq
or b ( ) (
0.0447 mm + 0.01 mm )

Inner race ball minimum film thickness:

uh0 21.47 m/s × 0.007 Pa ⋅ s


U= = = 1.01 × 10 −10 (18.33)
E ʹRir −b, x 231 GPa × 6.45 mm

G = aE ʹ = 1.54 × 10 −8 Pa −1 231 GPa = 3554 (18.34)


wr ,max 11,120 N
W= = 2 = 1.16 × 10 −3 (18.35)
E ʹRir2 −b, x 231 GPa × 6.45 mm( )

2 /p 2 /p
⎛ Rir −b, y ⎞ ⎛ 206.4 ⎞
k =⎜ =⎜ = 9.08 (18.36)
⎝ Rir −b, x ⎟⎠ ⎝ 6.45 ⎟⎠

H min = 3.63U 0.68G 0.49W −0.073 (1 − e −0.68k ) = 5.20 × 10 −5 (18.37)


hmin = Rir −b, x H min = 6.45 mm × 5.20 × 10 −5 = 0.335 mm (18.38)


hmin 0.335 mm
L= 2 2
= 2 2
= 7.32 (18.39)
Rq + Rq
ir b (0.0447 mm) + (0.01 mm)

In these equations, η 0 is the base viscosity, α is the Barus pressure–viscosity coefficient, and Rq is the
root mean square of surface roughness. Both the inner race and outer race contacts have a sufficient film
thickness and are operating in the EHD regime.
Elastohydrodynamic Lubrication 18-13

Spur gear: The kinematics of meshing gears is complicated for even the simplest gear system. The prob-
lem of meshing spur pinion and gear has been considered by Dowson and Higginson (1966). They show
that at the pitch line the entraining velocity is given by

pw g ⎛ M g l f ⎞
u= siny (18.40)
30 ⎜⎝ M g + 1 ⎟⎠

and the equivalent radius of the meshing gears is given by

Mgl f
R= 2 siny F (18.41)

(M g + 1)

where
Mg is the gear ratio
lf is the distance between the gear wheel centers
ωg is the angular velocity of the gear in rev/min
ψ is the pressure angle

As a particular example, consider a gear and pinion set having radii of Rg = 75 mm and Rp = 60 mm, a face
width of 20 mm, and a pressure angle of 20°. The angular velocity of the gear is ωg = 2500 rev/min and
the teeth transmit a load of 20 kN. The surface roughness of the gears is 0.3 μm. The lubricant viscosity
is 0.08 Pa·s and the pressure–viscosity coefficient is 2.0 × 10−8 Pa−1.
Taking typical properties for steel, the equivalent modulus of elasticity is

−1 −1
⎛ 1 − n12 1 − n22 ⎞ ⎛ 1 − 0.32 1 − 0.32 ⎞
Eʹ = 2 ⎜ + = 2⎜ + ⎟ = 231 GPa (18.42)
⎝ E1 E2 ⎟⎠ ⎝ 210 GPa 210 GPa ⎠

The gear ratio is

Rg 0.075
Mg = = = 1.25 (18.43)
Rp 0.06

and the distance between the wheel centers is lf = Rg + Rp = 0.075 + 0.06 = 0.135 m. Using Equations 18.40
and 18.41, the entraining velocity and equivalent radius are

pw g ⎛ M g l f ⎞ p × 2500 ⎛ 1.25 ⋅ 0.135 ⎞


u= siny = sin 20 = 6.72 m/s (18.44)
30 ⎜⎝ M g + 1 ⎟⎠ 30 ⎝ 1.25 + 1 ⎠

Rg l f 1.25 × 0.135
R= 2 siny = 2 sin 20 = 0.0114 m (18.45)

(R g + 1) (1.25 + 1)

Using Equation 18.12, the minimum film thickness can be determined as

uh0 6.72 m/s × 0.08 Pa ⋅ s


U= = = 2.04 × 10 −10 (18.46)
E ʹR 231 GPa × 0.0114 m

18-14 Theory and Practice of Lubrication and Tribology

G = aE ʹ = 2.0 × 10 −8 Pa −1 231 GPa = 4620 (18.47)


w 20, 000 N/0.02 m


W= = = 3.8 × 10 −4 (18.48)
E ʹR 231 GPa × 0.0114 m

H min = 1.714 U 0.694G 0.568W −0.128 = 1.067 × 10 −4 (18.49)

hmin = RH min = 0.0114 m × 1.067 × 10 −4 = 1.22 mm (18.50)


The lambda ratio is

hmin 1.22 mm
L= = = 2.87 (18.51)
Rqc2 + Rq 2f 2
(0.3 mm) + (0.3 mm)
2

Cam and follower mechanisms are used to convert rotational motion into translational motion. There
are a number of different cam and follower configurations (e.g., cam and flat faced follower, cam and
spherical faced follower, cam and roller follower, etc.). Figure 18.7 illustrates a cam and follower mech-
anism in contact. Diesel engine manufacturers commonly use the cam to operate the fuel injector sys-
tems. To determine the lubrication condition, the entraining velocity at the point of contact between
the cam and follower is needed. In this analysis, the cam is assumed to be operating at 1000 rpm, the
lubricant viscosity is 0.08 Pa·s, the pressure–viscosity coefficient is 1.54 × 10−8 Pa−1, and the normal load
at the point of contact is 2500 N. The entraining velocity for this configuration is given by Matthews
et al. (1996):

uc + u f
uE = (18.52)
2
R 40
0 (m
m)

R 5 (mm)

R 25 (mm)

(a) Isometric view (b) Front view (c) Side view

FIGURE 18.7  Cam and reciprocating follower mechanism.


Elastohydrodynamic Lubrication 18-15

where
uE is the entraining velocity
uc and uf are the cam and follower velocities, respectively, which can be calculated given the cam pro-
file, follower radius, and cam rotational velocity (Matthews et al. 1996; Gohar and Rahnejat 2008)

For the purposes of this example, assume the entraining velocity at the instant shown in Figure 18.7
is 1.8 m/s.
The equivalent elastic modulus and equivalent radius in the rolling direction can be calculated as

−1 −1
⎛ 1 − n12 1 − n22 ⎞ ⎛ 1 − 0.32 1 − 0.32 ⎞
Eʹ = 2 ⎜ + = 2⎜ + ⎟ = 231 GPa (18.53)
⎝ E1 E2 ⎟⎠ ⎝ 210 GPa 210 GPa ⎠

−1 −1
⎛ 1 1 ⎞ 1 1
Rx = ⎜ + ⎟ =⎛ + ⎞ = 4.17 mm (18.54)
⎝ Rc , x R f , x ⎠ ⎝ 25 5 ⎠

−1 −1
⎛ 1 1 ⎞ 1 1 ⎞
Ry = ⎜ + ⎟ =⎛ + = 400 mm (18.55)
⎝ Rc , y R f , y ⎠ ⎝ ∞ 400 ⎠

Rc,x and Rf,x are the radius of curvature of the cam and follower at the point of contact. In this case, the
cam is flat in the y-direction while the follower has a crown on Rf,y = 400 mm. Therefore, the minimum
film thickness can be determined as

uh0 1.8 m/s × 0.08 Pa ⋅ s


U= = = 1.5 × 10 −10 (18.56)
E ʹRx 231 GPa × 4.17 mm

G = aE ʹ = 1.54 × 10 −8 Pa −1 231 GPa = 3557 (18.57)


w 2500 N
W= 2
= 2 = 6.23 × 10 −4 (18.58)

E ʹRx 231 GPa × 4.17 mm ( )
2/p 2/p
⎛ Ry ⎞ 400 ⎞
k =⎜ ⎟ =⎛ = 18.3 (18.59)
⎝ Rx ⎠ ⎝ 4.17 ⎠


( )
H min = 3.63U 0.68G 0.49 W −0.073 1 − e −0.68k = 7.14 × 10 −5 (18.60)

hmin = Rx H min = 4.17 mm × 7.14 × 10 −5 = 0.3 mm (18.61)


The Rq for a typical cam and follower system are 0.12 and 0.11 μm, respectively; thus, the lambda ratio is

hmin 0.3 mm
L= = = 1.84 (18.62)
2 2 2 2
Rq + Rq
c f (0.12 mm) + (0.11 mm)

Thus, the cam and follower contact is operating in the mixed or partial lubrication regime.
18-16 Theory and Practice of Lubrication and Tribology

References
Ai, X. and Cheng, H. S., The influence of moving dent on point EHL contact, STLE Tribology Transactions,
1994, 37, 323–335.
Bair, S. and Winer, W. O., A rheological basis for concentrated contact friction, in Proceedings of 20th Leeds-
Lyon Symposium on Tribology, Lyon, France, 1994, pp. 37–44.
Bair, S. and Winer, W. O., A rheological model for elastohydrodynamic contacts based on primary labora-
tory data, Transactions of ASME F, Journal of Lubrication Technology, 1979, 101(3), 258–265.
Bamberger, E. N., Harris, T. A., Kacmarsky, W. M., Moyer, C. A., Parker, R. J., Sherlock, J. J., and Zaretsky,
E. V., Life Adjustment Factors for Ball and Roller Bearings, An Engineering Design Guide, ASME,
New York, 1971.
Barus, C., Isotherms, isopiestics, and isometrics relative to viscosity, American Journal of Sciences, 1893,
45, 87–96.
Barwell, F. T., The Founder of Modern Tribology (D. M. McDowell and J. D. Jackson, Eds.), Manchester
University Press, England, U.K., 1970.
Bedewi, M. A., Dowson, D., and Taylor, C. M., Elastohydrodynamic lubrication of line contacts sub-
jected to time dependent loading with particular reference to roller bearings and cams and fol-
lower, in Mechanisms and Surface Distress, Proceedings of 12th Leeds-Lyon Symposium on Tribology,
Butterworths, London, U.K., 1986, pp. 289–304.
Blok, H., Theoretical study of temperature rise at surfaces of actual contact under oiliness conditions,
Proceedings of the Institution of Mechanical Engineers General Discussion of Lubrication, 1937, 2,
222–235.
Brandt, A., Multi-level adaptive solutions to boundary-value problems, Mathematics of Computation,
1977, 31(138), 333–389.
Brandt, A. and Lubrecht, A. A., Multilevel matrix multiplication and fast solution of integral equations,
Journal of Computational Physics, 1990, 90, 348–370.
Cameron, A. and Gohar, R., Theoretical and experimental studies of the oil in lubricated point contact,
Proceedings of the Royal Society of London, 1966, A291, 520–536.
Castle, P. and Dowson, D., A theoretical analysis of the starved elastohydrodynamic lubrication prob-
lem for cylinders in line contact, in Proceedings of 2nd IMechE Symposium on Elastohydrodynamic
Lubrication, University of Leeds, England, 1972, pp. 131–137.
Chang, L., Cusano, C., and Conry, T. F., Effects of lubricant rheology and kinematic conditions on micro-
elastohydrodynamic lubrication, Transactions of ASME F, Journal of Tribology, 1989, 111, 344–351.
Chang, L. and Zhai, X., A transient thermal model for mixed film contacts, Tribology Transactions, 2000,
43, 427–434.
Cheng, H. S., A numerical solution of the elastohydrodynamic film thickness in an elliptical contact,
Transactions of ASME F, Journal of Lubrication Technology, 1970, 92(1), 155–162.
Cheng, H. S., Application of elastohydrodynamics of rolling element bearings, ASME Paper 74-DE-32,
American Society of Mechanical Engineers, New York, 1974.
Cheng, H. S., Calculation of elastohydrodynamic film thickness in high-speed rolling and sliding contacts,
Report No. MTI-67TR24, Mechanical Technology Inc., Latham, New York, May 1967.
Cheng, H. S. and Dyson, A., Elastohydrodynamic lubrication of circumferentially ground disks, ASLE
Transactions, 1978, 21(1), 25–40.
Cheng, H. S. and Sternlicht, B., A numerical solution for pressure, temperature and film thickness between
two infinitely long rolling and sliding cylinders under heavy load, Journal of Basic Engineering,
Transactions of ASME, 1965, 87(3), 695.
Chevalier, F., Lubrecht, A. A., Cann, P. M. E., Colin, F., and Dalmaz, G., Film thickness in starved EHL point
contact, Transactions of ASME F, Journal of Tribology, 1998, 120(1), 126–133.
Chittenden, R. J., Dowson, D., Dunn, J., and Taylor, C. M., A theoretical analysis of isothermal EHL con-
centrated contacts: Parts I and II, Proceedings of the Royal Society of London, 1985, A97(12), 245–294.
Elastohydrodynamic Lubrication 18-17

Christensen, H., Elastohydrodynamic theory of spherical bodies in normal approach, Transactions of


ASME, Journal of Lubrication Technology, 1970, 92, 145–154.
Clyens, S., Evans, C. R., and Johnson, K. L., Measurement of viscosity of supercooled liquids at high shear
rates with a Hopkinson torsion bar, Proceedings of the Royal Society of London, 1985, A381, 195–214.
Coy, J. J. and Zaretsky, E. V., Some limitations in applying classical EHD film thickness formulas to a
high-speed bearing, Journal of Lubrication Technology, 1981, 103(2), 295–304.
Crook, A. W., The lubrication of rollers, I, Philosophical Transactions of the Royal Society of London, 1958,
A250, 387–409.
Crook, A. W., The lubrication of rollers. II. Film thickness with relation to viscosity and speed, Philosophical
Transactions of the Royal Society of London, 1961, A254, 223–236.
Dawson, P. H., Effect of metallic contact on the pitting of lubricated rolling surface, Proceedings of the
Institution of Mechanical Engineers, Journal of Mechanical Engineering Sciences, 1962, 4(1), 16–21.
Deolalikar, N. and Sadeghi, F., Fatigue life reduction in mixed lubricated elliptical contacts, Tribology
Letters, 2007, 27(2), 197–209.
Deolalikar, N., Sadeghi, F., and Marble, S., Numerical modeling of mixed lubrication and flash temperature
in EHL elliptical contacts, ASME, Journal of Tribology, 2007.
Dowson, D., Elastohydrodynamic lubrication, interdisciplinary approach to the lubrication of concen-
trated contacts, Spec. Publication No. NASA SP-237, National Aeronautics and Space Administration,
Washington, DC, 1970, p. 34.
Dowson, D. and Ehret, P., Past, present and future studies in elastohydrodynamics, Proceedings of the
Institution of Mechanical Engineers, 1999, 213, 317–333.
Dowson, D., Harrison, P., and Taylor, C. M., The lubrication of automotive cams and followers, in
Mechanisms and Surface Distress, Proceedings of 11th Leeds-Lyon Symposium on Tribology,
Butterworths, London, U.K., 1986, pp. 305–322.
Dowson, D. and Higginson, G. R., A numerical solution to the elastohydrodynamic problem, Journal of
Mechanical Engineering Sciences, 1959, 1(1), 6–15.
Dowson, D. and Higginson, G. R., Elastohydrodynamic Lubrication, Pergamon Press, Oxford, U.K., 1966.
Dowson, D., Higginson, G. R., and Whitaker, A. V., Elastohydrodynamic lubrication—A survey of isother-
mal solutions, Journal of Mechanical Engineering Sciences, 1962, 4(2), 121.
Dowson, D., Higginson, G. R., and Whitaker, A. V., Stress distribution in lubricated rolling contacts, in
Proceedings of IMechE Symposium on Fatigue in Rolling Contact, Paper 6.66, London, England, 1963,
pp. 66–75.
Dowson, D., Taylor, C. M., and Zhu, G., A transient elastohydrodynamic lubrication analysis of a cam and
follower, Journal of Physics D, Applied Physics, Frontiers Tribology, 1992, 25, A313–A320.
Dowson, D. and Wang, D., An analysis of the normal bouncing of a solid elastic ball on an oily plate, Wear,
1994, 179, 29–37.
Dowson, D. and Whittaker, B. A., A numerical procedure for the solution of the elastohydrodynamic prob-
lem of rolling and sliding contacts lubricated by a Newtonian fluid, Proceedings of the Institution of
Mechanical Engineers, 1965, l80(3B), 57.
Dowson, U. and Higginson, G. R., Elastohydrodynamic Lubrication, Pergamon Press, Oxford, U.K., 1977.
Ehret, P., Dowson, B., and Taylor, C. M., On lubricant transport conditions in elastohydrodynamic con-
junctions, Proceedings of the Royal Society of London, 1998, A454, 763–787.
Ehret, P., Dowson, D., Taylor, C. M., and Wang, D., Analysis of isothermal elastohydrodynamic point
contacts lubricated by Newtonian fluids using multigrid methods, Proceedings of the Institution of
Mechanical Engineers, Part C, Journal of Mechanical Engineering Sciences, 1997, 211(C7), 493–508.
Ertel, A. M. Hydrodynamic lubrication based on new principles, Akad. Nauk SSSR. Prikadnaya Mathem­
atica i Mekhanika, 1939, 3(2), 41–52.
Evans, H. P. and Snidle, R. W., Inverse solution of Reynolds equation of lubrication under point contact
elastohydrodynamic conditions, Transactions of ASME F, Journal of Lubrication Technology, 1981,
103(4), 539–546.
18-18 Theory and Practice of Lubrication and Tribology

Evans, H. P. and Snidle, R. W., The elastohydrodynamic lubrication of point contacts at heavy loads,
Proceedings of the Royal Society of London, 1982, A382, 183–199.
Foord, C. A., Wedeven L. D., Westlake, F. J., and Cameron A., Optical elastohydrodynamics, Proceedings of
the Institution of Mechanical Engineers, 1969, 184(1), 487–505.
Fowles, P. E., The application of elastohydrodynamic theory to individual asperity–asperity collisions,
Journal of Lubrication Technology, Transactions of ASME, 1969, 91, 464–476.
Gao, J., Lee, S. C., Ai, X., and Nixon, H., An FFT-based transient flash temperature model for general
three-dimensional rough surface contacts, Journal of Tribology, 2000, 122, 519–523.
Glovnea, R. P. and Spikes, H. A., Elastohydrodynamic film collapse during rapid deceleration—Part I:
Experimental results, ASME Journal of Tribology, 2001, 123, 254–261.
Goglia, P. R., Conry, T. F., and Cusano, C., The effect of surface irregularities on the elastohydrodynamic
lubrication sliding line contacts. Part I—Single irregularities, Transactions of ASME F, Journal of
Lubrication Technology, 1981, 106, 104–112.
Goglia, P. R., Conry, T. F., and Cusano, C., The effect of surface irregularities on the elastohydrody-
namic lubrication sliding line contacts. Part II—Wavy surfaces, Transactions of ASME F, Journal of
Lubrication Technology, 1981, 106, 113–119.
Gohar, R. and Cameron, A., Optical measurement of oil film thickness under EHD lubrication, Nature,
1963, 200, 458–459.
Gohar, R. and Cameron, A., The mapping of EHD contacts, ASLE Transactions, 1967, 10, 214.
Gohar, R. and Cameron, A., Theoretical and experimental studies of oil film in lubricated point contact,
Proceedings of the Royal Society of London A, 1966, 291, 520–536.
Gohar, R. and Rahnejat, H., Fundamentals of Tribology, Imperial College Press, London, U.K., 2008.
Greenwood, J. and Kanzlarich, J., Inlet shear heating in elastohydrodynamic lubrication, Journal of
Lubrication Technology, Transactions of ASME, 1973, 95(4), 417.
Greenwood, J. A. and Morales-Espejel, G. E., The behaviour of transverse roughness in EHL contacts,
Proceedings of the Institution of Mechanical Engineers, Part J, Journal of Engineering Tribology, 1994,
208(J2), 121–132.
Grubin, A. N., Contact Stresses in Toothed Gears and Worm Gears, Central Scientific Research Institute for
Technology and Mechanical Engineering, Book No. 30, Moscow (DSIR English Translation No. 337.
As communicated by Prof. M. M. Krushchov to Prof. A. Cameron, Grubin’s contribution was origi-
nally studied by A. M. Ertel and after his death was seen through the press by his coworker Grubin
and is thus often known as Grubin’s name alone.)
Grubin, A. N., Fundamentals of the hydrodynamic theory of lubrication of heavily loaded cylindrical
surfaces, in Investigation of the Contact of Machine Components, Central Scientific Research Institute
for Technology and Mechanical Engineering (Moscow), Book No. 30 (DSIR Translation No. 337)
(Kh. F. Ketova, Ed.), 1949, pp. 115–166.
Grubin, A. N. and Vinogradova, I. E., Fundamentals of the Hydrodynamic Theory of Lubrication of Heavily
Loaded Cylindrical Surfaces, Central Scientific Research Institute for Technology and Mechanical
Engineering, Book No. 30. Moscow, Translation No. 337 into English by the Department of Science
and Industrial Research, U.K., 1949.
Guangteng, G., Cann, P. M., Olver, A. V., and Spikes, H. A., An experimental study of film thickness between
rough surfaces in EHD contacts, Tribology International, 2000, 33(3–4), 183–189.
Gupta, P. K., Cheng, H. S., Zhu, D., Forster, N. H., and Schrand, J. B., Visco-elastic effect in Mil-L-7808 type
lubricant, Part I: Analytical formulation, STLE Tribology Transactions, 1991, 34(4), 608–617.
Hamrock, B. J. and Brewe, D. E., Simplified solution for stresses and deformations, Journal of Lubrication
Technology, 1983, 105(2), 171–177.
Hamrock, B. J. and Dowson, D., Ball Bearing Lubrication. The Elastohydrodynamics of Elliptical Contacts,
John Wiley, New York, 1981.
Hamrock, B. J. and Dowson, D., Isothermal elastohydrodynamic lubrication of point contact. Part I—
Theoretical formulation, Journal of Lubrication Technology, 1976, 98(2), 223–229.
Elastohydrodynamic Lubrication 18-19

Hamrock, B. J. and Dowson, D., Isothermal elastohydrodynamic lubrication of point contacts. Part I—
Theoretical formulation, Transactions of ASME F, Journal of Tribology, 1976, 98(4), 223–229.
Hamrock, B. J. and Dowson, D., Isothermal elastohydrodynamic lubrication of point contacts.
Part II—Ellipticity parameter results, Transactions of ASME F, Journal of Tribology, 1976, 98(3),
375–383.
Hamrock, B. J. and Dowson, D., Isothermal elastohydrodynamic lubrication of point contacts.
Part III—Fully flooded results, Transactions of ASME F, Journal of Tribology, 1977, 99(2), 264–276.
Hamrock, B. J. and Dowson, D., Isothermal elastohydrodynamic lubrication of point contacts. Part IV—
Starvation results, Transactions of ASME F, Journal of Tribology, 1977, 99(1), 15–23.
Hamrock, B. J., Schmid, S. R., and Jacobson, B. O., Fundamentals of Fluid Film Lubrication, 2nd edn.,
McGraw-Hill, New York, 2004.
Hertz, H., On the contact of elastic solid, Journal für die reine und angewandte Mathematik, 1882, 92,
156–171.
Hoglund, E. and Jacobson, B., Experimental investigation of the shear strength of lubricant subjected to
high pressure, Transactions of ASME, Journal of Lubrication Technology, 1986, 108, 571–578.
Hoglund, E. and Larsson, R., Modelling non-steady EHL with focus on lubricant density, in
Elastohydrodynamics ‘96, Fundamentals and Application in Lubrication and Traction, Proceedings of
23rd Leeds-Lyon Symposium, Leeds, U.K., 1997, pp. 511–521.
Holmes, M. J. A., Transient analysis of the point contact elastohydrodynamic lubrication problem using
coupled solution methods, PhD thesis, Cardiff University, Wales, 2002.
Holmes, M. J. A., Evans, H. P., Hughes, T. G., and Snidle, R. W., Transient elastohydrodynamic point con-
tact analysis using a new coupled differential deflection method part 1: Theory and validation,
Proceedings of the Institution of Mechanical Engineers, Part J: Journal of Engineering Tribology, 2003,
217(4), 289–304.
Holmes, M. J. A., Evans, H. P., Hughes, T. G., and Snidle, R. W., Transient elastohydrodynamic point con-
tact analysis using a new coupled differential deflection method part 2: Results, Proceedings of the
Institution of Mechanical Engineers, Part J: Journal of Engineering Tribology, 2003, 217(4), 305–322.
Hu, Y. Z., Barber, G. C., and Zhu, D., Numerical analysis for the elastic contact of real rough surfaces,
Tribology Transactions, 1999, 42(3), 443–452.
Hu, Y. and Zhu, D., A full numerical solution to mixed lubrication in point contacts, Journal of Tribology,
2000, 122, 1–9.
Hua, D. V. and Khonsari, M. M., Application of transient elastohydrodynamic lubrication analysis for gear
transmissions, STLE, Tribology Transactions, 1995, 38(4), 905–913.
Hua, D. Y., Zhang, H. H., and Zhan, S. L., Transient elastohydrodynamic lubrication of involute spur gear,
in Proceedings of the Japan International Tribology Conference, Nagoya, Japan, 1990, pp. 1641–1646.
Ioannides, E., Bergling, G., and Gabelli, A., An analytical formulation for the life of rolling bearings, Acta
Polytechnica Scandinavica, Mechanical Engineering Series 137, The Finnish Academy of Technology,
Tekniikantie 12, FIN-02150 ESPP, Finland, 1999.
Ioannides, E. and Harris, T. A., A new fatigue life model for rolling bearings, Journal of Tribology, 1985,
107, 367–378.
Jacobson, B., On the lubrication of heavily loaded spherical surfaces considering surface deformations
and solidification of the lubricant, Acta Polytechnica Scandinavica Mechanical Engineering Series,
1970, 54.
Jacobson, B. O., A high pressure-short time shear strength analyser for lubricants, Transactions of ASME F,
Journal of Tribology, 1985, 107, 220–223.
Jaeger, J. C., Moving sources of heat and the temperature at sliding contacts, Journal of Proceedings of the
Royal Society of New South Wales, 1943, 76, 203–224.
Jiang, X., Hua, D. Y., Cheng, H. S., Ai, X., and Lee, S. C., A mixed elastohydrodynamic lubrication model
with asperity contact, Journal of Tribology, 1999, 121, 481–491.
Johnson, K. L., Contact Mechanics, Cambridge University Press, Cambridge, U.K., 1985.
18-20 Theory and Practice of Lubrication and Tribology

Johnson, K. L., Correlation of theory and experiment in research on fatigue in rolling contact, in Proceedings
of IMechE Symposium on Fatigue in Rolling Contact, London, England, 1963, pp. 155–159.
Johnson, K. L. and Cameron, R., Shear behavior of elastohydrodynamic oil film at high rolling contact
pressures, Proceedings of the Institution of Mechanical Engineering, 1967, 182, 307–319.
Johnson, K. L. and Greenwood, J. A., Thermal analysis of an Eyring fluid in elastohydrodynamic traction,
Wear, 1980, 61, 353–374.
Johnson, K. L., Greenwood, J. A., and Poon, S. V., A simple theory of asperity contact in elastohydrody-
namic lubrication, Wear, 1972, 19, 91–108.
Johns-Rahnejat, P. M. and Gohar, R., Measuring contact pressure distribution under elastohydrodynamic
point contacts, Tribotest, 1994, 1, 33–53.
Kaneta, M., Necessity of reconstruction of EHL theory, Japan Journal of Tribology, 1993, 38, 860–868.
Kaneta, M., Nishikawa, H., Kameishi, K., Sakal, T., and Ohno, N., Effects of elastic moduli of contact sur-
faces in elastohydrodynamic lubrication, Transactions of ASME F, Journal of Tribology, 1992, 114,
75–80.
Kannel, J. W., Measurements of pressures in rolling contact, Proceedings of the Institution of Mechanical
Engineers, 1965–1966, Elastohydrodynamic Lubrication, 1966, 180(3B), 135–142.
Kannel, J. W., Zugaro, F. F., and Dow, T. A., A method for measuring surface temperature between rolling/
sliding steel cylinders, Journal of Lubrication Technology, Transactions of ASME, 1978, 100(1), 100.
Kim, K. and Sadeghi, F., Non-Newtonian elastohydrodynamic lubrication of point contact, ASME Journal
of Tribology, 1991, 113, 703–711.
Kim, K. and Sadeghi, F., Three dimensional temperature distribution in EHD lubrication, ASME Journal
of Tribology, 1992, 114, 32–41.
Koye, K. A. and Winer, W. O., An experimental evaluation of the Hamrock and Dowson minimum film
thickness equation for fully flooded EHD point contacts, Journal of Lubrication Technology, 1981,
103, 284–294.
Kumar, A., Sadeghi, F., and Krousgrill C. M., Effects of surface roughness on normal contact compres-
sion response, Proceedings of the Institution of Mechanical Engineers, Part J, Journal of Engineering
Tribology, 2006, 220, 65–77.
Kweh, C. C., Evans, H. P., and Snidle, R. W., Microelastohydrodynamic lubrication of elliptical contact with
transverse and three-dimensional roughness, Transactions of ASME F, Journal of Tribology, 1989,
111, 577–584.
Lee, R.-T. and Hamrock, B. J., A circular non-Newtonian model: Part I—Used in elastohydrodynamic
lubrication, Transactions of ASME F, Journal of Tribology, 1990, 112, 386–496.
Lee, R.-T. and Hamrock, B. J., A circular non-Newtonian model: Part II—Used in microelastohydrody-
namic lubrication, Transactions of ASME F, Journal of Tribology, 1990, 112, 497–505.
Lubrecht, A. A., Numerical solution of the EHL line and point contact problem using multigrid tech-
niques, PhD thesis, Twente University, Enschede, the Netherlands, 1987.
Lubrecht, A. A., ten Napel, W. E., and Bosma, R., Multigrid, an alternative method of solution for
two-dimension elastohydrodynamically lubricated point contact calculations, Transactions of
ASME F, Journal of Tribology, 1986, 108(3), 551–556.
Lundberg, G. and Palmgren, A., Dynamic capacity of roller bearings, Acta Polytechnica—Mechanical
Engineering Series, 1952, 2, 1–32.
Majumbar, B. C. and Hamrock, B. J., Effect of surface roughness on elastohydrodynamics line contact,
Transactions of ASME, Journal of Lubrication Technology, 1982, 104(3), 401–409.
Martin, H. M., Lubrication of gear teeth, Engineering, London, 1916, 102, 199.
Matthews, J. A. and Sadeghi, F., Kinematics and lubrication of camshaft roller follower mechanisms, Society
of Tribologists and Lubrication Engineers, Tribology Transactions, 1996, 39(2), 425–433.
Matthews, J. A., Sadeghi, F., and Cipra, R. J., Radius of curvature and entraining velocity of cam follower
mechanisms, Society of Tribologists and Lubrication Engineers, Tribology Transactions, 1996, 39(4),
899–907.
Elastohydrodynamic Lubrication 18-21

Moes, H. and Bosma, R., Film thickness and traction in EHL at point contact, in Proceedings of 1972 IMechE
Symposium on Elastohydrodynamic Lubrication, University of Leeds, England, 1972, pp. 149–152.
Morales-Espejel, G. E., Greenwood, J., and Melgar, J. L., Kinematics of roughness in EHL, in Proceedings of
22nd Leeds-Lyon Symposium on Tribology, Lyon, France, 1996, pp. 501–513.
Mostofi, A. and Gohar, R., Oil film thickness and pressure distribution in elastohydrodynamic point con-
tacts, Journal of Mechanical Engineering Sciences, 1982, 24(4), 173–182.
Murch, L. E. and Wilson, W. R. B., A thermal elastohydrodynamic inlet zone analysis, Journal of Lubrication
Technology, Transactions of ASME, 1975, 97(2), 212–216.
Nagaraj, H. S., Sanborn, D. M., and Winer, W. O., Direct surface temperature measurement by infrared
radiation in elastohydrodynamic contacts and the correlation with the Block temperature theory,
Wear, 1978, 49(1), 43–59.
Nickel, D. A., Experimental methods for the assessment of tribological hard contacts at the device, compo-
nent, and surface scales, PhD thesis, Purdue University, West Lafayette, IN, 1999.
Okamura, H., A contribution to the numerical analysis of isothermal elastohydrodynamic lubrication,
in Proceedings of 9th Leeds—Lyon Symposium on Tribology, University of Leeds, England, 1982,
pp. 313–320.
Osborn, K. F. and Sadeghi F., Time dependent line EHD lubrication using the multigrid/multilevel tech-
nique, Journal of Tribology, 1992, 114(1), 68–74.
Pan, P. and Hamrock, B. J., Simple formulas for performance parameters used in elastohydrodynamically
lubricated line contacts, Journal of Tribology, 1989, 111(2), 246–251.
Patir, N. and Cheng, H. S., An average model for determining effects of three dimensional roughness on
partial hydrodynamic lubrication, Transactions of ASME, Journal of Lubrication Technology, 1978,
100, 12–17.
Patir, N. and Cheng, H. S., Effect of surface orientation on the central film thickness in EHD contacts, in
Proceedings of 5th Leeds-Lyon Symposium, Leeds (D. Dowson, C. M. Taylor, M. Godet, and D. Berthe,
Eds.), University of Leeds, England, 1978, pp. 15–21.
Persson, B. N. J., Sliding Friction: Physical Principles and Applications, Springer-Verlag, Berlin, Heidelberg,
New York, 2000.
Petrusevich, A. I., Fundamental conclusions from the contact-hydrodynamic theory of lubrication,
Izvestiya Akademii Nauk SSSR (OTN), 1951, 2, 209–223.
Ramesh, K. T., On the rheology of a traction fluid, Transactions of ASME F, Journal of Tribology, 1989, 111,
614–619.
Reynolds, O., On the theory of lubrication and its application to Mr. Beauchamp Tower’s experiments,
including an experimental determination of viscosity of olive oil, Philosophical Transactions of the
Royal Society of London, 1886, 177, 157–234.
Rodkiewicz, C. M. and Srinivanasan, V., EHD lubrication in rolling and sliding contacts, Journal of
Lubrication Technology, Transactions of ASME, 1972, 94(4), 324.
Roelands, C. J. A., Vlugter, J., and Waterman, H., The viscosity–temperature–pressure relationship of lubri-
cating oils and its correlation with chemical constitution, Journal of Basic Engineering, 1963, 85(4),
601–610.
Sadeghi, F., A comparison of the fluid models effect on the internal stresses of rough surfaces, Transactions
of ASME F, Journal of Tribology, 1991, 113, 142–149.
Sadeghi, F. and Sui, P. C., Compressible elastohydrodynamic lubrication of rough surfaces, Transactions of
ASME F, Journal of Tribology, 1989, 111, 56–62.
Safa, M. M. A., Anderson, J. C., and Leather, J. A., Transducers for pressure, temperature and oil film thick-
ness measurement in bearings, Sensors and Actuators, 1982, 3, 119–128.
Sibley, L. B. and Orcutt, F. K., Elastohydrodynamic lubrication of rolling contact surfaces, American Society
of Lubrication Engineering Transactions, 1961, 4(2), 234–249.
Smeeth, M. and Spikes, H. A., Central and minimum elastohydrodynamic film thickness at high contact
pressure, Transactions of the ASME, Journal of Tribology, 1997, 119(2), 291–296.
18-22 Theory and Practice of Lubrication and Tribology

Spikes, H. A., Mixed lubrication—An overview, in Proceedings of Symposium on Tribology-Solving Friction


and Wear Problems, Esslingen, Germany, 1996, pp. 1713–1735.
Stachowiak, G. W. and Batchelor A. W., Engineering Tribology, 2005.
Stanley, H. M. and Kato, T., An FFT-based method for rough surface contact, Journal of Tribology, 1997,
119, 481–485.
Tallian, T. E., On competing failure modes in rolling contact, Transactions of ASLE, 1967, 10, 418–439.
Tian, X. and Kennedy, F. E., Jr., Maximum and average flash temperatures in sliding contacts, Journal of
Tribology, 1994, 116, 167–174.
Venner, C. H., Multilevel solution of the EHL line and point contact problems, PhD thesis, Twente
University, Enschede, the Netherlands, 1991.
Venner, C. H. and Lubrecht, A. A., Numerical simulation of a transverse ridge in a circular EHL contact
under rolling/sliding, Transactions of ASME F, Journal of Tribology, 1994, 116, 751–761.
Venner, C. H. and Lubrecht, A. A., Numerical simulation of waviness in a circular EHL contact, under
rolling/sliding, in Lubricant and Lubrication, Proceedings of 21st Leeds-Lyon Symposium on Tribology,
Leeds, U.K., 1994, pp. 259–272.
Wang, Q. J., Zhu, D., Cheng, H. S., Yu, T., Jiang, X., and Liu, S., Mixed lubrication analysis by macro–micro
approach and a full-scale mixed EHL model, Journal of Tribology, 2004, 126, 81–91.
Wedevan, L. D., Optical measurements in EHD rolling-contact bearings, PhD thesis, University of London,
London, U.K., March 1970.
Wedeven, L. D., Evans, D., and Cameron, A., Optical analysis of ball bearing starvation, Transactions of
ASME, Journal of Lubrication Technology, 1971, 97, 321–363.
Weibull, W., The phenomenon of rupture in solids, Ingeniors Vetenskaps Akademien Handlingar, 1939, 153.
Whitehouse, D. J. and Archard, J. F., The properties of random surfaces of significance in their contact,
Proceedings of the Royal Society of London, 1970, A316, 97–121.
Winer, W. O., Temperature effects in elastohydrodynamically lubricated contacts, in Tribology in the 80’s,
NASA CP-2300, Vol. II, Cleveland, OH, 1983, pp. 533–543.
Xu, G. and Sadeghi, F., Thermal EHL analysis of circular contacts with measured surface roughness,
Transactions of ASME, Journal of Tribology, 1996, 118, 473–483.
Zhao, J. and Sadeghi, F., Analysis of EHL circular contact start up: Part II—Surface temperature rise model
and results, Journal of Tribology, 2001, 123, 75–82.
Zhao, J. and Sadeghi, F., The effects of a stationary surface pocket on EHL line contact start up, Journal of
Tribology, 2002, 126, 672–680.
Zhao, J., Sadeghi, F., and Hoeprich, M. H., Analysis of EHL circular contact start up: Part I—Mixed contact
model with pressure and film thickness results, Journal of Tribology, 2001, 123, 67–74.
19
Mixed Lubrication
19.1 Introduction..................................................................................... 19-1
19.2 Running-In (Break-In).................................................................... 19-3
19.3 Experimental Methods and Tests.................................................19-6
19.4 Wear and Durability under Mixed Lubrication
Conditions..................................................................................19-11
Scuffing under Mixed Lubrication Conditions  •  Rolling Contact
Fatigue under Mixed Lubrication Conditions  •  Micropitting
Martin N. Webster 19.5 Modeling Mixed Lubrication Problems..................................... 19-16
ExxonMobil Research 19.6 Mixed Lubrication: The Future................................................... 19-17
and Engineering References................................................................................................... 19-17

19.1  Introduction
As its name implies, mixed lubrication refers to conditions in which there are multiple mechanisms or
regimes of lubrication occurring within a single contact. Most frequently, there are two mechanisms:
one being hydrodynamic lubrication that relies on the rheological properties of the fluid and the result-
ing entrainment into the contact and the other being boundary lubrication that is governed by the
local contact mechanics and friction of asperity-level interactions. The hydrodynamic action can be in
the low-pressure regime that occurs between conformal or compliant surfaces or in the elastohydro-
dynamic lubrication (EHL) regime that occurs between nonconforming contacts. The nature of the
asperity contacts is determined by the surface mechanical properties, surface morphology, and any
environmental conditions that modify the surface chemistry within the contact. The latter includes the
effect of surface-active components brought into the contact by a liquid lubricant. Thus, in mixed lubri-
cation the load is supported by a combination of the fluid pressure typically generated by hydrodynamic
action and by the local stresses generated by the sum of the individual asperity contacts.
There is no universally accepted definition of the conditions under which mixed lubrication occurs.
However, it is most conveniently represented on a Stribeck curve as shown in Figure 19.1. As a general
guide, mixed lubrication occurs when the combined roughness of the mating surfaces is similar in
value to the thickness of the lubricant film. In EHL contacts, the specific film thickness (λ as defined in
Equation 19.1) is often used to define the condition of lubrication:

h
l= (19.1)
s

where
h is the lubricant film thickness
s = s 12 + s 22 , where σ1 and σ 2 are the root mean square (rms) surface roughness values for the two
respective surfaces in the conjunction.

19-1
19-2 Theory and Practice of Lubrication and Tribology

Boundary lubrication Mixed lubrication Liquid films


Extensive surface Some surface contact Good surface separation
contact

e
Friction

ns
spo
re
ic
am
d yn
dro
Hy nse
respo
EHL

Increasing load
Increasing speed and
lubricant viscosity
Lubricant film thickness
Surface roughness

FIGURE 19.1  Stribeck curve showing the relative friction response ranging from full film to mixed and bound-
ary lubrication conditions.

It is often assumed that λ values of 3 or greater correspond with fully hydrodynamic lubrication con-
ditions. However, depending on the distribution of asperity heights for each of the two surfaces, even
at this degree of separation there are still some surface interactions that take place within the contact.
Similarly, a l of <0.1 is typically assumed to correspond to conditions where boundary lubrication

dominates. However, even in this case the effects of running-in and surfaces that exhibit highly skewed
height distributions can substantially reduce the dominance of surface interactions. Thus, we must view
the transition from one mode of lubrication to another as part of a continuum for which precise defini-
tions of the boundaries cannot be easily quantified.
The mixed lubrication region has an important influence on many fundamental tribological pro-
cesses. It frequently marks the transition region between very low wear rates at high λ to the onset of
wear at lower l and even catastrophic failure if the conditions promote local film collapse. This is true

for almost all types of wear processes with the possible exceptions of the erosive and corrosive forms of
wear. The region also marks the transition in which surface chemistry and the formation of antiwear
and boundary films can start to influence the behavior of the contact. Likewise, it is also the region in
which the total friction developed within the contact relies on a combination of fluid film properties and
the contact mechanics of surface asperity interactions.
Studies of mixed lubrication problems have been advanced by developments in both experimental
and theoretical methods. The development of new measurement techniques, including the emergence of
so-called nanotools such as the atomic force microscope (AFM) and nanoindenters, enables us to define
in much greater detail the morphology and properties of surfaces. This includes the ability to study
idealized single asperity like contacts in isolation. Similarly, ever-increasing computational power has
spurred advances in models that can now predict tribological performance even in the highly complex
mixed lubrication regime.
In practice, the mixed lubrication regime represents one of the most challenging and complex
environments to study. It is also of great practical importance since it influences the fundamental
performance attributes of many everyday machine elements, such as bearings, gears, cams, piston
rings, etc. In this section, we will examine the areas where the study of mixed lubrication has con-
tributed to broadening our knowledge resulting in better practices, new specifications, and new test
methods.
Mixed Lubrication 19-3

19.2  Running-In (Break-In)


When an engineering component is manufactured, the surface topography has essentially been set by
the final finishing process. While this may have been tightly controlled to meet predetermined specifi-
cations, it is highly unlikely that it is going to conform perfectly with its mating surface. If the lubrica-
tion conditions fail to fully separate these new surfaces, there will be significant interactions between
the surface asperities. During these interactions, the surfaces will undergo significant changes in both
morphology and surface characteristics. A number of effects can come into play, including local plastic
deformation, removal of oxide or other films, abrasive wear, adhesive wear, or in some cases a micro-
scopic surface fatigue mode referred to as micropitting. These effects tend to be more severe under high
sliding/high load conditions. Another effect is related to the formation of surface films by the chemical
reaction of lubricant components, sometimes referred to as tribofilms. It is beyond the scope of this sec-
tion to fully describe the various kinds of surface films that might be formed, but a common feature is
that it takes a finite time for such films to form. Eventually, the changes in both surface morphology and
the generation of surface films can reach a near-steady state. Reaching this point is often called running-
in (or alternatively breaking-in) and can be a crucial factor in determining the overall durability of the
system components.
Manufacturers will often define the break-in procedure for their equipment. The most common prac-
tice is to follow a series of speed and load cycles that vary over time. The goal is to avoid catastrophic
failure or high wear rates while at the same time promoting sufficient wear to allow the surfaces and
parts to conform to one another. It is important that by the completion of this process the equipment
has been cycled over the range of operating conditions that will occur in service. This ensures that all
the load conditions and thermal strains have reached their operational limits and the moving parts have
been fully loaded against one another at the appropriate points.
Table 19.1 shows an example break-in sequence for a gasoline internal combustion engine. This
method is used to prepare engines for the ASTM sequence VID engine fuel economy test [1]. In this

TABLE 19.1  Sequence VID New Engine Cycle Break-In


Cycle
Parameter A B
Time at each step, min 4 1
Time to deceleration to Step A, s 15 max
Time to acceleration to Step B, s 15 max
Speed, rev/min 1500 ± 50 3500 ± 50
Power, kW 6 16.5
Torque, Nm 38.0 ± 5 45.0 ± 5
Oil gallery, °C 80 ± 2 80 ± 2
Coolant in, °C 80 ± 2 80 ± 5
Coolant flow, L/min 80 ± 5 80 ± 5
Intake air temperature and humidity control Control not required
Exhaust back pressure, kPa 105 Not specified
AFR record Record Not specified
Fuel pressure to fuel rail, kPa 405 ± 10 405 ± 10
Fuel temperature to fuel rail, °C 22 ± 2 22 ± 2
Fuel flow, kg/h Not specified Not specified
BSFC, kg/kWh Not specified Not specified
Source: ASTM D7589-11, Standard test method for measurement of effects of auto-
motive engine oils on fuel economy of passenger cars and light-duty trucks in sequence
VID spark ignition engine.
19-4 Theory and Practice of Lubrication and Tribology

TABLE 19.2  Break-In Sequence for Mack T-12 Diesel Engine Test
Step Time, h:mm:ss Speed, rpm Torque, N·m Comments
Prior to start Set injection timing to 21° BTDC and full EGR bypass, EGR valve
closed, VGT pressure below 414 kPa
1 0:00:00 Idle 0 Engine idle, waiting for oil pressure
0:00:10 Idle 0 Proceed if oil pressure >138 kPa
2 0:00:11 Idle 245 Engine idle, set torque to 245, hold conditions for 4 min 50 s
3 0:05:00 1200 245 Set speed to 1200, linearly ramp torque to 815 in 4 min
0:09:00 1200 815 End of torque ramp, hold conditions for 2 min 30 s
4 0:11:30 1200 815 Linearly ramp torque to 1085 in 2 min
0:13:30 1200 1085 End of torque ramp, hold conditions for 2 min 30 s
5 0:16:00 1200 1085 Linearly ramp torque to 2440 in 10 min. Open EGR valve, set EGR
0:26:00 1200 2400 End of torque ramp, hold conditions for 2 min 30 s
0:28:30 1200 2400 Hold conditions for 30 min
6 0:58:30 1200 2400 Linearly ramp torque to 1300 in 2 min
0:59:00 1200 Ramping Linearly ramp speed to 1800 in 2 min
1:00:30 Ramping 1300 End of torque ramp
1:01:00 1800 1300 End of speed ramp, hold conditions for 2 min 30 s
1:03:30 1800 1300 Set injection timing to 6° BTDC and EGR, hold conditions for
30 min
7 1:33:30 1800 1300 Proceed to shutdown sequence
Source: ASTM D7422, Standard test method for evaluation of diesel engine oils in T-12 exhaust gas recirculation diesel
engine.

case, there are two load cycles defined and a transition period between them. During the transitions, the
various engine surfaces are forced to conform to varying load conditions which facilitates the breaking-
in process. Table 19.2 shows an example of a progressive break-in sequence in which speed and torque
(i.e., load) are ramped over a fixed series of steps. This sequence is used to condition the engine for the
Mack T-12 diesel engine test [2].
In order to understand the break-in process, it is necessary to study the changes that occur at the
surface as this ultimately controls the friction and wear performance. This is most conveniently accom-
plished by running simple benchtop or equipment tests where it is relatively easy to remove components
for analysis. Katoh et al. [3] describes a systematic study using a vane-on-disk sliding wear test. In these
tests, the friction force was monitored over time in order to investigate the differences between longi-
tudinal and transverse roughness. This was accomplished by resolving the friction force as a function
of circumferential distance around the disk. Since the disk was ground in one direction, the relative
direction of the grinding marks and the sliding direction varied as the vanes rotated around the disk.
Figure 19.2 shows their friction results for the mixed lubrication regime. These data show the typical
reduction in friction as the break-in process takes place. Also revealed is the fact that the orientation of
surface topography has a significant influence on friction in this regime even when the nominal specific
film thickness is identical. The instantaneous friction measurements exhibit a significant amount of
scatter which is also typical for such experiments.
Tracking wear during break-in is more challenging. Katoh et al. [3] accomplished this by measuring
the surface finish at different stages of their wear test. They suggest that the asperity-level wear can be
calculated from the 3-D bearing area curves generated from the surface topography data collected at
each stage. Figure 19.3 shows the resulting wear results. Similar to the friction results, the wear rate
approaches a near constant value after a relatively short time. However, unlike the friction results there
is little difference between transverse and longitudinal roughness.
Clearly, surface morphology changes are key to the behavior of the system during the break-in period.
In a landmark work, Greenwood and Williamson [4] studied the contact between randomly rough
Mixed Lubrication 19-5

2.2
Load 15 N
Transverse
2.0
Longitudinal
1.8

Friction force, N
1.6
1.4
1.2
1.0
0.8
0 10 20 30 40 50 60 70 80 90
Sliding time, min
(b) Λ = 0.137

FIGURE 19.2  Friction versus time results comparing the effects of transverse versus longitudinal roughness.
(From Katoh, J. et al., Tribol. Trans., 44, 104, 2001. With permission.)

6000

5000 Λ = 0.06 – 0.07


Wear volume of disk, µm3/mm2

4000

Λ = 0.11 – 0.13
3000

2000

Longitudinal
1000
Transverse

0
0 10 20 30 40 50 60 70
Sliding time, min

FIGURE 19.3  Wear versus time results calculated from surface bearing area curve measurements. (From Katoh, J.
et al., Tribol. Trans., 44, 104, 2001. With permission.)

surfaces. They modeled the system as an equivalent rough surface, which combined the height variation
of both surfaces, loaded against a flat surface. Using an average asperity radius and Hertzian contact
analysis, they were able to define the plasticity index ψ given in the following equation:

EGW
ʹ ss
y = (19.2)
Hʹ b

where
1 1 − v12 1 − v22
= + , where E1 and E2 are the elastic modulus of surfaces 1 and 2, and ν1 and ν2 are
EGW
ʹ E1 E2
the Poisson’s ratio for surfaces 1 and 2
β is the average asperity radius
H′ is the hardness of the material
19-6 Theory and Practice of Lubrication and Tribology

Fresh soft roller Wom soft roller


Fresh hard roller Wom hard roller

2.60E–07 0.18

0.16
2.50E–07
0.14
2.40E–07
0.12
2.30E–07
0.1
Ra (µm)

Ra (µm)
2.20E–07
0.08
2.10E–07
0.06
2.00E–07 0.04

1.90E–07 0.02

1.80E–07 0
Uroll = 0.1 m/s Uroll = 0.2 m/s Uroll = 0.1 m/s Uroll = 0.2 m/s
(a) (b)

FIGURE 19.4  Before and after roughness measurements for hard and soft roller used in a break-in experiment.
(From Akbarzadeh, S. and Khonsari, M.M., ASME J. Tribol., 132, 2010. With permission.)

Values of ψ greater than 1 are associated with predominately plastic contacts while values <0.6 are
considered to correspond to elastic deformation of asperities. This can serve as a general guide for how
the material will behave during the initial phases of a break-in cycle.
Khonsari and Akbarzadeh provide some examples of the surface roughness changes that occurred
during the break-in period for a roller-disk test [5,6]. The authors studied the effects of speed, slide-to-
roll ratio (SRR), and specimen hardness. Figure 19.4 shows their data comparing the roughness changes
that occurred on hard and soft disks. The authors found that the changes take place preferentially on
the softer disk when there is a hardness difference between the two test disks. The data also show that
at higher speeds the roughness change is reduced. This is due to a thicker lubricant film which reduces
the number and average severity of asperity contacts. The authors go on to model the break-in process
by assuming that the changes are due to plastic deformation of the asperity tips. This is a challenging
problem since it is necessary to capture the changes in topography and update the profiles at each load-
ing sequence. As break-in proceeds, the relative proportion of plastic asperity contacts reduces until a
steady state is reached. Comparisons between predicted and measured friction and wear over the dura-
tion of the break-in show remarkably good agreement and suggest that their model has captured the
main physical process, at least for the test system investigated.
The break-in process can be summed up as a complex and transient interaction between the surface
materials and the lubricant which is highly dependent on the operating conditions. However, a few
simple rules are common to many situations. The rate and amount of surface change increases with
thinner lubricant films (i.e., lower λ), higher pressures, and lower material hardness. Recommended
break-in procedures should be sought from the equipment manufacturer and typically consists of a
predetermined series of operating conditions designed to fully prepare the surfaces for regular service.

19.3  Experimental Methods and Tests


There are many configurations of friction and wear tests that are available. Most of these are capable of run-
ning tests in the mixed lubrication regime. In the case of reciprocating tests, the lubricant film varies as a
function of position along the stroke such that it is possible that all three lubrication conditions (i.e., bound-
ary, mixed, and full film) are encountered along the rubbing track. In unidirectional sliding or mixed rolling–
sliding tests, it is possible to control conditions within the equipment limits to obtain the desired conditions.
Mixed Lubrication 19-7

For tests that are not defined by standards, it is largely up to the researcher to identify the desirable
test conditions. This requires an estimate of both lubricant film thickness and surface roughness.
The measurement and characterization of surface finish is covered elsewhere but an important factor
is to measure and characterize the finish over the appropriate wavelength. The concept of functional
filtering as described in Thomas [7] can be used to obtain roughness parameters that are scaled to the
dimensions of the contact. This concept was further developed in Sayles et al. [8] in which the ability of
the surfaces to conform to a range of different wavelength features is described. Using this approach, the
calculation of surface roughness should be based on data that have been filtered at a level that is some
fraction of the overall dimensions of the contact. The surface roughness based on such filtered data will
be significantly less than that calculated using the unfiltered data and will yield a more representative
estimate of the contact conditions as defined in Equation 19.1.
The lubricant film thickness is a function of load, speed, temperature, materials, and geometry and
can be estimated using models and equations based on smooth surface conditions. There are numer-
ous models that can be used to estimate film thickness under EHL and hydrodynamic conditions, as
described in other chapters. However, the use of different viscosity lubricants enables the film thickness
to be varied independently of the operating variables listed earlier. Thus, the researcher is faced with
an almost limitless series of options for designing a series of tests and some care is required to ensure
that the required conditions are met by the correct selection of test materials and operating conditions.
Typical machine configurations for conducting mixed rolling–sliding tests are described in [9,10]. In
both cases, the contact is formed between a ball and a flat disk. Since the ball and disk are driven and
controlled independently, it is possible to run tests at different entrainment velocities (UE) and SRRs as
defined in Equations 19.3 and 19.4:

U1 − U 2
SRR = 2 × × 100% (19.3)
U1 + U 2

U1 + U 2
UE = (19.4)
2

where U1 and U2 are the velocities of surfaces 1 and 2.


Using such equipment, it is possible to replicate a wide variety of contact conditions. Table 19.3 shows
examples of varying the entrainment velocity while maintaining the sliding speed. To be noted is the
speed reversal required to maintain a high sliding speed at reduced entrainment velocity. This allows
tests to be conducted in the mixed lubrication regime under relatively severe sliding conditions using
conventional lubricants.
The ball-on-disk methods described earlier are examples of tests that use relatively standardized
test parts. However, in some cases it is important to more closely replicate the conditions present in

TABLE 19.3  List of Speed Conditions to Vary Entrainment Speed at Constant Sliding Speed
SRR (%) Disk Speed (m/s) Ball Speed (m/s) Entrainment Speed (m/s) Sliding Speed (m/s)
25 4.5 3.5 4 1
100 1.5 0.5 1 1
200 1 0 0.5 1
1,000 0.6 −0.4 0.1 1
2,000 0.55 −0.45 0.05 1
10,000 0.51 −0.49 0.01 1
Source: PCS Instruments, United Kingdom.
19-8 Theory and Practice of Lubrication and Tribology

a particular application. Thus, the materials and finish of the parts need to be controlled and selected
such that the asperity interactions that take place in the test are similar to those in the real case. One
approach is to use equipment components. A good example is the study of piston ring–cylinder liner
contacts in internal combustion engines. The reciprocating motion of the ring typically means that
the mode of lubrication will transition between boundary lubrication around the stationary points at
the end of the stroke and mixed film or even full film lubrication at the center stroke corresponding
to the point of highest velocity. In order to obtain the correct mix of conditions to study friction and
wear performance, it can be convenient to test ring and liner parts directly. Hartfield-Wunch et al. [11]
describe the development of a reciprocating test using engine piston ring and liner parts. A further level
of sophistication is to instrument the parts in a fully operating engine [12]. While this provides the most
accurate representation of lubricating conditions, it becomes more difficult to vary conditions, materi-
als, and surface finish in order to examine the sensitivity of results to key variables.
Rolling element bearings and gears are components that, from a practical perspective, often operate
under mixed film conditions. In these cases, the hydrodynamic component is provided by EHL mecha-
nisms and the conditions can vary from near pure rolling to having a large sliding component within
the contact. There are a number of standardized tests that have been developed using gears and bearing
components directly to evaluate performance in the mixed film region. Perhaps the most common of
these is the FZG scuffing test [13]. Scuffing is a severe form of surface adhesive wear. The test is most
often used to determine the scuffing resistance of gear oils. The test can be run under different operating
conditions, but typically the gears are run at increasing load increments for a fixed period at constant
speed for each condition. The test is stopped and the gear surfaces are visually examined for the appear-
ance of scuffing. If the amount of surface scuffing is below a given threshold, the test is continued at
a higher load and reexamined. The test yields a numerical ranking corresponding to the highest load
before scuffing appears over an equivalent of one gear tooth width. Automotive gear oils are tested in a
similar fashion in the L37 test [14]. In this case, a full rear axle unit using hypoid gears is tested under
high load/low speed conditions in order to promote mixed lubrication conditions under high sliding.
In both the FZG and L37 tests, the ability of the lubricant to form durable antiwear films to protect the
surface during asperity collisions within the contact determines the outcome of the test.
Another test that has become increasingly popular over the last 10–15 years is the FVA micropitting
test originally described by Schonnenbeck [15]. In this case the gear finish and conditions are selected
to promote micropitting, a microscopic fatigue process associated with asperity interactions. This mode
of failure has become prominent in wind turbine gearbox systems and was the subject of a workshop
held at the National Renewable Energy Laboratory in 2009 and summarized by Sheng [16]. The FVA
test is now one of the tests used to evaluate the ability of gear oils to mitigate the onset and severity of
micropitting in gears that will operate under mixed lubrication conditions. However, some researchers
prefer to use simpler bench-scale tests and roller-disk machines can offer advantages for the study of
fundamental mechanisms [17–19].
Bearing tests are also used to evaluate lubricant performance. The FE-8 test as described by
Reichelta et al. [20] can be used in a number of configurations. When used in the thrust bearing con-
figuration at low speed, the resulting low film lubricant film thickness promotes surface interactions,
making it possible to study the influence of lubricant additives on the formation of surface films,
friction, and wear.
By its nature, the mixed film regime requires an understanding of fluid behavior under very thin
film conditions where the proximity of the surface may result in behavior that is different from what
is expected based on bulk properties. This aspect has been nicely reviewed by Spikes and Olver [21].
Israelachvili pioneered the use of surface force apparatus (SFA) to study the effects of fluids confined
within the gap between two ultra smooth surfaces [22]. Using this technique, it has been shown that
the normal force variation as a function of separation exhibits a periodicity that corresponds with
the molecular dimension of the fluid used. This suggests that a degree of ordering is imposed by the
Mixed Lubrication 19-9

proximity of the surface. Later work imposed a lateral force on the system allowing the measurement of
effective viscosity in the near-surface region [23]. These experimental platforms require a great deal of
care to set up, align, and control, and their use has been confined to a relatively small number of special-
ized research teams.
A more conventional approach has been to extend the application of optical interferometry that
was originally used to study the formation of EHL films [24]. The use of the so-called spacer layer
deposited on top of the semi-reflecting layer used on the transparent disk to form the optical interfer-
ence fringes effectively extends the measurement of films down to a few nanometers [25]. This system
has been used to test the limits of the validity of typical EHL models down to very low separations.
One such model is given by Hamrock and Dowson [26] which for a constant load can be expressed in
the simplified form:

0. 7
h = k (Uh ) (19.5)

where
h is the elastohydrodynamic film thickness
U is the average entrainment speed
η is the lubricant viscosity
k is a constant that depends on material, geometric and load conditions

For simple fluids, Equation 19.5 appears to hold down to the limits of the measurement resolution (i.e.,
a few nanometers). However, Guangteng and Spikes [27,28] have found that mixtures of polar and non-
polar fluids can exhibit significant departures from the expected behavior at small surface separation as
shown in Figure 19.5. It is suggested that the higher attraction of the polar fluid to the surface leads to

100

10
Film thickness (nm)

ESTL
1
PAOL

ESTL + PAOL

0.1
0.0001 0.001 0.01 0.1 1
Rolling speed (m/s)

FIGURE 19.5  Comparison between the elastohydrodynamic response of poly-alpha-olefin (PAOL), an ester
(ESTL), and a 10%–90% mixture of the two (ESTL + PAOL). The data suggest that the concentration of the
higher viscosity ester is increased within the contact. (From Guangteng, G. and Spikes, H.A., Wear, 200, 336,
1996. With permission.)
19-10 Theory and Practice of Lubrication and Tribology

an increased concentration in the near-surface region. If the polar material has a higher viscosity than
the parent blend, it can increase the lubricating film formed within the contact. In later work, the same
technique has been used to examine the influence of a number of different classes of lubricant additives
on near-surface effects within EHL contacts [29–31].
The near-surface issues described earlier are principally due to localized rheological effects. More
recently, a companion technique has been used to examine the behavior of chemically reacted films
that can form on rubbing surfaces [32]. These films are often referred to as tribofilms and are typically
formed by the reaction of antiwear and friction modifier components used in formulated lubricants. As
discussed in the section covering break-in, it takes a finite time for these films to form. However, they
serve a vital role in determining the ultimate impact that asperity interactions have on both local and
overall friction and wear performance in the mixed film regime. Figure 19.6 shows the author’s own
results using the technique described in [32]. Under full film conditions where there is little or no inter-
action of surface asperities, there is no evidence of tribofilm formation. Under mixed film conditions,
a film is clearly observed to cover most of the rubbing track. If conditions are changed to increase the
severity of the contact, we find that the tribofilm is diminished or completely removed. In Figure 19.6,
this has resulted in wear as evidenced by the distorted image formed when the interferometer surface
makes contact with the worn surface.
The formation of tribofilms depends on the nature of the interaction of surface asperities within the
contact. Surface roughness can also have an effect on the behavior of the lubricant film. The early experi-
mental work on EHL was limited to smooth surfaces [24]. While the study of random rough surfaces
remains a difficult experimental option, researchers have imposed individual features on the surface
and observed their behavior within EHL contacts [33]. Over the years, these studies have increased in
complexity to a point where transient effects can be observed on individual features and the impact of
rough surfaces can be measured [34,35]. This work, along with the associated developments in computer
models, led to the concept of micro-EHL. The concept behind micro-EHL is that opposing asperities
form their own individual contacts within the lubricated contact. The nature of these contacts is gov-
erned by the relative motion of the individual asperities and the rheology of the fluid at the elevated

Tribofilm response to different


Ball on disk running to surface roughness under mixed
create wear track rolling–sliding conditions
Microscope and spacer Images Film maps
RGB color Smooth disk
camera layer-coated disk
Steel ball Film
Steel disk map scale

Lubricant 250
Increasing surface contact

Heaters 0.15 µm Ra disk 200

150
Measurement of
tribofilm 100

50

0.5 µm Ra disk 0

FIGURE 19.6  Results showing the effect of varying l (Equation 19.1) on the formation of tribofilms (also known

as antiwear films) under mixed sliding and rolling conditions.
Mixed Lubrication 19-11

pressures and shear rates that exist during these transient events. This topic has been the subject of
numerous numerical studies and will be covered later.
This section has introduced a few of the experimental techniques and concepts used to study
mixed lubrication and is by no means exhaustive. However, it should be clear that the types of experi-
ments span the range between standardized tests, which often form part of a lubricant or equipment
specification, to highly specialized and novel methods developed to explore specific mechanisms that
occur within the lubricated contact. As our ability to measure events over increasingly small times
and dimensions increases, we will no doubt learn more about the nature of the transient events that
have such a large influence on the friction and wear performance of systems operating in the mixed
film regime.

19.4  Wear and Durability under Mixed Lubrication Conditions


Asperity interactions are a key element of the mixed lubrication regime. As stated earlier, surface con-
tact supports part of the total load and an inevitable consequence is that there will be some form of per-
manent damage to the surface that results from the highly stressed asperity interactions. In addition, the
small separation between surfaces can increase the chances that third body particles entrained into
the contact will create more damage than under thicker full film conditions. It is thus no surprise that
the mixed lubrication regime is associated with a number of wear processes and it is important to under-
stand the relationship between the operational and material variables with these different mechanisms
of wear. It is a general rule that nearly all forms of wear will tend to increase as l is decreased. However,
some forms of wear are more susceptible than others. In this section, we will cover three of the most
common and researched areas: scuffing, fatigue, and micropitting.

19.4.1  Scuffing under Mixed Lubrication Conditions


Scuffing is defined as a form of adhesive wear that results in solid phase welding between the opposing
surfaces. Other terms in common use include scoring, galling, and smearing. It most commonly occurs
under sliding or mixed sliding and rolling conditions where there is at least the potential for significant
interaction between the opposing surfaces. Scuffing problems have been reported and studied for a wide
range of different components including gears, engine cams and followers, rolling element bearings,
and piston rings. Scuffing failures can occur for a wide range of materials ranging from the traditional
ferrous-based materials to nonferrous metals and even ceramics. The scuffing failure itself is assumed
to be a thermally activated mechanism associated with the surface flash temperature concept proposed
by Blok [36]. Local surface yielding resulting in generation of wear debris occurs when the contact
stresses are greater than the yield strength of one of the materials in the contact. The increase in surface
temperature due to local friction serves to lower the yield stress and consequently increases the risk of
local yielding. The wear debris can then adhere to the opposing surface. The transfer of material from
one counter face to the other is often used as a criteria for categorizing a scuffing failure. Scuffing has
been studied over many years and there is a large body of literature that covers different aspects of this
failure mode. Here, we shall only concentrate on those factors directly related to the mixed lubrication
condition.
It is generally accepted that under lubricated conditions the collapse of the oil film is a prerequisite
for the onset of scuffing [37–39]. If macroscopic levels of scuffing occur, large areas of the surface are
distressed and it can progress rapidly to a catastrophic failure. A more localized microscopic form of
scuffing has also been identified [40] which is associated with a breakdown in the micro-EHL films
generated at the asperity level. Lee and Cheng [40] correlated rough surface EHL analysis with the onset
of microscopic scuffing from disk experiments. A summary of their data is given in Figure 19.7 which
plots the results in terms of the film thickness and roughness-related parameters at various levels of
Dyson’s fluid film/surface load sharing parameter Q [37]. The results confirm that the load at which
19-12 Theory and Practice of Lubrication and Tribology

10–3

S/R = 0.5396
S/R = 0.7826
S/R = 1.0769

E΄R
W
Load parameter,

Q = 0.8
Q = 0.7
Q = 0.6
Q = 0.5
Q = 0.4
Q = 0.3
Q = 0.2

10–4
0.1 1.0
2 η U
Hydrodynamic parameter, 48R 0 r
σ 2 E΄R

FIGURE 19.7  Predicted scuffing performance at various fluid film/asperity load sharing conditions (Q). (From
Lee, S.C. and Cheng, H.S., ASME J. Tribol., 113, 319, 1991. With permission.)

scuffing occurs is reduced for thinner lubricant films (i.e., lower Q). Zhang et al. [41] also found that
increasing roughness increased the maximum surface traction and asperity flash temperature, which
resulted in increasing their scuffing uncertainty factor in piston ring–cylinder liner contacts. Qua et al.
[42] presents transition diagrams showing the onset of scuffing as a function of surface roughness for
both steel–steel and steel–zirconia contact pairs.
Most lubricants contain surface-active antiwear and so-called extreme pressure (EP) additives. The
latter are specifically designed to help reduce the risk of scuffing and might be better described as anti-
scuffing additives. Such surface-active components form reacted tribofilms that can prevent direct
surface contact in the event that the lubricant film collapses. The presence and effectiveness of these
components has a significant impact scuffing performance. The FZG and L57 gears tests [13,14] are both
designed to evaluate the lubricant’s ability to protect against scuffing failure. Soejima et al. [43] pres-
ent results from cam-follower bench tests showing that both friction modifiers and zinc dialkyldithio-
phosphate (ZnDTP), a common antiwear component, reduce the risk of scuffing under any given load
condition. Their data also highlighted the fact that reducing the lubricant viscosity increases the risk
of scuffing. Jackson et al. [44] present data for both smooth and rough surface contacts for a variety of
different lubricants. They compared experimental results with both “flash temperature” and “frictional
power intensity” models and concluded that the presence of surface-active additives increases the effec-
tive critical temperature that must be reached prior to scuffing failure. Furthermore, their data also
support the concept that both EHL traction and friction due to asperity interaction affect the surface
temperature and hence the load at which scuffing will occur.
From a practical standpoint, it is difficult to establish an accurate prediction of the relative risk of
scuffing for a given piece of equipment. This is due to the complex interaction of multiple materials,
surface, operational and lubricant effects, the details of which are not always available to equipment
operators. However, research shows that scuffing should be considered as a possible failure mechanism
for contacts that operate in the mixed lubrication regime and exhibit moderate to high levels of sliding.
Under these conditions, the risk of scuffing can be reduced by increasing the lubricant film thickness,
decreasing the surface roughness, changing the material, or selecting a lubricant formulated to protect
heavily loaded systems against scuffing failure.
Mixed Lubrication 19-13

19.4.2  Rolling Contact Fatigue under Mixed Lubrication Conditions


The surface stresses associated with nonconformal contacts are very high and can easily exceed
1.0 GPa. Consequently, the subsurface volume of material in the immediate vicinity of the contact is
also subjected to a distribution of high stresses. In systems where repeated contact occurs, such as in
gears, cam/cam followers, and rolling element bearings, a given element of material is thus exposed to
cyclic stresses each time the contact passes over the surface. Over time, this can ultimately lead to the
initiation and propagation of fatigue cracks in a manner analogous to structural fatigue. Eventually,
the cracks will grow or coalesce to a point where a piece of material is ejected from the surface leaving
behind a pit. This process is known as rolling contact fatigue which sometime goes by the alternate
names of pitting or spalling.
Following the early work of Lundberg and Palmgren [45], the growth of rolling contact fatigue
theory and application has been closely associated with rolling element bearing and gear technology
development. It is beyond the scope of this section to comprehensively review rolling contact fatigue
phenomena. However, a widely accepted view is that fatigue cracks can be subsurface- or surface-
initiated. Subsurface initiation is associated with macroscopic subsurface stresses that can interact
with material inclusions. The surface-initiated cracks are generally assumed to result from stress con-
centrations that occur at local surface defects or roughness high spots. In the latter case, the lubrication
conditions can have a significant influence on the fatigue process and ultimately the effective life of the
component.
The calculation of rolling element bearing life is governed by a number of national and interna-
tional standards as well as the proprietary techniques developed by individual companies. In 1971, the
ASME published a design guide [46] to help establish a common approach to estimate bearing life. The
approach taken was to develop adjustment factors to account for the influence of different variables.
One of these is based on l defined in Equation 19.1. The adjustment factor exhibits a highly nonlinear
response to l and can result  in a 10-fold increase when l varies from 0.25 to 3 (i.e., spanning the range

which includes the mixed lubrication regime). In some cases,  l is replaced by the viscosity ratio κ which

is simply the ratio of the lubricant viscosity at the operating condition to the viscosity required to obtain
a reference condition (e.g., a certain EHL film thickness). The advantage of this approach is that it avoids
the need to measure and define the surface finish. The disadvantage is that the relationship holds for
standard mineral oils but is not valid for many synthetic oil-based lubricants due to differences in the
pressure–viscosity coefficient for these two classes of fluids.
The methods outlined in ISO 281 [47] has gained widespread acceptance as a updated approach to
the calculation of fatigue life. In this case, a single life modification factor aISO is defined as given in
Equation 19.6:

⎛e C ⎞
aISO = f ⎜ C u k ⎟ (19.6)
⎝ P ⎠

where
eC is the lubricant contamination factor
Cu is the fatigue load limit
P is the equivalent dynamic load
v
k = , where ν1 is the lubricant viscosity required to obtain a reference condition and ν is the
v1
operating viscosity

The factors used to estimate aISO are interdependent, such that their values can vary significantly based
on the value of l or κ, for example. This approach was taken a step further in the updated version of
the ASME bearing life design guide [48] where the λ (or κ)-related factors are replaced by an estimate of
19-14 Theory and Practice of Lubrication and Tribology

10
Symbol legend
50 Thru hardened gearing No distress
petroleum lube Distressed
Carb. pinions-nitrided gears No distress 5
20 polyolester lube Distressed 10
20
40
Specific film thickness, λ

1.0 80

0.5
% Probability
of distress
0.2

0.1

.05

.02

.01
1.0 10.0 100 1000 10,000 100,000
Pitch line velocity, FPM

FIGURE 19.8  Lines of constant probability of surface distress for gear surfaces plotted as a function-specific film
thickness (λ) and pitch line velocity. (From Wellauer, E.J. and Holloway, G.A., Trans. ASME J. Eng. Ind., 626, 1976.
With permission.)

the surface stresses used as input to a stress-life model. Also introduced are some rough guidelines to
account for the impact of various lubricant additive types.
There remains considerable debate over the relative merits of independent versus interdependent life
adjustment factors. However, the impact of the lubrication condition is a common factor and remains a
key variable in determining the ultimate fatigue life of bearings and has been the subject of wide rang-
ing research. The effect of lubricating condition on gear fatigue is also well established, although the
more complex geometry and presence of high sliding introduce key differences versus rolling element
bearings. In a key work, Wellauer and Holloway [49] applied simple EHL theory to the failure analysis
of gear performance. Their data, represented in Figure 19.8, show the combined influence of speed, as
defined by the pitch line velocity, and specific film thickness. Similar to the case with bearings, it is clear
that the response is highly nonlinear and there is a large variation within the mixed lubrication regime.
The effect of lubricant composition has also been an active area of study. Wang et al. [50] used a
four-ball tester to examine the fatigue life response of a variety of mineral and synthetic lubricants.
They confirmed that increased viscosity generally improved fatigue life but that there was consider-
able variation between the lubricant types. The viscosity effect can be attributed to increase surface
separation but the variation between lubricant types may be due to physical effects such as pressure–
viscosity and EHL traction or differences in surface adsorption characteristics. The study included
a comparison between an oil containing an EP additive with the equivalent nonadditized oil. The
EP-containing oil was found to improve fatigue life. In contrast, Torrance et al. [51] found that the EP
additive used in their tests resulted in a significant reduction in fatigue life. In both cases, the testing
was conducted at relatively high contact pressures (4.8–7.5 GPa) which is well above the practical limit
for traditional gear and bearing systems. In bearing tests, Wan et al. [52] found that an EP additive
reduced fatigue life under λ = 1.2 operating conditions. The effect of lubricant composition remains
a poorly understood phenomenon and accounts for the large range in the potential life adjustment
factors outlined in [48].
Mixed Lubrication 19-15

Given the cost of producing highly finished surfaces, it remains likely that for many practical appli-
cations bearings and gears will continue to operate under mixed lubrication conditions. The study of
surface-related fatigue phenomena remains an active area of research.

19.4.3  Micropitting
Another form of fatigue-related wear is micropitting. This mode of wear is sometimes referred to as gray
staining or frosting due to the matt surface appearance that it produces. Originally it was assumed that
this was the result of surface corrosion, but it is now widely accepted that it is a fatigue process that cre-
ates a multitude of microscopic pits on the surface. The matt appearance occurs once these pits coalesce
to form an almost continuous fractured surface that causes a diffuse reflection of light from the surface.
Micropitting has been recognized as a distinct form of wear for many years [53]. Research activity has
increased over the last 15–20 years due in part to problems that have arisen with the development of
gears for modern wind turbine systems. Today, micropitting has emerged as one of the key performance
requirements for advanced gear oils targeted for use in wind turbine gearboxes [54].
Micropitting is most often associated with mixed film lubrication conditions in which sliding is
present. As distinct from the scuffing failure mode, it appears in the form of progressive wear that
can spread relatively gradually across the surface. It is generally accepted that the fatigue cracks are
initiated and propagated under the influence of local asperity pressures associated with their inter-
action within the EHL contact. This gives rise to a near-surface localized stress distribution result-
ing from both elastic and plastic deformation of the asperities [55]. In practice, micropitting is most
often observed in case hardened steel gears and some types of rolling element bearings. Webster and
Norbart [17] demonstrated that it is possible to reproduce the failure using a relatively simple roller-
disk machine. They found that under mixed film conditions, micropitting is promoted by increasing
the amount of relative sliding above a given limit. Furthermore, their results showed that micropitting
preferentially occurs on the slower and softer of the two surfaces. It is presumed that the softer sur-
face undergoes a higher degree of plastic deformation. They also reported that polishing the surface
eliminated micropitting completely which confirms the role that asperity interaction plays in initiat-
ing micropitting.
In addition to the physical variables, there is a considerable body of evidence that surface-active
components, such as antiwear and EP additives, can also have an influence on micropitting [17,56–59].
While the effect varies considerably within a given class of additive, it appears that their presence is most
often detrimental when compared with results from test run on lubricants not containing such materi-
als. Unfortunately, gear oils especially are required to use antiwear or EP additives in order to provide
protection against other forms of wear such as scuffing. In order to design a practical gear oil, it is neces-
sary to achieve a good balance between these competing requirements [60]. The mechanism responsible
for the effect of additives remains an area of active research. However, Olver et al. [59] suggest that the
presence of antiwear additives protects the asperities from localized wear. Subsequently, the localized
stresses continue to occur due to asperity interactions. Conversely if the additives are not present, the
resulting removal of surface asperities eliminates the mechanism for generating the surface stresses
responsible for initiated the fatigue cracks. In related work, Lainé et al. [19] demonstrated that the use of
friction modifiers is effective in reducing micropitting even in the presence of antiwear additives. They
attribute this to a reduction in the local surface traction forces that occurs during asperity collisions
within sliding contacts. This modifies the stress distribution in the surrounding area to the extent that
fatigue cracks are no longer initiated and propagated. If this is the mechanism at work, it means that the
effect of additives on micropitting is due their influence on the running-in process rather than direct
action on crack initiation sites.
There are a number of practical steps that can be taken to reduce the potential for micropitting-
related failures. Many of these were outlined in the wind turbine workshop [16] and include the obvi-
ous steps of reducing roughness and increasing the lubricant film. Furthermore, many of the gear and
19-16 Theory and Practice of Lubrication and Tribology

bearing manufacturers have developed their own recommendations which should be referenced if it is
believed that a particular system is likely to be prone to micropitting problems.

19.5  Modeling Mixed Lubrication Problems


Models used to describe the lubrication process were initially based on the application of the Reynolds
equation with the assumption that the surfaces are smooth. Subsequently, analytical methods have been
developed that combine lubrication models with statistically based rough surface contact principles that
allow us to calculate the relative load sharing between the lubricant film and the surface asperities [4,62].
This approach is invaluable in terms of defining the overall conditions within a contact and estimat-
ing the relative load sharing between the hydrodynamic pressures developed in the lubricant film and
the pressures developed resulting from direct surface contact. However, such approaches yield average
properties and are based on a range of simplifying assumptions related to the nature of the topography
involved.
The application of numerical techniques allows some of these restrictions to be relaxed. The develop-
ment of numerical models has been broadly divided into two approaches: statistical and deterministic.
The first approach typically combines a statistical description of the roughness and a modified version
of the Reynolds equation accounting for the flow deviations due to roughness effects. Patir and Cheng
[63] give one of the best early examples of this approach, where they were able to generate results for
asperities arranged with different orientations relative to the entrained flow of lubricant. This approach
yielded many important results related to the generation of lubricant films and flow around asperities
within the contact and the potential for local stresses much higher than those predicted using smooth
surface assumptions.
The statistical approach benefits from relative numerical simplicity and was well suited for the limited
computational power available at the time of its development. However, with increasingly more pow-
erful computers and numerical solution methodologies, deterministic models are now the dominant
approach. These models use idealized geometric representations of the surface and were initially limited
to pure sliding conditions. The simplest cases are typically single features placed on the stationary sur-
face. This yields a steady-state solution for a given position of roughness within the contact. Goglia et al.
[64] contains results for a single furrow predicting high pressures around the shoulder of the irregular-
ity. In a companion work, Goglia et al. [65] present similar results for a model sinusoidal surface.
The concept of micro-EHL emerged around this time period and was driven in large part by results
from these models and the associated developments in the experimental arena [66]. It is important to
recognize that there is a clear distinction between micro-EHL and events leading to surface contact,
both of which may occur under mixed lubrication conditions. Micro-EHL still maintains a continuous
lubricant film that separates the surfaces. A challenging aspect of this is that the rheology of the lubri-
cant at the very localized high pressures and shear rates determines the nature of the stresses transmit-
ted from the lubricant into the surface. Even today, it is beyond our ability to measure fluid properties
under these extreme conditions and appropriate models must be selected to adequately describe the
fluid rheology. Once surface contact occurs, the shear stresses are determined by the friction at the
interface.
With increasing computational power, it has been possible to develop evermore sophisticated mod-
els. The application of multigrid methods by Lubrecht et al. enabled the more efficient solution of 3-D
problems associated with point contacts [67]. This approach was subsequently used to study roughness
effects within point contacts [68]. The ability to model finer features eventually lead to the possibility of
incorporating topography measured directly from a real rough surface, first for dry contacts [69] and
then for lubricated contacts [70]. Further advances allowed the use of non-Newtonian fluid models and
the inclusion of thermal effects. The latter added the requirement to simultaneously solve the energy
equation along with the Reynolds equation and surface displacements.
Mixed Lubrication 19-17

The development of numerical models has enabled us to study in detail how rough surfaces interact
within a lubricated contact. However, there is much more to be done—specifically to extend such mod-
els to allow prediction of the effect of asperity interactions on friction and wear, which are of utmost
practical interest. Such an objective presents some interesting challenges, such as how to account for
the possibility that some form of surface modification takes place while also incorporating a fatigue or
wear model component. This requires a dynamic modeling approach that is capable of predicting local
changes, such as material removal due to wear, and subsequently updating the surface profile before
proceeding to the next time step. Despite these hurdles, the most recent models have stepped in this
direction and have demonstrated the ability to account for competition between different modes of wear
and reproduce the progressive evolution of wear across a rough surface [71].

19.6  Mixed Lubrication: The Future


Our ability to study the complex interaction between the lubricating oil and rough surfaces has pro-
gressed tremendously over the last few decades. Every year yields models of increasing complexity. The
problem has been approached using continuum mechanics principles and more recently via molecular
dynamics simulations. It appears likely that these concepts will ultimately be combined so that the fric-
tion and wear processes that occur under mixed lubrication can be understood across all length scales.
Similarly our ability to conduct experiments has evolved as new techniques emerge. While in situ
measurements within real contacts remain a challenge, the development of so-called nanotools such as
AFM offers the potential to study effects at a scale relevant to the lubrication of rough surfaces. Similarly,
our ability to remove or deposit features of a specific geometry allows us to design experiments to test
new models and concepts.
For the most part operating in the mixed lubrication, regime is viewed as a necessary evil. The com-
bination of the need to operate over a wide range of conditions and the fact that it remains difficult and
costly to produce very smooth surfaces means that it is impossible to avoid this regime. However, it may
turn out that operating with thin lubricant films will become a desired and sought after condition. The
development of surfaces with precisely engineered structures now provides an opportunity to design
systems that take advantage of operating under very thin film conditions while providing mechanisms
to minimize surface contact and thus maintain control of friction and wear. The scientific principles at
work are likely to remain the same but perhaps instead we will be seeking to maximize performance
rather than avoid the consequences of working in this complex regime.

References
1. ASTM D7589-11, Standard test method for measurement of effects of automotive engine oils on fuel
economy of passenger cars and light-duty trucks in sequence VID spark ignition engine.
2. ASTM D7422, Standard test method for evaluation of diesel engine oils in T-12 exhaust gas recircu-
lation diesel engine.
3. Katoh, J., Satoh, T., Kamikubo, F., and Mizuhara, K., Analysis of running-in process under lubri-
cated conditions using combined time-space plot and three-dimensional bearing curves, Tribology
Transactions, 44(1), 2001, 104–110.
4. Greenwood, J. A. and Williamson, J. B. P., Contact of nominally flat surfaces, Proceedings of Royal
Society of London, A295, 1966, 300–319.
5. Akbarzadeh, S. and Khonsari, M. M., On the prediction of running-in behavior in mixed-lubrication
line contact, ASME Journal of Tribology, 132, 2010, 032102.
6. Khonsari, M. M. and Akbarzadeh, S., Experimental and theoretical investigation of running-in,
Tribology International, 44, 2011, 92–100.
7. Thomas, T. R., Rough Surfaces, Chapter 5, Section 5.4.1, Longman, London, U.K., 1982.
19-18 Theory and Practice of Lubrication and Tribology

8. Sayles, R. S., DeSilva, G. M. S., and Leather, J. A., Elastic conformity in Hertzian contacts, Tribology
International, 14, 1981, 315–322.
9. Wedeven, L. D., Method and apparatus for comprehensive evaluation of tribological materials, US
Patent 5679883, 1997.
10. Smeeth, M., Hamer, C., and Spikes, H., A study of antiwear additive film build up using the MTM
(mini-traction machine), Proceedings of the ASME/STLE International Joint Tribology Conference,
San Diego, CA, PART A, 2007, pp. 101–103.
11. Hartfield-Wunsch, S. E., Tung, S. C., and Rivald, C. J., Development of a bench test for the evaluation
of engine cylinder components and correlation, SAE Paper 932693.
12. Dearlove, J. and Cheng, W. K., Simultaneous piston ring friction and oil film thickness measurements
in a reciprocating test rig, SAE Paper 952470.
13. DIN 51354, Standard test method for load-carrying capacity of petroleum oil and synthetic fluid
gear lubricants.
14. ASTM D2161, Standard test method for evaluation of load-carrying capacity of lubricants under
conditions of low speed and high torque used for final hypoid drive axles.
15. Schonnenbeck, G., Testverfahren zur Untersuchung des Schmierstoffeinflusses auf die Entstehung
von Grauflecken, Antriebstechnik, 24(1), 1985, 31–37.
16. Sheng, S., Wind turbine micropitting workshop: A recap, Technical Report NREL/TP-500-46572,
February 2010.
17. Webster, M. N. and Norbart, C. J., An experimental investigation of micropitting using a roller disc
machine, Tribology Transactions, 38, 1995, 883–893.
18. Cardis, A. B. and Webster, M. N., Gear oil micropitting evaluation, American Gear Manufacturers
Association (AGMA), Fall Technical Meeting, Paper 99FTM4, 1999.
19. Lainé, E., Olver, A. V., Lekstrom, M. F., Shollock, B. A., Beveridge, T. A., and Hua, D. Y., The effect of a
friction modifier additive on micropitting, Tribology Transactions, 52(4), 2009, 526–533.
20. Reichelta, M., Gunst, U., Wolf, T., Mayer, J., Arlinghaus, H. F., and Gold. P. W., Nanoindentation, TEM
and ToF-SIMS studies of the tribological layer system of cylindrical roller thrust bearings lubricated
with different oil additive formulations, Wear, 268, 2010, 1205–1213.
21. Spikes, H. A. and Olver, A. V., Mixed lubrication—Experiment and theory, Boundary and Mixed
Lubrication: Science and Applications.
22. Israelachvili, J., Intermolecular and Surface Forces, 2nd edn., Academic Press, London, U.K., 1992.
23. Granick, S., Motions and relaxations of confined liquids, Science, 253, 1991, 1374–1379.
24. Snidle, R. W. and Archard, J. F., Experimental investigation of elastohydrodynamic lubrication at
point contacts, Proceedings of 1972 Symposium Elastohydrodynamic Lubrication, Paper C2/72,
Institute of Mechanical Engineering, London, U.K., 1972.
25. Johnston, G. J., Wayte, R., and Spikes, H. A., The measurement and study of very thin lubricant films
in concentrated contacts, Tribology Transactions, 34, 1991, 187–194.
26. Hamrock, B. T. and Dowson, D., Ball Bearing Lubrication: The Elastohydrodynamics of Elliptical
Contacts, John & Wiley, New York, 1981.
27. Guangteng, G. and Spikes, H. A., Fractionation of liquid lubricants at solid surfaces, Wear, 200, 1996,
336–345.
28. Guangteng, G. and Spikes, H. A., The control of friction by molecular fractionation of base fluid
mixtures at metal surfaces, Tribology Transactions, 40(3), 1997, 461–469.
29. Guangteng, G., Smeeth, M., Cann, P. M., and Spikes, H. A., Measurement and modelling of bound-
ary film properties of polymeric additives, Proceedings of Instrumentation Mechanical Engineering
Journal, 210, 1996, 1–15.
30. Smeeth, M., Gunsel, S., and Spikes, H. A., Boundary film formation by viscosity index improvers,
Tribology Transactions, 39, 1996, 726–734.
31. Anghel, V., Bovington, C., and Spikes, H. A., Thick film formation by friction modifier additives,
Lubrication Science, 11, 1996, 313–335.
Mixed Lubrication 19-19

32. Fujita, H., Glovnea, R. P., and Spikes, H. A., The study of zinc dialkyldithiophosphate anti-wear film
formation and removal process. Part I: Experimental, Tribology Transactions, 48, 2005, 558–566.
33. Wedeven, L. D. and Casano, C., Elastohydrodynamic film thickness measurements of artificially
produced surface dents and grooves, ASLE Conference, Minneapolis, MN, October 24–26, 1978.
34. Krupka, I., Hartl, M., Zimmerman, M., Houska, P., and Jang, S., Effect of surface texturing on elasto-
hydrodynamically lubricated contact under transient speed conditions, Tribology International, 44,
2011, 1144–1150.
35. Krupka, I., Svoboda, P., and Hartl, M., Effect of surface topography on mixed lubrication film forma-
tion during start up under rolling/sliding conditions, Tribology International, 43, 2010, 1035–1042.
36. Blok, H., Theoretical study of temperature rise at surfaces of actual contact under oiliness lubricating
conditions, Proceedings of the General Discussion on Lubrication and Lubricants, London, U.K., The
Institute of Mechanical Engineers, October 13–15, 1937, pp. 222–235.
37. Dyson, A., The failure of elastohydrodynamic lubrication of circumferentially ground discs,
Proceedings of Instrumentation Mechanical Engineers, 190(1), 1976, 52–76.
38. Castro, J. and Seabra, J., Scuffing and lubricant film breakdown in FZG gears. Part I. Analytical and
experimental approach, Wear, 215, 1998, 104–113.
39. Castro, J. and Seabra, J., Scuffing and lubricant film breakdown in FZG gears. Part II. New PV scuff-
ing criteria, lubricant and temperature dependent, Wear, 215, 1998, 114–123.
40. Lee, S. C. and Cheng, H. S., Correlation of scuffing experiments with EHL analysis of rough surfaces,
ASME Journal of Tribology, 113, 1991, 319–326.
41. Zhang, C., Cheng, H. S., and Wang, Q. J., Scuffing behavior of piston-pin/bore bearing in mixed
lubrication—Part II: Scuffing mechanism and failure criterion, Tribology Transactions, 47(1), 2004,
149–156.
42. Qua, J., Truhan, J. J., Blau, P. J., and Meyer, H. M., III, Scuffing transition diagrams for heavy duty
diesel fuel injector materials in ultra low-sulfur fuel-lubricated environment, Wear, 259, 2005,
1031–1040.
43. Soejima, M., Ejima, Y., Wakuri., Y., and Kitahara, T., Improvement of lubrication for cam and fol-
lower, Tribology Transactions, 42(4), 1999, 755–762.
44. Jackson, A., Webster, M. N., and Enthoven, J. C., The effect of lubricant traction on scuffing, Tribology
Transactions, 37(2), 1994, 387–395.
45. Lundberg, G. and Palmgren, A., Dynamic capacity of rolling bearings, Acta Polytechnic, Mechanical
Engineering Series 1, RSAEE, No. 3, 7, 1947.
46. Bamberger, E. N., Harris, T. A., Kacmarsky, W. M., Moyer, C. A., Parker, R. J., Sherlock, J. J., and
Zaretsky, E. V., Life Adjustment Factors for Ball and Roller Bearings—An Engineering Design Guide,
ASME, New York, 1971.
47. ISO 281, Rolling bearings—Dynamic load ratings and rating life.
48. Life ratings for modern rolling bearings/a design guide for application of International Standard
ISO 281/2, ASME, 2003.
49. Wellauer, E. J. and Holloway, G. A., Application of EHD oil film theory to industrial gear drives,
Transactions of ASME, Journal of Engineering for Industry, May 1976, 626–631.
50. Wang, Y., Fernandez, J. E., and Cuervo, D. G., Rolling-contact fatigue lives of steel AIS152100 balls
with eight mineral and synthetic lubricants, Wear, 196, 1996, 110–119.
51. Torrance, A. A., Morgan, J. E., and Wan, G. T. Y., An additive’s influence on the pitting and wear of ball
bearing steel, Wear, 192, 1996, 66–73.
52. Wan, G. T. Y., Amerongen, E. V., and Lankamp, H., Effect of extreme pressure additives on fatigue life
of roller bearings, Journal of Physics D Applied Physics, 25, 1992, 147–153.
53. Shotter, B. A., Micropitting: Its characteristics and implications on the gear test requirements of gear
oils, Proceedings of Instrumentation Petroleum, Vol. 1, London, U.K., 1981, pp. 91–103.
54. Brechot, P., Cardis, A. B., Murphy, W. R., and Theissen, J., Micropitting resistant industrial gear oils
with balanced performance, Industrial Lubrication and Tribology, 52(3), 2000, 125–136.
19-20 Theory and Practice of Lubrication and Tribology

55. Olver, A. V., Micropitting and asperity deformation, Numerical and Experimental Methods Applied
to Tribology, Dowson, D., Taylor, C. M., Godet, M., and Berthe, D., eds., Proceedings of the 10th
Leeds-Lyon Symposium on Tribology (Lyon, France, September 1983), Butterworth & Co. Ltd.,
London, U.K., 1984, pp. 319–323.
56. Cardis, A. B. and Webster, M. N., Gear oil micropitting evaluation, American Gear Manufacturers
Association (AGMA), Fall Technical Meeting, Paper 99FTM4, 1999.
57. Benyajati, C. and Olver, A. V., The effect of a ZnDTP anti-wear additive on micropitting resistance
of carburized steel rollers, American Gear Manufacturers Association, Alexandria, VA, AGMA
Technical Paper, 04FTM06, 2004, pp. 1–10.
58. O’Connor, B., The influence of oil additive chemistry on micropitting, Gear Technology, 5–6, 2005,
34–41.
59. Olver, A. V., Dini, D., Lainé, E., Beveridge, T. A., and Hua, D. Y., Roughness and lubricant chemistry
effects in micropitting, AGMA Fall Technical Meeting, Detroit, MI, October 2007, American Gear
Manufacturers Association, Alexandria, VA, Paper No. 07FTM14, 2007.
60. Cardis, A. B., Formulating modern industrial gear oils: New requirements for global markets,
Conference Proceedings, Lubrication Excellence 2007, Noria Corporation, Tulsa, OK.
61. Lainé, E., Olver, A. V., Lekstrom, M. F., Shollock, B. A., Beveridge, T. A., and Hua, D. Y., The effect of a
friction modifier additive on micropitting, Tribology Transactions, 52(4), 2009, 526–533.
62. Johnson, K. L., Greenwood, J. A., and Poon, S. Y., A simple theory of asperity contact in elastohydro-
dynamic lubrication, Wear, 19, 1992, 91–108.
63. Patir, N. and Cheng, H. S., Application of average flow models to lubrication between rough surfaces,
Transactions of ASME, Journal of Lubrication Technology, 110, 1978, 12–17.
64. Goglia, P. R., Conry, T. F., and Cusano, C., The effects of surface irregularities on the elastohydrody-
namic lubrication of sliding contacts. Part I—Single irregularities, ASME Journal of Tribology, 106,
1984, 104–112.
65. Goglia, P. R., Conry, T. F., and Cusano, C., The effects of surface irregularities on the elastohydrody-
namic lubrication of sliding contacts. Part II—Wavy surfaces, ASME Journal of Tribology, 106, 1984,
113–119.
66. Kaneta, M., Micro-elastohydrodynamic lubrication, Japanese Journal of Tribology, 35(1), 1990,
11–22.
67. Lubrecht, A. A., ten Napel, W. E., and Bosma, R., Multigrid, an alternative method of solution for
two-dimensional elastohydrodynamically lubricated point contacts calculations, ASME Journal of
Tribology, 109, 1987, 437–443.
68. Lubrecht, A. A., ten Napel, W. E., and Bosma, R., The influence of longitudinal and transverse rough-
ness on the elastohydrodynamic lubrication of circular contacts, ASME Journal of Tribology, 110,
1988, 421–426.
69. Webster, M. N. and Sayles, R. S., A numerical model for the elastic frictionless contact of real rough
surfaces, ASME Journal of Tribology, 108, 1986, 314–320.
70. Kweh, C. C., Patching, M. J., Evans, H. P., and Snidle, R. W., Simulation of elastohydrodynamic con-
tacts between rough surfaces, ASME Journal of Tribology.
71. Morales-Espejel, G. E. and Brizmer, V., Micropitting modelling in rolling–sliding contacts:
Application to rolling bearings, Tribology Transactions, 54, 2011, 625–643.
20
Boundary Lubrication
and Boundary
Lubricating Films
20.1 Origins of Boundary Lubrication.................................................20-1
20.2 Stribeck Curves and Boundary Lubrication................................20-2
20.3 Solid-Like and Viscous-Like Boundary Films............................20-3
Solid-Like Boundary Films  •  Viscous-Like Boundary Films
20.4 Detection and Measurement of Boundary Lubrication............20-5
Stribeck Curve Measurements  •  Thin Film Interferometry  •  Atomic
Force Microscopy  •  Chemical Analysis of Boundary Lubricating
Films
20.5 Film Replenishment........................................................................20-6
20.6 Types of Boundary Film.................................................................20-7
Oxide Films  •  Base Fluid Boundary Films  •  Adsorbing Polymer
Boundary Films  •  Organic Friction Modifier Films  •  Deposited
Solid Boundary Films  •  Reacted Solid Boundary Films
20.7 Friction Control by Boundary Films.......................................... 20-11
20.8 Wear Control by Boundary Films...............................................20-12
20.9 Correlation between Friction and Wear Properties of
Boundary Films.............................................................................20-13
20.10 Strength of Boundary Films.........................................................20-13
Hugh Spikes 20.11 Prediction of Boundary Lubrication Performance.................. 20-14
Imperial College London References...................................................................................................20-15

20.1  Origins of Boundary Lubrication


The hydrodynamic lubrication theory, developed in 1886 by Reynolds, suggested that the ability of a
lubricant to form a separating film within a sliding contact was controlled by a single physical property,
its dynamic viscosity. However, it was very quickly noticed that machines operated with most “natural
oils,” i.e., lubricants based on oils originating from plants and animals, gave lower friction and wear
than with corresponding mineral oils of the same viscosity. This led to the recognition in the 1890s of a
mysterious property of liquid lubricants called “oiliness.”
In 1919, it was found that the addition of oleic acid to mineral oils improved the latters’ friction and
wear performance to equal that of natural oils [1]. This sparked considerable research and in the early
1920s, Langmuir, Hardy, and others showed that the property of oiliness could be produced by mono-
layers of long-chain surfactants adsorbed or reacted on solid surfaces. In 1922, Hardy and Doubleday
coined the word “boundary lubrication” to describe this effect [2]. Nowadays, the superiority of natural

20-1
20-2 Theory and Practice of Lubrication and Tribology

oils in terms of boundary lubrication is considered to be due to their partial decomposition to form
long-chain surfactants, either carboxylic acids or partial esters, that act as “natural” friction modifiers.
In the 1930s and 1940s, other types of lubricant additives were developed to control wear and seizure
in engines and transmissions and the term boundary lubrication was extended to include the protec-
tion provided by these additives. More recently, it has been shown that the relatively weak adsorption
of functionalized polymers and other polar species in lubricants can reduce friction and/or wear and
this has broadened still further the concept of boundary lubrication. This breadth makes it difficult to
produce an encompassing and also useful definition of boundary lubrication. One suggestion is

A condition of lubrication in which friction, wear or seizure behavior is controlled by a chemical


or physical interaction between the rubbing surfaces and components of a liquid or vapor lubri-
cant and which results in a surface film with physical properties different from both the rubbing
surfaces and the bulk lubricant.

20.2  Stribeck Curves and Boundary Lubrication


Boundary lubrication can be most clearly identified by charting the way that the friction of a lubri-
cated contact varies with rubbing speed. In both the hydrodynamic and elastohydrodynamic regimes,
lubricant film thickness increases with the entrainment speed. Since the entrainment speed is the
mean speed of the two surfaces with respect to the contact, in a sliding contact the entrainment speed
is generally half the sliding speed. This means that a graph showing how friction varies with either
entrainment speed or sliding speed for a lubricated contact will actually show how friction varies as a
separating hydrodynamic or elastohydrodynamic film is developed. This type of graph has been vari-
ously described but is now most popularly, albeit loosely, referred to as a “Stribeck curve.” A typical
example for a lubricated, conformal contact is illustrated in Figure 20.1a and b.
As shown in Figure 20.1, at low speeds the friction is relatively high. This corresponds to the bound-
ary lubrication regime. In this, a negligible fluid film is present so the load is supported by any boundary
lubricating film present on contacting asperities. At high speeds, a full separating fluid film is developed
and friction arises entirely from the force needed to shear this fluid film. At intermediate speeds, in the
“mixed lubrication” region, the load is supported by a combination of asperity contacts and a partial
fluid film. Figure 20.1b is shown with a logarithmic x-axis since this has the effect of “opening out” the
boundary lubrication region compared to the linear axis representation shown in Figure 20.1a. In the
figure, friction within the boundary lubrication regime is shown to be constant with speed. As discussed
in Section 20.7, this is not always the case.
Since the underlying aim of these Stribeck-type curves is to illustrate how friction varies with fluid
film thickness, other x-axis parameters that describe the magnitude of fluid film thickness are often

Boundary
Friction coefficient
Friction coefficient

Mixed

Hydrodynamic
Hydrodynamic
Mixed

Boundary

(a) Speed (b) Log (speed)

FIGURE 20.1  Schematic Stribeck curves for a lubricated conformal contact: (a) linear speed scale and (b) loga-
rithmic speed scale.
Boundary Lubrication and Boundary Lubricating Films 20-3

used in place of entrainment or sliding speed. These include the product (speed × viscosity) or even the
ratio of calculated fluid film thickness to surface roughness, the lambda ratio. In lubricated conformal
contacts the parameter (speed × viscosity/load), often called the Hersey number, can be used but this is
not appropriate for nonconformal contacts where the prevailing fluid film lubrication regime would be
elastohydrodynamic.

20.3  Solid-Like and Viscous-Like Boundary Films


There are many different types of boundary lubricating film, having a wide range of structures and
mechanical properties, so it is difficult to produce useful generalizations. However, in the last few years
one quite helpful concept to emerge is a distinction between two extreme types of boundary lubricating
film, solid-like and viscous-like films. In a conceptual sense, these can be considered in terms of their
impact on friction in the Stribeck curve.

20.3.1  Solid-Like Boundary Films


The traditional model of boundary lubrication assumes that solid boundary films form on rubbing sur-
faces and the adhesion between these films and the shear strength of their interface are both lower than
those of the underlying surfaces. This results in reduced wear and friction. Bowden and Tabor have
provided a simple adhesive model of boundary friction which explains the friction reduction caused by
solid-like boundary films [3].
In the contact of rough surfaces of stiff materials, actual solid–solid contact occurs only at a few
asperity conjunctions (Figure 20.2a). At these asperities, the pressure is so high that plastic deformation
occurs, which means that the total area of these conjunctions, i.e., the real area of contact Ac, is propor-
tional to the applied load W and inversely proportional to the hardness H of the solid:

W
Ac = (20.1)
H

In the absence of any load support by a fluid film and also any significant contribution to boundary
friction from interpenetration of asperities (ploughing), the friction is simply the tangential force
required to shear the asperity conjunctions. It is thus proportional to the real area of contact and is
given by

F = Act (20.2)

where τ is the shear strength of the interface between opposing asperities.


Combining these two equations to eliminate Ac gives

F t
m= = (20.3)
W H

(a) (b)

FIGURE 20.2  Schematic contact of rough surfaces: (a) no boundary film and (b) boundary lubricating film
present.
20-4 Theory and Practice of Lubrication and Tribology

Friction coefficient
Hydrodynamic

Boundary

Log (speed)

FIGURE 20.3  Schematic diagram of influence of solid-like boundary film on friction in the Stribeck curve.

For a homogeneous, clean metal, τ/H ≈ 1/6, so Equation 20.3 predicts a friction coefficient of 0.16. In
practice, friction is usually higher than this due to plastic junction growth, which increases Ac as shear-
ing takes place [3]. However, the friction coefficient can be significantly reduced if a thin low shear
strength layer is formed on a hard substrate, as shown in Figure 20.2b. The low shear strength layer must
be thin enough not to deform significantly and so to increase Ac. The friction coefficient μ is then

tb
m= (20.4)
H

where τb is the shear strength of the boundary lubricating film.


Thus solid-like boundary films are able to reduce the shear strength of the interacting asperities while
not increasing the real area of contact. A second, related mode of action is to reduce adhesive junction
growth. Bowden and Tabor showed how junction growth can lead to a much higher friction than pre-
dicted from simple adhesion [3]. The earlier model of friction of solid boundary films has been extended
by numerous authors, to allow for junction growth, take account of ploughing, consider boundary films
whose mechanical properties change with pressure, etc.
The formation of a low shear strength, solid-like boundary film will change the Stribeck curve as
shown in Figure 20.3. The friction in the boundary lubrication regime will be reduced but there will be
no change in the location of the lubrication regime boundaries. Friction in the mixed regime is then
often considered to be a simple sum of the contributions of friction of the boundary film and the hydro-
dynamic film, i.e.,

F = W ( x mb + (1 − x )mh ) (20.5)

where
W is the applied load
μb and μh are the friction coefficients in the boundary and hydrodynamic regimes, respectively (at the
contact conditions of sliding speed, temperature, etc.)
x is the fraction of the load supported by asperities

20.3.2  Viscous-Like Boundary Films


Very early in the history of boundary lubrication, it was hypothesized that this type of lubrication might
arise from the oil becoming highly viscous next to a polar solid surface [4]. Recently it has been shown
that some liquid lubricants do, indeed, form very thin, viscous layers adjacent to metal surfaces. These
Boundary Lubrication and Boundary Lubricating Films 20-5

Friction coefficient
Hydrodynamic

Boundary/mixed

Log (speed)

FIGURE 20.4  Schematic diagram of influence of viscous-like boundary film on friction in the Stribeck curve.

layers are entrained between rubbing surfaces at lower speeds than the bulk lubricant, leading to a
change in the Stribeck curve shown schematically in Figure 20.4. As speed is progressively reduced,
the contact passes from being separated by a bulk lubricant film to being separated by a viscous surface
film before eventually entering the conventional, mixed lubrication regime. The distinction between the
boundary and mixed lubrication regime thus becomes blurred. A version of Reynolds equation in which
viscosity is allowed to vary across the film thickness can be used to explore the impact of viscous-like
boundary films on film thickness and friction [5,6].

20.4  Detection and Measurement of Boundary Lubrication


In the last few years, several experimental techniques and instruments have become available which
have greatly improved the ability to study and characterize boundary lubricating films. Some of these
are described in the following sections.

20.4.1  Stribeck Curve Measurements


In the past, boundary lubrication performance was usually measured under a single speed condition,
but over the last few years the benefit of obtaining Stribeck curves, i.e., of measuring friction over a range
of speeds, has been recognized. This ensures that the prevailing lubrication regime is correctly identi-
fied and can also reveal information about the physical properties of the boundary film, e.g., whether it
is solid-like or viscous-like. Such measurements are generally made in a mixed sliding–rolling contact,
where the sliding speed is kept quite low and the entrainment speed, which controls fluid film forma-
tion, is progressively increased. In principle it would be possible to obtain such Stribeck curves from a
pure sliding contact, but in high pressure conditions, thermal effects tend to dominate friction behavior
at high speeds. Some Stribeck friction curves to illustrate the impact of boundary lubricating films on
friction are shown in Section 20.7.

20.4.2  Thin Film Interferometry


Many boundary lubricating films tend to be <5 nm thick and they often only exist within a rubbing con-
tact. In recent years, optical interferometry has been quite widely used to measure such boundary films.
In this, a lubricated contact is formed between a transparent surface (usually glass or sapphire) and a
reflective metal ball or roller. Light is shone into the contact and produces an interference image that
can be used to accurately measure the separation between the two surfaces. This method can measure
separations between 1 and 1000 nm and, at the lower limit of this range, can detect boundary films [7].
It is particularly suited to study viscous-type boundary films.
20-6 Theory and Practice of Lubrication and Tribology

20.4.3  Atomic Force Microscopy


In the last few years, the development of the atomic force microscope (AFM) and related scanning nano-
probes has considerably strengthened the understanding of boundary lubricating films. In the AFM, a
tiny, sharp stylus is scanned across a solid surface at a constant applied load to produce a map of surface
topography. In principle this is not different from a conventional stylus surface profilometer, but the
AFM can traverse a surface very rapidly, which mean that topography maps can be acquired quickly,
with spatial resolution down to the nanometer level [8]. The AFM has proved to be extremely valuable
for exploring some types of solid-like boundary film. One practical advantage of AFM is that it works
well even when the surface is covered in a layer of liquid, which means that surfaces can be examined
without draining or rinsing them, which might destroy delicate boundary lubricating films.

20.4.4  Chemical Analysis of Boundary Lubricating Films


Many advances in our understanding of solid-like boundary films have resulted from advanced surface
analysis methods such as time-of-flight secondary ion mass spectrometry (TOF-SIMS) and x-ray analy-
sis methods such as x-ray adsorption near-edge structure (XANES). One very recent and valuable tool
is focused ion beam sputtering (FIB) which can cut tiny trenches in solid surfaces so as to expose a cross
section of the surface and subsurface for analysis by tools such as scanning electron microscopy (SEM)
and transmission electron microscopy (TEM) [9]. Figure 20.5 shows a TEM cross-sectional image
through the boundary film, steel surface, and steel subsurface of a bearing lubricated by an antiwear
additive-containing lubricant [10]. The method is able to reveal the thickness, morphology, and compo-
sition of the film with depth. This technique is enlarging our understanding of solid-like boundary films
by revealing clearly how structure and composition vary though their thickness.

20.5  Film Replenishment


A very important characteristic of all boundary lubricating films, whatever their nature is, should be
noted, that they are continually replenished. This means that when the film is damaged by rubbing, it can
reform by adsorption or reaction of species from the supernatant liquid lubricant or the environment.

(a)

Surface layer

(b)
0.25 nm

0.21 nm 0.15 nm
(c)

50 nm (211)
(110)

FIGURE 20.5  Schematic diagram of influence of solid-like boundary film on friction in the Stribeck curve. (From
Evans, R.D. et al., Tribol. Trans., 48, 299, 2005. With permission.)
Boundary Lubrication and Boundary Lubricating Films 20-7

In some cases the film may be removed by rubbing faster than it is formed, in which case the perfor-
mance benefits of boundary lubrication will be lost, but even so, replenishment will still be occurring.
This is an important characteristic of boundary lubrication and distinguishes it from dry lubrication by
solid coatings and treatments, which are applied to the surfaces of tribological components before use
and whose benefit is lost when they are worn away.

20.6  Types of Boundary Film


Boundary films have three main roles: to control friction, to control the rate of wear, and to prevent
large-scale adhesive damage during rubbing—usually called “scuffing” or “scoring.” Table 20.1 lists
some common types of boundary lubricating film. Boundary films can originate from a wide range of
types of bonding of active species to the rubbing surfaces, from weak van der Waals bonding to chemi-
cal bond formation. In general, boundary films that involve chemical (and thus relatively strong) bond-
ing to the rubbing surfaces are needed to control wear and seizure, while both physically and chemically
bonded boundary films may reduce friction.
In Table 20.1, except for the first example of oxide films, the films are listed in order of increasing
strength of the boundary film–substrate bonding. Each of these films is briefly described in the follow-
ing sections.

20.6.1  Oxide Films


If completely clean, similar metals are rubbed together, strong adhesion and junction growth occurs at
contacting asperities, resulting in very high friction. However, in practice almost all metals, including
steels, rapidly acquire a thin oxide coating when exposed to air. Most metal oxides are far less adhesive
than metals, so the friction is greatly reduced, typically to a friction coefficient value of between 0.2 and
0.4. As the oxide films are worn away, the exposed metal will react from oxygen in the air, which is effec-
tively acting as a gaseous boundary lubricant, albeit not a very effective one. In practice, in moist air, the
boundary lubricating film on steel will be hydrated iron oxide. Oxidized surface films also form on some
polymer and ceramic surfaces. It is important to note that solid surfaces on which boundary lubricating
additives adsorb and react to form protective films are thus not generally the same as the bulk substrate
material, but an oxidized and hydrated form.

20.6.2  Base Fluid Boundary Films


Research on the properties of liquid films between very smooth mica surfaces has found that when such
films are only a few molecules thick, the liquid molecules can become ordered and this can considerably

TABLE 20.1  Boundary Lubrication


Type of Boundary Film Example Film Formers Effect
1 Oxide films
2 Base fluid films Ester/hydrocarbon mixtures Reduce friction
3 Adsorbed polymer films Dispersant olefin copolymer, dispersant Reduce friction
polyalkylmethacrylate
4 Organic friction modifier films Olelylamide, glycerol monooleate Reduce friction
5 Deposited solid films Overbased calcium alkyl sulfonate detergent Reduce friction, reduce wear
6 Phosphate/polyphosphate films Organophosphorus-based antiwear additives Reduce wear
7 MoS2 nanocrystals Molybdenum dialkyldithiocarbamate Reduce friction
8 Iron sulfide films Sulfur- and sulfur/phosphorus-based Prevent scuffing
extreme pressure additives
20-8 Theory and Practice of Lubrication and Tribology

increase their viscosity [11,12]. This is known as a “confinement” or “solvation” effect. In practice, how-
ever, this phenomenon appears to occur only for pure, regularly shaped molecules such as linear alkanes
between atomically smooth surfaces, so its practical significance in engineering systems is disputed.
One situation where base fluids can show significant boundary lubricating properties is when the
fluid is a mixture, with one component being more polar than the other, e.g., an ester mixed with a
hydrocarbon. With such blends there is a tendency, due to van der Waals forces, for the more polar mol-
ecules to concentrate close to polar surfaces such as metals. If the more polar molecules have a signifi-
cantly higher or lower molecular weight and thus viscosity than the less polar ones, this effect will result
in the near surface fluid becoming more or less viscous than the bulk. This effect has been demonstrated
by both film thickness and friction measurements [13,14]. It has practical importance since a polar base
fluid is often blended with nonpolar, hydrocarbon base oils to control the swell of seals and to improve
the solubility of additives in the latter.

20.6.3  Adsorbing Polymer Boundary Films


Polymers are widely used as additives in lubricants. Although employed primarily to modify the rheo-
logical properties of their blends, it has been shown some polymers that contain polar monomer units
can also adsorb on polar surfaces such as metals to form boundary lubricating films.
The nature of these boundary films is shown schematically in Figure 20.6. Segments of polymer mol-
ecules adsorb while other segments remain in solution, to form a film whose thickness is of the order
of the polymer coil diameter. The resultant films contain a high concentration of polymer chains and
thus have much higher viscosity than the bulk solution. As such, they form viscous boundary films [15].
These films are effective at reducing friction in the mixed lubrication regime but tend to be less useful
at very low speeds, or during reversal, since they are rarely strongly enough bound to the surfaces to
survive very severe contact conditions.
The influence of polymer architecture on boundary film formation has been studied and it has been
found that block copolymer molecules, where there are several adjacent polar units, tend to adsorb and
form films more easily than random copolymers, where the polar monomer units are randomly distrib-
uted among the nonpolar units in the polymer chain [16].

20.6.4  Organic Friction Modifier Films


Organic friction modifier (OFM) additives are surfactant molecules, having a long, straight hydrocar-
bon chain with a polar group at one end. The polar group bonds with polar surfaces such as metal
oxides, while the alkyl chain remains in the relatively nonpolar, liquid lubricant. Initially it was sug-
gested that the molecules adsorb from oil solution to form single, complete monolayers on both surfaces
(Figure 20.7a). Low friction then results because of the easy slip between adjacent methyl groups at the
ends of pairs of monolayers. This model was later extended as shown in Figure 20.7b to include some
penetration of the monolayers at asperity conjunctions so that friction reduction arises both from the
low shear strength of the methyl–methyl group interface and from the beneficial effect of the absorbed

(a) At slow speed (b) At high speed

FIGURE 20.6  Presence of an adsorbed polymer boundary film in (a) thin and (b) thick fluid film conditions.
Boundary Lubrication and Boundary Lubricating Films 20-9

(a) (b)

FIGURE 20.7  Two models of boundary films formed by an organic friction modifier; (a) single, complete mono-
layers on both surfaces; (b) some penetration of the monolayers at asperity conjunctions.

monolayers of inhibiting junction growth [17]. This was based on the finding that some metal–metal
contact occurs at asperity contacts even when a friction-reducing OFM is present. In Figure 20.7, the
alkyl chains are shown aligned perpendicular to the surface. In practice, studies using the AFM indicate
that the molecules tend to tilt under load, and also the ends of the chains slightly interpenetrate [18].
Typical OFMs used today include long-chain amides, amines, partial esters, and phosphates. Often
unsaturated alkenyl chain compounds such as oleyl derivatives are preferred to saturated alkyl chains
because they are more soluble in hydrocarbon base oils.
There is still considerable debate as to whether the OFM molecules bond physically (i.e., reversibly)
or chemically with metal oxide surfaces. However, there is consensus that the main strength of OFM
boundary films originates from lateral van der Waals forces between adjacent, parallel alkyl chains
[19]. Bonding between the polar group of the OFM molecules and the surface position the molecules on
the surface initially, but the side chain forces hold the monolayer together under load. For this reason,
chains having at least 12 carbon atoms are generally needed to support high pressure and straight, rather
than branched, chains are most effective since they are able to pack more closely [19,20].
It should be noted that although the most widely held view is that OFMs form protective monolayers
on polar surfaces, it has also been suggested that some OFMs, in particular carboxylic acids, may form
considerably thicker films, possibly of metal carboxylates [21,22].

20.6.5  Deposited Solid Boundary Films


In recent years, there has been considerable interest in the use of nanoparticulate additives suspended
in lubricants to reduce friction and wear. Possible candidates include carbon nanotubes and inorganic
fullerenes made from MoS2 or WS2. Such additives have not yet reached general acceptance, but one
family of nanoparticulate additives that is widely used are the overbased detergents. These consist of
tiny (5 to a few hundred nanometers diameter) calcium or magnesium carbonate colloidal particles
stabilized by an oil-soluble detergent. Although their main roles are as a detergent and acid-neutralizing
agent, they also contribute to friction control and wear reduction. When compressed within rubbing
contacts, these particles adhere strongly to rubbed surfaces and coalesce to form a thick layer of metal
carbonate [23]. Figure 20.8 shows an AFM map and profile of a typical such film, which consists of
islands of calcite (CaCO3) about 100–300 nm thick [24].

20.6.6  Reacted Solid Boundary Films


Several different types of lubricant additives react chemically with rubbing ferrous surfaces to form
inorganic boundary films. Three of the most important are described in the following sections.

20.6.6.1  Zinc Dialkyldithiophosphates and Other Phosphorus-Based Additive Films


Organophosphorus additives have been used since the 1940s to reduce wear in engines and transmissions
[25,26]. These additives are now known to react with rubbing steel and some other metal oxide surfaces to
form quite thick films of a metal phosphate/polyphosphate glass-like material. By far the most widely used
such additives in crankcase engine oils are zinc dialkyldithiophosphates (ZDDPs). These form thick films
very rapidly because they have a readily available supply of zinc cations to help stabilize the phosphate
20-10 Theory and Practice of Lubrication and Tribology

100 µm
190.95 nm

50 µm
0.00 nm

0 µm
0 µm 50 µm 100 µm

500 nm

250

0
0 50 100 µm

FIGURE 20.8  AFM topographic image and profile of the boundary film formed on a steel surface by an overbased
calcium alkylsulfonate detergent. (From Topolovec-Miklozic, K. et al., Tribol. Lett. 29, 33, 2008. With permission.)

glass. Metal-free organophosphate additives react more slowly to give thinner films since growth of the
film requires the generation of iron cations. The films formed by ZDDP are typically 50–150 nm thick and
have a pad-like structure, quite similar to that of the calcite film shown in Figure 20.8.

20.6.6.2  Organomolybdenum Additive Films


Some soluble organomolybdenum compounds are very effective friction-reducing additives. Typical
examples are molybdenum dialkyldithiocarbamates and dithiophosphates. These additives, whose mol-
ecules contain both Mo and S atoms, have been shown to react in rubbing contacts to form tiny crystals
of MoS2, typically only about 1–2 lattice layers deep (<3 nm) and 20 nm across [27]. The layer-lattice
structure of MoS2 means that these crystals have very low shear strength. AFM has shown that they
form only on asperities, where, since the asperities support the applied load in low speed conditions,
they produce a large reduction in friction [28]. Organomolybdenum additives also form similar MoS2
crystals on the boundary films generated by ZDDP and other antiwear additives and so work well in
combination with these additives.

20.6.6.3  EP Films
Extreme pressure (EP) additives were introduced in the 1930s to control scuffing of hypoid gears. They
are now used in most transmission and metal cutting oils. They are predominantly oil-soluble organo-
sulfur compounds.
Scuffing is a catastrophic process that occurs in lubricated contacts operating at a combination of
high load and high sliding speed. It is triggered by the local breakdown of the lubricant films that sepa-
rate sliding asperities. This leads to an increase in asperity friction and thus temperature, which in turn
results into yet more lubricant film breakdown. This positive feedback cycle rapidly produces collapse of
the whole lubricant film and large-scale metal–metal adhesion [29]. EP additives are designed to react
with the hot steel asperities exposed during the onset of scuffing, to form mixed iron sulfide/oxide.
This film is quite easily abraded and a sequence of rapid reaction and film wear ensues, which results
Boundary Lubrication and Boundary Lubricating Films 20-11

in smoothing and improved conformity of the surfaces. The net effect is to lower the contact pressure,
surface temperature, and friction. All of these help halt the feedback process that leads to scuffing.

20.7  Friction Control by Boundary Films


A widely used instrument for measuring how friction varies with speed is the minitraction machine.
In this, a ball is loaded and rolled against the flat surface of a rotating disk while measuring the friction
force of the ball. The mean (entrainment) speed is gradually increased while holding the slide-to-roll
ratio constant to produce a Stribeck curve. Figure 20.9 shows results using this device and compares
the impact of a “solid-like” boundary film formed by the OFM olelyamide, which reduces friction at
low speeds, with a “viscous-like” functionalized polyalkylmethacrylate which postpones the onset of
boundary lubrication to very low entrainment speed, in this case to below 0.01 m/s.
For some solid-like monolayer boundary films formed by OFMs, it has been shown that the shear
strength, τb, of the interface depends on pressure p, temperature T, and sliding speed V according to

tb = to + ap (20.6)

tb = toʹ − bT (20.7)

tb = toʹʹ + qV (20.8)

where to, to�, to��, α, β, and θ are constants [30]. Since the boundary friction coefficient is the shear strength
divided by the pressure, this implies that, in the absence of any ploughing term, boundary friction of
OFMs depends on pressure according to a two-term equation:

to
mb = +a (20.9)
p

In practice, the first term is usually small so the boundary friction coefficient is essentially independent
of load except at very low loads.

0.16

Base oil
Olelyamide
0.12
Friction coefficient

D-PAMA

0.08

0.04

0
0.001 0.01 0.1 1 10
Entrainment speed (m/s)

FIGURE 20.9  Friction versus entrainment speed curves from the minitraction machine for a Group II base oil
and solutions of two friction modifier additives, olelyamide and a block dispersant polyalkylmethacrylate (PAMA)
in this base oil.
20-12 Theory and Practice of Lubrication and Tribology

0.16
Initial

Friction coefficient
0.12 5 min

60 min
0.08

0.04

0
0.001 0.01 0.1 1 10
Entrainment speed (m/s)

FIGURE 20.10  Friction versus entrainment speed curves for a solution of overbased calcium alkylsulfonate
detergent at three stages of a prolonged rolling–sliding test. (Adapted from Topolovec-Miklozic, K. et al., Tribol.
Lett. 29, 33, 2008.)

More recently, the boundary friction properties of OFM films have been related to their adhesion
properties and a fundamental relationship between adhesion hysteresis and boundary friction has been
proposed [31].
Figure 20.10 shows the impact of an overbased calcium alkylsulfonate detergent on friction. Three
Stribeck curve are shown, taken at different times during a test as the CaCO3 boundary film develops.
There are two distinct effects: (i) reduction of friction at low speeds and (ii) increase of the entrainment
speed at which the mixed lubrication regime begins. The first of these effects is normal OFM monolayer
friction behavior produced by the linear alkylsulfonate detergent molecules. In this case, as for many
saturated, linear alkyl chain OFMs, boundary friction increases linearly with log(sliding speed) in the
boundary lubrication regime, as predicted by Equation 20.8. The second effect, of raising the onset
of mixed lubrication to higher speeds as the boundary film develops, is believed to result from the
increased roughness of the surface due to the formation of a thick, pad-structured, calcite boundary
film. A similar shift in the Stribeck curve to higher speed in friction is also seen with ZDDP and other
organophosphorus-based antiwear films when these form thick, rough boundary films.
Organomolybdenum additives tend to form boundary films with very low friction coefficients in the
range 0.04–0.06, quite constant with speed.

20.8  Wear Control by Boundary Films


Almost all boundary films reduce wear compared to a hydrocarbon base oil, because these films par-
tially or wholly reduce the extent of asperity contact and adhesion that results in adhesive wear. The
extent of wear reduction then depends on the integrity of the boundary film and the extent to which it is
removed and replaced during rubbing. Antiwear additives such as ZDDP are very effective at separating
the underlying surfaces and also strong enough to resist rubbing. Also, since the outer layers of ZDDP
films contain very little iron, even if these films are removed there is little loss of ferrous substrate and
thus negligible wear.
A useful concept in understanding the impact of boundary additives on wear is the “adhesive–
corrosive balance.” This is shown schematically in Figure 20.11. As additive concentration (or chemi-
cal activity) is increased, the additive forms a boundary film more rapidly and this reduces the level
of adhesive wear. However, if this chemical film is being also removed by abrasion or fatigue, an
increase in reaction rate can also lead to a faster rate of wear due to rapid film removal. The result is
an optimal additive concentration at which there is minimum wear at a given condition of contact
severity.
Boundary Lubrication and Boundary Lubricating Films 20-13

Overall wear rate

Wear rate

Adhesive
wear

Corrosive wear

Additive concentration/chemical activity

FIGURE 20.11  Schematic diagram of the adhesive–corrosive balance in controlling wear rate.

20.9 Correlation between Friction and Wear


Properties of Boundary Films
Normally lubricants that show low wear also show low boundary friction and vice versa. This is because
effective boundary films act by preventing adhesion of the surfaces on which they are formed, so lubri-
cants that do this thereby reduce both wear and friction. The exception is when an additive is very
chemically reactive and forms boundary films that are quite weak and thus easily removed by abrasion,
such as EP additives. In this case, the additives can reduce friction but also give high wear. This concept
can also be deduced from the adhesive–corrosive balance model outlined in the previous section.

20.10  Strength of Boundary Films


One of the most problematic properties of boundary films is their “strength.” This is usually loosely
defined but essentially means the operating conditions of pressure, sliding speed, and temperature
above which the boundary films are removed from the rubbing surfaces and so cease to reduce friction
and/or wear. During severe conditions of high speed sliding and pressure, the result of this film loss is
usually scuffing. Under more mild conditions, it is simply evidenced as an increase in friction and wear
to levels expected due to surface adhesion in the absence of a boundary film and is often noted as a tran-
sition from “mild” to “severe” wear.
For OFMs, a loss of ability to reduce friction is seen above a critical temperature and this is ascribed
to desorption or melting of the adsorbed/reacted monolayers [32]. The transition from low to high
friction is more marked for unreactive surfaces, such as stainless steel, than for mild steel. The effect
of load on OFM boundary films is less clear. Hirst and Hollander have shown that the temperature at
which OFM boundary films fail depends on both the height and wavelength of the surface roughness
as well as the applied load [33]. Their results suggest that OFM film failure occurs when a critical level
of plastic deformation of the contact asperities is exceeded. It was also found that very smooth surfaces
failed at relatively low loads, possibly because large-scale junction growth occurs most easily for such
surfaces.
At a combination of very high sliding speed and very high load, most boundary films, including those
formed by friction modifiers and antiwear additives, fail, resulting in scuffing of the surfaces. Many
scuffing models are based on the principle that lubricating films fail at a critical temperature, which is
20-14 Theory and Practice of Lubrication and Tribology

2000

1600

Normal force Fn/(N)


III
1200 II

800

400 I

0
10–2 10–1 100
Sliding velocity V/(m/s)

FIGURE 20.12  Typical IRG transition diagram. (From Schipper, D.J. and de Gee, A.W.J., Lubr. Sci., 8, 27, 1995.
With permission.)

reached at high load and sliding speed [29]. This temperature is the sum of the out-of-contact surface
temperature of the rubbing bodies and the transient flash temperature which occurs within the contact
due to frictional heating. Such critical temperature and related models usually predict that scuffing
will occur when the product of applied load, W, and the sliding speed, V, raised to a constant power n,
exceeds a critical value k, i.e., PV n = k. This power normally lies between 0.5 and 1.5, depending on the
model chosen.
In most lubricated contacts, scuffing involves the breakdown of any elastohydrodynamic film
present followed by the boundary film. This is usefully encapsulated by friction transition diagrams
of load versus sliding speed such as that shown in Figure 20.12 [34]. At lower sliding speeds, as
the load is increased, first any elastohydrodynamic film is lost (transition I–II). Then at a higher
load, the boundary film fails (transition II–III). Both of these transitions are accompanied by a rise
in friction. At high sliding speeds, when the load is increased, the breakdown of the EHD film is
accompanied immediately by failure of the boundary film (transition I–III). Both transition curves
represent lines of constant PVn at which the respective elastohydrodynamic and boundary lubricant
films break down.

20.11  Prediction of Boundary Lubrication Performance


Although many mathematically based models of the boundary lubrication regime have been developed,
unlike for the hydrodynamic and elastohydrodynamic lubrication regimes, such models are rarely used
in the design or analysis of lubricated systems. This is partly because of gaps in current understanding
of the boundary lubrication regime, especially when lubricants containing several different surface-
active additives are used. Also boundary lubricating films are very diverse, and most models are only
applicable to a single type of film.
Instead, empirical or rule of thumb approaches are still largely employed to estimate friction and wear
performance in the boundary lubrication regime. For friction, values of friction coefficient of 0.08–0.1
for OFMs and 0.04–0.06 for soluble molybdenum–sulfur compounds are usually taken, although these
have to be confirmed by experiment for the lubricant–surface combination of interest. For wear, the
Archard wear equation is generally considered to be valid, with boundary films acting to reduce the
value of the wear coefficient to a value that has to be determined by experiment. Scuffing models, in par-
ticular the Blok flash temperature scuffing model [35], are quite often used to predict the likely impact
on scuffing performance of changes of gear geometry or operating conditions to gears with measured
performance. They are, however, not yet reliable enough to predict the impact of changes of lubricant
design on scuffing.
Boundary Lubrication and Boundary Lubricating Films 20-15

References
1. Wells, H.M. and Southcombe, J.E., The theory and practice of lubrication: The ‘germ’ process. J. Soc.
Chem. Lond. 39, 51T–60T (1920).
2. Hardy W.B. and Doubleday, I., Boundary lubrication—The paraffin series. Proc. R. Soc. A 100, 550
(1922).
3. Bowden, F.P. and Tabor, D., The friction and lubrication of solids, chapter V. Internal Series of
Monographs on Physics, Clarendon Press, Oxford, U.K. (1986).
4. Kingsbury, A., A new oil-testing machine and some of its results. Trans. ASME 24, 143–160 (1903).
5. Qingwen, Q., Yahong, H., and Jun, Z., An adsorbent layer model for thin film lubrication. Wear 221,
9–14 (1998).
6. Ingram, M., Noles, J., Watts, R., Harris, S., and Spikes, H.A., Frictional properties of automatic trans-
mission fluids: Part 2: Origins of friction-sliding speed behaviour. Tribol. Trans. 54, 154–167 (2011).
7. Johnston, G.J., Wayte, R., and Spikes, H.A., The measurement and study of very thin lubricant films
in concentrated contacts. Tribol. Trans. 34, 187–194 (1991).
8. Topolovec-Miklozic, K. and Spikes, H.A., Application of AFM to the study of lubricant additive
films. Trans. ASME J. Tribol. 127, 405–415 (2005).
9. Evans, R.D., More, K.L., Darragh, C.V., and Nixon, H.P., Transmission electron microscopy of
­boundary-lubricated bearing surfaces. Part II: Mineral oil lubricant. Tribol. Trans. 47, 430–439
(2004).
10. Evans, R.D., More, K.L., Darragh, C.V., and Nixon, H.P., Transmission electron microscopy of
boundary-lubricated bearing surfaces. Part II: Mineral oil lubricant with sulfur- and phosphorus-
containing gear oil additives. Tribol. Trans. 48, 299–307 (2005).
11. Granick, S., Motion and relaxations of confined liquids. Science 253, 1374–1379 (1991).
12. Klein, J. and Kumacheva, E., Simple liquids confined to molecularly thin layers. I. Confinement-
induced liquid-to-solid phase transitions. J. Chem. Phys. 108, 6996–7009 (1998).
13. Guangteng, G. and Spikes, H.A., The control of friction by molecular fractionation of base fluid
mixtures at metal surfaces. Tribol. Trans. 40, 461–469 (1997).
14. Bovington, C., Fuel economy additive and lubricant composition containing same. US Patent
5962381 (1999).
15. Smeeth, M., Gunsel, S., and Spikes, H.A., Boundary film formation by viscosity index improvers.
Tribol. Trans. 39, 726–734 (1996).
16. Fan, J., Stohr, T., Muller, M., and Spikes, H.A., Reduction of friction by functionalized viscosity index
improvers. Tribol. Lett. 28, 287–298 (2007).
17. Bowden, F.P. and Tabor, D., The friction and lubrication of solids, chapter X. Internal Series of
Monographs on Physics, Clarendon Press, Oxford, U.K. (1986).
18. Benitez, J.J., Kopta, S., Diez-Perez, I., Sanz, F., Ogletree, D.F., and Salmeron, M., Molecular packing
changes of octadecylamine monolayers on mica induced by pressure and humidity. Langmuir 19,
762–765 (2003).
19. Jahanmir, S., Chain length effects in boundary lubrication. Wear 102, 331–349 (1985).
20. Studt, P., The influence of structure of isometric octadecanols on their adsorption from solution on
iron and their lubricating properties. Wear 70, 329–334 (1987).
21. Allen, C.M. and Drauglis, E., Boundary layer lubrication: Monolayer or multilayer. Wear 14, 363–384
(1969).
22. Anghel, V., Bovington, C., and Spikes, H.A., Thick film formation by friction modifier additives. Lubr.
Sci. 11, 313–335 (1999).
23. Cizaire, L., Martin, J.M., Gresser, E., Truong Dinh, N., and Heau, C., Tribochemistry of overbased
calcium detergents studied by ToF-SIMS and other surface analysis. Tribol. Lett. 17, 715–721 (2004).
24. Topolovec-Miklozic, K., Forbus, T.R., and Spikes, H.A., The film-forming and friction properties of
overbased calcium sulphonate detergents. Tribol. Lett. 29, 33–44 (2008).
20-16 Theory and Practice of Lubrication and Tribology

25. Spikes, H.A., The history and mechanisms of ZDDP. Tribol. Lett. 17, 465–485 (2004).
26. Spikes, H.A., Low and zero-sulphated ash, phosphorus and sulphur anti-wear additives for engine
oils. Lubr. Sci. 20, 1893–1901 (2008).
27. Groissard, C., Martin, J.-M., Le Mogne, T., Inoue, K., and Igarashi, J., Friction-reducing mechanisms
of molybdenum dithiocarbamate zinc dithiophosphate combination: New insights in MoS2 genesis.
J. Vac. Sci. Technol. A 17, 884–890 (1999).
28. Topolovec Miklozic, K., Graham, J., and Spikes, H., Chemical and physical analysis of reaction films
formed by molybdenum dialkyldithiocarbamate friction modifier additive using Raman and atomic
force microscopy. Tribol. Lett. 11, 71–81 (2001).
29. Bowman, W.F. and Stachowiak, G.W., A review of scuffing models. Tribol. Lett. 2, 113–131 (1996).
30. Briscoe, B.J. and Evans, D.C.B., The shear properties of Langmuir–Blodgett layers. Proc. R. Soc. Lond.
A 380, 389–407 (1982).
31. Yoshikazakawa, H., Chen, Y.-L., and Israelachvili, J., Fundamental mechanisms of interfacial friction.
1. Relation between adhesion and friction. J. Phys. Chem. 97, 4128–4140 (1993).
32. Beltzer, M. and Jahanmir, S., Effect of additive molecular structure on friction. Lubr. Sci. 1, 3–25
(1988).
33. Hirst, W. and Hollander, A.E., Surface finish and damage in sliding. Proc. R. Soc. Lond. A 337,
374–394 (1974).
34. Schipper, D.J. and de Gee, A.W.J., Lubrication modes and the IRG transition diagram. Lubr. Sci. 8,
27–35 (1995).
35. Castro, J. and Seabra, J., Scuffing and lubricant film breakdown in FZG gears Part II. New PV scuffing
criteria, lubricant and temperature dependent. Wear 215, 114–122 (1998).
21
Additive Technology
21.1 Mechanism of Oxidation and Antioxidants................................ 21-1
Introduction  •  Oxidation  •  Antioxidants  •  Antioxidant
Synergism
21.2 Mechanism of Contamination and Deposit Control................. 21-9
21.3 Mechanism of Sludge and Varnish Control................................ 21-9
Preventing Sludge and Varnish Buildup
21.4 Mechanism of Wear Control by Antiwear and EP
Additives.................................................................................... 21-10
Antiwear Additives  •  Additive Interactions and Soot
Concerns  •  EP Additives
Elaine S. Yamaguchi 21.5 Mechanism of Corrosion Control............................................... 21-12
Chevron Oronite
Corrosion Inhibitors
Company LLC 21.6 Mechanism of Friction Modifiers............................................... 21-15
Physisorption: Formation of Absorbed Layers  •  Chemisorption:
Gaurav Bhalla Formation of Reacted Layers
Chevron Oronite 21.7 Mechanism of Viscosity Index Improver................................... 21-17
Company LLC
21.8 Mechanism of Foam Inhibitors................................................... 21-19
Vincent J. Gatto 21.9 Conclusions and Future Needs.................................................... 21-19
Albemarle Corporation References................................................................................................... 21-20

21.1  Mechanism of Oxidation and Antioxidants


21.1.1  Introduction
Lubricants are often exposed to severe environmental conditions that require the use of antioxidants
to inhibit oxidation. In recent years, the effectiveness of various antioxidants has been complicated by
a number of confounding factors. These include drastic base stock changes, environmental pressures
to reduce lubricant volatility and extend drain intervals, reductions in finished lubricant sulfur, ash,
and phosphorus, and the exceedingly fast pace of new and more demanding oxidation and deposit
control specifications for lubricants. This has resulted in a renewed interest in lubricant degradation
mechanisms and the additives (i.e., antioxidants) used to suppress this degradation. The purpose of
this chapter is to review the chemistry of lubricant oxidation and antioxidant function as it pertains
to the stabilization challenges encountered with today’s lubricants. The mechanism of function for the
most important classes of antioxidants will be discussed. In addition, the criteria for proper antioxi-
dant selection, synergistic interactions between different classes of antioxidants, and the impact hydro-
cracked and synthetic base stocks have on the performance of the various antioxidant classes will be
presented.

21-1
21-2 Theory and Practice of Lubrication and Tribology

21.1.2  Oxidation
The inhibition of oil oxidation is a complicated problem. This is mainly due to the fact that hydrocarbon
oxidation can be catalyzed by a wide variety of materials. Some byproducts of oxidation can sometimes
inhibit further oxidation, while other byproducts act as oxidation catalysts. Much of today’s under-
standing of hydrocarbon oxidation, and the role antioxidants play in preventing oxidation, has been
extensively researched and comprehensive reviews have been published [1]. There have also been reviews
specific to the oxidation and stabilization of lubricants [2,3].
An understanding of three aspects of oil oxidation is important when developing modern lubricants.
These are the mechanism of oil oxidation, the physical effects of oil oxidation, and the impact hydro-
cracked base stocks have on oxidation.

21.1.2.1  Chemical Degradation


21.1.2.1.1 Chain Initiation, Propagation, and Branching Steps
In the presence of heat and metal catalysts, hydrocarbons (RH) will react with oxygen to form alkyl
radicals (R·). This is the initiation step in oil oxidation (Reactions 21.1 and 21.2).

RH + O2 → R ⋅ + HO2 ⋅ (21.1)

2 RH + 1/2O2 → H2O + 2 R ⋅ (21.2)

These alkyl radicals are unstable and will react rapidly with oxygen present in the system. This leads to
the formation of alkylperoxy radicals (ROO·) (Reaction 21.3). The alkylperoxy radicals are also unstable
and will rapidly react with each other and alkyl radicals. However, in the early stages of oxidation there
are very few radicals in the oil. Alkylperoxy radicals are much more likely to react with other hydrocar-
bons in the oil. This reaction leads to additional alkyl radicals and hydroperoxides (ROOH) (Reaction
21.4). This represents the propagation step in oil oxidation.
A variety of chain-branching steps are possible based on the lubricant type and system temperature.
Some of these are shown in Reactions 21.5 through 21.8. Under high-temperature conditions, the hydro-
peroxides are unstable and decompose to peroxy (ROO·), alkoxy (RO·), and hydroxyl (HO·) radicals
(Reactions 21.5 and 21.6):

R ⋅ + O2 → ROO ⋅ (21.3)

ROO ⋅ + RH → ROOH + R ⋅ (21.4)

ROOH → HO ⋅ + RO ⋅ (21.5)

2 ROOH → RO ⋅ + ROO ⋅ + H2O (21.6)

RO ⋅ + RH → ROH + R ⋅ (21.7)

HO ⋅ + RH → H2O + R ⋅ (21.8)

Many other initiation, propagation, and chain-branching reactions have been proposed and excellent
reviews are available that cover these mechanisms and the kinetics of autoxidation [4,5].

21.1.2.1.2 Metal-Catalyzed Chain Initiation


Metals are the most common form of contamination in lubricants. The metals of most concern in a
lubricant system are copper and iron. Metal contamination is usually caused by corrosion and wear in
the lubricating system, which results in the formation of oil-soluble metallic components. What makes
Additive Technology 21-3

copper and iron so detrimental is that both these metals function in multiple oxidation states to decom-
pose hydroperoxides. Thus, even small amounts of copper or iron can promote the rapid degradation
of the relatively stable hydroperoxides back to the unstable peroxy and alkoxy radicals (Reactions 21.9
and 21.10):

Fe3 + + ROOH → Fe2 + + ROO ⋅ + H + (21.9a)


Fe2 + + ROOH → Fe3 + + RO ⋅ + HO− (21.9b)


Cu 2+ + ROOH → Cu + + ROO ⋅ + H + (21.10a)


Cu + + ROOH → Cu 2 + + RO ⋅ + HO− (21.10b)


As hydroperoxide levels increase, their decomposition to radical products begins to predominate.


The effects of metal-catalyzed oxidation are dependent on the lubricating system being stressed.
Metal-catalyzed oxidation and inhibition in polymeric hydrocarbon materials has been extensively
reviewed [6].

21.1.2.1.3 Formation of Aldehydes and Ketones


A critical sequence in the oxidation process is the formation of aldehydes and ketones. These are
unique chemical species since their subsequent reactions result in further degradation of the lubricat-
ing system. The formation of oligomers, polymers, and eventually sludge and deposits all initiate from
low-molecular-weight aldehydes and ketones initially generated during the early stages of oxidation.
Two of the more accepted mechanisms for aldehyde and ketone formation are shown in Reactions 21.11
and 21.12:

RRʹ HCO ⋅ → RCH=O + Rʹ ⋅ (21.11a)


RRʹRʹʹCO ⋅ → RRʹC=O + Rʹʹ ⋅ (21.11b)


2RRʹCHOO ⋅ → RʹRC=O + RCHOHR ʹ + O2 (21.12)


A key point regarding the early stages of oil oxidation is highlighted in Reaction 21.12. In this reaction,
the peroxy radical undergoes chain scission to form a carbonyl compound and alcohol of lower molecu-
lar weight. The physical effects of this chemical change will be discussed in detail in Section 21.1.2.2.

21.1.2.1.4 Formation of Water and Alcohols


Reaction 21.12 illustrates the formation of an alcohol during the oxidation process. In addition, an alk-
oxy radical (RO·) or hydroxy radical (HO·) can abstract a hydrogen from another oil molecule to pro-
duce alcohols and water (Reactions 21.7 and 21.8). Alcohols and water are considerably more stable in a
lubricating system compared to radicals. However, the formation of water can increase the potential for
corrosion under certain conditions. Also, alcohols may undergo secondary reactions with esters, acids,
and other species to further alter the physical and chemical properties of the lubricant.

21.1.2.1.5 Formation of Organic Acids


Carboxylic acids are formed by oxidation of aldehydes and ketones. The mechanism for this oxidation is
shown in Reactions 21.13 and 21.14. Ketones, in general, are oxidized more slowly than aldehydes. The
21-4 Theory and Practice of Lubrication and Tribology

ketone oxidation can proceed through a nonradical rearrangement that results in two lower-molecular-
weight organic fragments [7]. The aldehyde oxidation can proceed through an intermediate peroxy acid
that undergoes molecular decomposition to form the carboxylic acid [8]:

RCH=O + O2 → + R(C=O)OH + ½O2 (21.13)

R(R ʹCH2 )C=O + O2 → R(C=O)OH + R ʹCH=O (21.14)

21.1.2.1.6 Oligomerization and Polymerization


Condensation reactions start playing a significant role as the levels of aldehydes and ketones become
appreciable. These reactions are called Aldol condensations and are illustrated in Reaction 21.15 for
ketones [9]. In fresh oil systems, the rate of these reactions is very slow. However, as oil begins to oxi-
dize the amount of carboxylic acids increase. These organic acids are very effective catalysts for Aldol
condensation reactions. Thus the process of oxidation generates potent condensation catalysts that con-
vert the low-molecular-weight carbonyl compounds into higher-molecular-weight polar oligomers. The
molecular weight will continue to rise resulting in the formation of polar polymers. The physical effects
of these polymeric materials can be devastating on the lubricating system:

2R(R ʹCH2 )C=O → R(C=O)C(R ʹ)CH=C(R)CH2R ʹ + H2O (21.15)

21.1.2.2  Physical Effects of Oil Oxidation


A simplified representation of the physical effects of oil oxidation is shown in Figure 21.1. The lubricant
combines with fuel, blow-by gases, and metals, in the presence of heat and the combustion process, to
undergo a series of very complex chemical reactions. These reactions lead to a wide variety of oxidized
oil monomers. These are polar materials attracted to each other and the metal surfaces of the system.
In the early stages of oxidation, chain scission results in an apparent drop in molecular weight of the
oxidized oil monomers. Physically, this manifests itself as a reduction in oil viscosity. This phase of
oil oxidation is short lived as condensation reactions start to predominate, resulting in an increase in
polarity and molecular weight. On the much hotter metal surface, the low-molecular-weight oxidized
oil monomers will condense with each other in complex polymerization reactions that eventually lead
to varnish and deposits. Alternatively, the aggregated oxidized oil monomers in the bulk oil will con-
dense to form low-molecular-weight polar oligomers commonly characterized as varnish and sludge

Fuel
Heat and Complex Oxidized
combustion
Lubricant chemical oil monomers
reactions
Blowby gases
Varnish and
Metal sludge
surface precursors
adsorption Surface- Inorganic
active salts
Polymerization polymers

Varnish Sludge
Deposits

FIGURE 21.1  Physical effects of oil oxidation in crankcase oils.


Additive Technology 21-5

precursors. These oligomers still retain oil solubility. However, further reaction with inorganic salts and
other polar polymers in the oil eventually leads to the formation of very high-molecular-weight aggre-
gates that fall out of the oil as sludge. The polar sludge aggregates will eventually find their way to the
metal surface and form deposits.

21.1.2.3  Solvent-Refined versus Hydrocracked Oil Oxidation


There are major chemical differences between traditional solvent-refined oils and hydrocracked/
dewaxed oils that manifest themselves in the oxidation process [10]. Hydrocracked oils and polyalphao-
lefins (PAOs) have no naturally occurring nitrogen or sulfur and very low or no aromatics. The process
of removing aromatics, nitrogen, and sulfur changes oil oxidation in the following ways:
1. The removal of aromatics improves an oil’s oxidative stability because aromatics are more easily
oxidized than saturated hydrocarbons [11].
2. The removal of sulfur and nitrogen improves an oil’s oxidative stability because sulfur and nitro-
gen heterocycles can catalyze further oxidation [12].
3. The highly saturated hydrocracked oils have poorer solubility properties compared to tra-
ditional solvent-refined oils [13]. This means polar oil monomers will aggregate more
easily in hydrocracked oils leading to more rapid sludge and deposit formation once oxidation
begins.
4. Although aromatics, sulfur, and nitrogen promote oxidation, they also oxidize to form byproducts
that act as antioxidants. This means that although solvent-refined oil will oxidize more rapidly
than hydrocracked oil, it will not necessarily polymerize or thicken as quickly. Solvent-refined oil
generally exhibits gradual increases in viscosity as oxidation is partially suppressed by the natu-
rally occurring antioxidants present in the oil. Hydrocracked oil will initially show virtually no
oxidation or viscosity increase. However, once the oxidation process starts, very rapid viscosity
increases are seen [11].

21.1.3  Antioxidants
There are two general classes of antioxidants:
1. Radical scavengers (RscavH) function by converting alkylperoxy radicals to less-reactive hydro-
peroxides. In this process, the radical scavenger is converted to a stable radical that reacts less
readily with additional oil molecules.
2. Peroxide decomposers (PDs) function by converting unstable hydroperoxides to more stable
alcohols. In this process, the PD is oxidized.
Figure 21.2 provides a generalized mechanism of how these classes work together to inhibit oxida-
tion of a lubricant. A detailed description of the most common antioxidant classes used in lubricants
follows.

RO O . + RScavH RO OH + RScav .

O
PD + ROH RO OH + PD

FIGURE 21.2  General mechanism of antioxidant action. RScavH, radical scavengers (phenolics, diarylamines,
organocopper); PDs, peroxide decomposers (sulfur, phosphites, organomolybdenum).
21-6 Theory and Practice of Lubrication and Tribology

HO CH2 OH HO CH2 OH

CH2 OH
x

x= 0–4

(MBDTBP) (MBBP)

OH
R = CH3 (BHT)
HO R O
R = CH2CH2COR (ABHHC)

(DTBP)

FIGURE 21.3  Examples of hindered phenolic antioxidants.

21.1.3.1  Hindered Phenolics


Examples of typical hindered phenolic antioxidants used in lubricants are shown in Figure 21.3.
4,4′-Methylenebis(2,6-di-tert-butylphenol) (MBDTBP) is a highly effective high temperature–hindered
phenolic antioxidant. Recent studies have shown that this antioxidant is very effective at reducing
deposits in low-phosphorus passenger car engine oils [14]. Other studies have confirmed the ability of
this antioxidant to regenerate alkylated diphenylamines (DPA) in industrial oils at low temperatures,
thus prolonging the efficacy of the total antioxidant system [15].
Other classes of hindered phenolics include butylated hydroxytoluene (BHT), the alkyl-substituted
3,5-di-tert-butyl-4-hydroxyhydrocinnamates (ABHHC), and 2,6-di-tert-butylphenol (DTBP). These gener-
ally function by forming neutral peroxides that can decompose at high temperatures and promote oxidation.
One critical limitation of MBDTBP is its high melting point and limited solubility in severely
hydrocracked oils and PAOs. There are a number of antioxidants available that address this limita-
tion. Examples include the methylene-bridged tert-butylphenolics (MBBP) and the alkyl-substituted
ABHHC. Different versions of ABHHC are available that address specific handling restrictions and
performance needs [16–18].
BHT and DTBP are more volatile than the methylene-bridged and propionate-substituted derivatives
discussed earlier. These products are reserved for applications where volatility or extended lubricant use
is not a major concern. DTBP represents a special class of hindered phenolic antioxidant. Its unique struc-
ture makes it effective at low temperatures. DTBP and BHT have relatively similar volatilities and molecu-
lar structures, yet DTBP performs significantly better than BHT in a number of oxidation bench tests [19].
Recent studies have suggested that greater use of phenolics may be the key to resolving varnish issues
in certain gas turbines [20].

21.1.3.2  Diarylamines
Examples of typical diarylamine antioxidants used in lubricants are shown in Figure 21.4. Two types of
diarylamines are generally used: the alkylated diphenylamines (DPA) and the phenyl-α-naphthylamines
(PANA).
Additive Technology 21-7

NH R R NH R

(Monoalkylated DPA) (Dialkylated DPA)

NH NH R

(PANA) (Alkylated PANA)

FIGURE 21.4  Examples of diarylamine antioxidants.

Most commercial DPAs are mixtures of mono- and dialkylated DPAs. The most common alkyl
groups are octyl, nonyl, mixed butyl/octyl, and mixed styryl/octyl. The performance of these materials
in a variety of oxidation bench tests has recently been studied [21].
The performance properties of the DPAs are attributed to their total nitrogen content, their level of
mono-, di-, and trialkylation, and their type of alkylation. In commercial products, nitrogen contents
typically vary from 3.5 to 4.5 wt%. Alkylation ratios generally vary from 50:50 to 0:100 of monoalkyl-
ated to dialkylated components. Trialkylation is usually in the range of 0–10 wt%. High levels of trial-
kylation are considered detrimental to overall performance. In hydrocracked oils, it is desirable to have
a DPA with high nitrogen content. The DPAs also show very powerful responses to base stock type in
specific oxidation bench tests [22].
The effect of alkylation type on DPA performance has shown that dialkylation improves deposit con-
trol but hampers conventional oxidation control [23]. Recent research on new DPAs has focused on uti-
lizing different olefins, reducing volatility and improving deposit control performance [24,25]. Others
have focused on reducing the amount of undesirable trialkylation [26].
The PANAs are generally more effective antioxidants than the DPAs, especially in hydrocracked base
stocks. The general use of PANA has been limited due to cost and a tendency to cause sludge. The alkyl-
ated PANAs show a lower tendency to form sludge but are considerably more expensive. The most com-
mon use of the PANAs is in specialty applications such as aviation turbine oils and premium industrial
oils [27,28].

21.1.3.3  Sulfurized Antioxidants


Examples of typical sulfurized antioxidants used in lubricants are shown in Figure 21.5. The sul-
furized alkylphenols (SAP) and the sulfurized hindered phenols (SHP) represent examples of mul-
tifunctional antioxidants, having both radical scavenging and peroxide-decomposing capabilities.
Other sulfurized antioxidants include ashless dithiocarbamates (ADTC) and sulfurized olefins and
fats (SO).
Many mechanisms for sulfurized antioxidant function have been proposed [4]. The diversity of
these mechanisms is due to the highly varied chemical structures of these antioxidants and the
many possible oxidation environments to which they are exposed. Recent structure activity stud-
ies have shown that the chemical structure of sulfurized olefins strongly impacts antioxidant
performance [29].
Sulfurized antioxidants are usually characterized by two parameters: the total amount of sulfur and
the amount of active sulfur. The latter parameter is critical in formulating a finished lubricant because
active sulfur can be corrosive and destructive to nitrile seals. Products containing monosulfide, –S–,
bridges are generally considered noncorrosive and seal compatible. As the number of bridging sulfurs
increases, the active sulfur content increases. Different lubricant products have different tolerances for
21-8 Theory and Practice of Lubrication and Tribology

OH OH
S S S
x y
(R)2NCS CH2 SCN (R)2

R R

(SAP) (ADTC)

HO
R
HO (S)x
(S) CH CH
x y
(S)x OH R
y

(SHP) (SO)

FIGURE 21.5  Examples of sulfurized antioxidants.

active sulfur. The corrosive effects of active sulfur are most pronounced when there is significant metal
contact with oxygen and water. For effective antioxidancy, it becomes critical to minimize corrosion and
seal incompatibility effects while maximizing antioxidant performance. This can be done by using low
levels of sulfurized antioxidants and supplementing performance with less aggressive antioxidants such
as hindered phenolics and DPA. Trends toward low SAP engine oils may result in significant restrictions
on sulfurized antioxidant use in the future [30].

21.1.4  Antioxidant Synergism


Synergism exists when the total stabilizing effect of an antioxidant system is greater than the sum of
the effects caused by the individual components of that system. A comprehensive review is available
that effectively addresses the complicated topic of antioxidant synergism [31]. Examples of antioxidant
synergism taking place in lubricants are shown in the following sections.

21.1.4.1  Heterosynergism
Heterosynergism involves the combination of a radical scavenger and a PD as outlined in Figure
21.2. This is the most common antioxidant synergism practiced in lubricating systems. Examples of
radical scavengers that may be used include hindered tert-butylphenolics, alkylated phenolics, dia-
rylamines, and organocopper. Examples of PDs that may be used include sulfurized antioxidants,
phosphorus-containing antioxidants, zinc dialkyldithiophosphates (ZDDPs), and organomolybde-
num compounds.
The most common forms of synergism in this category are DPA and sulfurized antioxidants, DPA
and organomolybdenum compounds, and hindered phenolics and phosphite antioxidants [32–34].

21.1.4.2  Homosynergism
Homosynergism involves the combination of a kinetically fast-reacting radical scavenger, produc-
ing an unstable radical, with a kinetically slow-reacting radical scavenger, producing a very stable
radical. Examples of kinetically fast radical scavengers are diarylamines and less hindered pheno-
lics. The most common examples of kinetically slow radical scavengers are DTBP and its substituted
derivatives.
Additive Technology 21-9

The most common form of homosynergism exists between DPA and hindered phenolics [35]. The
net result in these transformations is the conversion of peroxy and alkoxy radicals into more stable
hydroperoxides and alcohols. Note that the DPA is regenerated at the expense of the hindered phenolic,
which is sacrificed. This synergism provides two benefits. First, it regenerates the faster-acting and more
effective DPA. Second, it suppresses aminic degradation and ultimately improves the oxidation stability
of the lubricant.

21.2  Mechanism of Contamination and Deposit Control


Detergents, that is, materials that contain basic metals, are able to control contaminants such as
acids produced on combustions in both diesel and gasoline engines by neutralization. The detergents
are available in low, medium, and highly overbased versions, and the mechanism of action is the
formation of micelles that solubilize one unit of liquid deposit precursor for each unit of additive
present. This allows them to control combustion acids and prevent rust. The detergents also peptize
the solids in the range of 0–20 nm, and they form films around deposit precursors to which they
are physically adsorbed. By surface charge interactions, these detergents are attracted to particles
in the size range of 50–1500 nm, and this charge mechanism prevents these particles from actually
coagulating and becoming bigger. Detergents come in many varieties, such as sulfonates, phenates,
and salicylates, and with different kinds of metals such as barium, calcium, and magnesium [36].
Recently detergents have been found to have beneficial properties in terms of film forming and fric-
tion control. They form calcium carbonate films (100–150 nm in thickness) on rough surfaces [37].
Thus, in the case of overbased sulfonate detergents, they also act as friction modifiers depending on
the organic side chain structure, while neutralizing acids and cleaning deposits on high-temperature
surfaces.

21.3  Mechanism of Sludge and Varnish Control


21.3.1  Preventing Sludge and Varnish Buildup
Ashless dispersants are used primarily to handle precursors to sludge and varnish at relatively low tem-
peratures by both peptization and solubilization [36]. The most common dispersants are succinimides,
succinate esters, or Mannich functional groups with a poly(isobutylene) tail, structures that are quite
capable of preventing sludge and varnish buildup and solubilizing acids [38–40].
In 1968, Fontana showed that ashless dispersants of the succinimide type are effective for solubilizing
pyruvic acid [41]. Hall followed with a similar conclusion regarding sulfuric acid [42], thus suggesting
that succinimides prevent the products of oxidation from becoming sludge and varnish.
Thus, sludge and varnish buildup are controlled by the following mechanisms:

1. Dispersion (or suspension)—Inhibition of particle–particle aggregation by adsorbed film of ash-


less dispersant, which occurs by either steric or electrostatic mechanism.
2. Solubilization of oil-insoluble liquid into the oil by ashless dispersant.
3. Neutralization by N-containing dispersants—The particles considered may be resin, soot, dirt,
water, any contaminant, as shown in Figure 21.6.

Resin is defined as oligomeric molecules of about 500–700 g/mol. Varnish (not easily wiped off engine
surfaces), lacquer, and deposits are of much higher molecular weight. Sludge can be wiped off the
engine surfaces since it is semisolid in appearance. Deposits are soot in combination with resin.
Lacquer is rich in resin, while soot is carbon rich [43]. Excess soot may result in viscosity increase of
the oil.
21-10 Theory and Practice of Lubrication and Tribology

Partially combusted fuel


Fuel NOx
Liquid nitrated and oxygenated monomers

Oil-in soluble products (resin)


Carbon
Lubricant Heat Water
Solids
Varnish Sludge

Resin Varnish
Accumulate as
Resin-coated deposits in areas
soot particles of low oil velocity
Resin + soot
Soot-coated Lacquer
resin particles

Resin + soot + oil + water Sludge

FIGURE 21.6  Mechanism of deposit formation. (Reprinted from Rizvi, S.Q.A., Additives and additive chem-
istry, in ASTM Manual 37: Fuels and Lubricants Handbook: Technology, Properties, Performance, and Testing,
Totten, G.E., Ed., Copyright ASTM International, West Conshohocken, PA, 2003, pp. 199–248, Chapter 9. With
permission.)

21.4  Mechanism of Wear Control by Antiwear and EP Additives


21.4.1  Antiwear Additives
Tribology concerns the conditions in a contact between two moving parts. Proper lubricating condi-
tions can maintain a piece of equipment for a lifetime. They are aided immensely by specific lubricant
additives. The mechanisms of wear controlled by antiwear (AW) and extreme pressure (EP) additives
will be considered here. In the case of AW agents, the most popular AW agent is the over 50 year old
ZDDP, protective against valve train wear. Recently, the possible reduction in use levels of ZDDP has
been much discussed, because ZDDP has been found to adsorb onto the three-way catalyst system sur-
faces of gasoline engines and thus render them less active [36,43–45], thereby reducing possible fuel
economy improvements.
ZDDP adsorbs onto and reacts with iron surfaces, especially, to reduce wear rates in mixed and
boundary lubrication conditions, but predominantly in the latter. When the ZDDP forms mature AW
films, the two surfaces are protected from direct contact, thus altering wear rates. The Ecole Centrale
de Lyon/Shell Corporation collaboration has produced excellent work on their unique surface force
apparatus to elucidate the mechanism by which ZDDP functions. ZDDP is thought to produce a thin
film of iron sulfide and zinc sulfide nearest the metal surface. Next there is a zinc polyphosphate layer
that is made up of long-chain zinc polyphosphates, and then soluble alkylphosphates are closest to the
oil layer [46].
Using inelastic electron tunneling spectroscopy (IETS), Yamaguchi and Ryason showed that second-
ary ZDDP is adsorbed much more readily than primary ZDDP, while alkaryl ZDDP is hydrolyzed on
adsorption onto aluminum oxide surfaces [47]. More recently, the usefulness of x-ray absorption struc-
ture spectroscopy (XANES) to study ZDDP has been shown [48]. The neutral structure of a ZDDP salt
was shown to be better adsorbed onto a steel surface than the basic version of the same ZDDP salt in tri-
bochemical and thermal films [49]. The length of the polyphosphate chain for the neutral versions ver-
sus basic versions of these model ZDDPs derived from isobutanol, isopropanol, and p-tert-octyl phenol,
respectively, was longer. This suggested that in commercial materials, the neutral salt might be a better
Additive Technology 21-11

wear inhibitor since longer polyphosphate chain lengths had been associated with better wear perfor-
mance. Indeed, this was observed in several Sequence V-E engine experiments in which the second-
ary neutral ZDDP was more effective in wear inhibition than the secondary basic ZDDP: three passes
versus three failures [50]. XANES also showed that the Lyon group’s intuition was correct: ZnS and FeS
have been observed from ZDDP films by XANES [51,52]. The AW films from ZDDPs are effective in
sustaining loads in the pressure range of about 100–350,000 psi. However, the next category of films, EP,
sustains loads at even higher pressures in the range of 200–500,000 psi [36], with no sharp distinction
between the AW and EP films. The ZDDPs for the most part are considered to be AW films and mild
EP films, and efforts to replace this molecule with other AW agents have been unsuccessful up to now.

21.4.2  Additive Interactions and Soot Concerns


There have been many papers written on the interaction of ZDDPs and ashless dispersants on metal
oxide surfaces. Some of the early papers written by Rounds in the mid-1970s came from a set of
Four-Ball experiments using mixtures of ZDDP and dispersant models or dispersants themselves. He
observed an increase in the amount of ZDDP necessary to produce a small wear scar diameter with
dispersants and/or model molecules [53,54]. Other researchers were able to observe similar results but
with different types of bench/engine tests, suggesting that dispersants/ZDDP interactions are not good
for wear.
Using the mono- or bis-succinimides, Yin et al. [55,56] in 1997 showed that the dispersants do have
an effect on ZDDP film formation, and in fact the ZDDP films are altered; that is, the normal AW poly-
phosphate film was not present in the tribofilm according to XANES analysis and was instead shortened
considerably. Shorter polyphosphates have been associated with poorer wear performance.
For the effect of soot in the presence of ZDDPs and polybutene succinimide dispersants, Gautam et al.
[57] reported that increasing wear rates were observed for soot-laden oils for various engine components
with the mixture of dispersant and ZDDPs. These experimentalists used engine soot to observe this
increase in wear. On the other hand, Rounds mostly used carbon black as a soot surrogate [58].
These works and others [59,60] indicate that an oil containing additives such as dispersant, detergent,
VI improver, and ZDDP provides a more complex situation than an oil with ZDDP alone or in a binary
mixture.

21.4.3  EP Additives
21.4.3.1  Sulfur-Containing EP Additives
EP additives are another group of boundary lubricants that can be used in engine oil formulations for
specific applications, such as gears, which are under high load conditions. Sulfur-containing additives
are probably the oldest (>100 years) known EP additives, and the mechanism by which they work con-
sists of breaking disulfide or better, polysulfide linkages, which are less stable than the monosulfide
bonds. The polysulfide materials, sulfurized olefins, and esters and fats can bond to the metal surface:
the sulfur bonds to iron, and iron sulfide is quickly produced from Fe-based surfaces. Iron sulfide gives
a low friction coefficient and is also able to carry high loads, such as from sulfurized isobutylene [61–65].
Other examples of sulfurized EP additives are the sulfurized fats or esters, xanthates, thiocarbonates,
and dithiocarbamates [36]. However, all EP agents have limits as indicated by Figure 21.7 [66].
Molybdenum dithiocarbamates (MoDTC) show good AW and EP protection and are also friction
modifiers, so they offer multifunctional behavior. Many papers have been written about the production
of molybdenum disulfide in the contact, both from the friction point of view and an EP wear point of
view [67–69]. In 1998, Grossiord et al. proposed a detailed mechanism of the formation of molybdenum
disulfide, from MoDTC in a tribometer combined with an XPS spectrometer under UHV [70].
However, there is other bench test evidence from Morina et al. [71] and Yamaguchi and Ruby [72] that
prolonged rubbing at high loads produces a higher friction due to the formation of MoO3 at the surface.
21-12 Theory and Practice of Lubrication and Tribology

I
0.5

Coefficient of friction
0.4
II
0.3

0.2

III
0.1
Tr
0
Temperature

FIGURE 21.7  EP action of additive (III) with threshold reaction temperature, Tr, as compared with base
oil (I) and fatty oil additive (II). Arrow represents scuffing. (Extracted from Sethuramiah, A., Lubricated wear:
Science and technology, in Tribology Series, Vol. 42, Dowson, D., Ed., Elsevier Science Publishers B.V., Amsterdam,
the Netherlands, 2003, p. 163. Used with permission.)

21.4.3.2  Phosphorus-Containing EP Additives


Phillips [73] has discussed phosphorus-containing additives without metals present. Some of the struc-
tures studied are shown in Figure 21.8. By conducting various bench tests and engine tests on these
materials, a few generalizations have been made (Figure 21.9).
One of the most tested compounds is tricresyl phosphate (TCP) [74–76]. Barcroft and Daniel [77]
showed that iron phosphides, which had been earlier proposed by Beeck et al. [74], did not describe
the mechanism exactly, since no evidence of iron phosphide was observed in a cam and tappet lubri-
cation contact using a Ford Consul engine scuffing test and TCP-containing lubricant. Instead, based
on adsorption experiments, they suggested that the ester initially adsorbs onto the metal surface
and decomposes to give acid phosphates, as proposed earlier by Godfrey [75], which then reacts
further to produce metal organic phosphates and then decomposes to metal phosphates such as iron
phosphates.
Davey studied primarily phosphates and phosphites [78]. He concluded that phosphites are superior
to phosphates as EP additives and also that the alkyl esters are superior to aryl esters using the Four-Ball
EP test in 1950.

21.4.3.3  Chlorine-Containing Agents


Chlorine-containing additives have been used as EP agents for the metalworking industry, and under
high loads, seizure has been prevented [79]. A correlation between chemical reactivity and load-carry-
ing capacity was worked out by Sakurai and Sato [76]. Today, the negative health and environmental
aspects of these chlorine-containing materials have resulted in declining use. However, recently several
fundamental surface studies have been accomplished by Tysoe and coworkers [80] using state-of-the-art
equipment. The main conclusion from his studies is that chlorinated EP additives produce iron chloride
on the rubbing surface, and this material is very effective at preventing direct metal-to-metal contact
and therefore seizure up to its melting point (950°K).

21.5  Mechanism of Corrosion Control


21.5.1  Corrosion Inhibitors
Corrosion inhibitors are used in various lubricant applications to protect against two types of corro-
sion. One is iron corrosion, or rust, and the other type is the corrosion of yellow metals like copper.
Additive Technology 21-13

O R O R

RO P OR O P O

OR O

R
Trialkyl phosphate Triaryl phosphate

O R O

RO P OH O P OH

OR OH
Alkly monoacid phosphate Aryl diacid phosphate

RO P OR O P O

OR O

Trialkyl phosphite Triaryl phosphite

RO P OH RO P R
OR OR

Dialkyl phosphite Dialkyl alkyl phosphonate

O O

RO P OH.H2NR RO P R

OR R
Amine phosphite Alkyl dialkyl phosphinate

FIGURE 21.8  Structures of common lubricating oil additives containing phosphorus. (Extracted from Phillips,
W.D., Ashless phosphorus-containing lubricating oil additives, in Lubricant Additives, Rudnick, L.R., Ed., Marcel
Dekker, New York, 2003, pp. 45–112, Chapter 3. Used with permission.)

Amine phosphites
Amine phosphates
Acid phosphites
Improvement in Improvement in Impact on
Acid phosphates
EP properties AW properties stability, etc.
Neutral phosphites
Neutral phosphates
Neutral phosphonates

FIGURE 21.9  Approximate ranking of the effect of structure on antiwear, extreme pressure, and stability prop-
erties of the base stock. (Extracted from Phillips, W.D., Ashless phosphorus-containing lubricating oil additives,
in Lubricant Additives, Rudnick, L.R., Ed., Marcel Dekker, New York, 2003, pp. 45–112, Chapter 3. Used with
permission.)
21-14 Theory and Practice of Lubrication and Tribology

Problem
O2 O2 Lubricant

Part A Fe2+
H2O
Rust
C e A e C Iron
metal

Cathodic reaction (C)


O2 + 2H2O + 4e 4OH–

Anodic reaction (A)


Fe Fe2+ + 2e

Solution

O2 O2 Lubricant
Part B
H2O
Protective
layer Iron
metal

FIGURE 21.10  Mechanism of rust inhibition. (Reprinted from Rizvi, S.Q.A., Additives and additive chemistry, in
ASTM Manual 37: Fuels and Lubricants Handbook: Technology, Properties, Performance, and Testing, Totten, G.E.,
Ed., Copyright ASTM International, West Conshohocken, PA, 2003, pp. 199–248, Chapter 9. With permission.)

The mechanism of iron corrosion involves water, oxygen, and electrolytes. It occurs because the iron
surface loses electrons and this gives rise to Fe2+. These Fe2+ ions then react with water to form OH− ions
and then eventually form ferrous hydroxide that is further oxidized to ferric hydroxide, which then
loses water and gives rise to rust (hydrated Fe2O3). A corrosion inhibitor would provide a protective
layer, as shown in Figure 21.10 [81]. Because engines are made predominantly out of steel, the concern
for rust must be taken seriously, since the presence of acids or bases will increase the rate of reaction of
rust formation. This requires employing a balanced formulation approach using various additives such
as polymethoxylated alkylphenols, sulfonates (both neutral and basic), the succinimide additives, and
phosphorus-containing additives such as alkylphosphites and phosphates and polymeric amines and
alkenolics [39].
For yellow metal corrosion, the type of corrosion inhibitor needed is different. This is because the
copper metal, for example, needs to be protected by a copper oxide film and then a deactivator film
needs to adsorb onto the surface and interact, giving rise to a very cohesive film that does not allow
the acids or water to penetrate. Luo et al. [82] showed that mercaptothiadiazole bisdisulfide derivatives
protect copper by reacting with active S to give product trisulfides, shown by Raman spectroscopy and
mass spectroscopy. These components were shown to inhibit copper corrosion by scavenging the active
sulfur and forming a passivating film that contains a thiadiazole ring on the copper surface. The mol-
ecules used to protect against copper corrosion are quite different from those listed for iron corrosion.
Additive Technology 21-15

21.6  Mechanism of Friction Modifiers


Friction modifiers or friction reducers can be described as chemicals that reduce the coefficient of
friction. These have been known since 1918 when Wells and Southcombe [83] showed the mysterious
property of “oiliness” when tiny amounts of fatty acids were added to the mineral oils. Since then they
have been applied to gear oils, automatic transmission fluids, tractor fluids, automobile lubricants, etc.
[83–86].
Good friction reduction and hence fuel efficiency can be achieved by two mechanisms: first, by choos-
ing a low-viscosity engine oil, thus depending on the property of the base oil. This is generally of choice
wherein an engine operates mostly in elastohydrodynamic regime (fluid lubrication) such as in bear-
ings. Modern engine oils have this feature now built into them. It is important in choosing the base oils,
to choose one with a low-pressure viscosity coefficient, since with a low-pressure viscosity coefficient, α,
there will be low friction in the contact, as long as the surface roughness to film thickness ratio remains
>3. Other features that must be present are low kinematic viscosity, high viscosity index, low “high tem-
perature high shear” (HTHS) viscosity; although it should be noted that one often has to adjust additive
formulations to enable the use of lower-viscosity base oils and still provide acceptable wear, oxidation,
and deposit control.
The second means of reducing friction is by using a friction-modifying additive, that is, friction mod-
ifier. The friction-modifying agents for automotive engine oils are meant to reduce friction in both the
mixed and the boundary lubrication regimes. Generally, friction modifiers work with either of these
mechanisms: physisorption (i.e., formation of absorbed layers) or chemisorption (i.e., formation of
reacted layers) (Figure 21.11). They may further act in a multilayer mode, where these form thick films
due to series of monolayers, such as in liquid crystal [85]. Friction modifiers most commonly used are

(A) (B) (C)

van der waals


interactions

Non
polar
head

Dipole–dipole
Polar interactions
head

Hydrogen bonding Covalent bonding

Surface

FIGURE 21.11  Friction modifier mechanisms: (A) absence of polar head results in randomly oriented molecules,
(B) physisorption, and (C) chemisorption.
21-16 Theory and Practice of Lubrication and Tribology

long-chain carboxylic acid derivatives, especially long-chain amides, organic partial esters (glycerol
monooleate, GLYMO), organic polymers, and phosphoric and phosphonic acid derivatives. Examples
of metal-containing friction modifiers include molybdenum dithiophosphate (MoDTP), MoDTC, etc.
[87–89]. Although, due to the ideal reduction of phosphorus in engine oils, phosphorous-containing
materials such as MoDTP and phosphoric or phosphonic acid derivatives are not preferred.
Broadly speaking, there are two mechanisms on how friction modifiers work: physisorption and
chemisorption.

21.6.1  Physisorption: Formation of Absorbed Layers


Friction modifiers form absorbed layers through a weak reversible adsorption process based on dipole or
van der Waal forces. This occurs due to the polar group attached to the molecule that gets attracted to the
metal surface. As a result, these molecules align among themselves (Figure 21.11B) [85]. No chemical bonds
are made. Some of the parameters that affect the frictional properties of a friction modifier include
• Polar group—Polarity is one of the important criteria for an effective friction modification. The
polar group should be able to hydrogen bond with the surface and be a hydrogen donor. For
example, nitroparaffins, which is a hydrogen bond acceptor, do not act as friction modifiers. A
subtle balance is required regarding polarity as it can compete with other surface agents such as
detergents [89].
• Chain length—Longer chains can increase the thickness of the absorbed films and hence can
affect friction [86].
• Molecular configuration—Structural properties such as straight chains can be preferred since
they are able to pack more closely. This should influence the nature of the film and frictional
properties.
• Temperature—Increasing the temperatures leads to the disturbance of the absorbed species that
might detach from the surface and affect the friction. Chemisorbed additives also do ultimately
detach with an increase in temperature.

21.6.2  Chemisorption: Formation of Reacted Layers


On the other hand, friction modifiers may react with the surface irreversibly and may involve ionic
interactions and/or make covalent bonds (Figure 21.11C) [90].
One example of this is the MoDTC and its interactions with other additives. The mechanism of action
with respect to the MoDTC is related to its ability to form molybdenum disulfide sheets on the surface
as shown in Figure 21.12 [79,88,89]. When force is applied, these sheets slip, like a deck of cards, hence
reduce friction. Friction modifiers can show some synergy with AW additives. One example is the use
of MoDTC with ZDDP. The mixture often shows synergy in wear and frictional properties [89,90].
Interestingly, Iwasaki showed that about 40% of the sulfur in MoS2 formed on the rubbing surface was
generated from ZDDP, by labeling with the 34S isotope and use of time-of-flight secondary-ion mass
spectrometry [91]. Recent work has shown that these succinimide posttreated dispersants, in the pres-
ence of ZDDP, also produced molybdenum disulfide in the contact, which gives rise to the low fric-
tion [92,93], and low wear in a bench test. Korcek et al. have shown that the MoDTC functions by an

Applied
Load (L) force Load (L)

FIGURE 21.12  Sliding of layers on applying force.


Additive Technology 21-17

exchange of functional groups with the ZDDP [94]. Morina et al. studied the effects of temperature and
ZDDP/MoDTC ratio on MoS2 film formation under boundary lubrication conditions on 52,100 steel
surfaces [71]. Using both energy-dispersive x-ray analysis and XPS, the authors found that the ratio of
MoS2/MoO3 in the wear scar governs the decrease in friction when using MoDTC additive. The ratio
was dependent on the temperature and amount of ZDDP in the original oil. However, the mechanism
of friction reduction based on the interaction of the two additives is still not fully understood. Friction
modifiers, in general, should also not promote rust, corrosion, oxidation, deposit formation, or wear.
According to some original equipment manufacturers (OEM), fuel savings of 3%–4% are achievable
today using various friction modifiers and lower-viscosity oils.
One other mechanism that has also been proposed to reduce friction is called “slip at the boundary”
[95]. Spikes and his colleagues have shown that an adsorbed friction modifier can promote slip of the
liquid lubricant at the stationary side [96].

21.7  Mechanism of Viscosity Index Improver


Viscosity is quite arguably the most important physical property of a lubricant. The lubricant must be
able to form a fluid film between moving parts of an engine, in order to seal combustion pressures. It
also controls engine friction and oil circulation, which influences oil consumption, oil leakage, engine
noise, engine coolant, fuel consumption, power output, starting, and warm up. Especially, at higher
temperatures, the film-forming properties diminish due to a drop in viscosity. Prior to usage of viscosity
index improvers (VIIs), the problem was overcome by seasonal oil changes. In winter low-viscosity oils
were used, and in summer high-viscosity oils were used. The invention of VII or viscosity modifiers led
to the introduction of multigrade oil and relieved the owner from seasonal oil changes. The effectiveness
of a VII is demonstrated in Figure 21.13.
VIIs are polymers such as olefin copolymer (OCP) types, polymethacrylates (PMAs), hydrogenated
styrene dienes (STDs), etc., that cause minimal viscosity increase at low temperature but considerable

103

5
SAE 30
3

102 SAE 10W-30


Kinematic viscosity (mm2/s)

2
SAE 10W
101

–20 0 20 40 60 80 100 120


Temperature (°C)

FIGURE 21.13  Viscosity–temperature characteristics of single grade and multigrade oil. (Reprinted from Rizvi,
S.Q.A., Additives and additive chemistry, in ASTM Manual 37: Fuels and Lubricants Handbook: Technology,
Properties, Performance, and Testing, Totten, G.E., Ed., Copyright ASTM International, West Conshohocken, PA,
2003, pp. 199–248, Chapter 9. With permission.)
21-18 Theory and Practice of Lubrication and Tribology

10,000
Base oil and
1,000 viscosity index improver

Viscosity, SUS
100
Base oil

40
0 100 210
Temperature, °F

FIGURE 21.14  Effects of VI improvers (VII) on temperature.

increase at high temperature. In other words, they minimize viscosity change with a change in temper-
ature. Thus, VII affects the change in viscosity with temperature and thereby increase the VI of the oil
(Figure 21.14). Base oil itself has high viscosity at low temperatures and thus the movement of the mol-
ecules is slow (see left curve and schematics in Figure 21.14). For the case of base oil + VII (Figure 21.14),
the VII molecules do not significantly add to the viscosity of the base oils at low temperatures as com-
pared to high temperatures. At high temperatures (right-hand side of Figure 21.14), the base oil mol-
ecules have higher kinetic energy and start to move around fast. However such movement of the base oil
molecules is affected by the presence of the VII, resulting in higher oil viscosity at high temperatures.
Thus for a given increase in temperature, the base oil + VII case improves the VI by thickening the oil
at higher temperatures as compared to the base oil only case. This is a practical means of extending the
operating range of oils to higher temperatures without adversely affecting their low-temperature fluid-
ity. Figure 21.15 illustrates the mechanism of oil thickening by viscosity modifiers. Excellent reviews
have been written about VIIs [81,97,98].
Some more studies have been carried out to determine other features of VII. From the initial obser-
vation of friction and wear reduction via boundary film formation from a PMA VII [99], Spikes and
collaborators did a series of elegant experiments regarding structure/performance relationships using
the ultrathin film interferometric approach. Müller et al. synthesized a series of functionalized poly-
alkylmethacrylates (PAMAs) and studied their friction and wear properties [100]. Although these
materials are primarily used as VIIs, the authors found that these materials can adsorb onto the sur-
faces and form thick boundary films in a rolling/sliding high pressure lubricated contact, reducing

Oil associated
with polymer

Polymer
Solubility molecule
Poor Good

Temperature
Low High

FIGURE 21.15  Mechanism of oil thickening by viscosity modifiers. (Reprinted from Rizvi, S.Q.A., Additives and
additive chemistry, in ASTM Manual 37: Fuels and Lubricants Handbook: Technology, Properties, Performance, and
Testing, Totten, G.E., Ed., Copyright ASTM International, West Conshohocken, PA, 2003, pp. 199–248, Chapter 9.
With permission.)
Additive Technology 21-19

TABLE 21.1  Viscosity Index Improver Types


General Class Dispersant Type Available Monomers Used
Hydrocarbon type
PIB (polyisobutylene) No Isobutylene
OCP (olefin copolymer) Yes Ethylene and propylene
SB or SI (olefin copolymers) No Styrene with butadiene or isoprene
Ester type
Styrene ester Yes Styrene and maleic ester
Polymethacrylate Yes Methacrylic acid and alcohol

friction as well as controlling wear. The test methods used were a high-frequency reciprocating rig
(HFRR) and the mini-traction machine (MTM). The MTM was also used in the bidirectional mode
in conjunction with inductively coupled plasma atomic emission spectroscopy (ICP-AES). The latter
monitors wear.
The newest trend in VIIs is for extra functionality to be built into these materials such as dispersant
functionality to the VIIs. Dispersant viscosity modifiers (DVMs) are made either by grafting or through
copolymerization to include basic nitrogen or surfactant-type oxygen-containing monomers. Examples
of DVMs are listed in Table 21.1.

21.8  Mechanism of Foam Inhibitors


Foam can form when air gets entrapped in the liquid. In every lubricant application, there is the splash-
ing action and/or the mechanical agitation of the crankcase oil that allows air and other vapors to be
whipped around, generating foam from the oil. In extreme cases, the oil actually can be displaced by the
foam. The entrainment of the air in the oil can also decrease the ability of the oil to provide an effective
hydrodynamic lubricating film because of the air bubbles that compromise the integrity of the film.
Viscosity and surface tension of the lubricant determine the stability of foam. Low-viscosity oils tend to
make large bubbles that break quickly, whereas high-viscosity oils generate stable foams that are difficult
to break [36,81].
Foam inhibitors function by altering the surface tension of the oil and by facilitating the separation of
the air bubbles from the oil phase. They are added in small amounts (10–200 ppm) as they have limited
solubility in oil. The control of foam in modern-day engines is accomplished by using silicon-based
foam inhibitors, particularly the polymethylsiloxanes. However, adding too much foam inhibitor leads
to foaming again, so it is important to maintain the optimum concentration.

21.9  Conclusions and Future Needs


The authors have attempted to provide the mechanisms of the performance of various additive classes.
In considering these mechanisms, one cannot help but look to the future engine oil development. It
appears that engine oil development will be influenced by many factors in the future; one of the factors
being the oil compatibility with the system such as after-treatment systems. Second, the environment
will play a big role since environmental concerns will require the use of less toxic raw materials. In
recent years, the importance of green chemistry initiatives, biodegradability, and minimal environ-
mental footprints have placed great demands on the performance of next generation lubricants. Finally,
the issue of sustainability will influence engine oil development going forward. Low-sulfur, phospho-
rus, and ash oils will continue to grow in popularity, as will biodegradable fluids and synthetic fluids
based on renewable raw materials. The combination of higher performance and minimal environmental
impact will lead to the development of new classes of additives.
21-20 Theory and Practice of Lubrication and Tribology

References
1. Scott, G., Atmospheric Oxidation and Antioxidants, Elsevier Science Publishers B.V., Amsterdam, the
Netherlands (1993).
2. Rasberger, M., Oxidative degradation and stabilization of mineral oil based lubricants, in Chemistry
and Technology of Lubricants (Mortier, R. M. and Orszulik, S. T., Eds.), Chapman & Hall, London,
U.K., pp. 98–143 (1997).
3. Migdal, C. A., Antioxidants, in Lubricant Additives Chemistry and Applications (Rudnick, L. R., Ed.),
Marcel Dekker, Inc., New York, pp. 1–28 (2003).
4. Al-Malaika, S., Antioxidants—Preventive mechanisms, in Atmospheric Oxidation and Antioxidants
Volume I (Scott, G., Ed.), Elsevier Science Publishers B.V., Amsterdam, the Netherlands, pp. 161–224
(1993).
5. Uri, N., Mechanism of autoxidation, in Autoxidation and Antioxidants Volume I (Lundberg, W. O.,
Ed.), John Wiley & Sons, New York, pp. 133–170 (1961).
6. Osawa, Z., Metal catalyzed oxidation and its inhibition, in Atmospheric Oxidation and Antioxidants
Volume II (G. Scott, Ed.), Elsevier Science Publishers B.V., Amsterdam, the Netherlands, pp. 327–362
(1993).
7. Sharp, D. B., Patton, L. W., and Whitcomb, S. E., Autoxidation of ketones. I. Diisopropyl ketone,
Journal of the American Chemical Society, 73, 5600–5603 (1951).
8. Maslov, S. A. and Blyumberg, E. A., Liquid-phase oxidation of aldehydes, Russian Chemical Reviews,
45(2), 303–328 (1976).
9. March, J., The aldol condensation, in Advanced Organic Chemistry, 3rd edn., John Wiley & Sons, Inc.,
New York, pp. 829–834 (1985).
10. Henderson, H. E., Advanced performance products from VHVI specialty base fluids, in Symposium
on Worldwide Perspectives on the Manufacture, Characterization and Application of Lubricant Base
Oils: IV, 218th Annual Meeting Preprint, American Chemical Society, New Orleans, LA, pp. 251–256
(August 22–26, 1999).
11. Igarashi, J., Yoshida, T., and Watanabe, H., Concept of optimal aromaticity in base oil oxidative stabil-
ity revisited, in 213th National Meeting of the American Chemical Society, Annual Meeting Preprint,
pp. 211–217 (1997).
12. Yoshida, T., Watanabe, H., and Igarashi, J., Pro-oxidant properties of basic nitrogen components
in base oil, in Industrial and Automotive Lubrication, I, Technische Akademie Esslingen, 11th
International Colloquium, pp. 433–444 (1998).
13. Stipanovic, A. J., Hydrocarbon base oil chemistry, in Synthetics, Mineral Oils, and Bio-Based
Lubricants—Chemistry and Technology (Rudnick, L. R., Ed.), CRC Press, Taylor & Francis Group,
Boca Raton, FL, pp. 169–197 (2006).
14. Moehle, W. E., Cobb, T. W., Schneller, E. R., and Gatto, V. J., Utilizing the TEOST MHT to evaluate fun-
damental oxidation processes in low phosphorus engine oils, Tribology Transactions, 50, 96–103 (2007).
15. Gatto, V. J., Moehle, W. E., Cobb, T. W., and Schneller, E. R., The relationship between oxidation sta-
bility and antioxidant depletion in turbine oils formulations with group II, III and IV base stocks,
Journal of Synthetic Lubrication, 24(2), 111–124 (2007).
16. Kristen, U., Muller, K., and Rasberger, M., Stabilized diesel engine oil, US Patent 5,523,007 (1996).
17. Gatto, V. J., Elnagar, H. Y., Cheng, C. H., and Adams, J. R., Preparation of sterically hindered hydroxy-
phenylcarboxylic acid esters, US Patent 7,667,066 (2010).
18. Abraham, W. D., Adams, P. E., Lamb, G. D., Denis, R. A., Kocsis, J. A., Roski, J. P., Carrick, V. A., and
Cowling, S. V., Lubricant compositions containing ester-substituted hindered phenol antioxidants,
US Patent 6,559,105 (2003).
19. Krishnamoorthy, P. R., Vijayakumari, S., and Sankaralingam, S., Effect of antioxidants and metal
deactivator on the oxidation of transformer oil, IEEE Transactions on Electrical Insulation, 27(2),
271–277 (1992).
Additive Technology 21-21

20. Livingstone, G. J., Thompson, B. T., and Okazaki, M. E., Physical, performance and chemical changes
in turbine oils from oxidation, in ASTM Symposium on the Oxidation and Testing of Turbine Oils,
Norfolk, VA (December 5, 2005).
21. Gatto, V. J., Elnagar, H. Y., Moehle, W. E., and Schneller, E. R., Redesigning alkylated diphenylamine
antioxidants for modern lubricants, Lubrication Science, 19, 25–40 (2007).
22. Migdal, C. A. and Abbott, R. D., The utilization of pressure differential scanning calorimetry as an
alternative to the Rotary Bomb Oxidation Test to measure the oxidative stability of lubricating oils
in the presence of antioxidants, in Industrial and Automotive Lubrication I, Technische Akademie
Esslingen, 11th International Colloquium, pp. 1967–1974 (1998).
23. Gatto, V. J., Grina, R. A., Tat, T. L., and Ryan, H. T., The influence of chemical structure on the physical
properties and antioxidant response of hydrocracked base stocks and polyalphaolefins, Journal of
Synthetic Lubrication, 19, 1–18 (2002).
24. Onopchenko, A., Alkylation of diphenylamine with polyisobutylene oligomers, US Patent 6,355,839
(2002).
25. Milne, N. J., Arrowsmith, S., Burrows, A. L., Wilson, P., Booth, C. J., Emert, J., and Chambard, L.,
Additives and lubricating oil compositions containing same, PCT Application WO 154334
(2008).
26. Aebli, B. M., Evans, S., and Gati, S., Nonylated diphenylamines, US Patent 6,315,925 (2001).
27. Yaffe, R., Synthetic aircraft turbine oil, US Patent 4,248,721 (1981).
28. Yano, A., Watanabe, S., Miyazaki, Y., Tsuchiya, M., and Yamamoto, Y., Study on sludge formation dur-
ing the oxidation process of turbine oils, Tribology Transactions, 47, 111–122 (2004).
29. Bala, V. and Hartley, R. J., The influence of chemical structure on the oxidative stability of organic
sulfides, Lubrication Engineering, 52(12), 868–873 (1996).
30. Woydt, M., No/low SAP and alternative engine oil development and testing, Journal of ASTM
International, 4(10), 1–13 (2007).
31. Scott, G., Synergism and antagonism, in Atmospheric Oxidation and Antioxidants, II (Scott, G., Ed.),
Elsevier Science Publishers B.V., Amsterdam, the Netherlands, pp. 431–459 (1993).
32. White, W. R. and Reale, J., Low ash, low phosphorus motor oil formulations, US Patent 4,330,420
(1982).
33. Gatto, V. J. and Devlin, M. T., Lubricant containing molybdenum compound and secondary diaryl-
amine, US Patent 5,650,381 (1997).
34. Cohen, S. C., Lubricating oil compositions containing novel combination of stabilizers, US Patent
4,652,385 (1987).
35. Gatto, V. J. and Grina, M. A., Effects of base oil type, oxidation test conditions and phenolic anti-
oxidant structure on the detection and magnitude of hindered phenol/diphenylamine synergism,
Lubrication Engineering, 55(1), 11–20 (1999).
36. Roby, S. H., Yamaguchi, E. S., and Ryason, P. R., Lubricant additives for mineral oil based hydraulic
fluids, Chapter 15 in Handbook of Hydraulic Fluid Technology (Totten, G. E., Ed.), Marcel Dekker,
New York, pp. 795–823 (2000).
37. Topolovec-Miklozic, K., Forbus, T. R., and Spikes, H., Film forming and friction properties of over-
based calcium sulphonate detergents, Tribology Letters, 29, 33–44 (2008).
38. Rizvi, S. Q. A., Dispersants, Chapter 5 in Lubricant Additives (Rudnick, L. R., Ed.), Marcel Dekker,
New York, pp. 137–170 (2003).
39. Rizvi, S. Q. A., Additives and additive chemistry, in Fuels and Lubricants Handbook: Technology,
Properties, Performance, and Testing (Totten, G. E., Ed.), ASTM International, West Conshohocken,
PA, pp. 199–248, Chapter 9 (2003).
40. Murakami, Y. and Hisamoto, A., Effects of NOx and unburned gasoline on low temperature sludge
formation in engine oil, SAE Paper 910747 (1991).
41. Fontana, B. J., Interaction of a polymeric detergent with surfactant micelles in hydrocarbon media,
Macromolecules, 1, 139–145 (1968).
21-22 Theory and Practice of Lubrication and Tribology

42. Hall, K. L., Iron corrosion by sulfuric acid solubilized in oil, presented at Symposium on Deposit, Wear,
and Emission Control by Lubricant and Fuel Additives, Division of Petroleum Chemistry, American
Chemical Society, New York, pp. A93–A101 (September 7–12, 1969).
43. Canter, N., Special report: Additive challenges in meeting new automotive engine specifications,
Tribology and Lubrication Technology, 10–19 (September 2006).
44. Papay, A. G., Antiwear and extreme pressure additives in lubricants, Lubrication Science, 10(3),
209–224 (1998).
45. McDonald, R. A., Zinc dithiophosphates, Chapter 3 in Lubricant Additives (Rudnick, L. R., Ed.),
Marcel Dekker, New York, pp. 29–43 (2003).
46. Tonck, A., Bec, S., Georges, J.-M., Coy, R. C., Bell, J. C., and Roper, G. W., The role of the interface and
surface layers in the thin film and boundary regime, in Proceedings of the 25th Leeds-Lyon Symposium
on Tribology, Lubrication at the Frontier (Dowson, D. et al., Ed.), Elsevier Science Publishers B.V.,
Amsterdam, the Netherlands, pp. 39–47 (1999).
47. Yamaguchi, E. S. and Ryason, P. R., Inelastic electron tunneling spectra of lubricant oil additives on
native aluminum oxide surfaces, Tribology Transactions, 36(3), 367–374 (1993).
48. Pawlak, Z., Tribochemistry of lubricating oils, in Tribology and Interface Engineering Series, No. 45
(Briscoe, B. J., Ed.), Elsevier Science Publishers B.V., Amsterdam, the Netherlands, pp. 1–214 (2003).
49. Fuller, M., Yin, Z., Kasrai, M., Bancroft, G. M., Yamaguchi, E. S., Ryason, P. R., Willermet, P. A., and
Tan, K. H., Chemical characterization of tribochemical and thermal films generated from neutral
and basic ZDDPs using x-ray absorption spectroscopy, Tribology International, 30(4), 305–315
(1997).
50. Yamaguchi, E. S., The relative wear performance of neutral and basic ZnDTPs in engines, Tribology
Transactions, 42(1), 90–95 (1999).
51. Kasrai, M., Fuller, M. S., Bancroft, G. M., Yamaguchi, E. S., and Ryason, P. R., X-ray absorption study
of the effect of calcium sulfonate on antiwear film formation generated from neutral and basic
ZDDPs, Part 2. Sulfur species, Tribology Transactions, 46(4), 543–549 (2003).
52. Zhang, Z., Yamaguchi, E. S., Kasrai, M., Bancroft, G. M., Liu, X., and Fleet, M. E., Tribofilms generated
From ZDDP and DDP on steel surfaces, Part 2, chemistry, Tribology Letters, 19(3), 221–229 (2005).
53. Rounds, F. G., Additive interactions and their effect on the performance of a zinc dialkyl dithiophos-
phate, ASLE Transactions, 21, 91–101 (1976).
54. Rounds, F. G., Some factors affecting the decomposition of three commercial zinc organodithio-
phosphates, ASLE Transactions, 18, 79–89 (1975).
55. Yin, Z., Kasrai, M., Fuller, M., Bancroft, G. M., Fyfe, K., and Tan, K. H., Application of soft x-ray
absorption spectroscopy in chemical characterization of antiwear films generated by ZDDP. Part I:
The effects of physical parameters, Wear, 202, 172–191 (1997).
56. Yin, Z., Kasrai, M., Bancroft, G. M., Fyfe, K., Colaianni, M. L., and Tan, K. H., Application of soft x-ray
absorption spectroscopy in chemical characterization of antiwear films generated by ZDDP. Part I:
The effect of detergents and dispersants, Wear, 202, 192–201 (1997).
57. Gautam, M., Durbaha, M., Chitoor, K., Jaraiedi, M., Mariwalla, N., and Ripple, D., Contribution of
soot contaminated oils to wear, SAE Paper 981406 (1998).
58. Rounds, F. G., Carbon: Cause of diesel engine wear? SAE Paper 770829 (1977).
59. Yamaguchi, E. S., Untermann, M., Roby, S. H., Ryason, P. R., and Yeh, S. W., Soot wear in diesel
engines, Journal of Engineering Tribology, Proceedings of the Institution of Mechanical Engineers Part
J, 220(J5), 463–469 (2006).
60. Devlin, M. T., Li, S., Burgess, T., and Jao, T.-C., Film formation properties of polymers in the presence
of abrasive contaminants, SAE Paper 2002-21-2793 (2002).
61. Forbes, E. S., The load-carrying action of organo-sulphur compounds—A review, Wear, 15, 87–96
(1970).
62. Forbes, E. S., Antiwear and extreme pressure additives for lubricants, Tribology, 145–152 (August
1970).
Additive Technology 21-23

63. Sellei, H., Sulfurized extreme-pressure lubricants and cutting oils, Petroleum Processing, 1003–1008
(1949).
64. Farng, L. O., Ashless antiwear and extreme-pressure additives, Chapter 8 in Lubricant Additives
(Rudnick, L. R., Ed.), Marcel Dekker, New York, pp. 223–258 (2003).
65. Braun, J., Additives, Chapter 6 in Lubricants and Lubrication (Mang, T. and Dresel, W., Eds.),
Wiley-VCH, Weinheim, Germany, pp. 88–118 (2007).
66. Sethuramiah, A., Lubricated wear: Science and technology, in Tribology Series, Vol. 42 (Dowson, D.,
Ed.), Elsevier Science Publishers B.V., Amsterdam, the Netherlands, p. 163 (2003).
67. Yamamoto, Y. and Gondo, S., Frictional characteristics of molybdenum dithiophosphates, Wear, 112,
79–87 (1986).
68. Martin, J.-M., Donnet, C., and Le Mogne, Th., Superlubricity of molybdenum disulphide, Physical
Review B, 48(14), 10583–10586 (1993).
69. Jensen, R. K., Johnson, M. D., Korcek, S., and Rokosz, M. J., Friction reducing and antioxidant
capabilities of engine oil additive systems under oxidative conditions. I. Effects of ligand exchange
between molybdenum dialkyldithiophosphate in hexadecane, Lubrication Science, 10(2), 99–120
(1998).
70. Grossiord, C., Varlot, K., Martin, J.-M., Le Mogne, Th., Esnouf, C., and Inoue, K., MoS2 single sheet
lubrication by molybdenum dithiocarbamate, Tribology International, 31(12), 737–743 (1998).
71. Morina, A., Neville, A., Priest, M., and Green, J. H., ZDDP and MoDTC interaction in bound-
ary lubrication—The effect of temperature and ZDDP/MoDTC ratio, Tribology International, 39,
1545–1557 (2006).
72. Yamaguchi, E. S. and Roby, S. H., Film formation of non-phosphorus wear inhibitors, Tribology
Transactions, 52, 706–716 (2009).
73. Phillips, W. D., Ashless phosphorus-containing lubricating oil additives, Chapter 3 in Lubricant
Additives (Rudnick, L. R., Ed.), Marcel Dekker, New York, pp. 45–112 (2003).
74. Beeck, O., Given, J. W., and Williams, E. C., On the mechanism of boundary lubrication. II. Wear pre-
vention by addition agents, Proceedings of the Royal Society of London, Series A, 968, 102–118 (1940).
75. Godfrey, D., The lubrication mechanism of tricresyl phosphate on steel, ASLE Transactions, 8, 1–11
(1965).
76. Sakurai, T. and Sato, K., Study of corrosivity and correlation between chemical reactivity and load-
carrying capacity of oils containing extreme pressure agents, ASLE Transactions, 9, 77–87 (1966).
77. Barcroft, F. T. and Daniel, S. G., The action of neutral organic phosphates as EP additives, Transactions
of ASME, 87, 761–770 (1965).
78. Davey, W., Extreme pressure lubricants: Phosphorus compounds as additives, Industrial and
Engineering Chemistry, 42, 1841–1847 (1950).
79. Myshkin, N. K., Kim, C. K., and Petrokovets, M. I., Introduction to tribology, Cheong Moon Gak,
Seoul, 1–32 and 204–220 (1997).
80. Kotvis, P. V., Huezo, L. A., Millman, W. S., and Tysoe, W. T., in Surface Science Investigations in
Tribology (Chung, Y. W., Homola, A. M., and Street, G. B., Eds.), American Chemical Society,
Washington, DC (1992).
81. Rizvi, S. Q. A., Additives—Chemistry and testing, in Tribology Data Handbook (Booser, E. R. Ed.),
CRC Press, Boca Raton, FL, pp. 117–137 (1997).
82. Luo, Y. H., Zhong, B. Y., and Zhang, J. C., The mechanism of copper-corrosion inhibition by thiadia-
zole derivatives, Lubrication Engineering, 51(4), 293–296 (1995).
83. Wells, H. M. and Southcombe, J. E., The theory and practice of lubrication: The “The Germ Process,”
Journal of the Society of Chemical Industry, 39, 51T–60T (1920).
84. Allen, C. M. and Drauglis, E., Boundary layer lubrication: Monolayer or multilayer, Wear, 14, 363–384
(1969).
85. Rudnick, L. R., Ed., Lubricant Additives, Chemistry and Applications, 2nd Edition CRC Press,
Taylor & Francis Group, Boca Raton, FL (2009).
21-24 Theory and Practice of Lubrication and Tribology

86. Hardy, H. B. and Doubleday, I., Boundary lubrication—The paraffin series, Proceedings of Royal
Society of London, A100, 550–757 (1922).
87. Arai, K., Yamada, M., Asano, S., and Yoshizawa, S., Lubricant technology to enhance the durability of
low friction performance of gasoline engine oils, SAE Paper 952533 (1995).
88. Muraki, M., Yanagi, Y., and Sakaguchi, K., Synergistic effect on frictional characteristics under roll-
ing-sliding conditions due to a combination of molybdenum dialkyldithiocarbamate and zinc dial-
kyldithiophosphate, Tribolology International, 30, 69–75 (1997).
89. Kasrai, M., Cutler, J. N., Gore, K., Canning, G., and Bancroft, G. M., The chemistry of antiwear films
generated by the combination of ZDDP and MoDTC examined by x-ray absorption spectroscopy,
Tribology Transactions, 41, 69–77 (1998).
90. Miklozic, K. T., Forbus, T. R., and Spikes, H. A., Performance of friction modifiers on ZDDP-generated
surfaces, Tribology Transactions, 50, 328–335 (2007).
91. Iwasaki, H., TOF-SIMS analysis of MoS2 formed on the rubbing surface for MoDTC and ZnDTP syn-
thesized by using 34S isotope, in Proceedings of Japanese Society of Tribologists Tribology Conference,
Takamatsu, Japan, pp. 359–360 (1999).
92. Zhang, Z., Yu, L. G., Yamaguchi, E. S., Kasrai, M., and Bancroft, G. M., Effects of Mo-containing dis-
persants on the function of ZDDP: Chemistry and tribology, Tribology Transactions, 50(1), 58–67
(2007).
93. Komvopoulos, K., Pernama, S. A., Yamaguchi, E. S., and Ryason, P. R., Friction reduction and antiwear
capacity of engine oil blends containing zinc dialkyl dithiophosphate and molybdenum-complex
additives, Tribology Transactions, 49(2), 151–165 (2006).
94. Korcek, S., Johnson, M. D., Jensen, R. K., and McCollum, C., Retention of fuel efficiency of engine
oils, in Technische Akademie Esslingen 11th International “Industrial and Automotive Lubrication,”
Colloquium, Vol. 2, pp. 1281–1288 (1998).
95. Hersey, M. D., Theory of Lubrication, John Wiley & Sons, Inc., New York (1936).
96. Choo, J. H., Forrest, A. K., and Spikes, H. A., Influence of organic friction modifier on liquid slip:
A new mechanism of organic friction modifier action, Tribology Letters, 27, 239–244 (2007).
97. Mang, T. and Dresel, W., Eds., Lubricants and Lubrication, Wiley-VCH, Weinheim, Germany,
pp. 23–33 and 94–106 (2007).
98. Marsden, K., Literature review of OCP viscosity modifiers, Lubrication Science, 1(3), 265 (1989).
99. Gunsel, S., Smeeth, M., and Spikes, H. A., Friction and wear reduction by boundary film forming
viscosity index improvers, SAE Paper 962037 (1996).
100. Müller, M., Fan, J., and Spikes, H. A., Design of functionalized PAMA viscosity modifiers to reduce
friction and wear in lubricating oils, Journal of ASTM International, 4(10) (2007).
II
Lubricants

22 Lubricants  Robert W. Bruce...........................................................................................22-1


Introduction  •  Fluid Lubricants  •  Grease  •  Solid Lubricants  •  Coatings  •  Water-Based
Lubricants  •  Specialty Lubricants  •  Conclusion  •  References
23 Incompressible Fluids  Ronald A. Reich and James R. Anglin....................................... 23-1
Introduction  •  Classes  •  Water-Based Lubricants  •  References
24 Base Oils  Joseph M. Perez and Kimberly Wain Fick......................................................... 24-1
Introduction  •  API Base Oil Categories  •  Refining  •  Base Oil
Applications  •  Performance  •  References
25 Additives for Lubricants  Leslie R. Rudnick...................................................................... 25-1
Introduction  •  Deposit Control Additives  •  Detergents  •  Dispersants  •  Film-Forming
Additives  •  Viscosity Control Additives  •  Corrosion Inhibitors  •  Miscellaneous
Additives  •  Future Trends  •  References
26 Rheology  Scott S. Bair.......................................................................................................... 26-1
Nomenclature  •  Introduction  •  Temperature and Pressure Dependence of
Viscosity  •  Shear Dependence of Viscosity, Generalized Newtonian Liquids  •  Non-
Newtonian Liquids: Normal Stress Differences, Elongational Viscosity  •  Liquid
Failure: Cavitation, Slip, Limiting Stress  •  Opportunities and Challenges for the
Future  •  Acknowledgment  •  References
27 Lubricant Application  Paul W. Hetherington and Evan S. Zabawski...........................27-1
Introduction  •  Gravity Feed Oilers  •  Constant Level Oilers  •  Grease
Dispensing  •  Centralized Lubrication Systems  •  Oil Mist Systems  •  Bibliography
28 Lubricating Grease  Paul A. Bessette.................................................................................. 28-1
Introduction  •  Grease Chemistry  •  Chemistry of Complex Grease  •  Thickener
Ramifications  •  Base Fluids Used in Lubricating Grease  •  Grease Tests  •  Grease
Rheology  •  Grease Specifications  •  Conclusions  •  References
29 Solid Lubricants  Robert W. Bruce...................................................................................... 29-1
Introduction  •  Common Solid Lubricants  •  Application Methods  •  Uses  •  References
30 Metalworking Lubricants  Gregory J. Foltz....................................................................... 30-1
Functions of Metalworking Lubricants  •  Types of Metalworking Fluids  •  Selection
of Metalworking Fluids  •  Metalworking Fluid Control  •  Metalworking Fluid
Application  •  Metalworking Fluid Failure Mechanisms  •  Health, Safety, and Regulatory
Aspects  •  References
31 Hydraulic Fluids  James B. Hannon....................................................................................31-1
Introduction  •  Oil Quality Required by Hydraulic Systems  •  Special
Application Considerations for Hydraulic Fluids  •  Hydraulic System and Fluid
Maintenance  •  References

II-1
II-2 Lubricants

32 Fluid Maintenance  Allison M. Toms and George J.W. Staniewski...........................32-1


Lubricant Selection Process  •  Procurement Process  •  Quality Assurance  •  Storage
Process  •  Lubricant Handling  •  Equipment Modification  •  Condition
Monitoring  •  Re-Greasing Program  •  Disposal  •  Managing the Lubricant
Database  •  Program Evaluation and Benchmarking  •  References
22
Lubricants
22.1 Introduction..................................................................................... 22-1
22.2 Fluid Lubricants............................................................................... 22-1
22.3 Grease................................................................................................22-4
22.4 Solid Lubricants...............................................................................22-5
22.5 Coatings............................................................................................22-5
22.6 Water-Based Lubricants..................................................................22-5
22.7 Specialty Lubricants........................................................................22-5
Robert W. Bruce 22.8 Conclusion........................................................................................22-6
GE Aviation References.....................................................................................................22-6

22.1  Introduction
Lubricants, substances that reduce friction, heat, and wear, are the most powerful implement in the
tribologist’s toolbox. Lubricants may reduce a friction coefficient from over 1.0 to <0.1, reduce oper-
ating temperature of a mechanism by tens to hundreds degree Celsius, and reduce wear by orders of
magnitude.

22.2  Fluid Lubricants


Fluid lubricants may separate surfaces completely to avoid wear and minimize friction coefficients
to <0.1. The most popular gaseous lubricant is air, because it is inexpensive, is abundantly available,
and has no disposal issues. The air turbine dentist drill is an example of air bearing technology, while
many hydrostatic air bearings are being used in the construction of positioning equipment. Of the
common gasses, hydrogen, ammonia, and ethane have viscosities about half that of air, and argon and
oxygen have viscosities slightly higher than air. The dynamic viscosity of air nearly doubles from room
­temperature to 250°C.
Liquid lubricants can minimize operating temperatures due to their ability to absorb heat. Most shaft,
bearing, and gear materials would be exposed to unacceptable temperatures if not for the cooling action
of liquid lubricants. This requires a system for application, collection, and maintenance of the lubricant/
coolant, but controls costs of construction materials. From mineral oil to water and mixtures of both, to
oil-derived chemicals, to liquid salts and metals, a very wide range of liquid lubricants are being used.
A comparison of viscous and cooling properties of fluids is shown in Table 22.1.
For petroleum-based lubricants, the additive technology has extended lubricant life from 500 h or a
season, to 15,000 miles in many cars. The critical properties of liquid lubricants are listed in Table 22.2.
The introduction of fill-for-life lubricants for cars appears to become a reality. The additive technology
for most formulated lubricants is very complex to meet the many specific requirements, so that it is not
advisable to make separate additive additions to those lubricants.

22-1
22-2 Lubricants

TABLE 22.1  Approximate Viscosity of Some Lubricants


Temperature Density Viscosity Viscosity Specific Heat
Material Type (°C) (g/cc) (cSt = mm2/s) (cP = mPa·s) Capacity (kJ/kg·K)
Air 20 0.00120 13.3–15.7 0.016–0.019 1.00
Air 100 0.00095 23.1–23.2 0.022 1.01
Watera 20 1.00 1.00 1.00 4.19
Watera 100 0.96 0.32 0.28–0.31 4.19
Ethanol 20 0.77 1.51 1.07–1.16 2.40
Glycerol 20 1.26 1183 1200–1490 2.40
SAE 10 Motor oil 20 0.87–0.93 115 65–140 1.7–2.2
SAE 10 Motor oil 100 0.84 5 4.2 1.9–2.4
SAE 40 Motor oil 20 0.90 900 320–900 2.20
SAE 40 Motor oil 100 0.84 12.5–16.3 3.7–13.7 2.50
AS5780 Turbine oil 40 0.99–1.02 >23.0 >21.9 1.5–1.9
AS5780 Turbine oil 100 0.94–0.97 4.9–5.4 4.5–5.0 1.9–2.4
cSt = cP/density.
a The conductivity of water is four times that of oil.

TABLE 22.2  Important


Liquid Lubricant Properties
Viscosity
Temperature
Pressure
Shear
Shear rate
Density
Volatility, vapor pressure
Solubility of gas, air
Hydrolysis, water miscibility
Flammability
Oxidative stability
Thermal stability
Foaming
Other
Electrical
Surface tension
EHS

Additive technology continues to improve oxidative and thermal degradation resistance are looking
for substitutes for zinc dithiophosphate (ZDDP) to reduce environmental impact of using this additive,
while maintaining lubrication of high-speed high-pressure concentrated contacts in bearings and gears.
To improve control of the lubricant properties, the increased use of synthetic base stocks continues.
Poly-alpha-olefins are increasingly used in the automotive field which also uses a wide range of addi-
tives to meet the requirements of the frequently, every 3–9 years, revised specifications. Ester-based
lubricants are used in jet engines, with lubricant specifications only in their fourth or fifth generation
in 60 years.
Other liquid lubricants worth mentioning are phosphate esters, silicones, polyphenyl ethers, perfluo-
ropolyethers (PFPE), and polyglycols, especially used for high-temperature applications (see Table 22.3).
Lubricants 22-3

TABLE 22.3  Base Oils


Temperature Temperature
Range Range
Density Viscosity (Minimum (Maximum Comments/
Name Use (g/cc) (cSt = mm2/s) °C) °C) Issues
Mineral oil 0.8 2–140 −25 150
Perfluoropolyethers Space, 1.9–1.9 3.3–4.3 −90 to −20 175–200 Pour point
(PFPE) magnetic −70°C, vapor
storage pressure
Phosphate esters Hydraulic fluid 1.1–1.2 3.2–8.5 −32 to −5 71–120 Nonflammable
Poly-alpha-olefins Automotive 0.80–0.86 1.8–100 −63 to −20 155–288
(PAO) synthetic
Polyalkylene glycols Hydraulic fluid 0.90–1.08 2.7–2500 −45 to −10 152–219 Pour point
(PAG) −25°C
Polyol esters Jet engine 0.98–1.03 3–5 −60 to −57 230–260 Water
lubricant contamination
Polyphenyl ethers High 1.2 13.1 −4 to 5 280–340 Stability to
temperature 400°C, vapor
pressure 10−8 Pa
Silicones Lubricate 1.0 5–100 −80 to −30 177–300 Use −40°C to
plastic and 200°C
rubber

The most important property of the liquid lubricants is viscosity and it must be recognized that vis-
cosity can be affected by entrained gas, shear, shear rate, thermal, and oxidative degradation. In con-
centrated contacts (>200 MPa), the increased pressure increased the lubricant viscosity manifold. The
pressure viscosity coefficient describing this behavior is shown here, with η being the dynamic viscosity:

h = h0 ⋅ exp(a ⋅ P )

The pressure viscosity coefficient for water is 0.00075 MPa−1, and ranges from 0.011 to 0.035 MPa−1 for
most fluids used as lubricants (Jacobson 1991).
Other requirements may include

Coke, sludge, and varnish formation


Compatibility with (polymeric) materials
Compressibility, bulk modulus
Corrosion protection
Density
Electrical conductivity
Environmental compatibility
Evaporation rate
Flammability
Flash point
Foaming, air entrainment
Gas solubility
High-temperature capability
Hydrolysis, stability in contact with water
22-4 Lubricants

Low temperature capability


Oxidative stability
Surface tension
Temperature range
Thermal stability
Vapor pressure
Viscosity index (relation to temperature)

One or more tests exist for determination of each of these attributes. The electrical conductivity of
the lubricant may rapidly increase by an order of magnitude during use, due to the increasing content
of metallic wear debris. Many of the aforementioned requirements may be addressed with additives,
requiring that the base oil is capable of solubilizing these additives.
Contamination of the lubricant during use is often a concern, a rating system for solid contaminant
particles can be found in NAS1638 (2001), ISO 4406 (1999), and SAE AS4059 (2005). Contamination by
liquids often requires the help of the original lubricant supplier to determine the nature of the contami-
nant, possibly the origin, and to find a suitable solution to the problem.
Regulations limit the use of volatile organic lubricants and disposal of all liquid organic lubricants. In
some cases, the disposal costs are close to the purchase price.
The numbers of available lubricant specifications and lubricant brands are incredibly high as the
number of available ingredients is high and mixing ratios are unlimited. See Chapter 24 for a review of
base oils.

22.3  Grease
Greases are a mixture of a liquid lubricant (usually 65–175 cSt) and a thickener that may be based on a soap
(often calcium- or lithium-based), a clay, porous silica, polyurea, silicone, or other materials. Soaps are
created by reacting a metal, mostly lithium but also calcium, sodium, aluminum, and others, with a fat or
fatty acid. Greases provide a generally inexpensive way to provide a liquid lubricant supply. The National
Lubricating Grease Institute (NLGI) lists grades of greases (Table 22.4). The high-temperature greases use
clay or fluorinated organic thickeners and can operate at 250°C. Calcium greases are known for water resis-
tance. Greases with different thickeners or base oils should not be mixed, without compatibility testing.
In addition to the usual additives for lubricating oils, greases may include solid lubricant particles,
tackiness additives, to increase adhesion to bearing surfaces, and fillers such as metal flakes and carbon

TABLE 22.4  Grades of Greases


Worked Penetration
NLGI Grade ASTM D217, D1403 Consistency
000 445–475 Semifluid
00 400–430
0 355–385 Soft
1 310–340
2 265–295
3 220–250
4 175–205
5 130–160
6 85–115 Hard
Lubricants 22-5

black to improve conductivity. And the grease may contain a dye for marketing purposes. See Chapter
28 for a detailed description of the grease manufacturing process.

22.4  Solid Lubricants


Solid lubricants can minimize friction to <0.3 where it may have been >1.0 without the lubricant. This
minimizes component temperatures and extends component life. Solid lubricants are finding increased
use to minimize maintenance costs. Teflon, molybdenum disulfide, and graphite are applied using
polymeric and ceramic binders, or physical vapor deposition (PVD) coating techniques and maintain
friction coefficients from 0.1 to 0.5. For temperatures >500°C (>1000°F), hexagonal boron nitride and
talcum powder are used to minimize friction, as well as coatings containing barium fluoride, calcium
fluoride, and silver.
Many other solid lubricants are no longer used because of environmental or health concerns. Lead
bearings and lead- or beryllium-containing bronzes and Babbitt alloys are disappearing, as are tellu-
rides, selenides, antimony, and lead compounds. Antimony antimonite and antimony oxide have disap-
peared as an adjuvant for graphite bearings to help those cope with low humidity conditions. Chrome
and cadmium plating use are greatly diminished. Lead oxide is disappearing from solid lubricant for-
mulations. See Chapter 29 for details on solid lubricants.

22.5  Coatings
Many coatings applied by plasma spray, PVD, or chemical vapor deposition mitigate wear, although
friction coefficients may be over 1.0. Wear can also be reduced by diffusion treatments such as nitriding,
carburizing, siliconizing, boriding, and aluminiding, which can penetrate 100 μm or multiples thereof
into the treated surface. See Chapters 38 through 41 for further details.

22.6  Water-Based Lubricants


Water-based lubricants can be found as metalworking coolants and also as hydraulic fluids. The ability
of water to penetrate deep into microcracks that affect fatigue has largely kept water-based lubricants
out of metal bearings. The thermal conductivity of water is four times that of oil and the specific heat
is two times, so water is able to keep contact temperatures low and it can thereby minimize wear and
facilitate use of cheaper materials of construction.
Water, a very thin layer of it, is also the lubricant that facilitates the low-friction coefficient associated
with ice skating, bobsledding, and curling. In curling the ice surface is speckled, small droplets of water
sprayed on the ice leave little bumps that reduce the contact area between the stone and the ice, thus
reducing friction. See Chapter 23 for details on water based lubricants.

22.7  Specialty Lubricants


The food industry uses lubricants approved for occasional contact with food by the Federal Food and
Drug Administration. Many of these lubricants also require rabbinical approval. Other lubricants are
approved for contact with skin. Reducing the friction with skin is paramount in improving shaving
comfort. Some materials are used to reduce the friction between strands of hair.
PFPE lubricants are applied to magnetic storage disks to accommodate occasional contact by the
slider that carries the magnetic sensor. Layer thickness may be less than a monolayer.
Various materials are used as lubricants or separating agents for molding and casting processes.
Many of these are based on organic materials, but may also consist of or incorporate salts, solid lubri-
cants, and inert solids.
22-6 Lubricants

Sticks of polymeric and grease-like ingredients are used to lubricate rails, especially where sharp
curves might cause excessive wear between the rail and railcar wheels.
Biodegradable natural oils and greases are increasingly specified for use in chain saws and other
equipment used in the field, like outboard motors and motors for personal watercraft.
Anti-seize compounds are formulated to facilitate loosening of bolted connections with long-term
exposure to high-temperature and aggressive environments. These may contain solid lubricants, fillers/
separating agents, and metal flakes.
Lubricants used for vacuum pumps and vacuum seals have to exhibit low vapor pressure, 10−6 to
10−8 Pa.
Teflon and other PFPE lubricants are used especially when their chemical inertness can be of
significance.
Silicone lubricants are used to lubricate rubber seals in aerospace applications. These are also used to
lubricate zippers of diving suits.
Nonconducting lubricants are used for electrical switches and in the manufacturing process of elec-
trical components.
Traction fluids are designed to maximize traction in elastohydrodynamic contacts. This facilitates
continuously variable transmissions.

22.8  Conclusion
Lubricants can be gaseous, liquid, or solid, or some combination thereof. The number of possible per-
turbations gets extremely large, and matches the number of possible requirements reasonably well.
Development of more environmentally friendly materials is ongoing as is the search for a low-friction
high-temperature >500°C solid lubricant.

References
ISO 4406, Hydraulic fluid power—Fluids—Method for coding the level of contamination by solid
particles, International Standards Organization, Geneva, Switzerland, 1999.
Jacobson, B.O., Rheology and Elastohydrodynamic Lubrication, Elsevier, Amsterdam, the Netherlands,
1991.
National Aerospace Standard, Cleanliness of parts in hydraulic systems, NAS1638, Aerospace Industries
of America, Washington, DC, 2001.
SAE, Aerospace fluid power—Cleanliness classification for hydraulic fluids, SAE AS4059Warrendale,
PA, 2005.
23
Incompressible Fluids
23.1 Introduction..................................................................................... 23-1
Reduce/Control Friction  •  Minimize Wear  •  Cooling  •  Manage
Debris
23.2 Classes...............................................................................................23-2
Petroleum-Based Fluids  •  Synthetic Fluids  •  Biobased
Ronald A. Reich Products  •  Glasses
Alcoa Inc.
23.3 Water-Based Lubricants..................................................................23-8
James R. Anglin Formulating Stable Water-Based Lubricants
Alcoa Inc. References................................................................................................... 23-18

23.1  Introduction
The use of liquids is a very appealing option for providing lubrication. Combining the relative ease of
application with the wide range of viscosity and chemistry options makes possible good solutions to
a great many tribological challenges. Using motor oils as an example, refinements in these lubricants,
in combination with improved engine materials, have enabled greatly improved performance over the
decades. The following four sections describe briefly some of the key attributes of liquid lubricants.

23.1.1  Reduce/Control Friction


While friction can be beneficial, for example, in the action of brakes, it is in general the enemy of energy
efficiency, and the heat generated when friction occurs can be troublesome. Liquid lubricants have the
advantage that under suitable conditions the ratio of the lubricant film thickness to the roughness of the
opposing surfaces is high enough that the surfaces do not make contact. This occurs in hydrodynamic
lubrication where friction can be very low. However, in many situations, contact between the opposing
surfaces is unavoidable and friction increases. In such instances, adsorbed or reacted layers on the sur-
faces, generated in the presence of the lubricant, can markedly reduce the friction. The precise control
of friction can be important in some applications. An example is the formulation of way lubricants that
lubricate the slide ways in precision machining equipment, where static and dynamic frictions need to
be well-matched to avoid stick–slip motion of the tooling.

23.1.2  Minimize Wear


Seemingly small amounts of wear can be very detrimental to operating equipment, whether in the deg-
radation of close tolerances, the generation of vibration in bearings, or the release of wear debris that
can lead to further wear and the catalysis of lubricant oxidation. It is of course important to have proper
materials selection, but the lubricant plays a key role. Avoiding contact through having hydrodynamic
lubrication is clearly one approach. However, where contact occurs, the use of a lubricant composi-
tion that enables formation of surface layers with antiwear or “extreme pressure” (EP) performance is

23-1
23-2 Lubricants

appropriate. The most suitable compositions can depend on the metallurgy of the surfaces, but, in gen-
eral, sacrificial films are formed on the metal surface that wear in a controlled rather than catastrophic
manner under moderate to heavy loading.

23.1.3  Cooling
Liquid lubricants, particularly in systems circulating fluid to the lubricated parts, can serve to remove
heat from those parts and limit the opportunity for severe wear such as galling. Given the relatively
rapid reduction in the viscosity of most lubricants with increasing temperature, their ability to maintain
an adequate film thickness will diminish as temperature increases. The use of a heat exchanger can then
enable control of the temperature of the lubricant to improve its cooling and minimize its chemical
breakdown by oxidative and thermal processes. In certain applications, proper cooling of the equipment
is necessary to maintain geometric tolerances. Of course, water-based lubricants are capable of relatively
high levels of heat removal compared to nonaqueous products.

23.1.4  Manage Debris


Debris in a lubricant can have a number of sources, including contaminants in the fluid as supplied, as
well as debris incorporated during usage. This can include wear debris, debris in the air and in fluid con-
taminants, and varnishes and sludges formed during usage. When a liquid lubricant is flooded into the
contact zone, it can serve to carry away the wear debris, which can then be removed, most commonly
by filtration. Modern hydraulic systems, operating at pressures that can exceed 300 bar, rely on excel-
lent filtration to minimize those of a size that can enter the very narrow clearances of these systems and
cause wear. Such systems should also be equipped with filters on breather ports to minimize airborne
dust ingress. Modern automotive lubricants contain dispersants whose function is to minimize the ten-
dency for soot and fine debris to agglomerate and potentially inhibit oil delivery throughout the engine.

23.2  Classes
In this chapter, anhydrous fluids will be discussed first, followed by water-based fluids, where certain
of the anhydrous fluid types are emulsified or dissolved in the water phase. Historically, many lubri-
cants were biobased, consisting of or containing natural fats and oils and their derivatives, but with the
increasing availability of petroleum-based fluids (mineral oils) by the latter half of the 1800s, petroleum
and products derived from it have become the primary source of lubricating fluids. In more recent
decades, a variety of synthetic products have found application in the marketplace. Two texts the reader
may consult for more detail on these products are Shubkin [1] and Mang and Dresel [2]. More recently
again, there has been renewed interest in biobased products as a result of the environmental benefits
associated with the use of products from renewable sources.

23.2.1  Petroleum-Based Fluids


A brief overview of these fluids is provided here for completeness. For more detail, the reader is referred
to Chapter 24.

23.2.1.1  General Characteristics


Petroleum-based fluids are primarily hydrocarbons that are refined from crude oil that is termed either
paraffinic or naphthenic in keeping with the predominant structural characteristics of the component
molecules. Paraffinic crudes are preferred for most lubricant applications, but naphthenics are preferred
in some applications such as emulsifiable oils and transformer oils. Base stocks ranging in viscosity
from ∼10 cSt to several hundred cSt at 40°C are generated by vacuum distillation in refineries. A range
Incompressible Fluids 23-3

of refining operations is available to decrease pour point, remove S, N, and O from the products, and
reduce the level of aromatic components. The latter operation, typically now performed by hydrotreat-
ing, can improve performance in oxidation tests, improve (raise) the viscosity index (VI), and, under
suitable conditions, reduce the content of polynuclear aromatic hydrocarbons (PAH), which are car-
cinogenic. Products suitably low in PAH can meet requirements for use in food-contact applications.

23.2.1.2  Formulation Considerations


The requirements of lubricant applications, particularly motor oils, have driven the industry toward
more severely hydrotreated petroleum oils with improved low temperature, viscosity, and oxidation
performance. This process narrows the gap between the performance of these products and fully syn-
thetic hydrocarbons, such as polyalphaolefins, which have petrochemical origins. These fluids, how-
ever, are only a basis from which further performance properties are met through additive technology,
enhancing such attributes as the viscometrics, oxidation performance, friction and wear properties,
detergency and dispersancy, and low temperature flow. Where the resulting performance still falls short
of requirements for specific applications, whether in technical, environmental, health, or safety aspects,
a variety of other fluid options are available, as discussed later.

23.2.1.3  Linear Paraffin Oils


A special case of petroleum-derived oils is presented by the linear paraffin oils. These products are
typically isolated from feedstock rich in linear paraffins by molecular sieve technology. While linear
paraffinic materials from the usual lubricant base oil fractions are waxes, those from kerosene–diesel
fractions with average chain lengths no greater than about 16 are liquids at ambient temperatures. Much
of this production has served as raw material for surfactant manufacture. However, other uses, includ-
ing use as process lubricants for metal processing including aluminum rolling, are well established.

23.2.2  Synthetic Fluids


Synthetic fluids offer the formulator the benefits of properties unavailable in petroleum base stocks or
biobased fluids. If a synthetic fluid choice uniquely meets requirements or its advantages outweigh its
disadvantages, including the expected increased cost, then a synthetic can be a good choice. Potential
advantages, depending on the choice of fluid, include
• Wider operating temperature range
• Greater range of available viscosities
• High VI
• Enhanced thermal and/or oxidation stability for greater service lifetime
• Improved fire resistance
• Low volatility
• Low staining tendency, for example, for metalworking fluids
Potential disadvantages to be overcome, again depending on the choice of fluid, can include
• Incompatibility with system materials and other fluids
• Poor additive solubility
• Greater chemical reactivity, including potential for hydrolysis (ester fluids)
• Environmental, health, and safety concerns
• Limited range of available viscosities
Certain synthetic fluids do meet the U.S. FDA requirements for use in lubricants used in the manu-
facture and packaging of food. Among these are certain polyalphaolefins, esters, and polyalkylene
glycols.
23-4 Lubricants

23.2.2.1  Polyalphaolefins
Polyalphaolefins are hydrocarbon fluids that are synthesized typically from C8–C12 olefin precursors
that in turn are synthesized from ethylene. Among the synthetics, they are the most similar to petro-
leum oils in composition and performance. The hydrotreated finished products are highly branched
fully paraffinic materials available in a wide range of viscosities based on their different proportions of
the dimer, trimer, tetramer, and heavier oligomers of the precursor olefins. The products have very good
thermal and oxidative stability, as well as good low temperature flow properties and good VI properties.
The VI values are typically above 120 and vary according to the specific olefin precursors used. A major
application has been in motor oils, but they are also used in a variety of machinery lubricants, especially
where wide ranges of temperature are encountered. Compared with typical petroleum base stocks, addi-
tive solubility and elastomer compatibility can be concerns, but these can be addressed by the addition
of ester to the formulation.

23.2.2.2  Esters
Esters are a very diverse class of compounds that are the reaction products of acids and alcohols. This
discussion is limited to products derived from organic acids and alcohols, but excludes triglycerides,
which will be covered under biobased products. A simple ester, such as isopropyl hexadecanoate (iso-
propyl palmitate), represents the combination of isopropyl alcohol, a simple alcohol with a single alcohol
function, with hexadecanoic acid, a simple fatty acid with a single acid function (Equation 23.1):

CH3 (CH2 )14COOH + (CH3 )2CHOH → CH3 (CH2 )14COOCH(CH3 )2 + H2O (23.1)

However, if the alcohol or acid in the reaction has multiple alcohol or acid groups in it, then the resulting
product can contain multiple ester groupings in a single molecule as is seen in the reaction of hexadeca-
noic acid with ethylene glycol (Equation 23.2):

2CH3 (CH2 )14COOH + HOCH2CH2OH →

CH3 (CH2 )14COOCH2CH2OOC(CH2 )14CH3 + 2H2O (23.2)


Further, polymeric products can result if the starting alcohol and the starting acid both have multiple
alcohol and acid groups, respectively, or a single starting material is used that contains at least one alco-
hol and one acid group in it.
Esters typically used as lubricants have VI values that can range 200 and above and they can impart
good boundary lubrication properties, especially with nonferrous materials. They can have a wide range
of viscosities depending on the choices of alcohol and acid precursors. Some are formulated to ISO
viscosities, especially in the 32–100 range for applications including hydraulic fluids and compressor
lubricants. With the proper precursor selection and purification, ester fluids can have relatively high
flash points of 250°C and above, leading to their recommendation where it is desirable to use fluids
that are difficult to ignite. Certain esters based on alcohol precursors such as pentaerythritol, with no
hydrogen atoms on the carbon atom adjacent to the one on which the alcohol group occurs (β position),
are known to have greater thermal stability. One drawback with esters is their tendency to undergo
hydrolysis, where, in the presence of water, they can react in the reverse direction to Equation 23.1 and
Equations 23.2 through 23.5 reform the precursor acid and alcohol. This tendency can be reduced in
esters synthesized from alcohols with branched structures.

23.2.2.3  Polyalkylene Glycols


Polyalkylene glycols used as lubricants are typically synthesized from propylene oxide or its mixtures
with ethylene oxide to give a polyether where each third atom in the backbone is an oxygen atom. The
Incompressible Fluids 23-5

resulting molecules have structures comprising –OCH2CHR– units, where R=H for ethylene oxide and
R=CH3 for propylene oxide as the precursor for that segment. Where mixtures of the two precursors are
in the product, it can be synthesized with random distributions of the components or with blocks (short
segments) of the pure polymers, imparting different properties. Commonly, the polypropylene glycols
have a butyl group at one end of the molecule; this may also be present in the products based on mix-
tures of the two precursors. The aqueous solubility of polyalkylene glycols displays temperature depen-
dence, such that a cloud point temperature is observed above which the product is no longer soluble.
The cloud point varies greatly with the structure and can be below room temperature for polypropylene
glycols, but above 100°C for pressurized solutions of polyethylene glycol.
Polyethylene glycols are noted to be room temperature solids for molecular weights of 600 and above.
However, polypropylene glycols and products with up to 50% of polyethylene moieties randomly dis-
tributed in polypropylene have quite low melting points. In the viscosity range of most hydraulic fluids,
pour points for these products are typically −30°C and lower and VI values are 160 and above. These
products also resist formation of sludges during usage. Among the main current lubricant applications
of these fluids are their use as compressor lubricants and as textile lubricants, where their ability to be
rinsed off is important. They are also used in the formulation of water–glycol fire-resistant hydraulic
fluids. Fire resistance is much reduced in the absence of water; however, one benefit is the reduced
energy released upon combustion because of the substantial oxygen content in the molecular structure.
Polyalkylene glycols can further be used in aqueous metalworking products, where the cloud point
behavior can contribute to the separation at elevated temperature of the polyalkylene glycol as droplets
that can enhance the tribological performance in the metal–metal contact. Polyalkylene glycols have
also been used as a carrier for introducing solid lubricants in situations where high temperatures cause
the carrier to decompose or burn off.

23.2.2.4  Alkylated Aromatics


These products are typically synthesized by the addition of alkyl groups to the aromatic compounds
benzene or naphthalene. Resulting products of interest are dialkylbenzenes (the monoalkylbenzenes are
preferred for surfactant manufacture) and products more generally termed alkylnaphthalenes. The alkyl
groups typically fall in the C8–C20 range and have linear or branched-chain structures. While properties
will vary with composition, performance for alkylated naphthalenes can include VI values in excess of
130, pour points in the −40°C range, and flash points in excess of 260°C. Applications include a vari-
ety of lubricant types where low temperature fluidity is important, including refrigeration compressor
lubricants, as well as transformer oils and heat transfer applications.

23.2.2.5  Phosphate Esters


These compounds have been used in a wide variety of applications that take advantage of the antiwear
characteristics and good fire resistance of this chemistry. Products based on trialkyl and triaryl phos-
phates are known as well as mixed alkyl aryl phosphates. These products have the composition P(O)
(OR)n(OAr)3n, where R is alkyl, Ar is aryl, and n = 0–3. The mixed alkyl aryl species are used in aircraft
hydraulic applications where low pour points are important. Industrial hydraulic fluids based on triaryl
phosphates with methyl, isopropyl, and tert-butyl substituents are available in the more common ISO
viscosity grades. Among the main classes of water-free hydraulic fluids, these products have the best
overall fire resistance performance in testing by the methods given in ISO 12922. Diligence needs to be
exercised in their usage, namely, to use gaskets and hoses compatible with the fluids, to avoid spillage
into waterways, and to avoid potential formation of toxic products from heating together with trimethy-
lolpropane esters. In use, it is important to keep the moisture content of these fluids low to minimize
formation of acid phosphate hydrolysis products. An ongoing program to remove these acidic prod-
ucts through treatment with activated alumina or ion-exchange resins can help enable good long-term
performance. Combinations of phosphate esters with selected organic esters afford the potential for a
somewhat reduced level of fire resistance at a lower cost compared to the neat phosphate esters.
23-6 Lubricants

23.2.2.6  Silicones
A wide range of chemistries and molecular weights is available for silicones (polysiloxanes), enabling
their use in a great variety of applications. The polymeric species used in lubricants typically have
repeating –OSiR 2– units terminated with –OSiR3 units, where R can be methyl, phenyl, and other sub-
stituents. In general, they have very good thermal and oxidative stability, with phenyl substituted prod-
ucts having good stability in air at 250°C, and they have low pour points, with polydimethylsiloxanes in
common viscosities being below −50°C. Silicones have relatively little viscosity change with temperature
such that the usual VI calculation is not appropriate for them. Silicones have limited affinity for many
other fluid types. Silicone polymers are noted to be self-extinguishing through their production of silica
ash upon burning. While antiwear performance is typically weak, R groups containing chlorine or fluo-
rine substituents can provide improved performance. Greases based on silicone fluids containing solid
lubricants can also have enhanced wear performance. Silicones typically are noted for their low surface
tension, below 25 dyn/cm for common polydimethylsiloxanes, and are commonly used as release agents.
However, traces of most silicones on surfaces can have a markedly adverse effect on the adhesion of coat-
ings and paints. Silicones are known to demonstrate relatively high compressibility under pressure, with
a typical silicone compressing 9.1% at 1380 bar.

23.2.2.7  Ionic Liquids


Research has occurred in the past decade [3] into the potential lubricant use of ionic fluids, particularly
those that exist as room temperature liquids. Such fluids are combinations of cations such as substituted
ammonium and phosphonium species with anions that include tetrafluoroborate, hexafluorophosphate,
and certain amides and sulfonates. The lubricating performance may be attributable to the formation
of ordered layers by the ionic liquids on the contacting surfaces. Potential advantages of the ionic liquid
base stocks include their negligible vapor pressure and low flammability. In addition, good thermo-
oxidative stability has been observed, including good tribological performance [7].
For certain products at 300°C, low volatility is of potential benefit for usage in clean rooms, vacuum,
and situations where emissions are currently a concern, such as rolling lubricants. Work is continuing
toward improving the lubricant performance of these materials through novel compositions or additive
technology, while maintaining good thermo-oxidative stability and minimizing corrosion.

23.2.2.8  Perfluoroalkylpolyethers
This class of lubricants, referred to as PFPE, consists of several families of relatively similar composition
that contain only the elements fluorine, carbon, and oxygen. The backbones of these families contain
carbon and oxygen, while the fluorine has replaced the usual hydrogen on the carbon atoms. PFPE fluids
are colorless and odorless, display excellent thermal and remarkable chemical stability, and are insoluble
in most common solvents. The fluids are essentially nonflammable and quite stable to radiation. The flu-
ids are considered to be very good lubricants. While relatively expensive, PFPE fluids have found a wide
variety of applications, especially under severe conditions where their chemical inertness and high-
temperature stability are advantageous. This can include lubricating valves and seals with reactive and
corrosive materials and the lubrication of computer disk drives.

23.2.2.9  Fischer–Tropsch Synthesis Products


Liquid hydrocarbons can be synthesized in a gas-to-liquid conversion by the Fischer–Tropsch process
using a mixture of carbon monoxide and hydrogen generated from raw materials such as coal and natu-
ral gas. While capital-intensive and practiced currently in a limited number of locations, the gradual
proliferation of large-scale plants will generate increased opportunities to use these products in lubri-
cant applications. Products include linear waxy materials that can be isomerized into high-quality par-
affinic lubricant feedstocks, as well as products in the kerosene–diesel boiling point range that have
Incompressible Fluids 23-7

the potential for use not only in fuels but also in metalworking applications where linear paraffins and
hydrotreated petroleum fractions are now used.

23.2.3  Biobased Products


As noted earlier, interest in biobased materials has greatly increased with the desire in recent years
to use products from renewable resources and to expand the use of environmentally considerate
products [4].

23.2.3.1  Biobased Base Oils


The large majority of biobased oils are triglycerides, where the three alcohol functions of the glycerol are
combined with linear fatty acids in the proportions typical of the source material. These fatty acids have
primarily even numbers of carbon atoms usually ranging from about 12 to 24, with many of the most
common oils from vegetable and animal sources being centered upon the C16 and C18 chain lengths.
Certain products, notably palm kernel oil and coconut oil, do have significant contents of triglycerides
based upon C12 and C14 fatty acids, and others, such as high erucic acid rapeseed oil, have high contents
of fatty acid chains with 22 carbon atoms. Most of these products have viscosities in the ISO 32 to ISO
46 range, making them good candidates for many hydraulic applications. Many also are suitable for the
formulation of food-grade lubricants.
While lubricants based on suitably refined vegetable and animal oils may not have all the technical
features of designed synthetic esters, they generally have some boundary lubricant performance and,
with VI values that can exceed 200, viscosity reductions at higher operating temperatures are reduced
compared with petroleum-derived oils. Their good biodegradability is an asset for use in outdoors loca-
tions or where waste treatment of mixtures with water may be needed. Further, their relatively high
flash points, typically above 250°C, and fire points help them resist ignition. A challenge in formulat-
ing with biobased oils is overcoming weaknesses in oxidative stability and in low temperature flow.
The increasing availability of vegetable oils with oleic contents of 75% and higher provides a technical
advantage. These high oleic products typically contain reduced levels of the polyunsaturated fatty acid
chains subject to rapid oxidative attack. At the same time, they can have low levels of saturated fatty acid
chains that give rise to elevated cloud and pour points. The development of suitable additive treatments
can also help overcome these limitations. As esters, in the presence of water, biobased oils are subject to
hydrolysis as discussed earlier, releasing fatty acid and proceeding through di- and monoglycerides that
have surfactant character.
Where products outside the usual viscosities are desired, lower viscosities are available with mid-
chain triglycerides that contain C8 and C10 fatty acid chains or with simple ester biobased materials, such
as jojoba oil. For increased viscosity, options include castor oil, with a higher viscosity of ∼230 cSt/40°C
attributable to hydrogen bonding of the 12-hydroxy group in the fatty acid chain, or various products
from thermal treatment processes that lead to increases in the molecular weight and thus the viscosity
of biobased esters.

23.2.3.2  Biobased Additives


A wide variety of additives have been generated from biobased precursors for use commonly in
­petroleum-based products. These can include the oil itself, used to provide boundary lubricant perfor-
mance in steam cylinder oils and compounded gear oils, metallic soaps derived from biobased products
and used to thicken oil into greases, and reaction products containing S and/or P that impart additional
EP character to lubricants. A host of oleochemical derivatives have been generated and have found
application, often as boundary lubricant additives, in nonferrous metalworking applications. These
include the component fatty acids from biobased oils as well as fatty alcohols, esters, and other products
synthesized from these materials.
23-8 Lubricants

23.2.4  Glasses
For metalworking operations at high temperatures, such as the hot forging of steel or titanium at tem-
peratures that can be in the 600°C–1500°C range, glass compositions can provide a chemically stable
liquid lubricant composition to control friction and enable more facile metal movement against die
surfaces. The glasses can consist of the usual network-forming oxides, such as silicon oxide and boric
oxide, but can be modified with additional oxides that serve to tailor the melting behavior and viscos-
ity to the process requirements. If lead-containing compositions are considered, the toxicity of the
nine materials needs to be recognized. The compositions can be applied to the workpiece in the form
of suspensions of the finely divided glass components and other formulation ingredients in water or
a suitable solvent and, upon melting to a thin layer, provide some protection to the hot metal from
surface reactions such as oxidation. As with any liquid lubricant, the film thickness and viscosity must
be suitable to retain an adequate film through the duration of the metalworking process. For opti-
mum performance, formulations can include insoluble solid lubricants such as boron nitride or finely
divided abrasives.

23.3  Water-Based Lubricants


Water-based lubricants consist of small droplets of oil and additives dispersed in a water phase. If the
diameters of the oil droplets formed are greater than 0.05 μm, the mixture is opaque and is called an
oil-in-water (O/W) emulsion. If the diameters of the oil droplets are <0.05 μm, the solution is translucent
and is called a microemulsion. A third possibility is that the components of the oil phase do not form
droplets at all. Instead, the components are miscible, forming a clear solution. Water-soluble polyethyl-
ene glycols rather than petroleum oil are required to form clear solutions.
There are several distinct advantages when using a water-based lubricant. First, water is more efficient
at removing frictional and deformational heat than oil due to water’s increased heat capacity and water’s
high energy of evaporation. This has a secondary benefit in reducing hydrocarbon emissions through
the corresponding reduction in the temperature of the lubricating process, which in turn, reduces the
vapor pressure of the oil phase.
The second advantage to using a water-based lubricant is its fire resistance. It is not possible, for
example, to hot roll aluminum without using an O/W emulsion; direct contact between oil and the
surface of the slab will set the oil on fire.
The final advantage to using a water-based lubricant is the ability to starve or supply less oil to a defor-
mation or contact zone than would occur if the lubricant was pure oil. Purposely starving a contact,
for example, increases friction in rolling lubrication [5,6] or decreases traction in elastohydrodynamic
lubrication [7].
There are two distinct disadvantages to using water-based lubricants. The first is corrosion in the
form of either pitting or crevice corrosion when using mild steel, for example. The second is that an O/W
emulsion makes an ideal environment for growing bacteria, mold, and fungus. A corrosion inhibitor or
the use of stainless steel is necessary to prevent the former and a biocide or pasteurization is necessary
to prevent the latter.

23.3.1  Formulating Stable Water-Based Lubricants


Forming a stable emulsion is a necessity when making a water-based lubricant. First and foremost, the
oil and water phases of the emulsion must not separate when being used and, in particular, if the appli-
cation of the emulsion is to starve the contact zone, the stability of the emulsion directly influences the
thickness of the lubricant film [6,7].
Incompressible Fluids 23-9

Mixtures of oil and water are not inherently stable. When mixed together, they quickly separate to
minimize the surface area between the oil and water phases. This separation reduces the Gibbs free
energy of the mixture according to Equation 23.3:

dG = g dA
(23.3)

where
dG is equal to the change in Gibbs free energy
γ is equal to interfacial tension between the O/W interface
dA is equal to the change in surface area between the O/W interface

Microemulsions are an exception to this rule. Here, surface-active agents (surfactants) are purposely
added to the mixture to reduce the interfacial tension (γ) of the hydrocarbon–water interface from
40–60 dyn/cm to <0.001 dyn/cm. Due to this exceedingly small interfacial tension, very small oil drop-
lets (<0.05) form when the two phases are mixed together. This solution is thermodynamically stable
because the decrease in free energy due to the increase in entropy from mixing a large number of small
oil droplets in the water phase (free energy of mixing) offsets the very slight but positive increase in
surface free energy of the O/W interface [8].
Adding a surfactant or combination of surfactants also stabilizes an O/W emulsion but not due to
thermodynamic stability but rather kinetic stability. In this case, the addition of surfactants to the O/W
mixture reduces the interfacial tension to only a tenth or a few dyn/cm rather than the ultralow value
seen for microemulsions. Therefore, the size of the oil droplets formed when mixing the oil and water
phases together is much larger, >0.05 μm, and the decrease in free energy from mixing a smaller num-
ber of larger oil droplets does not offset the increase in surface free energy. Instead, the addition of the
surfactant kinetically stabilizes the emulsion due to either electrokinetic or steric repulsion between oil
droplets.
Surfactants are the key to stabilizing emulsion. Why they behave the way they do and exactly how
they electrokinetically stabilize an emulsion will be described in the next two sections (Sections 23.3.1.1
and 23.3.1.2). How to choose the correct surfactant to stabilize an emulsion will be described in Section
23.3.1.3. Finally, how the stability of an emulsion affects its ability to starve a contact will be described
in Section 23.3.1.4.

23.3.1.1  Surfactants
By definition, a surfactant active agent or surfactant lowers interfacial tension or surface free energy by
spontaneously adsorbing at the O/W interface. This adsorption is due to both the elemental composition
and molecular structure of the surfactant molecule. All surfactants have the same general structure: a
long hydrocarbon chain attached to a polar head. The hydrocarbon chain specifically adsorbs in the oil
phase at the interface and the polar head adsorbs in the water phase. The hydrocarbon chains also line
up next to each other due to London dispersion forces, plus the polar ends of most surfactant molecules
hydrogen bond with the water molecules that surround the oil droplet (Figure 23.1).
The hydrocarbon chain is anywhere from 8 to 18 carbon atoms long. It can be saturated, unsatu-
rated, or branched. The polar head, on the other hand, consists of molecular structures such as amines,
sulfates, phosphates, carboxylates, hydroxyls, and ethoxylated moieties of the previous structures.
They form either dipoles or ions in the water phase and many are capable of hydrogen bonding with
water molecules due to the presence of oxygen or nitrogen atoms. Table 23.1 lists typical examples of
surfactants.
The number of surfactant molecules per area (Ns/A) adsorbed at an O/W interface is determined
using the Gibbs adsorption isotherm (Equation 23.4). This value is called the surface excess (Γ) and
23-10 Lubricants

Hydrocarbon
chain

Oil

δ+ δ+ δ+ δ+ Water
Polar head
δ– δ– δ– δ–

H H H
HO H HO
HO HO

FIGURE 23.1  Adsorption of surfactant molecules at the O/W interface.

TABLE 23.1  Surfactant Examples


Sulfates: ammonium lauryl sulfate, sodium lauryl sulfate
Phosphate: alkyl ether phosphate
Carboxylates: sodium stearate, potassium oleate
Amine: dioctadecyldimethylammonium bromide
Hydroxyls: lauryl alcohol, oleyl alcohol
Ethoxylated molecules: polyoxyethylene glycol alkyl ethers, polyoxypropylene glycol alkyl ethers

is equal to the total number of surfactant molecules in the system minus the sum of the number of mol-
ecules in the bulk water phase and the bulk oil phase:

Γ = N s / A = −dg / RTd lnC (Gibbs isotherm) (23.4)


where
γ is equal to the interfacial tension
R is equal to the universal gas constant
C is equal to the concentration of the surfactant in either of the bulk phases

Since interfacial tension decreases with the logarithmic concentration of the surfactant, the effective-
ness at lowering interfacial tension by adding additional surfactant molecules to the mixture decreases
as the concentration of the surfactant increases. Typical surfactant concentrations in O/W mixtures
vary between 5% and 20%. Rarely is the concentration of the surfactant greater than 20%.

23.3.1.2  DLVO Theory (Kinetic Stabilization of Emulsions)


In addition to lowering the interfacial tension between the O/W interface, surfactant molecules also
kinetically stabilize oil droplets due to either electrokinetic or steric stabilization. The DLVO theory is
used to explain why surfactant molecules stabilize an emulsion electrokinetically. This theory was devel-
oped by Derjaquin, Landau, Verwey, and Overbeek, and hence the name DLVO [9,10]. Because the theory
is complex, only a general overview will be presented here. However, a qualitative understanding of the
DLVO theory is necessary to fully comprehend what prevents oil and water from separating after mixing.
Incompressible Fluids 23-11

The concept behind the DLVO theory is simple. When two oil droplets approach each other in a
liquid media, their interaction is governed by two forces: an attractive force and an electrokinetic repul-
sive force. The attractive force is brought about by the natural attraction between molecules through
long-range dispersion forces or van der Waal forces and the electrokinetic repulsive force is due to the
buildup of an electric charge on the surface of the oil droplets when using ionic surfactants [11].
The potential energy of attraction between particles is due to van der Waal forces and is described by
Equation 23.5:

U (D) = − AR /12D (23.5)

where
D is equal to the distance between the particle
R is equal to the radius of the particle
A is equal to the Hamaker constant of the material that makes up the particle

A molecule’s Hamaker constant is a measure of its van der Waal energy of attraction [12]. Values
for many organic molecules are known and range from 10−21 to 10−20 J. Hamaker constants can also be
­calculated using Equation 23.6:

A = p 2C r1 r2 (23.6)

where
C is the particle–particle pair interaction constant
ρ is the particle atom density

Since the van der Waal force is attractive, the corresponding potential energy of attraction is negative
(Figure 23.2).

Potential energy of repulsion


U
0

U 0

Potential energy of attraction

FIGURE 23.2  Potential energy (U) diagrams of attraction and repulsion between two oil droplets (D is equal to
the distance between oil droplets).
23-12 Lubricants

Electrical potential

ψ Double layer

Zeta potential D

Oil phase Water phase

Surface potential (ψ)


Plane of zero shear, zeta potential (ζ)
Stern layer

FIGURE 23.3  Electrical double layer, zeta potential.

The electrokinetic repulsion between oil droplets is more complex but simply stated it is a repulsive
force that develops between oil droplets due to the buildup of an electric charge on the surface of the
oil droplets. For example, when an anionic surfactant is present at an O/W interface, the polar head
of the ionic surfactant dissociates to form an anion and a cation. The long-chain hydrocarbon anchors
the negative charge of the anion to the surface of the oil droplet. The positive-charged cation remains
in the water phase but very close to the surface of the oil droplet to balance the negative charge of the
anions. Many of these counter cations remain tightly bound to the surface, forming a Stern layer or
Helmholtz layer. The remaining ions adjacent to the Stern layer form a diffuse double layer as shown
in Figure 23.3.
The potential energy of repulsion due to the buildup of charge on the surface of each oil droplet is
described by Equation 23.7:

U (D) = (64pKTRrg 2 /k 2 )e −kD (23.7)


where
R is diameter of the oil droplet
K is equal to the Boltzmann constant
κ is a function of the ionic composition of the water phase
ρ is the concentration of ions in solution
γ is the reduced electrical potential at the surface of the oil droplet

Since the electrokinetic force is repulsive, the corresponding potential energy of repulsion is positive
(Figure 23.2).
The electrokinetic potential of the Stern layer cannot be measured but the potential at the plane of
slip between the oil droplet and water phase (Figure 23.3) can be measured by electrophoresis and is
known as the zeta potential (ζ). Unstable emulsions tend to have values less than −15 mV, semi-stable
emulsion between −15 and −45 mV, and very stable emulsions have values greater than −45 mV when
using anionic surfactants.
By combining Equations 23.5 and 23.7, a resulting potential energy diagram describing the interac-
tion between oil droplets as they approach each other can be drawn (Figure 23.4). As can be seen in
Figure 23.4, a secondary minimum first forms in which the oil droplets flocculate together but do not
completely coalesce to form a larger oil droplet. This minimum can be quite stable and it is not unusual
to see a multiple number of oil droplets flocculated together in this manner when viewing an emulsion
Incompressible Fluids 23-13

DLVO theory potential energy diagram

Potential energy

Energy barrier

U
D Secondary minimum
Flocculation

Coalescence

FIGURE 23.4  Combined potential energy diagrams of repulsion and attraction (D is equal to distance between
oil droplets).

using an optical microscope. For complete coalescence to occur, the Brownian motion of the flocculated
oil droplets must have enough thermal energy to overcome the energy barrier that leads to the primary
minimum in Figure 23.4.
It is possible using Equations 23.5 and 23.7 to calculate the values for the potential energy of interac-
tion between particles but usually this calculation is done to verify Equations 23.5 and 23.7 rather than
to predict the stability of an emulsion for everyday use. The real power of the DLVO theory is using the
resulting potential energy diagram to understand how to change the attractive and repulsive forces
between oil droplet to either form a stable emulsion and, if necessary, to form an unstable emulsion to
separate the oil and water phases for waste treatment. For example, it has already been stated that the
zeta potential must be increased to values greater than −45 mV to stabilize an emulsion. However, the
electrical double layer that balances the charge on the surface plays an important role in electrokinetic
stabilization as does the magnitude of the zeta potential. The reason for this is that the diffuse double
layer effectively screens the electrical charge on the surface of the oil droplets, allowing oil droplets to
approach much closer to each other before the electrokinetic repulsive force is felt. In effect, the double
layer decreases the height of the potential energy barrier that must be crossed before coalescence can
occur. Hence, purposely changing the characteristics of this double layer can dramatically change the
stability of an emulsion.
The decrease in the electric potential (Stern potential) due to the presence of the electrical double
layer is equal to

y = y oe −kD (23.8)

where
ψo is equal to the Stern potential
D is equal to the distance
κ is the inverse thickness of the electrical double layer, also known as its Debye length (κ−1)

This length is much smaller than the diameter of the oil droplet, a few nanometers at most. At a distance
of 2 Debye lengths, the electrical potential drops to around 2% of the potential at the surface of the oil
droplet.
A more important characteristic of the double layer is that even though the repulsive force between
oil droplets is described as being electrokinetic in nature, the real force comes from an osmotic pressure
23-14 Lubricants

that builds between oil droplets due to the overlap of their electrical double layers. The high concentra-
tion of ions between oil droplets draws water into the gap between them. This increases the pressure in
the gap, pushing the oil droplets apart.
If additional salt is added to the water phase of an emulsion, the density of cation ions surrounding
the negatively charged oil droplet at the Stern layer increases. This decreases the Debye length of the
double layer, which, in turn, increases the effectiveness at screening the charge on the surface of the
oil droplet. Oil droplets can approach much closer together before their double layers start to overlap,
thereby, increasing the potential energy of attraction between them. If the concentration of cations is
great enough, the double layer will completely collapse, allowing for rapid coalescence. In other words,
the height of the potential energy barrier in Figure 23.3 can be reduced to a very small value if enough
salt is purposely added to the emulsion.
The addition of salt to break an emulsion is called salting out. Cations that have double or triple
charges are much more efficient at breaking emulsions relative to cations that have a single charge. On
the other hand, using deionized (DI) water or reverse osmosis (RO) water is more effective at stabilizing
an emulsion relative to using potable water due to the difference in the hardness of the water.
The ionic strength of the water phase can also increase during the lifetime of a water-based lubri-
cant through the dissolution of ions from wear debris. When working with aluminum, the two most
common ions produced are triply charged aluminum ions and doubly charged magnesium ions. Left
unchecked, the stability of a water-based lubricant will deteriorate rather quickly due to the accumula-
tion of these ions.
Removing wear debris by filtration helps reduce the buildup of ions in the water phase. The ions in
the water phase can also be removed using ultrafiltration (membrane) to continuously separate part of
the water phase and the salt it contains from the emulsion. Adding back DI or RO water to the emulsion
will dilute the remaining ions. In extreme cases, it may be necessary to continuously throw part of the
emulsion away and make up with new oil and new water to maintain stability.
Besides changing the magnitude of the zeta potential or the thickness of the double layer, the stabil-
ity of an emulsion can also be increased by changing the potential energy of attraction rather than the
electrokinetic energy of repulsion (Equation 23.5). One way to do this that has already been described is
changing the Hamaker constant of the components of the oil phase, that is, changing either the base oil
or the additives. If changing the components of the oil phase is not an option, a second way to decrease
the potential energy of attraction is to change the size of the oil droplets. Decreasing interfacial tension,
for example, by either increasing the concentration of the surfactant or changing to a different surfac-
tant are two ways of doing this but increasing the mechanical shear rate when mixing the oil and water
phases together may be a more practical way. Homogenizing milk is an example that is most familiar to
everybody. In this case, the high shear rate of the homogenizer breaks the fat globules of the milk into
very small particles, preventing them from separating due to a decrease in the van der Waal forces of
attraction between droplets.
There are two other ways to stabilize an emulsion, which must be mentioned here. The first is the use
of polymeric nonionic surfactants rather than an ionic surfactant. Like an ionic surfactant, the hydro-
carbon chain or hydrophobic part of the nonionic polymer adsorbs in the oil phase and the polar end of
the polymer adsorbs in the water phase. The makeup of the polar group, however, consists of repeating
units of ethylene oxide (–CH2–CH2–O–), which can extensively hydrogen bond with water. The poten-
tial energy diagram of repulsion for a nonionic surfactant takes the same shape as the diagram when
using an ionic surfactant but in this case the repulsive force between two oil droplets is due to steric
repulsion when the polar ends of the nonionic molecules overlap.
The second way to either stabilize or destabilize an emulsion used as a lubricant is through the forma-
tion of wear debris. If, for example, the oil phase wets the surface of the debris particle, then one end
of the debris particle can attach itself to an oil droplet and another end can attach itself to another oil
droplet, causing the oil droplets to coalesce. On the other hand, if the debris particle is not wetted by the
oil phase then a major part of the debris particles adsorbed at the O/W interface remains in the water
Incompressible Fluids 23-15

Young’s eq. γw/m = γo/w cos θ + γo/m

γo/w

Water θ Oil

γw/m Metal γo/m

FIGURE 23.5  Contact angle of oil droplet at W(water)/M(metal) interface.

phase. It prevents coalescence in the same manner, as does a nonionic surfactant, through steric repul-
sion between debris particles attached to different oil droplets.
Whether a debris particle is wetted by the oil phase or water phase can be determined using a goni-
ometer to measure the contact angle of the oil–water–metal interface. How an oil droplet wets a surface
(Figure 23.5) is based on the balance of interfacial tensions between the O/W interface, the oil/metal
interface, and the water/metal interface (Equation 23.9). Contact angles <90° indicate that the particle
prefers oil to water:

gw/m = go/w cos q + go/m (23.9)


What is important here is that by purposely changing the interfacial tension of the O/W interface by
adding a surfactant or changing its concentration may also change how debris is wetted by the oil phase
and could potentially decrease the stability of the emulsion even if the emulsion is stable when no debris
is present. In general, a water-based lubricant contaminated by metal fines is less stable.
Finally, if an emulsion must be partially thrown away and replaced with new components to maintain
its stability then disposing large volumes of an emulsion can become an environmental problem. An
emulsion cannot simply be discharged into the municipal sewer due to increases in the chemical oxygen
demand (COD) and biological oxygen demand (BOD) of the waste stream from the decomposition of
the oil phase by either oxidation or bacteria. Before the water phase can be discharged, the oil phase
must be separated.
There are several ways or combination of ways to destabilize an emulsion that come directly from
changing the potential energy of repulsion between oil droplets. The first way is to decrease the charge
on the surface of the oil droplet by decreasing the pH of the emulsion. This neutralizes the negative
charge of an anionic surfactant if the anion is the conjugate base of a weak acid. If the pH of the solution
cannot be decreased, then the charge on the oil droplets can be neutralized by adding alum (Al 2(SO4)3)
to the emulsion. In this case, the aluminum ions from the alum not only neutralize the charge on the
oil droplets but the AL(OH)3 precipitate that forms causes further flocculation of the oil droplets as it
settles. A third possibility is to add a polyelectrolyte to neutralize the charge by the adsorption of an
oppositely charged polymer on the surface of the oil droplet. For an anionic surfactant, the most com-
mon polyelectrolyte used is a poly-quaternary ammonium salt. For most waste treatment facilities, it is
not unusual that a combination of these techniques is needed to successfully separate the oil and water
phases.
Choosing the appropriate surfactant to stabilize an emulsion, therefore, is considered the most
important task in developing a water-based lubricant. Not only must it stabilize the oil and water mix-
ture when first formulated but the emulsion must remain stable after several weeks if not months of use.
And, when the emulsion must be disposed, there must be a way to break the emulsion in order to sepa-
rate the oil phase from the water phase. But even though the DLVO theory can explain why a surfactant
stabilizes an emulsion, it cannot tell which surfactant to use. How to choose the correct surfactant for a
system will be described next.
23-16 Lubricants

23.3.1.3  Hydrophile–Lipophile Balance


There are many ways to choose a surfactant to stabilize an emulsion. For example, a chemist may develop
a feel for what type of surfactants stabilizes different types of oils but this methodology usually falls
short when working with something new. A more systematic way to choose a surfactant is to identify
the HLB (hydrophile–lipophile balance) requirement of the oil phase [13,14]. By choosing a number of
surfactants that have the same HLB number, there is a better chance of finding a surfactant or a combi-
nation of two surfactants to form a stable emulsion.
The HLB numbering scale was specifically developed for nonionic surfactants by Griffin [13,14] in the
1950s. HLB is a measure of whether the surfactant prefers the water phase (high HLB numbers) or the oil
phase (low HLB numbers). Nonionic surfactants, which have an HLB of 15 or higher, for example, have
a large number of repeating units of ethylene oxide. They prefer the water phase due to the formation of
hydrogen bonds between water and the ethylene oxide units of the surfactant molecule. Surfactants, which
have an HLB of 0, have no repeating units; therefore, surfactants with low HLB numbers prefer the oil phase.
Typical HLB values for surfactants and what they are used for are described in Table 23.2. The HLB
numbers of most surfactants are known and are listed by their manufacturer. Griffin’s original scale was
developed for nonionic surfactants where the molecular weight of the repeating units of ethylene oxides
in the surfactant molecule is divided by the total molecular weight of the surfactant and the resulting
percentage is divided by 5 to make a more workable scale. For example, using the aforementioned for-
mula, the HLB of lauryl alcohol with 10 repeating units of ethylene oxide is equal to
• Lauryl alcohol (H3C–(CH2))10–CH2–OH) molecular weight 186
• Ten repeating units of ethylene oxide (–CH2–CH2–O–) molecular weight 440
• H3C–(CH2)10–CH2–O–(–CH2–CH2–O)10H molecular weight 626
• HLB = (440/626) × 100/5 = 14.1
To determine what HLB is best for making a water-based lubricant, the preferred HLB value for the oil
phase must also be determined. This is done using a series of nonionic surfactants with known HLBs
that vary over a wide range. Each surfactant is then mixed at a 10% concentration level with 45% of the
oil phase and 45% of the water phase. Left standing for several hours, the HLB requirement of the oil
phase is the HLB of the surfactant used that has the least amount of separation for the aforementioned
mixtures. The most stable mixture will also have the lowest interfacial tension.
This process identifies the most likely HLB number, which will stabilize the oil phase. It does not nec-
essarily identify the surfactant or cosurfactant that will stabilize the oil phase of the water-based lubri-
cant unless by coincident the nonionic surfactant used to identify the HLB of the oil phase does form a
completely stable emulsion, no separation of oil and water after several days. To choose the appropriate
surfactant that will emulsify the oil phase, a variety of different ionic and nonionic surfactants can be
tried that have an HLB similar to the HLB requirement of the oil phase ±1 unit. Also, it is better to
choose two surfactants, one ionic and one nonionic, whose weighted average HLB number matches the
required HLB of the oil phase rather than using only a single surfactant.
Finally, in large systems that use a water-based lubricant, there is a high probability that the emulsion
will become contaminated by a second oil phase. Hydraulic leaks, for example, are a common cause

TABLE 23.2  HLB Numbers and Use


0 to 3 antifoaming agent
4 to 6 W/O (water-in-oil) emulsifier
7 to 9 wetting agent
8 to 18 O/W (oil-in-water) emulsifier
13 to 15 detergents
10 to 18 solubilizer
Incompressible Fluids 23-17

for contamination when operating machines such as a rolling mill. In this case, it is preferred that the
hydraulic oil has a different HLB requirement than the oil phase of the water-based lubricant. By having
different HLB requirements, the hydraulic fluid will not be emulsified by the surfactants that make up
the emulsion and instead will be rejected. Simple skimming can be used to remove the contamination.

23.3.1.4  Lubricant Starvation


Stabilizing an emulsion is a necessity to keep the oil and water phases from separating but it is also
known that factors that affect emulsion stability or are affected by its stability such as oil droplet size and
concentration directly influence the lubricating performance of an emulsion involving starvation [6]. It
is a formulator’s responsibility to optimize these factors to maintain consistent lubrication.
Emulsions are unique in that even though they are composed mostly of water, they behave in the lubri-
cating process as though they are made from pure oil except for their cooling capacity. The reason for this
discrepancy is that the oil phase is what passes through the deformation or contact zone in a lubricating
process. If this were not the case, emulsions would not be used as lubricants because unlike oil, water’s
viscosity does not significantly increase with pressure. And since an O/W emulsion is mostly water, it
would not form a thick film if it was the dominant phase passing through the deformation zone [15].
How an oil film forms in the deformation zone of a lubricating process, which uses an emulsion, is
of great importance and is related to an emulsion’s stability. Furthermore, manipulating the properties
of the emulsion to maintain a specific oil concentration and oil droplet size is necessary to maintain
consistent lubricating properties particularly when the contact is starved [6,7].
To form a lubricant film in a deformation or contact zone, the pressure in the pool of oil in front of
the deformation zone builds from atmospheric pressure to the deformation or contact pressure at the
start of the deformation or contact zone. This buildup of pressure helps lift the surface of the tool of the
workpiece and how far the two surfaces separate determines how thick the lubricant film between tool
and work piece becomes.
Film thickness equations for a variety of process such as rolling [16] and elastohydrodynamic lubrica-
tion [17] are known and illustrate this concept. For example, Wilson and Walowit derived a film thick-
ness equation (Equation 23.8) for rolling aluminum metal, starting with the Reynolds equation [16].
They showed that the thickness (hww) is dependent on a variety of things including roll and metal entry
speed (vr + vm), oil viscosity (n), oil pressure–viscosity coefficient (α), deformation pressure (y), entry
tension (s), and bite angle (Θ). Their derivation assumes an infinite distance between the start of defor-
mation zone in the roll bite and the location of the air/oil meniscus in front of the deformation zone:

hww = 3no a (vr + vm ) (Θ)−1 (1 − e −a (y − s ) )−1 (23.10)


For a mineral oil lubricant with a pressure–viscosity coefficient of approximately 10−8 m2/N, this
assumption holds. For an emulsion, however, this assumption is incorrect since the major component
of the emulsion that is located at what would be considered an infinite distance from the deformation
zone is water and not oil. Due to water’s low pressure–viscosity coefficient (approximately zero), pressure
will not increase in front of the deformation zone until a pool of oil forms. The continuous phase of the
emulsion, therefore, must invert to oil at some point in front of the deformation zone.
If the location of this inversion point is located far from the deformation zone, then the origi-
nal assumptions by Wilson and Walowit hold true for the derivation of the film thickness equation.
However, if the location of the inversion point is located close to the deformation zone, the buildup of
pressure in front of the roll bite takes place in an area much smaller than the area previously described
and therefore the force pushing the work roll up is less. The film thickness is less than what it would be
for a fully flooded process and therefore the contact is starved. The film thickness equation (Equation
23.11), in this case, includes the additional parameter, the height of the oil/water meniscus (h*) where
inversion occurs [5]:
23-18 Lubricants

hs = h* + h* /2hww − 1/2  (4h * /hww + h * /hww 2 )


2 3 4
(23.11)

The effect of starving a contact is to increase friction in rolling due to a decrease in film thickness [6].
However, in elastohydrodynamic lubrication, the effect of starvation is to lower friction or traction due
to a decrease in the volume of oil that is sheared in front of the contact zone [7].
Starving the contact zone using an emulsion is another way to control friction. It has been shown,
both in rolling lubrication [6] and elastohydrodynamic lubrication [7,18], that the thickness of the lubri-
cant film is dependent among other things on the concentration of the oil and the size of the oil drop-
lets, which in turn are dependent on the stability of the emulsion. Maintaining stability is critical to
maintaining the same friction when a contact is starved. Likewise, being able to change the percent oil
concentration and or the size of the oil droplet allows one to change the thickness of the lubricant in the
deformation zone. This is not possible with a mineral oil–based lubricant without changing the viscosity
of the lubricant.

23.3.1.5  Summary
Water-based lubricants have unique properties that make them preferred to oil-based lubricants in
many applications. Among these properties are improved heat extraction, lower hydrocarbon emis-
sions, and fire prevention. However, water-based lubricants that are O/W emulsions are thermodynami-
cally unstable and therefore must be kinetically stabilized by choosing a surfactant or combination of
surfactants, which have an HLB number or combined HLB number that matches the HLB requirements
of the oil phase. Other factors that affect the stability of an emulsion according to the DLVO theory are
oil droplet size, zeta potential, salt concentration, and Hamaker constant. In special lubricating process
such as rolling or elastohydrodynamic lubrication, the lubricant film that forms from an emulsion in the
contact or deformation zone can be purposely starved by controlling the oil droplet size and concentra-
tion of the emulsion, factors that are directly affected by the stability of the emulsion.

References
1. Shubkin, R.L., Ed., Synthetic Lubricants and High-Performance Functional Fluids, Marcel Dekker,
New York, 1993.
2. Mang, T. and Dresel, W., Eds., Lubricants and Lubrication, Wiley-VCH, Weinheim, Germany, 2007.
3. Bermúdez, M.-D., Jiménez, A.-E., Sanes, J., and Carrión, F.-J., Ionic liquids as advanced lubricant
fluids, Molecules, 14, 2888–2908, 2009.
4. Rensselar, J.V., Biobased lubricants: Gearing up for a green world, Tribology and Lubrication
Technology, 66, 32–41, 2010.
5. Reich, R., Panseri, N., and Bohaychick, J., The effect of lubricant starvation in the cold rolling of
aluminum metal when using an oil-in-water emulsion, Lubrication Engineering, 57, 15–18, 2001.
6. Reich, R. and Urbanski, J., The experimental support for the dynamic concentration theory of form-
ing and oil reservoir at the inlet of the roll bite by measuring the onset speed of starvation as a func-
tion of oil concentration and droplet size, Tribology Transactions, 47, 489–499, 2004.
7. Yang, H., Schmid, S.R., Kasun, T.J., and Reich, R.A., Elastohydrodynamic film thickness and tractions
for oil-in-water emulsions, Tribology Transactions, 47(1), 123–129, 2004.
8. Prince, L.M., Ed., Microemulsions: Theory and Practice, Academic Press, Inc., New York, 1977.
9. Derjaguin, B. and Landau, L., Theory of the stability of strongly charged lyophobic sols and of the
adhesion of strongly charged particles in solutions of electrolytes, Acta Physico Chemica URSS, 14,
633, 1941.
10. Verwey, E.J.W. and Overbeek, J.Th.G., Theory of the Stability of Lyophobic Colloids, Elsevier,
Amsterdam, the Netherlands, 1948.
Incompressible Fluids 23-19

11. Sparnaay, M.J., The electrical double layer, International Encyclopedia of Physical Chemistry and
Chemical Physics, D.H. Everett, Ed., Vol. 4, Pergamon, New York, 1972.
12. Hamaker, H.C., The London—van der Waals attraction between spherical particles, Physica, 4(10),
1058–1072, 1937.
13. Griffin, W.C., Classification of surface-active agents by ‘HLB,’ Journal of the Society of Cosmetic
Chemists, 1, 311, 1949.
14. Griffin, W.C., Calculation of HLB values of non-ionic surfactants, Journal of the Society of Cosmetic
Chemists, 5, 259, 1954.
15. Bett, K.E. and Cappi, J.B., Effect of pressure on the viscosity of water, Nature, 207, 620–621, 1965.
16. Wilson, W.R. and Walowit, J.A., An isothermal hydrodynamic lubrication theory for strip rolling
with front and back tension, Proceedings of the Tribology Convention, Institution of Mechanical
Engineers, London, U.K., pp. 164–172, 1971.
17. Dawson, D. and Higginson, G.R., A numerical solution to the elastohydrodynamic problem, Journal
of Mechanical Engineering Science, 2, 188–194, 1960.
18. Zhu, D., Biresaw, G., Clark, S.J., and Kasun, T.J., Elastohydrodynamic lubrication with O/W emul-
sion, Journal of Tribology, 116, 310–320, 1993.
24
Base Oils
24.1 Introduction.....................................................................................24-1
24.2 API Base Oil Categories..................................................................24-1
Joseph M. Perez Group I  •  Group II  •  Group III  •  Group IV  •  Group V​
Pennsylvania State Group I + Base Oils  •  Group II + Base Oils  •  Group III + Base Oils
University 24.3 Refining.............................................................................................24-3
24.4 Base Oil Applications......................................................................24-5
Kimberly Wain Fick
24.5 Performance.....................................................................................24-6
Pennsylvania State
University References.....................................................................................................24-9

24.1  Introduction
Most lubricants are composed of a base oil produced to specification by a manufacturer and a number of
additives to enhance the performance of the base oil. The American Petroleum Institute (API)* defines
the term base oil to mean either a single base stock or a mixture of two or more base stocks. The base oils
may be of petroleum or mineral oil origin, synthetic fluids, or natural oils. The API has categorized base
oils into five categories or groups. The base oils are petroleum base stocks or chemically engineered fluids
of petroleum intermediates that are refined or synthesized by a variety of processes including distillation,
solvent refining, catalytic dewaxing, hydrogenation, and hydrocracking. In addition, there are a number
of base oils from natural resources or chemically synthesized such as vegetable base oils, silicones, silicates,
fluorocarbons, polyphenyl ethers, etc. The base oil can be a single fluid or a mixture of two or more fluids.
The base oil is the major component in a lubricant formulation and it is selected for its physical,
chemical, and tribological properties by the lubricant manufacturer. Normally, additives amount to
20 vol% or less of the lubricant formulation. The number of different types of additives and the amount
of additives in the lubricant depend on the end use of the lubricant.

24.2  API Base Oil Categories


The criteria for API categories are found in Table 24.1. The categories are described as follows.

24.2.1  Group I
Group I base oils are the least refined of all the groups. They are a mix of various hydrocarbons some
of which contain sulfur, oxygen, and nitrogen constituents. The normal processing of the base oil is
fractional distillation followed by solvent extraction to remove wax to improve low temperature proper-
ties and improve additive susceptibility for better oxidation stability and performance. Most Group I

* The American Petroleum Institute was formed in 1919 to standardize specifications for petroleum products and is the
only national trade association that represents all aspects of U.S. oil and natural gas industry. There are ~400 corporate
members, producers, refiners, suppliers, pipeline operators, and marine transporters, and service and supply companies.

24-1
24-2 Lubricants

TABLE 24.1  API Base Oil Categories


Saturated
Category Sulfur, wt% Components, wt% Viscosity Index
Group I >0.03 And/or <90 80–120
Group II ≤0.03 And >90 80–120
Group III <0.03 And >90 >120
Group IV All polyalphaolefins (PAO)
Group V Various

base stocks have viscosity indexes (VIs)* <100. Group I base oils obtained from Pennsylvania crude oils
normally have VIs of 95 or higher.

24.2.2  Group II
Group II base oils are fractionally distilled and hydroprocessed. The hydroprocessing removes some of
the sulfur and other polar constituents. This produces quality base oil with improved additive suscepti-
bility and oxidation stability.

24.2.3  Group III


Group III base oils are subjected to the highest level of refining of the base oil groups. They are hydro-
cracked and hydrotreated to produce oils with good molecular uniformity, stability, and excellent
viscosity–temperature properties. Some lubricant manufacturers formulate them with additives and
market them as synthetic or semisynthetic products.

24.2.4  Group IV
Group IV base oils are fluids that are chemically engineered synthetic base stocks. Group IV contains all
polyalphaolefins (PAOs) and similar products produced through selective chemical reaction. The fluids
contain highly uniform chains that offer a wide range of improved tribological properties. Formulated
fluids have excellent stability, viscosity–temperature, and low temperature properties. Material compat-
ibility can be an issue that is resolved by blending with other base stocks.

24.2.5  Group V
Group V is the category for all remaining base oils not found in Groups I–IV. It includes mostly noncon-
ventional base oils including phosphate esters, silicones, silicates, polyglycols, polyphenyl ethers, fluo-
rocarbons, naphthenic pale oils,† monobasic, dibasic, and polyol esters, and natural oils. Some Group V
base oils are used in lubricant formulations both as base oil components and as additives.
Currently, the lines between the categories have become less distinct, but these categories are still
useful in describing base oils. The categories have been informally subdivided into Group I+, Group II+,
and Group III+. These oils have added quality but also are more expensive to produce.

* The API Base Stock Categories were established in the early 1990s to make it easier for lubricant blenders to interchange
one base stock for another. The original five categories were patterned on the base oils. The VI is an empirical, dimen-
sionless number that is an indication of the rate of change of the viscosity of oil with temperature. The higher the VI of
an oil, the less its viscosity will change with changes in temperature. The VI is calculated from the measured viscosity of
the oil at 40°C and 100°C according to ASTM Standard Test Method D2270 [1].
. il] (materials) A petroleum lubricating or process oil refined until its color (measured by transmitted light)
† Pale oil ['pāl o

is straw to pale yellow [2].


Base Oils 24-3

24.2.6  Group I + Base Oils


These still have higher sulfur and lower saturates, but processing conditions have been adjusted to pro-
duce base oils with higher VIs in the range of 100–105. Compared to Group I, these oils have better low
temperature properties and improved volatility performance.

24.2.7  Group II + Base Oils


These are produced by adjusting processing conditions to make oils with 110–120 VIs. This enables
significantly improved low temperature and volatility properties. These base oils are used in multi-vis-
cosity oils such as 5W-20 and 5W-30, GF-3 and GF-4 engine oils. As with Group I+, the extra quality
costs more to produce.

24.2.8  Group III + Base Oils


These have limited commercial availability. They can be made from gas to liquid processes. They have
VIs exceeding 140 similar to PAOs.
Recently, in Europe, a Group VI has been added. These base oils are similar to Group III+ and resem-
ble PAOs. They are designated as poly(internal olefins) (PIOs).

24.3  Refining
Base stocks are manufactured using a variety of different processes including but not limited to distil-
lation, solvent refining, hydrogen processing, oligomerization, esterification, and refining. In addition,
some base oils are re-refined. Refined base stocks shall be substantially free from materials introduced
through manufacturing, contamination, or previous use [3].
The first step in processing a crude oil is usually distillation that separates the crude oil into frac-
tions that are further processed into useful products. Figure 24.1 is a simplified diagram of distilla-
tion of quality Pennsylvania crude oil. Crude oils vary and distillation will produce different amounts
of specific fractions depending on the composition of the crude (Figure 24.2). The refinery processes
are designed to optimize specific products such as gasoline, aviation fuels, lubricant fractions, etc. The
degree of processing increases as product quality increases. In producing lubricant base stocks, the gas
oil fraction undergoes increased treatment depending on whether the product is a Group I, II, or III base
oil (Figure 24.3). There are numerous physical and chemical processes involved in producing good base
oils. The processes are optimized to produce a commodity for a specific application.

Gases

Gasoline

Solvents (Naphtha), jet fuel, kerosene (No. 1 Diesel)


No. 2 Diesel fuel (C16 Hc avg. range C12–C20)
Light lubricant fraction (80 neutral)
Medium lubricant fraction (120 neutral)
Heavy lubricant fraction (160 neutral)
Bright stock
Bottoms—asphaltenes (aromatic) or resins (paraffinic)

FIGURE 24.1  Petroleum distillation fractions.


24-4 Lubricants

100%
Gasoline
90%
80%
Kerosene
70%
60%
Light gas
oil 50%
40%
Heavy lube
oil 30%
Reduced 20%
crude 10%
0%
Pema. E. Texas Arabia Venezuela
Sulfur crude/residue 0.1/0.2 0.4/1.1 1.4/3.3 2.4/3.6
Res. octane no. gasoline 49 – 59 58 – 65 43 – 52 66 – 73

FIGURE 24.2  Typical yields of products, circa 1980.

Lubricant base oil production


Solvent Mild
Distillation extraction Dewaxing hydrofinishing
Crude oil
Gas

Oil Group I

Hydrocracker Dewaxing High pressure


Distillation Distillation hydrotreater
Crude oil
Gas L.O.
Group
Oil L,M,H II
Distillation Hydrocracker Wax High pressure
converter hydrotreater Distillation
Crude oil Gas L.O.
Oil L,M,H Group
III
Hydroisomerization

FIGURE 24.3  Simplified refinery layout.

General processing steps include


• Distillation—Involves separation by boiling point of the various fractions of the crude.
• Solvent or liquid extraction—Separates materials based on their relative solubility in two different
immiscible liquids. The goal is to reduce aromatics and wax in the base stock stream. Solvents
used in industry include propane, methyl ethyl ketone, and toluene.
• Hydroprocessing—Utilizes catalytic separation manipulation or hydrocracking technology to
reduce sulfur compounds, saturate double bonds, reshape molecules, and increase isoparaffin
content. Industrial processing in refineries started in the 1970s.
• Severe hydroprocessing—Cracking and restructuring of waxes (high purity waxes, Fischer–
Tropsch gas-to-liquid waxes), reduces sulfur, nitrogen, and oxygen content of the product.
Group I is still the most used base oil worldwide but its market share is slowly shrinking as product
performance requirements increase. Refining simply involves distillation followed by solvent extraction
of the vacuum gas oil to remove impurities and wax. Solvent extraction technology is an old process.
Propane was an early extraction solvent that was gradually replaced by methyl ethyl ketone, toluene, or
other solvents after World War II. Group I processing is the most cost-effective solution for many lubri-
cants (monograde engine oils, industrial lubricants, metalworking oils, etc.). Slight hydrotreating of the
Group I base stocks will result in improved quality Group I+ base oils.
Group II requires more expensive processing. Hydrocracking is used to reshape the gas oil mole-
cules and the product is catalytically dewaxed and hydrofinished using high pressure hydrotreating.
Base Oils 24-5

TABLE 24.2  Global Base Oil Refining 2006 and 2008 (Groups I, II, III, and Naphthenic Oils)
No. of Refineries
Location 2006 2008 2006 bbl/day 2008 bbl/day Change bbl/day
North America 24 27 254,800 264,700 9,900
South America 10 10 53,630 38,970 (14,660)
South Asia/Pacific/Australia 13 16 86,500 92,100 5,600
Western Europe 21 32 140,460 155,950 15,490
Central and Eastern Europe 24 23 138,450 132,835 (5,615)
China 20 22 85,220 86,990 1,770
Japan and Korea 12 14 96,650 109,550 12,900
Middle East and Africa 21 21 68,160 68,110 (50)
Source:  Lubes ‘n’ Greases, Special Edition, LNG Publishing Co., Falls Church, VA, 2006 and 2008.

In the process, paraffinic wax molecules are converted into isoparaffins and 98+% of the impurities are
removed, producing high-quality base oil with excellent additive susceptibility. Research by Fenske and
coworkers at Penn State in the 1960s produced “super-refined” base oils that were the early precursors to
Group II [4]. The major steps involved in the super-refining process included (a) hydrogenation and/or
acid extraction to remove polar impurities and saturate double bonds in hydrocarbons, (b) distillation
to produce a narrow boiling fraction, and (c) deep dewaxing to provide a wide liquid range [5].
Group III requires even higher severity hydroprocessing and requires high purity wax as a feed-
stock. The product has very low sulfur and oxidation susceptibility is better than Group II resulting in
improved lubricant formulations with excellent oxidation resistance. Group III+ has 140 or higher VI
and will compete strongly with Group IV base stocks. The cost in producing Group III base stocks is
high but less than producing Group IV base stock.
Group IV consists of synthetic hydrocarbons that are produced from catalytic polymerization of
n-alphaolefins. The chemical process starts with low-molecular-weight olefins, such as ethylene, to pro-
duce higher-molecular-weight alpha-olefins. PAO is synthesized by catalytic reaction of a 10-carbon
alpha-olefin to produce a series of unsaturated PAOs that are further refined by hydrogenation to pro-
duce isoparaffinic hydrocarbons that are separated into fluids ranging in viscosities from 2 to 100 cSt at
100°C and possessing high VI and excellent low temperature properties.
Group V contains many different compounds produced by various synthetic methods for specific
applications. In addition to synthetic fluids, Group V contains base stocks from natural sources that
undergo limited refining. Fluids in Group V include alkylated aromatics, organic esters, diesters and
polyol esters, polyglycol ethers, phosphate esters, silanes, silicones and silicate esters, halogenated
hydrocarbons, pale oils, polyphenyl ethers, and natural oils from renewable resources.
Worldwide refining capacity in 2006 and 2008 for petroleum base stocks is found in Table 24.2 [6].
Refineries of <1000 bbl/day are not included. The world increase was 25,395 bbl/day over the period with
only two areas showing a decrease in production. The production of nonpetroleum base stocks varies in
size from small batch to large facilities. Table 24.3 contains a few of the nonconventional base stocks [7].
The U.S. capacity for Group IV base stocks approaches 50% of the worldwide capacity. The production
of Group V fluids includes fluids, such as silicones, phosphate esters, etc., that are used as additives as
well as base stocks.

24.4  Base Oil Applications


The primary use of Groups I, II, and III are in automotive applications such as motor oils, hydraulic
fluids, transmission, process oils, industrial oils, and gear oils [8]. Automotive oils make up about half
of some 10 billion gallons of lubricants sold globally annually. Some 2 billion gallons of total lubricants
are marketed in the United States and Canada. Another 1.5–2 billion gallons are sold in the European
24-6 Lubricants

TABLE 24.3  Nonconventional Base Oils


Companies,
Base Stocka Worldwide Capacity Comments
Naphthenic 19 82,165 bbl/day USA—38,000 bbl/day
Group III 21 62,580 bbl/day USA—2,500 bbl/day
GTL 1 28,800 bbl/day Shell, Malaysia
PIB 21 944,000 tons/year (19,077 bbl/day)b USA—437,500 tons/year (8,841 bbl/day)
PAO and PIO 10 472,300 tons/year (9,544 bbl/day) USA—215,000 tons/year (4,345 bbl/day)
PAG 42 713,200 tons/year (14,412 bbl/day) USA—442,000 tons/year (8,932 bbl/day)
Esters 52 Not available Esters usually made in batches of 1–20 tons
Phosphate esters 15 160,000 tons/year (3,233 bbl/day) USA—40,000 tons/year (808 bbl/day)
Source: Lubes ‘n’ Greases, Special Edition, T. Sullivan, Editor, Lube Report (ISSN 1547-3392), Lubes ‘n’ Greases
Magazine and Lubricants Industry Sourcebook are published by LNG Publishing Co., Falls Church, VA, 2009. www.
LNGpublishing.com
a GTL, gas to liquid; PIO, poly(internal olefins); PIB, polyisobutylene; PAO, polyalphaolefin; PAG, polyalkylene glycols.

b Calculated (42 gals/bbl × 3.78 L/gal × 1000 cc/L × 0.88 g/cc × 1 kg/1000 = 139 kg/bbl; 944,000 tons/year × 1000 kg/ton × ​

1 bbl/139 kg × 1 year/356 days = 19,077 bbl/day).

market and over 4 billion gallons in the Asia Pacific region. The remaining 2 billion is marketed in the
Middle East, Africa, and Latin America. Industrial oils make up the next largest market. The use of base
stocks in automotive applications is governed by several different organizations.
For passenger cars, light duty and heavy duty vehicles, the API has two general classifications for
engine oils: S for spark ignition or “service” and C for compression ignition or diesel vehicles. API,
The Society of Automotive Engineers (SAE), Engine Manufacturers Association (EMA), and American
Standards and Testing Materials (ASTM) [1,8–10] are involved in the process of developing perfor-
mance standards in the United States.
The Coordinating Research Council (CRC) is a nonprofit organization that directs, through com-
mittee action, engineering and environmental studies on the interaction between automotive/other
mobility equipment and petroleum products [11]. CRC research programs are managed by five technical
committees, including the advanced vehicle, fuel, and lubricants committee. The Sustaining Members
of CRC are the API and a group of automobile manufacturer members (Chrysler, Ford, General Motors,
Honda, Mitsubishi, Nissan, Toyota, and Volkswagen) and the EMA.
The International Lubricant Standardization and Approval Committee (ILSAC) was jointly developed
and approved by the Japan Automobile Manufacturers Association, Daimler Chrysler Corporation,
Ford Motor Company, and General Motors Corporation as an international organization for the world-
wide development of standards for motor oils.
The Association des Constructeurs Européens d’Automobiles (ACEA) is a European counterpart to
the CRC body for fuel and lubricant testing and setting of standards in Europe.
The Japanese Automotive Standards Organization (JASO) is the Japanese counterpart to the SAE.
JASO sets the performance and quality standards for Japan and the Far East.

24.5  Performance
The performance of the base stocks in formulations depends on the specific use and the requirements for
the application. Other chapters in this book discuss these in detail. Other excellent sources of information
are found in References [12–15]. These are books and the editors are shown in the references. Table 24.4 is
a summary of the general ranking of the properties of various base oils neat (without additives). Group V
synthetic base oils include diesters, polyol esters, perfluorinated aliphatic esters, phosphate esters, polyg-
lycols, polyol ethers, polyphenyl ether, and silicones. The principal attributes of these fluids when used as
base stocks as reported by Klaus and Tewksbury [14] remain essentially the same as found in Reference [14]:
Base Oils 24-7

TABLE 24.4  Base Oil Characteristics—Groups I–V


Group Group
Property Group I Group II Group III Group IV V—Synthetics V—Natural Oils
Viscosity range Very good Very good Very good Good Poor–excellent Excellent
Viscosity index Moderate Moderate Good to Very good Poor–excellent Excellent
to good to good excellent
Pour point Good Good Good Very good Good–excellent Poor
Oxidation stability Moderate Moderate Moderate Good Moderate–excellent Moderate–poor
Thermal stability Moderate Moderate Moderate Very good Good–excellent Good
Hydrolytic stability Excellent Excellent Excellent Excellent Moderate–excellent Moderate
Corrosion prevention Excellent Moderate Moderate Excellent None–excellent Moderate
Lubrication Good Moderate Moderate Good Poor–excellent Very good
Volatility Moderate Moderate Moderate Good Good–excellent Good
Fire resistance Poor Poor Poor Moderate Moderate–excellent Moderate
Elastomer swell Low Moderate Moderate Shrink Shrink–low– Moderate
(BUNA) moderate
Paint and varnish Very good Good Good Excellent Poor–excellent Moderate
compatibility
Additive solubility Excellent Very good Very good Very good Negligible–very Very good
good
Improvement with AO Good Very good Very good Very good Negligible–very Very good
and AW additives good
Cost Low High Very high High Moderate–very high Moderate
Source: Handbook of Lubrication and Tribology, Volume II, Theory and Design, E.R. Booser, Editor, CRC Press, Boca
Raton, FL.

Group I, mineral oil Low cost


Groups II and III, mineral oil Improved performance
Group IV, synthetic hydrocarbons Low temperature fluidity
Group V
Vegetable oils Renewable resource
Organic esters Low temperature fluidity
Polyglycol ethers Excellent VI properties
Phosphate esters Low flammability
Silicones Excellent VI properties
Polyphenyl ethers Thermal stability
Perfluoropolyethers Oxidation stability
Halocarbons Nonflammable

The physical and chemical properties of these liquid lubricants such as viscosity, viscosity–temperature
and pressure properties, viscosity–shear relationships, bulk modulus, density, vapor pressure, thermal,
thermal stability, low temperature, electrical, surface tension properties are found in the chapter on
liquid lubricants in Reference 14.
Probably the most significant changes in base stock use in the past decade have been the use of Groups
II and III base stocks and the use of oils from renewable resources to replace the use of petroleum base
oils. Significant progress has been made in improving the quality and performance of engine oils by the
use of Groups II and III petroleum base oils while vegetable oils have made penetration in some markets
with the exception of automotive engine oils.
24-8 Lubricants

The use of vegetable oils to replace petroleum products has to overcome two major drawbacks: poor
low temperature properties and poor oxidation stability. The low temperature problem is improved by
the use of pour point depressants and blending vegetable oils with synthetic esters or low-molecular-
weight PAOs. Improving oxidation stability is a more difficult challenge but can be improved by the use
of the proper selection of additives, chemical modification of the triacylglycerides, or genetic modifica-
tion to reduce the unsaturation of the free fatty acid chains of the triacylglycerides.
The use of base fluids from renewable resources has been stimulated by the USDA Bio-Preferred
Program—Federal Bio-based Products Preferred Purchasing Program (FB4P). It was created in the
2002 Farm Bill, Section 9002. Subsequently, the USDA was directed to implement a program that estab-
lished guidelines and designates items for preferred purchasing of biobased products by government
agencies. All agencies are required to purchase biobased products that fall within designated item cat-
egories unless the products (1) are not reasonably available, (2) fail to meet performance standards, or
(3) are only available at an unreasonable price.
The USDA designates categories of biobased products for a Federal Procurement preference [16].
In the process, the minimum biobased content standards are established for each product category.
Currently, there are 64 bio-preferred designated product categories required for preferred Federal pur-
chasing. Under the category Industrial Lubricants, currently, there are some 239 products listed includ-
ing 71 metalworking fluids and 129 hydraulic fluids. The specific biobased content of these biobased
products has been established for a number of lubricants and ranges from 30% for greases to 95% for
transformer fluids (Table 24.5). Hydraulic fluids must contain 44% biodegradable fluids. The biode-
gradability of base oils in the CEC L33-A-99 test is found in Table 24.6. On the environmental side, the

TABLE 24.5  Required Biobased Content for Specific Lubricants


(CEC L33-A-99)
USDA Minimum Biobased
Lubricant Type Content (%)
Fluid-filled transformers 95
Multipurpose lubricants 88
Turbine drip oils 87
Chain and cable lubricants 77
Penetrating lubricants 68
Gear oils 58
Mobile equipment hydraulic fluids 44
Two-Cycle engine oils 34
Greases—rail track 30

TABLE 24.6  Base Stock Biodegradability—CEC L33-A-99


Base Stock % Biodegradation after 21 Days
Bright stock 5–15
Mineral oils 10–45
Hydrocracked mineral oils 25–80
White oils 25–45
PAOs 20–80
Polyols 5–100
Diesters 50–100
Vegetable oils 75–100
Source: Johnson, M. and Miller, M., Tribol. Lubr. Technol., 28,
28, 2010.
Base Oils 24-9

natural oils and some synthetics have a major advantage in biodegradability. The proper and careful
blending of base stocks can result in lubricants with outstanding performance.

References
1. ASTM International, West Conshohocken, PA.
2. McGraw-Hill Dictionary of Scientific & Technical Terms, 6E, Copyright 2003 by The McGraw-Hill
Companies, Inc.
3. API Document 1509, Appendix E.
4. Klaus, E.E., E.J. Tewksbury, and M.R. Fenske, Preparation, properties and some applications of super-
refined mineral oils, ASLE Transactions, 5(1), 1962, 115–127.
5. Klaus, E.E., E.J. Tewksbury, and M.R. Fenske, High-temperature hydraulic fluids from petroleum,
I&EC Product Research and Development, 2, 332, 1963.
6. Lubes ‘n’ Greases, Special Edition, L. Tocci (ed.), LNG Publishing Co., Falls Church, VA, 2006 and
2008.
7. Lubes ‘n’ Greases, Special Edition, T. Sullivan, Editor, Lube Report (ISSN 1547-3392), Lubes ‘n’
Greases Magazine and Lubricants Industry Sourcebook are published by LNG Publishing Co., Falls
Church, VA, 2009. www.LNGpublishing.com
8. American Petroleum Institute, NW, Washington, DC.
9. Society of Automotive Engineers, International, World Headquarters, Warrendale, PA.
10. Engine Manufacturers Association, Chicago, IL.
11. Coordinating Research Council, Alpharetta, GA.
12. Handbook of Lubrication and Tribology, Volume I, Application and Maintenance, E.R. Booser, Editor,
CRC Press, Boca Raton, FL.
13. Handbook of Lubrication and Tribology, Volume II, Theory and Design, E.R. Booser, Editor, CRC
Press, Boca Raton, FL.
14. Handbook of Lubrication and Tribology, Volume III, Monitoring, Materials, Synthetic Lubricants, and
Applications, E.R. Booser, Editor, CRC Press, Boca Raton, FL.
15. Synthetic Lubricants and High-Performance Functional Fluids, L.R. Rudnick and R. Shubkin, Editors,
Marcel Dekker, New York.
16. USDA BioPreferredⓇ Program. http:www.biopreferred.gov/Default.aspx
17. Johnson, M. and M. Miller, Eco-friendly fluids for the lubricants industry, Tribology and Lubrication
Technology, 28(10), 28–34, 2010.
25
Additives for Lubricants
25.1 Introduction..................................................................................... 25-1
25.2 Deposit Control Additives..............................................................25-2
Oxidation Inhibitors  •  Phenolics and Aromatic Nitrogen
Compounds  •  Hindered Phenols  •  Aromatic Amines  •  Zinc
Dithiophosphates  •  Dithiocarbamates  •  Sulfurized Paraffins,
Olefins, Carboxylic Acids, and Fats  •  Thiadiazoles  •  Selenides
25.3 Detergents.........................................................................................25-7
Metal Sulfonates  •  Overbased Metal Sulfonates  •  Metal Phenate
Sulfides  •  Metal Salicylates
25.4 Dispersants.......................................................................................25-8
Polyamine Succinimides  •  Polyamine Imidazolines
25.5 Film-Forming Additives.................................................................25-9
Antiwear and EP Additives  •  Friction Modifiers
25.6 Viscosity Control Additives.........................................................25-13
Viscosity Modifiers  •  Pour Point Depressant
25.7 Corrosion Inhibitors..................................................................... 25-14
Succinic Acids  •  Amine Phosphates  •  Aromatic Corrosion
Inhibitors  •  Metal Deactivators
25.8 Miscellaneous Additives............................................................... 25-16
Tackifiers and Anti-Misting Additives  •  Antifoam
Additives  •  Demulsifiers  •  Seal Swell Additives  •  Dyes,
Chemical Marking Agents, and Tracers  •  Food-Grade
Additives  •  Antibacterials/Biocides/Fungicides
Leslie R. Rudnick 25.9 Future Trends................................................................................. 25-18
Ultrachem, Inc. References................................................................................................... 25-18

25.1  Introduction
Lubricant additives can generally be considered chemical substances that are included in a lubricant for-
mulation to improve the performance or properties of the base fluids and/or to impart new properties or
performance features to the formulated oil. They can function by reducing the tendency of the base fluid
from undergoing deleterious processes such as oxidation and they can improve on limited performance
of the base fluid, for example, viscosity index, pour point, and wear protection.
Components that we currently call additives were originally used as the lubricant itself. For exam-
ple, animal fats (triglycerides) were used to grease the wheels of Roman chariots and to provide
friction reduction (friction modifiers) when moving or sliding heavy loads. The modern history of
lubricant additives began early in the twentieth century. The discovery of using drilling methods to
obtain petroleum, by Drake in 1859, and the development of refining processes ultimately provided
viscous liquids that could be used as lubricants and a new industry was born. The increased require-
ments being placed on lubricants created the need for additives that could improve the properties

25-1
25-2 Lubricants

and performance of the petroleum-derived lubricant fluids. Elemental sulfur and sulfur-containing
hydrocarbons and fatty esters and acids were used to improve the performance of lubricants to be used
under heavy loads [1].
There are far too many additives that have been used in lubricants to describe here. Many have been
described or referenced in recent books [2,3]. This chapter describes the most important classes of lubri-
cant additives and describes the more important structural types within these classes and refers the
reader to further literature. Additives can be pure chemical entities, for example, some phenolic anti-
oxidants, such as butyl hydroxy toluene (BHT), α-tocopherol (vitamin E), and the zinc dithiophosphates
(ZDDPs). Many other additives are complex mixtures, which can be essentially replicated in batch after
batch, and have a variety of active and inactive components.
Recent regulations, both national and international, are changing the range of additives that can
be used. Regulation, Evaluation, Authorization and Restriction of Chemicals (REACH) is a European
Community regulation of chemicals and their safe use. Under this regulation, manufacturers and
importers are required to collect information on the properties of substances and register the informa-
tion in the European Chemicals Agency (ECHA) database. Other countries, including the United States
and China, are moving toward their own versions of REACH.
Restriction of Hazardous Substances (RoHS), a directive, also adopted by the European Union,
restricts the use of hazardous substances used in the manufacture of electrical and electronic equip-
ment, such as cadmium, mercury, lead, hexavalent chromium, polybrominated biphenyls, and polybro-
minated diphenylethers.
California Proposition 65 is a Safe Drinking Water and Toxic Environment Act of 1986. The Prop­
osition requires the governor to publish a list of chemicals known by the state to either cause cancer or
have reproductive toxicity. This list must be published at least once annually. Components on the list
that are contained in lubricant formulations are required to be revealed on the Material Safety Data
Sheet (MSDS) for the product.
For the small manufacturer of lubricants, the quantities of a particular additive exported within a
calendar year will be sufficiently low that registration is not required. RoHS compliance requires docu-
menting the amount of any of the listed chemical substances within a lubricant formulation.
Additives that can be considered “food grade” are becoming more prevalent due to stricter regula-
tions in the food industry. Food grade does not mean that one can consume the additive as food but
rather that the additive in low concentration can be used in a lubricant that may have incidental contact
with food when used in a food-processing plant. In fact, Canada’s Canadian Food Inspection Agency
(CFIA) does not allow the term “food grade” on product labeling and prefers a statement, for example,
“for use in food-processing equipment.”
Additive toxicity and biodegradability are other factors that are considered to a greater extent in new
formulations.
The most important additive functions are controlling deposit formation, oxidation, corrosion, vis-
cosity, pour point, and other properties and performance features required to meet lubricant specifica-
tions or performance criteria. Other classifications based on lubrication regime have been described
[1]. Additives can be either chemically active or chemically inert. Chemically active additives may react
with metal surfaces to form protective films, or reactive species such as oxygen may react with additives
to render the reactive species inactive and therefore provide protection to the lubricant. Many additives
can have multiple functions.

25.2  Deposit Control Additives


Deposit control is important because deposit formation is usually the result of lubricant degradation.
Degradation can occur in several ways including oxidation of the oil and formation of insoluble compo-
nents. The insoluble components can be derived from oil breakdown or from reaction with or accumula-
tion of combustion products in the case of engine oils.
Additives for Lubricants 25-3

25.2.1  Oxidation Inhibitors


Since all hydrocarbons, natural and synthetic, oxidize under the conditions of use, antioxidants are
used in one form or other in virtually all oils, except those containing all C–F bonds where the C–H
bond has been replaced. These include engine and industrial oils. Common antioxidants used in lubri-
cants include phenolics, aromatic nitrogen-containing compounds, polysulfides, some heterocyclics,
transition metal peroxide decomposers, and a few specialized organometallic compounds.
The mechanism of oxidation of hydrocarbons as it relates to lubricant fluids has been extensively
studied and reported in the literature [1,4,5]. Oxidation is generally accepted to occur via free-radical
attack of oxygen or nitrogen oxides on the C–H bond of hydrocarbons to generate peroxy radicals that
undergo further deleterious reactions.
In practical terms, these deleterious chemical reactions result in oil thickening and varnish and
sludge formation. Acids are generally formed and can result in further degradation of the oil and odors
making the oil unfit for use.

25.2.2  Phenolics and Aromatic Nitrogen Compounds


These additives generally function as radical acceptors and transfer a hydrogen atom from the phenolic
oxygen or amine nitrogen to the hydrocarbon peroxy radical [6]. These radicals then react by radi-
cal recombination or electron transfer to give ionic compounds by formation of quinines and quino-
neimines. They may also react to terminate the radical chain mechanism often resulting in ketones,
aldehydes, and polyaromatic compounds. Sulfur- and phosphorus-containing antioxidants function by
decomposing peroxides [7].
The mechanism of oxidation as it relates to lubricants and additive chemistry has been recently
described [8].
Initiation:

RH →R⋅
Chain propagation:


R ⋅ + O2 → ROO ⋅


ROO ⋅ + RH → ROOH + R ⋅
Chain branching:


ROOH → RO ⋅ + ⋅ OH


RO ⋅ + RH → ROH + R ⋅

⋅OH + RH →H O +R⋅
2

Chain termination:


2R⋅ → R–R

R ⋅ + ROO ⋅ → ROOR
25-4 Lubricants


2ROO ⋅ → ROOR + O 2

Some of the main types of commonly used oxidation inhibitors are described in the following sections.

25.2.3  Hindered Phenols


Hindered phenols are one of the most diverse groups of compounds used as antioxidants. Hindered
phenols are commercially available as monocyclic and polycyclic structures. One of the most com-
mon hindered phenolic antioxidants is 2,6-di-tert-butyl-4-hydroxytoluene (BHT), a common additive
in food products and widely used in lubricants. Side chains are added to convert some of the phenolic
antioxidants from solids to liquids to facilitate formulation in commercial plants and to decrease the
volatility for use in high temperature applications. Some examples of phenolic and aromatic amine
antioxidants are shown as follows:

CH3 OH H3C
H 3C CH3
C C

H 3C CH3

CH3

2,6-Di-tert-butyl-4-methyl phenol

HO OC8H17

3,5-Di-tert-butyl-4-hydroxyhydrocinnamic acid, C7–C9-branched alkyl ester

OH

HO O OH

OH
S

2,2′-Thiodiethylene-bis[3-(3,5-di-tert-butyl-4-hydroxyphenyl)proprionate]
Additives for Lubricants 25-5

HO OH

4,4′-Thiobis(2-tert-butyl-5methyl phenol)

25.2.4  Aromatic Amines


Aromatic amines are some of the most effective antioxidants used in automotive engine oils and in
industrial lubricants. These materials are generally more effective for lubricants subjected to higher
temperatures. Aromatic amines can impart some color to the lubricant and generally darken due to
oxidation and exposure to light. Some examples include

Phenyl-α-naphthylamine

R N R

Alkylated diphenyl amine, R = typically C8 or C9

Phenothiazine
Phenolic sulfides, metal-containing phenolic sulfides, and metal salicylates are also used as antioxidants.

25.2.5  Zinc Dithiophosphates


ZDDPs are considered the most important and cost-effective group of antiwear additives. ZDDPs
have been extensively studied [5]. These additives also serve as antioxidants in lubricant formula-
tions. Secondary ZDDPs provide better wear protection than primary ZDDPs whereas primary
25-6 Lubricants

ZDDPs are better antioxidants than secondary ZDDPs. Aryl ZDDPs have excellent thermal stability
but are not as good as the primary or secondary ZDDPs as antioxidants or for wear protection. ZDDP
is proposed to function through deposition of a protective layer that wears away and gets replenished
during use.
Metal dialkyldithiophosphates (metal dialkyl phosphorodithioates) [1]

S S
R O O R

P M P

S S
R O O R

Zinc dithiophosphate, where M = zinc

25.2.6  Dithiocarbamates

S R

C N

M S R
X

Metal dithiocarbamate

25.2.7  Sulfurized Paraffins, Olefins, Carboxylic Acids, and Fats


S
R Sx
R S H R R
S S

C
HC
S
S
S
HC
C
R C

S H R R R

25.2.8  Thiadiazoles
N N

H
S C C
S
R S S

2-Alkyl-4-mercapto-1,3,4-thiadiazole
Additives for Lubricants 25-7

25.2.9  Selenides

H25C12 – Se – C12H25

Dilauryl selenide

25.3  Detergents
Detergents are polar molecules that provide a cleaning function, generally by removing substances
(deposits) from the metal surfaces [9,10]. Detergents and dispersants both have some of the same ability
to do both functions, surface cleaning and suspending contaminants, and perform these functions to
different extent depending on the structure and the polarity of the oil and contaminant. Some deter-
gents actually provide some antioxidancy. The metal-containing detergents produce ash. Metals can
include barium, calcium, lithium, potassium, or sodium [11].
Several classes of detergents include the following.

25.3.1  Metal Sulfonates


O O
R O O R
S M S

O O

O
O O
R
S M H

25.3.2  Overbased Metal Sulfonates


O O
R O O R
S M CO3 M S

x
O O y

25.3.3  Metal Phenate Sulfides


A M M A
M O O
O O

Sx

Sx

R R R R
25-8 Lubricants

25.3.4  Metal Salicylates

R O

C O

Metal thiophosphonates and overbased metal phenate sulfides are additional classes of additives used
as detergents [16]. Of all these types, the most common detergents are the sulfonates, phenates, and
salicylates.
Detergents are used in engine oils, automatic transmission fluids, industrial hydraulic oils, and trac-
tor hydraulics oils. Certain metalworking fluids contain detergents. Detergents can be either neutral or
basic (overbased). Neutral detergents are prepared by combining the acid component with a stoichio-
metric amount of base, whereas the overbased detergent is prepared with more base than needed in the
presence of carbon dioxide to produce a complex containing extra available base. Neutral detergents
form a film on the metal surface and inhibit rust and corrosion, while overbased detergents form micel-
lar structures around CaCO3. Due to their structure, detergents can thicken or gel in the presence of
water or acidic components in the oil.
In use, there are/can be stability issues when used for long periods of time. Detergents as well as
other additives can form deposits in high temperature applications or when high temperature excur-
sions occur in equipment due to faulty operation.
Sulfonates exhibit excellent detergency and can be overbased resulting in very high total base num-
bers (TBN) [17]. Phenates exhibit a good compromise of detergency and antioxidancy, but cannot be
overbased to the extent of sulfonate detergents. Salicylates exhibit both antioxidant performance in
addition to their detergent properties [18]. Salicylates have lower TBN than the other classes of deter-
gents [17].

25.4  Dispersants
Dispersants are polar molecules that provide a cleaning function, generally by keeping contaminants
and insoluble oil components suspended in the lubricant [19]. Dispersants function by keeping equip-
ment parts clean because the soots and sticky particulates and oxidized chemical components of the
oil or fuel are suspended rather than being allowed to deposit onto metal surfaces. By minimizing the
amount of agglomeration of particles, dispersants help to maintain oil viscosity rather than to allow the
normal viscosity thickening due to coalescence of particles [20]. Dispersants reduce the tendency for
motor oils to produce sludge.
Dispersants are included in lubricant formulation to retard the formation of deposits at low operat-
ing temperatures. They are generally “ashless,” meaning that their structures do not contain metals.
Dispersants function by having a polar end group and a nonpolar hydrocarbon tail [21]. Dispersant
function is provided by polar end group association with lubricant degradation products and any
combustion byproducts and soot. By maintaining these associations, the formation of deposits is
minimized.
Some common structural types of dispersants are as follows.
Additives for Lubricants 25-9

25.4.1  Polyamine Succinimides


O

R NH
N NH2

X
O

Alkyl polyamine succinimide

25.4.2  Polyamine Imidazolines


R

R C O R

N
N N N N

Polyhydroxy succinic esters and hydroxy benzyl polyamines are also used as dispersants.
The mechanisms of soot formation and particle agglomeration, especially as they relate to diesel engine
soot, have been extensively studied and reported in the literature [23–25]. Dispersants are important
additive components of diesel engine oils because of the necessity of keeping soot dispersed so that it
can be filtered out rather than depositing in the engine.
Polymethacrylate viscosity modifiers can also provide dispersancy in formulated lubricants. The
advantage of using polyfunctional viscosity modifiers is that formulations can be prepared using
fewer and lesser quantities of additives. This reduces additive–additive interactions and can improve
economics.

25.5  Film-Forming Additives


Film-forming additives are used either to reduce wear due to contact between asperities of two moving
surfaces or to reduce friction thereby lowering the heat generated due to the rubbing of the two surfaces
in the contact zone [26]. Antiwear and extreme pressure (EP) additives are used in automotive and
industrial applications. Antiwear additives are used to prevent ongoing wear caused by light or moder-
ate metal contact whereas EP additives are used to prevent seizure of surfaces caused by severe metal-
to-metal contact [27]. As mentioned earlier, zinc dithiophosphorate (ZDDP) is one of the most common
antiwear additives and is commonly found in motor oils. The phosphorus in ZDDP can adversely affect
metals used in catalytic converters. Antiwear additives are used in compressor oils, gear oils, antiwear
hydraulic oils, high-temperature chain oils, and many greases.
Wear is the loss of material from at least one of the two contacting tribosurfaces caused from the
mechanical action and/or relative motion of the surface(s). Wear mechanisms include adhesive and
abrasive wear, surface fatigue, corrosion, and fretting wear [28]. The microstructure of the surface is an
important factor in the process of wear. Within the uppermost few millimeters of the surface, several
phenomena occur that affect wear. In the top 5 Å, water and oxygen can be adsorbed. There is almost
always a metal oxide layer of up to ∼100 Å. Work hardening the surface affects the microstructure of the
surface with grain dislocations and deformations. Additional effects can come from the surface finish
of the substrate.
25-10 Lubricants

25.5.1  Antiwear and EP Additives


25.5.1.1  Antiwear Additives
Antiwear and EP additives provide protection by creating a thin film of the additive or a reaction prod-
uct that can be removed by shearing during operation of the equipment. These additives are most effec-
tive when the equipment is being operated under heavy load. These additives also protect against shock
load and wear caused by operation at high temperatures. Examples include zinc dithiophosphorates,
molybdenum and tungsten disulfides, and tricresylphosphate (TCP).
Antiwear additives have been synthesized to contain a wide variety of chemical structures and chemi-
cal elements; however, the most common elements used are phosphorus, sulfur [29–31], and combi-
nations of these elements [32]. These additives are used in many applications, including engine oils,
transmission fluids, and industrial fluids. Gear oils, some compressor oils, chain oils, hydraulic fluids,
and greases are commonly formulated with antiwear additives.
There are four main groups of commercially available EP additives. These are based on structures
containing phosphorus, sulfur, chlorine, and overbased sulfonates [33]. Phosphorus-, sulfur-, and
­chlorine-containing EP additives are activated by heat over a range of temperatures. Chlorine-containing
EP additives are typically activated between 180°C and 420°C. Phosphorus-containing additives oper-
ate at even higher temperatures and sulfur-containing additives operate in the range of 600°C–1000°C.
The species formed are iron chlorides, phosphates, and sulfides, respectively. These compounds form a
lubricious chemical film or layer that protects the metal surfaces and reduce friction and the probability
of welding.
Overbased sulfonates contain a colloidal carbonate that reacts with iron to form a thin-film barrier
layer between the tribocontacts, thereby protecting each surface from direct contact and welding [33].
Phosphorus-containing organic compounds are known to exhibit antiwear and/or load-carrying
performance when used as additives in lubricant formulations [34,35]. It is generally accepted that
the mechanism of action for triaryl phosphates is that they form a protective film composed of iron
­phosphate [33]. Dialkyl hydrogen phosphates, which also provide antiwear performance, have also been
reported to form an iron phosphate protective film [36].
Minami has shown that for di-n-hexyl and di-cyclohexyl hydrogen phosphates, the structure of
the phosphite is important. Di-n-hexyl hydrogen phosphite did not show good antiwear performance
whereas di-cyclohexyl hydrogen phosphite exhibited excellent performance [34].
TCP specifically reacts with iron to form compounds that serve as a friction modifier [37]. Another
approach, vapor phase lubrication, is becoming increasingly important in applications such as low heat-
rejection engines and where temperatures are in excess of 500°C, and also in microelectromechanical
systems (MEMS) which are negatively affected by liquid capillary condensation effects [38].

O P O

Tricresyl phosphate (TCP)


Additives for Lubricants 25-11

Phosphate esters are generally monoesters and diesters. These function by association of the polar phos-
phate group with the polar metal surface with long-chain hydrocarbon groups oriented into the bulk
liquid. This is similar to the function of sulfurized olefins where the sulfur is oriented on the metal
surface with the hydrocarbon chains oriented into the bulk lubricant.

25.5.1.2  Boundary Additives


Boundary additives are polar molecules, where the polar head of the additive associates with the metal
surface and the nonpolar portion of the molecule prefer to be oriented into the bulk lubricant. Boundary
antiwear additives function by first adsorbing onto the metal surface. This can occur either physically
or chemically depending on the chemical structure of the additive and the temperature of operation.
The film can prevent adhesive wear and metal contact through the film formation. These additives func-
tion to mitigate wear caused during boundary lubrication where higher speeds and loads cause asperity
contact.
Boundary additives are temperature-dependent or non-temperature dependent. For those bound-
ary additives that are temperature dependent, the reaction of the additive and the metal surface occurs
through temperature of operation and by the high temperatures produced due to heat caused by fric-
tion. Boundary additives based on phosphorus, sulfur, boron, and chlorine are activated by temperature
[39]. Overbased sulfonates operate by a different mechanism. Colloidal carbonate salts dispersed in the
sulfonates form a barrier film that protects the metal surface. This can occur at lower temperatures and
does not require heat to activate the process [39].
A variety of sulfur-containing additives is available [40]. The most commonly used are the sulfurized
olefins. Depending on the degree of sulfurization, these additives can be used with different metal-
lurgy. The use of certain sulfurized olefins also depends on the temperature and loads experienced in
the application. Many polysulfides decompose and promote wear or corrosion due to acidic species that
are produced. Thermal activity of trisulfides and pentasulfides are different. Trisulfides generate active
sulfur more slowly and at higher temperatures than pentasulfides. These are typically used in industrial
and automotive gear oils and in high-speed metalworking. Pentasulfides are used in low-speed metal-
working. Sulfurized and chlorinated additives can provide synergy in performance.

25.5.2  Friction Modifiers


Friction modifiers function by reducing frictional heat and, therefore, reduce the tendency for metal
contact to cause welding and adhesive wear. They compete for the surface with antiwear additives and
other polar additives and base fluids. Although there have been engine design improvements such as
surface modification, new materials, and the use of coatings, there is still a requirement for friction
modifiers as components in the formulated lubricant. Friction reduction improves operating efficiency
and is of benefit in automotive pumps, bearings, at pistons and piston rings. Friction modifiers are most
effective in boundary and mixed lubrication regimes. Friction modifiers improve fuel economy and
energy efficiency in engine oils and industrial oils, respectively.
Friction modifiers compete with polar additives and base oils for the surface. Dispersants, detergents,
EP, and corrosion inhibitors are all surface-active and, therefore, can affect the performance of friction
modifiers and vice versa. Examples of friction modifiers include long-chain fatty acids, such as oleic acid
[41], and esters, and glycerol derivatives.

25.5.2.1  Organic Friction Modifiers


Organic friction modifier structures include fatty acid esters or carboxylic acids and derivatives, amines,
and amides [42]. Phosphoric acid and phosphonic acid derivatives have been used and various organic
polymers have also been used as friction modifiers.
These materials function in different ways. Fatty acids and the phosphorus-containing acids func-
tion by formation of reacted layers on the tribosurface. Esters, amides, ethers, and imides function by
25-12 Lubricants

adsorbing on the surface in layers. Unsaturated acids, sulfurized olefins, and methacrylate polymers
form polymers on the surface.

25.5.2.1.1 Fatty Acids and Esters


Fatty acids and esters operate by adsorption onto the metal surface and provide a physical barrier
between the metal surfaces. These molecules have been shown to orient where the polar head is associ-
ated with the metal surface with the long hydrocarbon tail away from the surface. For linear hydro-
carbon chains, the packing density is improved because the molecules can align closely spaced to the
neighboring molecule. These are generally used in lower severity applications where the shear forces due
to load are not high. A common example is glycerol monooleate (GMO).
Fatty acid ester additives have been used in metalworking fluids to improve corrosion protection
and lubricity. These functions are important to preserve tool life and to minimize part rejection due
to insufficient quality of the machined part. Structures of fatty acid esters incorporate surface-active
functionality to the fatty acid molecule [43]. These materials are being used to replace the more toxic
chlorinated paraffins (CPs) [44].
Amides derived from fatty acids and GMO are examples of commercially available friction modifiers.

25.5.2.1.2 Chlorinated Paraffins


CPs used in lubrication are generally straight-chain aliphatic hydrocarbons having paraffinic chain
lengths ranging from C10 to C30. These CPs are generally classified as having paraffinic chain lengths
of short-chain C10–C13 (SCCP), mid-chain C14–C17 (MCCP), and long-chain C18–C30+ (LCCP). In North
America, mid-chain CPs are the most used with long-chain CPs the next most used.
CPs are generally declining in use mostly due to regulatory issues.
“Chlorinated paraffins remain excluded from U.S. federal hazardous waste regulations, however, for
example, in the state of Washington CPs in waste oils must be managed as ‘hazardous waste’ ” [45]. This
is partly due to the fact that EPA concluded, based on a 1990s study of SCCPs, that there was no need
to restrict use. They were, however, added to the Toxic Release Inventory in order to track future envi-
ronmental impact.
In Canada, the Environment Canada (EC) proposed that all CPs be added to the CEPA Toxic list
and also put all CPs up to C20 on the “Virtual Elimination” list. In the European Union, SCCPs used in
metalworking fluids have been added to the list of substances that must be authorized under the new
European REACH regulations. In the United Kingdom, a study concluded that there is no single sub-
stance that could adequately replace MCCP use in metalworking applications [45].

25.5.2.1.3 Carbon and Carbon Blacks


Carbon in one form or another has been used as a lubricant additive for many years. Various grades and
forms are available. Carbon black and graphite are black and insoluble in oil and are used especially in
lubricants for high-temperature applications where the lubricant is expected to burn off or evaporate.
The remaining graphite then serves as the lubricant [46]. More recently, other layered materials such as
boron nitride and tungsten disulfide with antiwear properties are being used.

25.5.2.2  Inorganic Friction Modifiers


25.5.2.2.1 Boron Nanotechnology
Boron nitride is well-known to provide reduction in friction and also serves as an effective antiwear
component in industrial lubricant formulations [47]. There are various particle sizes available com-
mercially and some varieties are classified as HX-1 for food processing applications. Boron, in the form
of nanosized (50–100 nm) boric acid, has been reported to be more easily dispersed in mineral oil base
fluids than micron-sized boric acid [48]. More recently, Canter has reported on the use of potassium
borate nanoparticles as a lubricant additive. Whereas nanoparticulate boric acid can agglomerate when
Additives for Lubricants 25-13

exposed to water, this additive dispersion in oil was unaffected by water [49]. It should be noted that
boric acid and certain other borate compounds are to be classified as Category 1B Reproductive Toxins
by the UEIL Health and Environment Committee as part of REACH. Boric acid is used as the precursor
to a variety of metalworking additives, although they can be present in very low concentrations [50].

25.5.2.2.2 Molybdenum Compounds


Various molybdenum compounds are used commercially for friction reduction in lubricants. Molybdenum
disulfide and molybdenum dithiocarbamate (MoDTC) are the most common [51]. Molybdenum disulfide
is insoluble in oil, whereas metallo-organic compounds are partially or highly soluble in lubricant oils.
MoDTC is the most common of the soluble molybdenum compounds and is used as lubricant addi-
tives. MoDTC is most likely active as molybdenum disulfide, which is produced in situ at the elevated
contact temperatures experienced at the tribocontact. MoDTC has also been found to have a synergistic
effect with ZDDP by ligand exchange, where the exchange products may be more effective than the
MoDTC [52]. These synergies are temperature dependent:

Mo(DTC)2 + Zn(DTP)2

Mo(DTC)(DTP) + ZN(DTP)(DTC)

Mo(DTP)2 + Zn(DTC)2

25.5.2.2.3 Tungsten Disulfide


Tungsten disulfide is the most lubricious material known. Tungsten disulfide in powder form has a
dynamic coefficient of friction of 0.03 and a static coefficient of friction of 0.07 [53].

25.6  Viscosity Control Additives


25.6.1  Viscosity Modifiers
Viscosity modifiers are used to provide multigrade oils in automotive applications, and to increase the
viscosity and viscosity index of industrial lubricants [54].
Viscosity modifiers work by improving the relationship between viscosity and temperature [55]. The
viscosity of a typical petroleum-derived base fluid varies inversely with temperature. When this occurs
to an appreciable extent, the viscosity of the lubricant is significantly different at low operating tempera-
tures than it is at high operating temperatures. To minimize this effect, viscosity modifiers are used. The
materials used are various polymers and copolymers that have the ability to coil under some conditions
and lengthen at other temperatures. At high temperatures, the viscosity index improvers (VII) uncoil
and increase the oil viscosity. At low temperatures a polymeric VII is coiled and, therefore, does not
impart much to the viscosity of the fluid. Thus, the viscometric difference between low and high tem-
perature is reduced and the fluid changes less with temperature. Examples of common commercially
available viscosity modifiers include as follows.

25.6.1.1  Polyisobutylenes [56]


CH3

H2
C C

CH3 X

Polyisobutylene
25-14 Lubricants

25.6.1.2  Olefin Copolymers


Ethylene–propylene copolymers are also used as viscosity modifiers [54,57]. These additives can vary
in the ratio of ethylene to propylene and also in molecular weight. Higher-molecular-weight versions
(50,000–200,000 MW) are more viscous and provide a greater degree of thickening. Functionalized
olefin copolymers have also been reported and provide dispersancy [58,59].

25.6.1.3  Polymethacrylates [60]


CH3
H2
C C
X
C

O O

Esters of polymethacrylic acid


Styrene butadiene copolymers and styrene maleic ester copolymers are also used as viscosity modifiers.
The former are hydrocarbons whereas the later have functionality and are capable of thickening and
have some dispersant functionality.
This is important in automotive applications for low temperature start-up in winter where the oil
would normally be quite viscous. Thickening efficiency is directly related to the molecular weight of
the viscosity modifier. Although not entirely dependent on structure, generally as the molecular weight
increases the propensity for the polymer to shear also increases. There are differences for different
polymer types, for example, straight-chain structures versus “star” polymers. The two main processes
that cause polymer degradation are shearing (mechanical) and oxidative cleavage of the polymer chain
(chemical). A formulator must balance the tendency of the polymer to shear with the expected viscosity
thickening due to oxidative processes and the viscosity thinning due to dilution with fuel.
For industrial fluids, viscosity modifiers are used to increase the viscosity of the final fluid by the use
of small amounts of these polymers.

25.6.2  Pour Point Depressant


Pour point depressants (PPDs) are used in automotive and industrial applications to lower the tem-
perature at which lubricants can flow. PPDs function by interfering with the wax molecules remaining
in the lubricant base fluid. For this reason, they are sometimes referred to as wax-crystal modifiers.
Petroleum-derived mineral oils are processed using a variety of solvent extraction processes and more
recently by hydroprocessing and hydrocracking methods [61]. These processes remove most of the wax
molecules from the lubricant base fluid, but not all. These wax molecules cause resistance to flow at lower
temperatures and the fluid will not pour, hence the term pour point. Several polymer types, including
polymethacrylates, polymeric alkylated naphthalenes and phenols, and chlorine-containing polymers
have been developed to interfere with the crystallization of these waxes and hence allow the lubricant
to flow (pour) at a lower temperature. By lowering the pour point of the fluid, the lubricant can be used
over a wider temperature range.

25.7  Corrosion Inhibitors


Rust and corrosion can affect machinery in most applications. Lubricants are formulated with additives
designed to mitigate the effects of water and acidic components that cause equipment damage. When
Additives for Lubricants 25-15

water and oxygen attack the surface of iron and iron-containing alloys, the result is rust. This is a surface
phenomenon and proceeds to build iron content in an oil as determined by analysis. Rust can also fuse
affected parts resulting in immobility thereby damaging the equipment. Corrosion affects nonferrous
metals. This phenomenon occurs through the action of acidic components generated by hydrocarbon
oxidation as described in Section 25.2.1. Other chemical species that can react with the metal surface
also cause corrosion.
Corrosion results in the loss of metal from equipment and, therefore, can affect the integrity and
operation of the equipment and result in unsafe operation. Lubricant degradation can be enhanced
due to the catalytic effect of metal in contact with the lubricant at operating temperatures. EP additives
in gear oils are known to be aggressive to metal especially at higher temperatures and therefore, gear
oils need to be formulated using corrosion inhibitors [62]. However, since corrosion inhibitors are also
surface-active, they compete with antiwear additives for the metal surface. The simplest examples of
corrosion inhibitors are the straight-chain fatty acids.

25.7.1  Succinic Acids


An example of a corrosion inhibitor that is commonly used is dodecenyl succinic acid [63].

C12H23 CH COOH

H 2C COOH

25.7.2  Amine Phosphates


Amine phosphates with varied hydrocarbon groups are also used [63].

(RO)2 P R’
OH NH
R’’

The amine phosphates are also used as EP/antiwear additives in gear oils. Both of these classes of cor-
rosion inhibitors are effective at low concentrations (0.05–1.0 by weight).
Commercial lubricants that typically contain rust and corrosion additives include the general class
of rust and oxidation (R&O) oils, specifically formulated to resist R&O [64]. Higher performance lubri-
cants contain R&O additives and also contain other additives to improve performance under more
severe conditions. These are used in most applications including automotive (engine oils, transmission
fluids, and automotive gear oils) and industrial oils (metalworking fluids, hydraulic fluids, industrial
gear oils, protective oils, and greases).

25.7.3  Aromatic Corrosion Inhibitors


Various other organic structures serve as corrosion inhibitors [63,65]. Triazoles, generally benzotriazole
or tolyltriazole, are commonly used in industrial lubricants. Alkylated versions of these heterocyclic
structures form the basis of many commercially available corrosion inhibitors. Corrosion inhibitors
based on benzotriazole can be used on copper, brass, bronze, and multimetal systems. These can be used
in hydraulic fluids, brake fluids, mineral oils, and metalworking fluids. Typical structures for commer-
cially available corrosion inhibitors are as follows [63,66]:
25-16 Lubricants

H
N

N
Benzotriazole

N N

H
S C C
S
R S S

2-Alkyl-4-mercapto-1,3,4-thiadiazole
Vapor phase corrosion inhibitors are specifically designed to transfer by vaporization the requisite cor-
rosion inhibitor components to spaces in equipment that are not easily reached by liquid transfer during
run-in. This is typically done by operating equipment containing a lubricant having a vapor-phase cor-
rosion inhibitor and then draining prior to shipment or long-term storage. These generally function by
forming a protective film that remains on the metal surface during storage or transport prior to addition
of lubricant.

25.7.4  Metal Deactivators


These additives operate by forming a chemically inactive film on the metal surface [65]. Since metal ions
can promote oxidation reactions on the surface, various additives are used to minimize this effect. The
main types of additives used for this application are phosphates, amines, and sulfur compounds in the
form of sulfides. These can include low concentrations of phosphoric acid, citric acid, and ethylenedi-
aminetetraacetic acid, which act by chelating iron, copper, and other metals. By chelating these metals,
their contribution to free radical chain oxidation of the lubricant is minimized. Thiadiazoles, pyrazoles,
and imidazoles have also been reported as effective metal deactivators [66].

25.8  Miscellaneous Additives


25.8.1  Tackifiers and Anti-Misting Additives
Tackiness additives are used to provide adhesion of a lubricant to metal surfaces and cohesion to itself
to minimize leakage, mist, or splatter of the lubricant. Many of these materials are viscosity modifiers,
thickeners, and aluminum salts of unsaturated fatty acids [67,68].
In many applications, mist can be a health issue and an environmental problem. Anti-mist additives
function by providing cohesion of the lubricant and keeping the lubricant from becoming airborne.

25.8.2  Antifoam Additives


Foaming occurs when gas is trapped within a lubricant [69,70]. This can occur due to high-speed mixing
or agitation of a lubricating fluid. If the gas cannot escape rapidly, then the foam acts to decrease the effec-
tiveness of the lubricant on the tribosurface and can result in rapid wear. Foam also results in carryover of
lubricant in, for example, compressor or vacuum pump operation where volatility is critical [71]. Foaming
also increases the surface contact between air (oxygen) and the lubricant and can increase the rate of oxi-
dation especially at higher temperatures. Liquids transfer heat more effectively than gases and foaming
reduces the effective cooling of equipment, one of the main functions of lubricating fluids. This can cause
problems of reduced equipment life and operational problems as a direct result of the foam itself.
Additives for Lubricants 25-17

Because these molecules have poor oil solubility, they remain associated with the oils surface. These
additives work by adsorbing on the foam bubble and thereby affect the bubble surface tension causing
coalescence and breaking of the bubble at the lubricant surface [72].
Antifoam additives are typically silicones of various molecular weight, functionalized polymers such
as alkyl methacrylates, and amine-containing additives. The most common types are polydimethylsi-
loxanes. Each supplier has specific diluent fluids in which the silicone fluid is dispersed. These fluids tend
to be mineral spirits, and light-to-medium viscosity paraffinic hydrocarbons. Other antifoam additives
types include polyethylene glycols, organic polyacrylate copolymers, and tert-butyl phosphates [72].

25.8.3  Demulsifiers
Demulsifiers are surface-active agents and are classified as surfactants [73]. These additives release mois-
ture from the lubricant when the fluid is subjected to interaction with water or in very humid environ-
ments. These additives are generally included in formulations intended for steel mills and paper mills.
Many food-processing operations also require demulsifiers. These additives must not only be operation-
ally effective but must also be safe for incidental contact with food.
Several types of surfactants have been used as demulsifiers. Block copolymers of ethylene oxide–­
propylene oxide (20%–50% ethylene oxide) are used. Anionic surfactants such as alkyl naphthalene
sulfonates and nonionic alkyl phenolic resins are also used.

25.8.4  Seal Swell Additives


Lubricant base fluids have a tendency to either swell or shrink seal materials [74,75]. Both can be a prob-
lem. Different seal materials are affected in different ways by lubricants and it is prudent to know how
the fluid performs prior to recommending it for service. Esters and aromatic structures are used to cause
moderate swelling of seals in formulations and hydrocarbons are used to shrink seals.
Seal swell agents are typically polar. Since saturated hydrocarbons tend to shrink seal materials,
esters, polyesters, and phosphorus-containing additives are commonly used. The addition of small
amounts (5%–15% by weight) of an ester base fluid or polyester can swell seals to effective levels and
result in improved additive solubility as well.

25.8.5  Dyes, Chemical Marking Agents, and Tracers


These materials are used for product identification, for tracing the use of product, and to determine
adulteration or mixing of products with competitive or lower-quality lubricants. Oil-soluble dyes are
commonly used to define a product line or to promote corporate identity. They are used to provide con-
sistent color and brand recognition. Chemical marking agents and tracers can have a specific elemental
signature that can be used to identify a lubricant using analytical techniques.

25.8.6  Food-Grade Additives


Additives suitable for use in lubricants that may come in contact with food are continuing to be regu-
lated to stricter standards [76]. There are several organizations that oversee the safety compliance of
additives used in these lubricants. The most common organization in the United States is the National
Sanitation Foundation (NSF) located in Ann Arbor, Michigan. Lubricant formulations to be used in
food processing plants near food processing operations are expected to be NSF certified. Lubricant
safety is assessed by evaluation of the individual components within the formulation. From known
information, submitted by the additive manufacturer about the toxicity of the particular additive, NSF
makes a decision as to its suitability for use in food processing plants. If approved, it is given the designa-
tion HX-1 indicating that the material is to be used in accordance with good manufacturing procedures
25-18 Lubricants

and that it is to be used in minimum quantities needed to achieve the desired technical effect. Generally,
individual additives have an upper limit of 0.5% in the lubricant formulation. Some are only as high as
0.1%. NSF maintains a searchable list of “food-grade” additives on their website, www.nsf.org under
nonfood compounds.

25.8.7  Antibacterials/Biocides/Fungicides
These materials are generally added to metalworking fluids due to the fact that these fluids are continu-
ously open to the air and are prone to promotion of bacterial and fungal growth [77]. Examples of mate-
rials used include triazines, morpholines, imidazolines, triazines, and thiazolines.

25.9  Future Trends


Further regulation related to safety, toxicity, and biodegradability is anticipated to become global in
their scope expanding well beyond the European Union, the United States, and Canada. Harmonization
of the regulations and requirements is a lofty goal for these issues. The costs to understand and imple-
ment requirements will become prohibitive for smaller companies and start-ups where innovation
typically occurs.
Research aimed at green synthesis of existing and new additives will become more important due to
the prevailing trend to eliminate toxic solvents and other chemicals that contribute to volatile organic
compounds (VOCs). Toxicity and biodegradability of both starting materials and final products will
affect the commercialization of new additive chemistry and cost of new product introductions to the
marketplace. VOCs will become a more regulated property of additives and formulations. Synthesis
of additives will be directed toward methods that are solvent-free and/or those where solvents can be
essentially completely recovered so as to minimize environmental impact.
Hurdles associated with getting new additives into commercial production because of regulatory
pressures and the cost of addressing safety concerns are increasing as REACH becomes more widely
adopted and other countries implement their own sets of regulations aimed at improving product safety
and environmental protection.
It is expected that there will be further consolidation within the industry in order to take advantages
of economy of scale, and global markets. Asia is expected to become a major producer and consumer of
lubricant additives.

References
1. J. A. O’Brien, Lubricating oil additives, in Handbook of Lubrication, Vol. II, E. R. Booser, ed., CRC
Press, Boca Raton, FL, 1983.
2. L. R. Rudnick, Lubricant Additives: Chemistry and Applications, 2nd edn., Taylor & Francis, Boca
Raton, FL, 2009.
3. D. Klamen, Lubricants and Related Properties, Verlag Chemie GmBH, Weinheim, Germany, 1984.
4. J. Dong and C. A. Migdal, Antioxidants, in Lubricant Additives, Chemistry and Applications, 2nd edn.,
L. R. Rudnick, ed., Taylor & Francis, Boca Raton, FL, 2009, pp. 1–50.
5. R. A. McDonald, Zinc dithiophosphates, in Lubricant Additives, Chemistry and Applications, 2nd
edn., L. R. Rudnick, ed., Taylor & Francis, Boca Raton, FL, 2009, pp. 51–62.
6. D. Klamen, Lubricants and Related Products, Verlag Chemie, Weinheim, Germany, 1984, p. 179.
7. M. Rasberger, Oxidative degradation and stabilization of mineral oil based lubricants, in Chemistry
and Technology of Lubricants, 2nd edn., R. M. Mortimer and S. T. Orszulik, eds., Blackie Academic
and Professional, London, U.K., pp. 98–143.
8. P. C. Hamblin, U. Kristen, and D. Chasen, Lubrication Sciences, 2(4), 287–318 and references cited
therein.
Additives for Lubricants 25-19

9. C. C. Colyer and W. C. Gergel, Detergents and dispersants, in Chemistry and Technology of


Lubricants, R. M. Mortier and S. T. Orszulik, eds., Blakie Academic and Professional, an imprint of
Chapman & Hall, London, U.K., 1997, pp. 77–95.
10. S. P. Srivastava, Advances in Lubricant Additives and Tribology, Tech Books International, New Delhi,
India, 2009, pp. 85–87.
11. S. P. Srivastava, Advances in Lubricant Additives and Tribology, Tech Books International, New Delhi,
India, 2009, p. 89.
12. C. C. Colyer and W. C. Gergel, Ashless phosphorous containing lubricating oil additives, in R. M.
Mortier and S. T. Orszulik, eds., Blakie Academic and Professional, an imprint of Chapman & Hall,
London, U.K., 1997, pp. 77–80.
13. S. P. Srivastava, Advances in Lubricant Additives and Tribology, Tech Books International, New Delhi,
India, 2009, pp. 88–89.
14. C. C. Colyer and W. C. Gergel, Ashless phosphorous containing lubricating oil additives, in
R. M. Mortier and S. T. Orszulik, eds., Blakie Academic and Professional, an imprint of Chapman &
Hall, London, U.K., 1997, pp. 80–82.
15. C. C. Colyer and W. C. Gergel, Ashless phosphorous containing lubricating oil additives, in
R. M. Mortier and S. T. Orszulik, eds., Blakie Academic and Professional, an imprint of Chapman &
Hall, London, U.K., 1997, p. 83.
16. C. C. Colyer and W. C. Gergel, Ashless phosphorous containing lubricating oil additives, in
R. M. Mortier and S. T. Orszulik, eds., Blakie Academic and Professional, an imprint of Chapman &
Hall, London, U.K., 1997, pp. 82–83.
17. S. P. Srivastava, Advances in Lubricant Additives and Tribology, Tech Books International, New Delhi,
India, 2009, p. 88.
18. S. P. Srivastava, Advances in Lubricant Additives and Tribology, Tech Books International, New Delhi,
India, 2009, p. 94.
19. C. C. Colyer and W. C. Gergel, Ashless phosphorous containing lubricating oil additives, in
R. M. Mortier and S. T. Orszulik, eds., Blakie Academic and Professional, an imprint of Chapman &
Hall, London, U.K., 1997, pp. 77–90.
20. C. C. Colyer and W. C. Gergel, Ashless phosphorous containing lubricating oil additives, in
R. M. Mortier and S. T. Orszulik, eds., Blakie Academic and Professional, an imprint of Chapman &
Hall, London, U.K., 1997, pp. 86–88.
21. C. C. Colyer and W. C. Gergel, Ashless phosphorous containing lubricating oil additives, in
R. M. Mortier and S. T. Orszulik, eds., Blakie Academic and Professional, an imprint of Chapman &
Hall, London, U.K., 1997, p. 87.
22. S. P. Srivastava, Advances in Lubricant Additives and Tribology, Tech Books International, New Delhi,
India, 2009, p. 100.
23. P. Colacicco and D. Mazuyer, The role of soot aggregation on the lubrication of diesel engines,
Tribology Transactions, 38(4), 1995, 959–965.
24. F. G. Rounds, Soots from used diesel engine oils—Their effects on wear as measured in 4-ball wear
tests, SAE Technical Paper 810499, February 1981.
25. M. Gautam, K. Chitoor, S. Balla, and M. Keane, Contribution of soot contaminated oils to wear—
Part II, SAE Technical Paper 1999-01-1519, May 1999.
26. S. P. Srivastava, Advances in Lubricant Additives and Tribology, Tech Books International, New Delhi,
India, 2009, p. 111.
27. M. Johnson, Tribology and Lubrication Technology, March 2008, 28–36.
28. R. Scott, Basic wear modes in lubricated systems, Machinery Lubrication, July–August 2008, 34–40.
29. S. P. Srivastava, Advances in Lubricant Additives and Tribology, Tech Books International, New Delhi,
India, 2009, p. 113.
30. S. O. Jones and E. F. Reid, Journal of American Chemical Society, 60, 1938, 2452.
31. D. E. Johnson et al., U.S. Patent 5,135,670, 1992.
25-20 Lubricants

32. S. P. Srivastava, Advances in Lubricant Additives and Tribology, Tech Books International, New Delhi,
India, 2009, p. 129.
33. N. Canter, Trends in extreme pressure additives, Tribology and Lubrication Technology, September
2007, 10–18.
34. I. Minami, H.-S. Hong, and N. C. Mathur, Antiwear mechanism of dialkyl hydrogen phosphites,
STLE presentation, 96-NP-7, 54th Annual Meeting, Las Vegas, NV, May 23–27, 1999.
35. W. D. Phillips, Ashless phosphorous containing lubricating oil additives, in Lubricant Additives:
Chemistry and Applications, 2nd edn., L. R. Rudnick, ed., Taylor & Francis, Boca Raton, FL, 2006,
pp. 63–122.
36. E. S. Forbes and Battersby, The effect of chemical structure on the load-carrying and adsorption
properties of dialkyl phosphites, ASLE Transactions, 17, 1974, 263–270.
37. E. E. Klaus, J. Phillips, S. C. Lin, N. L. Wu, and J. L. Duda, Tribology Transactions, 33, 1990, 25.
38. M. Abdelmaksoud, J. W. Bender, and J. Krim, Nanotribology of a vapor-phase lubricant: A quartz
crystal microbalance study of tricresylphosphate (TCP) uptake on iron and chromium, Tribology
Letters, 13(3), October 2002, 179–186.
39. J. Wright, Extreme pressure additives in gear oils, Machinery Lubrication, September–October 2008.
40. T. Rossrucker and A. Fessenbecker, Sulfur carriers, in Lubricant Additives, Chemistry and Applications,
2nd edn., L. R. Rudnick, ed., Taylor & Francis, Boca Raton, FL, 2009, pp. 251–279.
41. R. S. Fein, Boundary lubrication relations, in Tribology Data Handbook, E. R. Booser, ed., CRC Press,
Boca Raton, FL, 1997, p. 479.
42. Y. R. Jeng and R. Desing, Automotive engine lubricants, in Tribology Data Handbook, E. R. Booser,
ed., CRC Press, Boca Raton, FL, 1997, p. 254.
43. S. Q. A. Rizvi, Additives, chemistry and testing, in Tribology Data Handbook, E. R. Booser, ed., CRC
Press, Boca Raton, FL, 1997, p. 121.
44. A. Nilpawar, Lube Magazine, No. 94, December 2009, pp. 6–7.
45. T. Kelley, R. Fensterheim, and A. Jaques, Chlorinated paraffins, Compoundings, 59(4), April 2009,
21–23.
46. G. Mariani, Selection and application of solid lubricants as friction modifiers, in Lubricant Additives:
Chemistry and Applications, 2nd edn., L. R. Rudnick, ed., Taylor & Francis, Boca Raton, FL, 2009,
pp. 175–179.
47. G. Mariani, Selection and application of solid lubricants as friction modifiers, in Lubricant Additives:
Chemistry and Applications, 2nd edn., L. R. Rudnick, ed., Taylor & Francis, Boca Raton, FL, 2009,
pp. 180–181.
48. N. Canter, Friction—Reducing characteristics of nano-boric acid, Tribology and Lubricant
Technology, 64(2), 10–11.
49. N. Canter, Boron nanotechnology-based lubricant additive, Tribology and Lubricant Technology,
65(1), 12–13.
50. Lube Magazine, No. 94, December 2009, p. 16.
51. G. Mariani, Selection and application of solid lubricants as friction modifiers, in Lubricant Additives:
Chemistry and Applications, 2nd edn., L. R. Rudnick, ed., Taylor & Francis, Boca Raton, FL, 2009,
pp. 179–180.
52. D. Kenbeck and T. F. Bunemann, Organic friction modifiers, in Lubricant Additives: Chemistry and
Applications, 2nd edn., L. R. Rudnick, ed., Taylor & Francis, Boca Raton, FL, 2009, p. 204, 207.
53. Tungsten disulfide is sold by M. K. Impex Corporation, Mississauga, ON, Canada, www.lowerfric-
tion.com
54. M. S. Covitch, Olefin copolymer viscosity modifiers, in Lubricant Additives: Chemistry and
Applications, 2nd edn., L. R. Rudnick, ed., Taylor & Francis, Boca Raton, FL, 2009, pp. 283–313.
55. S. P. Srivastava, Advances in Lubricant Additives and Tribology, Tech Books International, New Delhi,
India, 2009, pp. 166–167.
Additives for Lubricants 25-21

56. S. P. Srivastava, Advances in Lubricant Additives and Tribology, Tech Books International, New Delhi,
India, 2009, p. 163.
57. S. P. Srivastava, Advances in Lubricant Additives and Tribology, Tech Books International, New Delhi,
India, 2009, p. 168.
58. S. P. Srivastava, Advances in Lubricant Additives and Tribology, Tech Books International, New Delhi,
India, 2009, p. 181.
59. R. L. Stambaugh, in R. M. Mortier and S. T. Orszulik, eds., Blakie Academic and Professional, an
imprint of Chapman & Hall, London, U.K., 1997, pp. 144–180.
60. B. G. Kinker, Polymethacrylate viscosity modifiers and pour point depressants, in Lubricant
Additives: Chemistry and Applications, 2nd edn., L. R. Rudnick, ed., Taylor & Francis, Boca Raton, FL,
2009, pp. 315–338.
61. J. G. Speight and B. Ozum, Petroleum Refining Processes, Marcel Dekker, New York, 2002.
62. D. Klamen, Lubricants and Related Properties, Verlag Chemie GmBH, Weinheim, Germany, 1984,
p. 212.
63. S. P. Srivastava, Advances in Lubricant Additives and Tribology, Tech Books International, New Delhi,
India, 2009, p. 215.
64. L. Z. Pillion, Performance of rust inhibitors in hydrocracked base stock, Petroleum Science and
Technology, 21(9–10), 2003, 1461–1467.
65. M. T. Costello, Corrosion inhibitors and preventatives, in Lubricant Additives: Chemistry and
Applications, 2nd edn., L. R. Rudnick, ed., Taylor & Francis, Boca Raton, FL, 2009, pp. 421–444.
66. M. T. Costello, Corrosion inhibitors and preventatives, in Lubricant Additives: Chemistry and
Applications, 2nd edn., L. R. Rudnick, ed., Taylor & Francis, Boca Raton, FL, 2009, pp. 434–435, S215.
67. J. A. O’Brien, Lubricating oil additives, in Handbook of Lubrication, Vol. II, E. R. Booser, ed., CRC
Press, Boca Raton, FL, 1983, p. 306.
68. S. P. Srivastava, Advances in Lubricant Additives and Tribology, Tech Books International, New Delhi,
India, 2009, p. 186.
69. S. P. Srivastava, Advances in Lubricant Additives and Tribology, Tech Books International, New Delhi,
India, 2009, p. 227.
70. J. Crawford, A. Psaila, and S. T. Orszulik, in R. M. Mortier and S. T. Orszulik, eds., Blakie Academic
and Professional, an imprint of Chapman & Hall, London, U.K., 1997, pp. 192–193.
71. C. Kajdas, in R. M. Mortier and S. T. Orszulik, eds., Blakie Academic and Professional, an imprint of
Chapman & Hall, London, U.K., 1997, pp. 243.
72. S. P. Srivastava, Advances in Lubricant Additives and Tribology, Tech Books International, New Delhi,
India, 2009, pp. 231–233.
73. S. P. Srivastava, Advances in Lubricant Additives and Tribology, Tech Books International, New Delhi,
India, 2009, pp. 191–212.
74. S. P. Srivastava, Advances in Lubricant Additives and Tribology, Tech Books International, New Delhi,
India, 2009, pp. 247–258.
75. J. K. Sceron and R. E. Zelinski, Seal swell agents, in Lubricant Additives, Chemistry and Applications,
2nd edn., L. R. Rudnick, ed., Taylor & Francis, Boca Raton, FL, 2003, pp. 363–369.
76. S. Lawate, Formulation components for incidental food-contact lubricants, in Lubricant Additives
Chemistry and Applications, 2nd edn., L. R. Rudnick, ed., Taylor & Francis, Boca Raton, FL, 2003,
pp. 511–523.
77. W. R. Schwingel and A. C. Eachus, Antimicrobial additives for metalworking lubricants, in Lubricant
Additives, Chemistry and Applications, L. R. Rudnick, ed., Taylor & Francis, Boca Raton, FL, 2003,
pp. 383–397.
26
Rheology
Nomenclature..............................................................................................26-1
26.1 Introduction.....................................................................................26-2
26.2 Temperature and Pressure Dependence of Viscosity.................26-3
26.3 Shear Dependence of Viscosity, Generalized Newtonian
Liquids...............................................................................................26-7
26.4 Non-Newtonian Liquids: Normal Stress Differences,
Elongational Viscosity..................................................................26-10
26.5 Liquid Failure: Cavitation, Slip, Limiting Stress.......................26-12
Scott S. Bair 26.6 Opportunities and Challenges for the Future...........................26-13
Georgia Institute Acknowledgment.......................................................................................26-13
of Technology References...................................................................................................26-13

Nomenclature
av thermal expansivity defined for volume linear with temperature, K−1
B Doolittle parameter
c polymer concentration
CF fragility parameter in the Johari and Whalley equation
DF fragility parameter in the Vogel equation
G liquid critical shear stress or material modulus associated with λ, Pa
K0 isothermal bulk modulus at p = 0, Pa
K 0� pressure rate of change of isothermal bulk modulus at p = 0
K00 K0 at zero absolute temperature, Pa
M molecular weight, kg/kmol
n power-law exponent
N number of flow units
N1 first normal stress difference, Pa
N2 second normal stress difference, Pa
p pressure, Pa
p∞ divergence pressure, Pa
pP Roelands pole pressure, Pa
Rg universal gas constant, = 8314.34 Pa·m3/kmol/K
S Roelands slope index
T temperature, K
TR reference temperature, K
T∞ temperature at which the viscosity diverges, K
u velocity in the x-direction (generally the rolling direction), m/s
v velocity in the y-direction, m/s

26-1
26-2 Lubricants

V volume at T and p, m3
VR volume at reference state, TR, p = 0, m3
V0 volume at p = 0, m3
V∞ occupied volume, m3
V∞R occupied volume at reference state, TR, p = 0, m3
w velocity in the z-direction (generally the cross-film direction), m/s
x coordinate in the direction of surface velocity, m
y coordinate along the surface normal to the surface velocity, m
z coordinate across the film, m
Z Roelands pressure–viscosity index
α local pressure–viscosity coefficient, Pa−1
α 0 initial pressure–viscosity coefficient, Pa−1
α * reciprocal asymptotic isoviscous pressure coefficient, Pa−1
αB Barus pressure–viscosity coefficient, Pa−1
β temperature–viscosity coefficient, K−1
β K temperature coefficient of K0, K−1
γ shear rate, s−1
μ limiting low-shear viscosity, Pa·s
μ∞ constant in the Johari and Whalley equation, Pa·s
μe elongational Newtonian viscosity, Pa·s
μ2 second Newtonian shear viscosity, Pa·s
μD constant in the free-volume formulation, Pa·s
μP Roelands pole viscosity, Pa·s
μR viscosity at the reference state, Pa·s
μV constant in the Vogel equation, Pa·s
η rate-dependent shear viscosity, Pa·s
ψ free-volume scaling parameter
ρ mass density, kg/m3
ρ 0 mass density at p = 0, kg/m3
ρR mass density, at reference state, TR, p = 0, kg/m3
σij the ijth component of total stress, tension is positive, Pa
τ shear stress, Pa

26.1  Introduction
Rheology is the science and study of deformation and flow of materials and, therefore, includes the
study of linear elasticity of solids and Newtonian viscous flow of liquids. However, in practice, rheol-
ogy research tends to be interested exclusively in the constitutive equations which describe the complex
behavior of non-Newtonian liquids. The flow conditions for tribological liquids are so severe that the
temperature and pressure dependence of the viscosity is often of overwhelming importance relative to
the non-Newtonian response. Therefore, it is proper to include temperature and pressure effects in any
discussion of rheology in lubrication. The temperature dependence takes on even more importance
when one considers that lubrication has a long history of mistaking viscous heating for non-Newtonian
response [1].
While it has been appreciated for some time that the dependence of viscosity on temperature and
shear is important to film predictions in hydrodynamic lubrication, rheology has a special significance
in the field of elastohydrodynamic lubrication (EHL). EHL friction and film thickness calculations must
always account for the pressure dependence of viscosity. With the exception of very low pressure and
very low sliding velocity, friction calculations must account for the shear and temperature variation of
Rheology 26-3

viscosity as well. For modern lubricants, particularly for small scale or high pressure, film thickness
calculations must also consider the dependence of viscosity on shear.
When rheology is included in a lubrication calculation, a mathematical model is required. There are
two important considerations regarding rheological modeling. First, a function must be chosen which
accurately reproduces the viscosity trends over the range of temperature, pressure, and shear stress
expected to exist within the contact. Second, the model must be parameterized accurately. For both
requirements, it should be clear that a viscometer is essential.
Consider a flow for which the velocity in the x-direction is u = γy and the velocities in the y- and
z-directions are v = w ≡ 0. In the case of a Newtonian liquid, the shear stress in the xy-plane is σxy = ηγ
and the other shear stresses are σyz = σzx ≡ 0. The viscosity, η, is always equal to the limiting low-shear
viscosity, μ, and does not vary with time (history) or γ except for the variations resulting from the effect
of temperature and pressure upon μ. The normal stresses are σxx = σyy = σzz = −p.
In the case of an elongational flow for which the elongation rate in the x-direction is ε = ∂u/∂x
and elongation rates in the y- and z-directions are equal, if the flow is incompressible, the elon-
gation rates in the other two directions must be −ε/2. The Newtonian elongational viscosity is
μ e = (σxx − σyy)/ε = 3 μ.
These Newtonian descriptions are valid for lubrication only over a restricted range of temperature,
pressure, and shear. The most well-known departure from Newtonian response occurs when the vis-
cosity is dependent upon the shear rate, γ (or shear stress, σxy). This shear thinning is, however, not
the most profound or interesting non-Newtonian effect and is sometimes as subtle as to be difficult to
measure with accuracy. For shear thinning, not only non-Newtonian liquids is η = η(γ), a decreasing
function of shear rate when a critical rate or stress is exceeded, but also the normal stresses are not
equal, σxx = σyy + N1 and σyy = σzz + N2. In the low-shear regime for which viscosity is constant, N1 ∝ g 2
and N2 ≈ −N1/10.
Another form of non-Newtonian response that can be observed in liquid lubricants is thixotropy.
This behavior is difficult to model because the viscosity is both time and shear dependent. When a
paraffinic mineral oil is cooled or compressed, a weak solid structure will sometimes form. This waxy
structure may be disrupted by flow resulting in shear dependence. The wax structure requires time to
solidify, hence the time dependence. Obviously, the viscosity will depend upon the temperature, pres-
sure, and shear history.

26.2  Temperature and Pressure Dependence of Viscosity


The 1953 report [2], commissioned by ASME Research Committee on Lubrication and Advisory
Board on Pressure Viscosity, has been a valuable source of data on the temperature and pressure
dependence of viscosity and density. Not only is the equation of state important to modeling of
lubrication with compressible fluids, but also a description of the temperature and pressure varia-
tion of volume (or density) is useful for calculations of viscosity under severe conditions. Here, an
example is made from one of the lubricants studied by the ASME committee, a silicone oil. The
pressure–viscosity isotherms for ASME sample 55H are displayed in Figure 26.1. The slopes of the
isotherms may be seen to begin to increase with pressure after the viscosity exceeds about 1 Pa·s and
the vertical spacing between the isotherms increases more rapidly with pressure at the higher pres-
sures. These effects represent fragility and the enhanced pressure and temperature dependence of
the viscosity as the glass transition is approached through cooling or compression. Lubricating oils
are fragile, glass formers.
A correlation of viscosity with temperature and pressure that is widely used in tribology is Roelands
third model [3], which assumes that all pressure–viscosity isotherms intersect at some negative pressure
(tension) p = pP where the viscosity is μ = μP. This model is appropriate for low temperatures and the low
pressures typical of hydrodynamic lubrication and the inlet zone of EHL:
26-4 Lubricants

1,000,000

Viscosity/mPa.s 10,000

ASME silicone oil 55H


100
T = 273 K
T = 311 K
T = 372 K
1 Roelands equation
J and W equation

0.01
–200 0 200 400 600 800 1000 1200
Pressure/MPa

FIGURE 26.1  Pressure–viscosity isotherms for sample 55H of [1] shown with two models: Roelands [3] for low
(hydrodynamic) pressures and Johari and Whalley [4] for high (elastohydrodynamic) pressures.

⎡ ⎛ p − p ⎞ Z ⎛ T −T ⎞ S ⎤
⎢ P R ∞ ⎥
⎛ m ⎞ ⎢⎜⎝ pP ⎟⎠ ⎜⎝ T −T∞ ⎟⎠ ⎥
m = mP ⎜ R ⎟ ⎣ ⎦
(26.1)
⎝ mP ⎠

where
Z is the pressure index
S is the slope index
μR = μ (TR, p = 0)
T∞ is the divergence temperature where the viscosity is unbounded

Roelands specified universal values of T∞ = 137 K, μP = 6.31 × 10−5 Pa·s, and pP = −0.196 GPa so that
Equation 26.1 with the universal parameters is a model with only three adjustable parameters. Equation
26.1 has been plotted in Figure 26.1 for Z = 0.49, S = 0.47, TR = 273 K, and μR = 0.164 Pa·s.
The most widely used pressure–viscosity model in high pressure physics is the Johari and Whalley
[4] equation:

⎛ C p ⎞
m = m∞ exp ⎜ F ∞ ⎟ (26.2)
⎝ p∞ − p ⎠

where the viscosity diverges at p = p∞. This model is plotted in Figure 26.1 for the fragility parameter,
CF = 27.8. The divergence pressure is p∞ = 1.87, 2.52, and 4.49 GPa and μ∞ = 11.5 × 10−14, 5.96 × 10−14, and
4.49 × 10−14 Pa·s for T = 273, 311, and 372 K, respectively. This model is appropriate for the Hertz region
of EHL contacts.
It is clear from Figure 26.1 that Equations 26.1 and 26.2 are useful at low and high pressures, respec-
tively, but neither is appropriate over the full range of pressures applicable to EHL from inlet to the
contact center. Further, Equation 26.1 is not capable of describing fragility. The strength of intermo-
lecular interactions is controlled by the intermolecular separation which depends upon both temper-
ature and pressure through the volume. Temperature activates the movement of molecules past one
another in liquids. It is therefore reasonable to correlate viscosity with temperature and volume rather
Rheology 26-5

1.1

1.05
55H
273 K
1
Relative volume

298 K
311 K
0.95 372 K
Tait

0.9

0.85

0.8
0 100 200 300 400 500 600 700 800
Pressure/MPa

FIGURE 26.2  The relative volume, V/VR, for ASME sample 55H and the Tait equation of state (Equation 26.3).

than temperature and pressure and introduce pressure through the equation of state. The relative vol-
ume, V/VR, is plotted for ASME sample 55H in Figure 26.2 where VR is the volume at reference state,
TR = 273 K, and p = 0. The modified Tait equation of state has been plotted as well:

V 1 ⎡ p ⎤
=1− ln 1 + (1 + K 0ʹ )⎥
V0 1 + K 0ʹ ⎢⎣ K 0 ⎦
K 0 = K 00 exp ( − bKT )

V0
= 1 + aV (T − TR ) (26.3)
VR

for K 0ʹ = 10.333 , K00 = 7.617 GPa, βK = 5.526 × 10−3 K−1, and aV = 9.28 × 10−4 K−1. Compressibility, the recip-
rocal of the bulk modulus, increases exponentially with temperature.
Free-volume theory sets the logarithm of viscosity proportional to the reciprocal of the free volume.
The basic relation is the Doolittle [5] equation:

⎛ V∞ ⎞
m ∝ exp ⎜ B (26.4)
⎝ V − V∞ ⎟⎠

for total volume, V, and occupied volume, V∞. The free volume is V − V∞. It is often assumed that the
occupied volume varies with temperature only as

V∞ = V∞R ⎡⎣1 + e(T − TR )⎤⎦ (26.5)



26-6 Lubricants

1,000,000 ASME silicone oil 55H

100,000 273K
311K
372K
Viscosity/mPa.s

10,000 Doolittle

1,000

100

10
0.8 0.9 1.0 1.1 1.2
(V/VR)/(1 + ε(T – TR))

FIGURE 26.3  The viscosity of ASME sample 55H plotted against the free-volume scaling parameter showing
collapse of all of the data of Figure 26.1 onto a master curve.

A free-volume scaling parameter may be defined as

⎛V⎞ 1
y =⎜ ⎟ (26.6)
⎝ VR ⎠ 1 + e(T − TR )

Now, if B is constant, ψ represents the part of the Doolittle equation which varies with temperature and
pressure. Then a plot of μ(ψ) should yield a master curve for all of the data in Figure 26.1. Figure 26.3
shows that the scaling parameter defined by Equation 26.6 results in a compact master curve of viscos-
ity for the occupied volume thermal expansivity, ε = −5.024 × 10−4 K−1. The Doolittle equation (26.4) may
then be written as

⎡ B(V∞R / VR ) ⎤
m = mD exp ⎢ ⎥ (26.7)
⎣ y − (V∞R / VR ) ⎦

and this relation is plotted in Figure 26.3 for μD = 1.855 mPa·s, B = 1.591, and V∞R/VR = 0.7425. The
free-volume correlation has been an essential part of the recent quantitative EHL simulations which
have accurately predicted film thickness and even friction behavior from the measured viscosity. For
constant pressure, the well-known Vogel equation for the temperature dependence

⎡DT ⎤
m = mV exp ⎢ F ∞ ⎥ (26.8)
⎣ T − T∞ ⎦

results from the Doolittle equation with the assumption that volume depends linearly on T.
The temperature and pressure dependence of the low-shear viscosity is quantified by the local
pressure–viscosity coefficient

a (T , p ) =
( )
∂ ln m
(26.9)
∂p

Rheology 26-7

and by the local temperature–viscosity coefficient

b (T , p ) = −
( )
∂ ln m
(26.10)
∂T

The film thickness formulas of EHL characterize the piezoviscous response along the inlet zone pressure
sweep with a pressure–viscosity coefficient. The definition of Equation 26.9 yields a coefficient which,
in general, depends upon pressure and is therefore not useful for this purpose. There has been no con-
sensus regarding the proper definition of a pressure–viscosity coefficient for EHL film-forming; there-
fore, two of the most popular are given here. The conventional pressure–viscosity coefficient is often
employed, defined by

a0 = ⎢
( ) ⎤⎥
⎡ d ln m
(26.11)
⎢⎣ dp ⎥⎦
p=0

The reciprocal asymptotic isoviscous pressure coefficient is given by

−1
⎡ ∞ m ( p = 0) dp ⎤
a* = ⎢

⎣0
∫ m ( p) ⎥


(26.12)

and is more useful for film thickness calculations when the integral converges within the pressure range
of the contact. Other definitions are in use, including a coefficient which is evaluated by fitting a classical
film thickness formula to film thickness measurements.
The pressure–viscosity coefficient, regardless of the definition, is extremely sensitive to chemical
structure. Many short branches will result in a greater value of the coefficient than a few long branches,
for an example. Because of this sensitivity to structure, tables of representative values are not useful since
the chemical structure of most lubricants is not well-defined. For very low-viscosity liquids, ordinary
lubricants at very high temperatures, and water–glycol solutions, the low-pressure EHL inlet behavior
follows the linear Barus [6,7] equation

m = m0 (1 + a B p ) (26.13)

requiring yet another definition of a pressure–viscosity coefficient, αB , for linear response.


Three pressure–viscosity coefficients are plotted for the example of the ASME oil sample examined
in Figure 26.4. These coefficients are the conventional pressure–viscosity coefficient, α 0, the reciprocal
asymptotic isoviscous pressure coefficient, α*, and the Barus pressure–viscosity coefficient, αB .

26.3 Shear Dependence of Viscosity,


Generalized Newtonian Liquids
In a shear flow, as molecules move past one another, the interaction stretches and aligns the molecules. If
the motion is sufficiently slow, thermal vibrations will reorient the molecules so that the next interaction
is unaffected by the last. This is the low-shear (Newtonian) regime for which viscosity is constant. When
the shear rate, γ, becomes comparable to the reciprocal of the rotational relaxation time of a molecule,
the viscosity will no longer be independent of shear rate. The molecular alignment, which increases with
26-8 Lubricants

20
Pressure–viscosity coefficient/(1/GPa) ASME sample 55H

15

10
α0
α*
αB

5
0 50 100 150 200 250
Temperature/°C

FIGURE 26.4  Three pressure–viscosity coefficients for ASME sample 55H: the conventional pressure–viscosity
coefficient, α 0; the reciprocal asymptotic isoviscous pressure coefficient, α*; and the Barus pressure–viscosity coef-
ficient, αB , plotted against the temperature.

the shear rate, eases the flow. Relaxation times are generally proportional to the low-shear viscosity, μ.
Therefore, shear dependency is to be expected when μγ reaches a critical value we will call G, that is to
say, when the shear stress becomes comparable to G. The Einstein–Debye equation [8] gives a theoretical
value for the relaxation time:

m mM
= (26.14)
G rRgT

For the case of polymer solutions, the relaxation time depends on the concentration, c, of the polymer:

m mM (26.15)
=
G c rRgT

The most widely used models for shear-thinning express viscosity as a function of shear rate. They are
the Cross [9]

m − m2
h = m2 + 1− n (26.16)
1 + mg /G

and Carreau [10]

m − m2
h = m2 + (1− n)/ 2 (26.17)
⎡1 + ( mg / G )2 ⎤
⎣ ⎦
Rheology 26-9

10

1
η/µ

24°C, 500 MPa


24°C, 552 MPa
25.5°C, 604 MPa
40°C, 618 MPa
22°C, 403 MPa
Carreau
0.1
0.0001 0.001 0.01 0.1 1
µγ/G

FIGURE 26.5  Shear-dependent viscosity in a mineral oil at the temperatures and pressures indicated in the
legend.

equations, with a second Newtonian viscosity, μ2, a shear-independent viscosity to describe the flow
curve at high shear rate. A second Newtonian will sometimes appear for polymer-thickened oil. For
μ2 = 0, the rate sensitivity is n = d ln τ/d ln γ in the limit of high shear rate. Another useful model for
lubrication expresses the viscosity as a function of the shear stress, τ. This is the Ellis equation [11]:

m − m2
h = m2 + (1 / n) −1)
(26.18)
1 + t/G (

The use of the Carreau model (26.17) is displayed in Figure 26.5 where, with G = 8.3 MPa, n = 0.29, and
μ2 = 0, the viscosity of a mineral oil is accurately described over a range of temperature, pressure, and
shear. The measurements were obtained with a pressurized rotating concentric cylinder viscometer.
Figure 26.5 also illustrates a powerful property of shear-thinning liquids: time–temperature–pressure
superposition. A single master curve can be generated from many flow curves by shifting the data along
the vertical axis according to the low-shear viscosity, μ, and along the horizontal axis by the relaxation
time, μ/G, where in this case G can be considered a constant. This implies that a single flow curve, mea-
sured at a convenient condition, may be used to describe the shear response over a range of temperature
and pressure.
The superposition principal is further illustrated in Figure 26.6 for a formulated gear oil. The flow
curve for ambient pressure was generated in a commercial high-temperature, high-shear (HTHS) vis-
cometer and the data are courtesy of Frederic Jarnias of Total Oil. The elevated pressure data were
obtained with the viscometer used in Figure 26.5. Again, the Carreau equation (26.17), with n = 0.875
and G = 3800 Pa, has been used to represent the master curve generated, this time, by plotting relative
viscosity versus shear stress, τ.
Equations 26.16 through 26.18 are known as generalized Newtonian models because they describe
behavior that is in all ways Newtonian, except for the shear-dependent viscosity. Often, in analyzing a
lubrication flow, the coordinate system cannot be aligned everywhere with the direction of shear. In this
case, neglecting elongation rates and with the cross-film direction being the z-direction, the appropriate
value of γ to be used in calculating viscosity is
26-10 Lubricants

Gear oil
10

150°C, 0.1 MPa, TBS viscometer

23°C, 400 MPa, pressure viscometer

37.5°C, 508 MPa, pressure viscometer

Carreau model
η/µ

0.1
1.E + 02 1.E + 03 1.E + 04 1.E + 05 1.E + 06
Shear stress/Pa

FIGURE 26.6  Shear-dependent viscosity in a gear oil at the temperatures and pressures indicated in the legend.
One flow curve is from a commercial HTHS viscometer and the other two from the pressurized Couette viscometer
used in Figure 26.5.

2 2
⎛ ∂u ⎞ ⎛ ∂v ⎞
g= ⎜ ⎟ +⎜ ⎟ (26.19)
⎝ ∂z ⎠ ⎝ ∂z ⎠

Similarly, neglecting normal stress differences, the appropriate value of τ to be used in calculating
viscosity is

t = s xz
2
+s 2yz (26.20)

26.4 Non-Newtonian Liquids: Normal Stress


Differences, Elongational Viscosity
The Reynolds equation is the foundation for nearly all of full-film lubrication. It is derived from Navier–
Stokes with simplifying assumptions regarding geometry and the possibility of pressure variations
across the film. Shear dependence of viscosity in lubricants arises from either thixotropy or molecular
stretching and alignment. The anisotropy from molecular alignment results in the directions of the
principal shear stresses not being the same as the directions of the principal shear rates. Then, in simple
shear, the three normal stresses cannot be equal and this is behavior which cannot be represented
by the Reynolds equation which assumes that load support comes from a local increase in isotropic
pressure.
This interesting phenomenon can have an effect on the load support in a thrust bearing as shown
in Figure 26.7. The thrust bearing consists of parallel disks of 19 mm diameter, one rotating and one
stationary, supported with a nominally fixed clearance. They are enclosed in a pressure vessel so that
the ambient pressure can be controlled. For Figure 26.7, two different lubricants are investigated: a
Rheology 26-11

8 Torque versus time

7 10W-40, 250 MPa


BCH, 150 MPa
6

5
Torque/N.cm

–1
0 0.2 0.4 0.6 0.8 1 1.2 1.4
(a) Time/s

40

35 Thrust versus time


10W-40, 250 MPa
30
BCH, 150 MPa
25

20
Thrust/N

15

10

–5
0 0.2 0.4 0.6 0.8 1 1.2 1.4
(b) Time/s

FIGURE 26.7  (a) The torque history of a thrust bearing, a rheogoniometer, at elevated pressure with a polymer
solution, 10W-40 motor oil, and a low-molecular-weight traction fluid base oil, BCH. (b) The thrust history for the
examples in (a).

low-molecular-weight base oil for traction fluids, BCH, and a commercial 10W-40 motor oil. Different
pressures, indicated in the figure, have been selected for each liquid so that the steady-state shear stress
generated by a single rotation at three revolutions per second will be the same, as shown in Figure 26.7a.
In other words, the steady shear viscosities are the same. The Reynolds equation would predict that no
load support or thrust may be generated in this geometry and in the case of the low-molecular-weight
liquid, BCH, this is true as shown in Figure 26.7b. However, there is a substantial thrust generated by
26-12 Lubricants

the motor oil in the same figure as a result of the elasticity contributed by the polymer VI improver.
The thrust is a consequence of the first normal stress difference [12]. In the low-shear regime for which
viscosity is constant, the first normal stress difference can be estimated from

2t2
N1 ⊕ (26.21)
G

and the effective shear modulus, G, may be estimated from Equation 26.14 for a base oil or Equation
26.15 for a polymer solution.
Another property of non-Newtonian lubricants which may benefit load support is the enhancement
of elongational viscosity. While the shear viscosity decreases with increasing shear rate, the elongational
viscosity increases with increasing elongation rate. For squeeze films, the deformation at the midplane is
pure elongation. For non-Newtonian lubricants, the enhancement of elongational viscosity in a squeeze
film can overcome the decrease of shear viscosity to result in a net increase in load over the Newtonian
case as found by Tichy and Winer [13].

26.5  Liquid Failure: Cavitation, Slip, Limiting Stress


One substantial requirement for an accurate rheological measurement is that the kinematics of flow
employed by the measurement technique must be reasonably predictable. For example, in a Couette
viscometer the flow is assumed to be Couette flow, the velocity varying in a linear fashion with position
moving from one boundary to the other. Any departure from linearity will render the measurement
useless because the rate of shear will not be known. Likewise, a calculation of the viscous friction in
a lubricant film cannot be successful if the velocity distribution cannot be accurately predicted. This
circumstance occurs, for example, when the flow cavitates. Hutton [14] described, as liquid failure, the
changes in the liquid other than structural which occur during flow and for which the rheological mod-
els do not apply, for example, fracture from tension in a flowing liquid.
The most obvious form of liquid failure for lubrication is the cavitation which occurs in the diverging
portion of the film where the pressure decreases below some critical value resulting in the appearance of
bubbles or gaseous streamers. Cavitation may be caused by a shear stress as well. In a Newtonian liquid,
one of the principal stresses will become tensile when the shear stress exceeds the pressure in the film.
This shear cavitation bedevils the ambient pressure measurement of high-shear viscosity and probably
influences the onset of cavitation in bearings. However, quiescent liquids may withstand a negative pres-
sure, a hydrostatic tension, of −10 MPa [15] and the pressure dependence of the viscosity is continuous
across p = 0 into the negative pressure regime.
Although there has recently been much speculation concerning slip at the liquid-to-solid boundary
from shearing of lubricant films, direct empirical evidence has not been forthcoming. However, slip
within the liquid film itself has been visually documented by at least two laboratories [16]. The slip
planes have two directions which may be explained by the Mohr–Coulomb failure criterion which
asserts that slip operates on a plane in the material where the shear stress reaches a value which varies
linearly with the normal stress on the plane. Shear bands may be most readily visualized in a squeeze
film. In Figure 26.8, a mineral oil has been compressed to 1 GPa in an optical cell. Then, in the micro-
graph of Figure 26.8, a 1.3 mm diameter pin at the top is being pushed toward an anvil at the bottom
of the picture with a force of 1.5 kN. The film was illuminated at a glancing angle so that only refracted
light is transmitted through the film. Shear bands are visible in Figure 26.8 operating in two direc-
tions within the film. Slip within the elastohydrodynamic oil film is the most reasonable explanation
of the limiting stress effect in full-film friction whereby liquid films display a characteristic friction
coefficient. The friction coefficient is directly related to the ratio of shear stress to normal stress in the
Mohr–Coulomb failure criterion.
Rheology 26-13

FIGURE 26.8  Shear bands operating in a squeeze film of mineral oil at high pressure (1 GPa).

26.6  Opportunities and Challenges for the Future


The concepts discussed in the previous sections, with the exception of shear banding, have been widely
accepted outside of EHL for >30 years. The application of realistic rheological models that have been
parameterized by a direct measurement of viscosity is, however, only a recent development [17] in EHL.
In 2007, film thickness and friction were accurately calculated from measurable properties [18] and
new details of the EHL film-forming response have been discovered prior to experimental observations
[19] by simulation with realistic rheology. We may now begin to discover the properties which control
film thickness and friction and, through simulation, begin to engineer the best lubricants for a particu-
lar application. For this purpose, nonequilibrium molecular dynamics [20] can provide a technique to
obtain the rheological properties of liquids which have not yet been synthesized for conditions that can-
not be attained in the present day viscometers.

Acknowledgment
This work was supported by the National Science Foundation under grant number EEC#0540834.

References
1. Hersey, M.D. and Zimmer, J.C., Heat effects in capillary flow at high rates of shear, J. Appl. Phys., 8,
1937, 359–363.
2. Kleinschmidt, R.V., Bradbury, D., and Mark, M., Viscosity and Density of Over Forty Lubricating
Fluids of Known Composition at Pressures to 150,000 psi and Temperatures to 425°F, ASME, New
York, 1953.
3. Roelands, C.J.A., Correlational aspects of the viscosity–temperature–pressure relationship of lubri-
cating oils, PhD thesis, University of Technology, Delft, the Netherlands, 1966, pp. 107–108.
4. Johari, G.P. and Whalley, E., Dielectric properties of glycerol in the range 0.1–105 Hz, 218–357 K,
0–53 kb, Faraday Symp. Chem. Soc., (6), 1972, 23–41.
5. Doolittle, A.K., Studies in Newtonian flow. II. The dependence of the viscosity of liquids on free-
space, J. Appl. Phys., 22(12), 1951, 1471–1475.
6. Barus, C., Note on the dependence of viscosity on pressure and temperature, Proc. Am. Acad. Arts
Sci., 27, 1891–1892, 13–18.
7. Barus, C., Isothermals, isopiestics and isometrics relative to viscosity, Am. J. Sci., Third Series,
XLV(266), 1893, 87–96.
26-14 Lubricants

8. Paluch, M., Sekula, M., Pawlus, S., Rzoska, S.J., Ziolo, J., and Roland, C.M., Test of the Einstein–Debye
relation in supercooled dibutylphthalate at pressures up to 1.4 GPa, Phys. Rev. Lett., 90(20), 2003,
175702-1-4.
9. Cross, M.M., Rheology of non-Newtonian fluids: A new flow equation for pseudoplastic systems,
J. Colloid Sci., 20, 1965, 417–437.
10. Carreau, P.J., Rheological equations from molecular network theories, Trans. Soc. Rheol., 16(1), 1972,
99–127.
11. Meter, D.M. and Bird, R.B., Tube flow of non-Newtonian polymer solutions: Part 1. Laminar flow
and rheological models, AIChE J., 10(6), 1964, 878–881.
12. Macosko, C.W., Rheology: Principles, Measurements and Applications, VCH, New York, 1994,
pp. 217–222.
13. Tichy, J.A. and Winer, W.O., An investigation into the influence of fluid viscoelasticity in a squeeze
film bearing, ASME J. Lubr. Technol., 100, 1978, 56–69.
14. Hutton, J.F., Theory of rheology, Interdisciplinary Approach to Liquid Lubricant Technology (P.M. Ku,
Ed.), NASA, Washington, DC, 1973, pp. 187–261.
15. Fischer, J.C., The fracture of liquids, J. Appl. Phys., 19, 1948, 1062–1067.
16. Bair, S. and McCabe, C., A study of mechanical shear bands in liquids at high pressure, Tribol. Int.,
37, 2004, 783–789.
17. Liu, Y., Wang, Q.J., Wang, W., Hu, Y., Zhu, D., Krupka, I., and Hartl, M., EHL simulation using the free-
volume viscosity model, Tribol. Lett., 23(1), 2006, 27–37.
18. Liu, Y., Wang, Q.J., Bair, S., and Vergne, P., A quantitative solution for the full shear-thinning EHL
point contact problem including traction, Tribol. Lett., 28, 2007, 171–181.
19. Krupka, I., Bair, S., Kumar, P., Khonsari, M.M., and Hartl, M., An experimental validation of the
recently discovered scale effect in generalized Newtonian EHL, Tribol. Lett., 33, 2009, 127–135.
20. Bair, S., McCabe, C., and Cummings, P.T., Comparison of non-equilibrium molecular dynamics with
experimental measurements in the nonlinear shear-thinning regime, Phys. Rev. Lett., 88(5), 2002,
058302.
27
Lubricant Application
27.1 Introduction..................................................................................... 27-1
27.2 Gravity Feed Oilers.......................................................................... 27-1
Paul W. 27.3 Constant Level Oilers...................................................................... 27-2
Hetherington 27.4 Grease Dispensing........................................................................... 27-3
Fluid Life
27.5 Centralized Lubrication Systems.................................................. 27-5
Evan S. Zabawski 27.6 Oil Mist Systems.............................................................................. 27-8
Fluid Life Bibliography............................................................................................... 27-11

27.1  Introduction
The successful operation of industrial equipment is an important requirement in meeting the opera-
tional goals in today’s reliability-driven production facilities. One of the key elements of that success is
the proper application of the right lubricant, at the right amount and at the right time. There are several
ways in which a lubricant can be added to a component. Considering all of the different equipment types
and requirements, it is important that an appropriate method for applying the lube is selected, but that
is also in a cost-effective way.

27.2  Gravity Feed Oilers


Gravity feed oilers have traditionally been a popular method of providing automatic oil lubrication to
industrial equipment such as bearings, gearboxes, motors, chains, and other interacting machinery
components. As the name implies, the flow of oil to the component is obtained through the use of
gravity. Gravity feed oilers offer some of the same benefits of more advanced automatic lubrication
systems without the high cost. These systems provide a simplistic, yet very reliable method of keeping
critical component parts properly lubricated. These typically single-point lubricators are designed to
maintain a constant flow to the required components. The required lubricant feed rate can be manu-
ally set by the plant operator by simply adjusting the needle valve and viewing the drip rate in the flow
sight indicator.
The design typically consists of a supply reservoir, which is used to feed the lubricant through tub-
ing or small bore piping, by means of gravity to the point of lubrication. Where required, these systems
can also be expanded to include a number of manifolds or distribution points to provide oil to several
lubrication points on the equipment. In this design, each point will have its own independent needle
valve or drip control mechanism. In either case, one advantage of this design is that the reservoir may be
located a fair distance away from the actual lubrication point allowing for convenient monitoring and
refilling of the reservoir.
A simple drip-feed oiler is the most popular style and simplest method of the gravity feed oiler design.
In this case, the oil is applied directly to the component as a steady drip or very slow stream without the
need for an additional applicator attachment.

27-1
27-2 Lubricants

Wick feed oilers combine the normal gravity feed benefits with a simple capillary action to provide
a constant and consistent flow of oil to the lubricated component. The capillary action is provided and
controlled by the use of various wick sizes. The rate in which the oil is applied is directly related to the
size of the wick and the viscosity of the lubricating oil.
Chain oilers are a specific form of gravity feed oilers that automatically apply a film of lubricant to
chains or similar components such as slides, and oscillating components. These systems are vital in
reducing friction, linkage wear, rust, and corrosion. The chain oilers are similar in operation to the drip-
feed oilers; however, they typically contain an applicator to supply lubricant to the source. Applicators
are available in a variety of styles providing both lubrication and cleaning of the chains, the most com-
mon being a brush applicator.
These reservoirs and applicators come in various styles and sizes. Most reservoirs are made of a thick
wall glass, while a more durable plastic (typically polycarbonate or similar) reservoir is available for use in
food-processing plants where there is the possible danger of broken glass coming in contact with the pro-
cess environment. One manufacturer includes a unique cap design allowing oil to be added to the reservoir
with a standard oil can and spout. Once the oil can spout is pulled back out, the hole closes preventing the
ingress of ambient contamination, thus providing a simple and effective way for topping up the oil reservoir.

27.3  Constant Level Oilers


The commonly used constant level oilers are designed to maintain a predetermined oil level in the com-
ponent. If the oil level in the component (generally a bearing housing, gear box, or other small oil sump
application) were to drop below the predetermined level, oil would automatically be provided from the
attached new oil reservoir, returning the level to the original height. The use of these relatively simple,
but effective, constant level oilers is an easy way to minimize maintenance and operational costs while
ensuring safe and reliable operation of the component and overall production facility.
The majority of constant level oilers available from many suppliers are easily adjustable. However,
there are some constant level oilers that provide no adjustment in the oil level. The proper installation
of these fixed units is critical as no adjustment is available after installation. The majority of constant
level oilers are typically mounted on the side of the component in front of the direction of shaft rotation
(Figure 27.1). In applications where there is significant turbulence created from high rotating speeds or
the use of slinger rings, improper feeding of the oil may be possible. In these cases, it may be appropriate
to connect the constant level oiler to the bottom of the oil sump (Figure 27.2). The advantage to this style
is that the level of the oil may be adjusted in the field to match the specific desired height.
Many of the constant level oilers are vented to the outside through air slots. This venting is critical to
the operation of this style of unit. Unfortunately, since these are vented to the outside atmosphere, it is
possible for contaminants in the environment to be drawn into the oil as the oil in the reservoir (bowl)

Direction of shaft rotation

FIGURE 27.1  Standard gravity feed oiler. (Courtesy Trico Corporation, Pewaukee, WI.)
Lubricant Application 27-3

Direction of shaft rotation

FIGURE 27.2  Bottom mounted gravity feed oiler. (Courtesy Trico Corporation, Pewaukee, WI.)

Direction of shaft rotation

FIGURE 27.3  Closed loop gravity feed oiler. (Courtesy Trico Corporation, Pewaukee, WI.)

decreases. Common atmosphere contaminants such as airborne dirt and moisture can repeatedly affect
the life of the oil and component.
In severe conditions or with critical components, it may be appropriate to use a closed system con-
stant level oiler. This style of constant level oiler prevents the ingress of outside contaminants from
entering into the new oil and into the system. Many of these closed loop systems contain a pressure bal-
ance line which is connected from an air chamber built on the body of the constant level oiler to the top
of the bearing cap or gear case (Figure 27.3).
A fourth style of constant level oiler is mounted directly on the centerline of the required oil level
within the system. These typically are also a closed loop system allowing for an exchange of air between
the oil sump and the oiler. Several models of this style will also incorporate a sight glass window in the
end of the oiler allowing for verification of the actual oil level in the system, as well as an indication of
the oil condition.
There are several different styles of constant level oilers available that will meet specific design needs
and applications. The somewhat simple design of these constant level oilers provides an easy but effec-
tive method of maintaining an adequate and proper oil level in many components.

27.4  Grease Dispensing


A significant number of bearing failures occur due to improper greasing and more specifically due
to lack of grease based on timing or over-greasing based on improper knowledge and application
27-4 Lubricants

techniques. In either case, significant component failures occur from the improper greasing of com-
ponents. There are several methods of applying grease to the desire component or point of lubrication.
The most common forms are manual, typically a grease gun, automatic single point, or a centralized
lubrication system.
Grease guns have traditionally been the most common method of applying grease to a bearing. Some
of the common pitfalls in using a standard grease gun are the lack of understanding by the technician as
to the volume of grease required and the actual amount being dispensed by the manual grease gun. In
practice, it is very important that the actual required grease volume and frequency of re-lubrication is
identified for each and every bearing. Furthermore, it is imperative that the technician understands the
actual volume dispensed with each and every “shot” or pump of the grease gun. One recent improve-
ment to the standard grease gun is the portable battery-powered version. This makes it much easier
on the technician when adding significant volumes of grease without the physical pumping action.
The second significant improvement to the manual system is the addition of a compact digital volume
metering device. This simple addition provides a rather inexpensive but effective way for the technician
to clearly know how much grease is being added to the bearing. With proper care and control, the use of
manual grease guns is still a very effective and simple method for applying the right amount of grease
lubricant.
Automatic single-point grease lubricators consist of a fixed storage device designed to slowly dis-
pense grease to the bearing or required point of lubrication, over a given period of time. These devices
are usually mounted directly on the bearing housing at the point where a common grease (Zerk) fit-
ting would be located. One of the major advantages to this style of lubricators is that a small amount
of new grease is constantly being applied to the bearing without the need for a routine visit by the
lube technician. These systems are an excellent option for remote lubrication points or where lubrica-
tion would normally require a shutdown of the system. In addition, these systems can be a significant
advantage over manual systems and where the costs of a more elaborate centralized system cannot be
justified. These lubricators are available in both single use styles as well as ones that can be refilled in
situ through a Zerk fitting attached to the housing reservoir. Several styles are available to control and
dispense the right amount of grease. Typically, these lubricators may be driven by one of the following
methods:
• Mechanical (spring loaded)
• Electrochemical
• Electromechanical
Mechanical devices contain a preselected spring that drives a plunger to feed grease to the bearing. The
size and type of spring is matched to the grade of grease and the operating temperature. Although this
is a very effective way for automatically applying the grease, field adjustments to the dispensing rate are
usually not available without replacement of the spring.
Electrochemical devices typically incorporate a battery-operated timer which activates an electro-
chemical cell producing expandable gas that in turn generates pressure against a piston. The length
of time it takes to empty the unit is set by the lube technician at the time of installation allowing for
precise and continual dispensing of the grease. Depending on the size and style of the unit, dispensing
rates typically range from 1 to 12 months. This setting may also be easily adjusted in the field should a
modification in the dispensing rate be required.
Electromechanical devices utilize either a self-contained battery or direct-wired timer and motor
that is attached to a small gearbox that drives a cam that activates a piston pump which injects a small
amount of grease. Similar to the electrochemical devices, these units can be adjusted and readjusted in
the field as deemed necessary. One advantage of the direct-wired version is that power may be taken
directly from the system electrical circuit preventing unnecessary dispensing when the component is
not operating.
Lubricant Application 27-5

Centralized grease systems are similar in design and share similar components to the oil systems and
are therefore discussed along with the oils in the following section.

27.5  Centralized Lubrication Systems


Maintaining properly lubricated machinery is an important part of any lubrication program. Without
administering lubricant at correct intervals with the proper amount, a machine can experience costly
failures. Centralized lubrication systems are a common tool used in industry to distribute a precise
amount of lubricant to specific locations at specific times through the use of programmable timers,
lubricant pumps, and lubricant injectors.
The primary purpose of a centralized lubrication system is to eliminate the tedious and sometimes
difficult task of manually lubricating all of the desired components, especially in large systems and oper-
ating plants. Centralized lubrication can deliver either oil or grease to a specific lubrication point. This
section is different than circulating oil systems which are discussed in chapter 52. Centralized lubrica-
tion systems (Figure 27.4) offer two main benefits in increased machine reliability and a reduced labor
cost compared to traditional manual lubrication. In addition, centralized systems also offer advantages
including reduced downtime, efficient use of lubricants, and less waste. A common pitfall of a cen-
tralized system is the misunderstanding of the proper operation of the system by the operations and
maintenance personnel, as well as the tendency to ignore or properly monitor and maintain the system.
However, when a centralized system is properly designed and maintained, significant improvements in
the overall machine reliability can be realized.
Centralized lubrication systems were first introduced in the 1930s. Since then, significant improves
have been developed and incorporated into the centralized systems of today, providing accurate delivery
and volumes for a wide range of industrial applications.
Centralized lubrication systems are first divided into one of two categories either direct or indirect.
The simplified direct system incorporates only a pump which is used to pump and pressurize the lubri-
cant. Metering of the lubricant volume is primarily determined by the output of the pump itself. The
more common indirect system, which is more complex, incorporates metering valves (injectors) within
the system to meter the required lubricant volume to the required lubrication point. These indirect
systems can be further divided into either a parallel or nonparallel system. In parallel systems, more
commonly referred to as nonprogressive, the system is pressurized by the pump and all injector valves
operate at the same time to provide lubricant to the lubrication points. The major disadvantage of this
parallel or nonprogressive system is that a failure of an individual injector may not be easily identified.
With a single injector failure, system pressure will still be maintained and lubricant will continue to
be supplied to all other lubrication points. In the nonparallel or progressive systems, the injectors are
installed in series. Once the pump is started, and the system pressurized, the first injector opens allow-
ing the lubricant to pass to the next injector in that line. With this system, if one injector fails it causes

Main line
Distributor

Line to the lub points

Lub points
Return line
Pressure relief valve
Pump
Container

FIGURE 27.4  Oil circulating system. (Courtesy Beka Lube, Gauteng, South Africa.)
27-6 Lubricants

Pressure
relief Distributor
valve

Nonreturn
valve

FIGURE 27.5  Single-line central lubrication systems. (Courtesy Beka Lube, Gauteng, South Africa.)

all further injectors and lubrication points downstream to also fail. The advantage to this system is that,
depending on the location of the failed injector, system pressure will increase clearly identifying that
there is a problem somewhere in the system.
Centralized lubrication systems are also classified based on the following types or styles:
• Single line
• Dual line
• Multiline
• Progressive
• Oil mist
Single-line lubrication systems are designed for oil, semifluid grease, and thicker greases up to typically
an NLGI 2. Single-line systems (Figure 27.5) are ideal for lubricating the numerous lubrication points
found on small- and medium-sized machines, and on systems operating on an intermittent basis. They
are also found on large commercial mobile fleet systems. The basic system operates with a pump which
supplies the desired lubricant to the individual components via the main supply line. The lubricant is
metered and fed to the lubrication points through the injector. The injection of the lubricant takes place
either during or after the pump is in operation, depending on the type of distributor used. In a pre-
lubrication distributor system, the pressure produced by the running pump causes the pre-lubrication
distributors to dispense a pre-stored quantity onto the lubrication points. Once the pump is turned off,
the main line is relieved of pressure and the distributors’ storage chambers refill themselves for the next
lubrication cycle. In a re-lubrication distributor system, the lubricant is pumped into the feeders’ stor-
age chambers, where it is then stored. Only after the pressure is relieved in the main line is this quantity
dispensed under spring tension to the lubrication points (re-lubrication effect).
Dual-line lubrication systems are designed for oil, semifluid grease, and possibly thicker greases up
to an NLGI 3. Dual-line systems (Figure 27.6) are ideal for lubricating the numerous lubrication points
found on large machine systems, over large distances, such as steel mills, power plants, mining, cement
plants, etc. The advantage of a dual-line system is that it supplies lubricant via two supply lines from

Switch-over
valve

Dual-line distributor

FIGURE 27.6  Progressive and dual-line system. (Courtesy Beka Lube, Gauteng, South Africa.)
Lubricant Application 27-7

a single pump. A four-way valve is used to direct grease alternately to each of the grease lines while
relieving pressure on the other lines. The second line provides a safety margin but involves additional
cost and complexity related to installation. The dual-line system can be combined with a secondary
progressive injector device, which can further increase the number of lubrication points supplied in a
dual-line system. Conventional dual-line systems operate on a fixed pressure differential principle. In
the first cycle, the lubricant is pumped into the first main line while the second main line is connected
to the relief line. Pressure in the first main line causes the injectors to stroke in one direction dispens-
ing lubricant to the desired group of lube points. Once all injectors have dispensed their lubricant from
the first line, the second cycle begins. In the second cycle, a four-way (reverser) valve directs pump flow
to the second main line and opens the first main line to the relief line. This allows pressure to build in
the second main line causing the injectors to stroke back to their original position dispensing lubricant
again to lube points. The system is a parallel type and each dual-line injector operates independently of
any other in the system.
Multiline lubrication systems are designed for oil, semifluid grease, and harder greases generally up
to an NLGI 3.
These systems are ideal for lubricating multiple lubrication points on construction equipment, com-
mercial vehicles, or for small machinery where the lubrication points are not too far apart. In the mul-
tiline system (Figure 27.7), the lubricant is delivered by the pump in metered quantities through several
outlets. The lubrication lines lead directly from the pump to the lubrication points or to a progressive
feeder that further divides up the respective delivery rates of the connected outlet ports. The design of the
lubrication pump limits the maximum number of lubrication points which can be supplied by a direct
line system. An increase in the number of lubrication points is possible through the use of progressive dis-
tributors. In addition, the maximum length of line is also limited based on the pressure of the pump, size
of feed lines, operating temperature, and the characteristics of the lubricant or grease (i.e., pumpability).
If the feed lines are too long (>20–40 m), the resistance of the line will exceed the available pump pressure.
Progressive systems include a distributor block which divides up and distributes the desired quantity
of lubricant to the individual lubrication points in a progressive sequence. The design and operation of
the progressive distributor provide a simple and reliable control or the desired lubrication quantities in
a lubrication system.
A complete distributor generally comprises of at least three and up to nine dosage units along with
the inlet and connection units. Each dosage unit has a hydraulically controlled piston which measures
and discharges the grease fed to it. The discharged quantity is determined by the diameter and stroke
of the piston. As the name implies, the pistons of the individual dosage units operate in a progressive
sequence, one after the other. Each individual piston cannot make its stroke until the preceding piston
has first completed its stroke. The advantage to this system is that if any piston is prevented from com-
pleting its stroke, the distributor stops immediately and ceases to operate, immediately identifying a
system problem. Once all the pistons in a distributor block have made a reciprocating movement (back
and forth), the distributor has completed a full cycle.

Adjustable delivery rate


(optional)

FIGURE 27.7  Multiline system. (Courtesy Beka Lube, Gauteng, South Africa.)
27-8 Lubricants

6a 5a 1a

6 5
4

Inlet

3 2
1

(a) 3a 2a 4a

6a 5a 1a

6
5 4

Inlet

3
2 1

(b) 3a 2a 4a

FIGURE 27.8  (a) Progressive distributor Stage 1. (b) Progressive distributor Stage 2. (Courtesy Assalub AB,
Atvidaberg, Sweden.)

Figure 27.8a and b shows an example of the progression sequence of the distributor block. Initially,
the pump pressure forces piston end 4 down forcing piston end 1 to discharge grease through outlet
1a. Once piston 4/1 has completed its stroke, end 4 of the cylinder has been filled with a predetermined
quantity of grease and the pump pressure then starts to act on piston end 5. Piston 5/2 (Figure 27.8b)
then makes its stroke and the volume of grease under piston end 2 is discharged through outlet 2a. Then
piston 6 becomes pressurized and starts its stroke and discharges the quantity of grease under piston
end 3 through outlet 3a. At this point, the pistons will start to move in the reverse direction. A full cycle
is completed when all the pistons have made a back and forward stroke.
Table 27.1 summarizes the type of central lubrication system applications and recommended
grades.

27.6  Oil Mist Systems


The first use of an oil mist system is actually quite older than most people understand with the origi-
nal development dating back to the 1930s in Europe. The original concern had to do with inability to
Lubricant Application 27-9

TABLE 27.1  Central Lubrication Systems


Type of System Oil Greasea Applicationa
Single line Yes NLGI 000 up to NLGI 2 Small- and medium-sized machines
Multiline Yes NLGI 000 up to NLGI 3 Small machines with a low number of lubrication
points (compressors, pumps, chain lubrication,
presses, etc.)
Dual line Yes NLGI 000 up to NLGI 3 On equipment where the lube points are widely
separated (paper machines, steel industry, mining
equipment, etc.)
Progressive Yes NLGI 000 up to NLGI 3 On equipment with several lubrication points within
small to medium distances
a Typical—Grease grades and applications will vary with individual suppliers.

satisfactorily lubricate high speed bearings. Spindle speeds were obviously too high for grease lubri-
cation and traditional oil bath generated too much heat through fluid friction. An oil mist system is
a different, however, very affective type of centralized lubrication system capable of reducing overall
oil consumption up to ~40% versus other methods. Oil mist is not a vapor but rather an aerosol or
mixture of air with finely dispersed oil droplets of 1–3 μm in size. Typical oil mist is about one part oil
to 200,000 parts air. Because these droplets of oil are so small, they are easily transported by the air
through small tubing over fairly long distances to bearings or lubrication point. Fortunately, due to the
very lean makeup of the air–oil mixture, the aerosol or oil mist system is not able to sustain combustion.
The aerosol is created by passing high velocity air through an orifice or venturi that draws the oil into
the air stream causing it to break up into the finely dispersed particles. Once the oil reaches the bear-
ing or lubrication point, it is then condensed into larger size particles which are now large enough to
provide an adequate oil film to the desired component. Oil mist systems provide a number of important
advantages and benefits, including

• Constant supply of fresh or filtered lubricant


• Reduced bearing failures and increase in mean time between failure (MTBF)
• Reduced contamination due to slight pressurization of bearing housing
• Reduced manpower required for oil level monitoring of individual components, top-ups, and oil
changes
• Reduced overall oil usage
• Reduced bearing temperatures due to significant reduction in internal friction from oil sump
• Reduced fire risk
• Improved mechanical seal life
• Reduced power requirements (typically 3%)

As mentioned in the earlier list, one of the biggest benefits of an oil mist lubrication system is the reduc-
tion in the operating temperature of the bearing or component. Reductions in operating temperature
of 10°C–20°C (20°F–35°F) are not uncommon. This reduction in temperature can have a significant
increase in a bearings fatigue life with a typical increase of 11% with every 5.6°C (10°F) reduction in
temperature. Improvements in mean time between repair (MTBR) or MTBF from 24 to 36 months to 48
to 60 months are not uncommon with oil mist systems.
Another significant advantage to the oil mist systems is the ability to lubricate several components
with a single system. Systems with as many as 50 components are not uncommon. Oil mist systems are
an excellent method of lubricating very high speed bearings operating in the 10,000–15,000 rpm range
where conventional splash lubrication is less desirable or ineffective. However, oil mist systems are an
acceptable alternative for bearings operating at the more common 1800–3600 rpm range.
27-10 Lubricants

A further significant, but less often considered, advantage to the oil mist systems is the ability of
the system to protect the internal components such as the bearings and gears while the equipment is
shutdown or in standby. This is a significant advantage compared to traditional sump systems where
the non-oil-wetted components are subject to corrosion primarily due to ingress contamination during
temperature fluctuations. A common failure mode of spared equipment is fretting or false Brinelling
of the components from low amplitude, high frequency vibrations from the building. However, as pre-
viously mentioned, an oil mist system maintains the component under a slight pressure preventing
ingress of contamination as well as providing a constant supply of oil to coat and protect all the com-
ponent surfaces. A further benefit may be obtained by incorporating and commissioning an oil mist
system early during the construction phase of a new facility or during warehouse storage to provide
long-term protection of idle equipment.
One of the main limitations and/or disadvantages of the oil mist lubrication systems is the general
lack of knowledge of how the system works and should be maintained. Obviously, maintaining the
proper suspension of the oil particles is very important in ensuring that the required component
receives the desired lubrication. Although gravity certainly has an effect on the settling of the oil
particles, this generally can take several minutes to occur. With the average oil mist velocities of
20 ft/s or higher, settling does not readily happen over relatively long distances. However, the proper
design of the piping systems (slope, pipe size, fittings, etc.) is important to ensure that premature oil
settling does not occur. In addition, one of the common errors of oil mist systems is the desire for
some operators to increase the flow rate of oil at the venturi due to the common misunderstanding
that more is better.
Industry guidelines, including API-610, clearly support and recommend the use of oil mist sys-
tems. Earlier additions allowed the use of injecting the oil mist into the center of the bearing cavity
(Figure 27.9). Starting with API-610 8th Edition, API recommends that the oil mist be dual injected
between the bearings and the seals (Figure 27.10). With the publication of the 10th Edition, the single-
point method is no longer included as an option.

Oil mist in

Oil mist Oil mist


out out

Oil mist Oil mist


out out

Oil mist and liquid


oil drain

FIGURE 27.9  Old style—oil mist introduced at midpoint of bearing housing. (Courtesy AESSEAL, Rockford, TN.)
Lubricant Application 27-11

Oil mist in

Oil mist in

Oil mist and liquid


oil outlet

FIGURE 27.10  API 610-compliant oil mist application at locations between the bearings and isolators. (Courtesy
AESSEAL, Rockford, TN.)

Bibliography
1. Totten, G. E., Handbook of Lubrication and Tribology, Volume 1, Co-Published STLE and CRC Press,
Boca Raton, FL, 2006, ISBN 978-0-8493-2095-8.
2. Gresham, R. M. and Totten, G. E., Lubrication and Maintenance of Industrial Machinery, Co-Published
STLE and CRC Press, Boca Raton, FL, 2008, ISBN 978-1-42008-935-6.
3. Ehlert, D., Oil mist lubrication in the hydrocarbon processing industry, Machinery Lubrication,
July 2001.
4. Bloch, H. and Shamim, A., Oil Mist Lubrication: Practical Applications, Fairmont Press, Lilburn, GA,
1998.
5. Reiber, S., Oil mist lubrication, in CRC Tribology Data Handbook, E. Booser (ed), CRC Press LLC,
Boca Raton, FL, 1997.
6. Alemite Corporation, Oil Mist Application Manual, Alemite Corporation, Charlotte, NC, 2004.
7. Towne, C., Practical experience with oil mist lubrication, Lubrication Engineering, 39, 496–502,
1983.
8. Trico Manufacturing Corporation, Product Catalogue, Pewaukee, WI, 2009.
28
Lubricating Grease
28.1 Introduction.....................................................................................28-1
28.2 Grease Chemistry............................................................................28-1
28.3 Chemistry of Complex Grease.......................................................28-3
28.4 Thickener Ramifications.................................................................28-5
28.5 Base Fluids Used in Lubricating Grease.......................................28-5
Chemistry of Polyalphaolefins  •  Ester Chemistry
28.6 Grease Tests......................................................................................28-8
28.7 Grease Rheology............................................................................ 28-11
Paul A. Bessette 28.8 Grease Specifications..................................................................... 28-14
TriboScience & 28.9 Conclusions....................................................................................28-15
Engineering, Inc. References...................................................................................................28-15

28.1  Introduction
All lubricating greases contain lubricating oil with various additives that carry most of the tribological
responsibility, and a solid thickening agent used to immobilize the fluid. Typical additives in grease
are antioxidants used to improve thermal and oxidative stability, rust inhibitors to mitigate corrosion,
and antiwear agents selected to ameliorate surface damage when loads, temperature, and speed induce
asperity contact. Commonly employed grease thickeners are shown in Table 28.1.

28.2  Grease Chemistry


Grease made with lithium 12-hydroxystearate (12-HSA) thickener is the most prevalent organic grease
type manufactured globally.

LiOH.H2O + HOOC–(CH2)10–CH(OH)–(CH2)5–CH3 Li+.–OOC–(CH2)10–CH(OH)–(CH2)5–CH3 + 2H2O


Lithium hydroxide 12-Hydroxystearic acid Lithium 12-Hydroxystearate Water
Monohydrate

The grease is prepared by reacting lithium hydroxide monohydrate with a stoichiometric amount of
12-HSA to produce lithium 12-HSA, a simple organic thickener, and water. Water is removed from the
vessel by heat used to accelerate the reaction.
The manufacturing procedure consists of the following steps.
The 12-HSA and a small amount of oil are added to the reaction vessel. The kettle contents are then
heated just sufficiently to melt the fatty acid. The lithium hydroxide monohydrate is mixed with deionized
water and added incrementally to the vessel to prevent the formation of uncontrollable amounts of foam.
The temperature is raised to ~200°C to dehydrate the vessel charge. Base oil is then added gradually with
sufficient time allowed between each successive addition for the kettle contents to thoroughly mix. The
oil needs to be absorbed by the thickener structure and this is best achieved through the gradual addition

28-1
28-2 Lubricants

TABLE 28.1  Grease Thickeners


Simple organic soap The most commonly used grease thickener
Dropping point usually below 200°C
Complex organic soap Produced by the reaction of a base with two or more carboxylic
acids. Dropping points are usually above 260°C
Polyurea Excellent thickener for high speed rolling element bearing
Application, low oil separation, metal-free thickener results in
improved oxidative stability at elevated temperature
Organo-modified clay Excellent thickener for high temperature greases
Heat is not required to produce grease, high dropping point
Thickener
Amorphous silica Highly efficient thickener for nonpolar fluids
Polytetrafluroethylene Excellent thickener for PFPE fluids and low temperature greases
Molybdenum disulfide Ideal for heavily loaded applications
Graphite Ideal for heavily loaded applications
Polymers Most recent advance in thickener technology
Wax In conjunction with other thickeners, useful to reduce oil
Separation

of the oil to the rigid thickener mass. If too much base oil is added to the kettle contents, a two-phase
system results that is incapable of proper mixing under the low shear agitation of grease-making vessels.
The grease is then cooled to ~100°C and specific additives are stirred into the grease. Post-reaction
processing may include milling, homogenization, deaeration, and filtration. Prior to being discharged
from the manufacturing vessel, a sample of grease is taken to determine unworked, P0 and worked, P60,
penetrations per ASTM D217. The results are used to grade the grease and make the necessary oil addi-
tions to bring the grease into the specification range.
Usually, manufacturers will deliberately formulate a grease on the firm side of a specification since
adding additional oil is a simple operation. It is important to realize that the consistency of any grease is
a function of the ratio of thickener to base oil. Typically, most grease formulations contain from 5% to
30% thickener. Using lithium 12-HSA as an example, a grease containing about 3% thickener would be
semifluid, a grease with ~8.5% thickener would be an NLGI Grade 2, and an NLGI Grade 3 grease would
result from using 12% thickener [1]. Once the thickener network has been established, the structure will
readily accept additional oil even at room temperature.
The reaction of lithium hydroxide monohydrate and 12-HSA is referred to as neutralization of a base
and a fatty acid. Since the chemistry implies 1:1 stoichiometry, the number of moles of each starting ingre-
dient must be equal to assure that no unreacted materials are present in the finished grease. Excessive base
or fatty acid would jeopardize the chemical and physical properties of the grease. Stoichiometry is usually
accomplished by determining the saponification number of the carboxylic acid according to ASTM D94.
The saponification number is determined by reacting a specific quantity of the carboxylic acid with
a known quantity of potassium hydroxide, KOH, and neutralizing the excess KOH with a standard-
ized solution of hydrochloric acid, HCl. The result provides an average molecular weight (MW) for the
composition of the carboxylic acid. Mathematically, the MW of the acid component derived from the
saponification number can be expressed as

56,100
MW =
SN

where
MW equals the average MW of the various carboxylic acids in the sample
SN is the measured saponification number
Lubricating Grease 28-3

Carboxylic acids like 12-HSA, are more expensive than triglycerides which are essentially esters of
long-chain carboxylic acids and the trihydric alcohol, glycerin. With a slight change in the reaction
stoichiometry, substantial quantities of grease are prepared from triglycerides and a suitable base. The
reaction chemistry can be illustrated as follows using hydrogenated castor oil as an example:

C17H34OHC(=O)–O–CH2
C17H34OHC(=O)–O–CH2 + 3 LiOH.H2O 3Li+.–O(O=)CHOC17H34 + C3H8O3 + H2O
C17H34OHC(=O)–O–CH2 Lithium 12-hydroxystearate Glycerin Water
Lithium hydroxide
Glycerol tri-(12-hydroxystearate) monohydrate

The earlier reaction illustrates how one molecule of triglyceride generates three molecules of lithium
12-HSA. When a triglyceride is reacted with a base like lithium hydroxide monohydrate, the process is
referred to as in situ saponification.

28.3  Chemistry of Complex Grease


Chemically, complex greases differ from other greases prepared from carboxylic acids and inorganic
bases in the number of acids employed to react with the base. It is customary to use both high- and
low-molecular-weight acids. Doing so greatly improves the thickener’s resistance to thermal stress.
Specifically, whereas a lithium 12-HSA grease may have a dropping point of 200°C, a lithium complex
grease will have a dropping point of 260°C or higher. A thermally robust thickener network allows the
grease to operate at higher temperatures without excessive softening or oil separation. To formulate an
aluminum complex grease, first determine the batch size and the percentage of thickener required. For
example, consider a 100 kg batch containing 10% aluminum complex thickener. The amount of thick-
ener required is determined by simple multiplication:

100 kg × 0.10 = 10 kg of thickener

Moreover, it is usually desirable to have 1.9 moles of carboxylic acid for each mole of aluminum. Stearic
acid, a saturated 18-carbon fatty acid, and benzoic acid are routinely used to formulate aluminum com-
plex greases. The reaction sequence is illustrated in Figure 28.1.
To obtain the required amounts of each ingredient necessary to satisfy the stoichiometry, proceed as
shown in Table 28.2 [2].

Initial step: A1-R’


A1-ROH
+ 3 C17H35COOH O O
O O
R’-A1 A1-R’
HO-R-AL A1-ROH Stearic acid O
O
Aluminum trimer

+ 3 R”-COOH 3 R’COO-A1-OOCR’”
Benzoic acid OH
Aluminum complex thickener
Where R’ = C17H35COO

and R’” = R”COO

FIGURE 28.1  Illustration of the chemistry used for the preparation of aluminum complex grease.
28-4 Lubricants

TABLE 28.2  Aluminum Complex Grease Thickener Requirements


Atomic or
Ingredient Molecular Weight Mole Ratio Unit Weight Kilograms
Aluminum 27 1.0 27 10 kg × 27/463 = 0.58
Stearic acid 279 1.3 1.3 × 279 = 363 10 kg × 363/463 = 7.84
Benzoic acid 122 0.6 0.6 × 122 = 73 10 kg × 73/463 = 1.58
Total 27 + 363 + 73 = 463 0.58 + 7.84 + 1.58 = 10

From Table 28.2, it can be determined that the chosen stoichiometry results in the desired 10 kg of alu-
minum complex thickener. However, additional work is required. Since the source of the aluminum is an
organic trimer and to formulate the thickener, only the aluminum portion of the molecule is needed, divide
the amount of aluminum determined in Table 28.2 by the percentage of aluminum in the trimer. The assay
of aluminum in the trimer is 5.3%. Therefore, the weight of trimer containing 0.58 kg of aluminum is

0.58 0.583 = 10.94 kg


Moreover, since an aluminum source that does not liberate isopropyl alcohol was chosen, the amounts
of stearic and benzoic acids must be reduced to compensate for the reduced activity of the aluminum.
This is accomplished by simple subtraction knowing that 2.9% of stearic acid has been added to the
trimer along with 0.65% benzoic acid. From Table 28.2, it is apparent that stearic acid is 78.4% of the
formulation and a reduction of 2.9% results in 75.5% stearic acid and the adjusted percentage of benzoic
acid is 15.1%. The final weight of each ingredient is shown in Table 28.3.
Note: Approximately 0.2 kg of stearic acid and 0.1 kg of benzoic acid are delivered by the trimer. Since
the kilograms of ingredients required for the thickener is 9.98 kg and the amount of thickener desired is
10 kg, the material balance assures the correct stoichiometry.
Formulating the grease requires that ~80% of the chosen base fluid be added to the grease vessel
along with the stearic and benzoic acids. The vessel contents are then heated to 90°C and care must be
exercised to prevent evaporation of the benzoic acid. When the acids are melted and the mixture is vis-
ibly clear, the aluminum can be added. The trimer should be warmed to 25°C to prevent the intractable
precipitation of aluminum dibenzoate. When all of the aluminum trimer has been added, the tempera-
ture is gradually increased to 200°C and held at temperature for ~30 min to complete the reaction. The
contents are then cooled to 100°C and the balance of the base oil is added along with additives. Post-
processing may consist of milling or homogenization.
Greases are characterized by the type of thickener used. For example, greases formulated with either
lithium 12-HSA or aluminum complex are referred to as organically thickened greases due to the pre-
ponderance of carbon atoms in these grease thickener systems. Inorganic thickeners are the other major
category of solid materials used to thicken lubricating fluids. Inorganic thickeners consist primarily of
organo-modified clay and amorphous silica. Grease prepared from inorganic thickeners exhibit high
dropping points, do not require heat in their preparation, and have a reduced tendency to separate oil
under thermal stress. It is the lamella morphology of hectorite or montmorillonite clay that is respon-
sible for their thickening efficiency [3].

TABLE 28.3  Adjusted Composition of


Aluminum Complex Grease Thickener
Ingredient Final % Weight
Aluminum 5.8 10.9 kg trimer
Stearic acid 75.5 7.6 kg
Benzoic acid 15.15 1.5 kg
Lubricating Grease 28-5

TABLE 28.4  Manufacture of Clay-Thickened Grease


1. Add ~90% of the required base fluid to the manufacturing vessel along with all of the clay thickener
2. Mix the two ingredients at high shear rate for ~30 min
3. Add the desired polar activator and continue mixing
4. Dissolve any additives in the remaining base oil and add to the mix
5. Mill or homogenize the mixture to form grease

Specifically, individual clay platelets have an aspect ratio of 1000:1. The sheets are 104 Å across and only
10 Å thick, a gram of clay has a surface area of ~800 m2. Moreover, since clay particles carry a negative
charge it is possible to neutralize the negative charge with oleophilic cations. This chemical modification
makes it possible to disperse the clay thickener in lubricating fluid and the affinity of the hydrocarbon
groups on the platelet surface reduces the likelihood of phase separation. Individual platelets of the
lamella structures need coaxing to sufficiently separate and disperse under high shear manufacturing
conditions. Polar additives typically employed for delamination include methanol, ethanol, and pro-
pylene carbonate. In addition, a small amount of deionized water is used to promote hydrogen bonding
between the dispersed clay sheets thus forming the thickener network. Although clay-thickened greases
do not require heating in their preparation, some heating is frequently beneficial especially with high
viscosity base fluids. Table 28.4 summarizes the manufacturing steps of clay-thickened grease.
One caveat about clay-thickened grease is that they are structurally vulnerable to certain additives.
For example, sulfonate rust inhibitors will destabilize the thickener network over time.

28.4  Thickener Ramifications


Since the vast majority of grease thickeners are discrete solids, greases are two-phase systems with oil as
the continuous phase. Although grease thickeners are selected to have an affinity for the oil, heat, pres-
sure, centrifugal force, gravity, density differences, and reorientation of the thickener network caused,
for example, by a rolling element bearing, all tend to induce agglomeration of the thickener.
Thickener agglomeration may promote lubrication starvation, noise, and premature metal fatigue.
For a given NLGI Grade of grease, less thickener is tribologically advantageous. Amorphous silica is
routinely used to thicken silicone and hydrocarbon fluids. Silica’s substantial surface area makes it an
extremely efficient thickener for a wide variety of nonpolar base oils. It has been used successfully to pre-
pare grease used in aerospace grade bearings, gels used as impact media in fiber optic cable, and auto-
motive connectors susceptible to fretting. Silica-thickened greases are not abrasive since amorphous
silica has no crystalline structure, based on x-ray analysis [4].
Polytetrafluoroethylene (PTFE) is also used as a grease thickener primarily for perfluoropolyether
fluids. PTFE is a highly crystalline fluoropolymer with a melting point in excess of 320°C and greases
prepared from PTFE exhibit unique physical properties. They are unsurpassed in terms of thermo-
oxidative stability, are nonflammable, and are inert to most chemicals and solvents, and depending on
the base fluid’s molecular architecture, PTFE-thickened grease is serviceable from −54°C to +250°C.
Grease formulated with PTFE and either PFPE or chlorotrifluoroethylene (CTFE) fluids are the only
lubricants suitable for use in contact with oxygen [5].

28.5  Base Fluids Used in Lubricating Grease


Fluids used in the manufacture of lubricating grease and some of their more common use are shown in
Table 28.5.
Since oils are the major constituent in a lubricating grease and shoulder most of the tribological
responsibility, a discussion of their properties is warranted. Petroleum and vegetable oils are derived
from natural sources and with modern refining techniques, petroleum oils are rendered suitable for an
28-6 Lubricants

TABLE 28.5  Partial Listing of Fluids Used to Manufacture Lubricating Grease


Fluid Use
Petroleum Full range of industrial applications when economics is the
prime consideration
Vegetable oil When environmental considerations prevail, e.g., railroads,
chain saws, and farm equipment
Synthetic esters Low temperature applications both military and automotive
Polyalphaolefins Numerous automotive and military applications where low
temperature serviceability is required. These fluids are
compatible with most plastics and elastomers
Silicones Bath fluids, damping greases, control cables, and numerous
industrial applications where plastics or elastomers are used
Polyglycols Gear oils and greases used in arcing electrical contacts
Polyphenyl ethers Grease prepared from PPE fluids possesses unsurpassed
resistance to ionizing radiation. Lubricants for gold contacts
Phosphate ester Hydraulic fluids and greases with improved flame resistance
Polybutenes VI improver for other fluids and greases that produce less
carbonaceous residue on thermal degradation
Perfluoropolyethers Nonflammable, fuel resistant, oxidatively stable lubricants for
aerospace, military, and numerous automotive applications
Alkylated naphthalenes Oil and greases with excellent resistance to oxidation
CTFE Nonflammable lubricants and good anti-seize properties

extensive range of industrial applications. These oils are economical and are the most common fluids
used globally to manufacture greases.
Synthetic oils are not found in nature and must be synthesized. Synthetic fluids are usually superior
to petroleum and vegetable fluids in one or more properties, for example, low temperature fluidity, ther-
mal and oxidative stability, flash point, resistance to ionizing radiation, and volatility.

28.5.1  Chemistry of Polyalphaolefins


Polyalphaolefins (PAOs) are synthesized via the oligomerization of monomers such as decene-1 in the
presence of a suitable catalyst. The reaction scheme can be illustrated as follows:
BF3 + ROH
H2C = CH – (CH2)7 – CH3 – (CH2 – CRH)X –
Decene-1
R = C7H14 and X = 5

If the MW of the fluid is about 529 g/mol, then X is ~5. This is determined by dividing the MW of the
PAO molecule, by the MW of the repeat unit. Therefore, this particular grade of PAO consists mainly
of tetramers and pentamers. The Lewis acid and alcohol catalyst are responsible for the narrow MW
distribution of PAOs. Less dispersity is conducive to improved low temperature fluidity and lower vapor
pressure. Post-synthesis processing consists of hydrogenation, to eliminate the last vestiges of unsatura-
tion, and distillation. PAOs are available in a range of viscosities as illustrated in Table 28.6.
Table 28.6 shows that viscosity and pour point increase as a function of MW [6]. Vapor pressure and
volatility decrease with increasing MW provided that distillation is done properly.
Greases made using PAOs exhibit excellent low temperature properties, have low volatility, and are com-
patible with the majority of plastics and elastomers. PAO greases can be readily fortified with antioxidants,
boundary additives, and corrosion inhibitors. Moreover, the molecular homogeneity of PAOs greatly facil-
itates grease making. Preform thickener technology further simplifies the process. The low temperature
fluidity of PAO-based lubricants is primarily due to the highly branched nature of the molecules.
Lubricating Grease 28-7

TABLE 28.6  Physical Properties of Various Grades of PAO


Fluid MW, g/mol KV at 100°C, mm2/s Pour Point, °C
PAO-2 287 2 Less than −70
PAO-4 437 4 −70
PAO-6 529 6 −68
PAO-10 632 10 −53
PAO-40 1400 40 −34
PAO-100 2000 100 −20

28.5.2  Ester Chemistry


Esters are synthesized by a completely different procedure than PAOs. Esters are formed by reacting
alcohol with a carboxylic acid in the presence of a suitable catalyst. The synthesis of bis(2-ethylhexyl)
sebacate may be illustrated as follows:

H+ CH3 – CH(C2H5) – C3H8– CH2 – OH + HOOC –(C6H12) – COOH R – OOC –(C6H12) – COO – R + 2H2O
2-Ethyl hexanol = R Octadecanoic acid Bis(2-ethylhexyl)sebacate and Water

H+ = p-Toluenesulfonic acid

Sebacate esters belong to a class of synthetic fluids known as diesters. These fluids have pour points
approaching −73°C and they have been extensively used for low temperature applications as formulated
oils and greases. Diesters have been supplanted by higher MW esters possessing lower volatility and
only slightly poorer low temperature usefulness. Esters have good solvency for additives and because of
their polarity have innate boundary lubricating characteristics.
Esters are relatively low cost, possess better thermal and oxidative stability than PAOs, but must not
be used with certain vulnerable plastics such as polycarbonate, polystyrene, polyvinylchloride, ABS,
and others. However, esters have been used extensively to provide seal swell with certain elastomers.
Perfluoropolyethers are the fluids of choice to formulate greases for severe service applications. PFPE-
based greases are nonflammable, may be used in the presence of pure oxygen, are resistant to dissolution
by fuels and hydrocarbon solvents, and are inert to mineral acids and strong bases. Although all PFPE
fluids are composed of carbon, oxygen, and fluorine, two distinct families exist: the branched or pen-
dant PFPEs and the linear PFPEs. Each fluid type possesses specific chemical and physical advantages
related to molecular architecture, but the linear fluids with their higher oxygen content, in this context
we can think of oxygen as acting as a molecular hinge, possess exceptional wide temperature service-
ability. Moreover, these fluids and their companion greases are able to function from approximately
−73°C to 250°C. Linear PFPEs have a viscosity index above 300 and fluids thickened with PTFE are
unsurpassed for their low temperature capabilities. Tables 28.7 and 28.8 list the physical properties of
several linear perfluoropolyether fluids.

TABLE 28.7  Physical Properties of Linear PFPE Fluids


Fluid Z03 Z15 Z25 Z60
KV at 100°C, mm2/s 5.6 28 49 98
KV at 40°C, mm2/s 18 92 157 355
VI 317 334 358 360
Pour point, °C −90 −80 −75 −63
Flash point, °C None None None None
Density, 20°C g/cc 1.82 1.84 1.85 1.85
Evaporation 22 h at 204°C, % n/a 1.2 0.4 0.2
28-8 Lubricants

TABLE 28.8  Typical Test Conducted on Lubricating Grease


Property Method Typical Value Comments
Color Visual Tan
Unworked penetration, P0 ASTM D217 260
Worked penetration, P60 ASTM D217 265–295 NLGI Grade 2 grease
Prolonged worked ASTM D217 >300 Lithium 12-HSA grease
penetration, P100K
Oil separation ASTM D6084 <5%
Evaporation ASTM D2595 <1% After 24 h at 100°C
PDSC ASTM D5483 >30 min at 210°C Grease with excellent oxidative stability
Wear prevention ASTM D2266 0.5 mm Grease fortified with antiwear agent
Water washout ASTM D1264 <5% at 79°C Good water washout

These fluids are identified as linear PFPEs by the prefix Z and the particular grade by the succeeding
number. Z25 refers to a linear fluid with a kinematic viscosity of 49 mm2/s at 100°C [7].
The molecular architecture of linear perfluoropolyether can be illustrated as

CF3 ‒O‒(CF2 ‒CF2 ‒O)X‒(CF2O)Y‒CF3

These fluids are copolymers and are prepared by the low temperature photooxidation of tetrafluoroeth-
ylene in the presence of oxygen. Unstable peroxides formed during the reaction are removed by subse-
quent fluorination and heating [8]. Another class of PFPE fluid frequently used for high temperature
grease applications are the K fluids. These fluids are made from the cationic polymerization of hexafluo-
ropropene oxide. The fluids are homopolymers and are further characterized as having a branched or
pendant molecular structure. The typical structure of a K fluid is illustrated as

CF3CF2CF2O‒[CF(CF3)CF2O]X‒CF2CF3

Similar to the linear fluids, the branched fluids do not contain hydrogen and are therefore nonflam-
mable since oxygen is unable to remove the more electronegative fluorine from the polymer backbone.

28.6  Grease Tests


Table 28.8 lists some common grease tests.
Greases are characterized by their chemistry and physical properties. For example, a grease formulated
from lithium 12-HSA and a PAO would be referred to as a lithium-thickened synthetic hydrocarbon.
Moreover, a grease formulated with amorphous silica and a dimethylpolysiloxane would be called an
inorganically thickened silicone. Grease is most often designated by the nature of the thickener and the
base fluid. Greases are more complicated physically than oils since they are composed of both solids and
fluids. However, numerous tests are available to characterize the physical attributes of various greases.
Table 28.8 is a partial listing of available tests. The color and appearance of a grease do not necessarily
relate to quality since trace level of color bodies or chromophores present in antioxidants can alter the
color of a grease with time. However, color and appearance should be mentioned in technical literature
since these attributes may vary with lot and line personnel are sensitive to such changes. Dyes are fre-
quently used to provide a distinctive color to a specific grease. The consistency of grease is measured in
accordance with ASTM D217 and similar standards issued by other countries. P0 refers to the unworked
penetration of a grease while P60 is the worked 60 stroke penetration. It is the P0 consistency that is expe-
rienced by dispensing equipment while the P60 is used to designate the NLGI Grade of the grease. For
example, a P60 from 265 to 295 represents an NLGI Grade 2 grease while a P60 from 310 to 340 designates
Lubricating Grease 28-9

a Grade 1 grease. The NLGI grease consistency classification extends from 000 to 6 representing nine
distinctive grades. Grease consistency is an extremely important property and is frequently not given
sufficient consideration when a grease is recommended for a particular application. Grade 2 and Grade
3 greases work well for numerous rolling element bearing applications; however, softer or even semifluid
greases are needed in sealed gearboxes.
The mechanical stability of grease is determined by prolonged working from either 10,000 strokes or
100,000 strokes per ASTM D217. ΔP100K-60 is the customary parameter used to asses grease mechanical
stability. Although some greases become firmer after P60 due to an improved dispersion of the thickener,
it is the nature of grease to soften after prolonged working.
Oil separation refers to the propensity of a grease to separate oil as a function of temperature, pressure,
and time. ASTM Method D6184 is the static cone and beaker test where ~10 g of grease is suspended in a
60-mesh cone supported above a Berzelius beaker. The test fixture is placed in a forced draft oven for 24 h at
100°C and the amount of oil that separates from grease structure is collected in the beaker and the percent
of separated oil is determined gravimetrically. Generally speaking, too much oil separation is detrimental
to the longevity of the grease since loss of the liquid component increases the consistency of the grease.
However, no oil separation can be tribologically disadvantageous for applications that require oil to sepa-
rate from the grease. For many greases, oil separation is self-limiting for a given temperature. The reason for
this behavior is that as oil is lost, the ratio thickener to oil changes resulting in a greater percentage of solids
that prevent additional oil loss. Figure 28.2 illustrates the oil separation behavior at some fixed temperature.
For an NLGI Grade 2 grease, it is unlikely that heat alone would result in 50% oil separation. Oil
separation is not usually conducted on soft grease due to the tendency for the thickener to also pass
through the 60-mesh cone.
Evaporation is the unwanted loss of material due to heat. All lubricants are composed of molecules
that are in constant motion at room temperature. The application of heat increases the molecular agita-
tion until some molecules are no longer able to remain in the liquid phase and escape from the surface
of the lubricant. Since greases are composed of molecules with varying MWs, low MW molecules are
the first to evaporate as temperature increases. Similar to oil separation, evaporation alters the ratio of
oil to thickener increasing grease consistency and giving the grease a dry appearance. Evaporation is
temperature and time dependent. It is also a function of MW and molecular homogeneity. Lubricants
formulated from ingredients with narrowly defined MWs are far less prone to incipient volatility prob-
lems caused by the loss of light ends. Grease volatility may be determined by ASTM D972 and ASTM
D2595. However, data on grease volatility may also be obtained using a Petri dish placed in a forced draft
oven and measuring weight loss gravimetrically.
Vapor pressure measures the tendency of a lubricant to evaporate at some temperature. Vapor pres-
sure is determined using an expression by Langmuir:

P = 17.14 G(T/M)1/2

10
9
8
% Oil separation

7
6
5
4
3
2
1
0
0 5 10 15 20 25
Days

FIGURE 28.2  Self-limiting oil separation with time.


28-10 Lubricants

TABLE 28.9  Physical Properties and Vapor Pressure of Base Oils


Base Oil Viscosity at 40°C, mm2/s Molecular Weight, g/mol Vapor Pressure, Torr
Naphthenic 100 335 4.2 × 10−4
PAO 32 529 4.6 × 10−8
Ester 57 528 1.4 × 10−8
PFPE 154 1600 9.1 × 10−10

where
P is the vapor pressure in Torr or mm of Hg
G is the mass loss in g/cm2/s
T is the absolute temperature in Kelvin, that is, K = °C + 273.15
M is the MW of the lubricant

Vapor pressure measured at three temperatures can be used to extrapolate a value at 25°C or some
other temperature. The data are linear if the extrapolated temperatures are not too high. Linearity
results when the vapor pressure is plotted against the reciprocal of the absolute temperature, K.
Table 28.9 presents physical data and vapor pressure on four lubricating fluids that are used to formu-
late lubricating grease [9].
Although lubricants are more volatile under vacuum conditions, there is no term for vacuum in the
Langmuir equation. A vacuum promotes greater lubricant volatility since it removes the atmosphere,
nitrogen, and oxygen molecules, from the surface of the lubricant. Since vapor phase molecules have
the same average kinetic energy, fewer lubricant molecules are returned to their source due to collisions
with molecules from the atmosphere.
The dropping point of a lubricating grease is the lowest temperature at which a drop of oil separates
from the grease structure under the conditions of the test. Greases with a soft consistency are prone to
low dropping points, regardless of the thermal robustness of the thickener, due to temperature-induced
oil separation. Moreover, a grease dropping is not its long-term upper use temperature. Many mod-
ern, industrial greases are formulated with thickener systems that resist melting at temperatures above
260°C during high temperature excursions in an application. However, simply because a thickener is
intact at 260°C, this does not imply that the grease, especially the base fluid, can withstand the ravages
of thermal and oxidative degradation at 260°C.
Pressure differential scanning calorimetry (PDSC) is a viable analytical tool to quickly assess the
thermo-oxidative stability of oils and greases. Testing exposes a few milligrams of grease to an atmo-
sphere of pure, dry oxygen maintained at 3500 kPa (500 psi) at 210°C or some other desired temperature.
Lubricant oxidation liberates prodigious amounts of energy that is readily detected by the calorimeter.
Failure is determined based on the ability of a particular grease formulation to endure the thermo-
oxidative stress as a function of time. Since calorimetry is sensitive to both heat liberation and absorp-
tion, it is also a useful instrument for determining the amount of thickener in grease provided that the
thickener melts at some temperature. When the amount of grease placed in the calorimeter is entered
into the instruments computer software, the endotherm due to the melting of the thickener can be
determined by peak integration. Since small quantities of grease are required, calorimetry is a technique
well-suited for monitoring grease behavior in rolling element bearings. If the grease is susceptible to
oxidation below the melting point of the thickener, a nitrogen purge through the calorimeter is required.
Figure 28.3 illustrates the exotherm resulting from the oxidation of a synthetic hydrocarbon after
42 min at 210°C in pure, dry oxygen maintained at 3500 kPa.
The wear-preventive characteristics of a lubricating grease are determined by ASTM D2266. The com-
monality of tribometry is that all these wear tests assess the tribochemical properties of the grease. The tests
are conducted under boundary lubrication where grease additives react with the metal surface producing
organometallic films that minimize mechanical wear. Additives used in grease that function as boundary
Lubricating Grease 28-11

Sample: MIL PRF 32014 lot JB070810 File: C....\MIL PRF 32014_JB070810.002
Size: 1.0000 mg DSC Operator: PAB
Method: MIL-PRF-32014_Method Run date: 14-Aug-2007 11:11
Instrument: DSC Q10 V9.4 Build 287
600

400
Heat flow (W/g)

200

42.88 min
0

–200
0 10 20 30 40 50 60
Exo up Time (min) Universal V4.3A TA instruments

FIGURE 28.3  Exotherm generated by thermal and oxidative degradation of grease.

lubricants typically contain phosphorus or sulfur. The fortification of grease with additives does not require,
as with oils, that the additive be soluble since the grease structure effectively suspends nonsoluble additives.
However, the optimum amount of any additive is subject to empirical verification. In the standard
procedure, ASTM D2266, three stationary steel balls are clamped in a test fixture and the test specimens
are covered with grease. An additional ball is attached to a rotating spindle and brought into contact
with the fixed balls. Testing is usually conducted for 1 h at 75°C under a load of 40 kg and a spindle speed
of 1200 rpm. Grease additive effectiveness is determined by microscopic examination of the three sta-
tionary balls. The wear scar on each specimen is determined in millimeters and the average is reported.
Wear scars below 0.3 mm are exceptional while a wear scar below 0.5 mm is good boundary lubrication
performance. Grease with scars above 1 mm may require further development.

28.7  Grease Rheology


One of the great attributes of grease is its rheology. Rheology is the study of the flow and deformation of
matter and grease behavior is rheologically unusual. The rheological world between Hookean solids and
Newtonian liquids is filled with strange materials including grease. Grease exhibits a rheological duality
in that it possesses both solid and liquid attributes depending on the shear field. The terminology used to
describe this behavior is viscoelasticity. The unique suitability of grease in many applications including
rolling element bearings is due to this rheological duality.
Consider the typical grease rheogram shown in Figure 28.4 illustrating the decrease in viscosity as a
function of shear rate.
The utility of grease, especially for rolling element bearings, is shown by the graph of apparent viscosity
versus shear rate. At low rates of shear, the grease behaves as a solid while at high shear rates its fluidlike
tendencies prevail. This behavior is responsible for maintaining the grease in place while a bearing is at rest,
but grease’s unique rheology reduces the amount of energy required to move the grease as speeds increase.
The behavior of grease in a rolling element bearing is rheologically ideal. The tendency of a grease to
become less viscous as a function of shear is known as shear thinning while thixotropy describes the
tendency of a grease to soften as a function of time under constant shear.
Most lubricating greases display shear thinning and thixotropy. A much less common rheological
phenomenon for grease is shear thickening. The condition is indicative of the grease structure forming
28-12 Lubricants

Apparent viscosity vs . shear rate


250

200
Apparent viscosity, Pa.s

150

100

50

0
0 5 10 15 20 25
Reciprocal seconds, S–1

FIGURE 28.4  Grease rheogram.

an improved network as a function of energy input. Rheologically, all greases are non-Newtonian mate-
rials possessing an apparent viscosity dependent on the rate of shear. Greases typically have a yield stress
below which they fail to move. However, at very high shear rates the apparent viscosity of a grease will
asymptotically approach the viscosity of the base oil. The contribution of the thickener at high rates of
shear may depend on how well the solid is dispersed within the base fluid.
The apparent viscosity of grease increases as a function of decreasing temperature. Both the base fluid
and thickener influence this behavior. Synthetic fluids are usually superior to petroleum base oils for
low temperature serviceability while PTFE is an excellent thickening agent used to formulate greases
intended for applications below −54°C. Figure 28.5 is an actual plot of apparent viscosity versus tem-
perature for two synthetic greases. Although it is difficult to state a priori whether a particular grease
will function as required in a subzero environment, it is clear that one of the greases shown in Figure
28.5 will be superior to the other in terms of low temperature serviceability at −40°C.
The chemical analysis of grease, base fluids, and additives is greatly facilitated today by Fourier
transform infrared analysis (FT-IR). Aside from translation due to thermal energy, molecules vibrate at

Apparent viscosity
5000
Apparent viscosity, Pa. s

4000
3000
2000
1000
0
–60 –40 –20 0 20 40
Centigrade, °C

FIGURE 28.5  Apparent viscosity of two synthetic grease with decreasing temperature.
Lubricating Grease 28-13

characteristic frequencies common to wavelengths found in the infrared region of the electromagnetic
spectrum. Since various thickeners have different absorption characteristics in the infrared, they can
be readily identified by FT-IR analysis as can base oils. For example, the FT-IR absorption spectrum of
a hydrocarbon fluid is very different from that of a silicone and even more so compared to a PFPE fluid
(refer to Figures 28.6 through 28.8).
FT-IR analysis is a viable analytical tool for both qualitative and quantitative work in the lubrication
laboratory.

0.20 PAO-4 2921.003


0.18 2955.381

0.16

0.14
2852.505
0.12
Absorbance

0.10

0.08
1464.996
0.06
721.35
0.04 1377.585

0.02

0.00
3500 3000 2500 2000 1500 1000

Wavenumbers (cm–1)

FIGURE 28.6  FT-IR spectrum of polyalphaolefin fluid.

1.3 Dimethylpolysiloxane-350 786.627


1.2
1.1
1.0
0.9 1010.178
0.8
Absorbance

0.7
0.6
0.5 1257.877
0.4
0.3
0.2 701.3
2962.394
0.1
0.0
3500 3000 2500 2000 1500 1000
Wavenumbers (cm–1)

FIGURE 28.7  FT-IR spectrum of silicone fluid.


28-14 Lubricants

PFPE fluid 1084.127


0.9 1184.910

0.8

0.7

0.6
Absorbance

0.5

0.4

0.3
684.
0.2
814.499
0.1

–0.0
3500 3000 2500 2000 1500 1000
–1)
Wavenumbers (cm

FIGURE 28.8  FT-IR spectrum of PFPE fluid.

28.8  Grease Specifications


Grease testing for specifications may be divided into two distinct categories. They are (1) testing required
on a batch basis and (2) testing done on an annual basis.
Some examples of testing that should be conducted on a batch basis are
• Unworked penetration
• Worked penetration
• Oil separation
• Dropping point
While the following tests could be run only on an annual basis:
• FT-IR
• Wear
• Oxidative stability
Grease manufacture, like any other industrial process, is subject to inherent variation and although
great care has been exercised in weighing ingredients into the grease vessel and the same lot of ingredi-
ents are used, grease consistency fluctuates on a lot basis and needs to be determined. The physical and
chemical properties monitored by annual testing are less prone to the vagaries of the manufacturing
process and are primarily intended to provide assurance that raw materials are behaving as anticipated.
For example, the proven behavior of an antioxidant or rust inhibitor would not be expected to crash
with lot number.
These comments must be tempered by the realization that the nature of the application should dictate
the level of routine testing. The lubricating grease in rolling element bearings used in a multimillion
dollar satellite requires much more attention than grease intended for a boat hitch.
Lubricating Grease 28-15

28.9  Conclusions
It has been said that lubricating grease represents a mature industry and to some extent there is validity
to that assessment. There has not been a great deal of dynamism in grease manufacture because it has
not been necessary. Lubricating grease services a wide range of industrial applications. Where grease
technology is anything but anemic is in the realm of new applications. Here the industry is exceedingly
vibrant. Forty years ago, automobiles had fittings for periodic relubrication while current vehicle has
none. This leap in technology from frequent lubrication to lube-for-life is the result of significant dyna-
mism in the products being continuously introduced.
Consider an automotive conductive plastic potentiometer where an inert PFPE grease is used due to
its resistance to fuel vapor, its ability to function from −40°C to 150°C, and its ability to reduce wear
between a plastic film and metal wipers without impeding current flow due to hydroplaning.
Great advances have also been made in producing clean grease through ultrafiltration. In the early
1980s, computer disk drives were supported by sintered metal bearings since they are easier to fabricate
to critical tolerances and are more economical than rolling element bearings. However, rolling element
bearings are advantageous for long-term, reliable lubrication.
When the technical difficulties of manufacturing a complement of rolling elements within the nec-
essary tolerances to accommodate the increased data density on storage media, the onus was on the
grease to maintain the required operational run-out. Spindles cannot wobble due to ball contact with
particulate contamination in grease if the read–write head is to access data sectors reliably. Removing
particulate contamination from grease down to the micron level helped to support the truism that the
computer industry refers to as Moore’s law.
Large and small grease manufacturers will continue using equipment that has proven so reliable
over the years but the greases produced will continue to improve and support technological innovation
everywhere.
With sustained grease innovation and improvements, we can prolong the service life of machines,
reduce energy consumption, and reduce the economic impact of friction and wear globally.

References
1. NLGI Grease Grading System.
2. H. Kruschwitz, Thickener systems for aluminum complex grease, NLGI Grease Education Course,
1992.
3. C. Crozier, Clay based grease, NLGI Grease Education Course, 1992.
4. Degussa Technical Literature.
5. DuPont Technical Literature.
6. Chevron Phillips Technical Literature.
7. Solvay Solexis Technical Literature.
8. G. Allen, S. Aggarwal, and S. Russo, Comprehensive Polymer Science, Pergamon Press, Oxford, U.K.,
1992.
9. P. Bessette, Vapor pressure of lubricants, NLGI Spokesman, 58(3), June 1994, 8–12.
29
Solid Lubricants
29.1 Introduction..................................................................................... 29-1
29.2 Common Solid Lubricants............................................................. 29-1
29.3 Application Methods....................................................................... 29-2
Robert W. Bruce 29.4 Uses....................................................................................................29-4
GE Aviation References.....................................................................................................29-5

29.1  Introduction
The key characteristic of solid lubricants is that these solids provide a low friction coefficient due to
their low shear strength, often along with low hardness. Numerous such materials were identified, but
for reasons of environmental, health, and safety concerns the number of viable solid lubricants in use
is diminished. Lead oxide PbO, antimony trioxide Sb2O3, and selenides are no longer considered suit-
able materials. Meanwhile the use of solid lubricants and the methods by which solid lubricants are
applied to surfaces are increasing, especially due to the use of physical vapor deposition (PVD) methods.
Increasing operating and component temperatures are limiting the use of grease or ability to relubricate
surfaces, leading to increased use of solid lubricant solutions. The effect of humidity on the effectiveness
of some solid lubricants remains problematic.

29.2  Common Solid Lubricants


The most commonly used solid lubricants are Teflon or polytetrafluoroethylene (PTFE), molybdenum
disulfide (MoS2), and various forms of carbon/graphite (C).
PTFE was invented/discovered in 1938. Big blocks of it were cast and machined into bushings. But
its mechanical properties are marginal. Now it is used dispersed in numerous polymeric wear mate-
rial formulations specifically polyamide (nylon), polyimide which has the higher PV limit (Blanchet,
1997), and polyetheretherketone (PEEK), as coatings both topical and 25% dispersed in nickel plating,
and dispersed in oil or grease. Its main limitations are 275°C maximum use temperature and limited
mechanical properties.
Molybdenum disulfide was used to lubricate axels of horse drawn wagons traversing the Rocky
Mountains on their way to California in the 1700s and 1800s. It is being used as a burnished film on
metal surfaces, dispersed in oil, and encapsulated in polymeric or ceramic binder systems to form dry
film lubricant (DFL) films. It is also applied by PVD methods. It is limited by a maximum temperature
of 350°C beyond which it oxidizes rapidly. It is less effective in humid environments. The structure
of MoS2 allows for easy shearing between the planes, allowing for the low coefficient of friction (see
Figure 29.1).
Tungsten disulfide has a similar structure to MoS2 and has a slightly higher temperature capability
at 400°C.

29-1
29-2 Lubricants

Mo

FIGURE 29.1  Structure of molybdenum disulfide. (Reprinted from Schey, J.A., Tribology in Metalworking,
Friction Lubrication and Wear, American Society for Metals, Metals Park, OH, 1983.)

Graphite was used as a solid lubricant in metalworking when around 1900 the synthetic carbon
manufacturing process produced the high purity grade required for lubrication of bearings and other
engineered sliding surfaces.
Graphite has a lamellar structure similar to MoS2 (see Figure 29.2). Some humidity or organic vapors
are required to facilitate the shearing of the plates, and the use of graphite in vacuum, space (Fusaro 1999),
or even at 10,000 m altitude leads to a doubling of friction coefficient and order of magnitude increase in
the wear rate (Savage and Schaefer, 1956; see Figure 29.3). Graphite oxidizes rapidly above 500°C.
Metal-impregnated carbon and graphite dispersed in metals have found widespread use in applica-
tions at high temperatures. Graphite can be dispersed in grease.
Carbon coatings applied by PVD and also chemical vapor deposition (CVD) methods are reducing
the risk of galling in liquid-lubricated contacts. Various versions of often metal-doped carbon films are
being used and investigated, but their properties are significantly affected by changes in humidity, and
have therefore found little use without the protection of liquid lubricants.
Properties of these most common solid lubricants are summarized in Table 29.1.
Other materials can exhibit low shear strength at temperatures close to their melting point. Some
of these are indium (In) (MP 155°C), boron oxide (B2O3) (MP 450°C), silver (Ag) (MP 961°C), and
gold (Au) (MP 1063°C). Still other solid lubricants are talc, hexagonal boron nitride (BN), and Babbitt.
Babbitted sleeves are used in hydrodynamically lubricated bearings to facilitate low friction during
start-up. Modern Babbitt formulations do not contain lead.

29.3  Application Methods


PTFE is widely used dispersed in polymeric materials to reduce friction. It can also be applied as coat-
ings over metal and polymeric surfaces. Using an entrapment plating technique, it can be dispersed
through a nickel plate. It can also be dispersed in a polymeric or ceramic binder to form a DFL coating of
about 25 μm thickness. The binder, and suitable surface pretreatment, assures bonding to the substrate,
and the binder can protect the dispersed solid lubricant from oxidation till it is exposed by wear of the
Solid Lubricants 29-3

c
B (6.70 )

3.35 

1.42 
a
(2.46 )

FIGURE 29.2  Structure of graphite. (Reprinted from Paxton, R.R., Manufactured Carbon: A Self-Lubricating
Material for Mechanical Devices, CRC Press, Boca Raton, FL, 1979.)

10–5 10
–4
10
–3
10
–2
10
–1

0.10 Ib/in.2

0.0035

0.08 Isopropanol Carbon


n-Heptane n-Pentane tetrachloride Methanol Propane 0.0030
Bromo-
pentane
Wear rate, mm/min

Water 0.0025
Wear rate, in/min

0.06

0.0020

0.04 0.0015

0.0010
0.02
0.0005

0.0
0.1 1 10 100 1000
Pressure, N/m2

FIGURE 29.3  Wear of graphite as function of vapor pressure. (Reprinted from Paxton, R.R., Manufactured
Carbon: A Self-Lubricating Material for Mechanical Devices, CRC Press, Boca Raton, FL, 1979.)
29-4 Lubricants

TABLE 29.1  Friction of Solid Lubricants between


Compatible Solids
Friction Maximum Continuous
Solid Lubricant Coefficient Use Temperature (°C)
PTFE 0.05–0.1 250–275
MoS2 0.1–0.25 300–350
WS2 0.1–0.5 350–400
Graphite 0.1–0.5 500

binder matrix and has to function as the low shear layer. DFLs can usually be applied by dipping, brush-
ing, or spraying, followed by air-dry or heat cure.
Molybdenum disulfide and tungsten disulfide can be burnished onto metal surfaces, dispersed
through metal wear surfaces by powder metallurgy methods, and dispersed in a polymeric or ceramic
binder to form a DFL coating of about 25 μm thickness. The ceramic matrix may be based on sodium
silicate (Na 2SiO3), aluminum phosphate (AlPO4), or titanium oxide (TiO2) in the order of increasing
resistance to water. MoS2 is also used to coat metal surfaces by applying PVD processes including sput-
tering, forming a columnar structure that requires some running in to generate optimal friction con-
ditions. The sputtered coating thickness is usually measured in micron or submicron, and benefits of
co-deposition or multilayer deposition with metal (Cr, Co, Ni) have been reported.
Carbon/Graphite bushings are used in high temperature applications. Graphite has also been dis-
persed through metal alloys, particularly bonzes to provide reduced friction. It can be used to lubricate
aluminum but can cause pitting corrosion of aluminum in humid environments. Porous graphite bush-
ings have been infiltrated with liquid metals to provide improved mechanical properties.
Graphite can also be burnished and used in DFLs, sometimes combined with MoS2 and/or PTFE.
PVD and CVD applied carbon coatings, especially diamond-like carbon (DLC) coatings are find-
ing application in liquid-lubricated contacts of automotive engines to minimize the chance of galling
during adverse hydrodynamic conditions. The properties of these often metal (Ti, W)-doped films are
sensitive to changes in humidity and have thus far not found application without the liquid lubricant.
Some work on diamond coating has shown that these coatings are creating very rough surfaces, suit-
able to provide abrading surfaces but not yet effectively used for long life wear applications without
expensive surface finishing operations.

29.4  Uses
PTFE coatings are used to reduce the friction and therefore pain associated with inserting a teflon
coated hypodermic needle, as a nonstick coating for frying pans, and as an ingredient in carpet cleaners
to help lift the dirt out of the carpet.
PTFE dispersed through polymeric materials and some bronzes help reduce friction and wear of
bushings and hinges.
Dry film lubricants using PTFE, MoS2, and graphite in polymeric and ceramic composites are
described in various military specifications such as MIL-PRF-23398, MIL-PRF-46147, MIL-PRF-835645,
for a large number of applications.
MoS2 is dispersed by powder metallurgy methods through metallic bushings.
Graphite bushings are used in high temperature conveyor systems and furnace doors and so are the
metal-impregnated graphite bushings.
Magnetic storage disks are coated with carbon films under the thin layer of liquid lubricant. Carbon
films are also used on oil-lubricated piston pins and fuel-lubricated carburetor parts that are increas-
ingly at risk of galling due to the reduced levels of sulfur in fuel. Carbon films are also used to reduce the
risk of galling in automotive gear systems.
Solid Lubricants 29-5

Babbitt, and especially new formulations without lead and antimony, is used to line the hydrody-
namic bearings supporting automotive drive shafts. This reduces friction and wear during start-up
when the shaft and bearing are in contact.
Boron nitride is used as a separating agent to prevent threaded fasteners from seizing during long-
term use at high temperature.
Gold continues to be used in sensitive electrical contacts of chips and switches.

References
Blanchet, T., Friction, wear, and PV limits of polymers and their composites, Tribology Data Handbook,
E.R. Booser, ed., CRC Press, Boca Raton, FL, 1997.
Fusaro, R.L., NASA Space Mechanisms Handbook, NASA/TP 1999-206988, Washington, DC, 1999.
Paxton, R.R., Manufactured Carbon: A Self-Lubricating Material for Mechanical Devices, CRC Press, Boca
Raton, FL, 1979.
Savage, R.H. and Schaefer, D.L., J. Appl. Phys., 27, 136, 1956.
Schey, J.A., Tribology in Metalworking, Friction Lubrication and Wear, American Society for Metals, Metals
Park, OH, 1983.
30
Metalworking Lubricants
30.1 Functions of Metalworking Lubricants........................................30-1
Metal Removal Fluids  •  Metal-Forming Fluids
30.2 Types of Metalworking Fluids.......................................................30-3
Mineral Oil–Containing Fluids  •  Non-Mineral-Oil-Containing
Fluids  •  Solids and Gases
30.3 Selection of Metalworking Fluids.................................................30-5
Type of Shop  •  Type of Machines  •  Operation  • ​Material  • ​
System/Filtration  •  Water  •  Contamination  •  Disposal/
Recycling  •  Side Effects  •  Chemical Restrictions  •  Cost
30.4 Metalworking Fluid Control..........................................................30-8
In-Plant Controls  •  Laboratory Tests
30.5 Metalworking Fluid Application.................................................30-12
Flood  •  High Pressure  •  Mist/MQL  •  Manual
30.6 Metalworking Fluid Failure Mechanisms..................................30-13
Corrosion  •  Foam  •  Rancidity  •  Stability  •  Residue  •  Lubricity
Gregory J. Foltz
CIMCOOL Industrial 30.7 Health, Safety, and Regulatory Aspects.....................................30-15
Products, LLC References................................................................................................... 30-16

Metalworking lubricants are broadly divided by function into two areas: those used for metal removal
and those used in metal forming. Metal removal operations consist of machining and grinding, where
metal is removed and chips are formed. Stamping and drawing are metal-forming operations where the
shape of the metal is changed and in most cases, chips are not generated.1

30.1  Functions of Metalworking Lubricants2


30.1.1  Metal Removal Fluids
The primary function of a metal removal fluid is to control heat. A cutting tool generates temperatures
of 375°C–750°C and the resulting chip, as it slides up the tool face, creates tremendous pressures (up to
1,379,000 kPa). About 75% of the heat is generated by deformation of the metal, and the other 25% by
friction between the chip and the tool. By controlling the temperature generated in the cut zone, tool
wear can be controlled and tool life increased.
As the cutting tool cuts (or the grinding wheel grinds), metal deforms by shear or plastic flow along a
shear plane extending from the top of the tool to the surface of the metal (Figure 30.1). Below the shear
plane is undisturbed metal; above it, the deformed metal forms a chip. Reduced friction at the chip–tool
interface increases the shear angle, produces a thin chip, and deforms less metal.
When a tool face is examined under a microscope, rough peaks and valleys can be seen (Figure 30.2).
These tiny projections collide with the chip as it slides up the tool face and weld to the chip under the

30-1
30-2 Lubricants

Thin
chip

Large
Thic
chip k
shear
angle
Tool
Small
shear
angle

Long path Short path


of shear of shear

Work piece

The amount of deformation of the metal is controlled by


the size of the shear angle and the amount of heat produced
is inversely proportional to the shear angle.

FIGURE 30.1  Shear angle.

Chip

Tool
Cutting
fluid

0.0001

FIGURE 30.2  Tool and chip surfaces.

conditions of very high heat and pressure. Continuous shearing of these welds results in tool wear, the
tip of the tool becomes cratered, and heat concentrates at this point. Small pieces of sheared-off metal
form a built-up edge on the face of the tool—a major cause of poor finish.
When a metal removal fluid is introduced between the tool and the chip, friction is reduced, the shear
angle increases, the chip becomes thinner, the power requirement is reduced, and less heat is generated.
Also, the built-up edge disappears, the finish smoothes out, and size control improves. The nascent
metal exposed under the high temperature and pressure conditions reacts with chemicals in the fluid to
form a low shear strength solid between the chip and the tool. The chip slides freely up the tool face, tools
last much longer, speeds and feeds can be stepped up, and much more work can be performed with each
tool. The cooling and lubricating mechanisms are dependent on the job: slow-speed, slow-feed opera-
tions need more lubrication, while high-speed, high-feed operations need more cooling.
The other major function of the fluid is to provide lubrication between the tool and the workpiece,
thereby reducing the heat generated by friction, controlling deformational heat of the chip, and reduc-
ing the energy required to perform the operation. Laboratory tests exist for evaluating and rating the
Metalworking Lubricants 30-3

lubricating properties of metal removal fluids. Many of these are just basic machining or grinding tests
run on an instrumented machine tool. Others are built around test machines3 and rely on a specific
operation such as tapping.

30.1.2  Metal-Forming Fluids2,4–6


The primary function of a metal-forming lubricant is to provide separation of the tooling and the work-
piece surface. This is necessary for reducing wear and protecting the dies, as well as limiting adhesion
(welding) of the die to the part. The lubricant must also control the friction between the part and the
die. Lowering friction is desirable; however, in some forming processes a limited amount of friction is
needed to allow for metal movement (rolling) or support (drawing). In many drawing operations, there
is a substantial enlargement of surface area, so the lubricant must be able to protect this new area as well
as the old.
Extreme pressure (EP), boundary, and hydrodynamic lubrication are involved in both metal forming
and metal removal and are described in more detail in other areas of this book.

30.2  Types of Metalworking Fluids7


Metalworking fluids can be differentiated in several ways, either based on performance properties or
based on composition. Performance rates fluids on the severity of the operations that they can handle,
usually from light duty to heavy duty. Composition can be as simple as whether they are designed to be
used neat (undiluted) or diluted with water. Typically, the types of fluids are defined by their composi-
tion, in terms of whether or not they contain petroleum (mineral) oil in the formulation, or if they are a
gas or a solid. ASTM Standard D2881-03 is the basis for these classifications.8

30.2.1  Mineral Oil–Containing Fluids


30.2.1.1  Straight Oils
Straight oils are composed of mineral oil and may include additives such as fats to help with wetting or
EP agents to improve lubricity. EP additives include sulfur, chlorine, and phosphorus and are used to
improve anti-weld properties for heavy-duty operations. Sulfur forms a better lubricant, but chlorine is
more reactive than sulfur and breaks down to form the EP lubricant at lower temperatures. Phosphorus
is effective at a lower temperature. Sulfur is frequently found in grinding fluids, chlorine in cutting
fluids, and phosphorus in forming fluids, although uses in all applications and combinations of these
materials are found in many fluids. Straight oils are often classified as active or inactive due to the sulfur
content: an inactive oil will not darken a copper strip immersed in it for 3 h at 100°C, while an active oil
will. Straight oils are used as supplied from the manufacturer, that is, they are not diluted.
Straight oils generally provide the excellent lubrication needed in low-clearance, low-speed opera-
tions, especially where a high quality surface finish is required. The straight oils do allow the buildup of
excessive heat, since oil dissipates heat only half as fast as water and because it is a more viscous liquid.
They have good ferrous corrosion control but need to be checked for stain on nonferrous materials.
Sump life of oil is relatively long, because rancidity is not a major issue, since most bacteria do not grow
in oils unless they are contaminated with water. These oils can also be a safety hazard, as they may
smoke and, in some cases, even burn. In addition, their high misting properties can cause parts and
surrounding areas, to become slippery and dirty if mist levels are not properly controlled.

30.2.1.2  Emulsifiable Oils (Soluble Oils)


These products generally contain greater than 30% of oil in the concentrate before blending with water
for use at a typical mix ratio of 1:10–20. The oil is made dispersible by the addition of emulsifying
agents, primarily sulfonates. The emulsified particles range in size from 80 to 200 μm, which are large
30-4 Lubricants

enough to reflect light and create a milky, opaque appearance when mixed with water. Premium grades
may contain corrosion inhibitors, EP additives, cleaning agents, microbicides, and other performance-
enhancing additives. Emulsified oils are typically used as general purpose products for all types of appli-
cations. They offer good lubrication because they contain oil, and have good cooling properties because
they also contain water. There are several disadvantages to emulsified oils. When mixed with hard water,
some products can form a precipitate and in extreme cases, the emulsion can be broken. At high concen-
trations, mists from a soluble oil can leave machine and work areas in a messy and slippery condition.
Water can support bacterial growth, leading to rancid odors and short sump life if proper bactericides
are not present.

30.2.1.3  Semisynthetics
These products have lower (<30%) mineral oil content than an emulsified oil. Semisynthetics, in use,
are diluted with water, also at a typical mix ratio of 1:10–20. They have a higher emulsifier content and
smaller oil droplet formation, which results in a mix that is translucent or transparent. Semisynthetics
will also contain many additives to enhance performance including corrosion inhibitors, EP lubricants,
wetting and cleaning agents, microbicides, and defoamers. They generally have enough lubricity for
moderate- to heavy-duty applications. Semisynthetics have better cleaning and settling properties than
emulsified oils, which contribute to a long and relatively trouble-free sump life. Because they contain
very little oil, these products rarely smoke and generate less oil mist. If hard water is used to make the
mix, these products, like soluble oils, may form a hard water scum. The enhanced cleaning action may
cause some semisynthetics to foam.

30.2.2  Non-Mineral-Oil-Containing Fluids


By definition, these fluids do not contain any petroleum-based oil. They may contain oil from another
source such as plants, trees, or vegetables, or may contain no oil at all. Soy, canola, and rapeseed oils are
some of the non-petroleum oils used in these types of fluids, as well as esters.

30.2.2.1  Solution Synthetics


These products are true solutions and are completely clear when mixed with water. They basically con-
sist of organic and inorganic salts dissolved in water and may include corrosion inhibitors, chemical
lubricants, wetting agents, and microbicides. Many of these fluids are designed to reject tramp oil that
may leak into them from exogenous sources, and are known as oil-rejecting products. Some synthetics
are designed specifically for nonferrous metal and have an operating pH in the 7s, while most metal-
working fluids are in the 8.5–9.5 range.
Very basic types of these fluids are used as light-duty grinding fluids, since they offer rust protection
and good heat removal. The more complex types are capable of heavy-duty operations as they offer
excellent lubrication as well. The transparency of these fluids is an advantage in many operations since
it allows the operators to see the workpiece. Solution synthetics are the least troubled by rancidity and
their superior settling and cleaning properties help extend fluid life. Their excellent cooling capability
makes possible high feeds and speeds, high production rates, and good size control. Solution synthetics
are stable, even in hard water.

30.2.2.2  Emulsion Synthetics


These products are the same as the oil-containing emulsified oils and semisynthetics except that min-
eral oil has been replaced by oil from another source. The formulations and performance are basically
the same as found with their oil-containing counterparts. The lubrication characteristics of these oils
and esters are excellent, but they must be balanced against their long-term oxidative and hydrolytic
stability.
Metalworking Lubricants 30-5

30.2.2.3  Straight Synthetic Oil


Straight synthetic oils are also the same as straight mineral oils, with the exception that the oil is non-
petroleum. These products are used neat, that is, not diluted with water. They can also be enhanced with
a package of EP lubricants.

30.2.3  Solids and Gases


Compressed air can be directed into the cut zone, both for heat removal and for removing chips. Inert
gases and carbon dioxide may also be used. In some forming operations, dry coatings and stearate
powders are used. In can drawing, the tin plating on steel can act as a lubricant. Dispersions of molyb-
denum disulfide and/or graphite in mineral oil are also used in some drawing applications. As a result
of advances in the field of nanotechnology, nanoparticles are being applied to enhance the lubricating
properties of both metal-forming and metal removal fluids.9

30.3  Selection of Metalworking Fluids2,10,11


Selection of the proper fluid is basically related to two parameters: the severity of the application (light,
moderate, heavy, or extremely heavy duty) and the type of material (cast iron, steel, aluminum, tita-
nium, etc.).12 Knowing these two factors will give you most of the requirements for selecting the correct
product, but there are many other factors as well.

30.3.1  Type of Shop


A general purpose fluid is the logical choice for job shops with their wide range of parts, materials,
machines, and operations. If one fluid can handle all of these variables, then fluid selection, inven-
tory control, and fluid management are much easier to accomplish. High production shops, with many
machines or lines doing the same operation on the same part, can use a fluid designed to meet very
specific requirements.

30.3.2  Type of Machines


The design of some machines requires the metal removal fluid to act as a lubricant for the slides and
other moving/rubbing parts. In other cases, the fluid may need to function as the gear box lubricant in
addition to being the cutting fluid. On some machines, and from some machine tool manufacturers,
comes the concern that only oil-containing fluids or high oil content fluids be used, and the machine
specification is written to that effect. The type of seals, elastomers, and plastic type components used
on the machine should also be checked for compatibility with the fluid.13 It is important to know these
machine considerations before deciding what fluid should be used.

30.3.3  Operation
The severity of the operation will dictate the lubricity requirements of the fluid. As severity increases,
it may also be necessary to increase the lubricity (physical or chemical) of the fluid. Stock removal
rates, feeds, speeds, and finish requirements of parts must also be considered. Metal removal operations
are divided based on severity: light duty (milling cast iron or surface grinding steel), moderate duty
(turning steel or aluminum), heavy duty (centerless grinding or drilling steel), or extremely heavy duty
(broaching, form grinding steel).
30-6 Lubricants

Metal stamping/drawing operations are also divided based on severity: light duty (lamination and
thin gauge sheet metal stamping), moderate duty (roll forming and sheet metal forming), heavy duty
(fineblanking and deep drawing), and extremely heavy duty (extrusion and pilger milling).
If performed regularly, the most severe operation in the shop will dictate the fluid selection.

30.3.4  Material
The type of material being worked (cast iron, steel, aluminum, titanium, copper, glass, carbide, plastic,
etc.) is very important in fluid selection. The corrosion control and/or staining properties of some fluids
may not be compatible with all materials. Since various components or additives in a fluid may stain
some alloys of a given metal, tests should be run to insure compatibility. Magnesium requires extra
caution to minimize the evolution of hydrogen gas and potential fire or explosion hazards. Only fluids
specifically formulated for magnesium applications should be used.
Technology advances to reduce weight and/or increase strength have lead to the increased use of
many different metals including titanium, nickel-based alloys, compacted graphite iron (CGI), and high
strength–low alloy (HSLA) steel. The fluid requirements for machining and forming these types of met-
als include high lubricity and excellent heat reduction, which are both found in many high performance
products, especially synthetic formulations. Because many of these parts are used in specialized appli-
cations such as aerospace and medical, there may also be chemical restrictions and specialized testing
needed to insure compatibility of the fluid used with the particular metal.
Machinability ratings12 can be used to differentiate metals and select fluids. Variations in microstruc-
ture, amounts of alloying materials, heat treatment, and many other factors can affect machinability.

30.3.5  System/Filtration14
In most applications, the metalworking fluid will reside in some sort of tank or container. There are
some non-recirculating systems where the fluid is applied and not contained for reuse. These are fre-
quently found in stamping applications where the fluid is simply put onto the part before the operation
or in metal removal operations where the fluid is misted into the work zone. The immediate perfor-
mance of the fluid in terms of lubricity and perhaps corrosion control are the only concerns.
Many applications utilize either a sump on each machine or a central filtration system where many
machines are linked to one large tank. Different types of operations will generate different types of
particulate/chips/fines that must be removed. Machining generally creates the largest chips and some
may be rather stringy, while grinding tends to create very fine small particles, and drawing operations
can make a very fine particulate.
Filtration methods can include simple settling by gravitation, some type of positive filter using media
(cloth, paper, and wire screens), or a separator such as a magnet, centrifuge, or cyclone. Whatever
method is used will affect fluid selection. Settling tanks obviously require a fluid that has excellent
chip-settling properties. Media filters require a fluid that will pass through the media without clogging
or being stripped of essential constituents. Generally, filtration through these types of filters will be in
the 20–50 μm range, with some cases of finer filtration needed in honing operations or through the tool
coolant applications. In some positive media applications, a cellulose floc or other material is added to
improve and tighten the filter cake making the particle size found in the fluid even more critical.
Separators such as cyclones and centrifuges require fluids that will not split or excessively foam under
the demands of these dynamic units.
The capacity of a settling system and most other filters as well should ideally allow a minimum of
10 min retention time (system size divided by flow rate) in the tank for the fines to settle out, tempera-
ture to stabilize, and to give any entrained air or foam a chance to dissipate.
Premix tanks, proportioners, and other mixing devices are frequently used in conjunction with these
tanks to keep the fluid mixture at the correct concentration and volume level in the tank.
Metalworking Lubricants 30-7

30.3.6  Water
Water will be the main component (90%–95%) of almost all water-based metalworking fluid when it
is diluted for use. Therefore, the quality of this water will be a key factor in the performance of the
fluid. Water quality is a measure of its mineral content, pH, and overall cleanliness. Water hardness
(cations, typically Ca 2+ and Mg 2+) greater than 200 ppm can produce mix stability problems with
emulsion type products. Water with high anion (chloride, nitrate, phosphate, or sulfate) levels (greater
than 150 ppm) can promote corrosion and/or rancidity. Soft water (<50 ppm total hardness) can lead
to foam in fluids not designed to work in this low hardness type water. Mineral content can vary
depending on the water source (river, lake, and well) and the time of year. It can also build up as sys-
tems turnover and age.
The pH of water in the United States normally varies from 6.4 to 8.9, depending on the area and the
source. The buffering ability of a metalworking fluid is far greater than that of any clean water supply.
Adjustments to pH are rarely needed.
Many plants will use city water for diluting their water-based metalworking fluids, but depending
on the mineral content, it may be need to be treated before use. Only in the case of corroded pipes (iron
contamination) or if using lake or runoff water is cleanliness a problem. If microorganisms or other
foreign material is in the water, it should be removed or treated before use.
It is very important to determine water quality before selecting a product. In many cases, especially
where high mineral content can be problematic, the use of deionized (DI) or reverse osmosis (RO) water
is beneficial to fluid performance and system life and will usually more than pay for itself in the long
run, based on the overall cost of manufacturing a part.

30.3.7  Contamination
Lubricating oils, way lubes, hydraulic fluids, floor cleaners, heat treat solutions, and rust inhibitors are
among the items often found in metalworking fluid systems. These contaminants can contribute to bac-
terial growth, rust, concentration control difficulties, residue and foam, and worker dermatitis. Many
water-based synthetic fluids are designed to reject most of the oil contamination. Other fluids, especially
the emulsion products, are likely to emulsify some of this leaked oil. Cleaners (high pH) can increase
alkalinity and lead to dermal issues or foaming. While most fluids can handle some contamination,
greater amounts will shorten the fluid life. It is beneficial to know not only what contaminants may get
into the fluid but also how much, so that a fluid can be chosen to best handle these conditions.

30.3.8  Disposal/Recycling15
In many plants before a fluid is selected for use in any operation, it must first pass tests to insure that it
can be waste-treated for disposal or recycled for reuse, when its useful time has ended. Economic and
efficient compatibility with these methods is a key performance factor. Waste treatment methods can
be some form of chemical treatment or splitting of the mix (separating the oil from the water) by use of
acid, alum, and other polymers. Physical treatments to achieve the same separation include ultrafiltra-
tion, evaporation, centrifuging, or RO. Many plants that generate insufficient waste to justify their own
treatment facility have it hauled away by licensed (and regulated) commercial treatment companies.
Recycling consists of passing the used fluid through a unit designed to remove solids and tramp oil
and, in some cases, treat for microorganisms. These units are typically based on filters, centrifuges, or
coalescers. They may be on-site or mobile units that go from plant to plant. It is important to remember
that no recycling unit completely restores a fluid to its original condition. Additives or fresh concentrate
should be used to improve the recycled fluid’s performance. It is also important that when a recycling
program is in place, the plant must standardize on very few, perhaps just one, fluid, in order to gain any
economic benefits.
30-8 Lubricants

30.3.9  Side Effects


Fluids should be safe and generally pleasant to use. In some cases, a particular product color or odor is
desirable, and in some operations a transparent (synthetic) fluid is preferred so that the tool–workpiece
interface can be seen by the machinist. Fluids should not leave objectionable residues on parts, machines, or
contiguous area surfaces. They should provide adequate corrosion protection on parts for at least 72 h under
normal conditions. The metalworking fluid should not damage machine tool paint or machine components.

30.3.10  Chemical Restrictions


Certain industries, such as aerospace or nuclear power, may have stringent restrictions on chemicals
used in the metalworking fluid, especially halogens. In some cases, there may not be a restriction but,
rather, a limit for the fluid not to exceed a certain ppm level in the concentrate or the mix.
Some companies have lists of particular chemicals that are restricted in their plants because of some
health, safety, environmental, or manufacturing concern. The same holds true of different countries
throughout the world. Some chemicals are not on every country’s approved domestic chemical inven-
tory and, therefore, cannot be used or exported into that country. It is important to be aware of these
restrictions when selecting a fluid.

30.3.11  Cost2
It is important to consider not only the initial purchase price of the fluid but also the impact the fluid
has on plant operations and performance. The first point to review is the use dilution of the fluid or the
cost per mixed gallon. A product that can be effectively used at a 5% dilution may cost more than the
one that must be used at 10%, so consider what it costs to make the diluted mix. In this case, it would be
less expensive to use the higher priced product.
Other factors such as tool or die life, part quality, tank life, waste fluid disposal costs, and shop clean-
liness should all be factored into the total cost of using a particular product or comparing one product
to another. Expenditures for tools, grinding wheels, and dies are generally a larger portion of a plant’s
budget than a metalworking fluid. However, by using a fluid with optimum performance, these costs
can all be reduced. Fluid contribution to a reduction in the number of scrap parts due to poor finish or
corrosion issues can also be a significant savings.

30.4  Metalworking Fluid Control2


Many plants have a defined procedure for selecting and evaluating new metalworking fluids. This can
consist of some lab experiments to evaluate parameters such as corrosion and foam control, as well as
waste treatability. A thorough health and safety review of the MSDS is required. Then it is very likely
that machine testing will take place. Parts are cut or stamped and the performance factors such as
tool/die life, part finish, and operator acceptance are all compared to the current fluid. It is important
in these types of tests that reliable data be measured and used to make a decision. Selecting the correct
fluid is only one step in achieving good performance.
Properly controlling and maintaining the fluid are also very important considerations. These activi-
ties are achieved with a series of in-plant procedures as well as with laboratory assistance from the fluid
supplier.

30.4.1  In-Plant Controls


A good control program16,17 is essential to prolonging fluid life. It is much more critical for water-based
fluids than with straight oil products where there is not as much chance for errors. The following basic
steps are part of an effective fluid control program:
Metalworking Lubricants 30-9

1. Assign the responsibility for control to one individual


2. Clean the system/tank thoroughly before charging with a fresh mix
3. Maintain the concentration of the fluid at the dilution needed for that operation
4. Keep the metalworking fluid free of chips, grit, and contaminants
5. Use good quality water
6. If microbial growth is a concern, aerate the fluid mix by keeping it circulated when not in use
7. Avoid chip and debris buildup in the machines and trenches
8. Employ good housekeeping practices
9. Minimize/remove tramp oil contamination

30.4.2  Laboratory Tests


Metalworking fluids, specifically the water-soluble types, are all formulated to operate within certain
parameters involving concentration, pH, dirt levels, tramp oil, bacteria, mold, and conductivity. For
some fluids, the operating range in these areas is wide and for others it is quite tight. Typically, soluble
oils have a much wider range than do solution synthetics, which must be more tightly controlled. The
frequency of these tests can vary from plant to plant. For many operations, a daily check of concentra-
tion and pH with perhaps a weekly evaluation of the other tests is sufficient. In can drawing operations,
it is customary to run a series of checks every 4 h.
While most of these tests are common practices, the exact procedures are often specific to a particular
manufacturer or product. End users should always check with their suppliers on test methods for con-
trolling their particular product.

30.4.2.1  Concentration
Water-based metalworking fluids are typically formulated to operate at concentrations of 2%–10% (see
Figure 30.3), although higher concentrations, especially in some stamping and drawing applications,
are not uncommon. Concentration is the most important parameter of the fluid to monitor and control.
It is not an absolute value but, rather, a value based on a particular component of the product. It is pos-
sible to measure the fluid concentration based on alkalinity, anionic or cationic content, oil levels, or
based on some specific raw material used in the formulation. Different materials will build up or deplete
at different rates based on a wide range of variables (see Figure 30.4). Concentration can be measured by
the following methods.

Too strong possible problems: foam


residue,cost,cleanliness, and dermatitis
10
Upper limit
9
Concentration %

5 Typical operating range

3 Lower limit

1 Too lean possible problems: rancidity,


lubricity, corrosion, and cleanliness

Time

FIGURE 30.3  Concentration parameters.


30-10 Lubricants

Typical buildup and depletion curves


for water-diluted, metalworking fluid mixes

ts ete
r
tan ity
+ urfac alin ron ctom
Alk Bo fra
ic s Re ent
nion c ont
No Oil
Relative concentration

Uniform
0
Anionic surfactants

pH
Anion
– Cationic surfactants* ic lub
rican
ts
Bacteri
cides an
d fungic
ides

Time
Cationic surfactants are only used with transparent, cationic-compatible, synthetic products.

FIGURE 30.4  Concentration versus time.

30.4.2.1.1 Refractometer
This is an optical instrument that measures the refractive index of a metalworking fluid. The read-
ing obtained from the refractometer is converted to a concentration value by use of a factor or graph
made from readings taken on known mixes of various concentrations. Refractometers work very well on
clean, fresh mixes, but as contamination in the fluid increases, it becomes much more difficult to get an
accurate reading. Refractometers are frequently used as in-plant control methods, especially in smaller
facilities, because of the simplicity of their use.

30.4.2.1.2 Titration
Chemical titrations can be developed to measure specific components of a fluid mix such as anionics,
nonionics, or alkalinity. Once again, standards are run on mixes of known concentration to establish
control values. Large plants, which are equipped with chemical laboratories, will typically run titrations
on their fluid mixes at least once per week.

30.4.2.1.3 Instrumental
This method is typically utilized by the metalworking fluid supplier, at their plant, to accurately check the
level of a particular material or component in the fluid, and then extrapolate a value for the concentration.
High-pressure liquid chromatography (HPLC), Fourier transform infrared (FTIR), or gas chromatogra-
phy (GC) are some of the methods used. The assumption is that all ingredients will stay within a relative
balance, which in some cases may not be true as there can be selective depletion of product components.

30.4.2.2  pH
Metalworking fluids are typically formulated to operate in a pH range of 8.5–9.5, although some prod-
ucts, especially ones designed specifically for aluminum applications, can run in the range of 7–8. pH
is a measure of the alkalinity or acidity of a metalworking fluid mix. Higher pHs are better for ferrous
corrosion control but could lead to skin mildness issues and staining on nonferrous materials. Lower
pHs are better for mildness and nonferrous corrosion control but could lead to rancidity and ferrous
corrosion issues. So the pH is really a compromise to achieve the best performance.
Metalworking Lubricants 30-11

pH is a good indication of the fluid’s condition. Low pHs are generally an indication of biological
activity and high pHs an indication of contamination, usually by some type of cleaner.
pH values are checked using pH indicator paper or a pH meter. It is important to insure that pH
meters are properly buffered and kept clean, as the oils found in fluids can easily foul the electrodes.
These meters should also be periodically calibrated according to manufacturer’s instructions or labora-
tory standard operating procedures.

30.4.2.3  Dirt Level


In use, metalworking fluid mixes will pick up dirt, which includes metal chips (of various sizes depend-
ing on the particular operation), wear particles from the tool, die, or grinding wheel, and other debris
from the shop environment. It is the function of the filtration system, whatever type it might be, on a
particular machine or group of machines, to remove this material. Recirculating dirt can clog fluid
supply lines, create finish problems on the part, and leave machines in a very dirty condition. In many
stamping applications, this is not a problem as the fluid is not recirculated.
It is important to check the dirt level to insure that the filtration system is properly working and that
the fluid is staying clean. Dirt can be measured by centrifuging a sample of the fluid in a calibrated cen-
trifuge cone. It can also be measured by filtering a sample through a specific pore size filter paper, dry-
ing, and/or sometimes ashing, and then weighing. Particle counters can also be used to determine size
and distribution of dirt. Dirt volumes greater than 500 ppm, or particles larger than 20 μm, can lead to
problems. If problems occur, consultation with the filter or filter media supplier can be beneficial. Fluid
additives are available to improve chip settling.

30.4.2.4  Oil Levels


Many metalworking fluids contain oil as part of their formulation. Just about any fluid in a recirculating
system will pick up some oil, known as “tramp oil,” which is leakage from hydraulic lines or gear boxes,
way lube systems, or residual material on incoming parts.
Tramp oil can be either emulsified into the product by chemical or mechanical means, or basically left
floating on the fluid mix as what is called “free” oil. Many synthetic fluids are designed to reject leaked oils,
so any that get into the fluids will be left separated or floating. Products that contain some oil in the for-
mulation have the ability to emulsify some of this leaked oil as well as the product oil, and some tramp oils
may even contain emulsifiers that allow them to emulsify into the water-based metalworking fluid mix.
Free oil can usually be removed through the use of skimmers (belts or ropes) that attract the fluid
and pull it out. Centrifuges and coalescers can also be used, especially to remove tramp oil that may be
loosely emulsified into the product.
It is of value to know the tramp oil level in a fluid mix, since performance factors like cleanliness,
mildness, corrosion control, and rancidity can all be adversely affected by too high a level. For some
products, 2% tramp oil can be a problem; for others, if the level gets to 5% or higher, steps should be
taken to reduce it. It has been shown that high levels of tramp oil can lead to increased misting from the
metalworking fluid.18 In addition to performance issues caused by a high level of tramp oil, it also causes
interference with refractometer readings.
Free oil is measured by letting a sample of the fluid stand in a graduated cylinder for 12–24 h and
observing how much separates. It can also be done by means of a laboratory centrifuge. The total oil
(that contributed by the product and the amount of tramp oil present) can be measured by completely
“splitting” a mix sample by means of sulfuric acid and then observing the total oil content. The contribu-
tion of product oil should be known from one of the previously described concentration check methods.
Subtract this number from the total oil and the amount of tramp oil is what is left.

30.4.2.5  Bacteria and Mold19


For water-based metalworking fluids, the water environment can be conducive to microbial growth.
Straight oils should not have microbial problems unless contaminated with water. Bacterial growth is
30-12 Lubricants

usually accompanied by the development of objectionable, rancid odors. Fungal growth is manifest by
slimy deposits in, and on, the fluid system as well as musty odors. When formulating these water-based
fluids, ingredients are typically included that will control this growth. Bactericides and fungicides are
added to kill any organisms that may get into the fluid.20 Other fluids, marketed as being “bioresistant,”
are formulated with materials that do not support biological growth. Other fluids are designed so that,
while microbial growth may occur, no offensive odors or performance problems develop. Many addi-
tives exist that can be added tankside to control microbial growth.
Bacteria and mold levels can be determined in several ways:
Petri dishes coated with specific agars are prepared and after diluting the used fluids, a small portion
is introduced onto the plates. After 48 h or more hours of incubation at 36°C, the colonies on the plates
are counted.
Special paddles inserted in plastic tubes have been developed that are coated with certain agars. These
paddles are dipped into the used fluids, returned to the tube, and once again incubated. After 48 h, they
are observed and compared to a chart that is calibrated to show the number of organisms.
Most suppliers have set 105 colony-forming units (CFU) as the maximum bacteria level and 0 CFU as
the maximum mold level. When these numbers are exceeded, some form of treatment may be needed.
Dissolved oxygen (DO) can also be used as an indication of biological activity. A circulating met-
alworking fluid mix will absorb a certain level of oxygen. As aerobic bacteria grow, they will consume
some of this oxygen and excrete gases that will drive out more. With this as the basis, a determination
can be made, using a DO probe to measure the oxygen concentration in the fluid mix. Typically, a value
of 3 ppm or less is indicative of a problem. Another DO method is to take an initial reading, followed
by one 2 h later. A difference of 2–3 ppm usually indicates a problem. DO is a quick method for problem
detection but does not give any indication of specific counts or microbial types. Other methods, such as
the use of innovative adenosine triphosphate assays, are under development.21

30.4.2.6  Conductivity
Another parameter that will give an indication of a water-based metalworking fluid’s condition is con-
ductivity. Conductivity is measured on a conductivity meter in units of milliSiemens per centimeter
(mS/cm). A typical metalworking fluid mixed with tap water will have a conductivity reading around
1.5 mS/cm. As the amount of dissolved material, water minerals, concentrate additions, and contami-
nates increase, the conductivity increases. The rate is largely dependent on system turnover and the
quality of water used.
When conductivity reaches 5 mS/cm and higher, possible problems include corrosion, residues, and
mix instability for emulsion products. Conductivity will increase as the fluid mix ages, but can usually
be kept under control with good quality water.

30.5  Metalworking Fluid Application


Whatever type of fluid is being used, it is essential that it be applied correctly and adequately onto the
part and into the work zone. Improved tool life, better surface finish, lower power consumption, and
in some cases, efficient chip flushing can be achieved when fluid is properly applied into the work zone.

30.5.1  Flood
In order to cool, lubricate, and in some cases move the chips, it is beneficial to apply a large amount of
metalworking fluid directly into the work zone. Nozzles are positioned so that the fluid will cover the
part or fill holes being machined, and penetrate the tool/wheel/die interface with the part. For metal
removal fluids, a good rule to follow on flow rate is
• General purpose machining and grinding, m3/s = machining kW/120
• High production machining and grinding, m3/s = machining kW/60–30
Metalworking Lubricants 30-13

30.5.2  High Pressure


In many applications, in order to better penetrate the work zone, the fluid is applied under high pressure
(50–300 bar). Typically, an additional fluid reservoir with a high pressure pump is positioned on the
machine tool so that these high pressures can be achieved.

30.5.3  Mist/MQL
Another option for applying fluid that can be used in stamping as well as in machining operations is
the use of a fluid mist or a minimum quantity lubricant (MQL), again applied through a special unit
attached to the machine or press. Fluid is applied as a mist at the rate of a few milliliters/minute. The
advantage of this type of system is that there is no large tank of fluid to maintain and usage rates are
greatly reduced. Disadvantages include the generation of mist that must be controlled (typically, by
enclosing the machine), as well as the difficulty in moving the chips in machining operations. Many
plants have found this system to be very advantageous in their operations. Current evaluations also
include the use of super critical carbon dioxide (SCCO2) and liquid nitrogen along with MQL.

30.5.4  Manual
Applying fluids manually is usually reserved for stamping and forming operations where brushes or
rollers are used to apply fluid onto coils of metal or parts before they enter a press. It is also found in
some tapping operations where a specialized tapping compound may be manually applied to the taps to
give enhanced performance.

30.6  Metalworking Fluid Failure Mechanisms2


When problems occur with a metalworking fluid, they can be related to a number of causes, many of
them, not directly associated with the fluid, such as improper machine setup, bad tooling, or dies, or
changing metal specifications. It is important to gather all the details and facts concerning the entire
operation before attempting to solve any problem.
Some of the problems commonly associated with metalworking fluid failure include corrosion, foam,
rancidity, stability, residue, and lubricity.

30.6.1  Corrosion
Corrosion, rust, or staining typically will occur on the finished part, but can also be found on the
machine tool. Some fluids are not acceptable for use on certain metals and that can result in staining. An
immersion test with the metal and fluid, prior to use, is the best way to prevent this. Lean concentrations
or the buildup of corrosive water minerals, especially chloride ions, can also contribute. It is advisable to
check for the length of protection expected from the fluid. Metal removal fluids are generally designed
for shorter-term corrosion protection, typically 3–4 days, and anything longer should be supplemented
by a rust inhibitor. Stamping fluids are usually designed for much longer-term protection. Testing is the
method to determine this.

30.6.2  Foam
Foam in a metal removal fluid can lead to many problems including safety issues if it overflows onto the
floors, tooling issues if instead of getting coolant into the cut zone air is applied, and excessive usage
issues if tanks continually overflow. Foam can be the result of too high a concentration of the fluid. It
can result from soft (<50 ppm) water being used to make the mix or from contamination such as cleaners
30-14 Lubricants

or washing fluids. High pressure pumps can cause some fluids to foam. Inadequate retention time in a
tank to allow entrained air in the fluid to escape can also lead to foam. If foaming is not controlled by
the formulation itself, the use of antifoams based on silicone and siloxane chemistries is helpful. In some
cases, materials such as calcium acetate or calcium nitrate are used to artificially harden the water used
for makeup.

30.6.3  Rancidity
Rancidity, typically recognized by objectionable odors in the metalworking fluid, is the result of micro-
bial growth.22 Almost every fluid is formulated with biocides and/or fungicides to kill these organisms
or with other materials capable of preventing their growth, rendering the fluid bioresistant. Much work
has been done in identifying the types of organisms typically found in metalworking fluids and any
related health effects that they might have. Low fluid concentrations and various contaminants can lead
to microbial issues. The microbes found in fluids are those typically found in the surrounding plant
environment so their presence is not unexpected. Many means of control exist including the use of EPA-
registered biocides as tankside additives.

30.6.4  Stability
Mixes of water-based oil-containing metalworking fluids can become unstable for several reasons. This
instability is typically manifest in a mix that is split, that is, the oil and water have separated. Freezing of
the concentrate during storage or shipment in cold weather months can be a cause. The most prevalent
reason is the buildup of hard water (calcium and magnesium) minerals to the level where the emulsi-
fier package in the fluid is tied up or stripped out, reducing the ability of the oil to remain emulsified.
Excessive tramp oil or contamination, especially from cationic containing materials, can also lead to
emulsion instability. Using low mineral content water and keeping the fluid free of contaminants are the
best ways to prevent stability issues.

30.6.5  Residue
A metalworking fluid is expected to leave behind some residual material on the part surface to mini-
mize corrosion. However, it is not expected to leave an objectionable residue. Residues are typically
found with water-based fluids, although dirty contaminated oils can also cause this condition. When
water evaporates from a fluid, what is left behind should be relatively easy to remove by rinsing with
the coolant. If it dries and becomes hard and tacky, then a problem exists. This can happen on parts as
well as machine tool components, especially on hot motor surfaces where evaporation is accelerated. In
situations where fluid concentrations that are too strong, buildups of hard water soaps exist, or there is
excessive contamination from tramp oil, chips or other materials can lead to residue issues. Frequent
machine cleaning will prevent buildups on machine tool surfaces.

30.6.6  Lubricity
Lubricity problems can be manifest in numerous ways including poor tool, punch or grinding wheel life,
unsatisfactory finish on the part, hot chips and parts, or excessive machining time due to slower speeds.
It is important to make sure that the fluid with the needed amount of lubricity for the particular job
is being used and that it is being run at the recommended concentration and correctly and adequately
applied into the work zone.
Contamination of the fluid with excessive amounts of dirt or tramp oil can also lead to lubrica-
tion issues especially in regard to finish. If there is a filtration system, it must be effective in removing
particulates.
Metalworking Lubricants 30-15

30.7  Health, Safety, and Regulatory Aspects23


The National Institute for Occupational Safety and Health (NIOSH) estimates that 1.2 million work-
ers are exposed to metalworking fluids in U.S. industries.24 Occupational exposures to metalworking
fluid products occur mainly through the inhalational and dermal routes, during routine machining
operations. There have been extensive research activities dealing with various health and safety aspects
of metalworking fluid usage. A common objective of many of these studies is to provide healthier
workplace environments for machinists by providing a better understanding of metalworking fluid
exposures and to develop and improve methods to control them. Significantly, metalworking fluid
stakeholders (fluid manufacturers, machining industries, health and safety specialists, worker rep-
resentatives, federal regulatory agencies, academic researchers, etc.) have organized well-attended
comprehensive health and safety symposia over the years to discuss common concerns and to share
state-of-the-art knowledge.25–27
There are no regulatory occupational exposure limits for water-soluble metalworking fluids. Non-
regulatory federal agency (NIOSH) has set a metalworking fluid recommended exposure limit (REL) of
0.4 mg/m3 of air (thoracic particulate mass) time-weighted average during a 40 h work week (corresponds
to 0.5 mg/m3 for total particulate mass).24 The U.S. Occupational Safety and Health Administration
(OSHA) has had a long-standing permissible exposure limit (PEL) for mineral oil mist of 5 mg/m3,28 and
the American Conference of Governmental Industrial Hygienists (ACGIH) (an association of industrial
hygienists and related professionals) has promulgated a threshold limit value (TLV) for mineral oil,
excluding metalworking fluids of 5 mg/m3 (“mineral oil” TLV includes paraffin oil, liquid petrolatum,
white mineral oil, and USP mineral oil).29 Due to the complexity of metalworking fluid mist constitu-
ents and the difficulty in evaluating exposure patterns, an integrated approach to protect workers from
inimical exposures has been proposed.30
Metalworking fluid mists are generated by a variety of machining mechanisms, especially during
grinding operations. The constituents of metalworking fluid mists are highly variable, due, in part, to
the following factors:
• Wide variety of metalworking fluid formulations available on the commercial market
• Polydispersion of environmental fugitive dusts and dirt and other contaminants in metalworking
fluid aerosols
• Migration of tramp oils and other exogenous lubricants into metalworking fluid sumps and cen-
tral systems
• Proliferation of microorganisms in water-soluble metalworking fluids
Susceptible workers who frequently inhale metalworking fluid mists may experience symptoms of
bronchitis or other respiratory disorders of an acute or chronic nature. However, the specific role(s) of
metalworking fluid aerosols in initializing breathing disorders is not clear, due to confounding factors
in workplace atmospheres. A goal of prudent industrial hygiene practices is to limit industrial emis-
sions in workplace atmospheres. The most effective method of reducing emissions is to use engineer-
ing controls, that is, mechanized area ventilation to efficiently dilute and remove mists in work areas.
A derivative methodology of the latter is to utilize portable local exhaust units that are positioned at
points of high concentrations of mist generation at individual work stations. In some cases, work-
ers may benefit from the use of respirators or administrative measures such as reassignment to non-
machining areas.
Symptoms of dermatitis are the most prevalent complaints among workers who come into direct
contact with metalworking fluids and other industrial lubricants. During machining operations,
the unprotected hands and forearms of workers regularly come into contact with parts, tools, and
machine surfaces coated with fluids. There are no regulatory occupational exposure limits for dermal
exposures, since (1) potentially problematic exposures are difficult to quantify, (2) individual sus-
ceptibilities to skin disease(s) vary considerably, and (3) idiopathic dermal responses to exposures.
30-16 Lubricants

An approach to minimizing dermal problems among machinists is to implement fluid management


programs:
• Monitor fluids (check physicochemical parameters regularly)
• Clean active fluids (filter dirt/debris, remove tramp oils, etc.)
• Condition recirculating fluids (replenish depleted key constituents including biocides to control
microbiological growth)
• Dump depleted fluids, clean sumps/central systems, recharge with fresh fluid
• Conduct regular employee training on the importance of prudent skin care
The diligent removal of metal swarf and fines from active fluids, especially, will help to minimize skin
abrasions that might become infectious. Personal hygiene regimens such as frequent hand washing, the
use of appropriate skin barrier creams, and removal of soiled clothing after work shifts can also be effec-
tive in preventing the onset of dermatitis. Personal protective equipment (PPE) such as non-permeable
gloves, arm shields, and aprons are effective in protecting vulnerable skin surfaces from excessive con-
tact with metalworking fluids.31
Metalworking fluids have been and continue to be an important part of the increasingly expanding
global industrial landscape. Historically, metalworking fluid formulations have evolved in parallel with
the demanding specifications of machining industries, for example, longer tool life, consistent parts fin-
ish, tighter tool, and parts tolerances. The advent of new technologies such as ultrahigh speed machin-
ing, MQL, and exotic alloys places extraordinary demands on metalworking fluid performance. In
modern-day terminology, “metalworking fluids” include products not traditionally considered “fluids,”
for example, vegetable oil–based (“green”) products, and modified fluids that are delivered in precise
quantities to the cutting zone by supercritical gases (SCCO2). The emerging science of nanotechnology
promises to radically change the way that metal parts are designed and manufactured. Consequently,
nanomaterial-based products will likely play an important part in research and development of radi-
cally new metalworking fluid formulations that exhibit unique lubrication and tribological properties.32
In this regard, nanotechnology will pose new challenges and considerations for metalworking fluid
health and safety practitioners and regulators.

References
1. ASTM Standard E-2523-06, Standard Terminology for Metalworking Fluids and Operations, ASTM,
West Conshohocken, PA, 2006.
2. Byers, J., Ed., Metalworking Fluids, Second Edition, Taylor & Francis, Boca Raton, FL, 2006.
3. Zimmerman, J. B., Takahashi, S., Hayes, K. F., and Skerlos, S. J., Experimental and statistical design
considerations for economical evaluation of metalworking fluids using the tapping torque test,
Journal of the Society of Tribologists and Lubrication Engineers, 59, March 2003, 17–24.
4. Lange, K., Ed., Handbook of Metalforming, McGraw-Hill Book Company, New York, 1985, Chapter 6.
5. Altan, T., Oh, S., and Gegel, H., Metal Forming Fundamentals and Applications, American Society for
Metals, Metals Park, OH, 1983, Chapter 6.
6. Canter, N., Challenges in formulating metal-forming fluids, Tribology and Lubrication Technology,
65, March 2009, 56–63.
7. Booser, E. R., Ed., Tribology Data Handbook, CRC Press, Boca Raton, FL, 1997, Chapter 74.
8. ASTM Standard D-2881-03, Standard Classification for Metalworking Fluids and Related Materials,
ASTM, West Conshohocken, PA, 2003.
9. Shen, B., Shih, A., and Tung, S., Application of nanofluids in minimum quantity lubrication grinding,
Tribology and Lubrication Technology, 65, March 2009, 73–80.
10. Byers, J., Selecting the perfect metalworking fluid, Tribology and Lubrication Technology, 65, March
2009, 29–36.
Metalworking Lubricants 30-17

11. ASTM Standard E1497-05, Standard Practice for Selection and Safe Use of Water-Miscible and
Straight oil Metal Removal Fluids, ASTM, West Conshohocken, PA, 2005.
12. Booser, E. R., Ed., Tribology Data Handbook, CRC Press, Boca Raton, FL, 1997, Chapters 71 and 72.
13. Rolfert, E., The influence of metalworking fluids on common elastomers, Journal of the Society of
Tribologists and Lubrication Engineers, 49, January 1993, 49–52.
14. Joseph, J., Coolant Filtration, Joseph Marketing, East Syracuse, New York, 1985.
15. Independent Lubricant Manufacturers Association, Waste Minimization and Wastewater Treatment
of Metalworking Fluids, ILMA, Alexandria, VA, 1990.
16. Canter, N., Development of Guidelines for Using and Maintaining Metalworking Fluids, A White Paper
Sponsored by the Society of Tribologists and Lubrication Engineers, STLE, Park Ridge, IL, May 2008.
17. Organization Resources Counselors, Inc., Metal Removal Fluids: A Guide to Their Management and
Control, ORC, August 1997.
18. Byers, J. and Turchin, H., Effect of oil contamination on metalworking fluid mist, Journal of the
Society of Tribologists and Lubrication Engineers, 56, July 2000, 21–25.
19. Booser, E. R., Ed., Tribology Data Handbook, CRC Press, Boca Raton, FL, 1997, Chapter 73.
20. Passman, F., Current trends in MWF microbicides, Tribology and Lubrication Technology, 66, May
2010, 30–38.
21. Passman, F., Egger, G., Hallahan, S., Skinner, B., and Deschepper, M., Real-time testing of biobur-
dens in metalworking fluids using adenosine triphosphate as a biomass indicator, Tribology and
Lubrication Technology, 66, May 2010, 40–45.
22. Passman, F., Metalworking fluid microbes: What we need to know to successfully understand cause-
and-effect relationships, Tribology and Lubrication Technology, 64, April 2008, 40–52.
23. Booser, E. R., Ed., Tribology Data Handbook, CRC Press, Boca Raton, FL, 1997, Chapter 87.
24. National Institute for Occupational Safety and Health (NIOSH), Criteria for a Recommended
Standard: Occupational Exposure to Metalworking Fluids, DHHS Publication 98–102 (1998),
Cincinnati, OH, 1998.
25. American Automobile Manufacturers Association, The industrial metalworking environment:
Assessment & control, Metalworking Fluid Symposium, Symposium Proceedings, Detroit, MI,
November 13–16, 1995.
26. American Automobile Manufacturers Association, The industrial metalworking environment:
Assessment & control of metalworking fluids, Metalworking Fluid Symposium II, Symposium
Proceedings, Detroit, MI, September 15–17, 1997.
27. Independent Lubricant Manufacturers Association, The industrial metalworking environment: The
3rd symposium on the assessment and control of metal removal fluids, Symposium Proceedings,
Dearborn, MI, October 5–8, 2008.
28. Occupational Safety and Health Administration (OSHA), Permissible exposure level for mineral oil
mist, 29 CFR 1910.1000, Table Z-1.
29. American Conference of Governmental Industrial Hygienists (ACGIH), Threshold limit value
(TLV) for mineral oil, excluding metal working fluids (unpublished).
30. Cohen, H. and White, E. M., Metalworking fluid mist occupational exposure limits: A discussion of
alternative methods, Journal of Occupational Environment Hygiene, 3, 2006, 501–507.
31. Independent Lubricant Manufacturers Association, Metalworking Fluids Dermal Assessment Guide
(DAG), 2008.
32. Canter, N., Friction-reducing characteristics of nano-boric acid, Tribology and Lubrication
Technology, 64(2), 2008, 10–11.
31
Hydraulic Fluids
31.1 Introduction..................................................................................... 31-1
31.2 Oil Quality Required by Hydraulic Systems............................... 31-1
Viscosity  •  Antiwear (Wear Protection)  •  Oxidation
Stability  •  Antifoam/Air Separation Characteristics  • ​
Demulsibility (Water-Separating Ability)  •  Rust
and Corrosion Protection  •  Hydrolytic Stability  • ​
Filterability  •  Compatibility  •  Performance
and Classification Standards
31.3 Special Application Considerations for Hydraulic Fluids......... 31-7
Mobile Equipment  •  Industrial (Stationary)  •  Aviation  •  Fire
Resistance  •  Biodegradable  •  Ecotoxicity (Aquatic Toxicity)  • ​
Energy Efficient  •  Food Machinery (NSF-H1) Hydraulics
31.4 Hydraulic System and Fluid Maintenance.................................31-11
Contamination Control  •  Filtration  •  Controlling
Temperatures  •  Reservoir Oil Levels  •  Oil Analysis  • ​
James B. Hannon Routine Inspections
ExxonMobil Corporation References................................................................................................... 31-15

31.1  Introduction
Hydraulic fluids can be steam, gas, water, oil, or many other fluid materials. These hydraulic fluids are
used in countless industrial and commercial applications that might range from a large, high-pressure
hydraulic press of a forge to sophisticated aerospace, tight tolerance robotics. From large to small, the
same hydraulic principals apply: the conversion of mechanical energy to fluid energy and then back to
mechanical energy.
Hydraulic systems can be designed to
• Transfer force
• Control speed
• Control motion
A hydraulic fluid will support these design purposes as well as provide rust and wear protection over
a wide range of temperatures and pressures. Specialized hydraulic fluids may support fire resistance,
environmental, or energy-efficiency concerns.

31.2  Oil Quality Required by Hydraulic Systems


Most often, satisfying the requirements of only the pump seems to be the primary consideration
for hydraulic fluid selection. Although the costs of hydraulic pump failures are generally one of the
more costly occurrences within hydraulic systems, erratic operation of valves and actuators due to

31-1
31-2 Lubricants

inadequate oil performance characteristics such as oil degradation (oxidation) causing deposits to
form in critical clearance areas can lead to costly production losses. With the close clearances, dif-
ferent metallurgies, various elastomers, and high pressures and temperatures, the service life and
performance of all of the system components depend on proper selection and maintenance of the
hydraulic fluids.
Hydraulic fluids perform many functions in addition to transmitting pressure and energy. These
include minimizing friction and wear, sealing close clearance parts from leakage, removing heat, mini-
mizing system deposits, flushing away wear particles and contamination, and protecting surfaces from
rust and corrosion. The important characteristics of a hydraulic fluid will vary by the components used
and the severity of service. The physical characteristics and performance qualities of hydraulic fluids
that are commonly required by most hydraulic systems are

• Viscosity
• Viscosity index (VI)
• Wear protection capability
• Oxidation stability
• Antifoam and air separation characteristics
• Demulsibility (water-separating characteristics)
• Rust protection
• Compatibility with equipment components and other fluids

Some specific applications may require


• Soluble oils
• High water content fluids
• Fire-resistant fluids
• Environmental performance
• Food machinery
• Wide temperature capability

31.2.1  Viscosity
The single most important physical characteristic of the hydraulic fluid is its viscosity. Viscosity is a
measure of the oil’s resistance to flow so in hydraulic systems that are dependent on flow, viscosity
is important from both the lubrication and energy transmission aspects. Although viscosity require-
ments will be somewhat dictated by the components (pumps, valves, motors, etc.), certain effects of
improper viscosity selection need to be recognized. Too low a viscosity can lead to excessive metal-to-
metal contact of moving parts, wear, and leakage. Too high a viscosity can result in excessive heating,
sluggish operation (particularly at start-up), higher energy consumption, lower mechanical efficiencies,
and increased pressure drops in transmission lines and across filters. Since viscosity decreases as tem-
perature increase, viscosity requirements are generally specified at operating temperature. If tempera-
tures are higher than those specified for normal operation, a higher viscosity oil may be required to
provide long service life of the components. If start-up and operating temperatures are lower than those
specified, then a lower viscosity oil may prove to be better for overall system performance. Systems
operating over a wide range of temperatures may require oil that exhibits high VI. Hydraulic systems
normally use oil with a viscosity range of 32–68 cSt at 40°C. Most hydraulic equipment builders require
that the viscosity at start-up temperatures not exceed 1515 cSt to ensure flow to the pump. Some build-
ers, however, limit the start-up viscosity to 860 cSt. Table 31.1 offers viscosity parameter guidance for
pumps and motors. Equipment specific guidance is recommended as these values are generic to pump
or motor type.
Hydraulic Fluids 31-3

TABLE 31.1  Pump and Motor Viscosity Guidance


Radial Axial Piston
Viscosity Requirements Piston Pump Pump Vane Pump Gear Pump Motor
Minimum viscosity (cSt) 21 14 10 12 12
Maximum viscosity with 160 100 400 160 160
load (cSt)
Maximum viscosity no 2000 2000 1000 1000 2100
load (cSt)

31.2.1.1  Viscosity Index


All oil viscosity varies substantially with changes in temperature. In some hydraulic systems, subjected
to wide variations in start-up and operating temperatures, it is desirable to use an oil that changes
relatively little in viscosity for a given temperature range. An oil that does this is said to have a high
VI. High VI can be achieved by using mineral oil base stocks that have been refined through a severe
hydroprocessing technique separately or in conjunction with the use of long-chain polymers called VI
improvers (VII). Mineral oil base stocks refined through conventional methods can also achieve wide
temperature range performance by the addition of VII. Synthetic hydraulic fluids with naturally high
VIs are an alternative for severe application temperatures—both high and low.
Hydraulic fluids are subjected to high shear rate conditions particularly as pressures and speeds rise.
VII are generally long-chain polymers, and depending on the type of polymers used, can shear over
time. As the VII shears, a loss of viscosity can result in poorer hydraulic system performance due to
increased potential for metal-to-metal contact, and increased leakage. Expected shear stability of a high
VI hydraulic fluid can be assessed in tapered roller bearing testing (Coordinating European Council
[CEC] L-45-A-99), often referred to as KRL, as percent viscosity loss. When selecting a high VI, mineral
oil–based, hydraulic fluid good shear stability performance should be required.

31.2.2  Antiwear (Wear Protection)


To assure satisfactory hydraulic component life, the hydraulic fluid must minimize wear. Wear results
in loss of mechanical efficiency as well as higher costs due to shorter component life. In some hydrau-
lic systems, such as low-pressure, low-temperature systems with gear pumps, antiwear additives are
not necessary. Other high-pressure, high-temperature systems using vane pumps will require antiwear
additives in the hydraulic fluid. Piston pumps may or may not require antiwear additives depending
on the metallurgy and design. A properly refined petroleum oil has naturally good wear protection
(without the use of antiwear additives). This quality, sometimes referred to as lubricity or film strength,
is present to a greater degree in oils of higher viscosity than those of lower viscosity. However, certain
additive materials are used to improve antiwear performance in hydraulic oils. These additives work by
chemically reacting with the metal surfaces forming a strong film preventing metal-to-metal contact
under boundary lubrication conditions. The use of effective antiwear additives may allow the use of
lower viscosity fluids without significant wear.
Antiwear fluids are generally required in gear and vane pumps operating at pressures above 1000 psi
(69 bar) and over 1200 rpm. Piston pumps may or may not require antiwear additives depending on
the specific manufacturer and the metallurgy used. For example, one manufacture prefers rust and
oxidation (R&O)-inhibited oils for their piston pumps, which typically use bronze piston shoes against
a steel swash plate. Another manufacturer requires the use of antiwear additives for their steel on steel
piston pumps. Steel against steel at high pressure will always require antiwear formulations. Premium
quality antiwear hydraulic fluids are formulated to provide performance in all pumps and hydraulic
systems. Due to the many different pumps and other components used in a given industrial plant, it
31-4 Lubricants

might be advisable to consolidate the number of hydraulic fluids by using antiwear formulations that
meet all of the requirements.

31.2.3  Oxidation Stability


Oil circulating in hydraulic systems is heated by the churning and shearing action in pumps, valves,
tubing, and actuators. Also, the energy released as the oil goes from high pressure to low pressure in a
relief valve is converted to heat, which raises its temperature. Oil temperature can be further increased
by convection or conduction heating while performing its work in applications such as the hot molds in
plastic injecting molding operations and continuous casters in steel mills.
The oil is in contact with warm air in the reservoir or sump. Air is also dissolved or entrained in the
oil. Because of this contact with air, oxygen is intimately mixed with the oil. Under these conditions
(exposure to temperature and oxygen), the oil tends to chemically combine with the oxygen creating
oxidation products. The tendency to oxidize is greatly increased as temperatures increase, as agitation or
splashing becomes excessive, and by exposure to certain materials that catalyze the oxidation reactions
such as iron, copper, nickel, and other metals.
Due to oxidation, both soluble and insoluble oxidation products are formed and the oil gradually
increases in viscosity. Some variation of viscosity, within 10% of the initial viscosity, that has proved sat-
isfactory in service is not necessarily harmful. However, viscosity higher than necessary is accompanied
by higher fluid friction and more heating. With much of today’s critical systems using electrohydraulic
servo-valves where clearances are extremely close, slight oxidation could result in deposits forming
on the servo-valve spools restricting their movement, resulting in production problems. Low quality
oils have poor resistance to oxidation under severe conditions. With such oils, troubles of the kind
just described often occur. In high-quality hydraulic oils, the natural ability of well-refined, carefully
selected base oils to resist oxidation is greatly improved by the use of additives that retard the oxidation
process.

31.2.4  Antifoam/Air Separation Characteristics


The positive and accurate motion of actuators within a hydraulic system is dependent on the virtual
incompressibility of the hydraulic fluid. Under high pressure, mineral oils can see a very slight reduc-
tion in volume (4.0% at 10,000 psi [690 bar] and 140°F [60°C]) and a corresponding increase in density.
For purposes of the vast majority of hydraulic systems, this is considered insignificant. Introduction of
air into the fluid can substantially change the compressibility. Air causes spongy or erratic motion that
will result in poor system performance particularly where precision positioning and repeatability are
required. Antifoaming and air separation characteristics are two different concepts although somewhat
connected. Air separation means that the entrained air is released from the oil while antifoam means
that the air bubbles getting to the surface of the oil are readily dissipated. Both aspects are important to
the performance of a hydraulic oil. External contamination can alter both of these characteristics so that
it is not only important to select an oil that will provide good antifoam and air separation performance
but also necessary to minimize contamination in order to maintain this good performance.

31.2.5  Demulsibility (Water-Separating Ability)


Water contamination is sometimes a problem in hydraulic systems. It may be present as a result of
water leaks in heat exchangers or wash-down procedures but more commonly it accumulates because
of condensation of atmospheric moisture. Most condensation occurs above the oil level in reservoirs as
machines cool during idle or shutdown periods. A clean oil of suitable type will have little tendency to
mix with water and, in a still reservoir the water will tend to settle at a low point. During operation, the
water may be picked up by oil circulation, broken up into droplets, and mixed with the oil, forming an
Hydraulic Fluids 31-5

emulsion. The water and oil in such an emulsion should separate quickly in the reservoir, but when solid
contaminants or oil oxidation products are present, emulsions tend to persist and to join with other
deposit-forming materials present to form sludge. The emulsion may be drawn into the pump and made
more permanent by the churning action of the pump and the mixing effect of flow at high velocities
through control devices.
To prevent this from occurring, it is essential that a hydraulic oil is able to separate quickly and com-
pletely from water. Properly refined oils have this ability when new, but only those oils having excep-
tional demulsibility are able to retain good water-separating ability over long service periods. Poorer
demulsibility can generally be expected when detergent dispersant fluids are used. When using such an
oil, every effort should be made to keep systems free of water, dirt, and other contaminants.

31.2.6  Rust and Corrosion Protection


Water and oxygen can cause rusting of ferrous surfaces in hydraulic systems. As air is always present
(except in specialized nitrogen blanket systems), oxygen and water vapor are available. The possibility
of rusting is greatest during shutdown when surfaces that are normally covered with oil may then be
unprotected and as they cool may gather condensation. This is particularly important where operations
are subject to high humidity conditions and temperature changes within the reservoirs.
Rusting results in surface destruction, and rust particles may be carried into the system where they
will contribute to wear and the formation of sludge-like deposits that will interfere with the operation of
pumps, actuators, and control mechanisms. Rusting of piston rods or rams causes rapid seal or packing
wear resulting in increased leakage and system contamination, when external contaminants can enter
the hydraulic fluid through worn seals or packing.
High-quality hydraulic fluids contain rust inhibitors, which have an affinity for metal surfaces. The
rust inhibitor plates out on the ferrous–metal surfaces, forming a barrier film that resists displacement
by water and, therefore, protects the surfaces from contact with water. The rust inhibitor must be care-
fully selected to provide adequate protection without reducing other desirable properties, especially the
water-separating ability.

31.2.7  Hydrolytic Stability


In many operating environments, water or moisture may come in contact with the hydraulic oil. Small
amounts of free water not only cause corrosion of the tight tolerance metal surfaces of servo-valves but
also degrade the oil. Unstable fluids will form acidic and insoluble contaminants that can cause hydrau-
lic system malfunctions or changes in oil viscosity. Increased acidity and copper corrosion due to the
influence of water contamination can be measured utilizing ASTM D2619 “Standard Test Method for
hydrolytic stability of hydraulic fluids (beverage bottle test).”

31.2.8  Filterability
Filterability is the relative performance of a hydraulic fluid to pass through a filter element without plug-
ging it. Relative performance can be assessed through ISO 13357—Determination of the filterability of
lubricating oils. This test method provides for testing of hydraulic oils in both wet, water added, and dry
conditions. Recommended filter sizes can be found in Table 31.2.

31.2.9  Compatibility
An often overlooked characteristic of the hydraulic fluid during the selection process is its compat-
ibility. Use of oils that exhibit undesirable reactions with system components such as metallurgy, elas-
tomers, paints, or gasket materials can result in leakage or contamination within the system that can
31-6 Lubricants

TABLE 31.2  Typical Critical Clearances, ISO Cleanliness, and Filter


Ratings for Hydraulic Fluid System Components
ISO Filter Absolute
Component Microns Cleanliness Micron (200 Beta)
Antifriction bearings 0.5 16/14/12 3
Vane pump—tip to vane 0.5–1 17/15/13 6
Piston pump—valve plate to Cyl 0.5–5 17/15/13 6
Gear pump—gear to side plate 0.5–5 19/17/15 12
Servo-valve spool 1–4 16/14/11 3
Journal bearings 100–300 18/16/14 12
Actuators (cylinder) 50–250 20/18/15 12

reduce overall performance. When changing from one oil to another such as going from a conventional
mineral oil to a fire-resistant fluid or a synthetic hydraulic fluid, attention to compatibility issues is
particularly important. Changing from one class of fluid to another requires a very thorough review of
associated risks.
Fluid-to-fluid compatibility can be assessed in different ways including visual observations and
performance testing. Visual inspections look for abnormalities of the mixed solutions, such as floc,
sediment, haze, or phase separation. Performance testing will look for potential detrimental impact of
mixtures by comparing key hydraulic fluid performance criteria versus the new fluids. A key element
of this testing should be filterability, ISO 13357-1 (wet method) understanding that the mixture should
pass equipment manufacturer requirements.
Seal (elastomer) compatibility can be confirmed in DIN 53538 or ISO 1817 testing where changes in
volume, hardness, dimension, surface area, and tensile stress–strain are evaluated.

31.2.10  Performance and Classification Standards


Industry standards and equipment manufacturer approvals can be used to qualify and classify hydrau-
lic fluids. These performance standards can be equipment manufacturer–specific tests that may require
product test approval or a slate of performance tests with acceptance limits. In addition, certain indus-
try standards have been created to classify hydraulic fluids.
Hydraulic fluid are classified and assigned alpha numerical codes as naming conventions by ASTM
D 6080, DIN 51502, and ISO 6743 standards. For example, ISO 6743 part 4 is dedicated to hydrau-
lic fluids and a common naming convention is ISO-L-HM-32 used on a monograde hydraulic fluid,
where
• ISO is International Standards Organization
• L is for lubricants
• HM designates a refined mineral oil improved for antirust, antioxidation, and antiwear properties
• 32 is the ISO viscosity grade, per ISO 3448
International standards, ASTM D6158, ISO 11158, DIN 51524, JCMAS HK, AIST 127, and SAE MS 1004,
offer test slates for hydraulic fluids with minimum acceptable limits. ISO also offers specifications spe-
cific to environmentally acceptable fluids, ISO 15830, and fire-resistant fluids, ISO 12922. Performance
claims are assessed by the oil supplier. Testing requirements will vary per specification but typically
include viscosity, VI, pour point, flash point, filterability, demulsibility, rust prevention, copper corro-
sion, oxidation stability, elastomer compatibility, air release, foam tendency/stability, and acid number.
Equipment manufacturer approvals, typically from pump and mobile equipment suppliers, often
involve testing that may be proprietary to the manufacturer or that has been adapted as an international
standard. As an example, ASTM D 6973 measures vane pump wear and was derived from a pump
Hydraulic Fluids 31-7

manufacturer fluid test rig. Performance acceptance levels are typically offered by industry standard or
equipment manufacturer specification.

31.3  Special Application Considerations for Hydraulic Fluids


Hydraulic fluids can be broadly categorized in the following general applications: industrial,
mobile equipment, aviation, fire resistant, biodegradable, low aquatic toxicity, energy efficient, and
food-processing machinery.

31.3.1  Mobile Equipment


The ability to reliably operate mobile equipment is complex due to exposure to wide ambient ranges and
component physical size and weight limitations. These design considerations require careful selection of
a hydraulic fluid that can maintain suitable viscosity while releasing air, water, and solid contaminants
under high pressure operation.
Mobile system hydraulic reservoirs are designed to be as small as practical and still allow for proper
contaminant release and temperature control. To minimize space and weight requirements, hydrau-
lic reservoirs are often constructed integral to the equipment structure, as in an excavator’s boom.
Reservoirs must be designed with baffles specific to each application to minimize air entrainment and
maximize contaminant separation. Some mobile equipment utilizes pressurized reservoirs to minimize
external contamination. Hydraulic fluids with good air release and good foam control are highly recom-
mended in mobile equipment. Entrained air and foam can impact valve, hydraulic motor, and actuator
operation. Entrained air under compression, sometimes called micro-dieseling, can generate localized
temperature excursions above 1000°F (538°C) that lead to oil degradation and reduced pump life.
Maintaining optimum oil viscosity during operation is greatly influenced by ambient temperature
peaks, high and low, as well as the lubricant cooling method. Temperature swings from cold start-up to
hot midday operation often require the use of high VI fluids, either synthetic oil or VI-improved mineral
oil. High VI hydraulic fluids, discussed in the previous section, can support cold start hydraulic fluid
flow and provide the necessary film strength at elevated operating temperatures. Mobile equipment can
be exposed to winter start-up conditions at −30°F (−34.4°C) and summer midday peak temperatures of
approaching 120°F (49°C). The use of high VI hydraulic fluids, often called all temperature hydraulic
fluids, supports equipment operation through this temperature range.
Mobile hydraulic fluids operate at reservoir temperatures approaching 200°F (93.3°C) and at pres-
sures that can exceed 5000 psi (345 bar), higher than other applications. These elevated temperatures
and pressures demand higher thermal and oxidation stability, as well as wear protection.
In many mobile equipment applications, the use of a hydraulic system pump that delivers fluid power
to a travel gear, used in tractor drives, hydraulic motors is common. In these systems, the hydraulic fluid
must be selected based on the viscosity requirements of both pump and the motor.

31.3.2  Industrial (Stationary)


Industrial (or stationary) hydraulic applications typically are afforded larger reservoirs, connectors, and
fittings than seen in mobile equipment. The larger reservoirs allow for better air, water, solids contami-
nation release. They also typically allow for better cooling. Larger connectors and fittings reduce inter-
nal friction offering improved efficiency and reduced cooling needs.
Often heat exchangers are water cooled providing better lubricant cooling than air cooling as found
in mobile equipment.
Some industrial hydraulic presses generate pressures up to 15,000 psi (1034 bar), but typical applica-
tions are below 3000 psi (207 bar). Industrial applications often utilize lower cost gear and vane pumps
versus higher pressure and cost piston pumps.
31-8 Lubricants

Industrial installations often have the benefit of using hydraulic tubing in place of hydraulic hoses as
in many applications there is minimal movement between the pump and the actuator housing.
Industrial hydraulic equipment that is either housed in a temperature-controlled environment or
exposed to minimal ambient temperature differential would typically use a monograde hydraulic
fluid. Equipment exposed to wide ambient or operating temperature swing may benefit from a high VI
hydraulic fluid.

31.3.3  Aviation
Military and commercial aircraft use hydraulic power for actuation of flight control surfaces, landing
gear control, air and wheel braking, and emergency power generation. An aircraft may have two to four
independent hydraulic systems to provide redundancy for the operation of flight-safety critical systems,
the actual number of systems depending on aircraft size, and the extent to which electrical systems are
used to provide backup functions.
Aircraft hydraulic systems most commonly operate at 3000 psi (207 bar). Some older and smaller
aircraft have 2000 psi (138 bar) systems. Newer commercial aircraft (Airbus A380 and Boeing 787) are
using 5000 psi (345 bar) systems. The military V-22 Osprey tiltrotor aircraft uses a dual 3000 psi/5000 psi
(138 bar/345 bar) system.
Hydraulic power is usually generated by variable flow axial piston pumps. The primary pumps are
either engine driven (via gearbox mounted on the jet engine) or AC power driven. As backup, some air-
craft systems have power transfer units, which use one hydraulic system to generate power in another;
these devices are essentially an axial piston hydraulic motor driven by one system sharing a shaft with
an axial flow piston pump on the second system. Finally, many aircraft have ram air turbines that are
released from a cabinet under the wing or the body of the aircraft to drive an axial flow piston pump
as emergency hydraulic power generator in case all aircraft engines fail; critical electrical power gen-
eration, flight and landing system controls, and brakes can thus continue to be actuated by hydraulic
power.
Reservoirs are typically much smaller than the ones used in industrial systems, because of the need
to minimize the weight of equipment and fluid carried by the aircraft. Residence times are typically a
fraction of a minute (reservoir capacity is a fraction of circulating flow in gallons or liters per minute). In
some designs, the fluid in the reservoir is open to air and engine bleed air pressurizes the reservoir fluid.
Hydraulic sensors and actuators are generally similar in design concept to industrial components.
However, special technologies have evolved to satisfy the low weight/high reliability/easy maintainabil-
ity requirements of aviation, resulting in significantly higher costs for aircraft components.
A distinguishing feature of hydraulic fluids designed for aircraft use is the need for sufficiently low
viscosity to allow operability at very low temperature.
Military aircraft and very small business and commercial aircraft tend to use hydrocarbon-based
hydraulic fluids. The US Air Force or Navy qualifies the fluids in accordance with military specifications.
MIL-PRF-5606 applies to mineral oil–based fluids and is suitable for operation down to −65°F (−54°C).
These fluids use light naphthenic base oils and polymeric VII/thickeners. MIL-PRF-83282 (suitable for
−40°F [°C] operation) and MIL-PRF-87257 (suitable for −65°F [−54°C] operation) are polyalphaolefin
(PAO)-based and offer significant improvements over the flammability properties of MIL-PRF-5606
fluids. They do not use polymeric VII.
Commercial aircraft use phosphate ester–based fluids, which offer fire resistance beyond that of the
PAO-based fluids. These hydraulic fluids meet the SAE Aerospace Standard AS1241 standard, and have
to be qualified by the manufacturer of the aircraft. Major commercial aircraft manufacturers have their
own specifications. These fluids use low-viscosity phosphate ester base oil blends and polymeric VII. The
SAE standard and aircraft manufacturer specifications define the fire-resistance requirements for air-
craft hydraulic fluids. The requirements specified do not coincide with the requirements of fire-resistant
industrial fluids.
Hydraulic Fluids 31-9

Military and commercial hydraulic system components use different type of elastomeric seal materi-
als, and using of the wrong type of fluid can cause significant problems. Both the military and com-
mercial hydraulic fluid specifications require low particulate contamination. The phosphate esters base
oils of commercial aircraft hydraulic fluids are very hygroscopic (readily absorb water), and since they
are esters, can hydrolyze to form corrosive phosphoric acid derivatives. Commercial aircraft hydraulic
fluids contain additives that neutralize the formed acids and prevent the corrosion of hydraulic system
components.

31.3.4  Fire Resistance


The need for fire-resistant hydraulic fluids in specified applications is a very important safety consider-
ation. Many types of these fluids are available, dependent on service requirements. Elevated fire poten-
tial risks exist in steel mill casting operations and applications near ovens, as well as in steam turbine
hydraulic controls. In most cases, the fire-resistant hydraulic fluids require special considerations in
system and pump design, and system compatibilities.
ISO 12922 identifies the following fire-resistant hydraulic fluid categories with operating temperature
guidance:
• Type HFAE: Oil-in-water emulsions, typically with >80% water content. 41°F–122°F (5°C–50°C)
• Type HFAS: Aqueous solutions, typically with >80% water content. 41°F–122°F (5°C–50°C)
• Type HFB: Water-in-oil emulsions. 41°F–122°F (5°C–50°C)
• Type HFC: Glycol solutions, typically with >35% water, polyalkylene glycol solutions, or water
glycols. −4°F to 122°F (−20°C–50°C)
• Type HFDR: Phosphate esters. −4°F to 158°F (5°C–70°C) with short-term peak of 302°F (150°C)
• Type HFDU: Polyol esters, polyalkylene glycols, and hydrogenated PAOs. −4°F to 158°F
(5°C–70°C) with short-term peak of 302°F (150°C)
Debate in hydraulic oil industry continues regarding the definition of “fire resistant” as fluid types
exhibit more fire resistance than others. There is no single test that will rate a fluids relative fire resis-
tance. Factory Mutual does qualify industrial hydraulic fluids as fire resistant if the fluid meets the
requirements of Factory Mutual Standard 6930.
Fire-resistant hydraulic fluids are not fire proof but do significantly reduce the potential fire hazard
as compared to mineral oil–based fluids. Some industry experts are recommending a three-tiered clas-
sification program: (1) fire resistant—HFAE and HFAS fluids, (2) fire retardant—HFB, HFC, and HFDR
fluids, and (3) less flammable—HFDU fluids. Care needs to be taken in fluid selection to achieve the fire
protection required for the application.
Compatibility with system materials should be understood prior to final product selection.

31.3.5  Biodegradable
The demand for biodegradable lubricants grows as environmental regulations increase and more equip-
ment is operated in unattended, remote locations. Every effort should be taken to eliminate oil spills
of any size. Routine maintenance with attention to hydraulic system connectors is the key element to
avoiding an environmental release. In the event of a release, the use of biodegradable lubricants can
minimize environmental damages. These fluids can degrade in days, whereas standard mineral oils can
last in the environment for years.
Biodegradability is defined in ASTM D7044 as any substance containing <10% wt O2 content that
undergoes 60% biodegradation as theoretical CO2 in 28 days and 67% biodegradation as theoretical O2
uptake in 28 days, or any hydraulic fluid containing 10% wt O2 content that undergoes 60% biodegrada-
tion as theoretical CO2 or as theoretical O2 uptake in 28 days. ISO 15380 offers a similar limit at 60%
biodegradability per ISO Standards 14593 or 9439.
31-10 Lubricants

In general, biodegradability is the chemical breakdown of materials by living organisms, or their


enzymes, in the environment.
Levels of biodegradability are classified as “primary biodegradation” and “ultimate biodegradation.”
Primary biodegradation measures the reduction of carbon and hydrogen bonds in CEC-L-33-A-932
testing. Ultimate (ready) biodegradation measures the percentage of a substance that undergoes com-
plete degradation to CO2 and water. ASTM D5864 and EPA 560/6-82-003 tests are used to determine
ultimate biodegradation.
The term inherently biodegradable is used to describe the relative degradability of hydraulic fluids
as obtaining greater than 20% biodegradation. The Official Journal of European Union (2005/360/EC)
defines inherently biodegradable as greater than 20% biodegradation in 28 days, based on OECD 301
testing.
Vegetable-based oils and synthetic esters will have superior biodegradability performance over high-
molecular-weight PAO synthetic oils and mineral oil.

31.3.6  Ecotoxicity (Aquatic Toxicity)


Similar to the considerations with biodegradable hydraulic fluids, demand for low aquatic toxicity fluids
is increasing. Care of equipment and connectors is the first line of defense against spills and the use of
low aquatic toxicity fluids provides environmental advantages over standard mineral oils.
Aquatic toxicity is the adverse response, typically mortality, of water-based organisms to materials, in
this case hydraulic fluid. Appropriate tests are
1. Acute fish toxicity, 96 h, lethal concentration 50%—ISO 7346-2
2. Acute daphnia toxicity, 48 h, effective concentration 50%—ISO 6341
3. Bacteria inhibition, 3 h, effective concentration 50%—ISO 8192
Per ASTM D 6046, the lowest (best) ecotoxicity designation is for fluids that achieve a lethal or effec-
tive concentration of 50% at or above 1000 wppm in a water solution. Most often, this qualification is
referred to as “low aquatic toxicity.”
Testing of various finished hydraulic oils indicates that additives have a greater impact on aquatic
toxicity performance that lubricant base stock.

31.3.7  Energy Efficient


As energy costs rise and energy-related environmental penalties are implemented, the demand for energy-
efficient hydraulic fluids will grow. Advances in energy-efficient hydraulic fluids have documented effi-
ciency gains approaching 10% versus monograde, mineral-based hydraulic fluids. Advancements in
shear stable VII permit closer to optimum operating temperature viscosity. Equipment operating at
or close to optimum viscosity will be more efficient as both mechanical efficiency, associated with the
energy to move a fluid, and volumetric efficiency, associated with internal pump leakage, combine to
offer improved overall efficiency. The effect of temperature and viscosity on hydraulic efficiency is shown
in Figure 31.1.
Projected efficiency and productivity gains can be confirmed in field or rig demonstrations. Extreme
care in replicating all phases of operation should be taken as the energy consumption of each fluid is
documented. A key consideration that confirms lack of equipment drift between fluid demonstrations
is to bracket the test fluid with the reference fluid, called A–B–A test sequencing. ASTM D02.NO.12 is
progressing a standard practice that will aid in the accuracy and relevance of future energy-efficiency
demonstrations.
Another key element of hydraulic efficiency is heat generated by frictional losses. Elimination or
reductions in system pressure drops through valves, fittings, and supply connectors will improve system
efficiencies. These losses can be quantified in BTUs per hour as follows:
Hydraulic Fluids 31-11

Hot operation Cold start-up

Volumetric efficiency ηve

Efficiency
Mec
hani
cal e
fficie
ncy η
Optimum mec
h
operating Over
all ef
ficien
range cy η
ov

Viscosity
Poor volumetric efficiency High frictional losses
Good cold start-up properties Poor cold start-up properties
Poor film thinkness Good film thinkness

. Oil viscosity has a significant impact on hydraulic efficiency.


. Overall efficiency is a balance between mechanical and volumetric.
. Shear stable high VI fluids enable increased hydraulic efficiency.
FIGURE 31.1  Impact of viscosity and temperature on hydraulic efficiency. (Courtesy of Evonik.)

BTU / h = Pressure drop ( psi ) × gpm × 1.5


or

Watt = Pressure drop ( bar ) × liter/min × 1.34


31.3.8  Food Machinery (NSF-H1) Hydraulics


Concerns over food safety have elevated awareness of the potential benefit to lubricating food-processing
machinery with food machinery–specialized lubricants. Oils qualified for use in food-processing
machinery are classified as H1 by NSF, previously National Sanitation Foundation. These oils are regis-
tered with NSF based on acceptable formulation chemistry and toxicity testing. NSF-H1 oils are suitable
for use in food-processing machinery but care should be taken to ensure that food is not contaminated.
Accidental contamination limits of 10 ppm into the food have been established.
Some NSF-H1 hydraulic fluids may carry additional food claims, such as Halal, Kosher, and Allergen-
free for wheat, nuts, or gluten. Some NSF-H1 hydraulic fluids may also carry industrial performance
claims.

31.4  Hydraulic System and Fluid Maintenance


The degree of system maintenance will be based on specific performance expectations, the fluid used,
and the system-operating parameters. The various hydraulic fluids ranging from mineral oil–based, to
synthetic, to water-containing fire-resistant fluids will demand various levels of maintenance to assure
performance. Water-containing fluids will require higher levels of maintenance to assure not only that
the fire protection properties are retained but also that the fluid will provide proper lubrication char-
acteristics while in service. Selecting the proper fluid needs an understanding of the limitations of that
fluid.
31-12 Lubricants

Once the proper fluid has been selected, then the equipment and operating conditions will dictate the
degree of maintenance required to keep that fluid in service for long periods of time while retaining its
lubrication characteristics. These maintenance procedure objectives should include
• Keeping fluid clean/controlling contamination
• Maintaining proper temperatures
• Maintaining proper oil levels
• Periodic oil analysis
• Routine inspections

Noise levels Shock loads Filtration


Vibration Leakage Temperatures
Pressures Fluid odor and color Foaming

31.4.1  Contamination Control


This starts at the time you receive the oil and store it through the period of time it is in service.
Contamination, such as moisture, can enter some sealed containers while in storage through nor-
mal expansion and contraction of the fluid due to temperature changes. This allows moisture to
enter containers. Contamination can also result while transferring the oil from storage (or contain-
ers) to the system. This could be due to dirty transfer containers or equipment used previously for
other materials such as gear oils, engine oils, coolants, etc. Dirty reservoirs and debris around the
fill location are also sources of contamination while filling or adding makeup oil to the system. In
critical systems, sometimes quick-disconnect fittings are installed on reservoirs or portable filter
carts used to facilitate adding oil to clean oil to these systems. This minimizes the potential for
contamination.
The fluid in service must be clean and the level of cleanliness depends on the system. Numerically
controlled (NC) machine tools, for example, require high levels of cleanliness (ISO Cleanliness Rating
16/14/11) due to the close tolerance servo-valves, whereas the hydraulics used to operate hydraulic lifts
in automotive repair shops can operate satisfactorily with minimal filtration. It should be noted that
conventional filters will not remove water- or oil-soluble contaminants. Special coalescing or desiccant-
type filters are available to remove limited amounts of water.

31.4.2  Filtration
Full flow filtration is the most common type used on hydraulic systems to control the levels of solid
contaminants. The filters are generally installed in the supply (pressure) line but can also be installed in
return (low pressure) lines to the reservoir. Full flow filter housings are generally equipped with bypass
valves, which will open when the pressure drop across the filter exceeds a predetermined level. This
assures that components will receive oil in the event of filter plugging or restriction of oil flow through
the filter due to start-up or cold oil when viscosities are high. When filters go on bypass, unfiltered oil
will be supplied to components. Some filters are equipped with condition indicators or differential pres-
sure gages to warn of restrictions or plugging.
Selection of appropriate filtration levels (fluid cleanliness) will be based on specific system compo-
nents and operation. Table 31.2 shows some of the typical clearances and recommended ISO cleanliness
levels in hydraulic system components. The vast majority of hydraulic systems will function properly on
10 μm filtration with filter efficiencies of 98.7% or greater. Filter efficiencies are often referred to as beta
ratios (β). A beta ratio of 75 for a 10 μm filter would mean that 98.7% of the particles in 10 μm and larger
range will be removed.
Hydraulic Fluids 31-13

The filtration ratio is calculated by dividing the number of particles entering the filter by the number
of particles exiting the filter. β5 represents the filtration ratio at 5 μm or the ratio of upstream to the
downstream particles larger than 5 μm.
Beta values can be directly related to efficiency. To determine the relative performance of two filters
with different beta ratios, the downstream particle count from each filter can be calculated using the
beta ratio and an assumed upstream particle count. The filter with the highest beta value will have the
lowest downstream particle count.
Bypass filters, sometimes referred to as polishing filters, generally are installed in an independent
system where from 5% to 15% of the systems oil capacity (in gpm or lpm) is filtered to a finer degree.
The oil is taken from a low point in the reservoir using an auxiliary pump, filtered, and returned to the
reservoir. With this type of system, oil purification can be continued whether the hydraulic system is
in operation or shutdown. An alternate to the independent system is to use a continuous bypass mode
where a percentage of the oil flow from the pressure or return line is passed through suitable purification
equipment and returned to the reservoir. Bypass purification equipment can be relatively small in size
since only a portion of the total oil capacity is handled.
Portable filters, sometimes called filter carts or buggies, are also used to supplement permanently
installed system filters. These units consist of a motor-driven pump and filter arrangement that circulate
fluid from the reservoir, through a fine filter and back to the reservoir. The suction and return hoses
should be connected to opposite ends of the reservoir with quick-disconnect fittings. Generally, portable
filters will operate for at least 24 h on each system to ensure that the full oil charge is filtered effectively.
Portable filter units can be used in place of bypass filters if periodic or as-needed filtration is sufficient
to maintain the desired levels of fluid cleanliness. Portable filtration units can be a simple arrangement,
as just discussed, or may be purification units consisting of motor-driven pumps, oil heating elements,
vacuum chambers, and fine filters. The advantage of reclamation units is that they can remove water and
some volatile contaminants (such as some solvents) in addition to removal of particulates.
Batch filtration may be used where the fluid volume is very large or is heavily contaminated. The large
volume of oil may be removed and reclaimed as a batch through settling processes, filtration, centrifug-
ing, and/or by use of reclamation units. The disadvantage of this process is that the machine must be
shut down for removal of the fluid charge.

31.4.3  Controlling Temperatures


Excessive temperatures in the presence of dissolved or entrained air, in addition to possibly reducing oil
viscosities to a point where metal-to-metal contact results, will oxidize the oil. Oxidized oil will lead to
varnish and sludge formation in the system. These deposits plug or restrict the motion of valves, plug
suction screens, and cause shortened filter life.
Heat develops as the fluid is forced through the pumps, motors, tubing, and relief valves. Temperatures
should be maintained between 120°F and 149°F (49°C–65°C) in conventional hydraulic systems. Some
variable-volume pump systems, closed loop hydraulic systems, and hydraulic transmissions can operate
up to 248°F (120°C) where premium fluids must be used or drain intervals shortened. Systems operating
on water-based fluids should be kept below 140°F (60°C) to prevent the water from evaporating.
To allow heat to radiate from the system, keep the outside of the reservoir clean and the surrounding
area clear of obstructions. Make sure the oil cooler is functioning properly and keep air-cooled radiators
free of dirt and debris. Keep the reservoir filled to the proper level to allow enough fluid residence time
for the heat to dissipate.
Oil degradation is even more critical in NC machine tools with electrohydraulic servo-valves. Because
of space constraints, these systems typically are designed with small reservoirs resulting in short resi-
dence times for the hydraulic fluid. With minimal rest times and high system pressures, entrained air
bubbles can cause extremely high localized temperatures due to the adiabatic compression of the air
bubbles as they pass from low suction pressure to the high discharge pressure. This results in nitrogen
31-14 Lubricants

fixation that, when combined with oil oxidation, can form deposits, which will plug oil filters and cause
servo-valve sticking.

31.4.4  Reservoir Oil Levels


Systems are designed to provide a certain amount of oil residence time in the reservoir. This allows the
fluid time for separation of air and water and for solid contaminants to settle. It also provides for cool-
ing. Operating with low oil levels reduces the effectiveness of these processes. In addition, if levels are
low enough, air could be pulled into the pump suction resulting in air entrainment. This could lead to
excessive foaming and cavitation. Foaming and air entrainment can also result from air leaks in the
suction, low fluid temperatures, or fluid too viscous to release air bubbles. Oxidation and contamination
increase the fluids tendency to foam and retain air. Entrained air is a major cause of destructive pump
cavitation. The intense pressures and temperatures created by the collapse of air bubbles erode metal
parts of pumps and valves, resulting in excessive wear. Pump or valve cavitation may cause irregular
operation or “spongy” response of the system. In addition to air leaks, a restricted suction can also cause
cavitation. Causes of this restriction could be plugged strainers, oil viscosity too high, or inadequate
suction design conditions. Cavitation is apparent by a high-pitched whine or scream in the pump or it
may sound as if there are marbles trapped in the pump housing.

31.4.5  Oil Analysis


Laboratory oil analysis can be set up as a routine testing and the fluid should be visually inspected. A
visual inspection may reveal the type and degree of contamination. Take a sample of the fluid and allow
it to settle over night in a clear container. Inspect the sample for color, appearance, and odor. If there is
no evidence of water, corrosion or excessive accumulation of deposits, sediment or sludge, and the fluid
has the color and odor of new fluid, generally, laboratory analysis is not necessary. A slight “burnt odor”
is common in conventional hydraulic systems using petroleum oils. However, a burnt oil odor in an
oil sample may be a cause for concern. In some cases, even conventional laboratory oil analysis cannot
determine the low levels of contamination, such as the nitrogen fixation discussed earlier under tem-
peratures that will cause system operational problems. These systems will generally require high-quality
oils as well as a specialized oil analysis program.
If the visual inspections cannot identify specific contaminants and potential sources, take a sample
for laboratory analysis. It is a good practice to establish a program to periodically submit oil samples
for laboratory analysis. In critical high-temperature machines, this may be as often as every 3 months.
Noncritical machines will only require laboratory oil analysis every 6–12 months, but specific schedules
should be based on operation, equipment, and oil supplier’s recommendations.

31.4.6  Routine Inspections


Routine inspections of operating systems can provide insight to potential problem areas leading to cor-
rective actions. These corrective actions can result in longer equipment life and lower cost operation of the
hydraulic system. The items listed in the following are commonly included in routine inspection programs:
Noise levels: Increases in noise levels may signal problems with cavitation. Noise levels may also
increase from excessive temperatures allowing the oil’s viscosity to become too low resulting in metal-
to-metal contact.
Vibration: Loose mounting or misalignment of components will cause vibration resulting in acceler-
ated wear or failure.
Shock loads: System components, such as hoses and fittings, subjected to shock loads due to abrupt
changes in flows and pressures can lead to leakage and failures. Where shock loads are experienced,
accumulators should be installed in the system.
Hydraulic Fluids 31-15

Leakage: Oil leakage can result in low reservoir oil levels and resulting poor system performance. In
addition, leakage can be costly as well as create safety problems. Use of fluorescent dyes can improve
hydraulic fluid leak detection. Leakage at a rate of one drop per second can result in hydraulic fluid
losses of 405 gallons (1533 L) per year.
Temperatures: As discussed earlier, controlling temperatures in the appropriate range is important
from both an oil life and system performance standpoint. Excessive temperatures could be caused by
plugged or dirty heat exchangers, excessive pressures or pressure drops, high rates of internal leakage,
and low oil levels. Use of fluids with too high (excessive shearing) or too low (inadequate films) a viscos-
ity will also result in higher temperatures.
Filtration: The condition of fill-screens, breathers, and filters (indicators or differential pressure gages)
should indicate the need for cleaning or replacement.
Foaming: A little foam on top of the oil in the reservoir is normal. Excessive foam may indicate air
leaks in the suction line, unsatisfactory contamination levels, or inadequate antifoam characteristics of
the oil.
Fluid odor and color: Although odor and color are not characteristics used to judge the oil’s ability to
provide proper lubrication, changes in these physical properties may indicate contamination (solvents,
wrong oils added, degradation, etc.) or the oil reaching the useful end of its service life. If in doubt as
to the causes of the color and odor changes, a sample should be submitted for laboratory oil analysis.
If these changes are accompanied by undesirable machine-operating characteristics, change the oil but
still submit a sample for laboratory oil analysis to better understand potential changes in oil chemistry.

References
1. Pirro, D.M. and Wessol, A.A., ExxonMobil L&S Lubrication Fundamentals, Second Edition, Marcel
Dekker, Inc., New York, 2001.
2. Eaton Corporation Training, Eaton Mobil Hydraulics Manual, Eden Prairie, MN, 2006.
3. Bloch, H.P., Practical Lubrication for Industrial Facilities, The Fairmont Press, Inc., Lilburn, GA, 2000.
4. Heald, C.C., Cameron Hydraulic Data, Ingersoll-Rand, Woodcliff Lake, NJ, 1988.
5. Leuger, L.O., The Practical Handbook of Lubrication, Maintenance Technology International Inc.,
Cochrane, Alberta, Canada, 2000.
6. Vickers, Hydraulic Hints & Troubleshooting Guide, Troy, MI, 1996.
7. Johnson, M. and Miller, M., Eco-Friendly Fluids for the Lubricants Industry, Tribology & Lubrication
Technology, Park Ridge, IL, October 2010.
8. Van Rensselar, J., The Quest for Fire Resistance, Tribology & Lubrication Technology, Park Ridge, IL,
September 2010.
9. Gere, R. and Hazelton, T., Rules for Choosing a Fire-Resistant Hydraulic Fluid, Penton Media, Inc. &
Hydraulics & Pneumatics Magazine, New York, 2010.
10. Honary, L.A., Biodegradable/Biobased Lubricants and Greases, Machinery Lubrication, Noria
Corporation, Tulsa, OK, 2001.
32
Fluid Maintenance
32.1 Lubricant Selection Process........................................................... 32-1
32.2 Procurement Process....................................................................... 32-2
32.3 Quality Assurance........................................................................... 32-2
32.4 Storage Process................................................................................. 32-3
32.5 Lubricant Handling......................................................................... 32-3
32.6 Equipment Modification................................................................32-5
32.7 Condition Monitoring.................................................................... 32-5
Allison M. Toms 32.8 Re-Greasing Program..................................................................... 32-7
GasTOPS Inc.
32.9 Disposal............................................................................................. 32-7
George J.W. 32.10 Managing the Lubricant Database................................................ 32-7
Staniewski 32.11 Program Evaluation and Benchmarking..................................... 32-7
Ontario Power Generation References.....................................................................................................32-8

Lubrication can significantly affect equipment reliability by reducing operational costs, improving
production, and increasing safety. Several factors impact machine and lubricant longevity. These are
choosing an appropriate lubricant for the machine and its application, properly applying the lubricant
to the machine, and maintaining the quality of the lubricant throughout its life in the machine and in
storage [1].
Procedures need to be in place for engineering, operation, and maintenance to manage lubricant
selection, assessment, procurement, storage, handling, condition monitoring, contamination control,
disposal and continued development to ensure safe, reliable and economic operation of lubricated
equipment. All of these activities are a part of fluid management process, which if properly implemented
can directly influence equipment reliability and control production costs.

32.1  Lubricant Selection Process


Two main principles govern the lubricant selection process. First, users generally follow the original
equipment manufacturer (OEM) recommendations particularly with respect to the lubricant type and
viscosity grade. It is of general belief that the OEM has determined the optimum lubricant characteris-
tics during the design process and confirmed this selection through testing. However in most cases, the
OEM recommendations are very conservative to ensure satisfactory performance under a wide range
of operating conditions and while under warrantee. Therefore, knowing the specific application and
operating conditions, lubrication recommendations can be modified, although this must be done with
caution.
Second, there is often the desire to consolidate lubricant types (brands, viscosities) to minimize the
possibility of lubricant misapplications and to control the storage and handling costs. Preference may
also be given to the best quality lubricants available locally [2].

32-1
32-2 Lubricants

Some industries may have additional, unique requirements that must be taken into account during
the lubricant selection process, for instance, special safety-related equipment that must operate dur-
ing and immediately after accidents to ensure a safe shutdown processes; preference for lubricant type
(i.e., low halogen content to reduce the stress corrosion cracking process on stainless steel applications
or nonmetal anti-seize pastes); and a preference for nontoxic and readily biodegradable lubricants for
systems having contact with bodies of water.
In general, only engineering groups experienced in lubrication should have the authority to approve
lubricant changes preferably after consultation with lubricant and machinery OEM. Both the initial
lubricant selection and the subsequent changes must be documented and filed for future reference.

32.2  Procurement Process


Once a lubricant has been selected, it is important to ensure that the purchased product is consistent
with respect to its stated specifications and performance. For lubricants delivered in closed containers,
the qualification process should be done on the evaluation of the published technical data (e.g., Product
Information Sheet and Material Safety Data Sheet). For lubricants in special applications such as safety-
related equipment, users may develop their own technical specifications, usually in consultation with
machinery and lubricant OEMs.
With respect to purchasing, end users have two options. One option is to purchase lubricants based
on the best price available at a particular time or select a lubricant supplier and engage in a special
commodity contract. The benefit of purchasing lubricants without any prearrangement allows access to
multiple lubricant manufacturers and thus competitive pricing.
The other option is a commodity agreement. This option may include access to
• A wide range of lubricants
• Professional support of lubrication-related services
• Lubricant databases on worldwide operating experience
• Training programs
• Technical support such as assistance in the lubricant selection process, support in understanding
lubricant degradation mechanisms and products
• Specialized lubricant-related software
A commodity contract may even include access to lubricant research centers to address specific lubri-
cant problems. Without a contract, such access is generally limited.

32.3  Quality Assurance


A lubricant quality assurance program encompasses four basic elements. First, the lubricant manufac-
turer should be ISO certified. This requirement ensures that the lubricant manufacturer implements a
recognized quality assurance program during purchase, manufacture, storage, and distribution.
Second, the lubricant manufacturer should have an effective in-house testing program to verify
critical characteristics of lubricants during each step of the manufacturing process. The manufacturer
should also have a proven record of maintaining critical characteristics of final products for a specified
shelf life.
Third, the products should be stored in appropriately sealed containers to prevent contamination. The
storage condition and the delivery process should be well-defined and maintained so that the lubricant’s
critical characteristics are not affected during transport and delivery.
Finally, the product certification statement specifies the lubricant properties at the time of shipment
and can be used to establish a direct link between the required configuration of the product specified by
the user and the quality of the delivered lubricant.
Fluid Maintenance 32-3

It is strongly recommended that the end user perform quality assurance tests during lubricant deliv-
ery and periodically during storage to verify the product’s characteristics. In cases where the test results
are outside the user specification, the entire batch should be quarantined and the test repeated. The
use of an independent laboratory for this testing is recommended. If the retesting confirms that the
lubricant quality is outside the user specification, the entire batch should be returned to the lubricant
manufacturer.
For lubricants supplied in bulk form, a two-phased screening program may be considered. The first
phase screening is performed upon delivery and consists of basic properties (i.e., viscosity at 40°C,
Fourier transform infrared spectroscopy (FTIR), elemental analysis, and ISO cleanliness). If test results
meet the specification, the oil is transferred to a temporary oil storage tank. The second phase of the
acceptance test is performed for historical trends and consists of all required physical and chemical tests
for the machine and application.
Regardless of the packaging type and size, the lubricant container must be tagged. The following
information is recommended:
• Lubricant name
• Lubricant lot or batch number
• Lubricant supplier name and address
• Manufacturing date
• Expiration date
• End user identification code for the particular lubricant

32.4  Storage Process


Lubricant inventory and housekeeping need to be maintained. Only approved containers can be used
for lubricant storage and preferably they should be located indoors in specially designated areas. The
primary objective during storage is to protect lubricants from contamination and chemical degrada-
tion. The storage facility should be temperature controlled, 18°C–27°C (65°F–80°F), have <30% relative
humidity and have an adequate ventilation system to prevent oil vapor buildup [3].
Large unopened drums are stored on a rack, preferably in horizontal position with bung orientations
at the 3 and 9 o’clock positions. An essential element of good lubricant storage is to properly rotate the
new drums to ensure “first in, first out” practice is applied. Carts should be available for moving con-
tainers from the delivery point to the rack or to the dispensing area.
In the dispensing area, the opened oil containers should be attached to desiccant breathers to reduce
dust and moisture ingression (Figure 32.1). The new, unused lubricants might have a poorer ISO cleanli-
ness than desired. Additional filtration units with variable speed pumps should be used to reduce new
oil contamination.

32.5  Lubricant Handling


Corporations with a dedicated lubrication crew usually achieve good to excellent equipment reliability.
These results are due to the continuous monitoring of oil levels and oil quality and by applying consis-
tent processes in lubricant application activities.
One such process is oil transfer from the original container to the equipment (Figure 32.1). The pref-
erable technique depends on the volume of the transferred oil. For larger volumes, using the original
containers is preferred. Once opened, a desiccant breather is placed on the container vent. The lubri-
cant is discharged through a portable filtration unit with water-absorbing and solid particulate removal
cartridges.
For transferring smaller quantities of oil, specially designed secondary containers are recommended
(Figure 32.2). These secondary containers are easily labeled, provide a safe and easy method for oil
32-4 Lubricants

FIGURE 32.1  Dispensing area: opened oil drums with desiccant breathers.

FIGURE 32.2  Storage area with secondary containers.

transferring and dispensing processes, and control oil quality during storage. The containers should only
be filled with pre-filtered oils so that no additional filtration is required during the oil top-up process [3].
If required, portable filtration units may also be used in situ to reduce solid contamination directly at
the equipment location to maintain proper oil cleanliness level. However, an oil-monitoring program is
necessary to identify cleanliness trends.
End users should be committed to continuous health, safety, and environmental improvements and
should apply best lubrication practices and technologies in order to minimize potential negative effects
of lubricants on the environment or on human health such as accidental oil spills and contamination to
groundwater. Despite the best prevention efforts, accidents and human error can occur. Therefore, end
users should be prepared for all accident scenarios and have applicable fire, safety, and cleanup operat-
ing procedures and supplies.
Fluid Maintenance 32-5

FIGURE 32.3  Typical oil-sampling port with quick disconnect, sight glass, and tube that inserts into the oil
reservoir.

32.6  Equipment Modification


Maintaining good oil conditions may require minor modifications to critical equipment. A design
review process should be performed on seals, filters, and air breathers for all equipment in order to
reduce external contamination and maintain oil cleanliness below the generic target for each equip-
ment type.
First, the equipment should have a breather, which prevents moisture and dust ingress into the oil
compartment.
Second, rubber seals may be replaced with special labyrinth type seals. These seals prevent dust and
moisture ingress as well as generation of small rubber particles due to inadequate lubrication during
start-up operation.
Some modifications are also required to consistently obtain representative oil samples and to per-
form satisfactory in situ filtration. It has been proven that installation of special oil sampling ports
(Figure 32.3) improve the oil monitoring and filtering processes. The quick disconnect fittings are criti-
cal to perform efficient and safe in situ filtration.
Improper practices such as varying sample location on a machine, taking cold oil samples, etc., can
generate erroneous sample data.
Oil-sampling locations should ensure that representative samples are obtained. A few examples of
suggested port locations are provided:
• In forced oil circulation systems, the preferable sample port should be located on the oil return
line (Figure 32.4). The concentration of wear particles in this line is greater than in other loca-
tions, thus increasing the probability of identifying changes in the component wear rate. In addi-
tion to the primary sample port, there may be a need for a number of additional secondary sample
ports also located downstream of critical components.
• In splash oil-lubricated equipment, the preferable sampling port location is on the sump wall,
usually with a stainless pitot tube extension to an area of high turbulence (Figure 32.5). A similar
option is a drain valve with a tube extension to allow sampling above the sump bottom.

32.7  Condition Monitoring


An effective fluid management program requires lubricant testing. The purpose of condition monitor-
ing is to identify the type and rate of change of contaminants and wear in lubricants and critical equip-
ment. Additional physical or chemical tests are also performed to monitor important oil characteristics.
32-6 Lubricants

Primary sampling port location


Secondary sampling port location

FIGURE 32.4  Recommended oil-sampling locations for circulation system.

Dipstick location

Oil level
Pitot tube

Drain plug

FIGURE 32.5  Recommended oil-sampling locations for splash lubrication applications.

Since a condition-monitoring program can be expensive and time-consuming, only selected critical
equipment is generally included in this program. In most cases, the selected equipment includes the
critical safety-related machines, equipment having critical impact on production, and the most expen-
sive machinery. In some cases, it is also beneficial to include equipment with a poor operation record.
The criticality and the accessibility of the selected equipment is reviewed to establish an initial
sampling frequency and to determine which oil tests should be performed. Initially, sampling fre-
quency may be higher in order to generate a reliable baseline for test methods. An oil baseline should
be taken at the beginning of the condition-monitoring program and after equipment overhauls or oil
changes. Baseline results should be obtained during normal operation conditions. The purpose of
valid baseline data is to identify the present condition of equipment. Overtime, the optimum alarm
and trend limits will be established to determine the rate of change of lubricants and component
deterioration. Specific testing and sampling frequencies will also be refined over time as determined
by data analysis [1].
Preferably, a representative sample can be obtained from equipment with minimum requirements for
flushing the sample line. However, all uncirculated oil in the sample line must be flushed prior to taking
the actual sample for testing. Sample volume should be kept to a minimum. A typical oil sample vol-
ume is approximately 100–200 mL depending on the oil-testing regime. Wherever possible, transparent
plastic sample bottles (i.e., polyethylene terephthalate [PETE]) should be used instead of opaque plastic.
ASTMI D4057 and D4177 are good reference guides [4].
With regard to contaminants, particle counts (ASTM D7647, D7416, and ISO 11171), trace metal anal-
yses (ASTM D6595), and FTIR (ASTM E2412, D7418) are generally performed. For solid contaminants,
Fluid Maintenance 32-7

ferrous particle count, patch test (ASTM D7670 and D7684), or ferrography tests (ASTM D7690) can
follow, as needed [4]. Additional moisture contamination testing may also be desired.
With regard to oil physical and chemical characteristics, the typical tests include viscosity (ASTM
D445), acidity, relative permittivity (ASTM D7416), antioxidant content (i.e., via RULER method,
ASTM 6971), FTIR (lubricant additives), and ASTM color (D1500). Additional tests may be added to
improve efficiency of the oil analysis program.
It is important that oil screening test results are within an acceptable repeatability and reproducibility
range. Wherever possible, standards (e.g., ASTM, ISO) should be followed. For instruments that are not
covered by standards, a quality process should be developed that includes repeatability and reproduc-
ibility limits.
Since the effectiveness of the condition-monitoring program depends on adequate and timely iden-
tification of initial changes in equipment condition, extensive knowledge and experience is required
from individuals involved in the analysis of oil test results. To provide a strong base for a learning
program, it is recommended that inspection of failed components include those who are involved in
the oil condition-monitoring program. All failure analysis information should be communicated with
everyone involved in the machinery maintenance and performance analysis programs and stored for
future reference.

32.8  Re-Greasing Program


Greasing is another critical element of the fluid management program. Statistics indicate that a signifi-
cant portion of grease-lubricated bearings failed prematurely mainly due to inadequate lubrication or
over-greasing. Although significant improvements have been recognized in the last decade through
both basic research and better field practices, most users still experience a high failure rate. Only a small
number of end-user organizations have an effective grease program, which allows them to control the
amount of grease during re-greasing process, and provide an in-service condition monitoring of the
grease and its lubricated components. Greased components should be included in every fluid manage-
ment program.

32.9  Disposal
Disposal of used lubricants should be in accordance with government and local state or province
requirements, facility procedures, and other applicable documents. Of particular importance is the seg-
regation of different lubricant types (i.e., synthetic vs. mineral) for easier recycling.

32.10  Managing the Lubricant Database


A computerized database of lubricants used in all equipment should be developed and properly con-
trolled and maintained. Such a database may become the official reference and takes precedent over
other information included in documents such as manufacturer manuals, drawings, or procedures. This
database should be available to all staff on a read-only basis. Separate databases are often used for con-
dition monitoring, machinery configuration, and failure analysis. If possible, these databases should
be linked. Instrument calibration and data validation processes must be included in the condition-­
monitoring database to ensure a proper quality assurance testing process is in place.

32.11  Program Evaluation and Benchmarking


Formal assessment should be performed on a periodic basis to identify gaps between current perfor-
mance and excellence, to improve safety and reliability, to drive down costs, and to verify effective-
ness of corrective actions. Usually, assessment consists of four basic steps: planning, gathering data,
32-8 Lubricants

documenting, and communicating. The planning step is to set clear objectives and scope the boundaries
of the assessment. Gathering information should include direct observation, document review, inter-
views, analysis, and conclusions. The documenting step should include methodology used, findings,
conclusions, recommendations, and an executive summary. It is important to provide information on
both the strength and adverse conditions of the program. Finally, the communication step should be
considered as a learning mechanism and therefore shared with all personnel involved in the program as
well as the major stakeholders and senior management.
To improve the process of gathering information, it is beneficial to set a list of performance criteria
for the lubrication program and group them in major categories. By assigning score processes for each
performance criteria and then normalizing them within each major group, the final assessment may be
more objective and easier for future trending.
Finally, it must be realized that the oil management and condition-monitoring program are expen-
sive. Therefore, cost benefit or cost avoidance analysis should be routinely performed showing the eco-
nomic benefit of these programs; in many cases, the value of uninterrupted production is paramount.
Lack of such cost/benefit calculation might reduce the scope or even eliminate the program.

References
1. Toms, L. A. and A. M. Toms, Machinery Oil Analysis—Methods, Automation & Benefits, 3rd edn.,
STLE, Park Ridge, IL, 2008, ISBN: 978-0-9817512-0-7.
2. Bohn, E., GM invests in lube program upgrades, Machinery Lubrication, September 2001, 48–53.
3. Johnson, M., Proper storage and handling of lubricants, Tribology and Lubrication Technology, 64(2),
February 2008, 22–30.
4. ASTM Standards, Petroleum Products, Lubricants, and Fossil Fuels, Vol. 05.01–05.05.
III
Wear Materials

33 Wear Materials  William A. Glaeser and Robert W. Bruce..........................................33-1


Introduction  •  Polymers  •  Carbon  •  Metals  •  Ceramics  •  Conclusion  •  References
34 Friction and Wear of Polymer Materials  Thierry A. Blanchet..................................... 34-1
Introduction  •  Friction  •  Wear  •  Pressure × Velocity Limit  •  Tribological Properties of
Polymers  •  References
35 Metals   Thomas W. Scharf.................................................................................................... 35-1
History and Introduction  •  Definition of Metallic-Based Friction and
Wear  •  Metals and Alloys  •  Shape Memory Alloys  •  Nanocrystalline Metals and
Alloys  •  Acknowledgments  •  References
36 Wear and Lubrication of Ceramics  Said Jahanmir........................................................ 36-1
Introduction  •  Wear under Sliding Conditions  •  Wear under Rolling and Abrasive
Conditions  •  Sliding Wear Models  •  Boundary Lubrication  •  Design with
Ceramics  •  References
37 Composite Materials  Li Chang and Klaus Friedrich.........................................................37-1
Introduction  •  Wear Testing for Polymer Composites  •  Developing Wear-
Resistant Polymeric Composites by Using Conventional Fillers  •  Wear of Polymeric
Nanocomposites  •  Summary  •  Acknowledgments  •  References
38 Coatings and Surface Treatments  Arup Gangopadhyay............................................... 38-1
Introduction  •  Thermal Spray Coatings  •  Electrochemical Coatings  •  Physical
Vapor Deposition  •  Chemical Vapor Deposition Process  •  Case Hardening of
Steels  •  Automotive Trends  •  References
39 Low Friction Coatings  Thomas W. Scharf........................................................................ 39-1
Introduction and Definition of Low Friction  •  Low Friction Coatings
and Mechanisms of Lubrication  •  Coating Processing Routes  •  Key
Applications  •  Acknowledgments  •  References
40 Wear Coating and Treatments  Gary L. Doll and Allan Matthews.............................. 40-1
Introduction  •  Wear Mitigation by Surface Treatments  •  Conclusions  •  References
41 Coatings and Surface Treatments: Interactions with
Lubricants  Staffan Jacobson.................................................................................................41-1
Introduction  •  Lubrication Modes: When and How Does the Presence of a Coating
Matter?  •  How and Why Is the Friction Modified?  •  Effect of Shifting the Curve or
the Borders in the Stribeck Curve  •  Consequences of the Complexity  •  Lubrication of
Low-Friction Coatings  •  Nitrides  •  Conclusive Summary  •  References

III-1
33
Wear Materials
33.1 Introduction..................................................................................... 33-1
33.2 Polymers............................................................................................ 33-2
33.3 Carbon............................................................................................... 33-2
33.4 Metals................................................................................................ 33-3
Aluminum  •  Bronze  •  Cast Iron  •  Cobalt  •  Diamond  • ​
Nickel  •  Steel  •  Titanium  •  Wear Treatments for Metals
William A. Glaeser 33.5 Ceramics...........................................................................................33-4
Battelle
CerMet  •  Ceramic  •  Glass
Robert W. Bruce 33.6 Conclusion........................................................................................33-6
GE Aviation References.....................................................................................................33-6

33.1  Introduction
Wear materials are generally really combinations of materials that result in low wear, low wear of one of
the materials in contact, or low wear of both. One such combination is that of bronze bushings support-
ing a steel shaft, which minimizes adhesive wear by means of the bronze transfer film developing on the
steel surface resulting in minimal total wear. Note that some bronzes are more wear resistant than oth-
ers and that it is beneficial to make the steel shaft hard and smooth. Plastic materials, especially Teflon
and polyimides, may show a similar behavior, and excellent wear life. And the combination of metal
surfaces with ceramics can also be aided by a transfer film of metal on the ceramic surface, minimizing
the wear rate.
Many wear-resistant materials are used in the form of coatings to create a wear-resistant surface
while maintaining the suitable mechanical and physical bulk properties of the component. Coatings
like T400 and T800 were originally designed to rub against surfaces covered with the same coating.
Tungsten carbide was developed by Krupp (Manchester 1968) industries of Germany as a cutting tool
material for high strength steel. Many of today’s physical vapor deposition (PVD) and chemical vapor
deposition (CVD) coatings were developed as coatings for cutting tools, requiring not only low wear
but also high shock and heat resistance. Therefore, most of these coatings will provide low wear against
steel counter faces. This likely means that diffusion of steel constituents into the coating is minimal to
keep adhesive wear low. How these materials rank in wear resistance against nickel, cobalt, and titanium
alloys is currently not public information. Generally available is the maximum use temperature limit
due to oxidation. Often heralded is the coating hardness, which is difficult to measure when the coating
is only microns thick, and may not have much correlation with wear resistance, let alone wear resistance
against any particular material.

33-1
33-2 Wear Materials

The selection of tribological materials depends on the following:

1. Chemical compatibility with the operating environment


2. Physical compatibility, especially thermal
3. Mechanical properties
4. Tribological suitability
5. Cost, increasingly this includes maintenance and/or life cycle cost

The wear resistance of materials is a function of the type of wear experienced. Some bronzes may show
good wear resistance when rubbing against steel under adhesive wear conditions, but may show poor
resistance in erosion or abrasive wear conditions.
Wear resistance does not necessarily imply low friction. Many wear-resistant material combinations
exhibit friction coefficients of 1.0 or even higher. Friction may be reduced by applying liquid lubricants
or solid lubricant coatings over the wear-resistant surface. Applying such friction coating on one side of
the contacting surfaces may be adequate; applying low-friction coatings to both surfaces may result in a
higher friction coefficient. A brief description of were material types follows (Glaeser 1992).

33.2  Polymers
Some polymers have excellent tribological properties. Main limitations are temperature capability,
velocity capability, and load-carrying capacity, as well as the pressure–velocity (PV) limit. The PV limit
is the product of load and speed, which, when multiplied with friction coefficient, relates to the frictional
heat generated. Polyimides tend to have a high PV limit and are therefore extensively used as polymeric
wear materials. Polymers suffer in time from reduced properties due to loss of plasticizers (new car
smell), oxidation, and thermal degradation, as well as corrosion due to pollution gasses, vapors, and
humidity (Jansen 2004). This degradation can result in exponential wear or oxidation after years of
service. Teflon or polytetrafluoroethylene (PTFE) can provide a friction coefficient <0.1 up to 250°C
(480°F); polyimides have a higher friction coefficient 0.2–0.4; and some will be able to operate at 315°C
(600°F) and high PV. In some cases, polyamide, nylon, can be successfully used, especially in contact
with water. However, water can be absorbed into the polymer, changing part dimensions, and after mul-
tiple freeze–thaw cycles damage the polymeric structure.
Frequently polymeric materials contain fillers, solid lubricants, mild abrasives, ceramics, glass, and
fibers to improve mechanical properties. The fillers can also affect the friction coefficient and wear rate.
The hardness and surface roughness of the metal counter surface can affect the wear rate of the poly-
mer significantly. In many cases, 0.1–0.2 μm (4–8 μin.) Ra is specified, and a hardness of >50 HRC.

33.3  Carbon
Carbon or graphite shows low adhesive wear against most metals; this is especially beneficial with
metals prone to galling, like aluminum, titanium, and stainless steel. Carbon is effectively used in
diamond-like coatings (DLCs), usually in conjunction with liquid lubricants. Carbon/graphite is a mix
of natural or man made graphite and carbon black heated to very high temperatures to from a hard solid
material. The more graphite in the mix, the softer the product. Some additives used in liquid lubricants
to protect iron surfaces seem to work satisfactorily on DLC surfaces.
Graphite–metal composites can provide low friction at temperatures up to 540°C (1000°F), the maxi-
mum operating temperature for graphite in the presence of oxygen. Graphite can be a source of pitting
corrosion of aluminum in humid environments.
Wear Materials 33-3

The roughness of a metal counter surface can affect the wear rate of the carbon/graphite; a roughness
of <0.6 μm (<24 μin.) is often recommended.

33.4  Metals
The wear resistance of metals, especially steels, has been related to their hardness. For some steels, a
5 HRC increase in hardness can reduce wear by 2×. Many steels provide excellent wear rates at 55–60
HRC. At some point however, the increase in hardness can lead to significant brittleness and an increase
in wear rate or decrease in fatigue life (Glaeser 1980).

33.4.1  Aluminum
Aluminum can be an adequate counter surface for polymeric materials. In contact with metals, it tends
to gall. A detailed description of alloys and heat treatments, Aluminum Standards and Data, is pub-
lished by the Aluminum Association. Most popular engineering alloys are 2024, 6061, 7075 and recently
the 2020, 2090, 2099, 2199 low-density lithium alloys.

33.4.2  Bronze
Bronze alloys are defined as mixtures of copper and tin, plus possibly other alloying elements. Bronze
bushings provide low wear against hard and smooth steel shafts. Some contain aluminum or up to 2%
phosphorous for strength. Aluminum bronze may contain only copper and aluminum. Performance
depends on whether or which liquid lubricant is used. The use of bronze is limited by the embrittlement
occurring while cooling down from elevated >288°C (>550°F) temperatures. The use of lead and beryl-
lium containing bronzes is greatly diminished for health reasons.

33.4.3  Cast Iron


Cast iron, with 2%–4% carbon, was used extensively at the start of the industrial age. Its wear proper-
ties are good due to the graphite lamellae in its structure and the nodular graphite inclusions in later
developed nodular cast irons. These inclusions produce a heterogeneous structure. Some cast irons have
significant chrome content to improve wear resistance. Nodular iron is often used as a substitute for
steel shafting.

33.4.4  Cobalt
Cobalt alloys are used for high temperature wear applications, and cobalt oxide is considered to reduce
friction at elevated temperature. Most popular alloys are HA25, HA188, and L605 alloys.

33.4.5  Diamond
Diamond coatings are effectively used on single-point cutting tools for aluminum, as aluminum
tends to adhere and build up on almost all other surfaces. DLCs are increasingly used to coat one
side of a liquid-lubricated contact to minimize the effect of asperity contact. This applies to magnetic
storage disks as well as an increasing number of automotive applications such as piston pins and fuel
injector parts.
33-4 Wear Materials

33.4.6  Nickel
Nickel-based alloys are used for high temperature and corrosive wear applications including aggressive
chemical environments. Most popular alloys are IN718 and IN625.

33.4.7  Steel
Carbon steels with 0.02%–1.7% carbon can be very hard and very wear resistant. AISI 52100 and
M50NiL are used for the manufacture of rolling bearings. Stainless steels with <0.02% carbon are softer
but protected for oxidation by a thin chrome oxide or in few cases aluminum oxide surface layer. A good
wear-resistant high-temperature stainless is 440C. Stainless steels are prone to galling, although some
specialty alloys have been created to minimize these galling tendencies. Nitriding and carburizing the
stainless alloys creates carbides and nitrides reducing the amount of chrome available for the protec-
tive surface layer, and thus resulting in rust. Tool steels and maraging steels are used as wear-resistant
materials.

33.4.8  Titanium
Titanium is fairly corrosion resistant and lightweight. Galling tendencies are similar to those of alumi-
num and stainless steel. Many different alloys exist, but Ti6-4 is most widely used.

33.4.9  Wear Treatments for Metals


Numerous treatments have been developed that can improve wear resistance of metals (see Table 33.1).
Many are effectively used by treating one or both of the metal surfaces in contact. Many increase the sur-
face hardness. Beyond some point, the brittleness can become a factor. Alternatively the chemical nature
of the surface can be modified by applying a coating, often consisting of ceramic material. During the
past 25 years, many surface cleaning, conversion coating, and plating processes are no longer acceptable
for environmental, health, and safety (EHS) reasons. Consolidation to the most effective methods has
occurred and most remaining processes are under industry and military specifications. Presently, the
PVD-coating processes are seeing rapid growth with increased numbers of coating materials, improved
performance, reduced coating temperatures. While these various treatments can improve wear per-
formance and sometimes corrosion resistance, they may negatively affect fatigue performance of the
substrate.

33.5  Ceramics
Wear resistance of ceramic materials seems more dependent on toughness than hardness, and tough-
ness is not only difficult to access, but also strongly dependent of manufacturing parameters. Ceramics
used for tribological applications:

33.5.1  CerMet
CerMets consist of pieces of ceramic materials bonded by a metal matrix. Examples are tungsten car-
bide, chrome carbide, chrome oxide, and titanium carbide, in binders of cobalt, nickel, or steel. Cobalt-
cemented tungsten carbide was invented by Krupp in the 1920s for use as cutting tools for high strength
steels in the armament industry. It is now also used for the punch in the deep drawing of aluminum
beverage cans, although sometimes replaced by nickel-cemented carbide to reduce the corrosive effect of
process lubricants or cleaning fluids. Some of these materials may be applied as coatings, providing the
wear resistance of the CerMet together with desirable properties of the substrate.
Wear Materials 33-5

TABLE 33.1  Treatments


Method Detail Process Temperature Materials (°C) Depth (μm) Comments
Boriding 940 Fe 50–200 B2O3 MP 450°C,
may cause rust
Carbonitriding 650–900 Fe 70–500
Carburizing 850–1000 Fe 300–6000 0.1%–0.3% C
steels
Cyaniding 700–800 Fe 250–750
Flame, induction 750–1000 Fe 300–6000
harden
Heat treating 600–1100 Fe through
Nitriding 500–600 Fe, Ti 125–450, 20–40
Siliconizing 950–1050 Fe 75–200
Shot peening 20 Work hardening 20–200
Conversion coating Anodize 6–40 Al, Ti 2–150, 0.1–0.3 Thicker layers are
more porous
Conversion coating Chromate 20–30; dry at 100–200 Al, Mg 0.025–1, 12–15
Conversion coating Dichromate 10–70; dry at 60–700 Mg 3
Conversion coating Phosphate 15–65 Fe, Mg, Ti 0.5–15
Plating Chrome 35–800 Metals 2–250
Plating Nickel 35–65 Metals 10–100
Plating Silver 15–30 Metals 3–50
Plating Zinc 60 Metals 3–120
Wear coating CVD 500–1000 Metals 5–50
Wear coating HVOF 50–600 Metals 75–200
Wear coating Plasma spray 50–600 Metals 75–200
Wear coating PVD 350–600 All 1–20

33.5.2  Ceramic
The most popular ceramic wear materials include alumina, silicon carbide, silicon nitride, and zirconia.
Partially stabilized zirconia has good wear resistance against the other listed ceramics, but the manu-
facturing process can have significant influence on performance. Uses may require electrical or ther-
mal-insulating properties, high temperature capability, and wear resistance. Thermal shock resistance
can be an issue (Glaeser 1997). With thermal shock resistance defined as tensile fracture stress divided
by both elastic modulus and thermal expansion coefficient, alumina may have a thermal shock resis-
tance of only 2% that of tool steel. The brittleness of ceramics can be addressed by increasing section
thickness, toughening the ceramic through compositional changes and mixes as in partially stabilized
zirconia, densifying the ceramic using appropriate sintering aids, or providing a structure that can cope
with stress patterns, as for instance the columnar structure of the PVD zirconia-based thermal barrier
coatings. The number of ceramic coating compositions is growing rapidly, increasing durability, tem-
perature capability, and oxidation resistance.

33.5.3  Glass
Glass-based ceramic materials are not known for their tribological properties except for corning’s
pyrex and pyroceram. These are however used as lubricants in some high temperature metalworking
operations, providing separation of surfaces by the squeeze film mechanism and the viscosity of the
hot glass.
33-6 Wear Materials

TABLE 33.2  Wear Material Combinations and Wear Modes


Material Abrasive Adhesive Cavitation Corrosive Erosion Fatigue Fretting Impact
Polymer Teflon on steel Rubber Polyimide
Polyimide on on steel
steel
Carbon Most metal Graphite
surfaces
Metal Hard Cast iron M50NiL Ni alloys M50NiL Bronze on
steel against cast steel
iron
Stellite 6B Hard steel NiAl
against bronze
bronze
Hard steel NiTi
against
WC–Co
Ceramic CerMet Si3N4 against Si3N4 WC– Si3N4 on WC–Co WC–Co on
WC–Co steel Co steel on steel WC–Co
Composite Carbon
composite
against self

33.6  Conclusion
Many surface treatments are available to improve wear resistance of metals. The rapid development of
coating materials and processes are leading to use not only on metal but also on polymeric surfaces.
Wear materials are really wear material combinations that provide low wear under certain wear mecha-
nisms and conditions of load, velocity, and temperature. Table 33.2 shows some of the proven material
combinations that minimize wear, organized by wear mechanism.

References
Aluminum Standards and Data, The Aluminum Association, Washington, DC, 2000.
Glaeser, W. Materials for Tribology, Elsevier, Amsterdam, the Netherlands, 1992.
Glaeser, W. Wear-resistant hard materials, Tribology Data Handbook, CRC Press, Boca Raton, FL, 1997.
Glaeser, W. Wear resistant materials, ASME Wear Control Handbook, ASME, Fairfield, NJ, 1980.
Jansen, J.A. The plastic killer, Advanced Materials and Processes, June 2004, pp. 50–53.
Manchester, W. The Arms of Krupp, Brown & Little, New York, 1968.
34
Friction and Wear of
Polymer Materials
34.1 Introduction.....................................................................................34-1
34.2 Friction..............................................................................................34-2
Viscoelastic Deformation  •  Deformation Contribution to Polymer
Friction  •  Adhesion Contribution to Polymer Friction
34.3 Wear...................................................................................................34-9
Abrasive Wear  •  Adhesive Wear
Thierry A. Blanchet 34.4 Pressure × Velocity Limit..............................................................34-12
Rensselaer Polytechnic 34.5 Tribological Properties of Polymers...........................................34-13
Institute References................................................................................................... 34-14

34.1  Introduction
Polymers are materials comprised of unit chemical groups, or mers, which repeat to build up extended
molecules. Depending on the conditions under which the polymerization reaction is conducted, the
resulting molecules may either be linear chains or also have branches off the main chains. Furthermore,
the collection of molecules comprising the polymer may either be amorphous, or semicrystalline with
domains in which collections of neighboring molecular segments are able to adopt a densely packed
arrangement, depending on the structure of the mer repeating unit and branching of the chain mol-
ecules, as well as conditions such as cooling rate and pressure of any prior thermal processing. In addi-
tion, the polymer may either be a thermoplastic where above a melt temperature the crystalline domains
dissociate and individual molecules are free to move relative to one another, thus enabling melt process-
ing, or a thermoset where chemical bonds are introduced between neighboring chains cross-linking
them into a network which may no longer melt.
As a class of materials, polymers offer many attractive attributes that have led to their widespread tech-
nological application, including tribological applications such as bushings for plain bearings, retainers
for roller bearings, seals, gears, and even hip and knee joint replacements. Such attributes may include
light weight, low cost per volume, ease and flexibility of processing, impact and vibration absorption,
and chemical inertness. Polymers may also be known for inherently low friction as with polytetra-
fluoroethylene (PTFE; i.e., Teflon®) or low wear rate as with polyamide PA “nylon,” though rarely both.
Thus a major activity in polymer tribology over the past several decades has been the addition of hard
reinforcement or solid lubricant fillers to address deficiencies in wear resistance or self-lubricity, though
such composite tribology will be covered in Chapter 37. Here, focus will instead be on the tribological
behavior of unfilled polymers. Several excellent reviews of this topic have been previously published,
including Briscoe and Tabor (1977) and Lancaster (1972).

34-1
34-2 Wear Materials

34.2  Friction
Friction is commonly considered to arise from two contributions: deformation and adhesion. The
adhesion contribution originates from the need to continuously shear the contact interface in order to
maintain a sliding motion, and thus its importance may be lessened by the use of lubricants to lessen the
shear strength of that interface or by instead using rolling to achieve the relative motion. Deformation
contributions may originate from counterface material that must be continuously deflected to accom-
modate the approaching counterbody, as well as from that counterbody itself if its motion is achieved
by rolling during which it may continuously cycle deformable material back into the contact.
Figure 34.1 portrays examples of various contact combinations of a deformable and a rigid body, with
relative motion of the moving body either by rolling or sliding, as well as the likely relative importance
of deformation and adhesion contributions to overall friction. For deformable material entering the
contact, energy must be input to its deflection, with a corresponding pressure distribution over this
entrance region creating a resistive force or moment opposing this sliding or rolling motion. Depending
on the nature of this deformable material which has just entered the contact, when it subsequently
enters the trailing portion of the contact it may relax back to its original state with its deformation
energy recovered and a corresponding pressure distribution over this trailing region urging the body
forward, countering the resistive force or moment. Otherwise, it may remain deflected with the input

Rigid roller Viscoelastic roller

Energy In
loss pu
te
ne
rg
Recovered y
energy
Recovered
y

energy
rg
ne
te
pu

Energy
In

loss

(a) Deformation-dominant (b) Deformation-dominant

Rigid slider Viscoelastic slider

Recovered
y

energy
rg
ne
te
pu

Energy
In

loss

(c) Deformation and adhesion (d) Adhesion-dominant

FIGURE 34.1  Various contact configurations of an upper body rolling (a and b) or sliding (c and d) relative to a
lower counterface, where in each configuration the upper body is either rigid (a and c) or viscoelastic (b and d) while
the mating counterface is the other. If the configuration continually introduces viscoelastic material into the con-
tact’s entrance, the path of energy input to its deformation there as well as the partitioning of this energy between
recovery to the contact at its exit or dissipation within the viscoelastic material is indicated by arrows. In each
configuration, the relative predominance of either this deformation or adhesion to the overall friction is indicated.
Friction and Wear of Polymer Materials 34-3

strain energy lost to dissipation of the deformation process, and friction thus resulting due to the lack
of a pressure distribution over the trailing region to balance that created over the entrance region of the
contact. Bodies deforming plastically will partition input energy more fully toward loss dissipated into
their bulk with significant deformation contributions to friction, while perfectly elastic bodies which
relax instantaneously toward their original state instead partition the input energy more fully toward
recovery with resultant deformation contributions to friction less significant. In the case of polymers as
the deforming bodies, their behavior is often viscoelastic where rather than instantaneous their defor-
mation will lag behind loading, and it is the extent of this lag that will dictate how heavily the input
energy is partitioned toward dissipative loss to the bulk rather than recovery toward the contact’s trail-
ing region, and in turn how heavily deformation will contribute to friction.

34.2.1  Viscoelastic Deformation


A simplified review of viscoelasticity may aid discussion of its contribution to polymer friction. As a
polymer is cooled below its melt temperature, it will transition from a viscous to a rubbery elastomeric
state, where there is ample free volume as well as thermal agitation for molecules to still be unkinking
and rotating about their bonds, and continuously rearranging. With further decreasing temperature,
a polymer will experience correspondingly linear decreases in specific volume indicative of a constant
thermal expansion coefficient, until eventually a point is reached at which there is neither adequate free
volume nor thermal agitation for molecules to rearrange themselves into specific volumes that con-
tinue to decrease with temperature at the same rate. Thus a discontinuity of slope occurs in the specific
volume–temperature behavior, a point referred to as the glass transition temperature, below which the
polymer is relatively stiff with its deformability prior to failure being more purely elastic.
The viscoelastic deformation behavior of polymers that exists above the glass transition temperature
is studied by dynamic mechanical analysis where either the strain or the stress is prescribed with time
and the relationship between these quantities, especially the extent by which the strain lags the stress,
is characterized. Deformation employed during such characterization is typically sinusoidal about a
zero mean, a full cycle of which is portrayed in Figure 34.2a, and a half-cycle of which (solid lines) may
analogously be approximated by viscoelastic material which is loaded then unloaded as it enters then
exits a contact. In the example portrayed in Figure 34.2a, strain is described by

e = e0 sin(w t )

where ω is the frequency of the oscillatory deformation (radians per time). As shown, the stress leads the
strain by a phase angle δ, which thus by two-angle trigonometric identity may be described by

s = s 0sin(w t + d) = s 0cos(d)sin(w t ) + s 0sin(d)cos(w t )

The phase angle may vary between zero (stress proportional to strain and response purely elastic) and
π/2 radians (stress proportional to derivative of strain and response purely viscous). From the previous
equations, if instead the storage modulus is used to describe the ratio of the in-phase component of
stress to the strain E′ = (σ / ε 0)cos(δ) and the loss modulus to describe the ratio of the out-of-phase com-
ponent of stress to the strain E″ = (σ 0 / ε 0)sin(δ), stress may alternately be expressed as

s = e0 E ʹ sin(w t ) + e0 E ʹʹcos(w t )

The path of stress and strain over the first half-cycle, where strain is increased to ε 0 then brought back to
zero forms a hysteresis semi-loop, depicted in Figure 34.2b for the case of δ = 0.3 radians. If the material
was instead purely elastic, with a phase angle of zero, this hysteresis semi-loop would collapse about a
34-4 Wear Materials

Normalized strain, ε/ε0 Normalized stress, σ/σ0


0.8

0.6

0.4 δ
0.2
Normalized time, ωt
0
0 1.57 3.14 4.71 6.28
–0.2

–0.4

–0.6

–0.8

–1

(a)
1
Normalized stress, σ/σ0

0.8
Loading
0.6

0.4

0.2 Unloading

0
0 0.2 0.4 0.6 0.8 1
–0.2 Normalized strain, ε/ε0
(b) –0.4

FIGURE 34.2  Viscoelastic response of a polymer to oscillatory loading, phase angle δ = 0.3 radians in this exam-
ple. (a) Stress and strain as a function of time during a fully reversing cycle and (b) hysteretic stress versus strain
behavior over a half-cycle (solid lines in (a)) during which strain is increased from then returned to zero.

line of slope E′ running from the origin to the point (σ, ε 0) and back, where the elastic energy per unit
volume U input to the deformation (i.e., area under the stress strain loading curve σ 0 ε 0/2 = E′ε 02/2) is
completely recovered upon unloading. However, in the viscoelastic case the energy per unit volume
D which is lost to dissipation during loading/unloading is equivalent to the area contained within the
hysteresis semi-loop, in other words the integral over the process of σdε. If the differential dε is replaced
.
by εdt, then the integral may instead be conducted over the time of the loading/unloading:

p /w
. p
D = s de = se dt =
∫ ∫ ∫ swe cos(w t ) dt = 2 Eʹʹe
0
2
0

Thus the fraction α of elastic input energy which is instead lost to dissipation if the deformation is
instead viscoelastic

D Eʹʹ
a= =p = p tan(d)
U Eʹ

is found to be directly proportional to the tangent of the phase angle, a useful characterization of
viscoelastic deformation thus commonly referred to as the loss tangent or loss factor.

34.2.2  Deformation Contribution to Polymer Friction


Considering a rigid body under normal load W against an elastic counterface, a pressure distribution
is created that when integrated over the area of the contact projected toward the top will balance W.
Friction and Wear of Polymer Materials 34-5

Likewise, integration of this pressure distribution instead over a side projection of the contact’s area
will produce a frontal force ϕ along the surface. For displacement of the rigid body imposed along the
counterface, ϕ would equivalently represent the elastic strain energy input into the counterface per unit
displacement. If the counterface was perfectly elastic this displacement would occur without friction, as
instantaneous full recovery of this strain energy over the trailing region of the contact would maintain
an equal rear force ϕ urging the rigid body forward. If the counterface was instead a viscoelastic poly-
mer, however, it would be anticipated that of the energy input to the counterface deformation a fraction
α would be lost. Thus, an additional force Fd = αϕ = πϕ tan(δ) must be externally applied from the rear
of the contact in order to compensate for lessened recovery and achieve displacement, where the defor-
mation contribution to friction force Fd is therefore dependent on the polymer’s viscoelastic properties
such as loss factor.
In the case of a conical rigid body of included semi-angle ψ. when normally loaded against an elastic
counterface the ratio of the areas of the side and top projections of their contact, and thus of their fron-
tal force ϕ and normal load W, is cot(ψ)/𝜋. independent of elastic properties. For a viscoelastic polymer
counterface, the deformation contribution to friction force would thus be predicted as

Fd = a f = W tan(d)cot(y )

If the rigid body is rather a sphere of radius R loaded against the elastic counterface, then the ratio of
side and top contact area projections, and in turn frontal force and normal load, becomes approximately
0.143(W(1 − v2)/R 2E)1/3 where E and ν are the Young’s modulus and Poisson’s ratio of the counterface
(Greenwood and Tabor 1958). In this case, the deformation contribution to friction force against a coun-
terface instead of viscoelastic polymer would thus be

1/3
⎛ 1 − n2 ⎞ tan(d) 4/3
Fd = 0.143p ⎜ 2 ⎟ W
⎝ R ⎠ Eʹ1/3

Greenwood and Tabor (1958) performed sliding experiments of both conical and spherical indentors
against rubbery counterfaces, well-lubricated so as to minimize adhesion contributions and instead
study the behavior of deformation contributions to friction. In the case of conical indenters, at a con-
stant load the cones of decreasing included semi-angle revealed a deformation friction force which
indeed increased in proportion to the correspondingly increasing cot(ψ) as predicted, through a semi-
angle of nearly ψ = 45°, though thereafter a discontinuity to higher friction levels occurred as sharper
cones initiated tearing of the counterfaces. In the case of spherical indenters, normal load was varied at
constant sphere radius and deformation friction force was confirmed to increase linearly with W4/3 as
predicted (Greenwood and Tabor 1958).
In the experiments earlier, however, the loss factor tan(δ) was not varied and thus the connection
between viscoelastic properties and friction not clearly demonstrated. In experiments which instead
minimized adhesive contributions by utilizing rolling motion of a rigid indenter over flat polymeric
samples, Ludema and Tabor (1966) studied the deformation contribution to friction under constant load
but at temperatures ranging from approximately −150°C to 200°C. Each of the four polymers ­studied—
PTFE, nylon, polychlorotrifluoroethylene (PCTFE), and polyoxymethylene (POM)—possessed at least
one and in most cases multiple maxima in loss factor over this temperature range, and correspond-
ingly displayed maxima in rolling friction at temperatures coincident with these loss factor maxima,
an example of which is shown in Figure 34.3. Furthermore, the relative magnitudes of these friction
maxima may be better predicted by the model if, in addition to the loss factor, the elastic modulus and
its decrease with temperature is also measured and taken into consideration. In a similar study using
rubbery elastomers, where loss behavior was less dramatic though still displaying a clear dependence
34-6 Wear Materials

40

Arbitrary units
30 Rolling friction µr

20

10

(a) –150 –100 –50 0 50 100 150 200


Temperature (°C)
40
Arbitrary units

30 Damping loss
20

10

(b) –150 –100 –50 0 50 100 150 200


Temperature (°C)

FIGURE 34.3  Coincidence of friction and damping loss behaviors of polychlorotrifluoroethylene (PCTFE) poly-
mer as a function of temperature. (a) Friction experienced by a steel ball rolling over it and (b) viscoelastic damping
loss raw characterization (solid line) as well as further divided by modulus raised to the 1/3 power (dashed line) per
model furthering nearness of shape to that of friction data. (From Ludema, K.C. and Tabor, D., Wear, 9, 329, 1966.
With permission.)

of monotonic decrease with increasing temperature, Flom (1960) demonstrated the predicted linear
relationship between rolling friction coefficient and loss factor when modified by division by modulus
raised to the 1/3 power.
The predicted dependence of deformation-based friction upon viscoelastic properties is thus demon-
strated, and a mechanical characterization of loss factor of a polymer as a function of temperature is
informative of the friction behavior to be anticipated. As viscoelastic response reflects the relative rates of
deformation imposed upon a body and of relaxation of it from that deformation, the net effect of the influ-
ence of increased temperature on relaxation time can equivalently be achieved by imposing deformation at
a reduced rate. Thus viscoelastic characterization techniques employing cyclic deformation are also often
performed by varying cycle frequency 𝜔. rather than temperature. Superposition of frequency-dependent
mechanical behavior data sets collected at different temperatures onto a master curve allows the equiva-
lence of increased temperature and decreased frequency to be described, in the case of rubbery polymers
(above their glass temperature Tg) by the temperature-dependent Williams–Landel–Ferry shift factor aT

−8.86(T − Ts )
log aT =
101.5 + T − Ts

The use of a reference temperature Ts that is ∼50°C above the glass transition temperature allows these
data sets collected at different temperatures to superimpose upon a master curve if instead plotted
as a function of the product aT ω. As the appearance of frequency ω results from its proportionality
to the rate of deformation in cyclic mechanical characterization, the sliding or rolling speed should
similarly reflect the rate of deformation in a tribological contact and its viscoelastic behaviors such as
deformation-based friction should thus fall upon a master curve if plotted as a function of aT ν. Grosch
(1963) demonstrated this superposition of speed- and temperature-dependent behavior of friction using
the WLF shift factor in experiments sliding isomerized natural rubber against rough silicon carbide
paper (Figure 34.4), where on the scale of the silicon carbide particles (∼0.1 mm) they are acting as rigid
Friction and Wear of Polymer Materials 34-7

–58°C

2
µ
2

1
1

90°C

0 0
–4 –2 0 –4 –2 0 –4 0 4 8
(a) Log10 V (b) Log10 aTV

FIGURE 34.4  Friction coefficient of isomerized natural rubber sliding against rough silicon carbide paper. (a) As
a function of sliding speed v to a maximum of 30 mm/s with each set of data collected at a common temperature
ranging from −58°C to 90°C and (b) as a function of aTv where superposition onto a single master curve results from
a shift factor aT of WLF temperature dependence described by a reference temperature of Ts = 20°C. (From Grosch,
K.A., Proc. R. Soc. A, 274, 21, 1963. With permission.)

sliders traversing across the rubber. It is concluded that the attainment of a master curve with a friction
maximum when plotted as a function of aT ν was reflective of viscoelastic deformation-based friction, as
this maximum persisted despite application of magnesium oxide powder lubricant which would have
altered adhesion-based friction contributions.
By such master curve superposition behavior, friction characterized at a combination of temperature
and speed implies that similar friction should exist at other combinations of temperature and speed pro-
ducing the same product aT ν. Thus if viscoelastic properties such as loss factor of a rubbery polymer are
characterized as a function of temperature, and correspondence is noted to the temperature dependence
of deformation-based friction behavior, through the use of the WLF shift factor its speed dependence is
in turn also described. Interestingly, even below the glass transition temperature where polymers are no
longer rubbery, it has been shown that temperature and speed dependencies of friction may still be super-
imposed onto master curves, though the shift factor must instead be described by an Arrhenius rather
than WLF temperature dependence (Tabor 1974).

34.2.3  Adhesion Contribution to Polymer Friction


In an effort to lessen deformation contributions to friction, applications utilizing polymer sliders are
often configured against a relatively rigid counterface as portrayed in Figure 34.1d. Sliding motion is
achieved by application of a tangential force which maintains continuous shear failure over the instanta-
neous area of contact. Thus adhesive contribution to friction force may be described as Fa = Ars, where Ar
is the real area of contact and s is shear strength. Depending on surface species (adsorbed gases and con-
taminants, oxides, purposeful thin film lubricants, etc.) and their effect on the strength of the interface
formed by the two contacting bodies relative to that of the polymer, the shear failure that allows sliding
motion may occur either at the interface or within the near-surface of the polymer. Correspondingly
34-8 Wear Materials

the shear strength s would be either that of the interface or the polymer, with failure within the polymer
leaving a transfer film upon the counterface which may affect the interfacial shear strength and thus
locus of shear failure and friction force during subsequent passes if the sliding application causes this
location of the counterface to be traversed repeatedly.
In sliding of polymethylmethacrylate (PMMA) at room temperature and thus below its glass transi-
tion temperature, Archard (1957) noted a region of uniform friction coefficient μ over lighter normal
loads that would eventually give way to a power law decrease μ ∝ W−m of friction coefficient with further
increases of normal load W. The insensitivity of friction coefficient at lower loads is understood through
consideration of the roughness of the polymer and that Ar over which shear must occur only exists at the
tips of contacting asperities. While those asperities already in contact will each be more heavily loaded
and increase in their individual contact areas as global normal load W is increased, additional asperities
newly coming into contact are more lightly loaded over smaller individual contact areas. The net effect,
regardless of whether asperity scale contacts are elastic or plastic, is that the average contacting asper-
ity and its individual loading and contact area remain relatively unchanged, and therefore, an increase
in global normal load W leads to a simple proportional increase in the number of asperities in contact
and in turn of the global real area of contact Ar. The proportionality of real area of contact also results
in proportionality of adhesive friction force Fa to normal load and thus the observed insensitivity of
friction coefficient.
Eventually with increasing normal load the real areas of asperity contact completely occupy the nom-
inal area of the global contact patch, which is typically elliptical in shape depending on the curvatures
of the contacting bodies, and any further increases in normal load continue to increase the size of this
nominal patch which is now more fully in real contact. If this nominal contact is elastic, the contact area
increases with normal load to the 2/3 power; however, if it instead is plastic, then its area continues to
increase in proportion to normal load. As this area is in real contact and must be sheared to continue
sliding, adhesive friction force is thus also proportional to normal load W raised to a power between 2/3
and 1, resulting in the observed high-load friction coefficient behavior μ ∝ W−m, where m is between 1/3
and 0 depending on the extent of elastic and plastic nature within the nominal elliptical contact. The
critical load at which insensitive friction coefficient transitions to a friction coefficient that decreases
with increasing normal load depends on the polymer roughness, with higher roughnesses requiring
higher loads to bring the contact to the transitional condition where the real area of asperity contacts
completely occupies that of the nominal contact patch available.
In cases where sliding motion occurs by shear within the polymer near-surface rather than at the
contact interface, a peculiarity observed in mechanical characterization of polymers that may addi-
tionally be given consideration in adhesive friction modeling is strengths that increase with hydrostatic
pressure (Briscoe and Tabor 1977). The effect in this case would be a shear stress that must be over-
come in the polymer of s = so + αp. In a case of real areas of contact at asperity tips, the global friction
coefficient would be the same as that at the average asperity contact, in other words, 𝜇. = s/p = so/p + α,
where p is either the flow pressure in cases of plastic asperity contact or in the case of elastic asperity
contact a pressure that remains constant despite increasing global normal load W, since as previously
described the average asperity contact remains approximately unchanged by increasing normal load.
It is interesting to note that polymers such as PTFE and high-density polyethylene (HDPE) having
the lowest values of the pressure coefficient (α ≤ 0.1) are those of smooth linear molecular profile
well-known to display low coefficients of friction, while those having higher values (α > 0.35) such as
PMMA, polyvinyl chloride (PVC), and polystyrene (PS) are those of less smooth molecular profile
having large atoms or side groups protruding from their backbone chain and correspondingly much
higher friction coefficients.
The connection of adhesive friction behavior to molecular structure may result from the ease with
which polymer chains may slide by one another in the shear process, and in turn it may also affect the
transfer behavior of polymers (Pooley and Tabor 1972). In the case of linear polymers, adhesion of
asperity tips with the counterface may lead to the drawing of fibrils between the polymer asperity and
Friction and Wear of Polymer Materials 34-9

the counterface as they move away from one another, with the extension of fibrils occurring by the slip
of smooth molecules by one another as they become aligned along the fibril. Eventual fibril rupture
leaves thin (nanoscale) transfer film of molecular orientation along the sliding direction atop the coun-
terface, as well as similarly oriented “running films” over the polymer surface, which for such “special”
polymers of smooth molecular profile will maintain reduced friction in subsequent sliding passes. In
contrast, “normal” polymers of rougher molecular profile thus lacking the ease of molecular drawing
instead form thick (microscale) lumpy transfer films and continue higher friction behavior.
In the case of polymers instead above their glass transition temperature and thus in a rubbery state,
they often display very high coefficients of friction (μ > 1) in unlubricated sliding. Such high friction
may occur even in contact configurations where deformation-based friction should be minimized, and
furthermore viscoelastic friction behavior thought to be limited to such deformation configurations
may still be observed. Rather than occurring by steady shear occurring throughout the entire contact
region, Schallamach (1971) instead observed that seemingly continuous sliding motion in such cases
was instead occurring by discrete waves of detachment passing through the contact along the interface,
achieving displacement not by sliding but instead in a manner similar to that by which an entire rug
may be given a displacement by buckling it away from the floor at one edge then transmitting this buck-
led discontinuity through the rug to its opposite edge. Of course, the detachments passing through the
contact are waves of deformation as the polymer buckles away from then reattaches to the counterface,
thus resulting in the viscoelastic characteristics of friction behavior otherwise unexpected in such con-
tact configurations.

34.3  Wear
The preceding description of frictional behavior twice provided examples prefacing this accompanying
description of wear behavior of polymers: investigation of deformation-based friction by sliding conical
indenters of varying included semi-angle caused greater tearing damage of the polymer counterface at
increased cone sharpness; in those cases of adhesive contact where the strength of the interface exceeds
that of the polymer the shear failure within the polymer’s near-surface such that externally applied fric-
tion force to maintain sliding motion also results in detachment of polymer that is transferred to the
counterface. Correspondingly, this overview of polymer wear will discuss deformation-related mecha-
nisms such as abrasive wear, as well as adhesive wear. Chemical wear mechanisms such as oxidative
wear that may be operative in other material systems are typically of lesser concern for polymers, and
the mechanism of erosive wear of materials due to impact rather than sliding will instead be overviewed
in Chapter 11.

34.3.1  Abrasive Wear


In addition to the product of normal load and sliding distance, basic models of the contact of a rigid
conical indenter abrading a counterface further describe the wear volume to be proportional to tan(θ),
the slope of the cone surface (in other words, θ is the complement of the cone’s included semi-angle α
from the previous description of deformation-based friction). Such models are based upon the indenter
sinking a sufficient depth into the counterface to provide an adequate top projection area of the contact
over which pressure may act to balance the normal load while translating along the surface. Material in
line with the frontal projection of the contact may be ploughed away as loose debris upon sliding, not
unlike chip production in a machining operation using a tool of negative rake.
When extended to multi-asperity surfaces abrading the counterface, presuming upon preparation
by roughening that heights of asperities increase more rapidly than their breadths, Lancaster (1972)
demonstrated the expected increase of wear rate with increasing roughness of the abrasive surface for a
broad variety of polymers (Figure 34.5). While it has been reasoned that the effective slope of an asper-
ity, and thus the abrasive wear rate, may increase with the square root of roughness (center line average),
34-10 Wear Materials

1 + 2
10–6

+
3

+ 4
–7
10 + 5

Wear rate cm3/cm. kg + 6 +


+
+
+

–8
10
+
7

10–9

10–10
0.025 0.075 0.25 0.75 2.5
Surface roughness. µm cla

FIGURE 34.5  Wear rate of various polymers during single traversals of sliding against steel surface of vari-
ous roughness (center line average CLA). 1: Polystyrene, 2: polymethylmethacrylate, 3: polyoxymethylene (acetal),
4: polypropylene, 5: polytetrafluoroethylene, 6: polyethylene, and 7: polyamide (nylon) 6/6. (From Lancaster, J.K.,
Friction and wear, in Polymer Science, Jenkins, A.D., Ed., North-Holland Publishing, Amsterdam, the Netherlands,
p. 959, 1972. With permission.)

only PTFE and polyethylene approximated such a dependence with all other polymers investigated
displaying increases of wear rate with increasing counterface roughness that were more rapid.
While each of the polymers displayed a wear rate that increased with counterface roughness, at the
higher values of roughness where abrasion would be the predominating mechanism, it may be further
expected that relative rank-ordering of the polymers by wear rate should be influenced by their tough-
ness, since by the ploughing removal described earlier material in the path of the asperity must be
brought to breaking failure. At the highest value of roughness (1.2 μm CLA) in the study portrayed in
Figure 34.5, Lancaster (1972) in fact observed abrasive wear rates of these polymers to be nearly linearly
proportional to 1/(Sεb), where S and εb are the breaking stress and the elongation at break, respectively.
If the asperities are less sharp and rather round at their tips, then deformations will be limited and the
asperity ploughing mechanism inoperative. Passage of each asperity will instead lead to energy dissipa-
tion and damage in the polymer subsurface, with accumulation leading to eventual debris liberation by
a fatigue wear mechanism. As reviewed in Lancaster (1972), a model surface commonly used to produce
fatigue wear experimentally is metal gauze, where the wire bends at their points of intersection serve as
mildly rounded asperities. Not unlike prediction of failure of rolling element bearings by surface fatigue,
the number n of passing asperities cyclically imposing subsurface stress level σ that leads to eventual
particle detachment may be described as n = (σ 0/σ)b, where b is an exponent of positive value. In contrast
to this implication that all reduced stress levels would simply require a correspondingly greater numbers
of cycles to the eventual particle liberation, it has been suggested instead that reducing surface roughness
below a threshold level may allow attainment of sufficiently low asperity stresses that the fatigue wear
mechanism no longer operates (Briscoe and Tabor 1977), similar to the concept of an endurance limit.
Friction and Wear of Polymer Materials 34-11

34.3.2  Adhesive Wear


Once mating surfaces have been made sufficiently smooth to deactivate all other wear mechanisms such
as ploughing abrasion and fatigue, adhesive wear is considered to be the one mechanism to remain. An
asperity contact-based model of adhesive wear, as by other mechanisms such as abrasion, predicts wear
volume V to increase in linear proportion to the product of normal load W and sliding distance Ls. The
wear performance of a polymer is typically characterized in tests measuring wear volume V produced
by an imposed WLs product, and from their quotient quantifying the dimensional wear coefficient k = V/
(WLs) expression of wear rate. In a nominal contact area over which pressure is uniform, this dimensional
quantity equivalently represents the depth of wear per unit product of pressure and sliding distance.
In some presentations of wear performance, a material’s dimensional wear coefficient may further-
more be multiplied by its hardness H for nondimensionalization, and specifically if multiplied by three
times hardness the resulting nondimensional wear coefficient K = 3Hk may be reasoned to have addi-
tional physical significance. Specifically, if at any asperity contact there is the potential upon subse-
quent relative motion of removing from the wearing body an adhered hemisphere of material of radius
corresponding to that of the circular contact spot, then K represents the probability of such particle
detachment in any single asperity interaction. As with other material classes, such nondimensional
wear coefficients for polymers are multiple orders of magnitude less than unity, implying that on aver-
age 1/K asperity interactions are required to detach the incipient particle. Though the mating surface
has been smoothened, such an adhesive wear mechanism seems much like another form of fatigue
wear, whose number of cycles to particle detachment n may be likened to 1/K. If returning the nondi-
mensional wear coefficient into an expression for wear volume, it can be observed that in addition to the
product of normal load and sliding distance the amount of wear is expected to be inversely proportional
to hardness, but relation to other polymer material properties is unclear. For example, as opposed to
the inverse relation demonstrated between abrasive wear and the product Sεb, against smoother mat-
ing surfaces, little correlation is observed between a polymer’s adhesive wear and Sεb (Lancaster 1972).
The adhesive wear mechanism as described earlier implies the transfer of adhered polymer to its
mating counterface, with the sliding system gradually transforming its interface from polymer/bare
counterface to polymer/polymer transfer, going through a corresponding transient “run-in” transfor-
mation of wear rate and friction coefficient until a steady state of transfer is attained. This transfer
modifies the topography as well as the chemistry of the counterface surface, and as the strength of the
polymer/polymer transfer interface drops to below that of the polymer the continued relative motion
may be more fully accommodated by interfacial sliding rather than shear failure within the bulk poly-
mer. In such cases where transfer does not continue to build up atop previous transfer, the wear process
is momentarily halted until previously deposited transfer is eventually detached, at which point the
bare counterface is again exposed driving additional transfer. The frequency of this cyclic transfer
detachment and redeposition process, and in turn this adhesion-based transfer wear rate, is thought
to be rate-limited by the detachment step and thus controlled by the adhesive strength of the polymer
transfer to the counterface (e.g., Briscoe et al. 1974; Briscoe et al. 1977). As such, it has become frequent
practice to add to polymers various filler particles thought to catalyze chemical interaction of the trans-
fer with the counterface in an effort to reduce the transfer wear rate. Chapter 37 will expand on this
topic of composites.
In cases of thick lumpy transfer which also modifies counterface topography by roughening it, the rate
of detachment may be more rapid as the transfer is a protrusion which will experience interference with
the polymer during subsequent sliding. Furthermore, if the adhesion of the lumpy transfer to the coun-
terface is strong and its residence time until eventual removal long, in the interim it may activate addi-
tional ploughing abrasive wear of the polymer. Thus, transfer may not necessarily always be considered as
beneficial. By performing sliding tests against relatively smooth counterfaces of 0.1 μm CLA roughness,
Lancaster (1972) assessed the benefit of transfer to various polymers by quantifying the ratio of wear rate
repeatedly sliding over the same counterface regions and thus any transfer developed there, to the wear
34-12 Wear Materials

rate when instead always sliding over new regions of counterface without transfer in single traversals. An
important mechanical property found to correlate with this wear rate ratio is ductility, with polymers
capable of at least 20% elongation at break experiencing a beneficial effect of transfer as this wear ratio
was less than unity, while those polymers of elongation <20% generally had ratios of wear rates with and
without transfer greater than unity. The ductile polymers are more likely to deform under continued slid-
ing into smoother films of transfer that will reduce wear rate relative to that occurring in their absence.
As previously discussed relative to friction, “special” polymers of smooth molecular profile such as
PTFE draw directly into oriented fibrils forming smooth thin transfer films upon combined adhesive
contact and relative motion with the counterface. As such drawing is viscous, increased sliding speed
and resultant strain rate, or decreased temperature, increase the forces associated with this deforma-
tion process. Under a critical condition the viscous force associated with fibril drawing will instead
pull from the polymer larger particles, transitioning from smooth film to lumpy transfer behavior and
correspondingly higher wear rates (Pooley and Tabor 1972; Tanaka et al. 1973; Blanchet and Kennedy
1992). In the case of PTFE, irradiation has been shown to resist these higher wear rates at more severe
sliding conditions, presumably as a result of a cross-linking of the molecules into a network that would
resist the deformation processes of such a transfer wear mechanism (Blanchet and Peng 1996). In the
case of ultrahigh molecular weight polyethylene (UHMWPE) used for hip and knee joint replacement
bearings, it has become common practice to irradiate the polymer under conditions that promote heavy
cross-linking, as it has been shown that its adhesive wear rate will decrease linearly with decreasing
distance along chains between cross-links (Muratoglu et al. 1999).
In closing, as adhesive contributions become predominant with smoother counterfaces while abra-
sive contributions predominate at higher counterface roughness, it has been noted for several polymers
that overall wear rate may experience a minimum at some intermediate roughness (Tanaka and Nagai
1985; Santner and Czichos 1989). Likewise, friction may become minimum at a similarly intermediate
roughness, with friction coefficients increasing again as counterfaces become too smooth and adhesive
friction contributions enhanced.

34.4  Pressure × Velocity Limit


As has been indicated in several instances, the tribological behavior of polymers can depend greatly on
temperature. Thus in sliding tests used to characterize the effects of any pertinent variables, including
normal load and sliding speed, and even temperature, efforts are generally made to conduct such tests
at low values of contact pressure and sliding speed, since their product when further multiplied by the
friction coefficient represents the frictional power per unit area being dissipated at the sliding interface.
If this pv product becomes too large, frictional heating will cause the temperature at the sliding contact
to increase above that prescribed in the background environment, thus becoming an uncontrolled vari-
able confounding the effects of those others intended for study.
Maintaining low pv conditions may be impractical in actual applications, and significantly increased
temperatures due to frictional heating must be anticipated. As polymers generally possess poor ther-
mal conductivity and thus have little ability to pull frictional heat away from sliding interface, they are
generally paired with counterface materials of higher thermal conductivity to provide the heat sink.
Temperatures will rise to a level at the surface where its gradients and resultant heat flux into the coun-
terface will balance the frictional power being dissipated in the sliding interface. As the pv level of the
contact is increased, a softening temperature and thereafter the polymer’s melt point will eventually be
reached. At such a “pv limit” a ceiling temperature has been reached on the interfacial temperature, as
with any further severity of sliding the polymer will experience phase change rather than any further
increase in temperature, and therefore, the full capacity of the counterface to conduct frictional heat
away from this sliding contact has been reached. Any additional pv and corresponding frictional power
input will instead be partitioned into the latent heat of the polymer’s phase change, with a melt wear
mechanism activated where molten polymer is ejected from the sliding contact as wear volume. Under
Friction and Wear of Polymer Materials 34-13

specific contact geometries involving large mutual overlap of the polymer and counterface surface, it is
possible to achieve a thermal friction control condition (Ettles and Hardie 1988), where molten poly-
mer is not easily ejected but instead trapped within the contact, serving as a fluid film bringing friction
coefficient down to a level just sufficient to maintain the polymer melt temperature without producing
further melt volumes beyond that needed to provide the lubricating effect.

34.5  Tribological Properties of Polymers


Instead of being intrinsic properties of polymers, as described earlier, quantities such as friction coef-
ficient or wear coefficient are strongly system attributes that may vary greatly with speed, temperature,
contact pressure, counterface topography, and so forth. Likewise, pv limit should also be considered
as a system attributed rather than a polymer property. The ability of the polymer’s mating counterface
to pull away from the interface frictional heat associated with a given pv condition will also obviously
depend on the counterface thermal properties and background environmental temperature, as well as
contact geometry and Peclet speed condition effects beyond the scope of this presentation, not to men-
tion the friction coefficient itself required to convert pv into a frictional power input and the dependen-
cies of friction coefficient on system attributes previously mentioned. For the reader seeking tabulated
tribological “properties” of various polymers, such compilations do exist (Jamison 1994; Blanchet 1997)
though caution should be exercised with equal attention paid to the references from which the values
were drawn and the experimental conditions to which they are thus specific. From tests run against
steel counterfaces of practically smooth 0.15 μm CLA surface roughness at common speed and load
conditions, Lancaster (1968) provides an introductory survey of the friction coefficient and wear rate of
a broad set of 13 different polymers (Figure 34.6). Also depicted is the prospect of improvement of this

0.8
Coefficient of friction

0.6

0.4

0.2
0
1 2 3 4 5 6 7 8 9 10 11 12 13
(a)
10–7

10–8
Wear rate (cm3 cm–1 kg–1)

10–9

10–10

–11
(b) 10 1 2 3 4 5 6 7 8 9 10 11 12 13

FIGURE 34.6  (a) Coefficient of friction and (b) wear rate sliding at a speed of 0.54 m/s against steel of 0.15 μm
CLA surface roughness of various polymers (dashed) and their composites when reinforced with carbon fibers.
1: Friedel–Crafts Mark2 polymer, 2: polyvinylchloride, 3: polytetrafluoroethylene, 4: polypropylene, 5: epoxy,
6: polymethylmethacrylate, 7: polyester, 8: polycarbonate, 9: phenolic, 10: polyamide (nylon) 6/6, 11: polyethyl-
ene, 12: polyoxymethylene copolymer (acetal M90), and 13: polyimide. (From Lancaster, J.K., Br. J. Appl. Phys.,
1, 549, 1968. With permission.)
34-14 Wear Materials

friction and wear performance for each polymer upon inclusion of carbon fiber reinforcement, serving
as an introduction to a subsequent chapter on composites (Chapter 37).

References
Archard, J.F., Elastic deformation and the laws of friction, Proc. R. Soc. A, 243, 190, 1957.
Blanchet, T.A., Friction, wear and PV limits of polymers and their composites, in Tribology Data Handbook,
Booser, E.R., Ed., CRC Press, Boca Raton, FL, 1997, p. 547.
Blanchet, T.A. and Kennedy, F.E., Sliding wear mechanism of polytetrafluoroethylene (PTFE) and PTFE
composites, Wear, 153, 229, 1992.
Blanchet, T.A. and Peng, Y.L., Wear-resistant polytetrafluoroethylene via electron irradiation, Lubr. Eng.,
52, 489, 1996.
Briscoe, B.J., Pogosian, A.K., and Tabor, D., The friction and wear of high density polythene: The action of
lead oxide and copper oxide filler, Wear, 27, 19, 1974.
Briscoe, B.J., Steward, M.D. and Groszek, A.J., The effect of carbon aspect ratio on the friction and wear
of PTFE, Wear, 42, 99, 1977.
Briscoe, B.J. and Tabor, D., Friction and wear of polymers, in Polymer Surfaces, Clark, D.T. and Feast, W.J.,
Eds., John Wiley & Sons, New York, 1977, p. 1.
Ettles, C.M. and Hardie, C.E., The friction of some polymers and elastomers at high values of
pressure × velocity, ASME J. Tribol., 110, 678, 1988.
Flom, D.G., Rolling friction of polymeric materials. I. Elastomers, J. Appl. Phys., 31, 306, 1960.
Greenwood, J.A. and Tabor, D., The friction of hard sliders on lubricated rubber: The importance of defor-
mation losses, Proc. Phys. Soc., 71, 989, 1958.
Grosch, K.A., The relation between the friction and visco-elastic properties of rubber, Proc. R. Soc. A, 274,
21, 1963.
Jamison, W.E., Plastics and plastic matrix materials, in Handbook of Lubrication and Tribology Volume III,
Booser, E.R., Ed., CRC Press, Boca Raton, FL, 1994, p. 121.
Lancaster, J.K., The effect of carbon fibre reinforcement on the friction and wear of polymers, Br. J. Appl.
Phys., 1, 549, 1968.
Lancaster, J.K., Friction and wear, in Polymer Science, Jenkins, A.D., Ed., North-Holland Publishing,
Amsterdam, the Netherlands, 1972, p. 959.
Ludema, K.C. and Tabor, D., The friction and visco-elastic properties of polymeric solids, Wear, 9, 329,
1966.
Muratoglu, O.K., Bragdon, C.R., O’Connor, D.O., Jasty, M., and Harris, W.H., The mechanism of marked
improvement of wear in highly crosslinked UHMWPEs, in Proceedings of the 45th Annual Meeting
Orthopaedic Research Society, February, Anaheim, CA, Vol. 817, 1999.
Pooley, C.M. and Tabor, D., Friction and molecular structure: The behavior of some thermoplastics, Proc.
R. Soc. A, 329, 251, 1972.
Santner, E. and Czichos, H., Tribology of polymers, Tribol. Int., 22, 103, 1989.
Schallamach, A., How does rubber slide? Wear, 17, 301, 1971.
Tabor, D., Friction, adhesion and boundary lubrication of polymers, in Advances in Polymer Friction and
Wear, Lee, L.-H., Ed., Plenum Press, New York, 1974, p. 5.
Tanaka, K. and Nagai, T., effect of counterface roughness on the friction and wear of polytetrafluoroethyl-
ene and polycthylene, in Wear of Materials 1985, Ludema, K.C., Ed., ASME, New York, 1985, p. 397.
Tanaka, K., Uchiyama, Y., and Toyooka, S., The mechanism of wear of polytetrafluoroethylene, Wear, 23,
153, 1973.
35
Metals
35.1 History and Introduction............................................................... 35-1
35.2 Definition of Metallic-Based Friction and Wear........................ 35-2
Friction of Metals  •  Wear of Metals
35.3 Metals and Alloys............................................................................35-4
Metallurgical Compatibility, Crystal Structure, and Stacking Fault
Energy  •  Nature of Surfaces and Subsurfaces  •  Surface and
Subsurface Microstructural Evolution during Wear  •  Role of Metal
Hardness and Other Properties  •  High Temperature Wear
35.4 Shape Memory Alloys...................................................................35-12
35.5 Nanocrystalline Metals and Alloys............................................ 35-13
Thomas W. Scharf Acknowledgments..................................................................................... 35-14
University of North Texas References................................................................................................... 35-14

35.1  History and Introduction


The first person to discuss a wear problem, that of material removal during polishing, was none other
than Newton [1]. His conclusion, that polishing is simply abrasion on a reduced dimensional scale,
was likely the first attempt to examine material wear. Not until the twentieth century did wear of
materials become a science. Holm [2], Samuels [3], Archard [4], Archard and Hirst [5], Lancaster [6],
Khruschov [7], Bowden and Tabor [8], Finnie [9], Rabinowicz [10], Kragelskii [11], Waterhouse [12],
Quinn [13], Ludema [14], Buckley [15], Childs [16], Rigney [17], Suh [18], Kuhlmann-Wilsdorf [19],
Czichos [20], Glaeser [21], Blau [22], Zum Gahr [23], Biswas [24], and a host of others have pioneered
the science and engineering behind metallic-based wear. Of all the material classes, the majority of
research in wear during the last 60 or so years has been in metallic systems, which is not surprising
since most engineering materials in moving mechanical assemblies are fabricated from metal-based
materials.
This chapter will be mostly based on sliding adhesive and abrasive wear, as opposed to rolling and
impact, since the vast majority of metallic wear research is sliding-based due to its relevance in many
engineering applications, testing equipment that is easy to use, cheap, and readily available. To quote
Blau [25] from a 1997 review on the wear of metals,

sliding wear is arguably more mechanistically complex than certain other forms of wear because
it involves not only the cutting and plowing included in abrasive wear, but also the adhesion of
asperities, third bodies (debris), subsurface crack initiation and growth (i.e., contact fatigue pro-
cesses), the transfer of material to and from the mating surfaces, subtle changes in surface rough-
ness during running-in, tribochemical film formation, and other processes.

35-1
35-2 Wear Materials

35.2  Definition of Metallic-Based Friction and Wear


35.2.1  Friction of Metals
Friction and wear mitigation, often interrelated, are typically accomplished by introducing a shear
accommodating layer (e.g., a thin film of liquid) between contacting surfaces. When the operating con-
ditions are beyond the liquid realm, attention turns to solids (bulk or coatings). According to the clas-
sical theory [8], friction force, F, is a product of the contact area and the shear strength of the solid
material, A · τ. Thus, the friction coefficient, μ, can be expressed by

F A ⋅t t t
m= = = = o +a (35.1)
L L PH PH

where
L is the normal force (load)
PH is the mean Hertz pressure
τo is the interfacial shear strength, a “velocity accommodation parameter” which is a property of the
interface
α represents the pressure dependence of the shear strength

The constant “α” is the lowest attainable friction coefficient for a given friction couple. In principle, a
hard material with a soft skin (e.g., soft metal or oxide film on the metal) ought to provide low friction
coefficient by reducing το and increasing PH (low A). Bowden and Tabor validated their concept by dem-
onstrating that indium metal when applied as a coating on a much harder steel substrate can indeed
reduce friction [8,26]. It is also interesting to note from their study that while the friction coefficients
for the unlubricated steel and for the steel lubricated with mineral oil obeyed Amonton’s first law of
friction (μ is independent of L), the friction coefficient of indium film on steel substrate decreased con-
siderably with increasing normal load (see explanation in Chapter 39). Here, the increased deformation
of the underlying steel resulting from an increase in normal load produced only a small increase in the
real contact area. Thus, there is only a slight increase in frictional force as the load is increased, which
results in a corresponding decrease in friction coefficient according to Equation 35.1. For thin layers and
soft metal coatings, the pressure is primarily supported by the substrate and increasing the substrate
modulus and hardness will decrease the contact area for a given normal load. Thus, the ideal scenario
for achieving low friction according to the Bowden and Tabor concept is to have an elastically stiff and
hard substrate support the normal load and keep the contact area small, while the surface provides
shear accommodation and reduces junction strength, until the substrate begins to yield and plasti-
cally deform. These are necessary but not totally sufficient conditions for most low friction metal-based
systems. In addition, compliant transfer films (third bodies) that adhere to the counterface or transfer
layers that adhere on the wear track, often tribochemically, can lower το further and provide long wear
life by preventing native metal–metal contact while accommodating interfacial shear [27]. For more
details, see Chapter 39.

35.2.2  Wear of Metals


For quantitative analysis of wear rates (factors), various equations have been developed, but the major-
ity have the general form based on the first relationship derived by Holm [2] and later by Archard for
sliding [4]:
Metals 35-3

⎛ k⎞
V = ⎜ ⎟ Ld (35.2)
⎝H⎠

where
V is wear volume loss
k is the wear coefficient
H is the hardness of the softer metal
L is the normal load
d is the total sliding distance

Archard’s law was initially established for adhesive wear involving shearing or breaking of adhesive
functions between contacting surfaces, but also applies to abrasive wear involving scratching by asperi-
ties or entrapped debris. In both modes, V is proportional to L and d and inversely proportional to H,
that is, the harder the metal, the lower the wear volume loss. With recent developments in nanocrys-
talline (nc) metals and alloys, hardness has been shown to increase due to the classical dislocation-
based Hall–Petch relationship that predicts increasing hardness and strength with decreasing grain
size. However, grain sizes in the critical range of ∼5–20 nm for hexagonal close packed (HCP) Mg [28],
face centered cubic (FCC) Cu [29], FCC Ni [30], and Ni–W alloy [31] have been shown as a threshold in
which this relationship breaks down, sometimes referred to as the inverse Hall–Petch effect when a peak
in hardness is observed. Instead of dislocation-dominated processes, there is a shift to grain boundary-
dominated processes [32,33]. As a result, the classical Archard wear Equation 35.2 breaks down for nc
metals and alloys, since hardening (or softening) in the wear tracks leads to deviations from the Archard
relation, which linearly relates hardness and wear resistance [32]. In addition, there are some more tra-
ditional microcrystalline (mc) grain size metal systems that do not obey Equation 35.2, such as Cu–Be/
steel [34] tribocouple, where the Cu–Be adhesive wear to the steel ball is inconsistent during sliding. The
amount transferred is independent of both the load on the tribocouple and the sliding distance.
In addition to hardness, as the only material property in Equation 35.2, other parameters such as the
H/E ratio and fracture toughness can be critical in metallic wear, since a material needs to be hard to
combat adhesive and abrasive wear while maintaining high strain tolerance and toughness to prevent
crack propagation. Besides plastic deformation, fracture and fatigue/delamination wear are often preva-
lent in the wear of metals and alloys. General metallic-based wear processes, appearances, and wear
rates/coefficients of these and other wear types are listed in Table 35.1 and References 35 and 36.
In Equation 35.2, k is a nondimensional wear coefficient, whereas a dimensional (specific) wear rate
(k), in units of mm3/Nm, can also be expressed as

V
k= (35.3)
Ld

The wear coefficient/rate is not always constant for the same material combination, since it depends on
the operating parameters (load, velocity, temperature, etc.) and the functioning wear regime/mode.
However, Equations 35.2 and 35.3 have been used as a general engineering guideline for the use of met-
als and alloys as wear-resistant materials. Figure 35.1, adapted from [37], and Table 35.1 present general
k values for several types of metal wear modes.
As will be seen from the following classes of metallic-based materials, environment, load, sliding
speed, etc., play a significant role in determining the tribological performance of metallic materials that
are designed to mitigate friction and wear. For instance, metallic materials that give extremely low fric-
tion and long wear life in one environment can fail to do so in a different environment.
35-4 Wear Materials

TABLE 35.1  Metallic Wear Types, Processes, Appearances, and Some Typical Wear Coefficient (k) Values for
Metal-on-Metal Sliding Systems
Wear Type Processes Appearances k
Adhesive Metal transfer to usually the harder Seizure, rough surfaces, adhesive 10−3 to 10−7
material; transfer film or particles debris pullout Severe galling 10−3
generally work harden; subsurface may Moderate wear 10−5
undergo plastic deformation Burnishing 10−7
Abrasive Asperities of a rough, hard surface/particle Scratching, plowing, wedge 10−2 to 10−5
slide on a softer surface and cause plastic formation, cutting Severe abrasion 10−2
deformation (ductile) or fracture (brittle) Polishing 10−5
Fatigue Subsurface or surface cracks, formation of Sharp edges around the spall or pit 10−4 to 10−8
pits in the surface (pitting) Brittle fracture 10−4
Surface fatigue 10−8
Fretting Involves all the aforementioned wear Roughening, formation of oxide
processes; powdery wear debris ridges
Impact Related to surface fatigue, small wear Fragmentation and pitting
particles or spalls
Erosion Abrasives in the fluid Waves, troughs
Corrosive Metal corrosion products in the presence of Dark bottomed 10−2 to 10−5
(oxidative) liquid/corrosive media. Corrosive wear in pits, with surface-originated Severe corrosion 10−2
air is called oxidative wear spalling Solid lubricant 10−5
Electrical Arcing causes craters, and any sliding shears Craters (holes) and sheared lips
(arc) or fractures the lips, leading to abrasion,
corrosion, and surface fatigue

Identical
Poor Good Excellent
metals Unlubed lube lube lube
Compatible Poor Good Excellent
Unlubed lube lube lube
metals
Adhesive
Partly compatible Unlubed Poor Good Excellent
wear lube lube lube
metals
Incompatible metals Unlubed Poor Good Excellent
lube lube lube
Nonmetal on metal or nonmetal Unlubed Lubed

Abrasive
2–body 3–body (high) (low)
wear
Corrosive Excessive Mild
wear
Fretting Unlubed Lubed

10–2 10–3 10–4 10–5 10–6 10–7 10–8


Wear coefficient (k)

FIGURE 35.1  Range of wear coefficients (k) for four types of metal wear. See Figure 35.2 and text for explanation
of different types of circles.

35.3  Metals and Alloys


35.3.1 Metallurgical Compatibility, Crystal Structure,
and Stacking Fault Energy
Tabor [38] discussed the formation of metallic junctions in relation to friction and wear. The interac-
tion energy between two flat parallel surfaces of clean metal as they are brought into atomic contact
shows that the van der Waals interaction operates for separations greater than 0.3 nm, while for
Metals 35-5

separations <0.2 nm, the energy is swamped by the formation of the metallic bond. Forming metallic
junctions in tribological contacts is clearly an undesirable phenomenon from the performance and
reliability perspective. The rupture of junctions usually leads to metal transfer from one surface to
the other. This transfer typically does not produce a wear particle immediately, but portions of the
transferred metal are removed in subsequent passes. The bonding between similar and dissimilar
metals has often been discussed in terms of mutual solubility [10]. The suggestion that mutually
soluble pairs show strong adhesive bonding and hence high friction and wear, while insoluble pairs
show weak adhesive bonding and hence low friction and wear, is useful as a general guideline, but
does not always hold rigorously, for example, incompatible pairs Ag–Fe, Pb–Fe, and Ag–Ni are able to
adhere strongly to each other. This solubility-metal adhesion approach has been used by Rabinowicz
[39] to develop a generalized compatibility chart, or map, shown in Figure 35.2. The map, based on
equilibrium binary phase diagrams, shows which metals can safely slide against one another and

W Mo Cr Co Ni Fe Nb Pt Zr Ti Cu Au Ag Al Zn Mg Cd Sn Pb

In

Pb

Sn

Cd

Mg

Zn

Al

Ag

Au

Cu

Ti

Zr

Pt

Nb

Fe
Low adhesion No liquid solubility; solid solubility <0.1%
Ni
Solid solubility >0.1%, but limited liquid solubility
Co
Full liquid solubility and soild solubility between 0.1 and 1%
Cr
High adhesion Full liquid solubility and soild solubility between >1%
Mo

FIGURE 35.2  Compatibility chart developed by Rabinowicz for selected metal combinations derived from binary
equilibrium diagrams. Chart indicates the degree of expected adhesion (and thus friction and wear) between the
various metal combinations. (Reprinted from Rabinowicz, E., ASLE Trans., 14, 198, 1971. With permission.)
35-6 Wear Materials

which metal couples should be avoided. Like metal pairs (data not shown) are assumed metallurgi-
cally compatible by definition, that is, show full solubility >1%. Thus, adhesive friction and wear
should decrease as one progresses from couples of the same material and those that form solid solu-
tions to insoluble couples and to metal–nonmetal pairs. Figure 35.1 shows this effect for the adhesive
wear coefficient (k). Furthermore, Ohmae and Rabinowicz [40] also looked at the sliding wear behav-
ior of nine noble metals that form little or no oxide in air. Noble metals and their alloys are often
used in sliding electrical contacts for applications such as slip rings. They determined that the lowest
wear coefficients were exhibited by hexagonal metals (Rh, Ru, and Os), incompatible metal pairs, or
both, which is in agreement with earlier studies by Buckley and Johnson [41] and Rabinowicz [42],
respectively.
A crystal structure factor influencing the compatibility effects is the number of active, independent
slip systems. With many available slip systems, such as 12 primary ones with FCC and body centered
cubic (BCC) metals, there is a greater tendency for the transition to multiple slip compared to a limited
number of slip systems, as in the three (basal and prismatic) or six (pyramidal) primary ones for most
HCP metals. Therefore, in general, the greater number of slip systems leads to higher true area of contact
for FCC and BCC metals resulting in higher friction and wear compared to HCP metals. Furthermore,
the abrasive wear of cubic metals is about twice the rate of hexagonal metals, which was attributed to
the lower work hardening rate of the hexagonal metals [43]. Another crystal structure contribution
observed in HCP metals is their c/a lattice ratio. In general, friction and wear decrease with increasing
c/a ratio, since the shear force is controlled by this ratio. Figure 35.3a shows this trend of friction coef-
ficient (in vacuum) for some HCP metals with Be being an anomaly [41]. The higher c/a ratio, lower fric-
tion, and wear HCP metals also tend to exhibit predominately basal plane slip, whereas the lower ratio,
high friction, and wear HCP metals tend to exhibit non-basal slip, for example, prismatic and pyramidal
slip planes. With non-basal slip predominating, a greater number of slip systems are operative during
wear, which can give rise to an increase in true area of contact and to work hardening. This results in
resistance to shear in sliding contacts because there is multiple, intersecting slip plane dislocations,
hence high friction, in metals such as Ti, Zr, Hf, Er, Y, Gd, etc., as shown in Figure 35.3a. In contrast for
HCP metals that exhibit predominately basal plane slip parallel to the wear surface, for example, basal
slip along the (0001) glide plane in Cd, this interaction and dislocation coalescence on slip planes is not
readily attained and the dislocations can pass out of the surface [44]. Figure 35.3b shows the same trend
in wear coefficient (dashed line). However, two metals Os and Ru (solid line) show that all HCP metals
plotted exhibit relatively similar wear coefficients with just the group 4 metals (Ti, Zr, and Hf) giving
very high wear [40]. Ohmae and Rabinowicz [40] concluded that metals or alloys containing Rh or Ru
10–2

Hf
0.8 Y DY c Zr
Ti
Ti Mg
Coefficient of friction

10–3
Wear coefficient

Gd Er a
0.6 Zr
Tl
Hf
10–4

Sm
0.4 Be Os Co
Tl La
Co
Ru La
0.2 Nd
10–5

c Ce
a Re
0
1.56 1.57 1.58 1.59 1.60 1.61 1.62 1.63 1.58 1.60 1.62
(a) Lattice ratio, c/a (b) c/a ratio

FIGURE 35.3  (a) Friction coefficient as a function of c/a lattice ratio for various HCP metals in vacuum sliding
against 440C steel balls at 10 N and 2 m/s. (Reprinted from Buckley, D.H. and Johnson, R.L., ASLE Trans., 9, 121,
1966. With permission.) (b) Wear coefficient (k) as a function of HCP metal c/a lattice ratio for self-mating sliding
pairs in air. (Reprinted from Ohmae, N. and Rabinowicz, E., ASLE Trans., 23, 86, 1980. With permission.)
Metals 35-7

gave the optimum performance for sliding electrical contacts in comparison to Au-, Pt-, and Pd-based
metals and alloys.
The role of metal and alloy stacking fault energy (SFE) in relation to wear has also been extensively
studied [44–47]. As mentioned earlier for metals that exhibit multiple intersecting slip systems, there
exists localized hardening within the grain interiors. Thus, there is a tendency for an increase in the
cross-slip of dislocations, which in turn is related to an increase in SFE. Thus, at small strains (<0.1),
low SFE materials should preferentially exhibit softer surface regions relative to similar materials
having high SFE. In high SFE materials (strain >0.1), the multiple slip region extends to grain inte-
riors, dynamic recovery occurs, and a dislocation cell structure (typically 0.1–5 μm) forms with the
cells tending to be strong barriers resembling grain boundaries [17,44,48]. In contrast for low SFE
materials, these cells may be absent, or the boundaries may tend to be semipermeable to dislocations
resulting in softer surface regions. A number of studies have been conducted on the contribution of
SFE to wear. Hirth and Rigney [44] have discussed the wear data of Pavlov et al. for a number of FCC
metals sliding against an Au–Cu alloy, in which there was a clear trend of lower wear for lower SFE
materials (Al, Pt, Pd, Ni, Cu, Ag, Au, Au–80 wt% Cu, Ni–40 wt% Co, and Ni–60 wt% Co). In addi-
tion, similar trends have been shown for other alloys including Cu-7Al [49], cobalt-based Stellites
[50], nickel-based Tribaloy [51], and copper alloyed with Si, Sn, and Cr [52]. The latter authors find
an opposite effect for Cu–Zn alloys which they attribute to grain boundary segregation effects domi-
nating changes associated with changes in SFE. These changes in SFE are complicated in alloys with
varying amount of solute content, since there can exist concomitant changes in other properties, such
as hardness and flow stress. However, by and large, there still exists a general trend of lower wear with
lower SFE materials.

35.3.2  Nature of Surfaces and Subsurfaces


Since strong metallic adhesive bonds form only at interatomic distances, the presence of a chemisorbed
or physisorbed monolayer of gas or contaminant layer on the surface (Figure 35.4a) may reduce the
interaction to that of van der Waals forces. Thus, the presence of impurities or thin oxide metal films
can strongly affect friction and wear. Without these layers, or the presence of a liquid lubricant, junction
formation can result and is often attributed to local adhesive bonding or welding and is exacerbated by

W film

Surface texture (no surface is perfectly flat)


Physisorbed, e.g.,
hydrocarbon contamination
and chemisorbed, e.g., water
vapor, layer (0.3 – 3 nm)
Chemically reacted layer,
e.g., metal oxide (3–100 nm)
Deformed layer, e.g., severely
sheared or strained (work (110) (211)
hardened) (1 – 40 µm); gradient
in deformation going from the
top to bottom
Undeformed material 500 nm

(a) (b)

FIGURE 35.4  (a) Schematic (not to scale) of typical polished metal surface and subsurface layers. Surface texture
in the vertical axis is magnified. (b) Cross-sectional TEM image from 8119 case carburized steel tapered roller
bearing cone after a surface grinding finishing operation. See text for details. A protective W film is on the surface.
(Reprinted from Evans, R.D. et al., Tribol. Trans., 47, 430, 2004. With permission.)
35-8 Wear Materials

increases in interfacial flash temperatures caused by the sliding of rough surfaces, since at no size scale
is the surface atomically flat. Because of the small tip size of asperities and the high local loads, the
instantaneous flash temperature can be quite high. Junction formation due to solid-state bonding (adhe-
sion) is the same as discussed earlier for friction. Thus, sliding adhesive wear should decrease as one
progresses from couples of the same material and those that form solid solutions to insoluble couples
and to metal–nonmetal pairs.
In addition to a “rough” contacting surface, the metallic surface itself consists of several zones having
physical and chemical properties different from bulk, also shown schematically in Figure 35.4a. Indeed,
surfaces are extremely complicated because of their texture/topography, chemical reactivity, and their
composition and microstructure, which may be very different from those of the bulk (undeformed) solid.
Frictional contact can induce additional adhesive contacts and result in significant microstructural
changes in the subsurface regions, which will be discussed in the next section. In the case of metallic
contacts, the native oxides can be worn away during the initial run-in period resulting in an unwanted,
high wear metal–metal contact. Also, manufacturing and finishing operations, such as polishing and
grinding, can lead to changes in the materials’ microstructure and chemistry even before being tribo-
logically stressed. For example, Figure 35.4b shows a cross-sectional transmission electron microscopic
(TEM) image from 8119 case carburized steel tapered roller bearing cone (inner raceway surface) after
a surface finish grinding operation, that is, before in-service wear [53]. The arrow indicates direction of
grinding which caused near-surface grain size gradient and microstructural distortion (high disloca-
tion density) due to large deformation strains. The near-surface region (depth <500 nm) consisted of
BCC iron ferrite nc grains (∼20 nm size), which was confirmed by inset selected area electron diffrac-
tion patterns. Larger deformed iron ferrite grains (subgrains or dislocation cells) existed deeper into the
subsurface up to another 500 nm, that is, ∼1 μm total depth of plastic strain. This gradient in deformed
structure agrees with the schematic shown in Figure 35.4a; however, no iron oxide existed on the surface
likely due to the grinding process removal. Thus, polishing, grinding, and other surface finishing steps
in metallic-based materials can lead to decreased grain size even before tribologically stressed. Also, it
was speculated and assumed for some time that these processes result in an amorphous “Beilby layer”
in metals, but evidence, such as the one given earlier, points to an ultrafine-grained material [54]. The
aforementioned processes introduce internal stresses and result in strengthening and hardening due to
cold working/strain hardening. However, there likely exists a threshold value in grain size leading to
softening (change from dislocation-driven deformation to diffusion-driven or grain boundary sliding/
rotation deformations).

35.3.3  Surface and Subsurface Microstructural Evolution during Wear


As Figure 35.5a illustrates, when tribological stresses are introduced, the surface and subsurface struc-
ture becomes even more complex due to the potential formation of ultrafine grain (<10 nm size) lay-
ers/zones, mechanically mixed layers (MML), tribochemically formed films, etc., during both dry and
lubricated sliding. Just like friction and wear are systems properties, the complex layers and zones that
microstructurally and chemically evolve during wear depend on the testing parameters and material
types. Some of the early work on sliding of ductile metals/alloys and examining cross sections of worn
samples, strain gradients, and transfer layers was conducted by Lancaster and coworkers [55–57]. They
used microhardness measurements inside brass worn surfaces to determine the hardness gradient as
a function of subsurface depth. Higher hardness values were determined closer to the surface due to
dislocation-mediated strain hardening. Rigney and coworkers also extensively investigated microstruc-
tural evolution during wear processes and are responsible for such terminology as ultrafine grain struc-
ture, highly deformed layer (HDL), and MML [17,58–62]. Their pioneering metallographic evidence
shows very large plastic strains and strain gradients exist in ductile metals and alloys that are respon-
sible for increased dislocation density and deformation substructures adjacent to the sliding interface.
Metals 35-9

Tribochemical films; transfer, oxide,


and/or mechanically mixed layers
(MML) with counterface material or
Sliding environment (50 – 500 nm)
Ultrafine (2 – 20 nm) grains and/or
dynamic recrystallized layer (1 – 5 µm)
High strain deformation cold-
worked zone gradient: subgrains,
dislocation cell boundaries, and/or
plastic flow (slip lines, elongated
grains, twinning) (1 – 40 µm)
Undeformed material

(a)

1 µm
Sliding
Slid MML: ultrafine nc-
ing Protective C on surface Ultrafine nc-Ni
Cu grains mixed
grain stucture
with oxides of Cu
and Fe (from steel Plastic flow of
ring) columnar,
Subgrains elongated grains
Undeformed
Dislocation cell material
boundaries 2 µm
(elongated grains)
(b) (c)

FIGURE 35.5  (a) Schematic showing various, possible constituents in the surface and subsurface layers/zones
present in metallic-based materials subjected to friction and wear. (b) Cross-sectional TEM image of worn surface
of OFHC Cu against 440C steel ring. (Reprinted from Heilmann, P. et al., Acta Metall., 31, 1293, 1983. With permis-
sion.) (c) FIB-prepared cross-sectional ion-induced SEM image of worn NiMn at 1.7 GPa stress. (Reprinted from
Jungk, J.M. et al., Acta Mater., 56, 1956, 2008. With permission.)

With each repeated pass during sliding, the amount of strain injected into the near-surface region accu-
mulates to very high levels. Shear strains of 1100% and strain rates as high as 105/s have been estimated
for the outermost layer [17].
Figure 35.5b shows a classic example of a longitudinal TEM cross-sectional image after unlubricated
sliding on oxygen-free high-conductivity (OFHC) polycrystalline (∼62 μm grain size) copper in argon
at 67 N normal load, 10 mm/s speed for 12 m distance in 20% RH argon [58]. From the image there is a
gradual change from loosely formed dislocation cells, to smaller and better organized cell structures,
and then, at the largest strains, to subgrains (∼0.3 μm thick) as the surface is approached. On the sur-
face, there exists a transfer layer or MML consisting of predominately ultrafine nc (∼5 nm) Cu grains
intermixed with oxides of Cu and Fe (from steel counterface). This transfer often occurs in the initial
stages of the sliding process. The plastic deformation involved leads to a rotation of the crystalline lat-
tice which may continue for the duration of sliding. Misorientations between adjacent substructure
features (grain and cell boundaries) gradually increase as the surface is approached where the strain
accumulation levels are the highest. The total surface displacement was ∼30 μm and the maximum
shear strain was ∼11.4. The top layers shown in Figure 35.5a, which are sometimes just compacted/
compressed wear debris, typically are not as dense as the underlying layers. Eventually, loose wear
particles can be formed by delamination of the transfer layer or MML material. The wear particles
often have the same structure and composition as these surface layers. Due to this complex sequence of
events and nature of the subsurface evolution, a single wear rate value should not be expected during
35-10 Wear Materials

these wear processes. In addition, frictional heating can complicate these processes, although it is dif-
ficult to separate the effects of strain rate from those of local temperature rise, which can be very large
at high sliding speeds.
Figure 35.5c shows another example of subsurface microstructural evolution during sliding wear for
the case of electrodeposited columnar grain NiMn sample [63]. As in the preceding example of OFHC
Cu, a layer of equiaxed, ultrafine nanocrytalline grains exists at the surface. Although not discussed, it
is possible that this layer also contains metal oxide (NiO/MnO), and SiO2 due to tribochemical wear of
Si3N4 ball counterface thereby making it an MML. Below it, there is clear evidence of plastic flow of the
NiMn columnar grains in the direction of sliding. Electron back scatter diffraction (EBSD) of this cross
section revealed that there was a significant degree of grain rotation in the high shear regions, similar
to OFHC Cu. EBSD studies in dry sliding of Ni pins on bronze (Cu–15Ni–8Sn) alloys have also revealed
that these highly deformed nc layers exhibit a high degree of crystallographic texture due to sliding [64].
In contrast to the work of Wheeler and Buckley [65] that suggested sliding wear-induced texture was due
to recrystallization, this study, along with others, points to texturing as purely a mechanical deforma-
tion process, since these tests are performed at low homologous temperatures and at low enough sliding
velocity to limit high flash temperatures at contacting asperities, which would suppress thermal recrys-
tallization. However, in some cases there is clear evidence of dynamic recrystallization, that is, recrystal-
lization that occurs during the deformation process, in sliding metallic wear contacts [66]. Interestingly,
for sliding against polycrystalline Cu under similar conditions as the previous one except in air at 50%
relative humidity, Ives [67] and Ruff et al. [68] showed a zone of elongated, larger (∼1 μm) recrystallized
grains formed with annealing twins inside them. This ∼3 μm thick recrystallized layer was underneath
the ∼1 μm thick ultrafine grain Cu MML layer proving that there can be clear demarcation between
these layers although the ultrafine-grained structure could also be recrystallized.
Under lubricated conditions transfer layers can still occur but tend to be thinner since the lubricant
will reduce the extent of the high strain region and the severity of extrusive wear [21]. Glaeser and
coworkers [48,60] determined that sliding steel on Cu–Al, Cu–Ni, and Cu–Al 2O3 alloys in the presence
of a stearic acid–octadecane, dibasic acid ester (MIL-L-7808), and dimethyl silicone lubricants resulted
in a thin transfer layer/MML (3–30 nm sized metal particles) whose surface and subsurface structures
are very similar to Figure 35.5b and c. Glaeser [60] also examined boundary lubrication regime sliding
on self-mated steels (mild carbon and 304 stainless) in the presence of stearic acid and zinc dialkyldi-
thiophosphates (ZDDP). The level of strain in the near-surface region under the wear scar was deter-
mined by TEM to contain three distinct zones, similar to the zones in Figure 35.5a. ZDDP will lubricate
304 stainless steel, forming a brittle sintered structure of reacted particles. Childs has also reviewed
boundary-lubricated wear of steels [16]. Boundary lubricants could form soaps on steel surfaces and
extreme pressure additives could react to form chlorine, sulfur, or phosphorus-containing films. A
more recent study on boundary-lubricated steels was conducted by Evans et al. [53,69]. Cross-sectional
TEM studies of the previously discussed (Figure 35.5b) steel tapered roller bearing cone surfaces were
examined after rolling contact in mineral oil with sulfur- and phosphorus-containing gear oil addi-
tives. The antiwear (AW) surface layers (∼20–100 nm thick) on the base steel/cone raceway, separated
by a sharp interface (approximately <10 nm), contained crystalline and amorphous regions. The sub-
surface grain deformation is similar in appearance to Figure 35.5b with distortion in the direction
parallel to the surface. The main difference between the images is the appearance and nature of the
AW layer, which is amorphous and consisted of Fe, O, and P in iron oxide and phosphate–phosphite
configurations. Under a more severe rolling condition (boundary lubrication BL regime), the AW layer
in some areas is approximately two times thicker than the previous AW layer, which was attributed to
higher strain deformation to near-surface material. This thicker AW layer was composed of different
regions of Fe, O, P, C, Ca, and S. These examples prove that similar subsurface deformation structures
exist between unlubricated and lubricated metal alloy contacts outside of the AW tribochemical sur-
face layers that form due to lubricant additive interactions.
Metals 35-11

35.3.4  Role of Metal Hardness and Other Properties


Higher metal wear resistance (lower wear rate) is expected with increasing metal hardness. This
trend was first studied by Khruschov and Babichev [7,70] and later by many others including
Sundararajan [71] and Kato [72], who compared the relative abrasive wear resistance (RAWR) of
numerous annealed metals as a function of their bulk hardness. Khruschov and Babichevs’ results
on pure metals showed that the slope is approximately linear, that is, RAWR ∝ H0.95. They also
found that correlations of wear rate and initial hardness were good if materials with similar bond-
ing type and microstructures were compared. In contrast, the RAWR of quenched and tempered
steels has a much weaker dependence on hardness compared to pure metals [71]. From the data of
several authors [73–75], the dependencies are RAWR ∝ H0.2 at lower hardness levels, RAWR ∝ H0.5
at higher hardness levels, and RAWR ∝ H0.4 at all hardness levels. Thus, hardness alone fails in pre-
dicting abrasive wear resistance since it does not include the interactions between abrasive particles
and the wearing materials sufficiently, which determine the formation of wear debris [76]. With
this in mind, Zum Gahr [76] plotted the abrasive wear resistance as a function of hardness of the
wear debris for pure metals, single phase alloys, and ferritic and austenitic alloys. The plot shows
a good linear fit between the abrasive wear resistance of these materials and hardness of the wear
debris. Deviations at higher values were due to a change in wear mechanism from microplowing/
microcutting to microcracking.
Rigney also explored the role of hardness gradients in the aforementioned (Section 35.3.3) heteroge-
neous deformed substructures, which affect transitions in friction and wear [77]. The nc material, often
making up the adhesive transfer layer and MML, can be either harder or softer than the underlying
material, which is strongly influenced by the morphology/texture of the wear track and the nature of
the wear debris. Not surprisingly, it is difficult to predict the trend since both adhesion and mechani-
cal properties are interrelated and contribute to the transfer and mixing processes. Local hardness on
the heterogeneous worn surface can vary due to the amount of deformation, strain hardening, and
transfer/mixing processes. In the majority of metallic-based materials that primarily work harden dur-
ing sliding, high friction and wear results, with increased hardness in the worn surface. For example,
if 304 stainless steel, which has a metastable austenitic structure, slides against a tool steel, such as
M2, the friction and wear are normal at first, but then strain-induced martensite is produced. This
transformation results in an extremely hard surface material (MML) that abrades the tool steel [78].
However, in other systems, the MML, despite its very fine grain size, can be softer than the adjacent
work hardened zone. Porosity in the MML may be responsible for this behavior. For example, when a
hard surface slides on a soft surface, for example, steel on Pb–Sn alloys [79], low friction and wear can
result due to softening on the surface, which is analogous to the Bowden and Tabor concept discussed
in Section 35.2.1.

35.3.4.1  Fracture Toughness


In general, fracture toughness is defined as the resistance to crack propagation. Crack formation during
sliding contact should be relatively easy in materials, for example, alloys, containing brittle phases, and,
thus, wear can be mostly controlled by crack growth. Hornbogen [80] studied the role of fracture tough-
ness in the wear of alloys to examine the trend of decreasing toughness resulting in higher relative wear
rates. When the applied strain is smaller than the critical strain (at which crack growth is initiated), the
wear rate is independent of toughness and Archard’s law is followed. However, when the applied strain is
larger than the critical strain, there is a greater probability of crack growth and thus a higher wear rate,
which, depends on the experimental conditions (pressure, strain rate, material properties, etc.). Similar
results were obtained by Zum Gahr and Doane for austenitic and martensitic irons [81]. However, in a
review article by Rosenfield [82], he determined that wear rates only depend on fracture toughness in
the case of hard, brittle alloys.
35-12 Wear Materials

35.3.4.2  Work Hardening and Ductility


As previously discussed, metallic-based materials are work (strain) hardened during wear processes, the
extent of which depends on material microstructure and operating conditions. In general, ductility is
reduced by an increase in yield strength relative to fracture strength, whereas the ductility can increase
for a given fracture strength when the work hardening rate increases relative to the yield strength. Zum
Gahr [23] summarized the effects of work hardening and ductility in metallic wear. Strain accumulates
during cyclic (repetitive) sliding wear and decays with increasing depth below the surface. The strains
will decay to a greater degree the higher the work hardening rate, that is, high strain rates will decrease
the work hardening rate. Higher surface and subsurface deformation strains result in increased friction
coefficient and applied stress. When the applied surface strains surpass the ductility, cracks can initiate
and propagate, especially at the aforementioned dislocation cell boundaries/walls.

35.3.5  High Temperature Wear


Unfortunately, wear of metallic-based materials at elevated temperatures has been investigated to a much
lesser degree than room temperature studies, and is compounded by the fact that the wear mechanisms
will change with increasing temperatures. High temperature metallic wear is important in such appli-
cations as internal combustion engines, aerospace propulsion systems, and metalworking equipment.
Peterson et al. [83] performed one of the earliest studies on high temperature metallic wear. Self-mated
sliding wear tests were performed for iron, copper, nickel, molybdenum, and chromium, where high
friction and considerable surface damage resulted until a transition temperature was reached result-
ing in lower friction and wear due to the formation of a softer, low shear metallic oxide. Stott [84] also
performed high temperature (up to 800°C) sliding tests on Ni–20%Cr base alloys (Nimonic 80A). At
higher temperatures, oxidation, compaction, and sintering of wear debris resulted in the formation of
hard oxide “glaze” surface layers that lead to lower wear. Fischer [85] conducted sliding abrasion tests
on numerous alloys (steels, Ni, and Co alloys) with and without hard carbide phases between room
temperature and 750°C. He determined that up to 650°C the wear rates decrease with increasing tem-
perature, which is more pronounced for materials with a softer metal matrix.
Recently, Blau [86] reviewed high temperature metallic wear and the effects of oxidation. Oxide reac-
tion products play a primary role in debris formation and microstructural evolution during metallic
wear. Also, the wear rate at elevated temperatures can either be enhanced or reduced depending on
contact conditions and nature of oxide layer formation. Some of his major conclusions are (1) as tem-
perature increases for metals, there are concomitant changes in their mechanical properties, coupled
with the role of oxidation can change the partitioning of frictional work into wear and surface dam-
age. (2) At least three factors affect the response of metallic interfaces at elevated temperatures: (a) the
mechanical properties of the reaction products on the surface, (b) the tendency to form stable transfer
layers/MML, and (c) the resistance of the bulk metals below the oxides and transfer layers/MML to
deformation and fracture.

35.4  Shape Memory Alloys


Shape memory alloys (SMA) are metal-based alloys that after deformation have the ability to return
to their original size and shape when temperature is changed. Deformation usually occurs at lower
temperatures, while shape memory occurs upon heating and/or cooling. SMAs that exhibit significant
amount of strain recovery are the NiTi alloys (Nitinol trade name) and some Cu-based (Cu–Zn–Al and
Cu–Al–Ni). SMA also exhibits high corrosion resistance, good fatigue properties, excellent biocom-
patibility, good ductility, and high mechanical damping. Like cobalt (HCP → FCC) and the austenitic
stainless steels (austenite → martensite), NiTi owes its superior properties to a stress-induced phase
transformation (austenite → martensite) upon cooling. Jin and Wang [87] were the first to examine
Metals 35-13

the wear behavior of NiTi, which they attributed to its pseudoelastic (superelastic) behavior caused by
the aforementioned stress-induced phase transformation. Li determined that NiTi exhibited high wear
resistance in comparison to other engineering materials such as steels, Ni-based and Co-based Tribaloy
alloys [88].
Most of the tribology studies on NiTi have concentrated on the equiatomic 50NiTi composition;
however, 55NiTi has been shown to exhibit superior wear resistance because of enhanced superelastic
behavior attributing to high E/H, high elastic recovery ratio, high yield load, and high transition strain
[89]. The dominant wear mechanism in superelastic TiNi was identified to be delamination wear, which
is initiated by the mismatch in elastic properties. Further improvements in friction and wear have been
determined by Dellacorte for 60NiTi composition (Nitinol 60) [90]. He determined that 60NiTi is hard,
wear resistant, non-galling, and tribochemically benign in the presence of liquid lubricants and is a
viable aerospace bearing material that is nonmagnetic, electrically conductive, hardenable, displays
favorable tribochemistry, and is non-corrosive. He also determined that Nitinol 60 has the ability
to boundary lubricate by the oils usually used on spacecraft mechanisms (Pennzane 2001A, Krytox
143AC, and Castrol 815Z), and is thus suitable for rolling and sliding contact applications such as bear-
ings and gears.

35.5  Nanocrystalline Metals and Alloys


Research in nc materials (<100 nm grain size), which are harder, stronger, but less ductile than their
coarse-grained counterparts, is hardly a new area. However, the tribological properties of nc metals and
alloys have received little attention until recently. There have been several recent reports on the friction
and wear properties of nc metals and alloys, such as Ni [33,91–95], Ti [96], Cu [96,97], 52100 steel [98],
and NiW [30,32]. As mentioned in Section 35.2.2, hardness has been shown to increase for nc metals due
to the classical dislocation-based Hall–Petch relationship that predicts increasing hardness and strength
with decreasing grain size. However, grain sizes in the critical range of ∼5–20 nm have been shown as a
threshold in which this relationship breaks down, sometimes referred to as the inverse Hall–Petch effect
when a peak in hardness is observed. Instead of dislocation-dominated deformation mechanisms, there
is a shift to grain boundary-dominated deformation mechanisms when grain sizes are in this range
[32,33]. As a result, the classical dimensional and nondimensional Archard wear Equations 35.2 and
35.3, respectively, break down for nc metals and alloys, since hardening (or softening) in the wear tracks
leads to deviations from the Archard relation, which linearly relates hardness and wear resistance [32].
Of the previous references, the majority [91,93–96,98] report that when nc metals are compared to
their mc metal counterpart, the nc metal exhibits improved wear resistance confirming that grain size
plays a significant role in hardness, plasticity, and wear. However, these improvements in wear are at
best only an order of magnitude different and are still considered relatively high wear materials since
they are metals, for example, wear rates of 1.3 × 10−4 mm3/Nm (mc-Ni) to 3 × 10−5 mm3/Nm (nc-Ni)
tested in air [93]. In addition, the softening or hardening inside the worn surface remains material-
specific and wear mode/mechanism-specific, and is somewhat controversial. For example, Qi et al. [94]
report increased nanoindentation hardness inside the wear track for electroplated mc-Ni compared
to unworn mc-Ni (i.e., 1.6–3.4 GPa) attributed to surface nanocrystallization and work hardening, as
opposed to slight softening for electroplated nc-Ni (7.8–6.7 GPa) attributed to grain-boundary-related
deformation modes. In contrast, Shafiei and Alpas [93] report microhardness values increase in the
worn surfaces for both mc and nc-Ni samples. In the case of mc-Ni, there is a steady increase from ∼120
to 180 Vickers hardness number (VHN) with sliding cycles due to increased amount of work harden-
ing; however, for nc-Ni there is no significant change in hardness (∼390 VHN) until 1000 sliding cycles
when the hardness increased to ∼490 VHN due to a hard oxidized tribolayer forming on the worn sur-
face. Unfortunately, based on these results, there will never be a universal wear model for comparing nc
metals to mc metals and alloys, and, thus the wear behavior and hardness will be specific for each mate-
rial and tribological operating condition. Lastly, Schuh et al. [30] compared the hardness and abrasive
35-14 Wear Materials

scratch resistance of nc-Ni to nc-Ni13W alloy and determined that they both increase for the alloy. His
group also studied the sliding wear behavior of the same alloy as a function of varying W content (3–28
at%) and across the full range of grain sizes (3–47 nm) over which the Hall–Petch breakdown occurs
[32]. At grain sizes of 15 nm and below there is a breakdown in Hall–Petch scaling, and at grain sizes
below ∼10 nm the Archard equation is also no longer obeyed with smaller grain sizes wearing less than
expected. This breakdown in Archard equation is correlated with wear-induced hardening in the wear
track, since wear induces grain boundary relaxation followed by grain growth (coarsening). Frictional
heating was determined to be too low and thus not responsible for the observed microstructural evolu-
tion (grain growth) during low speed sliding. Instead, deformation-driven microstructural evolution is
likely responsible for grain growth where the subsurface shear stress is at a maximum. Similar stress-
assisted grain growth in the HDL has also been observed for nc-Ni [33]. In summary, of all the experi-
mental research areas in tribology of metallic-based systems, nc metal and alloy tribology is likely to be
the most active in the future.

Acknowledgments
The author would like to acknowledge the financial support of the U.S. Air Force Research Laboratory
(AFRL, ISES Contract No. FA8650-08-C-5226) and the National Science Foundation (Grant No. CMMI-
0700828). The author would like to thank Bill Glaeser at Battelle Labs for reviewing the manuscript
and providing helpful comments. He would also like to thank colleagues Somuri V. Prasad, Brad L.
Boyce, Michael E. Chandross, Corbett C. Battaile, Thomas E. Buchheit, and Christopher R. Weinberger
from Sandia National Laboratories, Albuquerque, NM, for interesting and helpful conversations on the
tribology of nc metals/alloys and the hard versus soft question in highly deformed nc worn surfaces.

References
1. I. Newton, Opticks, Royal Society of London, 1704; reprinted by Dover publications, New York,
pp. 265–266, 1952.
2. R. Holm, Electric Contacts, Almqvist & Wiksells, Uppsala, Sweden, Section 40, 1946.
3. L.E. Samuels, The nature of some mechanically polished metal surfaces as evidenced by epitaxial
phenomena, Journal of the Institute of Metals, 85, 177–184, 1956.
4. J.F. Archard, Contact and rubbing of flat surfaces, Journal of Applied Physics, 24, 981–988, 1953.
5. J.F. Archard and W. Hirst, The wear of metals under unlubricated conditions, Proceedings of the Royal
Society of London. Series A, Mathematical and Physical Sciences, 236, 397–410, 1956.
6. J.K. Lancaster, The influence of temperature on metallic wear, Proceedings of the Physical Society.
Section B, 70, 112–118, 1957.
7. M.M. Khruschov, Principles of abrasive wear, Wear, 28, 69–88, 1974.
8. F.P. Bowden and D. Tabor, The Friction and Lubrication of Solids, Clarendon Press, Oxford, U.K.,
pp. 112 and 120, 1986.
9. I. Finnie, Erosion of surfaces by solid particles, Wear, 3, 87–103, 1960.
10. E. Rabinowicz, Friction and Wear of Materials, 2nd edn., John Wiley & Sons, New York, 1995.
11. I.V. Kragelskii, Friction and Wear, Butterworths, Washington, DC, 1965.
12. R.B. Waterhouse, Fretting Corrosion, Pergamon Press, Oxford, U.K., 1972.
13. T.F.J. Quinn, Review of oxidational wear I. The origins of oxidational wear, Tribology International,
16, 257–271, 1983.
14. K.C. Ludema, Friction, Wear, Lubrication: A Textbook in Tribology, CRC Press, Boca Raton, FL, 1996.
15. D.H. Buckley, Surface Effects in Adhesion, Friction, Wear and Lubrication, Elsevier, Amsterdam, the
Netherlands, 1981.
16. T.H.C Childs, The sliding wear mechanisms of metals, mainly steels, Tribology International, 13,
285–293, 1980.
Metals 35-15

17. D.A. Rigney, Sliding wear of metals, Annual Review of Materials Science, 18, 141–163, 1988.
18. N.P. Suh, Tribophysics, Prentice-Hall, Englewood Cliffs, NJ, 1986.
19. D. Kuhlmann-Wilsdorf, Dislocation concepts in friction and wear, in Rigney, D.A. (ed.), Fundamentals
of Friction and Wear of Materials, 1980 ASM Materials Science Seminar, American Society for Metals,
Metals Park, OH, pp. 119–186, 1981.
20. H. Czichos, Tribology—A Systems Approach to the Science and Technology of Friction, Lubrication
and Wear, Elsevier, Amsterdam, the Netherlands, 1978.
21. W.A. Glaeser, Microstructures associated with wear, Tribology in the 80’s, Proceedings of an International
Conference held at NASA Lewis Research Center, April, Cleveland, OH, pp. 219–235, 1983.
22. P.J. Blau, Friction and Wear Transitions of Materials, Noyes Publications, Park Ridge, NJ, 1989.
23. K.-H. Zum Gahr, Microstructure and Wear of Materials, Tribology Series, Vol. 10, Elsevier, Amsterdam,
the Netherlands, 1987.
24. S.K. Biswas, Wear of metals: A consequence of stable/unstable material response, Proceedings of the
Institution of Mechanical Engineers Part J: Engineering Tribology, 216, 357–369, 2002.
25. P.J. Blau, Fifty years of research on the wear of metals, Tribology International, 30(5), 321–331, 1997.
26. G.W. Stachowiak and A.W. Batchelor, Engineering Tribology, 3rd edn., Elsevier Butterworth-
Heinemann, Oxford, U.K., p. 430, 2005.
27. I.L. Singer, S.D. Dvorak, K.J. Wahl, and T.W. Scharf, Role of third bodies in friction and wear of pro-
tective coatings, Journal of Vacuum Science and Technology A, 21, S232–S240, 2003.
28. H.J. Choi, Y. Kim, J.H. Shin, and D.H. Bae, Deformation behavior of magnesium in the grain size
spectrum from nano- to micrometer, Materials Science and Engineering A, 527 1565–1570, 2010.
29. A.H. Chokshi, A. Rosen, J. Karch, and H. Gleiter, On the validity of the Hall–Petch relationship in
nanocrystalline materials, Scripta Metallurgica, 23, 1679–1683, 1989.
30. C.A. Schuh, T.G. Nieh, and H. Iwasaki, The effect of solid solution W additions on the mechanical
properties of nanocrystalline Ni, Acta Materialia 51, 431–443, 2003.
31. J.R. Trelewicz and C.A. Schuh, The Hall–Petch breakdown in nanocrystalline metals: A crossover to
glass-like deformation, Acta Materialia 55, 5948–5958, 2007.
32. T.J. Rupert and C.A. Schuh, Sliding wear of nanocrystalline Ni–W: Structural evolution and the
apparent breakdown of Archard scaling, Acta Materialia 58, 4137–4148, 2010.
33. S.V. Prasad, C.C. Battaile, and P.G. Kotula, Friction transitions in nanocrystalline nickel, Scripta
Materialia, 64, 729–732, 2011.
34. M.D. Sexton, A study of wear in Cu–Fe systems, Wear, 94, 275–294, 1984.
35. The Wear Control Handbook, M.B. Peterson and W.O. Winer, eds., ASME, New York, 1980.
36. Tribology Data Handbook, E.R. Booser, ed., CRC Press, Boca Raton, FL, 1997.
37. E. Rabinowicz, Wear coefficients—Metals, in Wear Control Handbook, M.B. Peterson and W.O.
Winer, eds., ASME, New York, pp. 475–506, 1980.
38. D. Tabor, Friction—The present state of our understanding, Journal of Lubrication Technology, 103,
169–179, 1981.
39. E. Rabinowicz, Determination of compatibility of metals through static friction tests, ASLE
Transactions, 14, 198–205, 1971.
40. N. Ohmae and E. Rabinowicz, The wear of the noble metals, ASLE Transactions, 23, 86–92, 1980.
41. D.H. Buckley and R.L. Johnson, Friction and wear of hexagonal metals and alloys as related to crystal
structure and lattice parameters in vacuum, ASLE Transactions, 9, 121–135, 1966.
42. E. Rabinowicz, The dependence of the adhesive wear coefficient on the surface energy of adhe-
sion, in Wear of Materials—1977, W.A. Glaeser, K.C. Ludema, and S.K. Rhee, eds., ASME, New York,
pp. 36–40, 1977.
43. P.J. Alison and H. Wilman, The different behavior of hexagonal and cubic materials in their friction,
wear and work hardening during abrasion, British Journal of Applied Physics, 15, 281–289, 1964.
44. J.P. Hirth and D.A. Rigney, Crystal plasticity and the delamination theory of wear, Wear, 39, 133–141,
1976.
35-16 Wear Materials

45. J.P. Hirth and D.A. Rigney, The application of dislocation concepts in friction and wear, in Dislocations
in Solids, Vol. 6, F.R.N. Nabarro, ed., North Holland, Amsterdam, the Netherlands, pp. 1–54, 1983.
46. D.A. Rigney and W.A. Glaeser, The significance of near surface microstructure in the wear process,
Wear, 46, 241–250, 1978.
47. J.J. Wert, The role of microstructure in subsurface damage induced by sliding contact, Key Engineering
Materials, 33, 101–134, 1989.
48. P. Heilmann, J. Don, T.C. Sun, D.A. Rigney, and W.A. Glaeser, Sliding wear and transfer, Wear, 91,
171–190, 1983.
49. J.J. Wert, S.A. Singerman, and S.G. Caldwell, An x-ray diffraction study of the effect of stacking fault
energy on the wear behavior of Cu-Al Alloys, Wear, 92, 213–219, 1983.
50. R.I. Blombery and C.M. Perrott, Adhesive wear processes occurring during abrasion of stellite type
alloys, Journal of the Australian Institute of Metals, 19, 254–258, 1974.
51. R.D. Schmidt and D.P. Ferriss, New materials resistant to wear and corrosion to 1000°C, Wear, 32,
279–289, 1975.
52. N.P. Suh and N. Saka, The stacking fault energy and delamination wear of single-phase f.c.c. metals,
Wear, 44, 135–143, 1977.
53. R.D. Evans, K.L. More, C.V. Darragh, and H.P. Nixon, Transmission electron microscopy of bound-
ary-lubricated bearings surfaces. Part I: Mineral oil lubricant, Tribology Transactions, 47, 430–439,
2004.
54. G.M. Scamans, M.F. Frolish, W.M. Rainforth, Z. Zhou, Y. Liu, X. Zhou, and G.E. Thompson, The ubiq-
uitous Beilby layer on aluminum surfaces, Surface and Interface Analysis, 42, 175–179, 2010.
55. M. Kerridge and J.K. Lancaster, The stages in a process of severe metallic wear, Proceedings of the
Royal Society of London. Series A, Mathematical and Physical Sciences, 236, 250–264, 1956.
56. W. Hirst and J.K. Lancaster, The influence of speed on metallic wear, Proceedings of the Royal Society
of London. Series A, Mathematical and Physical Sciences, 259, 228–241, 1960.
57. J.K. Lancaster, The formation of surface films at the transition between mild and severe metallic
wear, Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences, 273,
466–483, 1963.
58. P. Heilmann, W.A.T. Clark, and D.A. Rigney, Orientation determination of subsurface cells generated
by sliding, Acta Metallurgica, 31, 1293–1305, 1983.
59. D.A. Rigney, L.H. Chen, M.G.S. Naylor, and A.R. Rosenfield, Wear processes in sliding systems, Wear,
100, 195–219, 1984.
60. W.A. Glaeser, High strain wear mechanisms in ferrous alloys, Wear, 123, 155–169, 1988.
61. P.J. Blau, Mechanisms for translational friction and wear behavior of sliding metals, Wear, 72, 55–66,
1981.
62. D.A. Rigney, Large strains associated with sliding contact of metals, Materials Research Innovations,
1, 231–234, 1998.
63. J.M. Jungk, J.R. Michael, and S.V. Prasad, The role of substrate plasticity on the tribological behavior
of diamond-like nanocomposite coatings, Acta Materialia, 56, 1956–1966, 2008.
64. W. Cai, J. Mabon, and P. Bellon, Crystallographic textures and texture transitions induced by sliding
wear in bronze and nickel, Wear, 267, 485–494, 2009.
65. D.R. Wheeler and D.H. Buckley, Texturing in metals as a result of sliding, Wear, 33, 65–74, 1975.
66. J.H. Dautzenberg, The role of dynamic recrystallization in dry sliding wear, Wear, 60, 401–411,
1980.
67. L.K. Ives, Microstructural changes in copper due to abrasive, dry and lubricated wear, in Wear of
Materials—1979, K.C. Ludema, W.A. Glaeser, and S.K. Rhee, eds., ASME, New York, p. 246, 1979.
68. A.W. Ruff, L.K. Ives, and W.A. Glaeser, Characterization of wear surfaces and debris, in Fundamentals
of Friction and Wear of Materials, D.A. Rigney, ed., 1980 ASM Materials Science Seminar, American
Society for Metals, Metals Park, OH, pp. 235–289, 1981.
Metals 35-17

69. R.D. Evans, K.L. More, C.V. Darragh, and H.P. Nixon, Transmission electron microscopy of
boundary-lubricated bearing surfaces. Part II: Mineral oil lubricant with sulfur- and phosphorus-
containing gear oil additives, Tribology Transactions, 48, 299–307, 2005.
70. M.M. Khruschov and M.A. Babichev, Investigations of the Wear of Metals (in Russian), Academy
of Sciences of the USSR: National Engineering Laboratory Translation 889, National Engineering
Laboratory, East Kilbridge, Glasgow, U.K., 1960.
71. G. Sundararajan, The differential effect of the hardness of metallic materials on their erosion and
abrasion resistance, Wear, 162–164, 773–781, 1993.
72. K. Kato, Abrasive wear of metals, Tribology International, 30, 333–338, 1997.
73. R.C.D. Richardson, The wear of metals by hard abrasives, Wear, 10, 291–309, 1967.
74. J. Larsen-Basse, The abrasion resistance of some hardened and tempered carbon steels, Transactions
of the Metallurgical Society of AIME, 236, 1461–1466, 1966.
75. P.J. Mutton and J.D. Watson, Some effects of microstructure on the abrasion resistance of metals,
Wear, 48, 385–398, 1978.
76. K.-H. Zum Gahr, Wear by hard particles, Tribology International, 31, 587–596, 1998.
77. D.A. Rigney, The roles of hardness in the sliding behavior of materials, Wear, 175, 63–69, 1994.
78. Z.Y. Yang, M.G.S. Naylor, and D.A. Rigney, Sliding wear of 304 and 310 stainless steels, Wear, 105,
73–86, 1985.
79. X.J. Wang and D.A. Rigney, Sliding behavior of Pb–Sn alloys, Wear, 181–183, 290–301, 1995.
80. E. Hornbogen, The role of fracture toughness in the wear of metals, Wear, 33, 251–259, 1975.
81. K.-H. Zum Gahr and D.V. Doane, Optimizing fracture toughness and abrasion resistance in white
cast irons, Metallurgical Transactions A, 11, 613–620, 1980.
82. A.R. Rosenfield, Wear and fracture mechanics, in Fundamentals of Friction and Wear of Materials,
D.A. Rigney, ed., 1980 ASM Materials Science Seminar, American Society for Metals, Metals Park,
OH, pp. 221–234, 1981.
83. M.B. Peterson, J.J. Florek, and R.E. Lee, Sliding characteristics of metals at high temperatures, ASLE
Transactions, 3, 101–109, 1960.
84. F.H. Stott, High-temperature sliding wear of metals, Tribology International, 35, 489–495, 2002.
85. A. Fischer, Mechanisms of high temperature sliding abrasion of metallic materials, Wear, 152,
151–159, 1992.
86. P.J. Blau, Elevated-temperature tribology of metallic materials, Tribology International, 43, 1203–1208,
2010.
87. J.L. Jin and H.L. Wang, Research on wear resistance of NiTi alloy, Acta Metallurigica Sinica, 24,
A66–A69, 1988.
88. D. Li, A new type of wear-resistant material: Pseudo-elastic TiNi alloy, Wear, 221, 116–123, 1998.
89. Z.N. Farhat and C. Zhang, The role of reversible martensitic transformation in the wear process of
TiNi shape memory alloy, Tribology Transactions, 53, 917–926, 2010.
90. C. Dellacorte, Nickel-titanium alloys: Corrosion “proof ” alloys for space bearing, components and
mechanism applications, NASA/CP-2010-216272, 2010.
91. D.H. Jeong, F. Gonzalez, G. Palumbo, K.T. Aust, and U. Erb, The effect of grain size on the wear prop-
erties of electrodeposited nanocrystalline nickel coatings, Scripta Materialia, 44, 493–499, 2001.
92. C.A. Schuh, T.G. Nieh, and T. Yamasaki, Hall–Petch breakdown manifested in abrasive wear resis-
tance of nanocrystalline nickel, Scripta Materialia, 46, 735–740, 2002.
93. M. Shafiei and A.T. Alpas, Friction and wear mechanisms of nanocrystalline nickel in ambient and
inert atmospheres, Metallurgical Transactions A, 38, 1621–1631, 2007.
94. Z. Qi, J.C. Jiang, and E.I. Meletis, Wear mechanisms of nanocrystalline metals, Journal of Nanoscience
and Nanotechnology, 9, 4227–4232, 2009.
95. D.J. Guidry, K. Lian, J.C. Jiang, and E.I. Meletis, Tribological behavior of nanocrystalline nickel,
Journal of Nanoscience and Nanotechnology, 9, 4156–4163, 2009.
35-18 Wear Materials

96. P. Iglesias, M.D. Bermúdez, W. Moscoso, B.C. Rao, M.R. Shankar, and S. Chandrasekar, Friction and
wear of nanostructured metals created by large strain extrusion machining, Wear, 263, 636–642.
2007.
97. S. Tarasov, V. Rubtsov, and A. Kolubaev, Subsurface shear instability and nanostructuring of metals
in sliding, Wear, 268, 59–66, 2010.
98. L. Zhou, G. Liu, Z. Han, and K. Lu, Grain size effect on wear resistance of a nanostructured AISI52100
steel, Scripta Materialia, 58, 445–448, 2008.
36
Wear and Lubrication
of Ceramics
36.1 Introduction.....................................................................................36-1
Alumina  •  Silicon Nitride  •  Silicon Carbide  •  Zirconia
36.2 Wear under Sliding Conditions.....................................................36-4
Alumina  •  Silicon Nitride  •  Silicon Carbide  •  Zirconia
36.3 Wear under Rolling and Abrasive Conditions............................36-9
Rolling Contact Fatigue and Wear  •  Abrasive Wear
36.4 Sliding Wear Models.....................................................................36-10
Wear Rate Models  •  Contact Fracture Models  •  Transition Loads
in Sliding Contacts
36.5 Boundary Lubrication...................................................................36-15
Lubrication with Hydrocarbon Fluids  •  Lubrication with Water
Said Jahanmir
Mohawk Innovative 36.6 Design with Ceramics...................................................................36-17
Technology, Inc. References...................................................................................................36-18

36.1  Introduction
Ceramics are generally defined as inorganic nonmetallic solid materials. This definition includes not
only materials such as pottery, porcelains, refractories, cements, abrasives, and glass but also nonmetal-
lic magnetic materials, ferroelectrics, and a variety of other new materials. Interest in using ceramics
for engineering application is rooted in unique ceramics classified as advanced structural ceramics and
electronic and optical ceramics [1,2]. Structural ceramics are presently used in diverse applications as
tribomaterials [3–5] due to their unique properties that include resistance to wear and corrosion at ele-
vated temperatures, low density, and unique electrical, thermal, and magnetic properties. Applications
include precision instrument bearings, cutting tool inserts, prosthetic articulating joints, and engine
components.
This chapter focuses on two issues that determine performance of structural ceramics in tribologi-
cal applications: the transition from mild to severe wear and the influence of lubrication on wear and
particularly on wear transition.
Structural ceramics, by definition, are high-strength ceramics used in load-bearing applications. In
this section, the structure, processing methods, and properties of commonly used structural ceramics
are reviewed. Typical properties of different types of structural ceramics are listed in Table 36.1.

36.1.1  Alumina
Ceramics based on aluminum oxide or alumina have been used in commercial applications for many
years because of their availability and low cost. Alumina ceramics are often classified either as high

36-1
36-2 Wear Materials

TABLE 36.1  Typical Mechanical and Physical Properties of Structural Ceramics


Fracture Fracture Elastic Thermal Thermal
Density Strength Toughness Modulus Poisson’s Hardness Expansion Conductivity
Material (g/cm3) (MPa) (MPa m1/2) (GPa) Ratio (GPa) (10−6 m/m/C) (W/m·C)
Alumina
99.8 3.9 550 4.5 386 0.22 15 6.7 30
96 3.7 350 4.5 303 0.21 12 8.2 25
90 3.6 340 3.5 276 0.22 10.4 8.1 17
85 3.4 300 3.5 221 0.22 9.4 7.2 16

Zirconia
Cubic 5.8 245 2.4 150 0.25 11 8.0 1.7
PSZ 6.0 750 8.1 205 0.23 12 8.3 2.1
TZP 6.1 1150 12.0 210 0.24 13 10.6 3.0

Silicon nitride
Sintered 3.2 550 4.5 276 0.24 14 3.4 28
Hot-pressed 3.2 800 5.0 317 0.28 15 3.2 30
Reaction- 2.5 250 3.6 165 0.24 10 2.8 12
bonded
Sintered- 3.3 825 5.8 310 0.28 19 3.5 35
reaction-
bonded
HIP’ed 3.2 1000 6.0 310 0.28 20 3.5 32

Silicon carbide
Reaction- 3.0 350 3.5 385 0.20 25 4.4 125
bonded
Hot-pressed 3.2 550 3.9 449 0.19 25 4.5 115
Sintered 3.1 400 2.8 427 0.21 27 4.8 150
CVD 3.2 500 3.1 450 0.21 30 5.5 121
Note: Data in this table have been compiled from various sources including material data sheets from suppliers on the
Internet. The data listed represent average values when data variations were noticed due to specific composition and
processing.

aluminas having mass fractions >80% aluminum oxide, or as porcelains with <80% aluminum oxide
[1,2]. High aluminas are used in many mechanical devices and in electronics.
Pure aluminum oxide, Al2O3, has one thermodynamically stable phase at room temperature, des-
ignated as alpha phase. Often the term corundum is used for alpha alumina. Commercial aluminas
are densified by pressureless sintering, using MgO sintering aid for high-purity (>95%) aluminas or
silicates for less expensive, low-purity grades. Fracture strength improves as the percentage of alumina
is increased. However, cost increases because of processing difficulty. The thermal shock resistance of
aluminas is generally inferior to other structural ceramics (e.g., silicon nitride), because of the highly
anisotropic properties that depend on the crystallographic orientation of alumina grains.

36.1.2  Silicon Nitride


Silicon nitride, one of the strongest structural ceramics, has emerged as an important tribological mate-
rial, especially in rolling applications. It has low elastic modulus, high strength, and outstanding fatigue
resistance, excellent oxidation resistance due to a protective surface oxide layer, and very good thermal
shock resistance because of its low thermal expansion coefficient. In oxidizing environments, silicon
nitride is stable only at very low partial pressures of oxygen; in air, it rapidly forms a silicon oxide surface
Wear and Lubrication of Ceramics 36-3

layer. This layer protects against further oxidation; if it is removed, for example, by wear, oxidation
occurs rapidly.
Pure silicon nitride, Si3N4, exists in two crystallographic forms designated as alpha and beta. Since
the beta phase is thermodynamically more stable, silicon nitride materials are primarily in the beta
phase, but starting powder is usually in the alpha phase [2]. Commercial silicon nitride materials are
processed with various oxide-sintering aids.
Silicon nitride materials are classified according to processing techniques: sintered, hot pressed,
reaction bonded (or reaction-sintered), sintered reaction bonded, and hot isostatically pressed [1,2].
Variations in composition, microstructure, and properties depend on the processing route used in
fabrication.
Silicon nitride powder compacts are sintered to full density using combinations of rare earth oxides
and alumina-sintering aids. However, mechanical properties (i.e., strength, hardness, and fracture
toughness) of commercially available sintered silicon nitrides are inferior to those processed by hot
pressing, usually containing MgO- or Y2O3-sintering aids. Application of pressure during sintering
leads to materials with nearly full density and very good properties. Shapes that can be formed by hot
pressing, hot isostatic pressing (HIP), and gas pressure sintering (GPS), however, are limited and pro-
cessing cost is relatively high.
Reaction-bonded silicon nitride is made by compacting silicon powder and reaction-bonding with
nitrogen at high temperatures. While they are much less expensive than hot-pressed or sintered materi-
als, their porosity, which is greater than 10%, results in inferior mechanical properties. Adding oxide-
sintering aids to the starting silicon powder and a subsequent hot pressing, or HIP, reduces this porosity
and improves the properties. The principal advantage of reaction-bonded silicon nitride is its lower cost
of starting powder.

36.1.3  Silicon Carbide


Silicon carbide ceramics are widely used in applications requiring wear resistance, high hardness, reten-
tion of mechanical properties at elevated temperatures, and resistance to corrosion and oxidation. The
oxidation resistance is due to a protective silicon oxide layer, as with silicon nitride. The thermal shock
resistance is lower than that of silicon nitride. Silicon carbide ceramics are grouped into four types
depending on processing methods: reaction-bonded (or reaction-sintered), hot-pressed, sintered, and
chemical vapor deposition (CVD) [1,2].
In the reaction-bonding process, a mixture of silicon carbide powder, graphite, and a plasticizer is
pressed in a mold to prepare a preform or “green” compact. After the plasticizer is burned off to produce
a porous product, silicon metal is infiltrated into the pores as a liquid or vapor. The reaction between
silicon and carbon to form SiC is not usually complete, leaving some residual Si and C; excess Si is used
to fill the pores. The finished material has little porosity and contains a mixture of Si, C, and reaction-
formed SiC in between the original SiC particles. Mechanical properties depend on the amount of free
Si and C. Since the densification process does not produce shrinkage, tight dimensional tolerances are
more easily achieved than with sintering. The primary advantage of reaction-bonded silicon carbide is
its relatively low cost. Because components can be made to near net shape, little machining is required
after densification.
Hot pressing produces high-strength silicon carbide of nearly full density. In this process, boron and
carbon and sometimes alumina are used as sintering aids for processing of both alpha and beta silicon
carbide components. Although hot-pressed silicon carbide exhibits very good mechanical properties, its
use is limited by the high cost of finished components, due to difficulty in machining after densification.
Silicon carbide components are also produced by sintering without the application of pressure, using
carbon- and boron-sintering aids. The major advantage of this process is that most of the machining is
easily performed on the green compact. The densified component is then finish machined by diamond
grinding and polishing.
36-4 Wear Materials

CVD is used to produce a relatively pure and dense silicon carbide. This material is highly
anisotropic, due to the columnar structure developed during the deposition process. In addition to
anisotropy, high cost and residual stresses are major drawbacks against widespread use. Nevertheless,
CVD SiC is an excellent coating material where resistance to wear, erosion, or oxidation is required.

36.1.4  Zirconia
Zirconia ceramics are characterized by high strength and toughness at room temperature. Their major
limitation in tribological service is a low thermal conductivity, which causes wear by thermal shock at
high sliding speeds. Pure zirconia exists in three crystal structures: monolithic, tetragonal, and cubic
[1,2]. The monolithic phase is stable up to 1170°C, where it transforms to the tetragonal phase. The
tetragonal phase transforms to the cubic phase at 2370°C. The microstructure and properties can be
controlled by addition of various oxides such as MgO, CaO, Y2O3, and CeO2. The amount of oxide addi-
tives and thermal processing can be chosen such that the tetragonal and cubic phases become stable at
room temperature. Zirconia ceramics used in engineering applications are classified into three main
types: cubic, partially stabilized, and tetragonal zirconia.
Cubic zirconia is obtained by fully stabilizing the high-temperature cubic phase by addition of
about 10 mol% oxides. Its relatively low fracture toughness and strength limits its use in tribological
applications.
Partially stabilized zirconia, PSZ, has a two-phase microstructure consisting of cubic grains with
tetragonal and/or monoclinic precipitates, depending on the thermal processing history. It exhibits
increased fracture toughness over the cubic zirconia, and is therefore of importance in structural and
tribological applications. The compressive stress associated with an increase in volume in transforma-
tion of metastable tetragonal precipitates to the monoclinic phase at an advancing crack tip results in a
high strength and toughness. Typical commercial PSZ materials contain about 8 mol% MgO or CaO and
have a composition of about 58% cubic, 37% tetragonal, and 5% monoclinic.
Tetragonal zirconia polycrystal, TZP, is made by addition of about 2–3 mol% Y2O3 or CeO2 to stabi-
lize the tetragonal phase. This material is nearly 100% tetragonal at room temperature and exhibits the
highest toughness and strength among zirconia ceramics and other monolithic structural ceramics. The
toughening mechanism is similar to that for PSZ, namely, a tetragonal to monoclinic transformation
under stress. TZP ceramics are only suitable for tribological applications at room temperature because
of phase transformations and decrease in strength above 200°C.

36.2  Wear under Sliding Conditions


Various studies on tribological behavior of ceramics have shown that the wear mechanisms depend
on contact conditions. A common feature in most structural ceramics is that wear occurs through
a small-scale fracture (i.e., microfracture) process if the contact load is higher than a threshold
value specific to each material [5]. A second common feature associated with many ceramics is their
high reactivity with water vapor in the air environment forming oxide and hydroxide surface lay-
ers [6]. Such reactions are accelerated during wear testing, and thus are referred to as tribochemical
reactions. The terms “mild wear” and “severe wear” are used to distinguish the two extreme behav-
iors. Mild wear refers to a wear coefficient lower than 10 −4 and severe wear refers to unacceptably high
wear above this value. The wear coefficient is a dimensionless parameter that is obtained by multiply-
ing the measured wear volume by hardness of the ceramic and dividing this product by the sliding
distance and contact load [7]. This normalization implies that the wear volume is proportional to the
applied load and inversely proportional to the hardness of the material. Such linear relationships do
not generally hold for most ceramics under a wide range of contact conditions. Nevertheless, wear
coefficients serve as a universal parameter for comparison of wear data obtained under different test
conditions.
Wear and Lubrication of Ceramics 36-5

Wear mechanisms and concomitant wear rates depend on many contact parameters and fundamen-
tal properties as well as the ceramic microstructure. After the publication of wear mechanisms map for
steels by Lim and Ashby [8], many attempts were made to develop such maps for ceramics. Normalized
contact pressure and normalized velocity were used as the coordinates for the wear mechanisms map
of steel, along with analytical models to clarify the role of different parameters. A similar approach
was used by Kong and Ashby [9] to develop wear maps for alumina identifying different wear modes.
Wear maps for various ceramics have been constructed to delineate the effects of load, velocity, and
temperature [10–12]. The tribological behavior of ceramics depends on the composition, microstruc-
ture, and processing method of the material, as well as on the contact conditions. Therefore, published
data should be used with caution when extrapolated to predict the behavior of a different ceramic, since
wear performance of an “identical” ceramic from a different manufacturer or different processing steps
can vary greatly.

36.2.1  Alumina
Analysis of published data shows certain trends [4]. Although the variability in published results on
the effects of contact conditions on wear is large, the wear rate of alumina generally increases as either
sliding speed or load is increased. In self-mated tests at room temperature, the coefficient of friction of
high-purity aluminas (>95%) is lower than those of lower-purity grades [4]. The coefficient of friction
increases if the alumina counterface is changed to steel. The coefficient of friction in self-mated slid-
ing couples is reduced if the alumina counterface is replaced with zirconia, silicon nitride, or silicon
carbide.
The tribological behavior of a high-purity alumina (99.8%) sliding in air in self-mated, ball-on-
flat, unlubricated tests is displayed in Figure 36.1 [10] as a function of normal load and temperature.
The specific test conditions are listed in the figure caption. The Hertzian contact stress is indicated
on the right ordinate for reference. The wear coefficients, K, and the coefficients of friction, f, are
included in the figure. The wear transition diagram delineates the effects of load and temperature
and identifies the boundaries between regions dominated by different mechanisms under low sliding
speeds. The wear transition diagram can be used to predict the tribological behavior and to determine
the useable range of conditions (i.e., load and temperature). The tribological characteristics of this
alumina under low-speed sliding are divided into four distinct regions based on the fundamental
mechanisms involved in the wear process. At low temperatures (T < 200°C), tribochemical reactions
between alumina and water vapor in the environment control the behavior; friction coefficient and

100 2.50

80 2.33
Contact stress (GPa)
Contact load (N)

<IV>
60 2.11
f = 0.85 K > 10–4 <III>
< I>
40 1.84

20 f = 0.40 f = 0.40 1.47


K < 10–6 f = 0.60 <II> K < 10–6 K < 10–6

0
0 200 400 600 800 1000
Temperature (°C)

FIGURE 36.1  Wear transition diagram for 99.8% alumina. Self-mated, reciprocating, ball-on-flat tests in air; slid-
ing speed: 1.4 mm/s, ball diameter: 12.7 mm, polished surfaces, Ra: 0.1–0.2 μm. (From Jahanmir, S. and Dong, X.,
J. Tribology, 114, 403, 1992. With permission.)
36-6 Wear Materials

wear coefficient are relatively low at 0.40 and 10−6. Formation and removal of a thin hydroxide film
and formation of cylindrical wear debris control the tribological performance of this ceramic. At
intermediate temperatures (200°C < T < 800°C), wear behavior depends on contact load. At low loads,
wear occurs by plastic flow and plowing; friction coefficient and wear rates are low, similar to those
observed at the lower temperatures. At loads higher than 20 N, severe wear by intergranular fracture
results in a friction coefficient of 0.85 and a wear coefficient larger than 10 −4. A fracture mechanics
model [10] has confirmed that the observed mild to severe wear transition at the intermediate temper-
atures is controlled by propagation of microcracks from preexisting near surface flaws. The transition
from mild to severe wear is sensitive to the sliding speed [13,14]. For example, for a 99.7% purity alu-
mina tested under a constant load of 10 N, the wear increases by more than two orders of magnitude
as the speed is increased from 0.003 to 0.3 m/s [14]. At temperatures above 800°C, both the coefficient
of friction and wear coefficient are low because of the formation of a thin surface layer from diffusion
and viscous flow of the silica glass grain boundary phase. The wear rate and the wear transition are
both influenced by the grain size [15] and the glass content [13] due to the effect of these parameters
on the fracture toughness of aluminas.

36.2.2  Silicon Nitride


A clear distinction between the performances of different silicon nitride ceramics cannot be made due
to the large scatter in published data. In general, the wear resistance is improved as either the fracture
toughness or the hardness is increased through processing [4]. The wear rate increases as the load is
increased, and often a sudden increase in wear rate is observed at a specific transition load [16]. The wear
rate also increases with temperature, particularly above 800°C. The wear rate at higher temperatures
decreases as the speed is increased. Generally, replacing the silicon nitride counterface with steel or
alumina slightly increases friction, and a zirconia counterface can decrease the friction coefficient [4].
The wear transition diagram for a hot isostatically pressed silicon nitride in low-speed, self-mated,
ball-on-flat, unlubricated tests in air comprises five regions (Figure 36.2) [16]. For the test condi-
tions specified in the figure and in the figure caption, the contact stress ranged from 0.5 to 2.4 GPa.
The boundaries (the cross-hatched area) between the five regions are broader than those in alumina.
At low loads and relatively low temperatures (T < 400°C), the tribological behavior is controlled by
formation of silicon hydroxide on the wear track [17], with a friction coefficient of 0.30, and a wear
coefficient of ∼10−4. Cylindrical wear debris is formed on the wear track from the hydroxide film.

100

<V>
–2
f = 0.80 K = 10
Load (N)

10

<I>
<II>
–4
<III> f = 0.70
f = 0.30 K = 10 –4 f = 0.67 <IV>
f = 0.45 K = 10 –2
–3 K = 10
K = 10
1
0 200 400 600 800 1000
Temperature (°C)

FIGURE 36.2  Wear transition diagram for a hot isostatically pressed silicon nitride. Self-mated, reciprocating,
ball-on-flat tests in air; sliding speed: 1.4 mm/s, ball diameter: 12.7 mm, polished surfaces, Ra: 0.1–0.2 μm. (From
Dong, X. and Jahanmir, S., Wear, 165, 169, 1993. With permission.)
Wear and Lubrication of Ceramics 36-7

At higher temperatures (400°C < T < 700°C), selective oxidation of WC inclusions controls the wear


behavior. Formation of crystalline precipitates from amorphous magnesium silicate grain boundary
phase controls the wear process at higher temperatures (700°C < T < 900°C), where both the friction
coefficient and wear coefficient increase. Oxidation of silicon nitride controls the wear behavior at
elevated temperatures (T > 900°C); the friction coefficient is ∼0.70, and the wear coefficient increases
to 10−2. Similar to alumina, microfracture is the primary wear mechanism at high loads. The friction
coefficient increases to ∼0.80 and the wear coefficient is very high at 10−2. The critical load for onset of
microfracture decreases from about 14 to 3 N as the temperature increases to 900°C. This transition is
associated with an increase in wear by two orders of magnitude. Unlike alumina, the transition load in
silicon nitride is less sensitive to an increase in speed [18]. The high loads that cause microfracture must
be avoided because of large wear coefficients when wear occurs by microfracture.

36.2.3  Silicon Carbide


In general, the coefficient of friction of reaction-bonded silicon carbide is lower than the other types
because the excess carbon can act as a solid lubricant [4]. However, this effect depends on the specific
microstructure and amount of free carbon, as well as on contact conditions. The coefficient of friction
for sintered materials decreases with increasing temperature and speed. The wear rate for the reaction-
bonded materials tends to increase with temperature and speed. Both the coefficient of friction and wear
rate for hot-pressed materials are lower in sliding against alumina and zirconia than sliding against
silicon nitride or silicon carbide [4].
The wear transition diagram for a sintered silicon carbide in low-speed, self-mated, ball-on-flat, unlu-
bricated tests in air (Figure 36.3) shows four regions [19]. The contact stress ranged from 0.6 to 2.9 GPa.
At room temperature, high loads, and high relative humidities, friction coefficient is ∼0.23 and the wear
coefficient is 10−4. At temperatures near the room temperature, tribochemical reaction between water
vapor and silicon carbide controls the tribological behavior. The tribochemical reactions with water
form a continuous thin film on the wear track. As the humidity decreases from about 70% to 30%, the
friction coefficient increases to 0.70. Wear occurs by plowing and the friction coefficient is 0.63. Both
friction and wear decrease at higher temperatures due to oxidation of silicon carbide and formation of
large cylindrical debris on the wear track. At high loads, wear occurs by microfracture, resulting in a
high coefficient of friction and wear coefficient. When contact fracture occurs at lower temperatures, the

<I> –4
f = 0.23, K = 10
100

f = 0.78 f = 0.77
K = 10–3 K = 10–3
<IV> a <IV> b
Load (N)

10

<II>
<III>
f = 0.67
K = 10–4 f = 0.40, K = 10–4

1
0 200 400 600 800 1000
Temperature (°C)

FIGURE 36.3  Wear transition diagram for a sintered silicon carbide. Self-mated, reciprocating, ball-on-flat
tests in air; sliding speed: 1.4 mm/s, ball diameter: 12.7 mm, polished surfaces, Ra: 0.1–0.2 μm. (From Dong, X. and
Jahanmir, S., Tribology Int., 28, 559, 1995. With permission.)
36-8 Wear Materials

wear debris is loosely attached to the surface, whereas at higher temperatures, some of the debris form
compacted regions. The mild to severe wear transition and associated change in the wear mechanisms
to microfracture occurs at a load of 7 N, irrespective of temperature, and the wear coefficient increases
by one order of magnitude.

36.2.4  Zirconia
The wear rate of cubic zirconia is generally high and wear occurs by microfracture [20]. PSZ ceramics
are much better than the cubic zirconia with respect to both strength and wear resistance. The wear rate
of unlubricated PSZ is generally in the mild wear regime at room temperature [21]. Under mild wear
conditions, wear of zirconia is dominated by tribochemical processes and formation of amorphous
hydride films [22]. The wear rate increases at higher loads or speeds due to mild to severe wear transi-
tion [23,24]. The tribological properties of PSZ are sensitive to temperature. Wear becomes severe at
∼200°C due to (tetragonal-cubic) phase transformations but changes to mild wear above 300°C [25,26].
The same type of transformation can also occur due to increased sliding speed [27]. The coefficient of
friction for self-mated zirconia surfaces is usually higher (0.3–0.7) than what is needed for practical
applications [28]. The wear process (for both Mg-PSZ and Y-TZP) follows plastic deformation, delami-
nation, and microfracture [29]. The coefficient of friction of PSZ–steel sliding couple is lower than a
self-mated PSZ combination, and as the load is increased the coefficient of friction of PSZ–steel com-
bination decreases.
The fracture properties and the wear behavior of zirconia are substantially improved by addition of
Yttria to stabilize the tetragonal phase (Figure 36.4) [20]. The tribological performance of TZP is supe-
rior to that of PSZ at room temperature. The coefficient of friction of TZP is lowest at room temperature
when slid against alumina; the wear resistance, however, is greatest against silicon nitride or silicon

10–3

f = 0.7
10–4

0.7
Specific wear rate (mm3/N)

0.14
0.1
10–5

f = 0.6
–6
10
0.35
0.09
0.11
0.7
10–7 Cubic Tetragonal
Air
Water
Hexadecane
Stearic acid
Nitrogen
10–8
1 10 100 1000

Sliding distance

FIGURE 36.4  Wear rate of cubic and tetragonal zirconia in different environments. Friction coefficient f for each
condition is also shown. Tests performed at room temperature at low sliding speeds. (From Fischer, T.E. et al., Wear,
122, 133, 1988. With permission.)
Wear and Lubrication of Ceramics 36-9

carbide counterface [4]. The friction of TZP increases with increasing load or speed. The coefficient of
friction of TZP is higher than that of PSZ at higher sliding speeds.
Water has deleterious effects on wear of TZP and PSZ [26,29] and tends to accelerate wear in a process
akin to stress corrosion cracking. Even a small amount of water dissolved in lubricating oils can increase
the wear rate of Y-TZP.

36.3  Wear under Rolling and Abrasive Conditions


36.3.1  Rolling Contact Fatigue and Wear
The lower density, corrosion resistance, and wear resistance of high-strength ceramics make them par-
ticularly important for rolling element bearings and cam roller followers in internal combustion engines
and other applications [30]. An important issue in selection of brittle materials for rolling conditions
is the particular failure mode that the ceramic material exhibits. HIP silicon nitride, for example, fails
by spall formation, similar to high-strength steels used in rolling element bearings. Spalling describes
a specific form of surface damage and it is the most common mode of failure in rolling contacts. This
failure mode occurs by rolling contact fatigue, which is a surface damage process due to the repeated
application of stresses when the surfaces of two bodies roll over each other. Rolling contact fatigue is
encountered most often in rolling element bearings, gears, and rollers. The failure process of rolling con-
tact fatigue involves fatigue crack initiation and propagation, leading to the formation of a spall fracture
on the surface. The process is related to the characteristics of the surface quality, stress distribution, and
lubrication condition.
While formation of spall fracture limits the contact life, under certain conditions, small wear par-
ticles form through microfracture and enlarge the contact area. This form of wear by attrition is usually
not life limiting, and in fact can eliminate the rolling contact fatigue process through a reduction in
contact stresses. The wear mechanisms in rolling are similar to those observed during sliding, mainly,
microfracture and grain pullout [31].
Studies have shown that rolling wear rates of PSZ and reaction-bonded silicon nitride are much
higher than that of fully dense silicon nitrides [32]. Since most of the rolling wear studies are conducted
with hydrocarbon lubricants, the tribochemical effects observed for unlubricated sliding wear tests are
usually not observed in rolling. Even under unlubricated conditions in air, tribochemical reactions can
occur and wear is dominated by a microfracture process [33]. However, addition of sliding increases the
wear rate by several orders of magnitude and modifies the wear mechanism by adding the contribu-
tion of the tribochemical reaction between silicon nitride and moisture [34]. The increased wear due to
sliding often makes it difficult to assess rolling contact fatigue, especially when pure rolling conditions
cannot be established. For example, when a ball-on-rod fatigue test is used, the contact stress can rapidly
decrease leading to a reduced potential for failure by contact fatigue [35].
Other structural ceramics, for example, SiC, typically fail by catastrophic fracture due to their low
fracture toughness or severe wear by microfracture. Catastrophic failure can cause damage to the entire
bearing assembly. Nevertheless, silicon carbide has been successfully used as lightly loaded rolling ele-
ments in water pumps due to its resistance to abrasive wear and corrosion [30].

36.3.2  Abrasive Wear


Structural ceramics are generally resistant to abrasive wear due to their high hardness. However, they
can be subjected to abrasive wear if the contacting particles or the sharp counterface asperities possess
higher hardness than the ceramic material.
The abrasive wear mode is controlled by the penetration depth of the sharp particle (or counterface
asperities) and mechanical properties of the ceramic surface. This process can be best described by first
examining what happens when an indentation is made into the material similar to hardness testing with
36-10 Wear Materials

5000
20 N

Volume removed, V (µm3/µm)


4000 40 N

3000
(3)
(1) Damage accumulation
2000 (2) Grain dislodgement
(3) Lateral crack chipping

1000
(2)
(1) (2)
0
0 5 10 15 20
Number of passes, n

FIGURE 36.5  Wear volume and wear transitions of aluminum oxide sliding at low speeds against a diamond
indenter. (From Xu, H.H.K. and Jahanmir, S., Wear, 192, 228, 1996. With permission.)

a diamond indenter. Initially, the indenter deforms the surface by plastic deformation and produces an
indent [36]. Deformation, however, is not limited to the surface as the material below the indent is also
deformed either elastically or plastically, depending on the indentation load (and the hardness of the
ceramic). The subsurface plastic deformation is described as a “quasi-plastic” deformation process, espe-
cially in polycrystalline ceramics, where a network of intergranular and/or intragranular microcracks is
formed. Upon unloading, lateral or vent cracks are formed surrounding the indent due to residual stresses
associated with incompatibility between the deformation zone and surrounding elastic matrix [37]. If the
indenter is traversed along a linear path, the lateral cracks follow the indenter motion. Therefore, three
distinct mechanisms govern the removal of material in abrasive wear of ceramics under increasing load:
(1) small-scale plastic deformation and plowing, (2) grain-scale microfracture and grain dislodgement,
and (3) lateral crack chipping. The transition from the deformation-controlled to fracture-controlled
material removal depends also on the number of passes an abrasive particle makes with the surface since
each pass leaves behind a certain amount of plastic strain damage and residual stress (Figure 36.5) [15,38].
Various studies have confirmed that the abrasive wear rate of polycrystalline ceramics is gener-
ally related to both the hardness (controlling the indentation depth) and toughness (controlling the
crack propagation rate) [15,38,39]. Therefore, abrasive wear can be related to any microstructural fea-
tures (such as grain size, grain composition, or porosity) that control the mechanical properties of the
ceramic. It has been reported that the wear mode of fine-grained materials, for example, alumina, is
mostly by deformation and microfracture, but larger grain sizes promote lateral crack chipping [38,40].
Other parameters such as grain orientation and thermal expansion mismatch stresses due to second
phases and grain boundary phase also have pronounced effects on toughness and abrasive wear rate
[41]. Due to the microstructural effects on grain-scale toughness, it is often found that conventional
fracture toughness measurements (using precracked or notched specimens) for polycrystalline ceramics
cannot be used for prediction of abrasive wear rates. Instead, the toughness measured for small grain-
scale microcracks should be used if the material exhibits the so-called R-curve behavior (i.e., crack size
dependent toughness) [36,38].

36.4  Sliding Wear Models


36.4.1  Wear Rate Models
Various models have been suggested in the literature for the prediction of ceramic wear rates (see, for
example, [33,39,42]). Most models relate wear to the hardness, elastic modulus, and fracture toughness
Wear and Lubrication of Ceramics 36-11

of ceramics and model the wear process in the microfracture region. Most models rely on many assump-
tions with limited supporting information and include parameters and constants that can only be deter-
mined experimentally. It suffices to say that in general wear rates increase with the contact load or
pressure and decrease as the toughness is increased. The wear rate also depends on the coefficient of
friction and any attempt to reduce friction is usually beneficial. Although published models are use-
ful as they delineate the importance of microstructure and mechanical properties, after transition to
microfracture occurs, the wear rate is unacceptably high, and there is no need to predict the wear rate.
It is, however, crucial to predict the critical load for the onset of contact failure, that is, transition from
mild to severe wear.

36.4.2  Contact Fracture Models


The process of damage formation leading to microfracture in a sliding contact is similar to the types of
damage formed in Hertzian contacts, that is, normal indentation with a spherical body on a flat surface.
Ceramics exhibit either brittle or quasi-plastic behavior above a critical indentation load depending on
the microstructure, indenter radius, or the applied load [43–45].

36.4.2.1  Brittle Behavior


When a spherical solid is pressed against an opposing planar surface, the stress field is initially elastic,
and classical Hertzian stress calculations can be used to analyze the stress distribution within the stressed
half-plane. Beyond a critical load, the material undergoes irreversible deformation and/or fracture. In
brittle materials such as glasses and fine-grain polycrystalline ceramics, a ring crack is initiated in the ten-
sile region just outside the circle of contact produced by the indenting sphere [36]. This crack propagates
downward and outward into the material producing a surface-truncated cone-shaped crack (Figure 36.6).
The critical load, PC , for the formation of cone cracks, that is, brittle behavior, is determined by
assuming that the condition for crack initiation/extension is related to the stress intensity factor associ-
ated with the tensile stresses at the contact circle [45]:

r
Sphere

2a

C C
Y

Specimen

FIGURE 36.6  Types of damage formed in Hertzian contacts with brittle materials; C = cone cracks (brittle
behavior) and Y = plastic yield (quasi plastic behavior). (From Rhee, Y.W. et al., J. Am. Ceram. Soc., 84, 561, 2001.
With permission.)
36-12 Wear Materials

PC /r¢ = AT 2 /E¢ (36.1)

where
r′ is the radius of the ball in contact with a flat surface
T is the fracture toughness (i.e., KIc assuming single-valued toughness)
E′ is the modulus of elasticity

For a contact between two dissimilar materials and two curved surfaces, 1/E′ = 1/E1 + 1/E2 and
1/r′ = 1/r1 + 1/r2. In Equation 36.1, A is related to the Poisson’s ratio and has been obtained by experimen-
tal calibration (A = 8.6 × 103).

36.4.2.2  Quasi-Plastic Behavior


Hertzian indentations produced in coarse grain ceramics depart radically from the classical brit-
tle fracture pattern [43,44]. In these ceramics, the cone crack is suppressed in favor of distributed
damage below the contact. The subsurface damage, which occurs in regions corresponding to high
compression/shear expected for plastic deformation, consists of intragrain twin/slip bands and grain-
localized intergranular microfracture. Microcracks are nucleated by localized microplastic deforma-
tion. For the quasi-plastic mode, the critical load, PY, for onset of failure is calculated assuming that
damage formation is driven by the shear component of the Hertzian stress field and that hardness can
be used as a predictor for plastic yield [45]:

2
PY ⎛H⎞
2
= DH ⎜ ⎟ (36.2)
rʹ ⎝ Eʹ ⎠

where
H is the hardness
The dimensionless constant D is related to the Poisson’s ratio and has been determined experimen-
tally (D = 0.85)

Note that PY depends on r′2 while PC depends linearly on r′.


The transition from traditional brittle behavior (i.e., formation of cone cracks) to a quasi-plastic
behavior (i.e., distributed microfractures) depends on the microstructure of the material being stressed;
for example, in alumina, the transition occurs above a threshold grain size [43]. The competition
between brittle fracture and quasi plasticity can be expressed as a brittleness index: the ratio of critical
loads for quasi-plastic yield and initiation of cone cracks, that is the ratio of Equations 36.2 and 36.1
[45]. For a given ball radius, the response is brittle if PY/PC > 1 and quasi plastic if PY/PC < 1. Smaller ball
radius increases the Hertzian contact stress and promotes plasticity. This is an important phenomenon
since small radii of curvature of asperities in the contact between rough surfaces may promote plasticity
rather than contact fracture at the asperity scale.

36.4.3  Transition Loads in Sliding Contacts


When sliding occurs, the wear process below the critical indentation load is usually controlled by
a plowing mechanism and the wear condition is in the mild wear regime. However, above the criti-
cal load, wear occurs by either brittle fracture or quasi-plastic microfracture process. The brittle
and the quasi-plastic behavior have been alternatively described as “flake formation” and “powder
formation” based on observations of the wear debris [46]. The analyses for the prediction of criti-
cal loads for the two contact damage modes can be modified for sliding contacts to account for the
Wear and Lubrication of Ceramics 36-13

addition of friction. In a practical situation, analysis of stresses in a sliding contact should con-
sider both the stress distribution associated with the apparent contact and the asperity contacts.
Microcontact models [47–50] have shown that pressure at the real area of contact is amplified due
to surface roughness. Rough surfaces produce microcontact pressures that are much larger than the
calculated Hertzian pressure. However, the deviation of microcontact pressure from the Hertzian
pressure calculation is small for polished smooth surfaces [50]. Therefore, the following analysis,
while applicable to contacts between polished surfaces, should be used with caution if applied to
rough surfaces.

36.4.3.1  Brittle Behavior


Explicit equations developed by Hamilton and Goodman [51] for the stress field due to a circular
contact region under a hemispherical normal pressure and tangential traction have been refined and
simplified by Hamilton [52]. For a purely normal loading contact between a spherical body and a
semi-infinite plane, the maximum tensile stress occurs on the circumference of the contact circle.
Sliding with friction adds a compressive stress to the front edge and intensifies the tensile stress
at the trailing edge of the contact. Chiang and Evans [53] used linear elastic fracture mechanics
to analyze the propagation or extension of preexisting microcracks by the tensile component of
stress. Their analysis applies to a homogeneous, isotropic, and linear elastic semi-infinite plane sub-
jected to a combination of normal and tangential surface forces, similar to the stress state solution
of Hamilton and Goodman [51]. The analysis procedure consists of prescribing the state of stress
and assuming that a preexisting surface flaw with a semicircular crack plane is located at the trail-
ing edge of the contact circle. The crack is assumed to be located on a plane normal to the surface.
To predict the condition for crack extension, first the stress intensity factor, K, for such a crack is
determined. It is assumed that the crack propagates when K becomes equal to or exceeds the critical
stress intensity factor, Kc . The stress intensity factor is calculated by integrating the σx component of
prior stress over the crack. (Prior stress is defined as the stress at the crack location, in the absence of
the crack.) Allowing K = Kc for crack extension, an equation is obtained for the critical normal load
for the occurrence of contact fracture. This equation is based on maximum flaw size for unstable
crack growth and calculates a lower bound to the load necessary for fracture. The calculated load
is a linear function of ball radius r and inversely related to the elastic modulus E of the contacting
materials. The critical load Pc for the onset of brittle fracture [54] is similar in form to Equation 36.1:

Pc ⎛ K2 ⎞
= A⎜ c ⎟ g( f ) (36.3)
rʹ ⎝ Eʹ ⎠

The constant A is ∼7.5 × 103. An important feature of this equation is the dependence of contact fracture
load on the coefficient of friction. The form of this relationship based on the dependence of maximum
tensile stress on friction can be written as

−3
g ( f ) = (1 + Bf ) (36.4)

where B depends on the value of Poisson’s ratio. It can vary from 8 to 12 as the Poisson’s ratio increases
from 0.2 to 0.3. The cubic relationship in Equation 36.4 demonstrates a strong influence of friction and
Poisson’s ratio on the critical load for brittle fracture.
The critical load for onset of brittle fracture is sensitive to the fracture toughness (i.e., toughness
raised to the power of two). The value of fracture toughness is highly influenced by the measurement
technique and the initial crack size used for toughness measurement [54,55]. Therefore, published val-
ues of toughness should be used with great caution.
36-14 Wear Materials

It is well established [36] that many polycrystalline ceramics exhibit an R-curve behavior (i.e.,
dependence of toughness on the initial crack size). Experimental data clearly show that while homoge-
neous materials such as glasses exhibit single value toughness, microstructural inhomogeneities such
as coarse grains and second phases produce resistance toward an extending crack and result in an
increase in toughness [56]. In some cases, the toughness is four times higher in the long-crack regime
[57]. Since wear in most structural ceramics occurs by crack propagation at a scale equivalent to only
a few grain diameters, it is clear that the short-crack toughness should be used in the analysis of wear
transition.

36.4.3.2  Quasi-Plastic Behavior


The necessary condition for initiation of plastic deformation can be analyzed by using the stress distri-
bution to calculate the second stress invariant J2. The condition for plastic deformation, as well as the
location for the initiation of plastic flow, is then determined by invoking the von Mises yield criterion.
To analyze the conditions for initiation of plastic flow, the second stress invariant, J2, is first calculated
for a sliding contact, in this case between a ball and a flat surface. The elastic constants, E and ν, are
assumed to be independent of temperature to simplify the calculations. It is also assumed that the yield
stress can be approximated by a value equivalent to one-third of the indentation hardness. This assump-
tion is based on perfectly plastic slip-line field analysis, and there is ample evidence in its support for
metallic alloys [58]. It is used as a first approximation for ceramics, in the absence of other reliable
criteria.
It can be shown that the maximum yield parameter normalized with respect to the maximum contact
pressure (J21/2/P0) raised to the power of three is similar in form to Equation 36.2. Therefore, the transi-
tion load for quasi-plastic deformation with contact friction [54] can be expressed as

2
PY ⎛H⎞
= DH ⎜ ⎟ h( f ) (36.5)
r ʹ2 ⎝ Eʹ ⎠

The constant D depends on the value of Poisson’s ratio and decreases from 0.89 to 0.76 as the Poisson’s
ratio is increased from 0.2 to 0.3. The experimental results of Rhee et al. [45] on several ceramics sug-
gested that D = 0.85, equivalent to using an average Poisson’s ratio of 0.22.
In a sliding contact between a sphere and a flat plane, the location of flow initiation depends on the
magnitude of the coefficient of friction [52]. For f = 0, plastic flow initiates below the contact surface and
moves toward the surface as f increases. At f ≥ 0.33, plastic flow initiates on the surface at the trailing
edge of the contact circle, that is, at the same location as the maximum tensile stress. To a first approxi-
mation, h(f ) in Equation 36.5 can be written as

h ( f ) = 1 for 0 ≤ f ≤ 0.33 (36.6a)


−3
h ( f ) = (3f ) for f ≥ 0.33 (36.6b)

These equations imply that plastic flow is not sensitive to the coefficient of friction when plastic yield
occurs below the surface at 0 ≤ f ≤ 0.33. However, when the location of first yield occurs on the surface
at f ≥ 0.3, the critical load for onset of quasi plasticity decreases with friction.
It is important to note that the critical load for onset of quasi plasticity depends strongly on hardness
(i.e., hardness raised to the power of three). The measured values of hardness are sensitive to the spe-
cific test methods used for their determination and can decrease by 20%–30% as the indentation load
is increased [59]. Therefore, published hardness values should be used with caution, and comparisons
Wear and Lubrication of Ceramics 36-15

between different ceramics should be made only if the hardness has been measured using a consistent
and/or standard test method.
The competition between brittle fracture and quasi plasticity for sliding contacts can be expressed
as a brittleness index, that is, the ratio of Equations 36.5 and 36.3, similar to the approach used for
Hertzian contacts. The response is brittle if PY/PC > 1 and quasi plastic if PY/PC < 1. Note that brittle
behavior is promoted for a larger r, smaller toughness, lower elastic modulus, and higher hardness.
Quasi-plastic behavior is promoted as the coefficient of friction or the Poisson’s ratio is increased when
f ≤ 0.33. However, as the coefficient of friction increases, its influence diminishes relative to the effect of
Poisson’s ratio.
Similar analyses have been used to predict the onset of contact fracture in sliding by Hokkirigawa
[46] and Adachi et al. [11]. The proposed parameters that describe the severity of contact Sc and Sc∗ for the
prediction of “flake formation” (i.e., brittle fracture) and “powder formation” (i.e., quasi-plastic micro-
fracture), respectively, are based on hardness, fracture toughness, and the coefficient of friction. These
analyses have been extended to develop two contact severity parameters for failure induced by contact
stress and by thermal strain:

Sc,m =
((1 + 10m ) P max d 1/ 2 ) (36.7)
K Ic

1/ 2
⎛ gm ⎞ ⎛ nWHV ⎞
Sc,t = ⎜ (36.8)
⎝ DTs ⎟⎠ ⎜⎝ k rc ⎟⎠

where
μ is the coefficient of friction
Pmax is the maximum Hertzian contact pressure
K Ic is the fracture toughness
γ is the heat partition ratio
ΔTs is the thermal shock resistance
ν is the Poisson’s ratio
W is the normal load
HV is the Vickers harness
k is the thermal conductivity
ρ is the density
c is the specific heat

(In these equations, the original symbols used in the publication are used rather than symbols consis-
tent with those in this chapter to preserve the original forms of the equations.) The threshold values
for these contact severity parameters have been determined experimentally (Sc,m = 6 and Sc,t = 0.04) [11].
The critical fracture loads and the contact severity parameters provide estimates for prediction of wear
transition from mild to severe wear.

36.5  Boundary Lubrication


To maintain low friction and avoid severe wear, lubrication is often required. Tribochemical reactions
between ceramic surfaces and water vapor provide a natural opportunity for lubrication of ceramics
[60]. However, the adsorption tendency of water is reduced as the temperature is increased above 200°C
and the formation of oxides is favored over hydroxides. The temperature limit may be shifted to lower
temperatures at larger loads and higher speeds, due to increased contact temperatures.
36-16 Wear Materials

36.5.1  Lubrication with Hydrocarbon Fluids


The coefficient of friction can be controlled through boundary lubrication due to adsorption of hydro-
carbons or chemical reactions of certain chemical compounds with sliding surfaces [61–65]. Other
schemes used for lubrication of ceramics include solid lubricating coatings [66–68], supply of reacting
lubricants through a gas stream [69,70], deposition of carbonaceous layers by catalytic decomposition of
hydrocarbons [71–74], and self-lubricating ceramic matrix composites [75,76]. Pure hydrocarbon fluids
(e.g., hexadecane) and boundary lubricant additives (e.g., stearic acid) are effective in reducing the coef-
ficient of friction (to 0.10 or lower) and reducing the wear coefficient (to 10−5 or lower) by preventing the
microfracture wear mode [17]. The transition from mild to severe wear, however, can still occur under
certain conditions with hydrocarbon fluids [77]. An abrupt increase in wear rate occurs as the load
reaches a critical value that depends on the sliding speed. Observation of the wear scars before and after
this transition confirmed that severe wear had occurred through a microfracture process similar to the
unlubricated tests. The transition load was higher for the low-speed lubricated contacts compared with
unlubricated tests. The transition load decreased from 60 to 20 N as the rotational speed was increased
from 8 rpm (0.0031 m/s) to 1000 rpm (0.38 m/s). The reduction in the transition load is possibly due to
thermoelastic stresses as the contact temperature increases at higher speeds.

36.5.2  Lubrication with Water


Water, under certain sliding conditions, is also an excellent lubricant for silicon nitride and silicon
carbide. Tomizawa and Fischer [78] presented convincing results that the coefficient of friction for self-
mated silicon nitride in water decreases to a value <0.002 at sliding speeds higher than 60 mm/s. Since
the worn surfaces were extremely smooth, they attributed the low friction to the initiation of hydrody-
namic lubrication by a thin water film at the interface. Since no wear debris was observed in the water,
they suggested that wear of silicon nitride occurred through tribochemical reactions and dissolution
of reaction products (i.e., (SiOx) nH2O) in water. The formation of tribochemical reaction products on
the wearing surfaces and the role of the hydrated oxides on the tribological properties of silicon nitride
have been confirmed by several investigations [16,17,79]. While silicon nitride and silicon carbide are
lubricated with water, zirconia, and particularly PSZ, is negatively affected in water exhibiting high
coefficients of friction and wear rates [80,81].
The unique low friction behavior of silicon nitride has been confirmed by Xu et al. [82] and Chen
et al. [83] in a series of studies that included a narrow range of loads, speeds, and temperatures. They
reported a coefficient of friction below 0.03 after the run-in period. The low friction was only obtained
at higher speeds and lower loads. The coefficient of friction remained high at about 0.5 when the water
temperature was raised above 50°C [84]. The low friction was attributed to a combination of hydrody-
namic lubrication due to the smooth surfaces and boundary lubrication due to the presence of colloidal
silica on the wearing surfaces and was confirmed by introducing an additive to water that accelerated
the formation of silica and reduced the run-in period [84].
Data by Kato and Adachi [85] showed that the initial surface roughness must be 10 nm or less for
the low friction behavior to occur. Saito et al. [86] have shown that the run-in period necessary for the
transition from high friction to low friction behavior for self-mated silicon nitride depends on both the
surface roughness and sliding speed. The run-in period is shorter for a lapped surface compared to a
surface finished by grinding. They also showed that the friction remains high (about 0.8) at speeds above
130 mm/s for both surface finishes. The low friction behavior of silicon nitride in water, however, can be
unstable and occasional high friction spikes can occur under certain sliding conditions [87]. Addition
of ionic liquids to water can reduce the occurrence of the high friction spikes and decrease the required
running-in period [88]. The low coefficient of friction can also be achieved for silicon nitride sliding
against silicon carbide in steam, although the coefficient of friction increases to a high value when the
speed is raised above 1000 mm/s [89].
Wear and Lubrication of Ceramics 36-17

36.6  Design with Ceramics


Implementation of ceramics in engineering applications requires experience in designing with brit-
tle materials, design data, field performance data, and performance and failure prediction methods.
Selection of materials for tribological applications should be based not only on friction and wear behav-
ior but also on other application requirements such as strength, fatigue resistance, corrosion resistance,
dimensional stability, thermal properties, reliability, ease of fabrication, and cost.
A common feature for most ceramics in tribological applications is their excellent wear resistance—
often three to four orders of magnitude better than that of metals. This advantage, however, is lost at
higher contact loads because of severe wear and larger coefficients of friction. The transition from mild
to severe wear, for a given material, depends on the load, sliding velocity, and temperature, as well as on
the coefficient of friction, which can be modified by tribochemical reactions and lubrication. Therefore,
design of ceramic tribological contacts requires data regarding their tribological behavior and models
to predict the transition from mild to severe wear.
Typical values of coefficients of friction and wear rates for different ceramics in the mild wear regime
are listed in Table 36.2. The wear rate is listed as wear constant or specific wear rate in units of mm3/Nm,
which is wear volume divided by load and sliding distance. It is generally accepted that in the mild wear
regime the wear constant is <10−6 (or the wear coefficient <10−4). The data in Table 36.2 have been obtained
from unlubricated self-mated tests at room temperature. Note that the coefficients of friction associated
with these tests are generally high; however, wear is in the mild wear regime. For comparison, the wear
constants for different steels or aluminum bronze sliding against steel are higher and range from 10−5 to
10−4 [95] under similar unlubricated conditions. Since ceramics undergo transition from mild to severe
wear at higher loads and sliding speeds, it is important to design for sliding in the mild wear regime.
Reliable data and fundamental knowledge on the mechanisms of friction and wear are essential for

TABLE 36.2  Typical Coefficient of Friction and Wear Rates for Different Ceramics in the
Mild Wear Regime
Coefficient of Wear Constant Sliding
Material Friction (10−8 mm3/N·m) Speed (m/s) Load (N) Reference
Alumina
99.8 0.40 4 0.001 10–80 [10]
99.5 0.30 0.04 0.01 10 [90]
85 0.70 3 0.25 47 [91]

Zirconia
Cubic 0.70 200 0.001 10 [20]
PSZ 0.88 10 0.25 1–23 [92]
TZP 0.35 0.2 0.001 10 [20]

Silicon nitride
Sintered 0.42 4 0.20 10 [93]
Hot-pressed 0.60 20 0.25 47 [91]
Reaction-bonded 0.21 5 0.20 10 [93]
HIP’ed 0.30 100 0.001 1–10 [16]

Silicon carbide
Reaction-bonded 0.24 90 1.5 200–400 [94]
Hot-pressed 0.60 20 0.25 47 [91]
Sintered 0.63 50 0.001 1–10 [19]
Note: Room temperature, self-mated, ball on disk tests. Wear constant less than 10−6 is considered
as “mild wear.”
36-18 Wear Materials

developing guidelines for design and material selection. The wear transition diagrams and wear maps
serve as simple and easy-to-use tools for predicting the performance of different ceramics under ser-
vice conditions. These diagrams clearly show the varied tribological performance based on different sets
of fundamental mechanisms, and alert the designer to the conditions that could promote catastrophic
failures.
The mechanics analyses can be used to predict the onset of severe wear in ceramics under low-speed
sliding. It is, however, important to recognize the limitations of such analyses. The contact model is based
on the assumption of “smooth” contacts, that is, the presence of asperities is not included. The analysis
should be modified for rougher surfaces. The contact stress calculations are based on the assumption
of a homogeneous elastic material; it may not apply to composites and layered materials (e.g., films and
coatings). The model should be modified to include the effect of thin layers (e.g., tribochemical film)
on stress distributions below the contact. The analysis does not account for thermoelastic stresses that
may develop at high sliding speeds, nor does it consider the influence of residual stresses either due to
processing or developed during sliding. It is also important to note that transition to severe wear by
microfracture could occur by extension of several smaller cracks or by interactions between surface and
subsurface microcracks rather than by extension of a single crack, thus making the precise prediction
of contact fracture more complex.
Often ceramic components are mated with metallic counterparts. In such cases, the wear process
is governed by transfer of a metallic film onto the ceramic surface. Depending on the environment
and the sliding conditions, the metallic film can be subjected to oxidation and usually contains a
mix of material from both surfaces and oxide particles. The wear of transfer films also includes
delamination and abrasion by the ceramic surface (or wear debris). Wear rates and the stability of
such contacts are usually difficult to predict due to the complex nature of the transfer and mixing
process. It is a common practice to lubricate the metal–ceramic contacts to control the wear rate and
metal transfer.

References
1. D. W. Richerson, Modern Ceramic Engineering, Marcel Dekker, New York, 1982.
2. M. Schwartz, Handbook of Structural Ceramics, McGraw Hill, New York, 1992.
3. S. Jahanmir, Friction and Wear of Advanced Ceramics, Marcel Dekker, New York, 1993.
4. S. Jahanmir and T. E. Fischer, Friction and wear of ceramics, in Handbook of Lubrication and
Tribology, Vol. 3, E. R. Booser (Ed.), CRC Press, Boca Raton, FL, pp. 103–120, 1994.
5. K. Kato, Tribology of ceramics, Wear, 136, 117–133, 1990.
6. T. E. Fischer and H. Tomizawa, Interaction of tribochemistry and microfracture in the friction and
wear of silicon nitride, Wear, 105, 29–42, 1985.
7. J. F. Archard, Contact of rubbing surfaces, J. Appl. Phys., 24, 981–988, 1953.
8. S. C. Lim and M. F. Ashby, Wear mechanism maps, Acta Metall., 35, 1–24, 1987.
9. H. Kong and M. F. Ashby, Wear mechanisms in brittle solids, Acta Metall., 40, 2907–2920, 1992.
10. S. Jahanmir and X. Dong, Mechanism of mild to severe wear transition in alumina, J. Tribol., 114,
403–411, 1992.
11. K. Adachi, K. Kato, and N. Chen, Wear map of ceramics, Wear, 203–204, 291–301, 1997.
12. S. M. Hsu and M. C. Shen, Ceramic wear maps, Wear, 200, 154–175, 1996.
13. Y. M. Chen, B. Rigaut, and F. Armanet, Friction and wear of alumina ceramics at high sliding speed,
Lubr. Eng., 47, 531–537, 1991.
14. M. Woydt and K. H. Habig, High temperature tribology of ceramics, Tribology Int., 89, 75–88,
1989.
15. S. J. Cho, B. J. Hockey, B. R. Lawn, and S. J. Bennison, Grain-size and R-curve effects in the abrasive
wear of alumina, J. Am. Ceram. Soc., 72, 1249–1252, 1989.
16. X. Dong and S. Jahanmir, Wear transition diagram for silicon nitride, Wear, 165, 169–180, 1993.
Wear and Lubrication of Ceramics 36-19

17. S. Jahanmir and T. E. Fischer, Friction and wear of silicon nitride lubricated by humid air, water,
hexadecane, and hexadecane +0.5 percent stearic acid, ASLE Trans., 31, 32–43, 1988.
18. A. Skopp, M. Woydt, and K. H. Habig, Lubricated sliding friction and wear of various silicon nitride
pairs between 22 and 1000°C, Tribology Int., 23, 189–199, 1990.
19. X. Dong and S. Jahanmir, Wear transition diagram for silicon carbide, Tribol. Int., 28, 559–572, 1995.
20. T. E. Fischer, M. O. Anderson, and S. Jahanmir, Influence of fracture toughness on the wear resis-
tance of yttria-doped zirconium oxide, J. Am. Ceram. Soc., 72, 252–257, 1988.
21. R. H. J. Hannink, M. J. Murray, and H. G. Scott, Friction and wear of partially stabilized zirconia,
Wear, 100, 355–366, 1984.
22. W. M. Rainforth, The sliding wear of ceramics, Wear, 22, 365–372, 1996.
23. Y. Wang, F. J. Worzala, and A. R. Lefkow, Friction and wear properties of partially stabilized zirconia
with solid lubricant, Wear, 167, 23–31, 1993.
24. S. W. Lee, S. M. Hsu, and M. C. Shen, Ceramic wear maps: Zirconia, J. Am. Ceram. Soc., 76, 1937–
1947, 1993.
25. G. W. Stachowiak and G. B. Stachowiak, Unlubricated wear and friction of toughened zirconia
ceramics at elevated temperatures, Wear, 143, 277–295, 1991.
26. H. G. Scott, Friction and wear of zirconia at very low sliding speed, in Proceedings of Wear of
Materials Conference, K. C. Ludema (Ed.), ASME, New York, pp. 8–12, 1985.
27. Y. M. Chen, B. Rigaut, and F. Armanet, Wear behavior of partially stabilized zirconia at high sliding
speeds, J. Eur. Ceram. Soc., 6, 383–390, 1990.
28. G. W. Stachowiak and G. B. Stachowiak, Unlubricated friction and wear behavior of toughened
zirconia ceramics, Wear, 132, 151–171, 1989.
29. T. E. Fischer, M. P. Anderson, R. Salher, and S. Jahanmir, Friction and wear of tough and brittle
zirconia in nitrogen, air, water, hexadecane, and stearic acid, Wear, 122, 133–148, 1988.
30. R. N. Katz, Ceramic materials for rolling element bearing application, in Friction and Wear of
Advanced Ceramics, S. Jahanmir (Ed.), Marcel Dekker, New York, pp. 313–328, 1993.
31. J. F. Braza, H. S. Cheng, and M. E. Fine, Silicon nitride wear mechanisms: Rolling and sliding con-
tact, Tribol. Trans., 32, 439–446, 1989.
32. A. Gangopadhyay, H. S. Cheng, J. F. Braza, S. Harman, and J. M. Corwin, Tribological performance
of ceramic cam roller followers, in Friction and Wear of Advanced Ceramics, S. Jahanmir (Ed.),
Marcel Dekker, New York, pp. 329–356, 1993.
33. S. S. Kim, K. Kato, K. Hokkirigawa, and H. Abe, Wear mechanism of ceramic materials in dry rolling
friction, J. Tribol., 108, 522–528, 1986.
34. M. Akazawa, K. Kato, and K. Umeya, Wear properties of silicon nitride in rolling contact, Wear, 100,
285–293, 1986.
35. L. Y. Chao, R. Lakshminarayanan, N. Iyer, G. Y. Lin, and D. K. Shetty, Transient wear of silicon
nitride in lubricated rolling contact, Wear, 223, 58–65, 1998.
36. B. R. Lawn, Fracture of Brittle Solids, Cambridge University Press, Cambridge, U.K., 1993.
37. B. R. Lawn, A model for the wear of brittle solids under fixed abrasive conditions, Wear, 33,
369–372, 1975.
38. H. H. K. Xu and S. Jahanmir, Transition in the mechanism of material removal in abrasive wear of
alumina, Wear, 192, 228–232, 1996.
39. A. G. Evans and D. B. Marshall, Wear mechanisms in ceramics, in Fundamentals of Friction and
Wear of Materials, D. A. Rigney (Ed.), American Society for Metals, Metals Park, OH, pp. 439–452,
1981.
40. C. P. Dog and J. A. Hawk, Role of composition and microstructure in the abrasive wear of high
alumina ceramics, Wear, 225, 1050–1058, 1999.
41. C. P. Dog and J. A. Hawk, Microstructure and abrasive wear in silicon nitride ceramics, Wear, 250,
256–263, 2001.
42. Y. Wang and S. M. Hsu, Wear and wear transition modeling of ceramics, Wear, 195, 35–46, 1996.
36-20 Wear Materials

43. F. Guiberteau, N. P. Padture, and B. R. Lawn, Effect of grain size on Hertzian contact in alumina,
J. Am. Ceram. Soc., 77, 1825–1831, 1994.
44. H. H. K. Xu, L. Wei, N. P. Padture, B. R. Lawn, and R. L. Yeckley, Effect of microstructural coarsening
on Hertzian contact damage in silicon nitride, J. Mater. Sci., 30, 869–878, 1995.
45. Y. W. Rhee, H. W. Kim, Y. Deng, and B. R. Lawn, Brittle fracture versus quasi plasticity in ceramics:
A simple predictive index, J. Am. Ceram. Soc., 84, 561–565, 2001.
46. K. Hokkirigawa, Wear mode map of ceramics, Wear, 151, 219–228, 1991.
47. E. Ioannides and J. C. Kuijpers, Elastic stresses below asperities in lubricated contacts, J. Tribol., 108,
394–402, 1986.
48. J. I. McCool, Comparison of models for the contact of rough surfaces, Wear, 107, 37–60, 1986.
49. W. R. Chang, I. Etsion, and D. B. Bogy, An elastic-plastic model for the contact of rough surface,
J. Tribol., 109, 257–263, 1987.
50. T. Merriman and J. Kannel, Analyses of the role of surface roughness on contact stresses between
elastic cylinders with and without soft surface coating, J. Tribol., 111, 87–89, 1989.
51. G. M. Hamilton and L. E. Goodman, The stress field created by a circular sliding contact, J. Appl.
Mech., 88, 371–376, 1966.
52. G. M. Hamilton, Explicit equations for the stress beneath a sliding spherical contact, Proc. Inst.
Mech. Eng., 197C, 53–59, 1983.
53. S. Chiang and A. G. Evans, Influence of a tangential force on the fracture of two contacting elastic
bodies, J. Am. Ceram. Soc., 66, 4–10, 1983.
54. S. Jahanmir, Wear transitions and tribochemical reactions in ceramics, J. Eng. Tribol., Proc. Inst.
Mech. Eng., Part J, 216, 371–385, 2002.
55. G. D. Quinn, J. A. Salem, I. Bar-On, and M. G. Jenkins, The new ASTM fracture toughness of ceram-
ics standard: PS 070-97, Ceram. Eng. Sci. Proc., 19, 565–579, 1998.
56. S. K. Lee, S. Wuttiphan, and B. R. Lawn, Role of microstructure in Hertzian contact damage in sili-
con nitride: Mechanical characterization, J. Am. Ceram. Soc., 80, 2367–2381, 1997.
57. H. H. K. Xu, S. Jahanmir, L. K. Ives, L. S. Job, and K. Ritchie, Short-crack toughness and abrasive
machining of silicon nitride, J. Am. Ceram. Soc., 79, 3055–3064, 1996.
58. D. Tabor, Indentation hardness and its measurement, in Microindentation Techniques in Materials
Science and Engineering, ASTM-STP 889, P. J. Blau and B. R. Lawn (Eds.), American Society for
Testing and Materials, Philadelphia, PA, pp. 129–159, 1985.
59. G. D. Quinn, Indentation hardness testing of ceramics, in Materials Testing and Evaluation, Vol. 8,
Mechanical Testing, ASM International, Metals Park, OH, pp. 244–255, 2001.
60. T. E. Fischer, H. Liang, and W. M. Mullins, Tribochemical lubricious oxides on silicon nitride, Mater.
Res. Soc. Symp. Proc., 140, 339–345, 1989.
61. P. Studt, Boundary lubrication: Adsorption of oil additives on steel and ceramic surfaces and its
influence on friction and wear, Tribol. Int., 22, 111–119, 1989.
62. X. Z. Zhao, J. J. Liu, and T. E. Fischer, Effects of lubricant rheology and additive chemistry in the
wear of Si3N4 sliding on steel, Wear, 223, 37–43, 1998.
63. R. S. Gates and S. M. Hsu, Effect of selected chemical compounds on lubrication of silicon nitride,
Tribol. Trans., 34, 417–425, 1991.
64. R. S. Gates and S. M. Hsu, Silicon nitride boundary lubrication: Effect of phosphorus-containing
organic compounds, Tribol. Trans., 39, 795–802, 1996.
65. R. S. Gates and S. M. Hsu, Silicon nitride boundary lubrication: Effect of sulfonate, phenate, and
salicylate compounds, Tribol. Trans., 43, 269–274, 2000.
66. A. Erdemir, R. Erck, G. R. Fenske, and H. Hong, Solid/liquid lubrication of ceramics at elevated
temperatures, Wear, 203, 588–595, 1997.
67. A. Erdemir, D. E. Busch, R. A. Erck, G. R. Fenske, and R. Lee, Ion-beam-assisted deposition of silver
films on zirconia ceramics for improved tribological behavior, Lubr. Eng., 47, 863–872, 1991.
Wear and Lubrication of Ceramics 36-21

68. A. Erdemir, Lubrication of ceramics with thin solid films, in Friction and Wear of Ceramics,
S. Jahanmir (Ed.), Marcel Dekker, New York, pp. 119–162, 1993.
69. J. C. Smith, M. J. Furey, and C. Kajdas, An exploratory study of vapor phase lubrication of ceramics
by monomers, Wear, 181, 581–593, 1995.
70. E. E. Klaus, G. S. Jeng, and J. L. Dudda, A study of tricresyl phosphate as vapor delivered lubricant,
Lubr. Eng., 45, 717–723, 1989.
71. J. L. Lauer and S. R. Dwyer, Tribochemical lubrication of ceramics by vapors, Tribol. Trans., 34,
521–528, 1991.
72. J. L. Lauer and S. R. Dwyer, Continuous high temperature lubrication of ceramics by carbon gener-
ated catalytically from hydrocarbon gases, Tribol. Trans., 33, 529–534, 1990.
73. N. J. Barnick, T. A. Blanchet, W. G. Sawyer, and J. E. Gardiner, High temperature lubrication of vari-
ous ceramics and metal alloys via directed hydrocarbon feed gases, Wear, 214, 131–138, 1998.
74. J. L. Lauer and B. G. Bunting, High temperature solid lubrication by catalytically generated carbon,
Tribol. Trans., 31, 338–349, 1988.
75. A. Gangopadhyay and S. Jahanmir, Friction and wear characteristics of silicon nitride–graphite and
alumina–graphite composites, Tribol. Trans., 34, 257–265, 1991.
76. A. Gangopadhyay and S. Jahanmir, Self-lubricating ceramic matrix composites, in Friction and Wear
of Ceramics, S. Jahanmir (Ed.), Marcel Dekker, New York, pp. 163–198, 1993.
77. D. E. Deckman and S. Jahanmir, Wear mechanisms of alpha alumina lubricated with a paraffin oil,
Wear, 149, 155–168, 1991.
78. H. Tomizawa and T. E. Fischer, Friction and wear of silicon nitride and silicon carbide in water:
Hydrodynamic lubrication at low sliding speed obtained by tribochemical wear, ASLE Trans., 30,
41–46, 1987.
79. M. Woydt and J. Schwenzien, Dry and water-lubricated slip-rolling of Si3N4 and SiC-based ceramics,
Tribol. Int., 26, 165–174, 1993.
80. P. Andersson, Water lubricated pin-on-disk tests with ceramics, Wear, 154, 37–47, 1992.
81. J. K. Lancaster, A review of the influence of environmental humidity and water on friction, lubrica-
tion, and wear, Tribol. Int., 23, 371–389, 1990.
82. J. Xu, K. Kato, and T. Hirayama, The transition of wear mode during the running in process of sili-
con nitride sliding in water, Wear, 230, 55–63, 1997.
83. M. Chen, K., Kato, and Adachi, The comparisons of sliding speed and normal load effect on friction
coefficients of self-mated Si3N4 and SiC under water lubrication, Tribol. Int., 35, 129–135, 2002.
84. J. Xu and K. Kato, Formation of tribochemical layer of ceramics sliding in water and its role for low
friction, Wear, 245, 61–75, 2000.
85. K. Kato and K. Adachi, Wear of advanced ceramics, Wear, 253, 1097–1104, 2002.
86. T. Saito, T. Hosoe, and F. Honda, Chemical wear of sintered Si3N4, hBN and Si3N4–hBN composites
by water lubrication, Wear, 247, 223–230, 2001.
87. S. Jahanmir, Y. Ozmen, and L. K. Ives, Water lubrication of silicon nitride in sliding, Tribol. Lett., 17,
409–417, 2004.
88. B. S. Phillips and J. S. Zabinski, Ionic liquid lubrication effects on ceramics in water environment,
Tribol. Lett., 17, 1573–2711, 2004.
89. H. Heshmat and S. Jahanmir, Tribological behavior of ceramics sliding at high speeds in steam,
Tribol. Lett., 17, 359–366, 2004.
90. H. Kim, D. Shin, and T. E. Fischer, Mechanical and chemical aspects in the wear of alumina, in
Proceedings of the Japan International Tribology Conference, October 29–November 1, Nagoya,
Japan, pp. 1473–1479, 1990.
91. O. Ajayi and K. C. Ludema, Surface damage of structural ceramics, in Wear of Materials, K. C.
Ludema (Ed.), American Society of Mechanical Engineers, New York, pp. 349–359, 1987.
92. J. Denape and J. Lamon, Sliding friction of ceramics, J. Mater. Sci., 25, 3592–3598, 1990.
36-22 Wear Materials

93. M. Gee, C. S. Matharu, E. A. Almond, and T. S. Eyre, The measurement of sliding friction and wear of
ceramics at high temperature, Wear, 138, 169–175, 1990.
94. D. C. Cranmer, Friction and wear properties of monolithic silicon-based ceramics, J. Mater. Sci., 20,
2029–2036, 1985.
95. A. W. Ruff, Typical friction and wear data, in Tribology Data Handbook, E. R. Booser (Ed.), CRC
Press, New York, pp. 435–444, 1997.
37
Composite Materials
37.1 Introduction..................................................................................... 37-1
37.2 Wear Testing for Polymer Composites......................................... 37-1
37.3 Developing Wear-Resistant Polymeric Composites by
Using Conventional Fillers............................................................. 37-3
37.4 Wear of Polymeric Nanocomposites............................................ 37-6
Nanoparticle-Reinforced Polymer Composites  •  Carbon-
Li Chang Nanotube-Reinforced Polymer Composites  •  Hybrid
University of Sydney Nanocomposites for High Wear Resistance
37.5 Summary......................................................................................... 37-11
Klaus Friedrich Acknowledgments..................................................................................... 37-11
Technical University
Kaiserslautern References................................................................................................... 37-11

37.1  Introduction
Composite materials have been one of the most rapidly growing classes of engineering materials, com-
prising polymer-, metal-, glass-, and ceramic-based composites. The wide variety of different matrix and
filler materials permits the design of composites with unique properties for different kinds of applica-
tions. In the context of tribology, exactly how the composite materials must be designed depends on the
requirement profile of the particular application. Sometimes a high coefficient of friction, coupled with
low wear, is required (e.g., for brake pads or clutches). In most cases, however, it is of primary concern
to develop wear-resistant composites that possess low friction and low wear (e.g., for gears or bearings).
The principal composition of typical wear-resistant composite materials is demonstrated in Figure 37.1.
Among these diverse composite materials, advanced polymer-based composites have attracted con-
siderable interest in recent decades because of their superior property profiles, such as high specific
strength (strength/density) and high specific stiffness (modulus/density), self-lubrication and super
cleanliness, the ability to dampen shock and vibration, corrosion resistance, ease in manufacturing, and
low unit cost. These unique features make polymer composites an ideal candidate for high-performance
energy-efficient triboapplications in aerospace, automotive, and chemical industries. In many of these
applications, polymer composites are used as highly loaded triboelements sliding against a metallic
counterpart, for example, as in seals, bearings, and brake pads [1,2]. The main objective of this chapter is
to illustrate how to design such wear-resistant polymer composites, using different combinations of fill-
ers and reinforcements. More discussions on the design of other wear-resistant composites (e.g., metal-
and ceramic-based composites, as well as natural fiber composites) can be found in references [3–6].

37.2  Wear Testing for Polymer Composites


To characterize the sliding wear resistance of polymeric materials in the laboratory, the pin-on-disk
and block-on-ring are two commonly used tribomachines for wear testing. In many tribological

37-1
37-2 Wear Materials

Wear resistant composite materials

Continuous phase Reinforcing phase


(matrix) (fillers)

Fibrous fillers Particulate fillers


Polymer Glass Metal Ceramic

Thermosetting Continuous
Thermoplastic
(TS) (architecture) Short fibers
(TP) Internal Other particle fillers
Elastomer lubricants
Epoxy (incl. Nanoparticles)
POM
Random
polyimide (EL) PA6 Unidirectional Textile orientation
phenolic rubber PEEK Graphite TiO2
TPE PTFE PTFE SiC
MoS2 ZnS
Bronze

FIGURE 37.1  Systematic illustration of the structural components of wear-resistant composite materials.

experiments, metallic counterparts are used, often consisting of hardened and polished carbon steel
(German Standard 100 Cr6). The latter is the most frequently used steel for ball bearing applications.
Laboratory wear tests allow ranking of the most important tribological property of these materials, that
is, the specific wear rate, which is determined by the following equation:

Δm
ws = ⎡mm 3 / ( N ⋅ m )⎤ (37.1)
rFN L ⎣ ⎦

where
FN is the normal load applied on the specimen during sliding
Δm is the specimen’s mass loss
ρ is the density of the specimen
L is the total sliding distance

The specific wear rate can be considered as a kind of material property (as a basis for comparison
between different materials), as long as changes in contact pressure and sliding velocity do not influence
this value to a great extent. However, friction and wear are, in general, not real material properties but
depend strongly on the system in which the materials have to function [7]. When the pressure and/or
velocity increase under certain conditions the specific wear rate of a material may also increase signifi-
cantly, due to increases in the contact temperature and/or changes in the dominant wear mechanisms
[8]. To describe the wear behavior of materials under different sliding conditions, the time-related depth
wear rate, wt, is frequently used:

Δh
wt = k ∗⋅ pn = [nm/s ] (37.2)
t

with

k ∗ = w s ⎡⎣mm 3 / ( N ⋅ m )⎤⎦ (37.3)



Composite Materials 37-3

Improvement with

ws = k* = f ( p, v) • Higher temperature
polymer matrix
• Fiber reinforcement
• Internal lubricants
• Additional fillers, e.g.,
bronze or nanoparticles
wt (µm/h)

Scatter due to different


influences of p vs. v

(p·v)lim

( p·v)crit
p·v (MPa·m/s)

FIGURE 37.2  Schematic illustration of the course of the depth wear rate vs. pv-product, and the general tar-
gets in designing wear-resistant polymer composites, that is, to reduce the wear factor, k*, and to improve the
pv limit, (pv)lim . (Reprinted with kind permission from Springer Science + Business Media: Composite Materials,
Current and future applications of polymer composites in the field of tribology, 2011, 129–167, K. Friedrich,
L. Chang, and F. Haupert.)

where
k* is called the wear factor of the material, which is numerically the same as the specific wear rate
determined by Equation 37.1
p is the nominal pressure (normal load divided by the apparent contact area)
v is the sliding velocity
t is the test time
Δh is the height loss of the specimen

Depth wear rate is also a convenient measure for the design engineer to estimate the height loss of a
polymeric material in various pressure and velocity conditions (assuming that the wear rate of the steel
counterpart is negligible). This, in turn, can help to calculate the development of a certain clearance in
journal bearings with time and, thus, to predict necessary maintenance intervals in order to prevent
more catastrophic failures of complete machine units.
According to Equation 37.2, the time-related depth wear rate can be considered as a function of the
pv factor, resulting in two tribological parameters for evaluating the load-carrying capacity of bearing
materials: (1) the basic wear factor, k*, which remains almost constant within a certain range of pv fac-
tor, and (2) the “limiting pv,” above which the increase in the wear rate of materials is too rapid, so that
the material can no longer be employed. Such pv-dependent wear behaviors are schematically illustrated
in Figure 37.2. The general objectives in the design of wear-resistant polymer composites are to reduce
the slope of these curves, that is, the basic wear factor, and to enhance the limiting pv [9].

37.3 Developing Wear-Resistant Polymeric


Composites by Using Conventional Fillers
Nowadays, the use of polymeric materials in various tribological situations has become state of the art.
Nevertheless, many polymers cannot fulfill the tribological requirements in their pristine or bulk forms,
due to their relatively weak load-carrying capacity and poor durability, especially under extreme sliding
conditions. To overcome this inherent weakness, various fillers have been used to improve their friction
37-4 Wear Materials

PTFE
particles

PEEK
matrix
Carbon
fibers
100 µm

FIGURE 37.3  Light optical micrograph of a typical microstructure in wear-resistant polymer composites, for
example, of a PEEK-based composite reinforced by short carbon fibers and PTFE particles. (Reprinted with kind
permission from Springer Science + Business Media : Composite Materials, Current and future applications of
polymer composites in the field of tribology, 2011, 129–167, K. Friedrich, L. Chang, and F. Haupert.)

and wear behavior (see Figure 37.1). The integration of different types of fillers results in a complex micro-
structure in such composites, sometimes called hybrid composites. For instance, Figure 37.3 shows the
typical microstructure of a wear-resistant polymer composite. Under optimum composition, the wear
rate of the composite can be nearly 1 order in magnitude less than that of the pure polymer matrix [10].
The essence for the development of such hybrid composites is to provide synergistic actions between the
components during the wear process, which requires an in-depth understanding of the tribological role
of individual fillers. In the following paragraphs, the tribological effects of two important types of conven-
tional fillers, fibrous reinforcements and solid lubricants, are briefly discussed.
It is known that fibrous reinforcements can significantly improve the strength and the load-carrying
capacity of polymer composites, owing to their high aspect ratio (length-to-width ratio) that allows a
good transfer of load from the soft matrix to the hard fillers. The diameter of these fibers is typically in
the micrometer-size range. However, their length can vary between ∼100 μm (short fibers) and several
millimeters (long fibers) and up to meters in case of continuous fiber composites. In general, aligned,
continuous fiber-reinforced polymer composites can give good tribological service, for example, low
wear rates with high mechanical strength. This is not only because of the very high aspect ratio of con-
tinuous fibers but also because of the attainable high fiber content of continuous fiber-filled systems.
However, the extra cost entailed by the complex processing methods, for example, the slow autoclave
curing process, limits the use of this type of composite. In practice, the majority of fiber-reinforced
polymeric tribocomponents are made of discontinuous short-fiber systems, since they allow more
complex geometries, for example, by injection molding and better tailoring of their tribological char-
acteristics for particular applications. Among the commonly used fibers, such as glass fiber (GF),
aramid fiber (AF), and carbon fiber (CF), CF is normally considered to be most effective in improv-
ing the wear resistance of various polymers [11–13]. The beneficial effects of CFs have been attributed
to their excellent mechanical properties and their ability to transfer a smooth carbon film onto the
counterface [14,15]. Besides the fiber material, the wear performance of FRPs is also influenced by the
fiber orientation with respect to the sliding plane [16] and by the fiber/matrix interfacial bonding [17].
When designing high-wear-resistant FRPs, it is necessary to take all these factors into consideration
and to combine the favorable factors. For example, in a comparative study of sliding wear of unidirec-
tional continuous AF, CF, and GF composites, CF proved to be more favorable under in-plane parallel
orientation, whereas AF showed better results under normal orientation to the sliding plane [18].
Composite Materials 37-5

In polymer tribology, solid lubricants normally refer to particulate fillers with a lubricating effect,
that is, they can reduce the adhesion between the polymer composite and the metallic counterpart.
In particular, some lubricating fillers can also enhance the self-lubricating ability of the polymer matrix,
leading to the formation of a high-performance transfer film on the counterpart [19]. In this case, the
continuous transfer film acts as a solid “lubrication,” which normally contributes to low friction and
low wear of the polymer-on-metal sliding system [19,20]. For instance, it has been reported that with the
addition of CuO or CuS, polyamide 11 can form a uniform transfer film on a steel surface, as opposed to
the patchy character of the transfer film formed by unfilled polyamide [21]. As a result, the wear rate of
the filled polyamide 11 could be reduced to one-fourth that of the unfilled one. In addition to the cop-
per compounds (CuO, CuS, CuF2), some other particles, such as graphite (Gr), polytetrafluoroethylene
(PTFE), TiO2, SiC, and MoS2, have also been used as lubricating fillers in different polymer matrices
[22–26]. From the experimental results, it was concluded that these fillers can offer a wear reduction
only when the transfer film is thin and uniform, providing an effective coverage over the metal surface.
The enhancement effect of these fillers on the development of the transfer film was explained by the
improved bonding between the transfer film and the counterface, due to mechanical and/or chemical
reactions introduced by the fillers [22–26].
Figure 37.4 summarizes the friction and wear behavior of several polymer composite systems, using
different conventional fillers and their combinations [10,27–32]. It has been found that the lowest wear
rate was obtained with hybrid composites containing both solid lubricants and fiber reinforcements.
The fibers take most of the load during the sliding process and, consequently, improve the load-carrying
capacity of the polymer composite. At the same time, the solid lubricants are helpful in developing a
continuous transfer film on the metallic counterpart, protecting the fibers from severe abrasive action
caused by hard asperities of the counterface roughness. It is also interesting to note that with optimized
compositions, the wear behavior of the polymer hybrid composites is dominated by fillers, as the wear
rates of materials fall into the same range (∼0.5 × 10−6 mm3/N m), regardless of the matrix material. This
means that the optimal compositions can be considered as general guidelines for the design of wear-
resistant composite materials. Nevertheless, research work continues to improve existing formulations

1000
Neat PEEK (27)
PEEK-based composites(10,27,28)
Neat epoxy (29)
General trend
Specific wear rate (10–6 mm3/N.m)

Epoxy-based composites (29)


when adding Neat PEI (30)
100 Reinforcing and PEI-based composites (30,31)
Neat PA 66 (32)
lubricating fillers PA 66-based composites (32)
Neat PP
PP-based composites
10
+Gr.

+PTFE

1 +Cf

+Gr.+CF

+PTFE+Gr.+CF
0.1
0.2 0.4 0.6 0.8 1.0 1.2
Friction coefficient

FIGURE 37.4  Improvement in tribological properties of polymer composites by using different conventional tri-
bofillers and their combinations. All the data were obtained under the same pv condition of 1 MPa and 1 m/s, using
a pin-on-disk apparatus.
37-6 Wear Materials

by using new fillers to tailor the properties of materials for more extreme loading conditions and to
explore other fields of application for these materials.

37.4  Wear of Polymeric Nanocomposites


Polymer nanocomposites, that is, polymeric matrices reinforced with nano-sized fillers (at least one
dimension of the fillers being in the range of about 10–100 nm [33]), have been the focus of current
research in materials science [34–36]. The majority of material development in this respect has been
carried out to improve the mechanical properties of the composites, for example, the fracture tough-
ness and the creep resistance. Little has yet been published, however, on how the nanofillers influence
secondary properties, as for example, the composites’ friction and wear behavior. Nevertheless, research
has shown that the nanofillers can significantly affect the sliding wear behavior of polymers and com-
posites, even at a low volume fraction (e.g., less than 5 vol.%), being able to create a quite different
tribological performance of the material in comparison to a traditional polymeric composite [37,38].
Figures 37.5 and 37.6 summarize the wear results of different polymer-based nanocomposites, with data
taken from recently published papers [39–45]. There are mainly two types of nanofillers that have been
investigated in the area: nano-sized particulate fillers and carbon nanotubes (CNTs). In the following
section, approaches used in designing wear-resistant polymeric nanocomposites are briefly described,
with particular emphasis on recent efforts in developing hybrid nanocomposites, that is, combining
nanoparticles with traditional tribofillers. Here, the term “nanocomposites” is also used when the fillers
are in the submicron range, for example, 300 nm.

37.4.1  Nanoparticle-Reinforced Polymer Composites


To date, various inorganic nanoparticles, including ceramics, metal, and nonmetal oxides, have been
incorporated into different polymeric matrices to improve their wear performance [39–42]. It has been
reported that nanoparticles can generally contribute better to the improvement of tribological proper-
ties under sliding wear conditions than conventional micro-sized particles. The superior performance
of nanocomposites is normally attributed to the huge interface area between the fillers and the polymer

1000
PEEK filled with nano-Si3N4(39)
PTFE filled with nano-Al2O3(40)
PPS filled with nano-CuS (41)
Epoxy filled with nano-Al2O3(42)
Specific wear rate (10–6 mm3/N.m)

Epoxy filled with treated nano-Al2O3(42)


100

10

1
0 5 10
Nanofiller content (vol.%)

FIGURE 37.5  Summary of the wear results achieved with nanoparticle-filled polymeric systems.
Composite Materials 37-7

1000 PTFE filled with CNT (43)


PS filled with CNT (44)
PVDF filled with CNT (45)
Epoxy filled with CNT

Specific wear rate (10–6 mm3/N.m)

100

10

1
0 1 2 3 4 5 10 20 30
Nanofiller content (wt.%)

FIGURE 37.6  Summary of the wear results achieved with multiwall CNT-filled polymeric systems.

matrix, which potentially leads to a better bonding between the different phases and, therefore, to a better
property profile than that achieved with conventional polymer composites. In particular, it has been veri-
fied experimentally that nanoparticles can simultaneously improve the different mechanical properties
of polymer composites, whereas traditional micro-sized particles normally increase some properties, for
example, stiffness and strength, but have a detrimental effect on others, for example, toughness [46]. Such
general improvement in the mechanical properties of polymer nanocomposites, especially toughness, can
lead to a reduction in wear loss caused by subsurface fatigue/cracks and abrasion actions of counterpart
asperities. Another advantage of using nanoparticles in tribological applications is that the abrasiveness
of the nanoparticles decreases remarkably as a result of the reduction in angularity in comparison with
microparticles. Nevertheless, to realize those desirable nanoeffects, it is essential that the nanoparticles
are well dispersed in the polymeric matrix. In fact, agglomeration is considered as a general problem of
polymer nanocomposites, especially those with higher nanofiller content. There is, therefore, normally
an optimum amount of nanofiller beyond which the wear rate of the nanocomposites may even increase
due to these agglomeration effects. As shown in Figure 37.5, the optimum content of nanoparticles for
most polymers is in a range between 1 and 6 vol.%, except for PTFE-based nanocomposites.
To uniformly disperse nanofillers into polymer matrices, several processing routes have been devel-
oped according to the different types of polymers. For example, a twin-screw extruder, equipped with
appropriate feeders for the various fillers, is usually the best processing solution when using thermo-
plastics as the matrix for injection moldable tribology components [47], whereas a dissolver disc in
combination with a pearl mill has been successfully applied to disperse nanoparticles in thermosetting
resins [48]. In all cases, high shear energies are necessary to break up the agglomerates formed by the
nanoparticles. The aim is to separate the individual particles, wet them with the polymer fluid, and sta-
bilize their separated positions during the solidification of the matrix. The dispersion process can also
be supported by the use of additional chemical additives, such as surfactants (e.g., stearic acid or poly-
functional alcohols). Another technique, the irradiation grafting method, has been applied by Zhang
et al. [42,49–51] to disintegrate agglomerated nanoparticles. Several types of nanoparticles, for example,
SiO2, Si3N4, and Al2O3 were modified by covalently bonding polyacrylamide (PAAM) onto the particles.
It was found that the wear rate of the material could be effectively reduced by the use of nanoparticles
thus treated, at very low filler content (<1 vol.%) (see Figure 37.5).
37-8 Wear Materials

Despite much work, the size effect of nanoparticle fillers is not yet fully understood. For example,
some researchers have suggested that nanofillers may be ineffective in reducing the wear of PTFE [52].
It was proposed that fine fillers were unable to prevent transfer and large-scale destruction of the banded
structure of the PTFE matrix. However, other researchers have indicated that nanofiller-reinforced
PTFE composites can actually achieve a better tribo-property profile than that of conventional PTFE
microcomposites, due to the favorable effects of nanofillers on the surface morphology and transfer
films [53]. Nevertheless, there is no general agreement on the understanding of the unique wear behav-
ior of PTFE nanocomposites, namely, that the wear rate can decrease continuously with the increase in
nanofiller volume content (cf. Figure 37.5). More investigation of these materials is still needed.

37.4.2  Carbon-Nanotube-Reinforced Polymer Composites


Since their discovery in the 1990s, CNTs, including single-wall carbon nanotubes (SWCNTs) and
multiwall carbon tubes (MWCNTs), have attracted increasing attention throughout the academic and
industrial worlds owing to their excellent mechanical properties, their high electrical conductivity,
and their chemical stability [54,55]. To date, however, only MWCNTs rather than SWCNTs have been
predominantly used as functional or reinforcing fillers due to their relatively lower cost and better
availability [56].
Because of their excellent mechanical properties and the extremely large aspect ratio, CNTs were
expected to be a promising tribofiller that would significantly enhance the load-carrying capacity of
polymer composites. Yet it is only recently that studies have begun of the tribological properties of
CNT-filled polymer composites. Figure 37.6 shows wear results of different polymers filled with varying
proportions of MWCNTs. Again, it was found that the optimum nanofiller content is very low, that is,
normally less than 2 wt.%, except for PTFE composites. However, it has been evident that the improve-
ment in wear resistance obtained by CNTs is only modest in comparison to short CFs (cf. Figures 37.4
and 37.6). This can be partly explained by the fact that there are still manufacturing problems for CNT-
filled polymers [57]. In particular, the dispersion of nanotubes in a polymeric matrix remains a chal-
lenge, since this can significantly affect the wear behavior of CNT-reinforced polymers [58,59]. Although
large shear stresses can untangle a certain proportion of agglomerated tubes, they can also damage the
structure of nanotubes and, consequently, compromise their properties. Another problem is the dif-
ficulty in controlling the interfacial bonding between CNTs and the polymeric matrix. Such difficul-
ties have also been raised for carbon-fiber-reinforced composites. However, the problem becomes more
acute for CNTs, because of their atomically smooth surfaces and their huge surface area. To fully explore
the potential of CNTs in the tribology of polymers, it is essential to solve the aforementioned problems.

37.4.3  Hybrid Nanocomposites for High Wear Resistance


To develop high-wear-resistant polymer composites, it is a traditional route to integrate various func-
tional fillers (see Figure 37.1). To date, however, only a few studies have been directed toward the
development of hybrid nanocomposites by integrating nanofillers with traditional fillers. Wetzel et al.
[60] systematically investigated the sliding wear behaviors of epoxy composites reinforced by various
amounts of nano-Al2O3 (13 nm) and micro-CaSiO3 (4–15 μm). It was found that the combination of
micro- and nanoscale particles could induce some kind of synergistic effect and improve both the wear
resistance and the stiffness of the material. Cho et al. [61] also reported a synergistic action in the tri-
bological behavior of polyphenylene sulfide (PPS) composites reinforced with nano-CuO and conven-
tional short-fiber reinforcements. The lowest steady-state wear rate was obtained with hybrid composites
containing both nano-sized CuO particles and fiber reinforcements. The reduction in steady-state wear
rate and friction coefficient was attributed to the nanofiller and fiber materials modifying the transfer
film formation and, consequently, the wear mechanisms. Recently, Zhang et al. [62–64] indicated that
the incorporation of nanoparticles with optimized contents could further increase the wear resistance
Composite Materials 37-9

of carbon fabric composites. The beneficial effect of additional nanoparticles on the wear performance
of these composites was explained by the increased mechanical strength of the fabric composites and
better bonding strength of the transfer film developed on the metallic counterface. It is, thus, evident
that the incorporation of both nano-sized inorganic particles and traditional tribofillers in a polymer
can provide synergy in terms of improved wear resistance. This aspect has not yet been adequately
addressed.
To fully promote the synergistic effect of nanoparticles and conventional fillers on the tribological
performance of polymeric composites, systematic studies of the combinative effect of nanoparticles
with traditional fillers have been recently carried out by the authors and their colleagues [29,65–69].
The sliding wear behavior of a series of epoxy composites was studied. For the composite filled with
conventional fillers only, the best wear-resistant composition was found when a combination of fibrous
reinforcements with solid lubricants, that is, graphite and PTFE was used; this finding agreed quite well
with results found for other classic polymer composites (see Figure 37.4). Nevertheless, when nano-
sized TiO2 (300 nm) was used as an additional filler, the friction coefficient and the wear rate of the
composites could be remarkably reduced, especially under extreme sliding conditions, although the
nanoparticles alone did not greatly affect the wear behavior of the epoxy matrix [60]. Figure 37.7 shows
the time-related depth wear rates (calculated with Equation 37.2) for two epoxy composites (one filled
with traditional fillers only, and the other with both nanoparticles and traditional fillers) tested under
different loading conditions. It is obvious that under the same sliding conditions, the time-related depth
wear rate of the hybrid nanocomposite is generally lower than that of the composite filled only with
traditional fillers. In particular, it can be noticed that the wear factor of the composite with traditional
fillers increases rapidly when the pv-value exceeds 2 MPa·m/s (so that caution must be taken to operate
this material in such conditions).
The same tendency was also shown for polyamide-66-based composites filled with traditional tri-
bofillers without or with additional TiO2 nanoparticles [32]. A worn surface of such a composite filled
only with conventional fillers (tested at 2 MPa and 3 m/s) is shown in Figure 37.8. The surface damage
is characterized by numerous scratches, cracks, and sites of broken and removed fiber pieces. For the

25
PTFE+Gr.+SCF/Epoxy
PTFE+Gr.+SCF+nano-TiO2/Epoxy
20
Time-related depth wear rate (nm/s)

15

10

K* = 1 × 10–6 mm3/N.m

0
0 1 2 4 6 8 10 12
pv factor (MPa.m/s)

FIGURE 37.7  Comparison of the time-related depth wear rate of two epoxy-based composites filled only with
conventional fillers and also with additional nano-TiO2 as a function of pv-factor. Sliding conditions: pin-on-disk
apparatus.
37-10 Wear Materials

25kV 18WD 30 µm

FIGURE 37.8  SEM micrograph of a worn surface of a polyamide 66 composite filled only with conventional fillers
(tested at 2 MPa and 3 m/s).

25kV 18WD 30 µm

FIGURE 37.9  SEM micrographs of a worn surface of a polyamide 66 composite filled with both conventional fill-
ers and nanoparticles (tested at 2 MPa and 3 m/s).

composite with both nanoparticles and traditional fillers, however, the wear factor of the nanocom-
posite was relatively stable at ∼3 × 10−6 mm3/N m, even under much higher pv conditions, for example,
8 MPa·m/s. As shown in Figure 37.9, the worn surface of this polyamide 66 nanocomposite was relatively
smooth, even at such a high pressure. The fibers could still effectively carry most of the load and were
only gradually worn, in the classical sequence, that is, fiber thinning, fiber fracture, and removal of small
fiber segments. Such wear synergy effects between conventional fillers and nanoparticles have also been
found in other investigations of thermoplastic composites, for example, with a PEI matrix [30]. With
the addition of nanoparticles, both the friction coefficient and the wear rate of the composites filled
with conventional fillers (short CFs and graphite) were reduced, especially under more extreme wear
conditions such as higher normal pressures, faster sliding velocities, and elevated temperatures. The
advantages produced by additional nanoparticles can be attributed to the following effects: (1) the pres-
ence of nanoparticles in the contact region can reduce the adhesion between the contacting surfaces,
resulting in a lower coefficient of friction; (2) the stress concentration on individual fibers is reduced
with the dispersed nanoparticles in the contact region; (3) a possible rolling behavior of nanoparticles
significantly enhances the tribological performance of composites, especially under extreme sliding
conditions. Owing to these favorable effects, interface damage between fibers and polymeric matrix
Composite Materials 37-11

was to a great extent avoided, resulting in relatively smooth worn surfaces even under severe loading
conditions (cf. Figure 37.9). It was, therefore, concluded that the wear resistance and the “pv limit” of
conventional polymer composites, for example, short-fiber-reinforced polymers (SFRPs), can be signifi-
cantly improved by using nanoparticles as additional fillers, thus favoring further use of these materials
for triboapplication under more severe loading conditions.

37.5  Summary
This overview describes how to design polymeric composites that can operate under low friction and low
wear against metallic sliding counterparts, using different combinations of fillers and reinforcements.
Over recent decades, conventional micro-sized tribofillers such as short fibers and solid lubricants
have been successfully applied to develop wear-resistant polymeric composites for numerous indus-
trial applications. However, investigations are still under way to explore other fields of application for
these materials and to further enhance their properties for more extreme loading and environmental
conditions.
Recently, nano-sized fillers, including inorganic nanoparticles and CNTs, have come under consid-
eration. It has been found that these nanofillers can result in remarkable improvements in the friction
and wear properties of polymeric materials, even at very low volume contents (compared to conventional
micro-sized tribofillers). Further, this circumstance allows the use of these materials under more extreme
wear loading conditions, that is, at higher normal pressures, faster sliding velocities, and elevated tem-
peratures. However, the science of the wear mechanisms in polymer nanocomposites is still incomplete.
The “size effect” of nanofillers and their influence on the quality and stability of transfer films are not yet
fully understood. It is still a subject for continuous research on polymer tribology in the future.

Acknowledgments
Author Dr. L Chang wishes to thank the Australian Research Council for the financial support of his
research works. Additional thanks are to the Alexander von Humboldt Foundation for his previous
research stay at the Technical University Kaiserslautern in 2008.

References
1. J.K. Lancaste, Composites for aerospace dry bearing applications, in: K. Friedrich (Ed.), Friction and
Wear of Polymer Composites, Elsevier Science Publishers B.V., Amsterdam, the Netherlands, 1986,
pp. 363–396.
2. K. Friedrich, L. Chang, F. Haupert, Current and future applications of polymer composites in the
field of tribology, in: L. Nikolais, M. Meo, E. Miletta (Eds.), Composite Materials, Springer–Verlag,
London, New York, 2011, pp. 129–167.
3. S. Wilson, A. Ball, Performance of metal matrix composites under various tribological conditions, in:
K. Friedrich (Ed.), Advances in Composites Tribology, Elsevier Science Publishers, B.V., Amsterdam,
the Netherlands, 1993, pp.311–366.
4. B. Prakash, Friction and wear characteristics of advanced ceramic composites materials, in:
K. Friedrich (Ed.), Advances in Composites Tribology, Elsevier Science Publishers, B.V., Amsterdam,
the Netherlands, 1993, pp.405–450.
5. R.L. Deuis, C. Subramanian, J.M. Yellup, Dry sliding wear of aluminium composites—A review,
Composites Science and Technology 57(4), 415–435, 1997.
6. N. Chand, M. Fahim, Tribology of Natural Fibre Polymer Composites, Woodhead Publ., Cambridge,
U.K., 2008.
7. K. Friedrich, Z. Zhang, P. Klein, Wear of polymer composites, in: G.W. Stachowiak (Ed.) Wear—
Materials, Mechanisms and Practice, John Wiley & Sons, Ltd, Chichester, 2005, pp. 269–290.
37-12 Wear Materials

8. K. Kato, Wear in relation to friction—A review, Wear, 241(2), 151–157, 2001.


9. K. Friedrich, Wear of reinforced polymers by different abrasive counterparts, in: K. Friedrich
(Ed.), Friction and Wear of Polymer Composites, Elsevier Science Publishers, B.V., Amsterdam, the
Netherlands, 1986, pp. 233–287.
10. A.M. Häger, Polyaryletherketone für den Einsatz in Gleitlagern und Gleitelementen, PhD thesis, TU
Kaiserslautern, ISBN 3-8265-2094-7, 1996.
11. M. Cirino, R.B. Pipes, K. Friedrich, The abrasive wear behavior of continuous fibre polymer compos-
ites, Journal of Material Science, 22(7), 2481–2492, 1987.
12. M. Cirino, K. Friedrich, R.B. Pipes, The effect of fibre orientation on the abrasive wear behavior of
polymer composite materials, Wear, 121(2), 127–141, 1988.
13. K. Friedrich, Wear models for multiphase materials and synergistic effects in polymeric hybrid com-
posites, in: K. Friedrich (Ed.), Advances in Composite Tribology, Elsevier Science Publishers, B.V.,
Amsterdam, the Netherlands, 1993, pp. 209–273.
14. W. Bonfield, B.C. Edwards, A.J. Markham, J.R. White, Wear transfer films formed by carbon fibre
reinforced epoxy resin sliding on stainless steel, Wear, 37(1), 113–121, 1976.
15. V.K. Jain, Investigation of the wear mechanism of carbon-fibre-reinforced acetal, Wear, 92(2),
279–292, 1983.
16. H. Voss, K. Friedrich, On the wear behavior of short-fibre-reinforced PEEK composites, Wear,
116(1), 1–18, 1987.
17. H. Zhang, Z. Zhang, Comparison of short carbon fibre surface treatments on epoxy composites. II.
Enhancement of the wear resistance, Composites Science and Technology, 64(13–14), 2031–2038, 2004.
18. T. Tsukizoe, N. Ohmae, Friction and wear performance of unidirectionally oriented glass, carbon,
aramid and stainless steel fibre-reinforced plastics, in: K. Friedrich (Ed.), Friction and Wear of Polymer
Composites, Elsevier Science Publishers B.V., Amsterdam, the Netherlands 1986, pp. 205–231.
19. S. Bahadur, The development of transfer layers and their role in polymer tribology, Wear, 245(1–2),
92–99, 2000.
20. S. Bahadur, C.J. Schwartz, The influence of nanoparticle fillers in polymer matrices on the formation
and stability of transfer film during wear, in: K. Friedrich, A.K. Schlarb (Eds.), Tribology of Polymeric
Nanocomposites, Elsevier Science Publishers, Amsterdam, the Netherlands, 2008, pp. 17–34.
21. S. Bahadur, D. Gong, The role of copper compounds as fillers in the transfer film formation and wear
of nylon, Wear, 154(2), 207–223, 1992.
22. K. Tanaka, Effects of various fillers on the friction and wear of PTFE-based composites, in: K. Friedrich
(Ed.), Friction and Wear of Polymer Composites, Elsevier Science Publishers, B.V., Amsterdam, the
Netherlands, 1986, pp. 137–174.
23. J. Gao, Tribochemical effects in formation of polymer transfer film, Wear, 245(1–2), 100–106, 2000.
24. J.V. Voort, S. Bahadur, The growth and bonding of transfer film and the role of CuS and PTFE in the
tribological behavior of PEEK, Wear, 181–183(1), 212–221, 1995.
25. Q. Zhao, S. Bahadur, The mechanism of filler action and the criterion of filler selection for reducing
wear, Wear, 225–229(1), 660–668, 1999.
26. C.J. Schwartz, S. Bahadur, The role of filler deformability, filler-polymer bonding, and counterface mate-
rial on the tribological behavior of polyphenylene sulfide (PPS), Wear, 251(1–12), 1532–1540, 2001.
27. L. Chang, Z. Zhang, L. Ye, K. Friedrich, Tribological properties of high temperature resistant poly-
mer composites with fine particles, Tribology International, 40(7), 1170–1178, 2007.
28. J. Flöck, K. Friedrich, Q. Yuan, On the friction and wear behavior of PAN- and pitch-carbon fibre
reinforced PEEK composites, Wear, 225–229(1), 304–311, 1999.
29. L. Chang, Z. Zhang, C. Breidt, K. Friedrich, Tribological properties of epoxy nanocomposites: I.
Enhancement of the wear resistance by nano-TiO2 particles, Wear, 258(1–4), 141–148, 2005.
30. L. Chang, Z. Zhang, H. Zhang, K. Friedrich, Effect of nanoparticles on the tribological behavior
of short carbon fibre reinforced poly(etherimide) composites, Tribology International, 38(11–12),
966–973, 2005.
Composite Materials 37-13

31. G. Xian, Z. Zhang, Sliding wear of polyetherimide matrix composites II. Influence of graphite flakes,
Wear, 258(5–6), 783–788, 2005.
32. L. Chang, Z. Zhang, H. Zhang, A.K. Schlarb, On the sliding wear of nanoparticles filled polyamide
6,6, Composites Science and Technology, 66(16), 3188–3198, 2006.
33. H.S. Nalwa, Handbook of Organic-Inorganic Hybrid Materials and Nanocomposites, Vol. 2, American
Scientific Publ., Stevenson Ranch, CA, 2003.
34. E. Thostenson, C. Li, T. Chou, Nanocomposites in context, Composites Science and Technology,
65(3–4), 491–516, 2005.
35. J. Coleman, U. Khan, W. Blau, Y. Gun’ko, Small but strong: A review of the mechanical properties of
carbon nanotubepolymer composites, Carbon, 44(9), 1624–1652, 2006.
36. D. Paul, L. Robeson, Polymer nanotechnology: Nanocomposites, Polymer, 49(15), 3187–204, 2008.
37. Z. Zhang, K. Friedrich, Tribological characteristics of micro- and nanoparticle filled polymer com-
posites, in: K. Friedrich, S. Fakirov, Z. Zhang (Eds.), Polymer Composite—From Nano- to Macro-Scale,
Springer, Berlin, Germany, 2005, pp. 169–185.
38. B.J. Briscoe, S.K. Sinha, Tribological applications of polymers and their composites: Past, present
and future prospects, in: K. Friedrich, A.K. Schlarb (Eds.), Tribology of Polymeric Nanocomposites,
Elsevier Science Publishers, Amsterdam, the Netherlands, 2008, pp. 1–14.
39. Q.H. Wang, J. Xu, W. Shen, W. Liu, An investigation of the friction and wear properties of nanometer
Si3N4 filled PEEK, Wear, 196(1–2), 82–86, 1996.
40. W.G. Sawyer, K.D. Freudenberg, P. Bhimaraj, L.S. Schadler, A study on the friction and wear behavior
of PTFE filled with alumina nanoparticles, Wear, 254(5–6), 573–580, 2003.
41. S. Bahadur, C. Sunkara, Effect of transfer film structure, composition and bonding on the tribologi-
cal behavior of polyphenylene sulfide filled with nano particles of TiO2, ZnO, CuO and SiC, Wear,
258(9), 1411–1421, 2005.
42. M.Z. Rong, M.Q. Zhang, G. Shi, Q.L. Ji, B. Wetzel, K. Friedrich, Graft polymerization onto inor-
ganic nanoparticles and its effect on tribological performance improvement of polymer composites,
Tribology International, 36(9), 697–707, 2003.
43. W.X. Chen, F. Li, G. Han, J.B. Xia, L.Y. Wang, J.P. Tu, Z.D. Xu, Tribological behavior of carbon-nano-
tube-filled PTFE composites, Tribology Letters, 15(3), 275–278, 2003.
44. Z. Yang, B. Dong, Y. Huang, L. Liu, F.Y. Yan, H.L. Li, Enhanced wear resistance and micro-hardness of
polystyrene nanocomposites by carbon nanotubes, Materials Chemistry and Physics, 94(1), 109–113,
2005.
45. Q. Jacobs, W. Xu, B. Schädel, W. Wu, Wear behavior of carbon nanotube reinforced epoxy resin com-
posites, Tribology Letters, 23(1), 65–75, 2006.
46. B. Wetzel, F. Haupert, M.Q. Zhang, Epoxy nanocomposites with high mechanical and tribological
performance, Composites Science and Technology, 63(14), 2055–2067, 2003.
47. A. Gebhard, N. Knör, F. Haupert, A.K. Schlarb, Nanopartikelverstärkte Hochleistungsthermoplaste
für extreme tribologische Belastungen im Automobilbau. 48. Tribologie-Fachtagung, Band I,
S.22/1–22/11, 2007
48. B. Wetze, Mechanische Eigenschaften von Nanokompositen aus Epoxydharz und kera-
mischen Nanopartikeln, PhD thesis (ISBN 3-934930-65-4), Technical University Kaiserslautern,
Kaiserslautern, Germany, 2006.
49. M.Q. Zhang, M.Z. Rong, L. Shu, B. Wetzel, K. Friedrich, Effect of particle surface treatment on the
tribological performance of epoxy based nanocomposites, Wear, 253(9–10), 1086–1093, 2002.
50. G. Shi, M.Q. Zhang, M.Z. Rong, B. Wetzel, K. Friedrich, Friction and wear of low nanometer Si3N4
filled epoxy composites, Wear, 254(7–8), 784–796, 2003.
51. G. Shi, M.Q. Zhang, M.Z. Rong, B. Wetzel, K. Friedrich, Sliding wear behavior of epoxy containing
nano-Al2O3 particles with different pretreatments, Wear, 256(11–12), 1072–1081, 2003.
52. K. Tanaka, S. Kawakami, Effect of various fillers on the friction and wear of Polytetrafluoroethylene-
based composites, Wear, 79(2), 221–234, 1982.
37-14 Wear Materials

53. D.L. Burris, K. Santos, S.L. Lewis, X. Liu, S.S. Perry, T.A. Blanchet, L.S. Schadler, W.G. Sawyer,
Polytetrafluoroethylene matrix nanocomposites for tribological applications, in: K. Friedrich,
A.K. Schlarb (Eds.), Tribology of Polymeric Nanocomposites, Elsevier Science Publishers, Amsterdam,
the Netherlands, 2008, pp. 403–438.
54. S. Ijima, T. Ichihashi, Single-shell carbon nanotubes of 1-nm diameter, Nature, 363(6430), 603–604,
1993.
55. R.H. Baughman, A.A. Zakhidov, W.A. de Heer, Carbon nanotubes-the route toward applications,
Science, 297(5582), 787–792, 2002.
56. A. Moisala, Q. Li, I.A. Kinloch, A.H. Windle, Thermal and electrical conductivity of single- and multi-
walled carbon nanotube-epoxy composites, Composites Science and Technology, 66(10), 1285–1288,
2006.
57. P.M. Ajayan, J.M. Tour, Materials science: Nanotube composites, Nature, 447(7148), 1066–1068,
2007.
58. L.C. Zhang, I. Zarudi, K.Q. Xiao, Novel behavior of friction and wear of epoxy composites reinforced
by carbon nanotubes, Wear, 261(7–8), 806–811, 2006.
59. H. Chen, O. Jacobs, W. Wu, G. Rüdiger, B. Schädel, Effect of dispersion method on tribological prop-
erties of carbon nanotube reinforced epoxy resin composites, Polymer Testing, 26(3), 351–360, 2007.
60. B. Wetzel, F. Haupert, K. Friedrich, M.Q. Zhang, M.Z. Rong, Impact and wear resistance of polymer
nanocomposites at low filler content, Polymer Engineering and Science, 42(9), 1919–1927, 2002.
61. M.H. Cho, S. Bahadur, Study of the tribological synergistic effects in CuO-filled and fibre-reinforced
polyphenylene sulfide composites, Wear, 258(5–6), 835–845, 2005.
62. Z.Z. Zhang, F.H. Su, K. Wang, W. Jiang, X.H. Men, W.M. Liu, Study on the friction and wear proper-
ties of carbon fabric composites reinforced with micro- and nano-particles, Materials Science and
Engineering A, 404(1–2), 251–258, 2005.
63. F.H. Su, Z.Z. Zhang, K. Wang, W. Jiang, X.H. Men, W.M. Liu, Friction and wear properties of carbon
fabric composites filled with nano-Al2O3 and nano-Si3N4, Composites Part A: Applied Science and
Manufacturing, 37(9), 1351–1357, 2006.
64. F.H. Su, Z.Z. Zhang, W.M. Liu, Mechanical and tribological properties of carbon fabric composites
filled with several nano-particulates, Wear, 260(7–8), 861–868, 2006.
65. Z. Zhang, C. Breidt, L. Chang, F. Haupert, K. Friedrich, Enhancement of the wear resistance of
epoxy: Short carbon fibre, graphite, PTFE and nano-TiO2, Composites Part A: Applied Science and
Manufacturing, 35(12), 1385–1392, 2004.
66. L. Chang, Z. Zhang, Tribological properties of epoxy nanocomposites: II. A combinative effect of
short carbon fibre and nano-TiO2, Wear, 260(7–8), 869–878, 2006.
67. L. Chang, Z. Zhang, L. Ye, K. Friedrich, Tribological properties of epoxy nanocomposites: III
Characteristics of transfer film, Wear, 262(5–6), 699–706, 2007.
68. L. Chang, K. Friedrich, Enhancement effect of nanoparticles on the sliding wear of short fibre-
reinforced polymer composites: A critical discussion of wear mechanisms, Tribology International,
43(21), 2355–2364, 2010.
69. N.K. Myshkin, M.I. Petrokovets, A.V. Kovalev, Tribology of polymers: Adhesion, friction, wear, and
mass-transfer, Tribology International, 38(11–12), 910–921, 2005.
38
Coatings and Surface
Treatments
38.1 Introduction.....................................................................................38-1
38.2 Thermal Spray Coatings.................................................................38-1
Plasma Wire Arc Process  •  High Velocity Oxy-Fuel Coating  • ​
Plasma Spray Process
38.3 Electrochemical Coatings...............................................................38-6
Nickel Composite Coatings  •  Plasma Electrolytic Oxidation
Coatings
38.4 Physical Vapor Deposition.............................................................38-8
Vacuum Evaporation  •  Cathodic Arc Vapor Deposition  •  Sputter
Deposition  •  Ion Plating
38.5 Chemical Vapor Deposition Process............................................38-9
38.6 Case Hardening of Steels.............................................................. 38-10
Nitriding  •  Carburizing  •  Carbonitriding  •  Induction Hardening
Arup Gangopadhyay 38.7 Automotive Trends........................................................................38-13
Ford Motor Company References...................................................................................................38-13

38.1  Introduction
Coatings are materials deposited on surfaces to enhance performance, that is, resist wear, erosion, cor-
rosion, lower friction, etc. The coating material may have a very different chemical composition than
the base material. Since wear, corrosion, and friction phenomenon are restricted on the surface, deposi-
tion of a coating allows the use of less expensive base material while protecting the surface. Generally,
ferrous-based material offers improved wear resistance compared to other engineering alloys. However,
under extreme conditions, the wear resistance of ferrous material may need to be improved further
through deposition of coatings. Aluminum alloys, titanium alloys, and magnesium alloys generally do
not offer adequate wear resistance and therefore require a coating for wear protection. There are wide
varieties of materials available for deposition and the choice of a material depends on the application,
compatibility with the base material, base material properties, and the deposition process. Similarly,
there is a variety of coating deposition techniques and the selection of a particular technique depends on
the substrate material. The following sections will describe various deposition techniques.

38.2  Thermal Spray Coatings


Thermal spray coatings are widely used in aerospace, petrochemical, gas turbine, and automotive indus-
tries to improve wear resistance and corrosion resistance. Thermal spray process is a family of processes
that uses a concentrated heat source either by using an electric arc or plasma arc or by combusting a
gas mixture, passing gas with high kinetic energy through the heat source and melting the materials

38-1
38-2 Wear Materials

to be deposited. The molten material is carried by the gas and deposited on to the substrate where they
rapidly solidify. The kinetic energy is provided by constricting the flow of the gas through a nozzle.
Thermal spray process could experience extremely high temperature (3,500°C to as high as 20,000°C)
and the kinetic energy of the gas is high enough to transport molten particles at supersonic speed to the
substrate. This leads to very high deposition rate and allows deposition of thick coatings (100–300 μm).
Also, thermal spray is a “line-of-sight” process where the stream of molten droplets deposits only on
to the surface that is directly in line with the spray stream. One of the advantages of thermal spray
process is that the substrate undergoes minimal heating or some local melting. Therefore, the substrate
chemistry or mechanical properties like hardness do not change appreciably. The bonding between the
substrate and coating tends to be mechanical interlocking caused by the impingement of molten drop-
lets at supersonic speed on the substrate surface. Therefore, surface preparation is key to achieving the
desired bond strength. Thermal spray process is capable of working in wide temperature range, particle
velocities, and atmospheric conditions allowing deposition of wide range of materials. The processes are
differentiated depending on how the heat source is provided, how the material is fed into the system,
etc. This section will describe some of the spray processes in detail. Crawmer [1] provided a detailed
comparison of various thermal spray processes.

38.2.1  Plasma Wire Arc Process


The plasma wire arc process is suitable for depositing coatings on both cast iron and aluminum sur-
faces. The deposition rate achieved by this process is much higher than that achieved with conventional
plasma spray process and also at a lower power level. The lower heat input to the deposited coating leads
to reduced residual stress allowing deposition of thick coatings. In automobile engines, cast iron cylin-
der blocks and cylinder heads have been used for a long time. Over the last 20 years, the cast iron heads
have been replaced with aluminum heads for weight saving. Similarly, cast iron cylinder blocks have
been replaced with aluminum. However, aluminum alloys do not show acceptable wear behavior in con-
tact with ferrous-based piston rings under engine operating conditions. Therefore, a cast iron liner or
sleeve is either cast-in or pressed-in on aluminum cylinder blocks. The current trend is toward replace-
ment of cast iron liners with deposition of an iron-based coating on aluminum cylinders to reduce mass
and improve fuel economy [2]. Figure 38.1 shows a setup for deposition of coating using plasma wire
arc process. It consists of a torch head attached to a spindle, which is rotated by a motor. The torch head
consists of a nozzle through which a plasma gas flows surrounding a copper cathode. The plasma gas
could be argon or nitrogen. The wire feedstock acts as anode, and is pulled within the torch head per-
pendicularly and in front of the nozzle cone. The plasma is created by applying a voltage of 100–120 V
and 60–100 mA current between the tip of the cathode tube and the anode feedstock. These conditions

Wire
feedstock
Plasma
gas

Cathode

Nozzle

Atomizing gas

FIGURE 38.1  A schematic drawing of plasma-transferred wire arc process.


Coatings and Surface Treatments 38-3

raise the temperature high enough to melt the wire materials. The flow of the plasma gas and the atom-
izing gas, which could be compressed air detaches the particles from the anode tip and forces it to the
substrate on which the coating is desired. The flow of plasma gas and the atomizing gas is maintained
such that the particles travel at a high speed (100–130 m/s) creating a dense coating. However, for coating
the cylinder bore of an automotive engine, the spindle needs to rotate vertically at the center of cylinder
bore and then honed to the desired specification. The coating deposition process requires mechanical
roughening of the substrate to improve the bond strength. In a slightly different setup, two different
wires could be used as well and the method is called twin wire arc process.
The typical coating thickness is about 200 μm, although coating thickness as high as 3 mm can be
deposited for certain applications [3]. Since the coating thickness increases with time, a layered structure
is developed with low porosity (2%–8%), and, therefore, the as-deposited coating tends to be rough. The
coating composition and hardness are controlled primarily by the composition of wire feedstock, which
is typically low to medium carbon steel for automotive engine applications. Other metallic wires can also
be used, that is, copper, aluminum, and zinc. Cored wires containing powders can also be used to deposit
composite coatings [4]. If the atomizing gas is air, which is used in most cases, the coating contains
oxides. The amount of oxide and porosity levels depends on the air flow rates [5]. The substrate tempera-
ture also plays a significant role in coating properties. Increased substrate temperature tends to decrease
the amount of porosity at the coating–substrate interface improving bond strength and deposition effi-
ciency [6]. Increased temperature also changes the morphology of splats (deposited material), being quite
irregular at lower temperature to more round shape with a smooth edge at higher temperature.
The plasma-transferred wire arc process is also widely used in electrical and electronic industry for
its ability to deposit adherent metallic coating on nonmetal substrates without causing damage. The
process is used for making conductive terminals of resistors and capacitors and in the application of
conductive solderable and brazeable terminal joints on ceramic substrates. Nickel is the typical metal
deposited on alumina substrate [7,8].

38.2.2  High Velocity Oxy-Fuel Coating


The high velocity oxy-fuel (HVOF) process consists of feeding a mixture of oxygen and an oxy-fuel at
high pressure into water-cooled combustion chamber. Following combustion, the resulting flame/gas
stream is passed through a confined nozzle. The coating material is fed through a carrier gas either at
the tip of the nozzle in a powder form or through the center of the gun depending on the design. The
coating powders melt at the high temperature generated at the nozzle tip and molten droplets are carried
by the gas stream at a high velocity (300–600 m/s) toward the substrate material. The molten powder
particles solidify upon impingement on the substrate surface. Figure 38.2 shows a schematic diagram of
the process. The high kinetic energy of particles leads to extremely dense coatings [9,10]. The oxy-fuel
gases include hydrogen, propane, propylene, acetylene while the carrier gas for powders could be air,
nitrogen, etc.

Compressed air

Fuel gas Coating

Powder
with
carrier gas

FIGURE 38.2  A schematic drawing of high velocity oxy-fuel process.


38-4 Wear Materials

HVOF process is attractive for improving wear and corrosion resistance of Al alloys and aluminum
metal matrix composites. The low temperature and the supersonic speed of particles being deposited
avoid overheating of substrates, thereby allowing the deposition possible in materials with low melting
temperatures [11]. This process replaces chrome plating used in aerospace and automotive industries
[12,13]. WC-based materials are commonly used in HVOF process for improving wear and corrosion
resistance. The addition of Co binder phase in WC matrix offers high hardness together with high duc-
tility [14]. Other materials used in combination with WC are Ni, Co–Cr, Ni–Cr, etc. [15]. There is a wide
range of materials including fluoropolymers [16–18] that can be used in this process because of com-
mercially available powders. The coating may be amorphous or microcrystalline in nature or could be
fine-grained mixture of metastable or stable phases. In the case of WC–Co coating, the primary coating
constituent is WC–Co, while the secondary constituents such as W2C, CoW9C, and Co8W9C 4 phases are
identified [19]. In the case of Ni–Cr system, the microstructures have been found to consist of high and
low Cr γ-Ni metallic regions along with NiO and NiCr2O4 oxides [20].
The higher gas flow associated with HVOF process leads to high compressive residual stress leading
to improved coating adhesion [9,21]. The residual stress increases with increased coating thickness,
and decreased spray distance [22,23]. The mechanical properties of HVOF coatings could be quite dif-
ferent from the bulk material because of the thermal changes involved during the deposition process.
Oxidation and the boiling of the individual components in the powder mixture could very well lead to
coating composition quite different from initial powder mix leading to different mechanical properties.
Varis et al. [24] observed lower Young’s modulus and fracture strength of WC–CoCr, NiCr, and Al2O3
coatings than bulk material while Hu et al. [25] measured higher hardness and lower Young’s modulus
of NiAl coating than the bulk NiAl intermetallic compound.
The HVOF process has been used for the deposition of various coatings (oxides based on Al, Ti,
Mo, and mixture of these) on the cylinder bore of an automobile engine [26]. The cylinder block is
rotated around its vertical axis while the torch angle with respect to the bore also varied to deposit
uniform coating thickness. The torch angle with respect to the bore is maintained at 90° to deposit
coating near the top dead center area while an angle of 30° is required to cover bottom dead center
area. The use of propane/oxygen fuel mixture introduced larger compressive stress due to larger gas
flow rate and less degree of particle fusion than acetylene/oxygen fuel mixture. HVOF process is also
used to deposit hard coating of nickel chromium/chromium carbide on diesel engine piston rings.
A special cooling arrangement was required to avoid overheating of the piston rings [27,28a,b]. This
coating is developed to replace chromium plating on rings, which has adverse health and environ-
mental effects.

38.2.3  Plasma Spray Process


In this process, plasma is created between a tungsten cathode and anode, which is also a nozzle, by
applying a voltage. The ionization gas could be nitrogen, hydrogen, argon, and helium. The diatomic
gases generally have higher enthalpies than monatomic gases and therefore higher temperatures can be
attained. In plasma arc, the gas heating is sufficient to raise the temperature of the plasma to 10,000°C.
The hot gas is passed through a constriction (nozzle) to increase its kinetic energy. Material to be depos-
ited is fed in powder form right in front of the nozzle where it melts and molten droplets are carried by
the gas stream on to the substrate surface. Figure 38.3 shows a schematic drawing of the plasma spray
process. The important differences between plasma spray process and HVOF process are that the plasma
spray process offers higher temperatures but lower particle velocities [29]. The extreme high temperature
of the plasma flame allows deposition of high melting point materials, like oxides and ceramic com-
pounds. A wide variety of materials have been attempted for deposition. The deposition of a few of the
most promising materials will be discussed here.
Plasma spraying of Al2O3–TiO2 coatings has been studied by several authors [30–34] because of its
high hardness, corrosion, and erosion resistance. The coating is used for improving wear resistance,
Coatings and Surface Treatments 38-5

Powder
injection

Plasma gas

Cathode

Nozzle
(anode)

FIGURE 38.3  A schematic drawing of plasma spray process.

corrosion protection in shipboard and marine applications, and for electrical insulation. It has been
reported that the composite coating containing 13% TiO2 offers the best wear resistance. It has been
observed that nanostructured alumina–titania coatings offer even better wear resistance because of
the presence of partially melted regions. Plasma spraying of Al2O3–B4C composite is attractive for cut-
ting tool applications because of extreme high hardness of B4C. Also, high thermal conductivity of B4C
makes it a candidate material for nuclear reactors [35].
Plasma spray process has been explored in automotive and aerospace industries. In the automotive
area, the application has been primarily focused on engine bores. In the case of diesel engine applica-
tions, the focus was on spraying thermal barrier coatings on cast iron cylinders. The efficiency of diesel
engine increased from 36% in 1980s to 43% in early 1990s [36]. Further increase in efficiency requires
effective thermal management by better utilization of exhaust heat. This can be achieved by insulating
the combustion chamber and raising the engine operating temperature from 150°C to 300°C. One of the
ways to achieve this is by providing a thermal barrier coating on the bore to prevent heat loss. Several
types of materials have been tried including zirconia [37,38]. In the case of gasoline engines, thermal
spray coatings have been considered on engine bore to allow use of parent aluminum materials and
elimination of cast iron liner. There are two primary tooling requirements for depositing plasma spray
coatings on cylinder bores: (1) the plasma torch has to rotate inside a stationary block and (2) the plasma
torch has to fit inside the bore with sufficient spray distance for satisfactory coating deposition with the
required properties and reliability [29]. Different types of materials can be used for coating aluminum
bores including metallic alloys, metal matrix composites, metal–solid lubricant composites, and metal
ceramic composites. Deposition of low alloy steel has been investigated in detail. However, for this coat-
ing the formation of FeO and Fe3O4 needs to be controlled. These oxides can also function as solid lubri-
cant. The formation of Fe2O3 should be avoided because it can act as an abrasive. The coating hardness is
generally in the range of 350–650 HV similar to that of cast iron material. Rao et al. [39] explored deposi-
tion of 434 stainless steel-based material coupled with a solid lubricant, which could be Ni-encapsulated
boron nitride or calcium fluoride. The coatings were deposited on a production engine block where the
cast iron liners were replaced with coated aluminum liners. Fleet test results showed reduced ring and
bore wear. Plasma spray coating has been in use for piston rings in gasoline engines. The most com-
monly used coating in North America is nickel chrome/chromium carbide alloy blended with molyb-
denum powder. Addition of molybdenum powder improves scuff resistance and coating cohesion [27].
Surface preparation is critical for achieving adequate substrate/coating bond strength, which can
be done by grit blasting with alumina-based material [29], abrasive jet machining, or laser technology.
Another important step following coating deposition is honing of the coated surface to maintain the
required bore diameter and surface roughness pattern. The first introduction of plasma spray coating
was in a four-cylinder aircraft engine in 1966. Later, plasma spray coating was applied in the production
of gasoline and diesel engines in Europe.
38-6 Wear Materials

38.3  Electrochemical Coatings


38.3.1  Nickel Composite Coatings
A series of nickel composite coatings were developed in early 1990s to address corrosion issues related
to the use of alternate fuels. In automotive engines, use of M85 (a blend of 15% gasoline and 85% metha-
nol) revealed higher wear of rings, cylinder bore, and bearings than gasoline-fueled engines [40]. These
coatings were also developed to be applied directly on an aluminum engine cylinder bore to avoid use of
cast iron liners. These coatings were appropriate because of their higher wear resistance and lower fric-
tion coefficient than cast iron. Nickel composite coatings are attractive because of precision-controlled
low-temperature operation, low cost and low energy requirements, high deposition rate, and the ability
to deposit on complex geometry [41].
Nickel composite coatings are deposited by electroplating technique. It consists of a plating bath with
an anode and a cathode with an electrochemical solution. The anode is made of nickel and the cathode
is made of the material on which the coating is to be deposited. The electrochemical solution is made up
of nickel sulfate or ammonium sulfate, nickel chloride, and boric acid. The bath is maintained typically
at 50°C, and a current density of about 10 A/dm2 is maintained while the bath is continuously stirred. A
second phase is often added to tailor the properties of the coating for improved wear resistance, friction
coefficient, and corrosion resistance. The second phase could be ceramic particles like silicon carbide,
boron carbide, silicon nitride, etc., hard oxides like Al 2O3, SiO2, TiO2, or solid lubricants like boron
nitride, molydisulfide, graphite. The most common material is electroplated nickel with embedded sili-
con carbide particles and is commonly known as Ni–SiC or Nikasil coating. The coating deposition rate
can be increased by increasing the pH of the electrochemical solution [42], and the amount of the second
phase particle in the solution. Zimmerman et al. [43,44] deposited Ni–SiC coating by pulse electrode-
position technique and characterized its microstructure and mechanical properties. The deposits had
a very smooth and shinny surface having a thickness in the range of 200 μm. The size of SiC particles
were micron or submicron scale. The typical microstructure consisted of irregular-shaped SiC par-
ticles in a nanocrystalline Ni matrix. The hardness of nanocomposite coating was about 600 kg/mm2,
which is four times higher than a conventional Ni–SiC metal matrix composite. Addition of phospho-
rous in Ni–SiC coating further increases hardness [45–47]. Phosphorous is added in the coating by
addition of phosphoric acid in the electrolytic bath. Hardness can be increased further by heat treatment
above 400°C due to the precipitation of Ni3P phase.
Nickel composite coatings can also be deposited without any electricity and coating is called electro-
less nickel composite coatings [48]. The coating is deposited by controlled chemical reduction of nickel
ions on a catalytic surface. The reaction continues as long as the surface is in contact with the solution. A
typical electroless nickel bath consists of nickel sulfate and sodium hypophosphate [49]. The advantage
of electroless coating lies in its ability to deposit uniform coating thickness and it can also be applied
on internal surfaces of a part, which cannot be accomplished easily with electrolytic process. It is pos-
sible to incorporate specific particles in the coating to impart specific properties similar to electrolytic
process. The particles could be Si3N4 [50] B4C [49], and SiC [51] for improved wear resistance property
and PTFE, BN, etc. for friction reduction. Electroplated nickel composite coatings can be deposited on
a variety of materials including aluminum alloys, steel, and magnesium alloys. Engine components like
piston rings, pistons are coated for two-stroke, motorcycles, marine, and snow-mobile engines [48].

38.3.2  Plasma Electrolytic Oxidation Coatings


The anodizing of aluminum alloys has been utilized for several decades and the process forms a thin
hard layer of aluminum oxide on the surface of an aluminum part. This is an electrolytic process using
an electrochemical bath where the electrolyte could be sulfuric acid, phosphoric acid, oxalic acid,
etc., while the anode is made of aluminum alloy. Typical current density and voltage in the range of
Coatings and Surface Treatments 38-7

0.5–1.5 A/dm2 and 10–100 V are used, respectively [52,53]. A simplified description of the formation of
aluminum oxide can be written as

2Al 3 + + 3H2O → Al 2O3 + 6H −

The coating forms by concurrent formation and dissolution of the oxide. The coating also contains
anions from the electrolyte and water [54]. Generally, the anodized layer consists of a thin barrier layer
adjacent to the aluminum surface, the thickness of which could be tens of nanometers followed by a
thicker functional layer. Anodizing process is widely used to improve wear and corrosion resistance of
aluminum alloys.
When the voltage during the anodizing process is increased above 250 V in the presence of suitable
electrolytes, there are plasma discharges at the interface. These discharges are short-lived, often in the
range of nanoseconds to milliseconds and the sizes of these discharges are very small, in the range of
100–300 μm [55]. The process is also described as plasma anodizing, microarc anodizing, spark or arc
anodizing. The coating growth takes place in the regions of microdischarges. Figure 38.4 shows a sche-
matic description of the plasma electrolytic process. Yerokhin et al. provided a detailed review of the
process [56]. The coating generally forms two layers: the top layer (10–20 μm) is called a loose layer con-
taining large pores and can be removed by machining or could be impregnated with other materials to
impart specific property like reduced friction by incorporating solid lubricants like graphite or molyb-
denum disulfide [57]. The layer closer to the substrate, which could be as thick as 100–200 μm, is the
actual functional layer for tribological application or corrosion protection with low porosity. The coat-
ing is essentially alumina and the hardness of the intermediate layer is about 900–2000 HV. The coat-
ing may also contain silicon or phosphorous from the electrolyte. The electrolyte used for this process
contains sodium silicate and potassium hydroxide, although phosphates and fluorides could be used as
alternate electrolytes. During the PEO process, the classical electrochemical process involving charge
transfer at the electrode–electrolyte interface is combined with strong ionization and charge transfer
across wet plasma gas film between electrolyte and the electrode. This results in plasma thermochemical
reactions [58]. Aluminum alloys subjected to plasma electrolytic oxidation process is currently used in
automotive pistons and brake components for improved wear and corrosion characteristics. Some of the
aluminum alloys used for this process include Al–Si, Al–Cu–Mg, Al–Mg, and Al–Zn–Mg alloys. The

 +

Cathode Anode
Cations
2H+ H2 O2– O2
Me0 Me
Cat Cat0 Anions

Electrolyte

FIGURE 38.4  A schematic drawing of plasma electrolytic process.


38-8 Wear Materials

plasma electrolytic oxidation process is applicable to titanium and titanium alloys (Ti–50Al, Ti–50Zr,
Ti–6Al–4V) and magnesium alloys (AZ31, AZ91D, AZ92) to improve their corrosion resistance [59–64].

38.4  Physical Vapor Deposition


Physical vapor deposition process is a technique for depositing thin coatings on a surface atom by atom.
In simple terms, the coating material is vaporized from a solid or liquid source, then guided through a
vacuum or partial vacuum on to the surface to be coated. The target surface is maintained at a negative
potential so that the atoms or ions are attracted to the surface. The vacuum maintained inside the chamber
could be in the range of 10−1 to 10−5 Pa. The coating thickness is generally in the range of few nanometers
to few micrometers. The coating material could be an element, a compound, a single layer or multilayered,
or a composition-graded layer. The process can have different names depending on the source of heating
the coating material and the way the plasma is created. The process can be divided into four categories
[65,66]: (1) vacuum evaporation, (2) arc vapor deposition, (3) sputter deposition, and (4) ion plating.

38.4.1  Vacuum Evaporation


This is one of the simplest techniques for coating deposition. The material to be deposited is heated
inside a vacuum chamber to a high enough temperature that vapors are generated. A vacuum of 10−2 Pa
is required. A higher vacuum may be necessary to avoid contaminants. The vapors condense on the
substrate directly on the line of sight. The heating source could be an electron beam or resistive heating.
This process works well for depositing a single element but in the case of an alloy, the element having a
higher vapor pressure will tend to deposit more than the other resulting in a coating composition, which
may differ from the original alloy composition. Since it is a line-of-sight process, the coating thickness
may not be uniform. However, the coating deposition rate is generally high.

38.4.2  Cathodic Arc Vapor Deposition


In cathodic arc vapor deposition process, an electric arc is created by two closely placed electrodes and
by passing a high current 60–120 A. The arc discharge current is concentrated at the cathode surface
with extremely high current density (of the order of 1010 A/m2) creating a concentrated ionized plasma.
The plasma consists of the cathode material that is made of the material intended for deposition. The sur-
face to be coated is maintained at a negative bias voltage, which could be of the order of 250 V. The ions
from the plasma move at a supersonic velocity to the surface to be coated and get deposited. The cathode
and the anode are kept inside a vacuum chamber where a pressure of 10−1 Pa is maintained. The coat-
ing density depends on how fast the ions hit the substrate surface. Slower deposition rate coupled with
higher substrate temperature favors denser coating because of time available for the depositing atoms to
rearrange themselves leading to a microstructure similar to the bulk material [67]. Thick coatings can
be deposited by continuous operation in DC mode, used for commercial operation. One of the issues
with cathodic arc deposition technique is the formation of macroparticles, which gets incorporated in
the coating and degrades coating properties. The macroparticles are the droplets or debris particles
originating from the cathode surface. The amount of macroparticle should be reduced by increasing
the cathode–substrate distance, reducing the current level, and lowering the substrate temperature [66].
Macroparticles can also be effectively reduced by using a duct filter or a knee filter where the plasma is
bent out of line of sight by a magnetic coil when the charge neutral macroparticles are deposited on the
duct wall while the charged plasma gets to the substrate surface.
The cathodic arc deposition technique can be used for deposition of TiN, ZrN, TiC, NbN, CuO, amor-
phous carbon coatings, etc. Deposition of oxides or nitrides can be achieved by introducing oxygen or
nitrogen gas in the vacuum chamber [68–70].
Coatings and Surface Treatments 38-9

38.4.3  Sputter Deposition


In this process, materials are evaporated by the bombardment of the surface by energetic ions. The ener-
getic ions are created by plasma between the substrate to be coated and the target material. The substrate
represents an anode and the target material represents a cathode. The energetic ions are typically argon
gas. In the simplest form, a DC diode is used for generation of a glow discharge. But the deposition rate
is low and creates high thermal heating because of the localization of the plasma very close to the target
surface [71]. Application of a magnetic field (magnetron sputtering) allows the electrons to circulate in
a circular path around the target increasing plasma density and therefore the deposition rate. However,
the disadvantage is the nonuniform coating thickness. The use of unbalanced magnetron configuration
enables passage of high ion current toward the substrate allowing deposition of dense coating but non-
uniformity in coating thickness remains. A further development is the use of multiple magnetrons for
deposition of uniform coating thickness [72,73]. Radiofrequency (RF, 13.56 MHz) plasma discharge is
introduced to deposit insulating materials like oxides of aluminum, silicon, etc. It is possible to deposit
an element or compound by using them as a target material. It is also possible to deposit oxides or
nitrides by allowing the metal to react (reactive sputtering) with a gaseous phase (oxygen or nitrogen,
respectively) or carbides [74] and co-sputtering with a carbon target material. However, this process
runs the risk of poisoning the target material due to coating deposition.

38.4.4  Ion Plating


In this process, the substrate surface and the film is bombarded either continuously or intermittently
with energetic ions during film growth. This allows adjustments of atoms while the film is being depos-
ited resulting in a denser film, and also alters surface morphology, grain size, and porosity. The film
properties can approach that of the bulk material. The energetic ions can be redirected from the plasma
source or can be supplied through an ion gun attached to the vacuum chamber. The energetic ions can
be used to clean the substrate surface prior to film deposition, which improves film adhesion [66].

38.5  Chemical Vapor Deposition Process


Chemical vapor deposition involves dissociation or chemical reaction of gaseous reactants in activated
state inside a chamber resulting in the deposition of a film atom by atom on a substrate. The source of
energy could be in the form of heat, plasma, or a laser. The advantage of chemical vapor deposition is the
ability to form a dense high quality (pure) coating. Another advantage of this process is that it is not a
line-of-sight process enabling deposition of coating uniformly on a complex shape. The process is used
extensively in semiconductor industry for deposition of silicon and silicon dioxide layers and also in
manufacturing industry for deposition of wear and corrosion protection layers.
The critical steps involved in chemical vapor deposition process include (a) transport of gaseous reac-
tants inside the chamber, (b) absorption of gaseous reactants on to the heated substrate and gas phase
reactions at the gas–solid interface forming a film, (c) diffusion of the deposit along the substrate surface
forming crystallization centers and film growth, and (d) removal of the gaseous by-products from the
chamber. An excellent review of CVD process can be found in Reference 75.
Depending on the source of energy, CVD processes can be divided into thermal-activated CVD,
plasma-assisted CVD, or photo-assisted CVD. In thermal-assisted CVD, the energy can be supplied
by resistive heating, RF, and infrared frequency. Thermal-assisted CVD can be accomplished at atmo-
spheric pressure, low pressure (0.01–1.0 kPa), and high vacuum (<10−4 kPa). At atmospheric pressure,
the mass transfer rate and surface reaction rate are similar. But at lower pressure, the mass transfer rate
becomes much higher than the surface reaction rate resulting in uniform film thickness [76]. Thermal-
assisted CVD technique has been used for the deposition of polycrystalline Si films (from SiH4), silicon
nitride films (from SiH2Cl2 and NH3), SiO2 films (from SiH2Cl2 and N2O), W films (from WF6 or WCl6
38-10 Wear Materials

Target Hydrocarbon
Plasma vapor
Lon flux

Substrate
–Vb
Capacitor
RF power
supply

FIGURE 38.5  A schematic drawing of plasma-assisted CVD process.

and hydrogen), and Mo films (from Mo(CO)6) for the semiconductor industry. In the cutting tool indus-
try, the process can be used for deposition of TiN, TiC, and Ti(C,N) films.
Plasma-enhanced CVD, also known as glow-discharge CVD, uses electron energy (plasma) as the
activation method by supplying high electrical power at high voltages to a gas mixture under reduced
pressure (<1.3 kPa). The electrons and ions in the plasma interact with gaseous chemical reactants result-
ing in breaking down of the gas(es) and generation of chemically active ions and radicals. These ions and
radicals undergo chemical reaction at or near the heated substrate surface and deposit films. Figure 38.5
shows a schematic diagram of a plasma-enhanced CVD process. Since activation method is energetic
electrons instead of thermal energy, the substrate temperature is lower. Therefore, it is more suitable for
deposition of films on steel substrate without tempering or loss of hardness, and aluminum alloys for
wear and corrosion protection. However, this technique has two disadvantages. It is difficult to deposit
high purity films because of the incomplete desorption of by-products and the presence of unreacted
species like hydrogen in the film. The second disadvantage is the presence of high compressive stress
that limits the deposition of thick films. Plasma-enhanced CVD process is widely used for deposition of
thin films of TiN, TiC, Ti(C,N), BN, carbon nitride, diamond, and diamond-like carbon (DLC) films.
The PVD and CVD processes are used widely in automotive and cutting tool industries. In the auto-
motive industry, thin film of CrN is used in piston ring application, and DLC thin films are used on
various engine components for racing applications. Hydrogen-free DLC films are deposited on bucket
tappets in engine valvetrain [77]. Piston pins, fuel injectors, and fuel pumps are coated with DLC coat-
ing in diesel engines. TiN coatings on bucket tappets showed a reduction valvetrain friction [78]. In the
cutting tool industry, TiN, TiAlN, TiCN, and CrN coatings deposited by PVD are widely used.
Table 38.1 shows a comparison of thermal spray, electrochemical coating, and vapor deposition pro-
cesses in terms of material used for coating deposition, rate of deposition, and coating properties. The
purpose of the table is to show the overall perspective on differences between various coating deposition
processes.

38.6  Case Hardening of Steels


38.6.1  Nitriding
Nitriding is a surface diffusion process where nitrogen is diffused on the surface to increase surface hard-
ness. The depth of nitrided layer could vary from tens of micrometers to couple of hundred micrometers.
The increased hardness is due to the formation of hard nitride phases. All steels are capable of forming
iron nitrides but some of the alloying elements in steel are strong nitride formers like Al, Cr, etc. Other
nitride formers are vanadium, molybdenum, and tungsten. For nitriding, nitrogen is introduced while
TABLE 38.1  The Differences between Various Coating Deposition Techniques
Thermal Spray Coatings Chemical Coatings Vapor Deposition
Plasma Ni Composite Electrolytic
Coatings and Surface Treatments

Wire Arc High Velocity Oxy-Fuel Plasma Spray Coatings Oxidation PVD CVD
Energy source Plasma Fuel–oxygen mixture Plasma Electrolytic bath Electrolytic bath Energetic ions Energetic ions
Process 6000 3,000–25,000 15,000 50–90 20–50 70–1500 300–2400
temperature, °C
Environment Air Air 10−1 to 10−5 Pa 0.01–100 kPa
Deposition rate, 60 0.5–3 5.0–50 0.02–0.25
μm/min
Coating hardness, GPa 5.0–12 7.0–17 10.0–30 6.0–10 8.0–20 20–40
Coating thickness, μm 200–300 250–400 100–400 2.0–20 20–200 <1–10
Surface roughness 15–45 2.5–5.5 0.5–1.5 0.5–1.0 1.5–2.5 Conforms to substrate roughness
(Ra), μm
Deposition materials Carbon NiCr–CrC, WC–Co, Al2O3, Al2O3–TiO2, ZrO2, Ni–SiC, Ni–BN, Ni Oxides of parent Carbides and nitrides of refractory
steel, Ni, Al WC–CoCr, Cr2O3, TiO2, Cr2O3, Cr2O3–TiO2, Al2O3, Ni–P–B4C, metal metals like Ti, W, Cr, V, Nb, Zr,
Cr2O3–TiO2, Cr3C2–NiCr Cr3C2–NiCr, Ni–50Cr Ni–P–Si3N4 and BN. B4C, SiC, C
38-11
38-12 Wear Materials

the steel is heated between 500°C and 550°C. This process leads to minimum part distortion since no
quenching is required. There are three nitriding processes: (1) gas nitriding, (2) liquid nitriding, and
(3) plasma nitriding [79,80]. Gas nitriding is accomplished in an ammonia environment. There are two
types of gas nitriding process: single stage and double stage. The single-stage gas nitriding is done at
495°C–525°C, resulting in the formation of hard, brittle nitrogen-rich layer. The double-stage gas nitrid-
ing consists of the single-stage nitriding process as the first step followed by raising the temperature to
550°C–565°C. The second stage reduces the case hardness and increases case depth. Liquid nitriding
involves a fused salt bath containing cyanide and cyanates maintained at 565°C. The bath contains
a mixture of sodium salt (60–70 wt%), containing NaCN, Na 2CO3, and NaCNO, and potassium salt
(30–40 wt%), consisting of KCN, K 2CO3, and KCNO. The process can be used for finished parts because
of excellent dimensional stability. The plasma nitriding process utilizes a glow-discharge technology
where elemental nitrogen is introduced on the surface. The process is carried out in a low pressure
vacuum chamber (30–400 Pa) containing a mixture of nitrogen and hydrogen gases. Plasma (consisting
of electrons and ions) is created by applying a high voltage (about 500 V) between the substrate (cathode)
and the wall of the furnace (anode). The interaction of electrons in the plasma with neutral nitrogen
atoms produces nitrogen ions, which are accelerated by the voltage difference toward the substrate and
impinge on the surface resulting in heating the substrate to the nitriding temperature. This results in
the formation of hard thin dense layer of Fe4N and Fe2–3N on the surface [81]. Generally, medium carbon
and chromium-containing alloy steels, 300 and 400 series stainless steels, and precipitation-hardening
stainless steel are nitrided.

38.6.2  Carburizing
Carburizing is a case-hardening process, similar to nitriding, but instead of nitrogen, carbon is diffused
on the surface of steel. Carburizing temperature is typically in the range of 850°C–950°C, although tem-
perature as high as 1095°C is used. The high temperature is required to bring the steel in the austenitic
phase to allow dissolution of more carbon. A carbon gradient is formed from the surface downward
resulting in high hardness at the surface and as the carbon content decreases deeper into the surface
layer, the hardness drops. The amount of carbon that can be diffused at the surface depends on the car-
bon potential of the environment surrounding the steel part. Also, higher the temperature, higher the
diffusion rate of carbon and thus the process time could be shortened.
Carburizing can be done by three processes: (1) gas carburizing, (2) salt bath carburizing, and
(3) pack carburizing. In gas carburizing, the source of carbon is gaseous hydrocarbons, like meth-
ane, propane, or butane. The case depth in gas carburizing is generally in the range of 0.5–1.5 mm.
The majority of the carburized parts are directly cooled in an oil medium forming a martensitic
phase. The rapid cooling creates some distortion of parts and a light machining may be required for
dimensional control. Therefore, it is best to carburize near-finished parts to minimize carburizing
time. Salt bath carburizing involves dipping the parts in a molten salt bath comprising principally
sodium cyanide, sodium carbonate, and sodium chloride in the temperature range 845°C–955°C. The
case depth is generally 0.15–0.25 mm and it contains some level of nitrogen. Nitrogen content can be
avoided by using a non-cyanide bath where a special grade of carbon particles is added in a molten salt
(carbonates). It is believed that the reaction between carbon particles and carbonates generate carbon
monoxide, which reacts with steel surface. In pack carburizing, the parts are packed around carbu-
rizing powders consisting primarily of barium carbonates and calcium carbonates, although sodium
carbonates can also be used. The parts are held at temperature range 815°C–955°C. The chemical
reaction produces carbon monoxide, which decomposes at the surface of the part into nascent carbon
and carbon dioxide. The nascent carbon diffuses into the part and carbon dioxide reacts with the car-
bonaceous material in the carburizing compound to form carbon monoxide. The case depth could be
order of millimeters. Typical steels for carburizing are low carbon steels and alloy steels (4027, 4140,
4620, 8620, and 9310).
Coatings and Surface Treatments 38-13

38.6.3  Carbonitriding
Carbonitriding is a modified form of gas carburizing where ammonia is added in the furnace atmo-
sphere. Like gas nitriding, ammonia dissociates on the surface of steel parts and nascent nitrogen diffuses
into the parts along with carbon resulting in the formation of a hard wear-resistant case. Carbonitriding
process is similar to cyaniding but without the hazards of disposal of toxic molten salts. Carbonitriding
temperature is lower temperature than carburizing, typically in the range of 775°C–855°C and the case
depth is thinner, 0.05–0.75 mm. The process is typically applied to low carbon steels, medium carbon
steels, and alloy steels (4140, 5140, 8640, and 4340).

38.6.4  Induction Hardening


Induction hardening is a process of heat treating steel parts where electrical energy is used for heating.
The basic for induction heating lies in its ability to induce electric current without touching the part.
Therefore, it has the ability to heat selected areas of a component. The electric current should be high
enough to heat parts to austenitic temperature, which is in the range of 725°C–775°C depending on the
carbon content. This is followed by rapid cooling either in water or oil allowing transformation to mar-
tensitic phase. This results in the formation of hard thin wear-resistant case, the depth of which could be
as high as 2 mm. Further details about induction hardening can be found elsewhere [82,83].

38.7  Automotive Trends


The coatings for automotive applications in the near future will be driven primarily by fuel economy
improvement and durability requirements. There are considerable efforts in Europe and North America
for the development and implementation of plasma-transferred wire arc process for deposition of
ferrous-based materials on aluminum cylinder bores replacing cast iron liners. In addition to weight
savings, other benefits are believed to be reduced bore distortion leading to lower oil consumption.
Reduction in bore distortion also enables use of lower tension rings leading to friction reduction and
improved fuel economy. This type of coating has already been introduced by Nissan in limited volume
production [84] and Ford also plans to introduce it in 2011 [85]. This technology is being used today
for remanufacturing of heavy duty diesel engine cylinders [85]. Efforts are also being made to identify
suitable ring face coating for friction reduction. Thermal spray coating on ring face is common in North
America. Several PVD and CVD coatings are being actively looked into for friction reduction including
CrN, and DLC films [86,87]. Applicability of nitrided rings is also being explored [87]. DLC films were
evaluated in the past for racing applications and are now finding applications in mass produced engines.
DLC-coated mechanical bucket tappets in direct active valvetrain architecture are in use for friction
reduction and also for improved wear resistance [77]. DLC-coated piston pins, cam follower system in
fuel pumps, fuel injectors are examples of applications in diesel engine. DLC coatings proved beneficial
in several other applications but implementation is limited due to cost consideration. Plasma arc oxida-
tion of aluminum alloys is also finding applications on aluminum pistons for improved durability.
There is also a trend of increased use of renewable fuel, namely ethanol. The proportion of ethanol
mixed with gasoline fuel is expected to increase in the future. There are evidences of corrosion of some
engine components coming in contact with liquid fuel mixture of ethanol and gasoline and their com-
bustion products. This effect, if observed more widely, may offer more opportunities for use of coatings.
Coatings being developed today should consider compatibility with alternate fuels [88,89].

References
1. D.E. Crawmer, Thermal spray process, Handbook of Thermal Spray Technology, J.R. Davis, Ed., ASM
International, Materials Park, OH, pp. 54–76, 2004.
38-14 Wear Materials

2. K. Bobzin, F. Ernst, K. Richardt, T. Schlaefer, C. Verpoort, and G. Flores, Thermal spraying of cylinder
bores with plasma transferred wire arc process, Surface and Coatings Technology, 202, 4438–4443,
2008.
3. M.P. Planche, H. Liao, and C. Coddet, Relationships between in-flight particle characteristics and
coating microstructure with a twin wire arc spray process and different working conditions, Surface
and Coatings Technology, 182, 215–226, 2004.
4. Q. Hou, Y. He, and J. Gao, Microstructure and properties of Fe–C–Cr–Cu coating deposited by
plasma transferred arc process, Surface and Coatings Technology, 201, 3685–3690, 2006.
5. D.P. Chakravarthy, D.N. Barve, D.S. Patil, P. Mishra, C.S.R. Prasad, and R.K. Rout, Development and
characterization of single wire arc plasma sprayed coatings of nickel on carbon blocks and alumina
tube substrates, Surface and Coatings Technology, 202, 325–330, 2007.
6. A. Abedini, A. Pourmousa, S. Chandra, and J. Mostaghimi, Effect of substrate temperature on the
properties of coatings and splats deposited by wire arc spraying, Surface and Coatings Technology,
201, 3350–3358, 2006.
7. D.H. James, The properties and applications of arc sprayed coatings, Transactions of the Institute of
Metal Finishing, 60(2), 49–53, 1982.
8. M.L. Thorpe, Thermal spray industry, Advanced Materials and Processes, 143(5), 50–61, 1993.
9. M. Buchmann, R. Gadow, and D. Lopez, Advanced hypersonic flame spray coatings for cylin-
der liners in light metal engines, Society of Automotive Engineers Paper No. 2003-01-1099, SAE
International.
10. J.R. Davis, Ed., Handbook of Thermal Spray Technology, ASM International, Materials Park, OH,
2004.
11. J.A. Picas, A. Forn, R. Rilla, and E. Martin, HVOF thermal sprayed coatings on aluminum alloys and
aluminum matrix composites, Surface and Coatings Technology, 200, 1178–1181, 2005.
12. H. Mindivan, Wear behavior of plasma and HVOF sprayed WC–12Co+6% ETFE coatings on
AA2024-T6 aluminum alloy, Surface and Coatings Technology, 204, 1870–1874, 2010.
13. M.P. Nascimento, R.C. Souza, I.M. Miguel, W.L. Pigatin, and H.C. Voorwald, Effects of tungsten car-
bide thermal spray coating by HP/HVOF and hard chromium electroplating on AISI 4340 high
strength steel, Surface and Coatings Technology, 138, 113–124, 2001.
14. L.G. Yu, K.A. Khor, H. Li, K.C. Pay, T.H. Yip, and P. Cheang, Restoring WC in plasma spraying
WC–Co coatings through spark plasma sintering, Surface and Coatings Technology, 182, 308,
2004.
15. L.-M. Berger, S. Saaro, T. Naumann, M. Weiner, V. Weihnacht, S. Thiele, and J. Suchianek,
Microstructure and properties of HVOF-sprayed chromium alloyed WC–Co and WC–Ni coatings,
Surface and Coatings Technology, 202, 4417–4421, 2008.
16. C. Mateus, S. Costil, R. Bolot, and C. Coddet, Ceramic fluoropolymer composite coatings by ther-
mal spraying—A modification of surface properties, Surface and Coatings Technology, 191, 108,
2005.
17. Y. Wang and S. Lim, Tribological behavior of nanostructured WC particles/polymer coatings, Wear,
262, 1097–1101, 2007.
18. U. Akin, H. Mindivan, R. Samur, E.S. Kayali, and H. Cimenoglu, Tribological behavior of combined
coatings deposited on aluminium, Key Engineering Materials, 280–283, 1453, 2005.
19. E. Celik, O. Culha, B. Uyulgan, N.F. Ak Azem, I. Ozdemir, and A. Turk, Assessment of microstruc-
tural and mechanical properties of HVOF sprayed WC-based cermet coatings for a roller cylinder,
Surface and Coatings Technology, 200, 4320–4328, 2006.
20. J. Saaedi, T.W. Coyle, S. Mirdamadi, H. Arabi, and J. Mostaghimi, Phase formation in a Ni–50Cr
HVOF coating, Surface and Coatings Technology, 5804–5811, 2008.
21. J. Stokes and L. Looney, Residual stress in HVOF thermally sprayed thick deposits, Surface and
Coatings Technology, 177–178, 18, 2004.
Coatings and Surface Treatments 38-15

22. L. Zhao, M. Mauer, F. Fischer, R. Dicks, and E. Lugscheider, Influence of spray parameters on the par-
ticle in-flight properties and the properties of HVOF coating of WC–CoCr, Wear, 257, 41–46, 2004.
23. M. Hassan, J. Stokes, L. Looney, and M.S.J. Hasami, Effect of spray parameters on residual stress
build-up of HVOF sprayed aluminum/tool-steel functionally graded coatings, Surface and Coatings
Technology, 202, 4006–4010, 2008.
24. T. Varis, E. Rajakami, and K. Korpiola, Mechanical properties of thermal spray coatings, Thermal
Spray 2001, New Surface for a New Millennium, C.C. Bendt, K.A. Khor, and E.F. Lugscheider, Eds.,
ASM International, Materials park, OH, pp. 993–997, 2001.
25. W. Hu, M. Li, and M. Fukumoto, Preparation and properties of HVOF NIAl nanostructured coat-
ings, Materials Science and Engineering A, 478, 1–8, 2008.
26. M. Buchmann and R. Gadow, Tribologically optimized ceramic coatings for cylinder liners in
advanced combustion engines, Society of Automotive Engineers Paper No. 2001-01-3548, SAE
International.
27. F. Rastegar and A.E. Craft, Piston ring coatings for high horsepower diesel engines, Surface and
Coatings Technology, 61, 36–42, 1993.
28. J.A. Picas, A. Forn, and G. Matthaus, Hard chrome coating alternatives for pistons and valves, Society
of Automotive Engineers Paper No. 2004-05-0211, SAE International.
29. G. Barbezat, The state of the art of the internal plasma spraying on cylinder bore in AlSi cast alloys,
Society of Automotive Engineers Paper No. 2000-05-0068, SAE International.
30. E.P. Song, J. Ahn, S. Lee, and N.J. Kim, Microstructure and wear resistance of nanostructured
Al2O3–8 wt% TiO2 coatings plasma-sprayed with nanopowders, Surface and Coatings Technology,
201, 1309–1315, 2006.
31. M. Gell, E.H. Jordan, Y.H. Sohn, D. Goberman, L. Shaw, and T.D. Xiao, Development and imple-
mentation of plasma sprayed nanostructured ceramic coatings, Surface and Coatings Technology,
146–147, 48–54, 2001.
32. P. Bansal, N.P. Padture, and A. Vasiliev, Improved interfacial mechanical properties of Al2O3–13 wt%
TiO2 plasma-sprayed coatings derived from nanocrystalline powders, Acta Materialia, 51(10), 2959–
2970, 2003.
33. E.P. Song, J. Ahn, S. Lee, and N.J. Kim, Effects of critical plasma spray parameter and spray distance
on wear resistance of Al2O3–8 wt.% TiO2 coatings plasma sprayed with nanopowders, Surface and
Coatings Technology, 202, 3625–3632, 2008.
34. A. Ibrahim, Z. Abdel Hamid, and A. Abdel Aal, Investigation of nanostructured and conventional
alumina–titania coatings prepared by air plasma spray process, Materials Science and Engineering A,
527, 663–668, 2010.
35. A. Datye, S. Koneti, G. Gomes, K.-S. Wu, and H.-T. Lin, Synthesis and characterization of alu-
minum oxide–boron carbide coatings by air plasma spraying, Ceramics International, 36, 1517–
1522, 2010.
36. M. Tricard, J. Hagan, P. Redington, K. Subramanian, and M. Haselkorn, Development of plasma spray
coated cylinder liners, Society of Automotive Engineers Paper No. 960048, SAE International.
37. P. Scardi, L. Lutterotti, and E. Galvanetto, Microstructural characterization of plasma sprayed zirconia
thermal barrier coatings by x-ray diffraction full pattern analysis, Surface and Coatings Technology,
61, 52–59, 1993.
38. P.M. Pierz, Thermal barrier coating development for diesel engine aluminum pistons, Surface and
Coatings Technology, 61, 60–66, 1993.
39. V.D.N Rao, D.M. Kabat, R. Rose, D. Yeager, R. Brandt, and D.Y. Leong, Performance of plasma spray
coated bore 4.6L–V8 aluminum block engines in dynamometer and fleet vehicle durability tests,
Society of Automotive Engineers Paper No. 980008, SAE International.
40. K. Funatani, I. Kurusawa, A. Fabiyi, and M.F. Puz, Engine performance improvements, Automotive
Engineering, I, 15–20, 1995.
38-16 Wear Materials

41. L. Shi, C.F. Sun, F. Zhou, and W.M. Liu, Electrodeposited nickel–cobalt composite coating containing
nano-sized Si3N4, Materials Science and Engineering A, 397, 190–194, 2005.
42. H.-K. Lee, H.-Y. Lee, and J.-M. Jeon, Codeposition of micro and nanosized SiC particles in a nickel
matrix composite coatings obtained by electroplating, Surface and Coatings Technology, 201, 4711–
4717, 2007.
43. A.F. Zimmerman, D.G. Clark, K.T. Aust, and U. Erb, Pulse electrodeposition of Ni–SiC nanocom-
posite, Materials Letters, 52, 85–90, 2002.
44. A.F. Zimmerman, G. Palumbo, K.T. Aust, and U. Erb, Mechanical properties of nickel silicon car-
bide nanocomposite, Materials Science and Engineering A, 328, 137–146, 2002.
45. K.-H. Hou, W.-H. Hwu, S.-T. Ke, and M.-D. Ger, Ni–P–SiC composite produced by pulse and direct
current plating, Materials Chemistry and Physics, 100, 54–59, 2006.
46. J. Alexis, B. Etcheverry, J.D. Beguin, and J.P. Bonino, Structure, morphology and mechanical properties
of electrodeposited composite coating Ni–P/SiC, Materials Chemistry and Physics, 120, 244–250,
2010.
47. D.H. Cheng, W.Y. Xu, L.Q. Hua, Z.Y. Zhang, and X.J. Wan, Electrochemical preparation and
mechanical properties of amorphous nickel–SiC composites, Plating and Surface Finishing, 85(2),
61–64, 1998.
48. L. Ploof, Electroless nickel composite coatings, Advanced Materials and Processes, May, 36–38,
2008.
49. S.M. Monir Vaghefi, A. Saatchi, and M. Ebrahimian-Hoseinabadi, Deposition and properties of
electroless N–P–B4C composite coatings, Surface and Coatings Technology, 168, 259–262, 2003.
50. C.M. Das, P.K. Limaye, A.K. Grover, and A.K. Suri, Preparation and characterization of silicon
nitride codeposited electroless nickel composite coatings, Journal of Alloys and Compounds, 436,
328–334, 2007.
51. A. Grosjean, M. Rezrazi, J. Takadoum, and P. Bercot, Hardness, friction and wear characteristics of
nickel–SiC electroless composite deposits, Surface and Coatings Technology, 137, 92–96, 2001.
52. G.C. Wood and J.P. O’Sullivan, The anodizing of aluminum in sulphate solutions, Electrochimica
Acta, 15, 1865–1876, 1970.
53. J.P. O’Sullivan and G.C. Wood, The morphology, and mechanism of formation of porous anodic
films on aluminum, Proceedings of Royal Society of London A, 317, 511–543, 1970.
54. J. Rasmussen, New insights into the microhardness of anodized aluminum, Metal Finishing, 99,
46–51, September, 2001.
55. E. Matykina, R. Arrabal, A. Mohamed, P. Skeldon, and G.E. Thompson, Plasma electrolytic oxida-
tion of pre-anodized aluminium, Corrosion Science, 51, 2897–2905, 2009.
56. A.L. Yerokhin, X. Nie, A. Leyland, A. Matthews, and S.J. Dowey, Plasma electrolysis for surface
engineering, Surface and Coatings Technology, 122, 73–93, 1999.
57. A. Kuhn, Plasma anodized aluminum—A 2000/2000 ceramic coating, Metal Finishing, 100, 44–50,
November/December, 2002.
58. N. Godja, N. Kiss, Ch. Locker, A. Schindel, A. Gavrilovic, J. Wosik, R. Mann, J. Wendrir, A. Merstallinger,
and G.E. Nauer, Preparation and characterization of spark-anodized Al alloys: Physical, chemical
and tribological properties, Tribology International, 43, 1253–1261, 2010.
59. I.L. Song-Park, T.G. Woo, W.Y. Jeon, H.H. Park, M.H. Lee, T.S. Bae, and K.W. Seol, Surface char-
acteristics of titanium anodized in the four different types of electrolyte, Electrochimica Acta, 53,
863–870, 2007.
60. C.E. Barchiche, E. Rocca, C. Juers, J. Hazan, and J. Steinmetz, Corrosion resistance of plasma-anodized
AZ91D magnesium alloy by electrochemical methods, Electrochimica Acta, 53, 417–425, 2007.
61. Y.L. Cheng, H.I. Wu, Z.H. Chen, H.I. Wang, Z. Zheng, and Y.W. Wu, Corrosion properties of
AZ31 magnesium alloy and protective effects of chemical conversion layers and anodized coatings,
Transactions of Nonferrous Metals Society of China, 17, 502–508, 2007.
Coatings and Surface Treatments 38-17

62. R. Arrabal, A. Pardo, M.C. Merino, M. Mohedano, P. Casajus, E. Matykina, and P. Skeldon, Corrosion
behavior of a magnesium matrix composite with a silicate plasma electrolytic oxidation coating,
Corrosion Science, 52, 3738–3749, 2010.
63. M.-R. Yang and S.K. Wu, The improvement of high temperature oxidation of Ti–50Al by anodic
coating in the phosphoric acid, Acta Materialia, 50, 691–701, 2002.
64. E.A. Ferreira, R.C. Rocha-Filho, S.R., Biaggio, and N. Bocchi, Corrosion resistance of the Ti–50Zr
at.% alloy after anodization in different acidic electrolytes, Corrosion Science, 52, 4058–4063, 2010.
65. W.D. Sproul, Physical vapor deposition of tool coatings, Surface Coatings and Technology, 81, 1–7,
1996.
66. D.M. Mattox, Handbook of Physical Vapor Deposition Processes, Noyes Publications, New York, 1998.
67. D.M. Sanders and A. Anders, Review of cathodic arc deposition technology at the start of millen-
nium, Surface and Coatings Technology, 133–134, 78–90, 2000.
68. N. Cansever, Properties of niobium nitride coatings deposited by cathodic arc physical vapor
deposition, Thin Solid Films, 515, 3670–3674, 2007.
69. R.A. MacGill, S. Anders, A. Anders, R.S. Castro, M.R. Dickinson, K.M. Yu, and I.G. Brown, Cathodic
arc deposition of copper oxide thin films, Surface and Coatings Technology, 78, 168–172, 1996.
70. S.K. Kim, P.V. Vinh, J.H. Kim, and T. Ngoc, Deposition of superhard TiAlSiN thin films by cathodic
are plasma deposition, Surface and Coatings Technology, 200, 1391–1394, 2005.
71. J.L. Vossen and W. Kern, Eds., Thin Film Processes II, Academic Press, London, U.K., 1991.
72. G. Brauer, B. Szyszka, M. Vergohl, and R. Bandorf, Magnetron sputtering—Milestones of 30 years,
Vacuum, 84, 1354–1359, 2010.
73. R.D. Arnell and P.J. Kelly, Recent advances in magnetron sputtering, Surface and Coatings Technology,
112, 170–176, 1999.
74. G. Contoux, F. Cosset, A. Celerier, and J. Machet, Deposition process study of chromium oxide thin
films obtained by d.c. magnetron sputtering, Thin Solid Films, 292, 75–84, 1997.
75. K.L. Choy, Chemical vapor deposition of coatings, Progress in Materials Science, 48, 57–170, 2003.
76. R.F. Bunshah, Ed., Handbook of Hard Coatings, Noyes Publications, Park Ridge, NJ, 2001.
77. S. Okuda, T. Dewa, and T. Sagawa, Development of 5W-30 GF-4 fuel saving engine oil for DLC
coated valve lifters, Society of Automotive Engineers Paper No. 2007-01-1979, SAE International.
78. M. Masuda, M. Ujino, K. Shimoda, K. Nishida, I. Marumoto, and Y. Moriyama, Development of tita-
nium nitride coated shim for a direct acting OHC engine, Society of Automotive Engineers Paper
No. 970002, SAE International.
79. H.E. Boyer, Ed., Case Hardening of Steel, ASM International, Materials Park, OH, 1987.
80. R. Davis, Ed., Surface Hardening of Steels, JASM International, Materials Park, OH, 2002.
81. T. Schuelke and T. Krug, Plasma assisted surface technologies to improve tool properties, Society of
Automotive Engineers Paper No. 982334, SAE International.
82. S.L. Semiatin and D.E. Stutz, Induction Heat Treatment of Steels, American Society of Metals,
Materials Park, OH, 1986.
83. T.H. Spencer, Induction Hardening and Tempering, American Society of Metals, Materials Park, OH,
1964.
84. H. Ichikawa, N. Nakada, and J. Yajima, The new high performance V6 gasoline turbocharged engine
from NISSAN, Society of Automotive Engineers Paper No. 2009-01-1067, SAE International.
85. Ford launches plasma-sprayed cylinder bores on 2011 Shelby V8, Automotive Engineering online,
February 17, 2010.
86. D.L. Martinez, M. Valverde, R. Rabute, and A. Ferrarese, Ring packs for friction optimized engines,
MTZ, 71(7–8), 26–29, 2010.
87. R. Wakabayashi, K. Mochiduki, and H. Yoshida, Lubricating condition of piston ring and cylinder for
significantly reducing piston friction loss, Society of Automotive Engineers Paper No. 2009-01-0188,
SAE International.
38-18 Wear Materials

88. P.J.R. Mordente and R.A. Bruno, Top ring technology to low fuel consumption flex fueled engines,
Society of Automotive Engineers Paper No. 2009-36-0128, SAE International.
89. A. Ferrarese, G. Marques, E. Tomanik, R. Bruno, and J. Vatavuk, Piston ring tribological challenges
on the next generation of flex-fuel engines, Society of Automotive Engineers Paper No. 2010-01-
1529, SAE International.
39
Low Friction Coatings
39.1 Introduction and Definition of Low Friction.............................. 39-1
39.2 Low Friction Coatings and Mechanisms of Lubrication...........39-3
Transition Metal Dichalcogenides  •  Carbon-Based Materials  • ​
Polymers  •  Soft Metals  •  Other Low Friction Coatings
39.3 Coating Processing Routes........................................................... 39-14
39.4 Key Applications............................................................................ 39-14
Thomas W. Scharf Acknowledgments..................................................................................... 39-15
University of North Texas References................................................................................................... 39-15

39.1  Introduction and Definition of Low Friction


Friction mitigation is typically accomplished by introducing a shear-accommodating layer (e.g., a thin
film of liquid) between contacting surfaces. When the operating conditions are beyond the liquid realm,
attention turns to solid coatings. According to the classical theory of Bowden and Tabor [1], friction force,
F, is a product of the contact area and the shear strength of the lubricant material, A·τ (see Figure 39.1).
Thus, the friction coefficient, μ, can be expressed by

F A ⋅t t t
m= = = = o +a (39.1)
L L PH PH

where
L is the normal force (load)
PH is the mean Hertzian contact pressure
τo is the interfacial shear strength, a “velocity accommodation parameter” which is a property of the
interface
α represents the pressure dependence of the shear strength

The constant “α” is the lowest attainable friction coefficient for a given friction couple. In principle, a
hard material with a soft “skin” ought to provide a low friction coefficient by reducing το and increasing
PH (low A). The Bowden and Tabor concept was verified for steel sliding against a soft metallic coating
of indium on steel substrate (see Chapter 35 for more details).
For a “sphere-on-flat” elastic contact, which is known as the Hertzian elastic contact model, this fric-
tion coefficient can be expressed as

39-1
39-2 Wear Materials

L Counterface
Environment
Wear debris

Wear surface Transfer film

Coating A
F

FIGURE 39.1  Schematic illustration of a hemispherical ball sliding on a coated substrate with the wear surface in
contact with the transfer film on the counterface ball.

2 /3
⎛ 3R ⎞
m = to p ⎜ ⎟ L−1/ 3 +a (39.2)
⎝ 4E ⎠

where
R is the sphere radius
E is the equivalent (composite) Young’s modulus

Instead of the Amontonian first law of friction, where μ is independent of L, the Bowden and Tabor
analysis for Hertzian contacts predicts

m∝L−1/ 3 (39.3)

Thus, when contact deformation is elastic, the friction coefficient will decrease with increasing nor-
mal load (or mean Hertz pressure). The linear relationship between L−1/3 and μ (Equation 39.2) has
been experimentally verified for a number of low friction coatings, including diamond-like nanocom-
posite (DLN) [2], MoS2/Sb2O3/Au [3], and ultrananocrystalline diamond (UNCD, unpublished) (see
Figure 39.2). For thin and soft coatings, the pressure is primarily supported by the substrate and increas-
ing the substrate modulus and hardness will decrease the contact area for a given normal load. Thus,

0.25
50% RH air τo= 73 MPa; α = 0.02
Dry nitrogen DLN
Least squares fit
0.20 τo= 38 MPa; α = 0.057
MoS /Sb O /Au
Friction coefficient

2 2 3

0.15 τ = 42 MPa; α = 0.022


o
UNCD
τo= 26 MPa;
0.10 α = 0.015 UNCD
τo= 20 MPa; α = 0.007
MoS2/Sb2O3/Au

0.05 τo= 9 MPa; α = 0.007


DLN

0.00
0 2 4 6
Inverse Hertzian pressure (1/GPa)

FIGURE 39.2  Linear regression fits for friction coefficient as a function of inverse Hertzian pressure fitted to
μ = (τo/PH) + α with values of shear strength (τo) shown for three low friction, environmentally robust coatings:
Si3N4 on diamond-like nanocomposite (DLN) coating, Si3N4 on MoS2/Sb2O3/Au, and ultrananocrystalline diamond
(UNCD)-coated Si3N4 on UNCD, and in both dry nitrogen and humid air.
Low Friction Coatings 39-3

the ideal scenario for achieving low friction is to have an elastically stiff and hard substrate support the
normal load and keep the contact area small, while the surface coating provides shear accommodation
and reduces junction strength, until the substrate begins to yield and plastically deform.
These are necessary but not totally sufficient conditions for most low friction and wear coatings.
In addition, compliant transfer films (third bodies) that adhere to the counterface (see Figure 39.1)
or transfer layers that adhere on the wear surface (track), often tribochemically, can lower το further
and provide long wear life by preventing native counterface-coating contact while accommodating
interfacial shear. The role of the transfer film is twofold. First, a mixture of inference and in situ fric-
tion and wear testing evidence [4] suggests that the transfer film covers the area of contact preventing
direct contact between the sliding surfaces and allowing the counterface (e.g., sphere) to behave as an
ideal “spherical tip,” which makes the tip behave as an ideal elastic sphere according to Equation 39.2.
Second, and more importantly, the low friction and long wear life rely on the transfer film to provide
low shear at or near the sliding interface. If the transfer film wears away, then the friction coefficient
increases and the contacting surfaces can eventually seize, gall, etc., unless of course the transfer film
is replenished. Therefore, low friction coatings do not necessarily have to exhibit low inherent shear
strength in order to impart low friction. While necessary, the coating’s low shear strength cannot
account for either the Hertzian behavior (Equation 39.2), or the long wear life of a low friction coating.
Though most low friction, solid lubricant coatings function by transfer film processes, there are a few
exceptions, for example, near frictionless carbon (NFC) [5], that do not appear to form transfer films
to sustain low friction.
Friction and wear do not just rely on physical and mechanical properties of individual materials that
come into contact; instead, they are systems properties involving interactions within pairs of contacting
surfaces and between them and the environment. Many tribological contacts result in the aforemen-
tioned transfer of material from the coating surface to the counterface and surface chemical reactions
with the surrounding environment (Figure 39.1), resulting in wear surfaces whose chemistry is signifi-
cantly different from the bulk, known as tribochemical reactions. As can be seen from the following
classes of low friction coatings, environment, load, sliding speed, etc., play a significant role in deter-
mining their tribological behavior. Coatings that give extremely low friction and long wear life in one
environment can fail to do so in a different environment; for example, MoS2 exhibits long wear life in
dry nitrogen but wears away in humid air. In addition, no low friction coating, as of yet, can act as a
solid lubricant in all environments and under all operating conditions. Details of commonly used low
friction coatings, some of which are referred to as solid lubricant coatings, are listed in Table 39.1, and
are described in further detail in the following sections.

39.2  Low Friction Coatings and Mechanisms of Lubrication


There are many classes of low friction coating materials: (a) transition metal dichalcogenides such as
MoS2 and WS2; (b) carbon-based materials like graphite, diamond-like carbon (DLC), and nanocrystal-
line diamond; (c) polymers, such as polytetrafluoroethylene, PTFE; and (d) soft metals like silver, tin,
gold, and their alloys. Barring a few exceptions, such as PTFE and its composites, most of these materi-
als can be applied as thin coatings on tribological components (bearings, seals, magnetic hard drives,
etc.) to reduce friction, wear, and debris generation. Low friction coatings can be broken down into
single-phase (elements or compounds) and multiphase composites.

39.2.1  Transition Metal Dichalcogenides


39.2.1.1  Molybdenum Disulfide and Tungsten Disulfide
Among the various members of the transition metal dichalcogenides family of compounds, MoS2
and WS2 are well known for their low friction/solid lubricating behavior and are widely used in many
TABLE 39.1  Details of Commonly Used Low Friction Coatings, Some of Which Are Referred to as Solid Lubricant Coatings (See Chapter 29)
39-4

Common Typical Steady-State


Coating Friction Coefficients, μ Approximate
Deposition Thickness (Environment/Normal Operating Ranges
Low Friction Coating Method(s) Range (μm) Load) (°C) Applications Comments
Single phase (elements or compounds)
Graphite (sp2 bonding)* Evaporation, 0.2–5 0.5–0.6 (dry nitrogen or −200 up to 400 in Bearings, seals, pump and *Often referred to as a glassy-
pyrolysis of high vacuum/5 to 1 N)** air*** valve parts, piston skirts, carbon coating
hydrocarbon 0.1–0.2 (humid air/5 to brushes for motors **Basically only effective in the
polymers 1 N) presence of water vapor and
oxygen
***Increasing μ values with
increasing temperatures
Diamond-like carbon rf and dc sputtering, 0.005–1 Extremely variable** −200 up to 400 in Magnetic media (protect *DLC is an amorphous (a-C) or
(mixed sp2/sp3 ion-beam, chemical 0.6–0.7 a-C (dry nitrogen air*** head slider and disk), hydrogenated amorphous
bonding with majority vapor deposition or high vacuum/10 N) razor blades, bearings, carbon (a-C:H) coating
being metastable sp3 (all involve highly 0.001-0.05 a-C:H (dry bushings, seals, gears, car (typically 10–50 atomic % H).
unless it is stabilized energetic ions) nitrogen or high engines (e.g., fuel The various types are typically
with C–H bonds)* vacuum/10 N) injectors, camshafts, valve classified according to their
0.1–0.2 a-C (humid air tappets, and piston pins/ fraction of sp2, sp3 bonding and
30% RH/ 10 N) rings), femoral heads for hydrogen from a ternary phase
0.2–0.3 a-C:H (humid air hip implants diagram
30% RH/10 N) **Of all the low friction coatings,
DLCs are the most dependent
on loading and environment
***Increasing temperature from
300°C to 400°C causes increase
in friction in air likely due to
oxidation of carbon
Tetrahedral amorphous Mass selected ion 0.01–1 0.7 (dry nitrogen/10 N) −200 up to 400 in Mechanical valve parts, *Known as ta-C coatings unless
carbon (sp3 bonding)* beam, filtered 0.1 (humid air 50% air seals, cams, etc. in the presence of hydrogen
cathodic arc, pulsed RH/10 N) gas→ta-C:H that are deposited
laser deposition with very high density plasmas
(ion or plasma generated by electron cyclotron
beams with a high resonance, inductively coupled
ion fraction and ion plasma sources
energy of ∼100 eV)
Wear Materials
Ultrananocrystalline Microwave plasma 0.5–1.5 0.05–0.13 (dry Unknown Mechanical shaft seals in *UNCD coatings have 2–5 nm
diamond (sp3 chemical vapor nitrogen/1 N)** rotary equipment, such grain sizes, hence the term
bonded)* deposition 0.03–0.1 (humid air 45% as pumps and mixers; “ultranano”
RH/1 N) AFM tips; rf MEMS **Slightly higher μ values in dry
resonators; piezoresistive nitrogen than humid air.
microcantilever; However, μ values in dry
biomedical devices/ nitrogen are considerably lower
sensors than DLC, graphite, and other
carbon-based coatings
Low Friction Coatings

MoS2 and WS2 rf and dc sputtering, 0.2–2 0.02–0.06 (dry nitrogen −150 up to 350 Satellite components *Increasing normal load results
air spray (resin or high vacuum/5 to (MoS2) and 400 (bushings, shafts, reaction in decreasing μ
bonded), 1 N)* (WS2) in air. In wheels, gyroscopes, solar **Basically only effective in the
impingement (N2 0.15–0.25 (initial) then N2 or vacuum, arrays, antenna drives, absence of water vapor and
blast), burnishing, increasing values due to up to 800*** gears, pumps, actuators, oxygen
evaporation, eventual coating failure latches, releases, etc.) ***Temperatures in the extreme
chemical vapor (humid air >1% RH/5 to ranges (300–400 in air and
deposition, atomic 1 N)** lower −150 in N2 or vacuum)
layer deposition generally result in increasing μ
values; while in intermediate
ranges from −50 to 100 μ is
athermal
Polytetrafluoroethylene Powder jet milling, 2–10 0.02–0.15 (humid air −70 up to 200 in Bearings (e.g., cages in *Major glass transition
(PTFE)* spray processes, 35%–45% RH, N2 and air, N2 or some types of rolling temperature is ∼126°C and
sputtering, hot vacuum/1 N)** vacuum*** element bearings), release melting temperature is ∼327°C
filament chemical molds, satellite **Low friction, but wears;
vapor deposition components (revolute moderate load capacity and
joints in gears), higher friction at elevated
automotive (seat belt temperatures. Also higher
clips, fasteners, throttle friction at high sliding speeds,
body shafts) e.g., 14 mm/s, and low loads
***Friction coefficient increases
(∼0.07 to 0.21) with decreasing
temperature from 400 to 200 K
in N2 (2%–6% RH)
(continued)
39-5
39-6

TABLE 39.1 (continued)  Details of Commonly Used Low Friction Coatings, Some of Which Are Referred to as Solid Lubricant Coatings (See Chapter 29)
Common Typical Steady-State
Coating Friction Coefficients, μ Approximate
Deposition Thickness (Environment/Normal Operating Ranges
Low Friction Coating Method(s) Range (μm) Load) (°C) Applications Comments
Nanocrystalline ZnO* Pulsed laser 0.1–1 0.1–0.25 (humid air 50% 25–400 in air*** So far very limited *Type of lubricious oxide with
deposition, rf RH and dry nitrogen/5 commercial use in microstructurally engineered
sputtering, atomic to 1 N)** tribological applications, defective grain structure to
layer deposition but may be a suitable achieve low friction
high temperature **μ values are typically insensitive
lubricant if no grain to environment. Increasing
coarsening and limited normal load results in
diffusion occur decreasing μ
***400°C annealed coatings
resulted in lower μ values
Hexagonal boron Ion plating, rf 2–5 0.2–0.3 (humid air 40% 25–500 in air Spray coatings used in *Similar crystal structure to
nitride (h-BN)* sputtering, RH/2 N) steel processing rails, die graphite but has higher friction
chemical vapor 0.6 (vacuum/0.1 N) castings, hardware in air environments. Not to be
deposition, aerosol fasteners in vacuum confused with the much harder
spray polymorphic cubic BN
Boric acid (H3BO3)* Forms on boric B2O3:1–2 0.08–0.2 (humid air 50% 25–750 in air Widely used as a lubricant *Boric acid is a crystalline-
oxide (B2O3) H3BO3: RH/10 to 1 N)*** (μ values or grease additive. Also layered solid lubricant
coatings deposited 0.01–0.05 Note: vacuum and dry decrease with being explored as an in **Boric acid lamellar films form
by evaporation or conditions are not increasing situ powder delivery and from a spontaneous chemical
atomic layer applicable since water is temperature as a nanoparticle additive reaction between water
deposition** necessary for H3BO3 from 450 to 750) to lubricants molecules and boric oxide
reaction (B2O3) coatings. Films are very
thin (tens of nm) but can
self-replenish
***Increasing normal load results
in decreasing μ
Wear Materials
Soft metals (Ag, Au)* Electrochemical, dc 0.2–20 0.2 (vacuum), 0.43 (air, Active at higher Ag films on high speed *Soft metals are often
sputtering, ion 40% RH), 0.23 (dry temperatures ball bearings used in co-deposited with other phases
plating, burnishing, nitrogen)/1 N (300–600) in air x-ray tubes in vacuum; to act as a high temperature
evaporation, Ion-plated Ag film dopant used in satellite solid lubricant due to increased
electroplating 0.12–0.2 (humid air/18 N applications diffusivity. Elemental or doped
at 650°C) Pb was once considered but is
Ag film 50 μm thick being phased out due to
0.15 (vacuum) environmental concerns
0.39 (air, 40% RH)
Low Friction Coatings

0.48 (dry nitrogen)/1 N


Ion plated Pb film

Multi-phase composites
MoS2/Sb2O3/Au* rf and dc sputtering, 0.5–2 0.01 (dry nitrogen/1 N)** −100 up to 100 in Satellite components (e.g., *A variation of this coating is Au
air spray (resin 0.15 (humid air 50% N2 or launch lock mechanism) replaced with C
bonded), RH/1 N)*** vacuum**** **Increasing normal load results
impingement (N2 in decreasing μ
blast) ***Effective in both dry and
humid environments although
Au islands form inside wear
tracks in humid air leading to
increased μ
****μ is athermal in the range of
−100°C to 100°C
W-C-S* Sputtering (WS2) 0.5–2 0.05 (vacuum/1 N) Unknown Potentially in satellite *Typical composition (at.%)
and pulsed laser 0.15–0.2 (humid air 50% components W20C60S20, whose structure =
deposition (C) RH/1 N)** WC, WS2, and DLC
**Effective in both dry and humid
environments although
conclusive synergistic effects of
C and WS2 remain unknown
(continued)
39-7
39-8

TABLE 39.1 (continued)  Details of Commonly Used Low Friction Coatings, Some of Which Are Referred to as Solid Lubricant Coatings (See Chapter 29)
Common Typical Steady-State
Coating Friction Coefficients, μ Approximate
Deposition Thickness (Environment/Normal Operating Ranges
Low Friction Coating Method(s) Range (μm) Load) (°C) Applications Comments
YSZ-Au-MoS2-DLC* Hybrid magnetron 1–2 0.05 (dry air and N2 4% 25–500 in air*** Potentially in satellite *Typical composition (at.%)
sputtering (Au) and RH, and vacuum/6 N)** components YSZ0.34Au0.24(MoS2)0.17C0.24
pulsed laser 0.1–0.15 (humid air ** Effective in both dry and
deposition (YSZ, 25–45% RH/6 N)** humid environments
MoS2, and DLC) ***μ increases slightly to ∼0.2 at
500°C and remains there after
returning to 25°C since Au
islands form inside the wear
tracks
PTFE/nano-Al2O3* Powder jet milling 5–10 0.16–0.2 (humid air 50% Unknown Potentially in satellite *∼40 nm Al2O3 particles. Difficult
RH/250 N = 6.3 MPa components to bond as a coating
stress)** **While μ increases with Al2O3
addition, and other fillers in
general (e.g., Ni), the wear
resistance is better than pure
PTFE, although still a moderate
wear composite
VN-Ag* Unbalanced dc 2–3 0.35, 0.30, 0.10, and 0.20 25–1000 in air** Potentially in aerospace *∼30–40 at.% Ag
magnetron [at 25°C, 350°C, 700°C, components that require **High temperature
co-sputtering and 1000°C, respectively high temperature 500°C–1000°C in situ Raman
(humid air ∼60% protection tribometry testing revealed
RH /2 N)] tribochemical silver vanadate
phases (i.e., Ag3VO4, AgVO3)
form and are responsible for
μ∼0.1–0.2
Ni-Mo-Al/Cr2O3/Ag/ Plasma spray 200–400 0.31, 0.16, and 0.21 [at 25–650 in air Aerospace components, *Denoted as NASA PS400 with
(Ba,Ca)F2 eutectic* 25°C, 500°C, 650°C, e.g., turbine engine shaft, Ni-Mo-Al as binder/matrix
respectively (humid air that require high **Corresponding wear rates are
∼50% RH/5 N, and 3 m/s temperature protection 1180, 6.3, and 7.6 (×10–6) mm3/
sliding velocity)]** Nm, respectively
Wear Materials
Doped DLC (light elements)
Si-DLC Plasma-enhanced 0.5–2 0.07 6 atomic % Si (humid Unknown Camshafts, piston pins, *Increasing atomic % Si above 6%
chemical vapor air 50% RH/10 N)* and valve tappets results in increasing μ
deposition,
co-sputtering
SiO-DLC* Plasma-enhanced 0.5–2 0.02 (dry nitrogen <1% Unknown Camshafts, piston pins, *These amorphous coatings are
chemical vapor RH/0.1 N) and valve tappets often referred to diamond-like
deposition, 0.05 (dry air <1% nanocomposite (DLN) coatings
Low Friction Coatings

co-sputtering RH/0.1 N) with typical composition of


0.22 (humid air 50% (CH 0.15)0.7 (SiO0.3)0.3
RH/0.1 N)** **Increasing normal load results
0.05 (humid air 50% in decreasing μ
RH/10 N)**
N-DLC* Plasma-enhanced 0.01–1 0.16 (humid air 30% 25 up to 300 in Magnetic media (protect *Often referred to as carbon
chemical vapor RH/0.1 N) air head slider and disk) nitride (CNx) coatings
deposition, 0.01–0.05 (dry nitrogen
sputtering, ion or high vacuum/0.1 N)
beam assisted
deposition

Doped DLC (transition metals)


W-DLC, Ti-DLC, Hybrid reactive 0.2–2 W-DLC 0.15 (humid air 25 up to 300 in Bearings (inner/outer *Transition metal doped DLCs
Cr-DLC* magnetron 50% RH/10 N) air races and rollers) coatings that contain
sputtering 0.04 (dry air 2% RH/10 N) nanocrystalline metal (M)
(transition metal Ti-DLC carbides (WxCy,TixCy,CrxCy)
carbide) and 0.13 (humid air 50% incorporated into the a-C:H
chemical vapor RH/10 N) matrix. These coatings are
deposition (DLC) 0.05 (dry air 2% RH/10 N) referred to as MC/a-C:H,
M-C:H, or M-DLC. An
advantage of MC/a-C:H films
over other hard coatings, such as
metal nitrides, is their lower
friction and elastic modulus
Note: The friction coefficient values are from room temperature testing, unless stated, and sliding speeds in the range of ∼1 to 4 mm/s. Friction coefficient values in the presence of a liquid
lubricant are not provided, or they are unknown/unmeasured, but typically result in decreased friction coefficient values unless in the boundary lubrication regime. This list is by no means
exhaustive but does represent the most widely used low friction coatings as well as some novel exploratory coatings especially for use at high temperatures.
39-9
39-10 Wear Materials

Strong Strong

Weak

6.7 Å
6.15 Å
S
Mo (or W)
S
(a) (b)

FIGURE 39.3  Lamellar crystal structures of (a) Mo (or W) disulfide and (b) graphite. Both have hexagonal lat-
tices with basal planes held together with strong covalent bonds, while in between planes are held with weak van
der Waals cohesive forces. This gives rise to interlamellar mechanical weakness and hence easy shear between the
planes.

Shear

(a) Basal plane Basal plane (b)

FIGURE 39.4  Schematic representations of (a) two crystallographic textures with basal planes perpendicular and
parallel (preferred) to the substrate and (b) the process of shear-induced reorientation of perpendicular basal planes
parallel to the sliding direction to achieve low friction.

practical applications. Their lubricating behavior stems from their intermechanical weakness which is
intrinsic to their crystal structure. For example, MoS2 crystallizes in the hexagonal structure where a
sheet of molybdenum atoms is sandwiched between two hexagonally packed sulfur layers with a high
c/a ratio (c = 12.29 Å, a = 3.16 Å), as shown schematically in Figure 39.3a [6,7]. The bonding within the
S–Mo–S sandwich is covalent, while weak van der Waals forces hold the sandwich together resulting in
interlamellar mechanical weakness. Thus, under a shearing force the basal planes slide over one another
by intracrystalline slip and transfer to the rubbing counterface. The main mechanisms for imparting low
interfacial shear are (a) creation of the (0002) basal plane by separating the weakly bonded sandwiches
and subsequent reorientation parallel to the sliding direction and (b) the development of a transfer film
on the counterface to accommodate interfacial sliding [3,4,7–11]. The ability of MoS2 and WS2 to form
transfer films on the counterface implies that is not necessary to coat both surfaces of the sliding couple;
coating one contacting surface would suffice to generate low friction after the initial run-in period. Also,
it is not necessary to have fully crystalline films and with the preferred parallel crystallographic texture,
shown in Figure 39.4a. Several experimental studies confirmed that friction would induce crystallinity
into MoS2 and WS2 films that lack long range order [3,12], and (re)orient the (0002) planes parallel to the
sliding direction during the run-in period [3,10,12,13], schematically shown in Figure 39.4b. Tungsten
disulfide behaves in a similar fashion, but it is more expensive to synthesize compared with the naturally
occurring molybdenite. However, WS2 is thermally stable to ∼400°C, while the performance of MoS2
begins to deteriorate above ∼300°C [14,15]. Both MoS2 and WS2 coatings exhibit extremely low friction
(μ ∼ 0.02 or less) and long wear life (several million sliding cycles) when employed in either dry inert gas
or in ultrahigh vacuum [6,14,15]. However, when sliding in humid air, dangling or unsaturated bonds
on the edge of basal planes react with moisture and oxygen in the environment to form tribo-oxidation
products, such as MoO3 and WO3, resulting in higher friction (0.15–0.2) and extremely short wear life
[7,14,15]. Table 39.1 summarizes many of these characteristics for MoS2 and WS2 under the heading of
single-phase compounds.
Low Friction Coatings 39-11

39.2.1.2  Composite MoS2 and WS2


There is an increasing demand for environmentally robust low friction coatings that can adapt them-
selves to different environments [10]. For instance, even if the targeted application is friction mitiga-
tion in space, often times the satellites wait for extended periods of time in humid environments prior
to launch, potentially exposing the moving mechanical assemblies to humidity. In view of this, there
have been major studies aimed at developing multiphase materials known as adaptive lubricants and
chameleon coatings [10]. A number of metal or oxide dopants in MoS2 have been successfully tried.
Notable examples of dopants include Ti, Al, Ni, Au, Pb, PbO, and Sb2O3. The presence of these dopants
can lead to increased coating density, hardness, and oxidation resistance in humid environments com-
pared to pure MoS2. Amongst these, Sb2O3 and Au-doped MoS2 films [3,16,17], shown in Figure 39.2,
and Ti-doped MoS2 films [18] are gaining acceptance as robust coatings for commercial use, including
moving mechanical assemblies in satellites. Table 39.1 summarizes many of these characteristics for
composite MoS2 and WS2 under the heading of multiphase composites.

39.2.2  Carbon-Based Materials


39.2.2.1  Graphite
Graphite crystal structure, shown in Figure 39.3b, is a layered solid with a hexagonal lattice and high
c/a ratio (c = 6.71 Å, a = 2.46 Å) [6]. Similar to MoS2, the carbon atoms in its basal planes are held with
strong covalent bonds while the basal planes themselves are held together by weak van der Waals cohe-
sive forces, resulting in interplanar slip. The presence of water vapor and oxygen in the environment
are believed to facilitate the interlamellar shearing of graphite crystals. These close-packed basal planes
exhibit low surface energies and have little adhesion amongst them [19]. However, when high energy edge
sites of the basal plane lamellae are exposed, they bond strongly to other edge sites causing increased
adhesion. Low friction is maintained when these reactive edge sites are neutralized (passivated) by the
adsorption of water, or other condensed vapors [6]. Unlike the case of MoS2 and WS2, graphite needs
moisture, or adsorbed gases, in the environment (>100 ppm) to passivate the dangling covalent bonds
in order to lubricate [20]. In vacuum and in dry environments, graphite without additives exhibits high
friction—a phenomenon known as “dusting,” first observed in the late 1930s when graphite brushes in
aircrafts experienced accelerated wear at high altitudes. A recent report refutes the traditional claim
that the interlamellar weakness in graphite results from chemisorption or intercalation of vapor mole-
cules near the surface leading to an increase in the interlayer spacing between near-surface basal planes
[21]. Instead the authors determined that since there is no change in graphite’s interlayer spacing in
tested environments of vacuum, ambient air, or water vapor saturated air, then there must be another
mechanism operative to explain graphite’s low friction. They rationalized, as stated earlier, that vapor
molecules are required to saturate dangling bonds at the low surface energy, basal plane reactive edges
sites in order to maintain low friction. The most widespread use of graphite (besides electrical contact
brushes) in antifriction applications is in metal- and polymer-matrix self-lubricating composites as a
“built-in” solid lubricant. Sometimes, graphite is also applied as a resin-bonded coating. Table 39.1 sum-
marizes many of these characteristics of graphite under the heading of single-phase elements.

39.2.2.2  Diamond-Like Carbon


DLC coatings are typically amorphous with short-range ordered phases of mixed sp3-type tetrahedral
bonding (diamond hybridization) and sp2-type trigonal bonding (graphitic hybridization). These are
known to exhibit an unusual combination of tribological and mechanical properties: low friction coef-
ficients and low wear rates, relatively high hardness, and high elastic modulus [22]. DLC materials doped
with hydrogen (∼10–50 at%) are commonly referred to as hydrogenated DLC. Alternatively, DLC can be
doped with N2, Si, and SiOx, as well as transition metals, such as Cr, W, and Ti, which form nanoscale
hard metal carbide phases, to improve mechanical strength, hardness, and wear resistance. The friction
39-12 Wear Materials

coefficients of DLC coatings range from 0.001 to 0.5 depending upon the test conditions (i.e., contact
stress, sliding velocity, temperature) and the environment, as shown in Table 39.1. In addition, the
chemical bonding and hydrogen content of the coatings profoundly influence the friction coefficient.
For instance, hydrogen-free DLC coatings work best in humid air where low friction coefficients (∼0.1)
can be achieved for long durations, while hydrogenated coatings perform better in dry or inert gas
environments.
Synthesizing one single DLC material to achieve low friction in both dry and humid environments
can be a challenging task. Recent research shows promise for environmentally robust tribological nano-
composite coatings, such as DLN, whose structure has been conjectured to consist of two amorphous
interpenetrating networks, a diamond-like (a-C:H) network and a quartz-like (a-Si:O) network with
minimal bonding between the two networks [2]. The mutual stabilization of these networks prevents the
growth of graphitic carbon at high temperatures as well as serves to enhance the adhesion and reduce
the internal, compressive stress to approximately −0.5 GPa. The composition of DLN can vary but is
typically around (CH0.15)0.7(SiO0.3)0.3.
There are many similarities in the tribological behavior of DLN and MoS2. For instance, DLN (as
well as many other DLCs) transfer a thin layer of material from the coating to the counterface forming
a transfer film. Thus, it is not always necessary to coat both surfaces of the friction couple. Second, DLN
also exhibits non-Amontonian behavior with friction coefficient decreasing with increasing Hertzian
contact stress, according to Equation 39.2 (see Figure 39.2); the interfacial shear strength values τo, com-
puted from these plots, are 9 MPa in dry nitrogen and 78 MPa in humid air [2]. Chemical mapping by
time-of-flight secondary ion mass spectroscopy (ToF-SIMS) aided by automated expert spectral image
analysis (AXSIA) software showed that the transfer film in dry nitrogen was predominantly a mixture of
long range carbon and hydrogenated carbon while in humid air it was composed of mostly silicon oxide
species. It appears that by forming transfer films of long range carbon and hydrogenated carbon in dry
nitrogen, and predominantly silicon oxide species in humid air, DLN is able to adapt itself to both dry
and humid environments, thus becoming an environmentally robust low friction coating.
Besides the inherent tribological behavior and the environmental effects discussed earlier, coating–
substrate interface plays a critical role in governing the coating performance. For instance, DLCs are
known to have adhesion issues with substrate materials that contain non-carbide-forming elements.
In such cases, a thin Ti adhesion layer is typically applied prior to depositing DLC. Focused ion beam
(FIB) microscopy and finite element analysis modeling (FEM) are important tools to study coating–
substrate interface reliability. FIB sections of wear scars are routinely made to visualize friction-induced
subsurface deformation and to validate FEM [23,24]. For example, when contact stresses are increased
beyond a limit that the substrate begins to plastically yield, Hertzian elastic contact theory is no longer
valid. This is illustrated in Figure 39.5 for the case of DLN coating on a variety of substrates [24]. All the
data points in the shaded regions, with friction coefficients exceeding the predicted values by an order
of magnitude, correspond to the cases where the underlying substrates deformed plastically. This is
confirmed by FIB microscopy of cross sections of wear surface shown on the left side of the figure. This
underscores the need to design multilayer coating architecture (e.g., a hard coating sandwiched between
DLN and the softer substrate) to withstand higher operating stresses. Table 39.1 summarizes many of
these characteristics for DLC under the heading of single-phase elements.

39.2.2.3  Nanocrystalline Diamond


Diamond films offer many attractive properties such as high hardness, stiffness, thermal conductivity,
and resistance to wear. Diamond when applied as a coating with ultrananocrystalline structure, that is,
∼3–5 nm grains, is referred to as UNCD. UNCD exhibits low surface roughness (∼13 nm RMS) and low
friction in both dry nitrogen and humid air. Like MoS2/Sb2O3/Au and DLN coatings, the friction coef-
ficients of UNCD (self-mated configuration) decrease with increasing Hertzian contact stress according
to Equation 39.2 (see Figure 39.2). However, unlike MoS2/Sb2O3/Au and DLN coatings, the friction coef-
ficients and interfacial shear strengths (τo) are lower in humid air than in dry nitrogen, which is likely
Low Friction Coatings 39-13

0.8 FIB-cut
Model fit Protective Pt coating
DLN coating
0.6 Substrate plastic
deformation

COF
Ni substrate
0.4
2 µm
2 µm Slope = 47 MPa (S0)
0.2

0
0 0.5 1 1.5 2 2.5
Inverse Hertzian contact stress (1/GPa)

FIGURE 39.5  The friction coefficient decreases with increasing contact stress (i.e., increases linearly with inverse
contact stress) in agreement with the Hertzian elastic contact model (straight line). The shaded region corresponds
to the case where the contact stress induces plastic deformation in the underlying substrate (see the focused ion
beam cross section of the wear scar on the left) causing fracture and delamination. Wear scar image on the right
shows no substrate plasticity. (Reprinted from Jungk, J.M. et al., Acta Mater., 56, 1956, 2008. With permission.)

due to adsorbed water passivating the dangling bonds on the UNCD surfaces [25]. Furthermore, it is
desirable to coat both surfaces of the friction couple and use it in self-mated configuration to avoid wear
of the uncoated counterface [26]. Table 39.1 summarizes many of these characteristics for UNCD under
the heading of single-phase elements.

39.2.3  Polymers
Among polymeric materials, PTFE (i.e., Teflon®) is well-known for its antifriction property [6]. This fol-
lows from its smooth molecular profile and low intermolecular cohesion [27]. PTFE has no unsaturated
bonds and is nonpolar. During sliding contact, it forms a thin transfer film of itself on the counterface.
Unfortunately, the low intermolecular cohesion responsible for easy drawing of molecular chains out
of the crystalline portions of the polymer, which gives rise to low friction (∼0.1 against a steel counter-
face), results also in unacceptable amounts of wear [27,28]. Thus, it is impossible to achieve the desired
combination of low friction and wear using PTFE alone. In fact, polymers either exhibit low friction as
in PTFE, or low wear rate as in polyamide “nylon,” though rarely both (see Chapter 34 for more details).
Without sacrificing the characteristic low friction of PTFE, its wear resistance can be improved (by up to
a factor of 1000) by adding fillers to the PTFE matrix [28]. The improved wear performance of filled PTFE
may be due to the formation of a continuous and strongly adhering transfer film on the counterface.
PTFE also shows decreasing friction coefficients with increasing contact stress, similar to the low friction
coatings shown in Figure 39.2 [27]. PTFE is used as a matrix in self-lubricating composites as well as solid
lubricant fillers (fibers and powders) in composites, for example, electroplated Ni-PTFE coatings, for low
friction applications. However, attempts to deposit thin coatings of PTFE by physical vapor deposition
(PVD) techniques often result in cross-linking the polymer and loss of its antifriction characteristics.
PTFE/nano-Al2O3 and other polymer nanocomposites have also shown very promising friction and wear
behavior under certain conditions [29]. Table 39.1 summarizes many of these characteristics.

39.2.4  Soft Metals


Soft metals, such as lead, tin, indium, silver, and gold, when applied as coatings on relatively hard sub-
strates can result in low τo and hence low friction, μ ∼ 0.1. The widespread use of Pb, Sn, and In is in
39-14 Wear Materials

bearing alloys. Examples include Babbit metals based on needle-shaped intermetallics in a tin-rich
matrix, leaded bronzes containing islands of lead in a bronze matrix and Al–Sn alloys. These are applied
as thick overlays in steel backings to fabricate the bearings. Silver is sometimes used as a filler in high
temperature self-lubricating composite coatings (see Table 39.1), such as Ni–Mo–Al/Cr2O3/Ag/(Ba,Ca)F2
[30,31], or a dopant in hard coatings such as VN where high temperature tribochemical silver vanadate
phases (Ag3VO4, AgVO3) form and are responsible for low friction, μ ∼ 0.1–0.2 [10,32]. Thin layers of
soft metal can either provide shear accommodation, as explained earlier, or their low melting eutec-
tics can result in melt lubrication at asperity contacts. Typically, soft metal lubrication occurs by shear
within the coating, rather than by surface transfer films as in MoS2, WS2, and DLC coatings, and hence
the soft metals exhibit higher friction. Coating thickness is a very important parameter. Thicker coat-
ings result in increased real area of contact and thus increased μ values, while thinner coatings cause
an abrupt increase in μ since they do not prevent the substrate from interacting with the counterface.
Therefore, metal coatings often have a threshold in coating thickness, such as 1 μm for indium. Metal
coating lubrication is useful mainly for rolling rather than sliding applications and is most relevant at
high temperatures. Silver and barium films have been used successfully on lightly loaded ball bearings
in high-vacuum x-ray tubes, while silver and gold films have been tested successfully for high-vacuum
use in spacecraft applications. Also, soft metals are good conductors of heat and electricity and are stable
in vacuum or when exposed to nuclear radiation. Table 39.1 summarizes many of these characteristics
for soft metals.

39.2.5  Other Low Friction Coatings


Other examples of low friction coatings include tetrahedral amorphous carbon, nanocrystalline zinc
oxide, hexagonal boron nitride, boric acid, W–C–S, and YSZ–Au–MoS2–DLC (see Table 39.1 for more
details on these coatings).

39.3  Coating Processing Routes


Low friction coatings can be applied by PVD techniques, such as dc and rf sputtering, ion beam
deposition, pulsed laser ablation, and evaporation. Also, they can be deposited by chemical vapor
deposition (CVD) techniques, such as metal organic CVD, plasma-enhanced CVD (PECVD),
microwave plasma CVD (MPCVD), and atomic layer deposition (ALD) [33]. Also, burnishing and
impingement of coatings are common. The PVD processes are line-of-sight, making it difficult to
coat surfaces shadowed from the target, or uniformly coat sidewalls of three-dimensional or high
aspect ratio structures. PVD coating thickness may be reduced to 50% in blind holes with a depth
equal to the diameter. While the CVD techniques can coat shadowed surfaces, high temperatures
that are required for absorption and decomposition of chemical precursors on the surface can result
in softening the substrate. PECVD, which uses electron energy as the activation method to enable
deposition to occur at lower temperatures, can provide a good balance between achieving conformal
coatings without exposing the substrates beyond 150°C–200°C. The use of planetary rotating stages
in PECVD processing may circumvent most of the line-of-sight issues [34]. The low processing tem-
perature ALD [9], though a research tool at the moment, appears to have potential for coating fully
assembled rolling element bearings [11]. Table 39.1 lists common deposition methods for various low
friction coatings.

39.4  Key Applications


Low friction coatings are used in numerous engineering applications in unlubricated contacts and/or
in conjunction with liquids or grease. In the latter case, bearings, gears, and magnetic hard-disk drives
Low Friction Coatings 39-15

often use DLC, or doped DLC, coatings with liquid lubricants, for example, perfluoropolyether (PFPE),
to enhance performance. Doped MoS2, for example, MoS2–Sb2O3–Au, low friction coatings are used in
satellite components, like shafts, gears, and bushings. UNCD coatings are used in mechanical shaft seals
in rotary equipment, such as pumps and mixers. Pure and Si- or W-doped DLC coatings are also being
studied for potential use in the automotive industry (camshafts, valve tappets, and piston pins/rings)
and the biomedical industry (femoral heads) [22]. Table 39.1 lists common applications for various low
friction coatings.

Acknowledgments
This work was supported in part by the National Science Foundation (Grant No. CMMI-0700828). The
author also acknowledges the UNT CART facilities. The author would like to give special thanks to
Somuri V. Prasad from Sandia National Laboratories (SNL), Albuquerque, NM, and Irwin L. Singer
from Naval Research Laboratories (NRL), Washington, DC, whose knowledge of low friction coatings is
unsurpassed and their guidance and mentoring during the author’s time at these labs is immeasurable.
Also thanks to graduate students Hamidreza Mohseni and Benedict Mensah at UNT for nanocrystal-
line ZnO data, Ryan Evans and Gary Doll at Timken Co., and former colleagues Kathy Wahl at NRL,
Michael Dugger, Rand Garfield, Liz Huffman, Tony Ohlhausen, Bonnie McKenzie, Michael Rye, Joe
Michael, Paul Kotula, Dick Grant, and Gary Bryant all from SNL.

References
1. F.P. Bowden and D. Tabor, The Friction and Lubrication of Solids, Clarendon Press, Oxford, U.K.,
p. 112, 1986.
2. T.W. Scharf, J.A. Ohlhausen, D.R. Tallant, and S.V. Prasad, Mechanisms of friction in diamond-like
nanocomposite coatings, Journal of Applied Physics, 101, 063521-1–063521-11, 2007.
3. T.W. Scharf, P.G. Kotula, and S.V. Prasad, Friction and wear mechanisms in MoS2/Sb2O3/Au nano-
composite coatings, Acta Materialia, 58, 4100–4109, 2010.
4. I.L. Singer, S.D. Dvorak, K.J. Wahl, and T.W. Scharf, Role of third bodies in friction and wear of
protective coatings, Journal of Vacuum Science and Technology A, 21, S232–S240, 2003.
5. A. Erdemir, O.L. Eryilmaz, and G. Fenske, Synthesis of diamondlike carbon films with superlow
friction and wear properties, Journal of Vacuum Science and Technology A, 18, 1987–1992, 2000.
6. E.R. Braithwaite, Solid Lubricants and Surfaces, Clarendon Press, Oxford, U.K., p. 139, 1964.
7. S.V. Prasad and J.S. Zabinski, Super slippery solids, Nature, 387, 761–763, 1997.
8. I.L. Singer, Solid lubrication processes, in Fundamentals of Friction: Macroscopic and Microscopic
Processes, I.L. Singer and H.M. Pollock, Eds., Kluwer Academic, Dordrecht, the Netherlands, p. 237,
1992.
9. T.W. Scharf, S.V. Prasad, M.T. Dugger, P.G. Kotula, R.S. Goeke, and R.K. Grubbs, Growth, structure,
and tribological behavior of atomic layer deposited tungsten disulphide solid lubricant coatings
with applications to MEMS, Acta Materialia, 54, 4731–4743, 2006.
10. C. Muratore and A.A. Voevodin, Chameleon coatings: Adaptive surfaces to reduce friction and wear
in extreme environments, Annual Review of Materials Research, 39, 297–324, 2009.
11. T.W. Scharf, D.R. Diercks, B.P. Gorman, S.V. Prasad, and M.T. Dugger, Atomic layer deposition of
tungsten disulphide solid lubricant nanocomposite coatings on rolling element bearings, Tribology
Transactions, 52, 284–292, 2009.
12. K.J. Wahl, D.N. Dunn, and I.L. Singer, Wear behavior of Pb–Mo–S solid lubricating coatings, Wear,
230, 175–183, 1999.
13. J.J. Hu, R. Wheeler, J.S. Zabinski, P.A. Shade, A. Shiveley, and A.A. Voevodin, Transmission electron
microscopy analysis of Mo–W–S–Se film sliding contact obtained by using focused ion beam micro-
scope and in situ microtribometer, Tribology Letters, 32, 49–57, 2008.
39-16 Wear Materials

14. W.A. Brainard, The thermal stability and friction of the disulfides, diselenides, and ditellurides of
molybdenum and tungsten in vacuum, NASA TN D5141, 1969.
15. H.E. Sliney, Solid lubricant materials for high temperatures—A review, Tribology International, 15,
303–315, 1982.
16. M.R. Hilton and P.D Fleischauer, Applications of solid lubricant films in spacecraft, Surface and
Coatings Technology, 54–55, 435–441, 1992.
17. J.S. Zabinski, M.S. Donley, S.D. Walck, T.R. Schneider, and N.T. McDevitt, The effects of dopants on
the chemistry and tribology of sputter-deposited MoS2 films, Tribology Transactions, 38, 894–904,
1995.
18. D.G. Teer, New solid lubricant coatings, Wear, 251, 1068–1074, 2001.
19. R.F. Deacon and J.F. Goodman, Lubrication by lamellar solids, Proceedings of the Royal Society of
London, Series A, 243, 464–484, 1958.
20. R.H. Savage, Graphite lubrication, Journal of Applied Physics, 19, 1–10, 1948.
21. B.K. Yen, B.E. Schwickert, and M.F. Toney, Origin of low-friction behavior in graphite investigated
by surface x-ray diffraction, Applied Physics Letters, 84, 4702–4704, 2004.
22. A. Erdemir and C. Donnet, Eds., Tribology of Diamond-like Carbon Films: Fundamentals and
Applications, Springer, New York, 2008.
23. S.V. Prasad, J.R. Michael, and T.R. Christenson, EBSD studies on wear-induced subsurface regions
in LIGA nickel, Scripta Materialia, 48, 255–260, 2003.
24. J.M. Jungk, J.R. Michael, and S.V. Prasad, The role of substrate plasticity on the tribological behavior
of diamond-like nanocomposite coatings, Acta Materialia, 56, 1956–1966, 2008.
25. A.R. Konicek, D.S. Grierson, P.U.P.A. Gilbert, W.G. Sawyer, A.V. Sumant, and R.W. Carpick, Origin of
ultralow friction and wear in ultrananocrystalline diamond, Physical Review Letters, 100, 235502-5,
2008.
26. D.S. Grierson, A.V. Sumant, A.R. Konicek, M. Abrecht, J. Birrell, O. Auciello, J.A. Carlisle, T.W. Scharf,
M.T. Dugger, P.U.P.A. Gilbert, and R.W. Carpick, Tribochemistry and material transfer for the
ultrananocrystalline diamond–silicon nitride interface by X-PEEM spectromicroscopy, Journal of
Vacuum Science and Technology B, 25, 1700–1705, 2007.
27. B.J. Briscoe and D. Tabor, The effect of pressure on the frictional properties of polymers, Wear, 34,
29–38, 1975.
28. S. Bahadur and D. Tabor, The wear of filled polytetrafluoroethylene, Wear, 98, 1–13, 1984.
29. D.L. Burris, B. Boesl, G.R. Bourne, and W.G. Sawyer, Polymeric nanocomposites for tribological
applications, Macromolecular Materials and Engineering, 292, 387–402, 2007.
30. C. Dellacorte and H.E. Sliney, Tribological and mechanical comparison of sintered and hipped
PM212: High temperature self-lubricating composites, Lubrication Engineering, 48, 877–885, 1992.
31. C. DellaCorte, B.J. Edmonds, and P.A. Benoy, Thermal processing effects on the adhesive strength of
PS304 high temperature solid lubricant coatings, NASA TM-210944, 2001.
32. C. Muratore, J.E. Bultman, S.M. Aouadi, and A.A. Voevodin, In situ Raman spectroscopy for
examination of high temperature tribological processes, Wear, 270, 140–145, 2011.
33. G.L. Doll, B.A. Mensah, H. Mohseni, and T.W. Scharf, Chemical vapor deposition and atomic
layer deposition of coatings for mechanical applications, Journal of Thermal Spray Technology, 19,
510–516, 2010.
34. S.V. Prasad, T.W. Scharf, P.G. Kotula, J.R. Michael, and T.R. Christenson, Application of diamond-
like nanocomposite coatings on LIGA microsystem parts, Journal of Microelectromechanical Systems,
18, 695–704, 2009.
40
Wear Coating and
Treatments
40.1 Introduction.....................................................................................40-1
40.2 Wear Mitigation by Surface Treatments......................................40-3
Gary L. Doll
Adhesive Wear  •  Abrasive Wear  •  Fretting Wear  •  Fatigue
University of Akron
Wear  •  Erosive Wear  •  Cavitation Wear  •  Corrosive Wear
Allan Matthews 40.3 Conclusions....................................................................................40-11
University of Sheffield References...................................................................................................40-12

40.1  Introduction
When surfaces moving relative to each other operate in a well-lubricated condition, they are fully sep-
arated. In these cases, mechanical wear usually does not occur. Unfortunately, such ideal operating
conditions are rarely achieved in mechanical systems and wear usually limits the operational life of
a component or system. It has been estimated that the adverse economic impact of wear accounts for
as much as 7% of the gross national product (GNP) in many industrialized nations.1 Because the eco-
nomics of wear are of such huge proportions, a considerable amount of resources have been devoted to
developing surface treatments that can mitigate the effect of wear on the reliability and performance of
mechanical systems.2
Typically, surface treatments that address root causes provide the most effective solutions to wear
problems. Surface treatments can be categorized into those that modify either the microstructure,
the composition, or the topography of a surface, although some surface treatments may be classified
in to multiple categories. Microstructural modifications can increase the hardness of a material (e.g.,
steel alloys) and/or increase the compressive stress of the surface. Surfaces with high residual compres-
sive stresses can inhibit crack propagation, which is beneficial to mitigating surface-initiated fatigue. 3
Examples of surface treatments that improve the wear resistance of a surface through microstructural
modifications are flame hardening, hardening by induction heating, and shot peening. Topographical
modifications do not usually increase the hardness of a surface, although they can sometimes increase
the compressive stress of materials. The primary benefit of topographical modifications is that they can
reduce the roughness of a surface in terms of both long wavelength roughness and asperity geometry.
In elastohydrodynamic lubrication, a decrease in the rms roughness of contacting surfaces results in an
increase in the lambda value (Λ), that is, the ratio of the lubricant film thickness to the composite rms
roughness of the interface. An increased Λ usually reduces the risk of adhesive and fatigue wear, and
abrasive wear if particles that are entrapped between the moving surfaces are less likely to plow through
asperities or cause them to fracture. Gillespie has provided a compilation of technologies designed to
modify the topographies of surfaces.4 Compositional modifications can also provide hardness and com-
pressive stress to a surface; however, they can also offer additional protection from adhesive wear by
creating a dissimilar surface that resists the microscopic welding of asperities in the contact.5 Examples

40-1
40-2 Wear Materials

TABLE 40.1  Frequently Used Wear-Mitigating Surface Treatments


Thickness Elastic Application
Range Modulus Temperature
Surface Treatment/Type Example Substrates (μm) Hardness (HV) (GPa) (°C)
Carburizing Steel and low alloy 1000 700–850 No change 880–980
steel
Nitriding Low alloy and tool 500 500–1600 No change 450–570
steel (plasma:
300–800)
Nitrocarburizing Steel 300 400–1250 No change 580
Boriding Steel and stainless 100 1500–2100 300 800–1050
steel
Hard anodizing Aluminum alloy 50 >450 (depending — 10
on alloy)
CVD TiN Steel 5 2000–2500 256 850–900
TiC Steel 5 3500–4000 458 950–1000
CrC Steel 5 2000 370 850–1000
PVD TiN Steel 3 1500–2000 356 450
CrN Steel 3 1500–2000 255 100
WC/aC:H Steel 3 1000–1500 180 100
Thermal Molybdenum Steel, stainless steel, 250 1100–1800 30–40 50–200
spraying copper, aluminum
Aluminum oxide Steel, aluminum 100 700–1000 40 ≤200
alloys
WC/Co Steel 100 700–1000 60–220 ≤200
Plating Electroplated Steel 40 850–1050 140 40–60
hard chromium
Electroless nickel Steel 2–30 500–1000 120–190 60–100
Electroless nickel Steel 5–10 350–550 — 95
PTFE

of surface treatments that improve the wear resistance of a surface through compositional modifications
are shown in Table 40.1 and include case carburizing, nitriding, ion implantation, and coatings.
If the desirable surface treatment includes a coating, Matthews et al.6 have provided a coating selec-
tion methodology. Rickerby and Matthews7 have suggested that coatings can be classified according to
their deposition processes into the following categories:

1. Gaseous state processes, including chemical vapor deposition (CVD) and physical vapor
deposition (PVD)
2. Solution state processes, including chemical solution deposition, electrochemical deposition,
sol–gel processing, and plasma electrolysis
3. Molten or semi-molten state processes, including laser surface treatments, thermal spraying,
cladding, and welding

Since many specific technologies exist within the three surface treatment categories, the process of
selecting the optimum surface treatment to address a specific wear issue can be difficult. However, the
process can be simplified by (1) identifying the type and root cause of the wear, (2) being cognizant of
the surface properties that are beneficial to mitigating the wear type and root cause, (3) selecting surface
treatments that provide the beneficial surface properties, and (4) down-selecting the surface treatments
that can meet the application conditions.
Wear Coating and Treatments 40-3

40.2  Wear Mitigation by Surface Treatments


40.2.1  Adhesive Wear
Adhesive wear has been defined as wear by transference of material from one surface to another dur-
ing relative motion due to a process of solid phase welding.8 When the rubbing of surface asperities on
metallic surfaces causes surface oxide films to be broken, intimate contact between surfaces occurs and
adhesive junctions can be formed. Adhesive wear occurs if material is either transferred or liberated
because of these junctions. For a given set of application conditions, surfaces that have reduced asper-
ity interactions, frictional heating, and diffusion capability are less likely to experience adhesive wear.
Surface treatments that improve the topography of a surface from its as-machined state reduce asperity
interactions. For example, chemically accelerated mass finishing processes reduce the rms roughness
of a surface (Rq), thereby increasing the average separation of the surfaces in a lubricated contact (i.e.,
increased Λ).9 Furthermore, the process also decreases the asperity slope (Δq) which describes the ratio
of the average asperity height to width. Therefore, surfaces with smaller Δq values are less likely to form
asperity junctions than surfaces with higher Δq values. Surface treatments that modify the composition
of a surface can be used to reduce the diffusion rate across contacting asperities as well as the frictional
heating and shear stresses. A judicious choice of a compositionally modified surface to inhibit adhesive
wear is a dissimilar material that does not form adhesive junctions with the untreated mating surface,
that generates low frictional heating and stresses in sliding contact, and that is harder than the mating
surface such that it reduces the Rq and Δq of the mating surface during operation.
Scuffing is often described as an adhesive wear phenomenon that occurs when the lubricant film does
not adequately separate the contacting surfaces of gears or roller bearings, for example, and typically
manifests itself when the components are subject to high loads and high sliding velocities. Contacting
surfaces moving relative to each other can generate frictional heating and elevate the temperature in the
area of the contact. When a critical local temperature is surpassed, plastic flow of the surface material
can occur. Scuffing, or the plastic flow of material, is therefore directly related to the local tempera-
ture, or flash temperature occurring in the contact area through frictional heating. Technologies that
increase the Λ ratio and reduce the plasticity index are known to be beneficial to combating scuffing
wear. Topographical and compositional surface modifications have been used in concert to mitigate
scuffing wear in gears. Snidle et al.10 demonstrated that the combination of plasma nitriding, superfin-
ishing, and a tungsten containing diamond-like carbon coating (WC/aC:H)11 provided about a 2.7 times
increase in the scuffing load in disk on disk experiments. In actual gear tests, Ribaudo et al.12 showed
that whereas superfinishing the gear pairs provided about a 20% increase for the load required to cause
scuffing, the combination of superfinishing and a WC/aC:H coating increased the load for the onset
of scuffing by about 90%. In these experiments, superfinishing reduced the Ra values of the as-ground
gears from about 430 to 120 nm, correlating with increasing the Λ value by more than a factor of three,
from 0.6 to 2.2. Application of the WC/aC:H coating did not alter the Ra values of the superfinished
gears but increased the load to scuffing by reducing the frictional heating in the contact and inhibiting
the ability of interacting asperities to micro-weld. Scuffing of the gears only occurred after the WC/aC:H
coating had worn from the flanks of the teeth.
Roller bearings can also experience scuffing in a number of applications. Sliding contact (with a very
small amount of rolling) exists between the large ends of rollers and rib faces in tapered roller bear-
ings. In the event of a loss of lubrication or excessive rotational speeds and thrust loads, the lubricant
film thickness can become insufficient to separate the asperities on the roller ends and the rib faces and
scuffing can ensue.
Metal carbide containing diamond-like coatings are nanocomposite materials, comprised of metal
carbide precipitates (with dimensions typically <5 nm) and amorphous matrices. The precipitates are
dispersed within the matrices, enhancing the toughness and hence the fracture strength of the coating.13
Another important aspect of these coatings is that they do not adhesively interact with steel, which
40-4 Wear Materials

4000

3500

3000

2500
L50 life (s)

2000

1500

1000

500

0
Base TiC/aC TiC/aC:H WC/aC:H

FIGURE 40.1  Oil-out lifetimes of a tapered roller bearing part number with different types of diamond-like
carbon coatings applied to the critical surfaces of the bearing. Base refers to the tapered roller bearings without
any coatings, TiC/aC is a titanium-doped amorphous carbon coating, TiC/aC:H is a titanium-doped amorphous
hydrocarbon coating, and WC/aC:H is a tungsten-doped amorphous hydrocarbon coating. The horizontal bars
denote the L50 values and the boxes range from the upper and lower 65% confidence points. (From Doll, G.L.,
Improving the performance of rolling element bearings with nanocomposite tribological coatings, 2006-01-3555,
SAE 2006 Commercial Vehicle Engineering Congress & Exhibition, October, Rosemont, IL, 2006. With permission.)

makes them effective barriers to the adhesive wear that arises during starved lubrication conditions.
To make the coatings strongly adhere to their steel substrates, thin metallic interfaces (∼0.1 μm) such as
Cr and Ti are used. In a bench test designed to mimic an oil-out condition faced by bearings in a rotor-
craft gearbox, three nanocomposite, diamond-like carbon coatings were applied to the large roller ends
of tapered roller bearings and the time to the onset of scuffing from a termination in the flow of lubricant
was measured.14,15 Specifically, the three coatings tested were TiC/aC, TiC/aC:H, and WC/aC:H, and
the results of these tests are shown in Figure 40.1. The WC/aC:H coating on the ends of tapered rollers
provided more than a 50 times increase in the oil-out time to scuffing of the uncoated tapered roller bear-
ings. Tapered roller bearings with WC/aC:H coatings on the large ends of the rollers are currently being
used to provide extended oil-out capability to tapered roller bearings in rotorcraft and other gearboxes.

40.2.2  Abrasive Wear


The American Society for Testing and Materials (ASTM) defines abrasive wear as the loss of material
due to hard particles or hard protuberances that move along a solid surface.16 The economic cost of abra-
sive wear has been estimated to range from 1% to 4% of the GNP of an industrialized nation.17 Abrasive
wear can be classified as being due to either two- or three-body interactions. Two-body interaction
occurs when hard particles or asperities remove material from a surface, while three-body interaction
occurs when hard particles trapped between two surfaces remove material. In a two-body abrasive wear
environment, the general approach is to apply a coating to the workpiece that is harder than the abrasive
particles or counterface asperities. Two-body abrasive wear occurs in many different situations such
as with the blades and buckets of earthmoving equipment, slurry pumps or pipelines, rock drilling,
extruders, or chutes.18 Tungsten carbide coatings applied by plasma transfer arc or high velocity oxyfuel
(HVOF) powder spray are considered to have excellent abrasion-resistant properties. Surfaces with high
hardness and yield strength are more desirable in a three-body abrasive wear environment. One solu-
tion to the three-body abrasive wear problem is to create one high strength surface that is harder than
the abrasive material while leaving the other surface softer. In this strategy, the abrasive particles can
become embedded in or generate plastic flow of the softer surface and the wear rate of the system will be
determined by that of the harder surface.
Wear Coating and Treatments 40-5

Rolling element bearings like those used in gas turbine engines are comprised of inner and outer
rings having raceways, rolling elements, and a cage. The primary purpose of the cage is to separate and
uniformly space the rolling elements around the periphery of the raceways. Inner and outer ring race-
ways of angular contact bearings have shoulders that are referred to as land surfaces. For high-speed
(e.g., 1.5 million DN) operation, cages are designed to operate in close proximity to either outer or inner
ring land surfaces keeping the cage nearly coaxial with the bearing raceways and minimizing radial
vibration. Gas turbine engines utilize electroplated Ag coatings (25–50 μm thick) on cages which are
believed to function both as an adhesive wear barrier between the cage and the rolling elements and
as a solid lubricant if the flow of the lubricating oil to the bearing is interrupted.19 In some situations,
very fine alumina and silicon carbide particles that invade the oil supply system are carried to the bear-
ing and become embedded in the Ag-coated cage. The protruding particles cause three-body abrasive
wear to initiate during transient cage–land contacts. Steel from the lands soon transfers to cover the
particles, and adhesive wear ensues. The loss of material from the land surfaces that guide the cage
can lead to large amounts of frictional heating, thermal runaway, and a lockup of the bearing that cre-
ates catastrophic consequences to the turbine engine and the aircraft. Full-scale bearing contamination
endurance tests showed that bearing rings fabricated from M50NiL and Pyrowear 675 steel alloys suffer
excessively high land wear due to the abrasive debris particles embedded in the Ag-coated cage.20
Fisher et al.21 have shown that a conventional, hard PVD titanium nitride (TiN) coating on the race-
way lands of a rolling element bearing is effective in preventing the three-body abrasive wear of the
raceway lands. Rai et al.22 have proposed the use of a composite TiN–Ag coating (deposited by utilizing a
large area filtered arc deposition system) that was found to be about 50% harder than a conventional TiN
coating (18–21 GPa). The results of tests designed to simulate the abrasive wear of bearing lands in gas
turbine engines showed that bare M50NiL bearing lands experienced almost 20 times as much material
removal as M50NiL bearing lands coated with the TiN–Ag coating.23
Doll et al.24 have proposed an alternative approach that may more directly address the root cause of
the abrasive land wear in gas turbine engine bearings. In their approach, the 25–50 μm thick electro-
plated Ag coating on the cages has been replaced with a duplex surface treatment consisting of a base
layer of a 1–3 μm thick adhesive wear-resistant coating such as diamond-like carbon and a second layer
of a solid lubricant such as MoS2 (<5 μm thick). Unlike the electroplated Ag, this surface treatment is too
thin to trap debris particles, so the threat of abrasive wear to the bearing lands from debris embedded
in the cage is removed. Additionally, this surface treatment provides significantly more adhesive wear
protection to the cage-guiding land surfaces as well as between the rolling elements and the cage pockets
than the conventional Ag coating.

40.2.3  Fretting Wear


If two surfaces in loaded contact are subject to minute relative motion by vibration or some other force,
natural oxide films on the surfaces can be removed and metal-to-metal contact can result in adhesion of
surface asperities. Small amplitude (<100 μm) motion creates wear debris that may remain in the contact
area and cause three-body abrasive wear. Oxidation from environmental exposure can accelerate the
wear process. Fretting fatigue occurs if small amplitude displacements initiate cracks that propagate
into and weaken the material.
Since fretting is initiated by an adhesive wear mechanism, a compositionally modified surface with a
dissimilar material that does not form adhesive junctions with the untreated mating surface and gener-
ates low frictional heating in sliding contact can address the root cause of the wear. If fretting fatigue
is encountered, a microstructural modification process like shot peening could also be used to inhibit
crack propagation.
Fretting wear and fretting fatigue commonly occur at the blade–disk interface in the compressor sec-
tion of a jet turbine engine. This interface, also called the dovetail joint, is often fabricated from Ti6Al4V
because of its high strength-to-weight ratio and corrosion resistance. Titanium materials are notorious
40-6 Wear Materials

for their propensity to gall (severe adhesive wear), and the small amplitude vibratory motion in the
dovetail joint creates an environment with a high risk of fretting wear. Moreover, the combination of
the rotating disk and the airflow through the engine imposes centrifugal forces and oscillations on the
blades causing cyclic stresses that can create fretting fatigue.
A widely followed strategy used to mitigate fretting wear and fatigue of dovetail joints is to use
Cu–Ni–In or aluminum bronze with solid lubricant topcoats: typically, MoS2 or graphite dispersed in
an epoxy or acrylic binder.25 These coatings can be applied by a variety of methods including thermal
spray, plasma spray, CVD, or PVD. The roughness of the thermal- and plasma-sprayed coatings helps to
lock the solid lubricant topcoat into place. Solid lubricant components of these surface treatments are
effective at operating temperatures less than ∼315°C. Freimanis et al.26 have reported that after the solid
lubricant has worn away, Cu–Ni–In-coated dovetails caused significant wear damage to the uncoated
disk. This damage was caused by the transfer of titanium to the surface of a Cu–Ni–In-coated dovetail
via galling with the uncoated disk. Since the aerospace industry continuously demands higher perfor-
mance engines, the operational temperatures of newer engine components are exceeding the capabili-
ties of these surface treatments to inhibit dovetail fretting. Therefore, coatings that are more robust are
needed to improve the reliability of bladed disk assemblies as well as other areas of aircraft engines that
experience fretting wear. Hager et al.27 have proposed a fretting wear mitigation strategy based upon
the high temperature potential of nickel coatings to produce lubricious oxides and the low temperature
use of graphitic solid lubrication to prevent galling at the metallic interface over a broad temperature
range. Those authors evaluated thermally sprayed nickel graphite composite coatings with 5%–20%
graphite content. Their results showed that the embedded graphitic particles reduced the friction of the
thermal-sprayed nickel coatings during both low and high temperature fretting wear experiments, and
that wear on the mated Ti6Al4V surfaces was reduced by the formation of uniform transfer films that
were identified as graphitic at room temperature, and NiO-based at 450°C.
Under conditions where a rolling element bearing is oscillating or subject to vibrations, lubricant can
be squeezed from between the rolling elements and the raceways and the relative motion of the surfaces
is too small for the lubricant to be replenished. Fretting wear between the vibrating rolling elements and
the raceway creates the macroscopic damage during this period referred to as false brinelling.28 An image
of a roller bearing raceway that has experienced fretting or false brinelling wear on its raceway is shown
in Figure 40.2. In the 1980s, Boving et al.29 developed a CVD titanium carbide (TiC) coating for AISI 440
C balls to solve a specific false brinelling problem being encountered by ball bearings in orbiting gyro-
scopes. In addition to providing a barrier to the adhesive wear mechanism responsible for false brinelling,
laboratory testing revealed that the harder TiC-coated balls polished the raceways surfaces to Ra ∼ 30 nm,
well below the surface roughness that can be achieved by typical raceway manufacturing processes.30
Hard WC/aC:H coatings applied to the bearing rollers have also been found to be successful at form-
ing barriers to the adhesive wear mechanism responsible for false brinelling. 31 Although uncoated

FIGURE 40.2  Damage to roller bearing raceway caused by fretting or false brinelling.
Wear Coating and Treatments 40-7

raceways in contact with the harder coated rollers can experience wear through a mild abrasive wear
process during vibration, a significant prolongation of the useful bearing life is obtained. Ball bearings
used in the pitch and yaw systems of wind turbines are especially susceptible to false brinelling since
they tend to dither and not continually rotate. Although WC/aC:H coatings on steel balls would prob-
ably mitigate this life-limiting wear mode, hybrid bearings consisting of steel raceways and ceramic
balls may be the preferred approach in new bearing designs.

40.2.4  Fatigue Wear


Highly loaded cyclic contacts experience local surface and subsurface shear stress fields, plastic defor-
mation, and dislocations that can eventually lead to the nucleation of cracks at defects or inclusions in a
microstructure. Once nucleated, the cracks tend to propagate and compromise the load-bearing capability
of a material. This mechanism of crack initiation and propagation is referred to as fatigue, a wear process
that can sometimes be responsible for the removal of material, leaving behind macroscopic pits. Pitting
fatigue is wear mode experienced by bearings and gears. In gears and splines, the roots of teeth experience
alternating high tensile and compressive stresses and are therefore preferential sites for crack initiation. As
the fatigue crack propagates, the cracked tooth is deflected, thus increasing the loading on adjacent gear
teeth. Larger loads on these teeth create larger stresses at the corresponding root radii and lead to further
fatigue crack initiation.32 As a result, bending fatigue usually leads to failure of a number of adjacent gear
teeth and a catastrophic failure of the gear. Cycles of alternating tensile and compressive stress can also
be responsible for bending fatigue in rotating shafts especially when there is misalignment or unbalanced
loading.
Surface treatments that can inhibit the initiation and propagation of cracks can be very effective
at combating surface-initiated fatigue wear. Residual compressive stresses imparted to a material by
microstructural and compositional modifications can inhibit crack propagation. Topographical modi-
fications that reduce shear stresses and remove surface defects can be effective at hindering crack ini-
tiation. However, sometimes the greatest benefit at combating surface fatigue of a critical component
comes from the action performed by an engineered surface on a mating surface.
Planetary gear sets are comprised of a sun gear, a ring gear, and a set of planetary gears that engage
the teeth of the ring and sun gears. The sun gear can be exposed to heavy periodic loads each revolution,
the frequency of which is determined by the number of planetary gears in the gear set. As a result, the
sun gear is the gear set element that typically experiences fatigue wear due to its high cycle loading. Hard
coatings on the mating planetary gear teeth can polish the uncoated sun gear teeth to ultrafine finishes,
thereby increasing Λ, reducing the surface shear stresses, and increasing fatigue life.33
Historically, rolling element bearing life has been limited by the statistical probability of cracks ini-
tiating at and propagating from nonmetallic inclusions residing near the region of the largest Hertzian
contact pressures. Because of improved steel making processes, many of today’s bearing steels are usu-
ally able to withstand many millions of stress cycles. That is, the predicted L10 lives of bearings made
from high quality bearing steels are very large for most applications.32 Consequently, the useful life of
a rolling element bearing is seldom determined by inclusion-originated fatigue of the steel, but by wear
associated with poor lubrication or occasionally debris damage. In a well-designed and well-lubricated
rolling element bearing application, the maximum shear stresses occur well below the rolling-contact
surfaces. If the rolling elements also undergo a substantial amount of sliding motion on the raceways,
the maximum shear stresses move closer to the surface and surface-initiated fatigue wear can occur.
The addition of frictional shear stress increases the bulk contact stress and brings the maximum values
closer to the surface,34 magnifying the localized stresses under the asperity contacts. The combination
of thin film lubrication, roller/raceway sliding, and friction shear stresses can create a surface-initiated
fatigue mechanism known as micropitting.
Spherical roller bearings are commonly used to support main shafts in many wind energy turbine
modular designs because of their ability to accommodate misalignment between the shaft and the
40-8 Wear Materials

3
Depth-µm

120 150
100

80 100

60
40 50
µm 20 µm
0 0

FIGURE 40.3  Digital representation of a roller surface that has been superfinished. The superfinishing process
has reduced the roughness of the roller surface in terms of both long wavelength roughness (Ra,Rq) and asperity
geometry (Δq).

bearing housing. Although these bearings are designed to operate for decades, in actual operation they
experience thin film lubrication, roller/raceway sliding, and frictional shear stresses that combine to pro-
mote micropitting wear that can end useful bearing life in as little as 3–5 years. The thin film lubrication
condition is a consequence of the slow rotation speed, and the roller/raceway sliding (or Healthcote slip)
is intrinsic to spherical roller bearings.32 One solution that has been particularly successful at inhibiting
micropitting wear in main shaft spherical roller bearings is a specific combination of topographical and
compositional modifications applied to the rollers. A superfinishing process described by Hashimoto
et al.9 has been used to produce surfaces such as that shown in Figure 40.3 in which most of the machin-
ing features associated with grinding have been removed. Metal carbide-containing diamond-like
carbon coatings have been shown to work well at reducing and eliminating wear in rolling element
bearings, and the WC/aC:H coating appears to be particularly well-suited for inhibiting wear in most
roller bearing applications.35,36 The combination of superfinishing and the WC/aC:H coating provides
the following benefits to the ring raceways that address the root causes of micropitting. Superfinished
surfaces reduce the opportunities for asperity interactions and the shear stresses that interactions gen-
erate. Second, since the WC/aC:H coating is about twice as hard (∼10–13 GPa) as of the ring raceways
(∼7 GPa), the coated rollers polish the raceways during operation to surface finishes (Ra ∼ 20–30 nm)
well beyond those attainable with conventional machining processes. Therefore, although the lubri-
cant film thickness is unchanged, Λ is much greater. Finally, the WC/aC:H coating reduces the traction
coefficients between the roller and the ring raceway that are responsible for the frictional shear stresses
by almost 10%. Hence, the engineered surface on the rolling elements reduces the shear stresses from
asperity contact and simultaneously increases Λ in the contact, conditions that are desirable to the
avoidance of low-cycle micropitting in bearings. Without these asperity interactions, the shear stresses
associated with thin film–lubricated roller/raceway sliding are greatly reduced, theoretically delaying
the onset of micropitting to beyond the bearing’s predicted L10 life.
Whereas WC/aC:H coatings have been often used to mitigate life-limiting wear modes of rolling ele-
ment bearings, until recently, coatings have been rarely employed to increase the performance of bearings
beyond their designed capabilities. For example, although the WC/aC:H-coated rollers described earlier
prolong bearing life by mitigating the effects of life-limiting wear modes, the actual life of these bearings
Wear Coating and Treatments 40-9

does not statistically exceed the designed L10 life of similar bearings with uncoated rollers. Observations
indicate that although WC/aC:H coatings are capable of providing barriers to the assorted wear mecha-
nisms in the important early stages of bearing operation, the coatings tend to suffer from high rates of wear.
Recently, the mechanism responsible for the limited durability of a reactively sputtered WC/aC:H
coating has been isolated, and deposition process refinements have been made that appear to eliminate
the defect responsible for that mechanism.37 Bearings with the durable WC/aC:H coating on rollers have
an L10 life more than three times that of similar bearings with uncoated or standard WC/aC:H-coated
rollers. The statistical increase in life provided by the durable WC/aC:H coating on bearing rollers may
be one of the most significant advancements in the fatigue life extension of all-steel roller bearings.

40.2.5  Erosive Wear


Material removal from a surface because of the impingement of solid particles or small drops of liquid
or gas is called erosive wear. The severity of erosive wear depends upon the angle of the impinging par-
ticles or droplets. If the impingement angle is <45°, hardness is the best attribute of an erosion-resistant
surface. On the other hand, if the impingement angle is greater than 45°, surfaces that are tougher and
more elastic experience less wear. In addition to the impingement criteria, surface treatments usually
have to withstand the rigors of additional environmental demands. Erosion-resistant surface treatments
in gas turbine engines have to function at high temperatures so they must be resistant to oxidation. If
the carrier medium of erosive particles is a liquid, an erosion-resistant surface treatment should also
be resistant to the potential corrosion from that liquid. Other examples of applications that use surface
treatments to protect against erosive wear include valves and pipelines that pass or carry abrasive mate-
rials, gas turbine engine components, and blades from helicopters.
Erosion-resistant coatings are widely used on gas turbine blades. However, the compositions of the
coatings are usually dictated by how they withstand the effects of temperature in the locations within
the turbine engine. If the blades are operating at temperatures between 600°C and 850°C, Nicholls
et al.38 have shown that Ni–Cr–Al coatings withstand this environment well. If the blades are operating
at temperatures between 750°C and 950°C, the most suitable materials are PtAl 2–(Ni–Pt–Al) coatings
and MCrAlY coatings (M = Co or Ni) containing up to 25 wt% Cr and 6 wt% Al.39
According to Pomeroy, two categories of coatings are used on gas turbine blades: diffusion coatings
and overlay coatings.40 Although diffusion coatings are often formed by CVD, their deposition involves
interdiffusion between deposited Al or Cr and the substrate onto which they are coated. For example, an Al
deposition onto a nickel-based superalloy gives rise to an NiAl (β-nickel aluminide) coating. A particular
benefit of this approach is the cleanliness of the coatings that yields improved oxidation resistance41 and
that the internal, external, and cooling hole surfaces can all be coated. Overlay coatings are applied by ther-
mal spraying, electron beam evaporation, ion plating/sputtering, and electroplating.42 These coatings are
typically comprised of β ⨥ γ aluminide in a γ matrix with typical compositions of (Ni, Co)—15–28 wt% Cr,
4–18 wt% Al, and 0.5–0.8 wt% Y.43 The relative amounts of Ni and Co depend upon the coating ductility
requirements (greater Ni content) and corrosion resistance (greater cobalt content). The excellent oxida-
tion and corrosion resistance of the coatings is afforded by the formation of highly adherent alumina scales
that grow slowly because of the presence of yttrium and the significant Cr content. Although thermal spray
processes are still frequently used, the dense columnar microstructures of electron beam deposited overlay
coatings are less prone to oxidation than the open microstructures of thermal-sprayed coatings.

40.2.6  Cavitation Wear


Cavitation is a mechanism in which vapor bubbles in a fluid grow and collapse due to local pressure
fluctuations. Cavitation wear describes a process in which material is removed from a surface because
of the collapse of gas bubbles in a fluid. The collapse of the gas bubbles generates shock waves that
produce impact forces on solid surfaces, which causes work hardening, fatigue, and pits. In cavitation
40-10 Wear Materials

wear, microscopic cracks initiate and propagate to weaken the surface, facilitating an easier removal of
material. Like most fatigue processes, the microscopic cracks usually form at features or defects on the
surface that function as stress risers. Therefore, it follows that surface treatments beneficial to fatigue
wear are good candidates as a protection against cavitation damage. Topographical modifications that
reduce long and short wavelength surface roughness can reduce the stress risers and therefore the crack
initiation sites. High amounts of compressive residual stress produced from microstructural modifica-
tions can inhibit crack propagation. Additionally, materials that protect against cavitation wear usually
have a uniform microstructure with an absence of large mechanical differences between phases.
Examples of applications that experience cavitation wear damage are marine propellers, hydrofoils,
mechanical heart valves, pumps, cast iron cylinder liners, and hydraulic components. Coatings applied
by arc welding, thermal spraying, CVD, PVD, and other techniques can be effective means to protect
against cavitation wear.44 Several studies have examined the cavitation wear resistance of thermally
sprayed cermets and alloys applied to stainless steels. HVOF-deposited Fe–Cr–Si–B–Mn,45 cermets
with tungsten carbide particles in a cobalt matrix,46 and a 41WC/Ni/Cr/Co flame thermally sprayed
coating47 applied to stainless steels have all demonstrated increased cavitation wear resistance.
Pumps are used for movement of liquid in both industrial and nonindustrial processes. The cen-
trifugal pump is a widely used type that can be found in the chemical industry for transporting liquids,
and for pumping potable water, storm water, sanitary and industrial wastewater, boiler feed, condenser
circulation, and other applications. A centrifugal pump consists of an impeller with a number of blades
mounted on a rotating shaft that projects through a casing. Despite proper design and operation, cavi-
tation wear can occur inside pumps and damage components, especially the impeller blades. Thermal
spray coatings such as Stellite® 6 applied by HVOF are used to improve the performance of pumps sub-
jected to erosion and subsequent cavitation resulting from surface roughening.

40.2.7  Corrosive Wear


In corrosive wear, surfaces may first react with the environment, forming reaction products on surface
asperities, and then material is more easily removed due to other wear modes like abrasion, adhesion,
erosion, cavitation, and fatigue. The synergistic effects of wear and corrosion can result in total material
losses that are much greater than the additive effects of each process taken alone. Corrosion-enabled
wear causes aggressive damage in a number of industries, such as mining, mineral processing, chemical
processing, pulp and paper production, and energy production. Obviously, desirable surface treatments
are the ones that can improve the resistance of a component to the corrosive elements while at the same
time providing protection against other possible wear mechanisms.
Lives of dies used in the casting operations of aluminum alloys are limited by corrosive wear, erosive
wear, and thermal fatigue. Corrosion is caused by the chemical reaction between the casting alloy and
the mold material. Iron in the die material dissolves into the melt whereas aluminum and other elements
in the melt diffuse into the die. As a result, an intermetallic layer is formed on the die surface. Erosion is
caused by the physical impingement of the incoming liquid aluminum and thermal fatigue results from
the change in stress caused by alternate heating and cooling of the die surface during the casting pro-
cess. Under the combined effects of these failure mechanisms, the die will crack; fragments are broken
off the die, necessitating die removal and a consequent increase in process costs.
To protect die surfaces from this corrosion-accelerated wear, a number of surface treatments have
been evaluated. Ferritic nitrocarburizing has resulted in easier separation of the Al die cast product
from the die; less frequent cleaning of the die core system; and up to 50% increase in die life.48 PVD and
CVD coatings consisting of transition metal nitrides and carbides, and duplex treatments that combine
surface modification of the die with a coating have been reviewed.49 Lin et al. have outlined a thor-
ough approach to designing an optimum surface treatment for aluminum die casting material. 50 They
recognized that an optimized surface treatment for an aluminum die casting material would need to
be (a) non-wetting with liquid aluminum, (b) wear- and oxidation-resistant, (c) able to accommodate
Wear Coating and Treatments 40-11

Working layer (Al,Cr)2O3


Al–Ti–N or Cr–Al–N Al
Intermediate layer graded
TiN CrN
Adhesion layer
Cr or Ti
Surface-modified H13
Ion-nitrided/ferritic-
nitrocarburized H13

Quenched and
tempered H13
H13 tool steel
substrate

FIGURE 40.4  Design of an optimized surface treatment for H13 tool steel used in aluminum die casting. (From
Lin, J. et al., Surf. Coat. Technol., 201, 2930, 2006. With permission.)

the thermal residual stresses during the pressure die casting process, (d) adherent to the die material,
and (e) able to delay the onset of thermal fatigue cracking. Based upon their experimental results, the
authors designed an optimized surface treatment such as that shown in Figure 40.4.

40.3  Conclusions
Wear modes and some of the surface treatments designed to mitigate their root causes have been dis-
cussed and are summarized in Table 40.2. Surface treatments were categorized into those that modify

TABLE 40.2  Summary of Types of Wear, Description, Application Examples, Beneficial Surface Properties,
and Surface Treatments
Application Beneficial Surface Beneficial Surface
Wear Mode Description Examples Characteristics Example
Abrasive Loss of material by the Earthmoving Hardness Hard facing of metals
(2-body) passage of hard particles equipment,
over a surface conveyors, chutes
Abrasive Loss of material from hard Pivot pins, turbine One hard surface, one TiN coating on
(3-body) particles trapped between bearings softer cage-guiding lands of
two moving surfaces gas turbine bearing
Erosive Loss of material due to impact Turbine blades, High hardness and Cobalt-based alloy
of particles of solid or liquid pump impellers toughness coatings
Cavitation Loss of material from a solid Centrifugal High fatigue resistance Shot peened surface,
surface through the cyclic pumps, impellers high compressive
formation and collapse of stress, Stellite
bubbles in a liquid
Adhesion Loss of material from surfaces Gears, bearings Small asperity slopes, Diamond-like carbon,
that interact, form junctions, dissimilar material super finishing
then break apart
Corrosive Loss of material due to a Al die casting Inert material in the Chromium, chromium
chemical attack contact nitride
Fatigue Loss of material due to crack Rolling element High compressive Diamond-like carbon,
initiation and growth and bearings, gears stresses, small asperity shot peening, super
subsequent fracture under slopes, reduced finishing
cyclic loading traction forces
Fretting Loss of material due to small Dovetail joints, Dissimilar material, Boric acid, diamond-
amplitude, reciprocating rolling element small asperity slopes, like carbon,
sliding contact bearings transfer films superfinishing
40-12 Wear Materials

composition, microstructure, and topography of a material, although a number of the treatments may
fall into more than one category. Surface properties that are generally beneficial at inhibiting the specific
wear modes were summarized, and specific applications that utilize surface treatments to mitigate wear
were provided.

References
1. Anon, Roadmap of the European technology platform for advanced engineering materials and
technologies, EuMAT Roadmap, Version 27, June 2006, p. 63.
2. Martinella, R., Selection and application of wear resistant materials to increase service life of
components, Ceram. Int. 19 (1993), 375–389.
3. Schutz, W., Shot peening of components to improve fatigue strength, Proceedings of the Third
International Conference on Shot Peening (ICSP-3), Garmisch-Partenkirchen, Germany, October
12–17, 1987, pp. 501–506.
4. Gillespie, L., Mass Finishing Handbook, Industrial Press, New York, 2006.
5. Kreith, F. and Goswami, D.Y., eds., The CRC Handbook of Mechanical Engineering, 2nd edn., CRC
Press, Boca Raton, FL, 2005, pp. 3–130.
6. Matthews, A., Holmberg, K., and Franklin, S., A methodology for coating selection, Proceedings of
the 19th Leeds-Lyon Symposium on Tribology: Thin Films in Tribology, September 8–11, Leeds, U.K.,
1992.
7. Rickerby, D.S. and Matthews, A., eds., Advanced Surface Coatings: A Handbook of Surface Engineering,
Blackie, Glasgow, U.K., 1991.
8. Bowden, F.P. and Tabor, D., The Friction and Lubrication of Solids, Oxford University Press, London,
England, 1958.
9. Hashimoto, F., Melkote, S.N., Singh, R., and Kalil, R., Effect of finishing methods on surface char-
acteristics and performance of precision components in rolling/sliding contact, Int. J. Mach.
Machinability Mater. 6 (2009), 3–15.
10. Snidle, R.W., Evans H.P., Alanou M.P., and Holmes M.J.A., Understanding scuffing and micropitting
of gears, RTO AVT Specialists’ Meeting on The Control and Reduction of Wear in Military Platforms,
June 7–9, Williamsburg, VA, 2003, and published in RTO-MP-AVT-109.
11. Dimigen, H., Hubsch, H., and Memming, R., Tribological and electrical properties of metal-
containing hydrogenated carbon films, Appl. Phys. Lett. 50 (1987), 1056.
12. Ribaudo, C.R., Aylott, C., Hofmann, D., Darragh, C.V., Evans, R.D., Cooke, E.P., and Doll, G.L.,
Engineered surfaces for improved gear scuffing resistance, Proceedings of DETC’03 ASME 2003
Design Engineering Technical Conferences and Computers and Information in Engineering Conference,
September 2–6, Chicago, IL, 2003, pp. 923–930.
13. Evans, R.D., Nanocomposite tribological coatings for rolling element bearings, Mater. Res. Soc. Symp.
Proc. 750 (2003), 407–417.
14. Doll, G.L., Evans, R.D., and Johnson, S.P., Providing oil-out protection to rolling element bear-
ings with coatings, Society of Vacuum Coaters: 48th Annual Technical Conference Proceedings, 2005,
pp. 595–598.
15. Doll, G.L., Improving the performance of rolling element bearings with nanocomposite tribological
coatings, 2006-01-3555, SAE 2006 Commercial Vehicle Engineering Congress & Exhibition, October,
Rosemont, IL, 2006.
16. ASTM Standard G40-10b, Standard Terminology Relating to Wear and Erosion, ASTM International,
West Conshohocken, PA, 1987. DOI:10.1520/G0040-10B, www.astm.org.
17. Tylczak, J.H., Abrasive wear, ASM Handbook Vol 18: Friction, Lubrication, and Wear Technology,
ASM International, Materials Park, OH, 1992, p. 184.
18. Neale, M.J. and Gee, M., Guide to Wear Problems and Testing for Industry, William Andrews
Publishing, New York, 2001, p. 20.
Wear Coating and Treatments 40-13

19. Brown, P., Bearing retainer material for modern jet engines, ASLE Trans. 13 (1970), 225–239.
20. Johnson, M., Laritz, J., and Rhoads, M., Thin dense chrome bearing insertion program—Pyrowear
675 and cronidur wear testing, AFRL-PR-WP-TR-1998-2110, 1998, pp. 1–51.
21. Fisher, K.L., Carter, R.R., Johnson, M.G., Schell, J.D., Nahm, A.H., and Price, M.J., Rolling element
bearing having wear resistant race land regions, U.S. Patent 5,165,804, 1992.
22. Rai, A.K., Bhattacharya, R., and Trivedi, H.K., Wear resistant coatings for race land regions of bearing
materials, U.S. Patent US2008/0107917A1, May 8, 2008.
23. Rai, A.K., Trivedi, H.K., Bhattacharya, R.S., Klenke, C.J., and Forster, N.H., An approach to evaluate
the abrasive land wear of high-speed bearings, Tribol. Lett. 39 (2010), 143–149.
24. Doll, G.L., Evans, R.D., and Ribaudo, C.R., Coated rolling element bearing cages, U.S. Patent
6,994,475, February 7, 2006.
25. Betts, R.K., Copper–nickel–indium, a classic basis for coating development, 17th AeroMat Conference
and Exposition, May 15-–18, Seattle, WA, 2006.
26. Freimanis, A., Segall, A., Conway, J., Jr., and Whittney, E., Elevated temperature evaluation of fretting
and metal transfer between coated titanium components, Tribol. Trans. 43 (2000), 653–658.
27. Hager, C.H., Jr., Sanders, J., Sharma, J., and Voevodin, A.A., The use of nickel graphite composite
coatings for the mitigation of gross slip fretting wear on Ti6Al4V interfaces, Wear 267 (2009),
1470–1481.
28. Kotzalas, M.N. and Doll, G.L., Tribological advancements for reliable wind turbine performance,
Philos. Trans. R. Soc. A 368 (2010), 4829–4850.
29. Boving, H., Hinterman, H., and Stehle, G., TiC-coated cemented carbide balls in gyro application
ball bearings, Lubr. Eng. 39 (1983), 209–215.
30. Savan, A., Boving, H.J., Fluehmann, F., and Hintermann, H.E. Increased performance of bearings
using TiC-coated balls, J. De Phys. IV Colloque C7 3 (1993), 943–948.
31. Harris, T.A. and Kotzalas, M.N., Rolling Bearing Analysis, 5th Ed.: Essential Concepts of Bearing
Technology, Taylor & Francis, Boca Raton, FL, 2007.
32. Fernandes, P.J.L., Tooth bending fatigue failures in gears, Eng. Failure Anal. 3 (1996), 219–225.
33. Lev, L.C., Weiner, A.M., and Harris, S.J., Gear tooth smoothing and shaping process, U.S. Patent
6,170,156, January 9, 2001.
34. Harris, T. and Yu, W., Lundberg–Palmgren fatigue theory: Considerations of failure stress and
stressed volume, Trans. ASME, J. Tribol. 121 (1999), 85–89.
35. Doll, G.L. and Osborn, B.K., Engineering surfaces of precision steel components, Proc. Annu. Tech.
Conf. Soc. Vacuum Coaters, 44 (2001), 78–84.
36. Doll, G.L., Ribaudo, C.R., and Evans, R.D., Engineered surfaces for steel rolling element bearings and
gears, Mater. Sci. Technol., 2 (2004), 367–374.
37. Doll, G.L. and Evans, R.D., Solving wind turbine tribological issues with materials science,
Proceedings: Materials Science and Technology 2010, October 17–21, Houston, TX, 2010,
pp. 2158–2169.
38. Nicholls, J.R., Simms, N.J., Chan, W., and Evans, H.E. Smart overlay coatings—Concept and practice,
Surf. Coat. Technol. 149 (2002), 236–244.
39. Goward, G.W., Progress in coatings for gas turbine airfoils, Surf. Coat. Technol., 109 (1998), 73–79.
40. Pomeroy, M.J., Coatings for gas turbine materials and long term stability issues, Mater. Des. 26 (2005),
223–231.
41. Warnes, B.M. and Punola, D.C., Clean diffusion coatings by chemical vapor deposition, Surf. Coat.
Technol. 94–95 (1997), 1–6.
42. Stringer, J., Role of coatings in energy producing systems: An overview, Mater. Sci. Eng. 87 (1987),
1–10.
43. Sims, C.T., Stoloff, N.S., and Hagel, W.C., Superalloys II, Wiley, New York, 1987.
44. Bernecki, T., Surface science, Handbook of Thermal Spray Technology, J.R. Davis, ed., ASM
International, Materials Park, OH, 2004.
40-14 Wear Materials

45. Yuping, W., Pinghua, L., Chenglin, C., Zehua, W., Ming, C., and Junhua, H., Cavitation erosion
characteristics of a Fe–Cr–Si–B–Mn coating fabricated by high velocity oxy-fuel (HVOF) thermal
spray, Mater. Lett. 61 (2007), 1867–1872.
46. Escaler, X., Farhat, M., Avellan, F., and Egusquiza, E., Cavitation erosion test on a 2D hydrofoil using
surface-mounted obstacles, Wear 254 (2003), 441–449.
47. Sugiyama, K., Nakahama, S., Hattori, S., and Nakano, K., Slurry wear and cavitation erosion of
thermal-sprayed cermets, Wear 258 (2005), 768–775.
48. Srivastava, A., Joshi, V., Shivpuri, R., Bhattacharya, R., and Dixit, S., A multilayer coating architecture
to reduce heat checking of die surfaces, Surf. Coat. Technol. 163–164 (2003), 631.
49. Mitterer, C., Holler, F., Reitberger, D., Badisch, E., Stoiber, M., Lugmair, C., Nobauer, R., Muller, Th.,
and Kullmer, R., Industrial applications of PACVD hard coatings, Surf. Coat. Technol. 163–164
(2003), 716.
50. Lin, J., Carrera, S., Kunrath, A.O., Zhong, D., Myers, S., Mishra, B., Ried, P., and Moore, J.J., Design
methodology for optimized die coatings: The case for aluminum pressure die-casting, Surf. Coat.
Technol. 201 (2006), 2930–2941.
41
Coatings and Surface
Treatments: Interactions
with Lubricants
41.1 Introduction..................................................................................... 41-1
41.2 Lubrication Modes: When and How Does the Presence of a
Coating Matter?............................................................................... 41-2
41.3 How and Why Is the Friction Modified?..................................... 41-3
41.4 Effect of Shifting the Curve or the Borders in the Stribeck
Curve................................................................................................. 41-4
41.5 Consequences of the Complexity.................................................. 41-5
41.6 Lubrication of Low-Friction Coatings.......................................... 41-6
DLC-Type Coatings
41.7 Nitrides............................................................................................ 41-11
Staffan Jacobson 41.8 Conclusive Summary.................................................................... 41-12
Uppsala University References................................................................................................... 41-12

41.1  Introduction
The surfaces of tribological components must often endure extreme conditions. The combinations of
stresses, temperatures, and chemical loads are typically much more severe than for other components.
To achieve sufficient performance under these conditions—with respect to wear life, surface damage,
friction, and mechanical capacity—the surfaces must typically be given special treatments or coatings.
Numerous types of surface treatments and coatings are used to improve tribological surfaces [1,2]. The
development of new and improved processes and coating materials is very dynamic, and today the total
number of commercially available coatings and treatments is enormous.
Several coating types are primarily used on different types of forming and cutting tools. These coat-
ings include metal nitrides (TiN, TiAlN, etc.), carbides (TiC, TaC, etc.), oxides (Al 2O3, etc.), borides
(TiB2, etc.), and multilayer combinations of these. Also diamond-like and diamond coatings have been
successful. This category of applications usually employs various types of cutting fluids, rolling fluids,
pressing oils, and other fluids whose primary task is to allow the forming process to proceed without
severe wear or galling due to overheating or seizure, rather than keeping friction low and minimizing
adhesion.
In recent years, the number of wear-resistant coatings that also offer low-friction properties has risen
sharply. These coatings are now frequently applied on machine components, for demanding automo-
tive and other uses. Typically, they are based on carbon (diamond, diamond-like carbon (DLC), metal
carbides, etc.) or lamellar solids of the graphite and MoS2 types, including several MeS2 and MeSe2

41-1
41-2 Wear Materials

compounds (where Me denotes a metal). They are typically very thin (1–5 μm) and produced using
various physical vapor deposition (PVD) techniques or PVD/chemical vapor deposition (CVD) hybrid
techniques. These families of coatings are rapidly growing, and the commercial variants often involve
multilayered structures to facilitate simultaneous optimization of several properties, that is, high hard-
ness, high thermal stability, high wear resistance, and low friction.
The literature dealing with lubricant action at non-Fe-based surfaces has until recently been very
sparse. However, it is now growing rapidly as more researchers appreciate the importance of optimiz-
ing surface/additive performance, and more low-friction, wear-resistant-component coatings come into
industrial use. In his review of engineered tribological interfaces for improved boundary lubrication,
Erdemir [3] emphasizes the importance of considering the surface and lubricant as a system.
He comments on the recent development and increased use of various smart surface engineering
and coating technologies in combating friction and wear. Particularly, advances in physical and CVD
technologies have led to coatings that can be tailored in such a way that some of the phases can favor-
ably react with additives in oils to provide much lower friction than available on steel surfaces, while
another phase can provide low wear. The self-lubricating coatings (i.e., metallic, DLC, or lamellar
solids) can also be used on lubricated sliding surfaces to act as a backup lubricant [3].
This is an area under rapid development. Much has been learnt, and some commercially successful
examples already exist best represented by the fact that DLC coatings are vital for all modern high-
pressure diesel injection systems [4], and the area carries a potential for energy saving, resource saving,
and performance boosting combinations of matched coatings and lubricants.
This chapter focuses on the lubrication of modern low-friction coatings used in machine elements,
starting from the basic questions regarding the mechanisms of how the surface material influences the
friction and following up with examples of research into the tribochemical interaction of additives and
DLC-type coatings, that so far are the most promising low-friction coatings used in lubricated machine
components and automotive parts.

41.2 Lubrication Modes: When and How Does


the Presence of a Coating Matter?
The surface material will influence the lubrication effectiveness differently in different lubrication
regimes. Generally, it can be stated that once in a full film lubrication situation, the compositions of the
mating surfaces have no or marginal influence on the friction. However, the surfaces may have a distinct
influence on whether the full film can form or not. Analogously, the surfaces will influence whether
there is mixed or boundary lubrication.
Mechanical components may operate in any lubrication regime, and less frequently even unlubri-
cated. When run unlubricated, it is easy to understand the beneficial effect of a low-friction coating.
However, it has frequently been shown that the low-friction coatings may be very beneficial also
under boundary and mixed lubrication conditions [5–8]. Under these conditions, which are the most
critical situations for machine elements, different oil additives are used to prevent metal-to-metal
contact by boundary film formation, thus reducing friction and wear of contact surfaces. Since the
formation of boundary films often involves reactions and at least involves bonding between different
additive constituents and the lubricated surface, additive compounds are usually formulated for a
specific application and specific surface material. The vast majority of oil additives in use today were
formulated for iron-based surfaces. With the introduction of modern low-friction coatings, notably
the DLC coatings, the major concern is the compatibility of these coatings with existing oils and
additives [9].
So it is in the boundary lubrication and mixed regimes that the friction- and wear-reducing addi-
tives exhibit their functionality. Neville et al. [10] state that in this regime the load-bearing capac-
ity and lubricity of solid lubricants provide a backup function to lubricants. However, it has not
Coatings and Surface Treatments 41-3

been common to purposely design the surface and lubricant to be used in partnership to optimize
performance.

41.3  How and Why Is the Friction Modified?


It may seem strange that the material under the lubricant layer has a strong influence on the friction;
there are actually several such mechanisms. Most of these mechanisms can be understood from the
schematic illustration in Figure 41.1.
1. Even if normally separated by a boundary layer of lubricant molecules, both boundary lubricated
and mixed lubricated conditions involve at least some direct contact between the mating surfaces.
This corresponds to the most severely loaded areas where the lubricant molecules become—at
least temporarily—removed. The removal may be due to either mechanical scraping off or due
to the associated local frictional heating resulting in immediate desorption. In these local and
temporary spots, the underlying surfaces come into direct contact, and their intrinsic properties
will influence the friction. Once the stripped surfaces become exposed to the lubricant again, the
bare surface sites will soon be covered by surface active molecules again.
2. The strength of the bond (physisorption, chemisorption, or reaction) between the lubricant mole-
cule and the surface differs between different surface compositions. In this way, the surface mate-
rials will influence to what degree the lubricant molecules become scraped off.
3. Many lubricant additives form tribofilms that require reactions with the solid surfaces. One exam-
ple is the antiwear additive ZDDP, which forms a glassy phosphate film on top of an oxy/sulfide
layer via complex chemical reaction routes, which also requires tribological contact [10–12]. In
this process, iron-oxide-wear debris from steel and iron surfaces are proposed to play an impor-
tant role [13]. Since the solid surfaces, or wear particles from them, are involved in the formation
of the tribofilms, the composition of the surface is decisive for the tribological performance, even
if the surfaces do not directly interact once the films have formed. Therefore, whenever iron- or
copper-based component materials are coated by low-friction coatings, the tribofilm formation
should be expected to change or even in some cases disappear [10,14].

FIGURE 41.1  Schematic illustration of a boundary lubricated contact, where physisorbed lubricant molecules
become scraped off at the most severely loaded part of an asperity contact. Within the less loaded zone of the con-
tact (to the left), the molecules stick to the surface and provide a low-friction boundary between the solids. In front
of the sharper part, the conditions are too severe for the molecules to stick, and the solid surfaces come into direct
contact. The degree of scraping off is determined by the severity of load (as determined by the topography of the
surfaces, their hardness, and degree of wear) and the bond strength of the boundary molecules (as determined by
the type of molecules, the surface composition, and temperature). Note that the thickness to width proportion is
not very representative, but strongly exaggerated for clarity. Typically asperity contacts are micron-wide while the
monolayer of molecules is in the nanometer range.
41-4 Wear Materials

re
Failu
Wear volume

tate
-in

dy-s
Efficient running-in

ning
With low-friction coating

Stea
Lower µ once run-in

R un
Lower steady state wear rate

Coefficient of friction
Longer life to failure
Reduced energy loss
Reduced working temperature

With low-friction coating

Time, number of cycles, sliding distance, etc.

FIGURE 41.2  For contacts in the mixed or boundary lubricated regimes, low-friction coatings may influence the
tribological conditions in several ways.

Neville further emphasizes that although it is known that wear and friction performance in the
boundary lubrication regime is controlled mainly by the film-forming lubricant additives, surface
treatments and coatings also play important roles as they can, in fact, eliminate the benefits of the
additives [10]. Therefore, knowing the details of how surfaces and additives react is paramount in
understanding how to achieve optimal lubrication in the boundary (and mixed) regimes.
4. After deposition, the coating has a topography that differs from that of the underlying material.
If the coating has sharper asperity geometries, this will result in higher contact pressures, which
lead to higher risks of having the lubricant molecules scraped off.
5. In a contact between two coated surfaces, the hardness is typically higher and the wear resistance
better than those of the underlying material. Higher hardness will lead to less plastic deformation,
and higher wear resistance will lead to less wear rounding of the asperities. As in the previous
point, sharp asperity geometries lead to higher contact pressures, which lead to higher risks of
having the lubricant molecules scraped off.
6. In a coating against uncoated surface contact, the coating is typically more wear resistant than the
countersurface. After running-in, the coating will, therefore, also influence the asperity geom-
etries on the uncoated side and, hence, the risk of having the lubricant molecules scraped off.
The latter points may affect the time for running-in as well as the wear rate and the life expectancy (see
Figure 41.2). Efficient smoothening and shape adaptation during running-in may shift the conditions
from the boundary to the mixed regime, with substantial reductions of both friction level and wear rate.
In cases where only one side is coated, the coating may facilitate efficient smoothening of the uncoated
surface, thereby resulting in a rapid and efficient running-in.

41.4 Effect of Shifting the Curve or the


Borders in the Stribeck Curve
The friction modifications in lubricated contacts achieved by modifying the solid surfaces can be
explained by the corresponding alterations in the Stribeck diagram (see Figure 41.3). If the friction
coefficient of local solid areas that come into direct (non-lubricated) contact is reduced, this directly
Coatings and Surface Treatments 41-5

Boun- Mixed Full film


dary lubrication regime
A

Coefficient of friction
0.1

C
0.01

0.001
C D

Load × viscosity/speed

FIGURE 41.3  Addition of a low-friction coating to a surface may affect the coefficient of friction in several ways.
This is here visualized through the corresponding shifts of the friction curve and shifts of the boundary lines
between the different regimes of a Stribeck diagram.

lowers the friction within the boundary regime (arrow A) but also to a smaller extent within the mixed
regime (arrow B).
If the interface becomes smoother, the borders between the regimes are shifted to the left (arrows C
and D), and the step in the friction curve follows. This is because smoother surfaces can be separated by
a thinner lubricant film.

41.5  Consequences of the Complexity


As soon as coatings are used in tribological contacts, the possible variations are very significant even in
unlubricated conditions. The coatings may become worn through, they may form films, and they may
modify the countersurface, and so on. The analysis of friction and wear curves often becomes very intri-
cate. Add the action of lubricants and film-forming additives and a good analysis becomes extremely
demanding.
Arguably, the most successful lubricating oil additive ever invented—zinc dialkyldithiophosphate
(ZDDP)—for use on steel and cast-iron surfaces was first introduced in engine oils in the late 1930s.
Despite the enormous amount of research and development spent on the topic over the last 8 decades,
it can be concluded that we now know a great deal about the properties and morphology of ZDDP
antiwear films but still relatively little about the reaction pathways that lead to ZDDP film formation or
about the kinetics of ZDDP film generation and removal [11].
A lack of understanding prevents theoretically based selection of the best coatings and lubricants for
low friction and long wear life for mechanical components. When we add a coating on an established
component, the chemical conditions for film formation on the surfaces change, and perhaps the lubri-
cant should best be modified to give optimum performance of this component. However, typically, the
same lubricant must also lubricate other contacts that possibly have not been coated. So the original
lubricant will still be used, or if it is modified this must not impair the performance of the uncoated
parts.
The need to understand the complex phenomena occurring during boundary lubrication has given
surface analytical science an important role. The interactions between the lubricant additives and
the various surfaces materials are complex and lead to the formation of inhomogeneous multilayer
protective surface films. Surface analysis by advanced vacuum techniques such as x-ray photoelec-
tron spectroscopy, Auger electron spectroscopy, and secondary ion mass spectrometry has allowed
model structures for these films to be derived, through which the behavior of the contact can be
understood [15].
All these techniques require that the specimen be introduced into vacuum, which necessitates
removal of adherent lubricant. Obviously, this leads to an uncertainty regarding whether the surface
analyzed is a relevant representative of the surface when in use. Techniques such as infrared microscopy
41-6 Wear Materials

or micro-attenuated total reflectance spectroscopy (ATR) give chemical information without the need
for vacuum, but they do not offer the spatial resolution and surface sensitivity of the same level.
Scanning probe techniques such as AFM and STM can give very high-resolution information in air,
or even in lubro, together with important mechanical property data, but it is difficult to extract informa-
tion on the chemical nature of the surface from these methods alone.

41.6  Lubrication of Low-Friction Coatings


The interaction and possible tribofilm formation between lubricants and nonferrous surfaces have
attracted a lot of interest with the increasing popularity of low-friction coatings [16]. The most popular
and successful type of low-friction coatings for sliding and rolling contact is the large family of DLC
coatings [17]. Kalin et al. recently presented an overview of boundary lubricated DLC contacts, ana-
lyzing the behaviors, and suggested mechanisms from published studies [18].

41.6.1  DLC-Type Coatings


41.6.1.1  H-DLC against Cast Iron Tested with Six Simple Oils
Haque et al. [19] studied H-DLC coatings in lubricated reciprocating pin-on-plate tests using cast-iron
pins against coated plates. The tests were intended to simulate typical cam/follower contact in gaso-
line engines. The friction was recorded during the last hour of 6 h tests. Six different oils were tested,
including poly-alfa-olefin (PAO) with ZDDP, PAO with dinuclear molybdenum dialkyldithiocarbamate
(MoDTC), known under the name moly dimer, PAO with trinuclear MoDTC (moly trimer), and two
mixtures including moly dimer and ZDDP or moly trimer and ZDDP.
They found a positive effect of the antiwear additive (ZDDP) along with the friction modifiers
(moly dimer/moly trimer) and concluded that this effect is crucial to ensure good coating durability
for the studied H-DLC/cast-iron system. They found low-friction MoS2 sheets derived from the moly
dimer/moly trimer and ZnO/ZnS compounds derived from the ZDDP. These films at the interface
offered enhanced coating durability.
The MoS2 compound, originating from moly dimer and moly trimer, was observed in the tribo-
films formed on the wear scars of the cast-iron pins. The moly dimer additive gave 3.5 times lower
MoS2/Mo-oxide ratio and 50% higher friction than the moly trimer additive (μ ≈ 0.044 and 0.067,
respectively). However, for both oils the coating experienced catastrophic failure, indicating that the
presence of low-friction species and providing low friction were not sufficient to protect the surface.
For PAO with ZDDP, a zinc phosphate antiwear film was detected on the cast-iron pin, which actu-
ally gave lower wear than those of the moly dimer and trimer additivated oils, but in this oil the H-DLC
coating showed severe delamination.
On the other hand, the two oils containing both moly dimer and ZDDP or moly trimer and ZDDP
showed both low-friction MoS2 sheets and ZDDP-derived compounds like ZnO/ZnS at the interface.
This provided both low friction and improved wear protection. Unlike PAO with just ZDDP, no con-
ventional zinc phosphate antiwear compound was observed, but the presence of ZnO/ZnS provided
improved coating durability.
Thus, it is apparent that neither low-friction species alone nor zinc phosphate antiwear compound
alone was effective for protecting the coating. Rather, the presence of components derived from moly
dimer or moly trimer and from ZDDP was necessary for improved durability of the H-DLC coating.
The failure of the coating occurred in two steps, namely, gradual thinning by polishing wear followed
by delamination. The thinning of the coating was suggested to be due to the removal of a graphitic layer
resulting from the pressure-induced phase transformation of the H-DLC coating. The delamination
then occurred because of the poor load-bearing capability of the thinned coating.
Based on the absence of Fe and P in the tribofilms formed on the H-DLC coating and the success-
ful formation of ZDDP tribofilm containing Zn/ZnO/ZnS, the authors concluded that a different
Coatings and Surface Treatments 41-7

0.2
PAO
PAO+MoDTC
PAO+MoDTC+ZnDTP
0.15
Friction coefficient

0.1

0.05

0
steel/steel a-C:H/steel Ti-C:H/steel a-C/steel a-C:H/a-C:H Ti-C:H/Ti-C:H

FIGURE 41.4  Steady-state friction coefficient for DLC-coated flat/steel cylinder couples tested under boundary
lubricated conditions (reciprocating sliding, 350 N load). The steel/steel couple is shown for comparison. (Adapted
from de Barros’Bouchet, M.I. et al., Tribol. Int., 38(3), 257, 2005.)

mechanism of decomposition and wear protection of ZDDP may be active on carbon-rich surfaces.
Finally, they stated that a novel approach is needed for the optimal design of additive packages for this
type of low-friction coatings. The additives need to readily supply components required to stop the thin-
ning process of the coating by providing low friction as well as good wear protection.
De Barros’Bouchet [20] investigated the effects of the combination of MoDTC and ZnDTP additives
in their comparison of steel/steel, DLC/DLC, and steel against three types of DLC coatings in boundary
lubricated contact. When testing in pure PAO, the friction was clearly lower for all combinations involv-
ing at least one DLC than for the steel/steel contact (see Figure 41.4). When testing the additivated oils,
the steel/steel contact typically exhibited a slightly lower friction than the other combinations. After
XPS and EDS analysis, the authors concluded that MoDTC and ZDDP additives can react directly on
amorphous carbon surfaces. The composition of the tribofilm appears similar to that of the tribofilm
formed on steel surfaces in the same lubrication conditions. The friction and wear performance was
improved by coating both counterfaces compared to only one surface. MoS2 was detected in the contact
area only, while zinc phosphate was detected in the wear debris collected inside the tribofilm.
They also concluded that for amorphous carbon surfaces the principal role of antiwear ZDDP agent
seems to be to enhance the formation of MoS2 sheets.
Jia and colleagues reported [21] some additives that may improve the friction over the MoDTC. In
their tests, three synthesized benzotriazole-containing borate esters were separately added into PAO as
additives, using MoDTC as the comparison. The friction and wear behavior of Ti-DLC and Ti/Al-DLC
coating sliding against AISI 52100 steel under the lubrication of PAO containing the various additives
were evaluated in a reciprocating ball-on-disk apparatus. All three benzotriazole-containing borate
ester additives showed better tribological properties than the reference MoDTC.
Equey et al. studied the tribofilm formation from ZnDTP on DLC coatings of a-C:H type [22]. When
comparing DLC/DLC with steel/steel contacts after low wear sliding in the additivated oil, they noticed
that tribofilms formed on both pairs, but the morphology was clearly different. Furthermore, the film
formed on the DLC adhered much weaker than that formed on the steel, and it was totally removed from
the surface upon washing the sample in an ultrasonic bath with cyclohexane.

41.6.1.2  Lubrication of ta-C DLC Coatings against Steel


Kano [23] found a state of extremely low friction when sliding hardened steel pins on a hydrogen-free
DLC film (ta-C) lubricated with a PAO oil containing 1 weight% of an ester (GMO, glycerol mono-
oleate) additive.
41-8 Wear Materials

0.1200

0.1000

Friction coefficient
0.0800 a–C:H 5W–30
ta–C 5W–30
0.0600 Needle bearing 5W–30
ta–C PAO–ES1
0.0400 a–C:H PAO–ES1

0.0200

0.0000
10 200 400 600 1000
Rotational speed, rpm

FIGURE 41.5  Friction properties of the ta-C/steel pair lubricated with PAO + GMO as a function of the sliding
speed compared with those of a needle bearing. (From Kano, M. et al., Tribol. Lett., 18(2), 245, 2005. With permission.)

H H O O

e
ur
erat
mp
Te

R
O H O
H
R H H
H O O
O

Friction

Before friction After friction

FIGURE 41.6  Schematic view of friction-induced hydroxylated terminated carbon surfaces in agreement with
static SIMS data. (From Kano, M. et al., Tribol. Lett., 18(2), 245, 2005. With permission.)

This combination showed a friction coefficient of 0.006 at a sliding speed of 0.1 m/s, which may be the
first demonstration of the fact that the rolling contact friction level of roller bearings can be obtained in
sliding contact under a boundary lubrication condition (see Figure 41.5).
The good results lead to plans of applying this DLC coating technology to valve lifters lubricated with
a newly formulated ester-containing engine oil in actual mass-produced gasoline engines. A friction
reduction of more than 45% was expected at an engine speed of 2000 rpm [23].
By using SIMS analysis, the ultralow friction of this system was shown to be due to the formation of
a very thin low-shear-strength tribofilm on the ta-C sliding surface [24,25].
The tribofilm was believed to be formed by the tribochemical degradation of GMO, generating OH spe-
cies on the friction surface especially by reaction with the carbon dangling bonds. The origin of superlu-
bricity in these conditions was attributed to the low-energy Van der Waals interaction between two sliding
hydroxylated carbon surfaces (OH-terminated carbon film surface), as schematically shown in Figure 41.6.

41.6.1.3  Lubrication of H-DLC Compared to Metal-Doped (Ti, WC) DLC Coatings


Most published investigations on the reactions between additives and DLC coatings have treated
non-doped DLCs. To broaden the understanding, Kalin et al. have made careful investigations of
Coatings and Surface Treatments 41-9

reactions between metal-doped DLC coatings and oil additives, under reciprocal sliding conditions
[26]. They investigated the reactivity of H-DLC and metal-doped (Ti, WC) DLC coatings with the
extreme-pressure (EP) dialkyl dithiophosphate additive. Specifically, to make sure that the studied
chemical activities were not modified by the presence of steel, they performed their studies in DLC/
DLC contacts only.
Using surface-sensitive analytical techniques, they could confirm chemical reactions between the EP
additive and all the DLC coatings during the tribological contacts. The chemical activity was estimated
to be roughly 10 times higher for the Ti-doped DLC than for the WC-doped or non-doped DLC. The
higher film-forming activity also agreed with Ti DLC having a clearly lower wear rate. All the coatings
also showed high oxygen content in the wear-scar regions, indicating partial oxidation.
Their data thus show, in contrast to most publications, that some direct chemical activity between the
W-DLC and the dialkyl dithiophosphate EP additive is possible also without any iron catalytic effect.
However, the chemical changes of this coating were significantly smaller, which allowed for the affirmed
graphitization, which also was suggested to be one of the reasons for the WC-doped DLC showing 50%
higher wear rate than the Ti-doped DLC.
The authors finally stated that the use of the EP additive is essential to reduce the adhesive contact of
the metal-doped DLC coatings, which otherwise resulted in severe plastic deformation. In contrast to
the non-doped H-DLC, the metal-doped DLC behaved more “metal-like,” in terms of wear mechanism
and friction behavior.
Related to this, Krzan et al. described that for gear tests with tungsten carbide-doped DLC
(W-DLC)-coated gears, the lubricants have been mostly commercially available gear oils. Since these
were developed or tailored for ferrous materials, they were probably far from optimal for amorphous
W-DLC layers. In their work [27], the influence of lubricant chemistry on the friction and wear in
W-DLC-coated contacts was investigated using a reciprocating test device. They included six conven-
tional base stock lubricants plus selected test blends in combination with different surface-active single
additives. Their experimental results with the W-DLC showed that the additives were not necessary
for the friction performance, while wear resistance was clearly related to the additive chemistry. The
authors concluded that different mechanisms of friction-induced interaction between lubricants and
W-DLC-coated layers are possible.

41.6.1.4  Transfer to the Uncoated Side When Only One Side Is DLC Coated
Furthermore, addition of a low-friction coating onto only one of the countersurfaces in a boundary-
lubricated component may modify the tribological conditions by producing a solid tribofilm on the
uncoated surface by transfer of the coating material [17]. DLC coatings sliding against a metal exemplify
this. In well-behaving systems, the transfer layer can protect the metallic from further wear. Of course,
adhesion and cohesion of the transfer layer is critical and is also influenced by the lubrication and factors
such as the relative humidity.
When DLC is running against a hard and chemically inert counterface, such as sapphire, the forma-
tion of a transfer layer is not observed but low wear is still obtained. This is attributed to the lubricating
properties of the graphitic wear residues inside the interface [17].
It has also been shown that a tribofilm may form on an uncoated steel counterface not by direct trans-
fer, but by combining elements from the coating with elements from lubricant additives, thus forming
a new compound. One such interesting film was shown to significantly reduce the friction [8,9,28,29].
Here, steel sliding against tungsten (W)-doped DLC coating under boundary lubrication conditions
with oils containing S-based additives resulted in significant friction drops after some running-in
period. It was shown that this was caused by the formation of a WS2 tribofilm, which is a well-known
lamellar-type low-friction material.
From their investigations on commercially available non-doped hydrogenated DLC coatings, and
W-doped DLC coatings, Podgornik and coworkers [9] concluded that for non-doped DLC coatings, the
tested traditional oil additives gave very minor effect on the frictional behavior. Similar trends can be
41-10 Wear Materials

0.3
20°C
0.25 50°C
100°C

Coefficient of friction
0.2 150°C

0.15

0.1

0.05
200°C
0
0 5,000 10,000 15,000 20,000 25,000 30,000 35,000
Sliding cycles

FIGURE 41.7  Coefficient of friction curves for W-DLC coating lubricated with PAO oil containing 1% of EP addi-
tive as function of testing temperature. Contact pressure 1.5 GPa, sliding speed 0.02 m/s. (From Podgornik, B. and
Vizintin, J., Tribol. Mater. Surf. Interfaces, 4(4), 186, 2010. With permission.)

observed for W-DLC coatings when used in combination with antiwear and friction modifier additives.
Low additive concentrations lead to similar friction values as observed for pure oil, and high concentra-
tions lead to conditions typical for uncoated steel surfaces.
However, for contacts between uncoated steel and W-DLC lubricated with oils containing S-based
additives, improved tribological performance caused by WS2 tribofilm formation was observed.
Depending on the additive concentration and contact conditions (speed and load temperature), the level
of improvement varied significantly (see Figure 41.7).
Of course, in all cases where the tribofilm formation depends on the transfer of material from a low-
friction coating, it is necessary that the film formed is very stable and that the transfer rate can become
very low once the film is established. In other cases, the coating will soon be consumed and the effect
will be of short duration.
Most published investigations on the friction properties of DLC coatings treat unlubricated slid-
ing, typically evaluated in simplified model tests. In an effort to broaden the knowledge to more
application-like conditions, Gangopadhyay et al. investigated the lubricated friction between DLC-
coated and uncoated surfaces in cylinder against flat tests and in valvetrain tests. The fact that reduced
friction is not a given when adding a DLC coating was clearly demonstrated [30].
When tested under dry conditions, their silicon-doped DLC coating showed significant friction
reduction compared to steel sliding on steel, possibly due to the formation of a transfer film. However,
in the presence of fully formulated GF-4 engine oil, the friction was slightly higher than in the absence
of the engine oil, possibly due to the lack of a transfer film formation. Under boundary lubrication con-
ditions, the wear of the counterface steel cylinder in contact with a DLC-coated flat was higher than that
against uncoated steel flats, probably because of high hardness of DLC coating and also because of the
absence of any lubricant-derived antiwear films on the coated substrate.
In the valvetrain tests, bucket tappets with as-produced surface finish were compared with polished
buckets and polished plus DLC-coated buckets. Under the mixed lubrication conditions in the val-
vetrain (using the same GF-4 oil), both polished and DLC-coated steel bucket tappets showed lower
friction than the rougher production bucket tappets. However, no significant friction difference was
observed between the polished and the DLC-coated buckets in this application.
Under the mixed lubrication regime in motored valvetrain tests, patchy additive-derived antiwear
films were detected on the DLC coating. Again, it was shown that the film formation depends on the
surface material. These films primarily consisted of calcium phosphate, in contrast to the phosphate-
type inorganic compounds and carbonates observed on corresponding uncoated steel surfaces [30].
Coatings and Surface Treatments 41-11

100,000
Starved
10,000 Dry

Boundary lubricated
2570

Cycles to failure
990
1,000
240
100

10 5
2
1

DLCA

DLC A

DLC B

DLC C

DLC C
Uncoated
FIGURE 41.8  DLC coatings may significantly extend life in starved lubrication conditions. The failure criterion
corresponds to keeping the coefficient of friction below 0.2. The starved lubrication is compared to normal bound-
ary lubricated conditions and dry, unlubricated conditions (From Hanson, M., Coatings for machine components
in starved lubricated sliding contact, Department of Applied Materials Science, 2001.). DLC-coated steel against
steel or steel against steel in reciprocal sliding contact. Each value is a mean of three parallel tests.

41.6.1.5  Effects on Prolonging Low Friction in Starved Lubricated Situations


An interesting but little discussed aspect of the protective and sometimes life-extending effect of DLC
coatings was reported by Hanson [31]. In a reciprocating sliding rig, he found that in cases with starved
lubrication (only a film of a few micrometers covered the surfaces from the start) that the three com-
mercial DLC coating types kept a low friction level some 50–500 times longer than the uncoated steel
reference (see Figure 41.8).
It was suggested that the very limited amount of lubricant lasted longer in the DLC-coated cases due
to the lower production of wear debris. Each detached particle that is transported out from the wearing
interface brings some lubricant film along. Eventually, too much of the lubricant is removed for the sur-
faces to be covered, and the friction accordingly increases. This would correspond to a much improved
safety margin against temporary lubricant loss in machine components.

41.7  Nitrides
Metal nitride coatings are typically used not for their low-friction properties in sliding contact, but
rather for their surface protection ability in severe contact situations, such as forming and cutting oper-
ations. When used in lubricated sliding contact, they typically give friction levels on par with steel and
work well with the lubricants designed for steel [32].
Recently, the environmental issues with ZDDP have inspired the search for phosphorus-free alterna-
tives to lubricate metal nitride coatings. Blanco et al. [33] studied the use of a specific ionic liquid (IL) as
an additive to PAO 6 for CrN coatings. The presence of 1 wt% IL resulted in a friction reduction, but the
conventional ZDDP additive provided a better behavior in all cases. They concluded that even if it was
shown that the tribological behavior of the IL is slightly worse than that of ZDDP, the use of a “green
additive” may be interesting from an environmental point of view. In a similar investigation, the authors
also demonstrated that IL may be used as additives for lubricating TiN, but again with higher friction
than when using ZDDP of the same concentration [34].
In a study related to their H-DLC studies mentioned earlier [19], Haque et al. [16] studied the perfor-
mance of three lubricants on the sliding friction of CrN against cast iron. The oils were PAO with ZDDP,
PAO with ZDDP plus dinuclear MoDTC, known under the name moly dimer, and finally PAO with
ZDDP plus trinuclear MoDTC (moly trimer). They found that the PAO + moly dimer+ ZDDP provide
the lowest friction coefficient.
41-12 Wear Materials

Both moly dimer and moly trimer formed MoS2 and MoO3 compounds in the tribofilm on the coat-
ing. ZDDP derived components were also identified in the tribofilms formed. Further, a significant
amount of Fe was detected in the tribofilm of the CrN coating. The stable antiwear film formed on the
CrN coating was believed to be due to the supply of iron oxide from the high pin wear. The wear of the
coating and the counterbodies was not found to be sensitive to the selection of friction modifiers.

41.8  Conclusive Summary


Rapid progress is taking place for the development of various low-friction coatings. A lot of the testing
and research has been performed in unlubricated tests (e.g., in dry air, humid air, nitrogen, etc.), and
very nice low-friction behavior has been demonstrated by numerous types of coatings. However, most
low-friction coatings will be used in lubricated systems, and the lubricants typically also have to lubri-
cate uncoated surfaces being part of the same lubricant system.
Therefore, the recent research and development trends are very important and promising. Today, we
see an escalation of research into the action of low-friction coatings in traditionally formulated oils, in
parallel, an intensified search for additives that will better match the modern coatings without disturb-
ing the rest of the system, and finally the development of DLC coatings that tribochemically act more
like steel surfaces in tribological contact with traditional additives.

References
1. Holmberg, K. and A. Matthews, Coatings Tribology: Properties, Mechanisms, Techniques and
Applications in Surface Engineering. 2nd edn. Tribology and Interface Engineering Series. 2009:
Elsevier: Amsterdam, the Netherlands.
2. Stachowiak, G. and A.W. Batchelor, Engineering Tribology. Materials and Mechanical. 2000: Elsevier
Butterworth-Heinemann: Boston, MA, pp. 1–750.
3. Erdemir, A., Review of engineered tribological interfaces for improved boundary lubrication.
Tribology International, 2005. 38(3): 249–256.
4. Gåhlin, R., M. Larsson, and P. Hedenqvist, ME-C:H coatings in motor vehicles. Wear, 2001. 249(3–4):
302–309.
5. Ronkainen, H., S. Varjus, and K. Holmberg, Friction and wear properties in dry, water- and oil-
lubricated DLC against alumina and DLC against steel contacts. Wear, 1998. 222(2): 120–128.
6. Jacobson, S. and S. Hogmark. On the tribological character of boundary lubricated DLC coated com-
ponents, in Boundary and Mixed Lubrication: Science and Applications, 28th Leeds-Lyon Symposium
on Tribology, 2001. Elsevier: Vienna, Austria.
7. Podgornik, B. et al., Tribological behaviour of WC/C coatings operating under different lubrication
regimes. Surface and Coatings Technology, 2004. 177–178: 558–565.
8. Podgornik, B. et al., Combination of DLC coatings and EP additives for improved tribological
behaviour of boundary lubricated surfaces. Wear, 2006. 261(1): 32–40.
9. Podgornik, B. and J. Vizintin, Action of oil additives when used in DLC coated contacts. Tribology—
Materials, Surfaces and Interfaces, 2010. 4(4): 186–190.
10. Neville, A. et al., Compatibility between tribological surfaces and lubricant additives—How fric-
tion and wear reduction can be controlled by surface/lube synergies. Tribology International, 2007.
40(10–12): 1680–1695.
11. Spikes, H., The history and mechanisms of ZDDP. Tribology Letters, 2004. 17(3): 469–489.
12. Martin, J.M. et al., The two-layer structure of zndtp tribofilms Part 1: AES, XPS and XANES analyses.
Tribology International, 2001. 34(8): 523–530.
13. Martin, J.M., Antiwear mechanisms of zinc dithiophosphate: A chemical hardness approach.
Tribology Letters, 1999. 6(1): 1–8.
Coatings and Surface Treatments 41-13

14. Neville, A., T. Haque, and A. Morina, Friction and its importance in fuel economy—Probing the
nanoscale characteristics of surfaces in order to understand lubricant/surface interactions. Journal
of Physics—Condensed Matter, 2008. 20(35).
15. Smith, G.C., Surface analytical science and automotive lubrication. Journal of Physics D: Applied
Physics, 2000. 33(20): R187–R197.
16. Haque, T. et al., Non-ferrous coating/lubricant interactions in tribological contacts: Assessment of
tribofilms. Tribology International, 2007. 40(10–12 SPEC. ISS.): 1603–1612.
17. Hauert, R., An overview on the tribological behavior of diamond-like carbon in technical and
medical applications. Tribology International, 2004. 37(11–12): 991–1003.
18. Kalin, M. et al., Review of boundary lubrication mechanisms of DLC coatings used in mechanical
applications. Meccanica, 2008. 43(6): 623–637.
19. Haque, T. et al., Effect of oil additives on the durability of hydrogenated DLC coating under
boundary lubrication conditions. Wear, 2009. 266(1–2): 147–157.
20. de Barros’Bouchet, M.I. et al., Boundary lubrication mechanisms of carbon coatings by MoDTC and
ZDDP additives. Tribology International, 2005. 38(3): 257–264.
21. Jia, Z. et al., Tribological behaviors of different diamond-like carbon coatings on nitrided mild steel
lubricated with benzotriazole-containing borate esters. Tribology Letters, 2010. 41(1): 247–256.
22. Equey, S. et al., Tribofilm formation from ZnDTP on diamond-like carbon. Wear, 2008. 264(3–4):
316–321.
23. Kano, M., Super low friction of DLC applied to engine cam follower lubricated with ester-containing
oil. Tribology International, 2006. 39(12): 1682–1685.
24. Kano, M. et al., Ultralow friction of DLC in presence of glycerol mono-oleate (GNO). Tribology
Letters, 2005. 18(2): 245–251.
25. De Barros Bouchet, M.I. et al., Superlubricity of diamond/glycerol technology applied to automo-
tive gasoline engines, in Superlubricity. 2007, Elsevier Science B.V.: Amsterdam, the Netherlands,
pp. 471–492.
26. Kalin, M. et al., Metal-doped (Ti, WC) diamond-like-carbon coatings: Reactions with extreme-
pressure oil additives under tribological and static conditions. Thin Solid Films, 2010. 518(15):
4336–4344.
27. Krzan, B., F. Novotny-Farkas, and J. Vizintin, Tribological behavior of tungsten-doped DLC coating
under oil lubrication. Tribology International, 2009. 42(2): 229–235.
28. Stavlid, N., On the formation of low-friction tribofilms in Me-DLC—Steel sliding contacts. PhD
thesis. 2006, Uppsala University: Uppsala, Sweden.
29. Podgornik, B., S. Jacobson, and S. Hogmark, Influence of EP and AW additives on the tribological
behaviour of hard low friction coatings. Surface and Coatings Technology, 2003. 165(2): 168–175.
30. Gangopadhyay, A. et al., Friction, wear, and surface film formation characteristics of diamond-like
carbon thin coating in valvetrain application. Tribology Transactions, 2011. 54(1): 104–114.
31. Masters degree diploma work report, Uppsala University, Department of Engineering Sciences,
UPTEC F01031, 2001.
32. Huang, Z.P., Y. Sun, and T. Bell, Friction behaviour of TiN, CrN and (TiAl)N coatings. Wear, 1994.
173(1–2): 13–20.
33. Blanco, D. et al., Lubrication of CrN coating with ethyl-dimethyl-2-methoxyethylammonium
tris(pentafluoroethyl)trifluorophosphate ionic liquid as additive to PAO 6. Tribology Letters, 2011.
41(1): 295–302.
34. Blanco, D. et al., Use of ethyl-dimethyl-2-methoxyethylammonium tris(pentafluoroethyl) trifluoro-
phosphate as base oil additive in the lubrication of TiN PVD coating. Tribology International, 2011.
44(5): 645–650.
IV
Design for
Lubrication
and Tribology

42 Design for Lubrication and Tribology  Robert W. Bruce............................................... 42-1


Introduction  •  Wear Modes  •  Materials  •  Environment  •  Temperature  •  Geometry,
Kinematics, and Load  •  Developments  •  Answers to Frequently Asked
Questions  •  Acknowledgment  •  References
43 Fluid Film (Hydrodynamic) Lubrication  Andras Z. Szeri........................................... 43-1
Nomenclature (Dimensional, Nondimensional)  •  Introduction  •  Reynolds
Equation  •  Boundary Conditions  •  Finite Journal Bearings  •  Dynamic Properties
of Lubricant Films  •  Linearized Force Coefficients  •  Stability of a Flexible
Rotor  •  References
44 Journal Bearings  John C. Nicholas.................................................................................... 44-1
Nomenclature  •  Introduction  •  Fixed Geometry Journal Bearing Stability  •  Tilting
Pad Journal Bearings  •  Tilting Pad Pivots  •  Tilting Pad Bearing Lubricating
Flow  •  Advanced Tilting Pad Journal Bearings  •  Conclusions  •  References
45 Thrust Bearings  Scan M. DeCamillo and Bruce R. Fabijonas....................................... 45-1
Introduction  •  Thrust Bearing Types  •  Alignment and Equalization  •  Design
Criteria  •  Low-Speed Performance Equations  •  Industrial Applications  •  Direct Lube
Bearings  •  Hydroelectric Applications  •  Acknowledgments  •  References
46 Hydrodynamic Step and Wedge Bearings  Theo G. Keith, Sorin Cioc,
and L. Moraru............................................................................................................................ 46-1
Nomenclature  •  Design Variables  •  Step and Wedge Bearings  •  References
47 Compliant Foil Bearing Technology: An Overview  Hooshang Heshmat...................47-1
Nomenclature  •  Introduction  •  Analysis of Gas-Lubricated Foil
Journal Bearings  •  Advancement in Foil Bearing Design Analysis  • ​
Acknowledgment  •  References  •  Bibliography
48 Components with Nonconforming Contacts  Andrew V. Olver................................... 48-1
Scope  •  Design Process  •  Lubrication Analysis  •  References

IV-1
IV-2 Design for Lubrication and Tribology

49 Lubrication of Rolling Element Bearings  E. Ioannides and


Guillermo E. Morales-Espejel.................................................................................................. 49-1
Nomenclature  •  Introduction  •  Rolling Element Bearing Types  •  Elements of Bearing
Theory and Performance  •  Bearing Lubrication Methods  •  Calculations of Effective
Bearing Lubrication  •  Calculation of the Lubrication Effect in Bearing Life  •  Lubrication
Contamination Effects  •  Bearing Friction Torque Calculation  •  Closure  •  References
50 Gear Lubrication  Robert F. Handschuh............................................................................ 50-1
Introduction  •  Gear Failure Modes  •  Gear Types  •  Lubricant Selection and
Delivery Method  •  Liquid Lubricant Flow Rate Requirements  •  Film Thickness
Calculation  •  Flash Temperature Calculation  •  Lubrication Starvation  •  Summary
Comments on Gear Lubrication  •  References
51 Cams  Andrew V. Olver...........................................................................................................51-1
Nomenclature  •  Introduction  •  Unsteady Lubrication  •  Cams and
Tappets  •  Discussion  •  References
52 Lubrication Oil Systems  Jan Ploszaj, Hooshang Heshmat, and
George J.W. Staniewski........................................................................................................... 52-1
Noncirculation Oil System  •  Circulating Oil System  •  References
53 Surface Texturing  Izhak Etsion...........................................................................................53-1
Introduction  •  Background  •  Basic Modeling  •  Optimization  •  Microreservoirs and
Microtraps  •  Recent Studies  •  Summary  •  References
54 Sliding Bearings  Timothy Alan Parsons and Jianpeng Feng.......................................... 54-1
Introduction  •  Types of Sliding Bearings  •  Bearing Types and Example Ratings  •  Sliding
Bearing Design  •  Load and Pressure  •  Sliding Speed  •  PV Values  •  Design Parameters
and Measurement  •  Bearing Life  •  Failure Modes and Robustness  •  References
55 Magnetic Bearings  Alan B. Palazzolo, Zhiyang Wang, Jung Gu Lee,
Albert F. Kascak, and Andrew J. Provenza............................................................................ 55-1
Overview  •  Actuators  •  Sensors  •  Control  •  Power Electronics  •  Finite
Element Magnetic Field Simulation  •  Auxiliary/Backup/Catcher Bearings  • ​
Acknowledgments  •  References
56 Face Seals  Tom W. Lai.......................................................................................................... 56-1
Introduction  •  Face Seals  •  References
57 Lip Seals  Robert K. Flitney....................................................................................................57-1
Introduction  •  Basic Lip Seal Design  •  Dynamic Sealing Mechanism  •  Pumping
Aids  •  Performance Limits  •  PTFE Seals  •  Excluders  •  V-Ring Seals  •  Bearing
Seals  •  Lip Seals for Pressure  •  Seals for Fluctuating Pressure Applications  •  Plastic Lip
Seals for Pressure Applications  •  Seal Materials  •  References
58 Brake and Clutch  Roberto C. Dante, Carlo Navire, and Bruno Tron........................... 58-1
Introduction  •  Brakes as Kinetic Energy Dissipaters  •  Brake Types and Design  •  Clutch
Types and Design  •  Brake Noise and Vibrations  •  Friction Materials’ Compositions,
Processes, and Properties  •  Friction Couple: Brake Pads and Rotors  •  Future
Trends  •  References
59 Automotive Tribology  Edward P. Becker......................................................................... 59-1
Introduction  •  Engine  •  Transmission  •  References
60 Turbomachinery Tribology  William D. Marscher......................................................... 60-1
Mechanical Component Design and Analysis  •  Mechanical Testing, Diagnostics, and
Prognostics  •  Closure  •  References
61 Natural and Artificial Human Joints  Francis E. Kennedy and
Douglas W. Van Citters.............................................................................................................61-1
Natural Human Joints  •  Artificial Human Joints  •  References
Design for Lubrication and Tribology IV-3

62 Nuclear Reactor Power Station Lubrication  Ken J. Brown and Steven Lemberger......62-1
Introduction  •  Effect of Radiation on Lubricants  •  Condition Monitoring  •  Maintenance
Scheduling  •  Maintenance Rule  •  Other Considerations  •  Effect of Radiation on
Elastomers  •  Design Considerations for Equipment  •  Summary  •  References
63 Space Mechanism Lubrication  Stuart Loewenthal......................................................... 63-1
Mechanism Requirements  •  Solid versus Liquid Lubrication  •  Mechanism
Components  •  Concluding Remarks  •  References
64 Magnetic Storage  Nan Liu and David B. Bogy................................................................. 64-1
Introduction  •  Historical Background: Head-Disk Interface and Air-
Bearing Sliders  •  Generalized Reynolds Equation  •  Slider’s Static
Performance  •  Slider’s Dynamic Performance  •  Thermal Flying-Height Control
Sliders  •  Outlook  •  Conclusion  •  Acknowledgments  •  References
65 Diagnostics  Richard S. Cowan............................................................................................ 65-1
Introduction  •  Asset Management  •  Nondestructive Evaluation  •  Condition
Monitoring  •  Tribosystem Applications  •  References
66 Tribology Testing  Terry L. Merriman............................................................................... 66-1
Types of Testing  •  Testing Delineated by Assembly Complexity  •  Preparing for
Tribology Testing  •  Parameters  •  Selection of Type of Tester  •  Metallurgical
Techniques Supporting Tribology Testing  •  Tribology Test (Other than Friction and
Wear Testing)  •  Design of Experiment  •  Break-In or Wear-In Phenomena  •  Accelerated
Testing  •  Statistical Analysis of Friction and Wear Data  •  Role of Modeling in
Tribology Testing  •  References
42
Design for Lubrication
and Tribology
42.1 Introduction.....................................................................................42-1
42.2 Wear Modes......................................................................................42-2
Abrasive Wear  •  Adhesive Wear  •  Cavitation  •  Corrosive
Wear  •  Erosion  •  Fatigue, Rolling Fatigue  •  Fretting  •  Impact
42.3 Materials............................................................................................42-5
Solid Lubricant Coatings  •  Polymeric Materials  •  Carbon,
Graphite  •  Metals  •  Ceramics  •  Composites  •  Coatings,
Surface Treatments
42.4 Environment................................................................................... 42-11
42.5 Temperature................................................................................... 42-11
42.6 Geometry, Kinematics, and Load...............................................42-12
42.7 Developments.................................................................................42-12
Answers to Frequently Asked Questions...............................................42-13
Robert W. Bruce Acknowledgment....................................................................................... 42-14
GE Aviation References................................................................................................... 42-14

42.1  Introduction
Tribology is concerned with friction and wear and, therefore, the energy efficiency and durability of
mechanisms.
The friction coefficient is independent of load and apparent area of contact. This is primarily due to
the small areas of real contact that carry the load at a stress level near the yield strength of the weaker
material. As a result, the real area of contact is often about 104 times smaller than the apparent area of
contact, and doubling the load just doubles the apparent area of contact. Shearing these real contact
areas required a force close to the shear strength of the weaker material. So the friction coefficient
being the friction force divided by the normal force approximates the shear strength divided by the
yield strength of the weaker material, or, depending on Von Mises or Hencke, 0.33 or 0.5. Surface
oxides and contamination can mitigate these values. Friction is a property of a combination of two
materials; the friction between a steel skate and ice is different from that of the same steel skate with
rubber or wood.
Wear is affected by so many parameters that it is very difficult to predict. Additionally, the varia-
tion in wear test results is very large; it appears that the variation is slightly smaller for the better wear
couples. When all other forms of wear have been eliminated or mitigated, the remaining form of wear
is generally adhesive wear. Adhesive wear rates are generally lower between dissimilar materials, as for
most materials they weld best to themselves. Exceptions include some ceramic materials and coatings.
Wear resistance is also a function of the two materials in contact; the wear resistance of steel against
steel is different from that against wood.

42-1
42-2 Design for Lubrication and Tribology

There is little correlation between low friction and low wear. Tungsten carbide sliding against tungsten
carbide shows very low wear but a friction coefficient of 0.9; a soft carbon pencil against paper shows a
friction coefficient of 0.2, but high wear.
Lubrication is one of the tribologists’ most powerful tools. It can reduce friction and wear to 1/1,000–
1/10,000 times the original value. Most lubricants are based on mineral oil or consist of chemicals
derived from mineral oil. But water and water-based fluids are making significant inroads, especially
when cooling requirements are significant.
Tribological design parameters, in decreasing order of importance, generally are as follows:
• Materials, including lubricants
• Chemical environment
• Physical environment, especially temperature
• Mechanical considerations, geometry, kinematics, load, etc.
This chapter aims to address design selections and the methods to indentify how and how much existing
systems may be best improved. Many of the most recent improvements are a result of rapid development
of related technologies such as materials, physical vapor deposition (PVD) coatings, numerical tech-
niques and speeds, and surface finishing technology.
The importance of doing the initial tribological design thoroughly cannot be stressed enough, as
component choice (i.e., type of bearing), materials, and geometry are often the most significant fac-
tors in the resulting component life and maintenance cost. A review of all the available options is often
short-circuited by applying that what has been used before. But continuous development of component
design, materials, and lubricants can sometimes lead to significant improvements through the use of air
bearings, plastic and ceramic ball bearings, ferrofluidic lubricants, etc. Design principles for all com-
mon tribological components are available in the literature or from vendors.
Design that allows for wear and/or repair should reduce life cycle costs. Make the small and easily
replaced part or easily repaired part of the least wear-resistant material of the two materials in contact.

42.2  Wear Modes


For each specific application, the determination of the expected wear mode(s) is key in determining
what wear rate and friction forces to expect. This will then help determine choice of materials and
determination of maintenance interval or component life. The wear modes are extensively described in
Part I, but some summary information is provided in the following text. In many cases, determination
of the wear mode is simple, but occasionally, unexpected abrasive wear particles or corrosive effects lead
to unexpected high wear rates. Some characteristics are provided here.

42.2.1  Abrasive Wear


Characterized by a hard, rough, surface, wearing a softer surface (two-body) or hard particles between
solid surfaces (three-body).
Tests: For two-body abrasive wear, block-on-ring ASTM D2714-68, pin-on-disk ASTM G133-05, and
flat-on-flat test configurations. For three-body abrasive wear, ASTM G105-02 wet and ASTM G65-04
dry rubber wheel tests. ASTM G8197a.
Wear coefficient: Two-body: 50–500 × 10−4, three-body: 0.5–5 × 10−4 (Rabinowicz 1980).
Recognized by gouges in the wearing surface of the size of the particles/roughness asperities, in the
direction of motion of the hard surface.
May be mitigated by reducing number and size (<10 μm) of abrasive particles/asperities, nitriding
steel 50–250 μm deep may provide 2–3× wear resistance, and applying wear-resistant coatings like
Stellite 6 or high-velocity oxy fuel (HVOF)–applied tungsten carbide 125–250 μm thick may provide
4–10× wear resistance.
Design for Lubrication and Tribology 42-3

42.2.2  Adhesive Wear


Characterized by relatively low roughness <2.5 μm Ra wear scar, when suitable material pairs are used,
usually with an initial roughness of 0.5–1 μm Ra. The most severe types of adhesive wear are galling and
seizing, which lead to very high wear rates and components welding together; aluminum, titanium, and
stainless steel are notorious for galling and seizing, which can lead to much rougher scars.
Tests: Block-on-ring ASTM D2714-68, ASTM G77-05, pin-on-disk ASTM G99-05, flat-on-flat test
configurations, ASTM 196-08 compression-twist test for galling.
Wear coefficient: 0.1–100 × 10−4 (Rabinowicz 1980).
Recognized by minor surface scratches in the direction of motion.
May be mitigated by better choice of dissimilar materials (beware of galvanic effects), coating/treating
one or both surfaces. Steel or IN718 against Al-Fe-Ni bronze or HVOF tungsten carbide or advanced
nitride coatings. Nitronic 60 is a stainless steel with low propensity for galling.

42.2.3  Cavitation
Characterized by fluid environment, especially water, that is being accelerated close to solid surfaces.
Test: ASTM 134-95 for liquid jet, ASTM G32-09 for vibratory agitation.
Wear rates: Difficult to quantify, use control tests with conventional materials for comparison.
Recognized by a mottled sandblast like surface finish of the worn surface.
May be mitigated by elimination of water ingress. Use M50NiL steel, IN718, or nickel aluminum
bronze.

42.2.4  Corrosive Wear


Characterized by the presence of a medium that is corrosive to the substrate.
Test: ASTM B117-09 salt spray, ASTM 1654-08 coatings on panels.
Wear rates: 1–100 × 10−4.
Recognized by corrosion products in the wear scar.
May be mitigated by the use of chemically compatible materials, nickel-based alloys, or titanium.

42.2.5  Erosion
Characterized by solid or liquid particles impacting a solid surface.
Tests: ASTM G73-10 for liquid impingement, ASTM G76-07 solid particle gas jet, abrasive jet DIN
50332A8.57, ASTM F1864-05 dust.
Wear rates: 0.0002–0.2 g/g solid particulate. Correct choice of the eroding particulate for testing is
critical. For ceramic coatings, the erosion rate is sometimes expressed as a function:

V = c × v 2.3 × D1.3 × H − 0.18 × E0.75 × F −1.65

where
V is the wear volume
c is the erosion coefficient
v is the particle impact velocity
D is the particle size
H is the coating hardness
E is the coating elastic modulus
F is the coating fracture toughness

Recognized by a grit blast–type surface appearance and often eroding material remaining stuck in
the wear scar. Very fine grit may build up on the surface instead of eroding it.
42-4 Design for Lubrication and Tribology

May be mitigated by reducing the velocity and particle size or coating with elastomeric materials,
neoprene, polyurethane, rubber (especially at near-perpendicular impact angles), IN625, or hard mate-
rials like tool steel, Stellite 6. Most sensitive angle of impact is 30° for metal and 90° for ceramic surfaces.

42.2.6  Fatigue, Rolling Fatigue


Characterized by fine cracks developing in the direction perpendicular to motion at ∼45° angle with
the surface.
Test: No standard test specification exists; 40 mm bearings are routinely used.
Wear rates: Vary depending on conditions.
Recognized by cracks in the cross section of the wear scar.
May be mitigated by optimizing film thickness to roughness ratio, compressive stress in bearing sur-
face, and improved material cleanliness. Use M50NiL, nitride.

42.2.7  Fretting
Characterized by small amplitude <1 mm reciprocating motion or smaller than the apparent area of
contact.
Tests: ASTM G204-10, ASTM D4170-10, ASTM D7594-10.
Wear coefficient: 0.1–1 × 10−4 (Rabinowicz 1980).
Recognized by wear debris remaining between the wear surfaces, possibly reacting with the envi-
ronment; iron-based surfaces generate a red-brown powdery wear debris of iron oxide (Vingsbo and
Soderburg 1988).
May be mitigated by higher interface forces to reduce/eliminate motion, gluing instead of press fit,
eliminating the environmental/chemical effect by sealing out the environment with grease. Figure 42.1
shows wear and fatigue life as functions of sliding amplitude. Use suitable material combinations for
low adhesive wear.

42.2.8  Impact
Characterized by motion perpendicular to the wear surface and no directional orientation to the wear
scar. Initial surface roughness of 0.5–1 μm Ra is often recommended.

Mixed stick Gross slip


and slip Reci-
Fatigue life (number of cycles)

10–14 procat. 107


Stick sliding
Wear (m3/Nm)

10–15 106

10–16 105
1 3 10 30 300 1000
∆ (µm)

FIGURE 42.1  Fretting map in which amplitude is plotted against wear rate. (From Vingsbo, O. and Soderburg, S.,
On fretling wear maps, Wear, 126, 131, 1988.)
Design for Lubrication and Tribology 42-5

Test: U.S. Patents 3,312,100 and 4,375,762; also Engel and Yang (1995).
Wear rates: 2.5–250 μm/100,000 cycles.
Recognized by lack of directional features in the wear scar; may look like hammer impact on a smaller
scale. Wear volume relates to the number and energy of impacts endured.
May be mitigated by using material combinations suitable for adhesive wear. HVOF tungsten carbide
or above 400°C Tribaloy 800.
Significant wear due to one wear mode may facilitate another wear mode, as for instance increased
clearance due to adhesive wear may facilitate impact wear and corrosion, which may reveal itself by
noise.

42.3  Materials
Design is an iterative process, and the design of a tribological component may start with the determina-
tion of operating conditions leading to various exclusions and solutions as in Table 42.1.
Usually, the operating temperature (range) and chemical environments are given, but sometimes,
tribological components can be shielded from severe temperature or environment. This then deter-
mines the most suitable construction materials and surface treatments. Such selection then needs to be
refined based on possible friction and wear requirements, as well as other material properties as shown
in Table 42.2. Note that fatigue strength of materials may be 0.25–0.6× the tensile strength.
For sliding contacts, the wear mode is adhesive and/or abrasive, or fretting if the sliding distance is
small (<1 mm) or much smaller than the apparent contact area. Table 42.2 lists some examples of fric-
tion coefficients and wear coefficients (Rabonowicz 1980). The wear coefficients are determined based
on wear test results and the equation:

K = WH/PL

TABLE 42.1  Method to Determine Critical Tribological


Component Design Features
Parameter Determines
Operating temperature Suitable materials, surface treatments
Chemical environment Suitable materials, surface treatments
Friction requirement Need for lubricant, lubricant type
Expected types of wear Wear life and lubrication requirements
Relative surface speeds
Interface forces
Desired operating life

TABLE 42.2  Examples of Friction and Wear Coefficients


Materials Temperature Environment Friction Coefficient Wear Coefficient
Metal–metal Room temp. Dry 0.2–0.8 0.4–70 × 10−4
Metal–metal 400°C Dry 0.3–1.1 4.0–70 × 10−4
Metal–ceramic Room temp. Dry 0.3–1.1 0.04 × 10−4
Solid lubricant coatings Room temp. Dry 0.1–0.6 0.6 × 10−4
Metal–PTFE Room temp. Dry 0.06–0.1 0.25 × 10−4
Steel–steel Room temp. Lubricated 0.08–0.2 0.03–0.2 × 10−4
Hydrodynamic lubrication Room temp. Liquid <<0.1 0.01–0.6 × 10−4
42-6 Design for Lubrication and Tribology

TABLE 42.3  Common Wear Modes


Wear Mode Recommended Materials Selected Test Specifications
Abrasive Hardened steel, Stellite 6, tungsten carbide ASTM G65-4, G105-02, D4060-10, B611-85
Adhesive Hard steel against PTFE or bronze or ceramic ASTM G77-05, G99-05, G196-08,
D2714-94, D2981-94
Cavitation M50NiL, IN718 ASTM D7583-09, G134-95
Corrosive Nickel or Cobalt alloys ASTM G119-09
Erosion Elastomer, titanium, or nitride coating ASTM G73-10, G76-07
Fatigue 52100, M50NiL ASTM STP771
Fretting Tungsten carbide against tungsten carbide ASTM G204-10
Impact Tungsten carbide vs. self, Tribaloy 800 vs. self

In this equation, the characters stand for the following:


K is the wear coefficient
W is the wear volume (m3)
H is the hardness of softer material (N/m2), ∼3.2× yield strength
P is the load (N)
L is the total sliding distance (m)

Unfortunately, there is no generally accepted method to estimate the supposedly lesser wear of the
opposite, harder, surface from this approach, so it is used in Table 42.3 to indicate gross differences
between various systems. In practice, testing of material combinations is used to determine wear rates
of both surfaces, generally with the objectives to
1. Minimize the total wear: sum of the wear of each surface
2. Minimize the wear of the more expensive or harder to replace component
3. Provide an acceptable friction coefficient
Rolling contacts can reduce the adhesive wear rate by an order of magnitude compared to pure sliding.
But even 5% slip can increase the wear rate by 100% over pure rolling.
If friction coefficients of ≪ 0.1 are required or extremely low wear, a liquid lubricant system is required
and a mechanism that builds enough pressure in the lubricant film to separate the solid surfaces. Such
mechanisms may be hydrodynamic lubrication or elastohydrodynamic lubrication, where the motion
of surfaces generates the pressure or hydrostatic lubrication in which an external pump generates the
lubricant pressure.

Liquid Lubricated
Liquid-lubricated journal bearings usually have a length equal to half the shaft diameter and can
be readily optimized for minimum friction or maximum lubricant film thickness using Moes
and Bosma (1971) and Bosma and Moes (1970) for thrust bearings. If surface speeds are inad-
equate to separate the surfaces, a hydrostatic bearing may be required. Rowe and O’Donoghue
(1971) provide a simple approach to design of hydrostatic bearings. In some lightly loaded appli-
cations, the bearing may be given a lobed shape to maintain stability (Ten Napel and Bosma,
1980), also (Lubrecht et al. 1988), or at very low loads, journal bearings may be lubricated using
air as the fluid medium (DellaCorte et al., 2008). Each of these should result in very low friction
coefficients, <0.08, and very long wear life. Babbitt or Al20Sn linings may be applied to accom-
modate starting friction for oil-lubricated bearings. Similarly in concentrated contacts as in
rolling bearings or gears, liquid lubricants and elastohydrodynamic lubrication can provide low
friction coefficients, <0.08, and low wear (Winer and Cheng 1980).
Design for Lubrication and Tribology 42-7

Selection of the lubricant starts with the ability to generate enough film thickness to separate the
surfaces based on viscosity, velocity, pressure, and temperatures using hydrodynamic or elasto-
hydrodynamic theory. In Chapter 24, an overview of available lubricant base oils is presented.
If minimal film thickness is generated, additives like viscosity improvers and/or antiwear addi-
tives and boundary lubricants may be required for the application.
Liquid lubrication systems carry initial costs and require space and continued maintenance:
• Circulating lubricant systems are discussed in Chapter 52 of this handbook. Their ben-
efits include their corrosion prevention and cooling effects on components, thereby pos-
sibly facilitating the use of less expensive component materials and also removal of debris.
But the costs associated with procuring and maintaining liquid lubricants and of dispos-
ing of filter media and skimmed waste can be significant.
• Oil mist lubrication system design principles can be found on the Internet. The environ-
mental effects, especially inhalation of mist, are under scrutiny.
• Grease and automated grease application systems are widely applied.
The compatibility of these lubricants with component and seal materials and residual machin-
ing lubricants or preservation oils must be established.

Without Liquid Lubrication


If no liquid lubricants are required, or liquid lubricants are only marginally effective, the next
step is determination of the wear mode(s) that may be expected, as listed in Table 42.3 and
described in Section 42.2. These wear modes will give an indication of the amount of wear that
may be expected and the materials that may be selected.
For each of these wear modes, the materials and surface treatment are the most important
factors to mitigate wear, while also reducing surface speeds, sliding distance, and interface
forces can be effective. The operating temperature range determines which materials may be
used, and the substrate materials selected determine the maximum coating/surface treatment
temperature.

42.3.1  Solid Lubricant Coatings


Solid lubricant films based on solids with low shear strength like PTFE, molybdenum disulfide,
tungsten disulfide, and graphite (see Table 42.4) can provide friction coefficients of 0.05–0.6 to miti-
gate friction under dry wear conditions. Note the effect of temperature in Figure 42.2. Dry film
lubricant (DFL) coatings should only be applied between material combinations that show good
adhesive wear resistance, as adhered wear particles may plow through and thus destroy the solid
lubricant films. The DFL often uses a polymeric or ceramic binder, which adheres to the substrate
surface and through which the solid lubricant particles are distributed. Some of the binders keep
the solid lubricant particle from oxidizing till wear of the DFL surface reaches the individual solid
lubricant particle, thus replenishing the interface with solid lubricant. Applying such coating to one
of the surfaces in contact usually suffices; sometimes, applying these soft coatings to both contacting

TABLE 42.4  Friction of Solid Lubricants between Compatible Solids


Solid Lubricant Friction Coefficient Maximum Continuous Use Temperature (°C)
PTFE 0.05–0.1 250
MoS2 0.1–0.2 300
WS2 0.1–0.6 350
Graphite 0.1–0.5 500
42-8 Design for Lubrication and Tribology

0.6

Coefficient of friction
TiN MoS2
0.4
Graphite

WS2
0.2

PbO
PTFE
0.0
200 600 1000 1400
Temperature (ºF)

FIGURE 42.2  Effect of temperature on friction for soft solids. (From Peterson, M.B. and Ramalingam, S., Effect
of temperature on friction for soft solids. Coatings for tribological applications, Fundamentals of Friction and Wear
of Materials, ASM, Metals Park, OH, p. 341, 1980. With permission.)

surfaces increases the friction. DFLs may be deposited by spray, dip, and brush, and curing may be
from air-dry to several hours of heat treatment.

42.3.2  Polymeric Materials


The polymeric materials most often used in tribological application are PTFE, nylon (polyamides),
and polyimides. These may be mixed with various other polymeric materials and even metal and
ceramic powders for improved friction or wear resistance or conductivity. Fiber materials may be
used for improved strength or wear. Often, a pressure-velocity PV limit is provided, established by
multiplying apparent interface pressure with sliding velocity. The PV constraint relates to the fric-
tional energy dissipated (load × cof × velocity) and the low thermal conductivity of polymeric materi-
als, which together causes the polymer at the interface to heat up. So the use of polymeric materials is
limited by a combination of pressure, velocity, and temperature. Polyimide tends to show the better
PV, wear resistance, and temperature capability. Maximum recommended temperatures are shown
in Table 42.5; generally, maximum pressure is <100 MPa, and maximum velocity <10 m/s. Table 42.5
also shows property estimates for many non-polymeric materials as these may be found in refer-
ences: Aluminum Standards and Data 2000, ASM Handbook, ASME Wear Control Handbook, Neal’s
Tribology Handbook, vendor brochures and web sites. Properties of the commercial polymeric mate-
rials depend on the fillers that are mixed in with the resin. Fillers can be solid lubricants but also liquid
lubricants, fibers, and ceramics. Polymeric materials often fail from environmental degradation due to
the loss of plasticizers or mold release agents and the negative effect of chemical contaminants in the
environment. The net effect may be a rapid catastrophic failure after years of adequate performance.
The preferred counter surface of polymeric materials is a smooth (<0.5 μm Ra), hard (>50 HRC) steel.
Nickel and some aluminum and titanium counter surfaces may also work.
Friction coefficients are generally 0.3 and up, except for Teflon/PTFE which maintains friction coef-
ficients below 0.1 (Neale, 1996). Liquid lubricants such as oil and water may help reduce friction and
wear. The low elastic modulus may facilitate elastohydrodynamic lubrication.
When designing with polymeric materials, the high thermal expansion, the low strength, low elastic
modulus, and low conductivity of the polymeric material may require extra attention. Gluing, especially
when the polymeric material is positioned in a recess in a metal surface, can be an effective fastening
method. Some polymeric materials can be applied as coatings. When sliding a soft polymeric material
against a hard metal surface, it is usually advantageous to ensure that the edges of the hard surface
extend well beyond those of the soft material.
Design for Lubrication and Tribology 42-9

TABLE 42.5  Approximate Material Property Ranges


Expansion Yield Ultimate Elastic Maximum Use
Coefficient Strength Strength Modulus Density Temperature
Material (μm/m/C) (MPa) (MPa) (GPa) (g/cc) Hardness (°C)
Plastic
Polyamide 80–110 45 75–76 2–4 1.1 85 Shore D 120
Polyethylene 200 26–33 27 0.2–2.0 0.90–0.98 41–69 Shore D 100–150
Polyimide 45–56 69–90 45–250 2.5–3.2 1.33–1.43 87–90 Shore D 150–240
Polypropylene 91 12–43 20–80 1.5–2 0.9 73–78 Shore D 90–120
PVC 66 1.9–57 1.4–55 0.002–4.8 0.55–1.85 30–65 Shore D 50–90
Rubber 53–70 7–74 0.003–9.6 1.00–1.60 43–76 90–240
Rockwell E
Teflon, PTFE 130–150 21–35 0.5 2.16 56–65 Shore D 260
Carbon 2–3 5000–6000 0.2–0.26 2.3 1–2 Mohs 450
Graphite // 6.5–8.2 0.26 20–200 8–15 1.3–2.2 0.05–0.5 GPa 450
Graphite _|_ 0.5–1.2 4–5.1 0.7–1.5 450
Metal
Aluminum 16–25 30–450 70–485 69–73 2.6–2.8 30–140 HB 100–370
Bronze 15–18 82–690 200–830 96–120 7.5–8.2 65–190 HB 200–360
P-bronze 17.8 100–840 295–880 111–112 8.86 78 HRB 125–200
Cast iron 10–13 120–290 69–480 58–170 7.0–7.3 110–500 HB 120–190
Cobalt 10.5–15 500–830 618–1345 209–235 8.3–9.1 42–49 HRC 760–1050
Copper 15–18 55–330 230–380 117–130 8.4–8.9 30–51 HRB 100–500
Nickel 12–17 100–725 380–790 170–213 8.3–9.2 30–40 HRC 760–1050
Steel 11–17.3 250–700 400–800 155–165 7.4–8.0 100–600 HRB 35–550
Steel, stainless 10–19 275–760 500–1280 163–190 7.8–8.0 <100 HRB 400–600
Steel, tool 9.5–13.5 170–960 500–2250 190–210 7.8–8.9 20–70 HRC 200–600
Steel, 440°C 10.1–11.7 200–1900 760–1970 200 7.65 22–57 HRC 100–600
Titanium 9–13 470–1170 380–1100 105–120 4.5 24–28 HRC 360–600
Ceramic Fracture Hardness Toughness
strength (GPa) (MPa Vm)
(MPa)
Al2O3 6–8 300–550 220–435 3.6–4.0 9–15 3–4
SiC 3.1–5.6 350–500 304–480 3.2 22 4–5
Si3N4 2.6–3.6 250–1000 120–483 3.2 13.9 5
WC-Co 4–7 2100–3500 450–700 15.7–16 15.5 20
ZrO2 6–11 750–1150 100–220 6.0 13.2 7–8

42.3.3  Carbon, Graphite


Graphite bushings and metalized carbon bushings can be used at temperatures above grease suitability
but below 540°C (1000°F), when the oxidation rate of the carbon becomes excessive. The expansion coef-
ficient is 6.5 × 10−6/°C parallel to −0.5 × 10−6/°C perpendicular to the platelet direction. Fastening is not
without problems (Paxton 1979).

42.3.4  Metals
Metal against metal is the most common tribological contact material combination, although metal
against polymer and metal against ceramic can reduce wear rates by an order of magnitude, albeit at
lower stress levels. Cost and manufacturing ease are main reasons for the metal-to-metal preference.
42-10 Design for Lubrication and Tribology

Too often, metal components in contact are made from the same alloy. This may work if a good lubricant
is present, but in most cases, contacts of dissimilar metals are advisable to reduce adhesive wear and
galling tendencies (Rabinowicz 1965). No metal welds better than to itself, and adhesive wear is based on
welding of asperities. But beware of galvanic effects, in contacts of dissimilar materials.
Increasing surface hardness may improve wear resistance, but it can also make the surface more
brittle and reduce wear resistance. Nitriding and carburizing can increase corrosion of stainless steel, as
these processes form chrome nitrides and carbides, respectively, leaving less chrome available to form
the protective chrome oxide surface film. Applying a coating to one of the components can reduce the
friction or wear rate significantly.
For elevated temperature applications, some cobalt-, iron-, and nickel-based alloys are available for
use at 1000°C (1800°F).

42.3.5  Ceramics
Design with ceramics has to allow for the brittleness and difficulty of fastening the ceramic surface to
metal components also because of thermal expansion coefficient mismatch. Another issue can be the
variability of wear properties from one manufacturer to another; wear rates of seemingly similar mate-
rials can readily vary by a factor of 4×. Ceramic materials generally show high friction, but low wear,
especially in contact with metals or other ceramics. Routinely used tribological ceramics are Al 2O3,
Si3N4, SiC, and WC-Co (Jahanmir and Fischer 1994). These are generally designed to greater thick-
ness than metal components to mitigate brittleness and reduce costs. Hardness is even less important
than with metals, as ceramic toughness correlates better with wear performance. When designing with
ceramic materials, it is important to avoid any tensile stress concentrations in the ceramic surface and
apply generous fillets and radii. See Chapter 36 for details on ceramic materials.

42.3.6  Composites
Cermets, or cemented ceramic metal composites such as tungsten carbide cemented with 10% cobalt
(WC-10Co), developed by Krupp in the 1920s, provide very low wear when rubbing against themselves
or hard steel. Such wear rates may be 10× to 1000× better than metal to metal. As the cobalt cement can
be washed out by water-based cleaners, nickel binder may be used.
Glass-fiber- and carbon-fiber-reinforced plastic composites are routinely used to enhance load car-
rying capacity of the polymeric material. Ceramic and glass fillers can further enhance the wear per-
formance, but care must be taken as some of these fillers can scratch and wear a metal counter surface.
One popular composite material is the carbon–carbon composite, with a mesh of carbon fiber that
is infiltrated over months with carbon matrix material to form a carbon-carbon brake material. This
material is very lightweight, very temperature capable, but expensive. Designing with these materials is
challenged by the thermal mismatch with metals, two elastic moduli in different directions, and fasten-
ing issues. See Chapter 37 for details on composite materials.

42.3.7  Coatings, Surface Treatments


Coating technology is rapidly adding greater wear resistance and reduced friction coatings to the aid
of designers. Coating adhesion to metal substrates is seldom a cause for failure anymore; temperature
capabilities are increasing, while often costs are being reduced. When choosing a coating type, impor-
tant considerations include wear resistance, coating thickness, and misalignment of parts, as well as
corrosion resistance.
Arc welded and sprayed wear coatings, flame, plasma, and HVOF, are readily available from many
sources and applied to thicknesses of 75–250 μm. Most of these coatings have to be machined to meet
dimensional and roughness requirements. PVD coatings are similarly available from many sources and
Design for Lubrication and Tribology 42-11

generally up to 5 μm thick, although some can be 50 μm thick. PVD wear coatings can often be applied
effectively at 5 μm thickness without having to change drawing dimensions or tolerances.
Chemical vapor deposition (CVD) coating development has resulted in much reduced processing
temperatures (400°C), thereby facilitating a wider range of substrate materials.
Low-friction DFL coatings can be applied by PVD but also by spray and dry/cure and are generally
from 10 to 30 μm thick.
Efforts to mix low-wear with low-friction materials have resulted in a lack of both. This may not be
surprising when considering that most wear-resistant materials are hard and dense (>98% dense), while
low-friction materials are soft.
Coatings can affect the fatigue strength of substrate material; not all do. Test to verify.
Carburizing, nitriding, and boriding of metal surfaces can produce increased wear resistance to
depths of 250 μm. Many heat-treatment facilities offer these surface enhancements. Flame and induc-
tion hardening may reach a depth of 5 mm.
Babbitt coatings continue to be applied to hydrodynamic bearings to mitigate start-up friction and
wear. Lead-free alloys are now available.
With coatings as well as when rubbing a hard against a soft material, it is desirable to have the
harder, more wear-resistant materials extend beyond the edges of the softer surface. This avoids creat-
ing a deep wear track, the edges of which can create instability in the friction force and the motion of
the contact. See Chapter 38 through 41 for details on coatings and surface treatments.

42.4  Environment
The environment can have a significant effect on both friction and wear, some examples are listed here.
• Salty air causes increased corrosive wear and can cause some stainless steel alloys to rust at
elevated temperatures.
• Fine sand from dry desert areas can erode compressor airfoils of jet engines and the canopies of
helicopters, especially in dust storms. Formula 1 engines are experiencing wear due to sand inges-
tion at desert circuits. However, if the sand gets fine enough, <<3 μm, its effect between sliding
surfaces diminishes.
• Use seals to keep contaminants from bearings and seals. Sometimes grease can be effective.
• Reduced sulfur levels in gasoline are causing increased wear of fuel injector parts, which can be
mitigated by carbon or other coatings.
• Alcohol additions to gasoline can affect polymeric seal materials and be the cause of leakage.
• Wear test results from a tester with a gas-fired heater can change significantly when the combus-
tion environment is changed from oxidizing to reducing.
• Plastics can show a sudden and dramatic increase in wear rate after years of service due to envi-
ronmental fatigue, i.e., the effects of environmental gases and vapors over long periods, and the
loss of plasticizers (new-car smell) to the environment.
• O-ring seals can be attacked by lubricants, especially at higher temperatures.
• Graphite’s friction and wear increase dramatically in the absence of water vapor at 10,000 m
altitude.

42.5  Temperature
A 10°C increase in temperature increases the chemical reaction rate to 2×, and thus, increased tempera-
ture increases the effects of environmental reactions. Increased temperatures also increase diffusion
rates and thus adhesive wear. But increased temperatures also increase the oxidation of wear surfaces,
and that can reduce the wear rate. Some materials used effectively at elevated temperatures show poor
wear at room temperature.
42-12 Design for Lubrication and Tribology

Especially in high-velocity contacts with high friction coefficients, the frictional energy being dissi-
pated in small areas of real contact can cause asperity temperatures to be hundreds of degrees above the
bulk temperature. This “flash temperature increase” can be calculated using Dunaevsky (1997).

42.6  Geometry, Kinematics, and Load


The effect of geometry on wear is most dramatic when the contact geometry is changing during opera-
tion and facilitating a different wear mode. A shaft in a bushing subjected to adhesive or abrasive wear
can increase the clearance to facilitate impact wear in addition to the adhesive or abrasive wear. Also, the
reduced size of the contact may increase surface stress to where one or both sides of the contact experi-
ence plastic deformation leading to a significant, possibly 10×, increase in wear rate.
The normal load on a contact generally has a linear effect on the resulting wear. The depth of wear
can thus be decreased by increasing the apparent surface area or vise versa. But there are limits to this
behavior when approaching the yield or softening point of the materials in contact. Plastic deformation
will bring increased friction due to van der Waals’ forces, possibly leading to increased surface tempera-
tures, faster diffusion, more severe adhesion, and galling failure.
Very smooth and flat surfaces in contact can experience so much van der Waals’ adhesion that it can
be difficult to separate those surfaces.
When velocities get very high (Mach 1 or 340 m/s), energy dissipated in contacts causes significant
heating of the wearing surfaces, hundreds of degrees above their bulk temperature, as demonstrated
by Green and Bushin (1991) and Griffioen and Winer (1985) and proven by casting structures in the
wear zone.
In lubricated rolling bearings, the combined effect of viscosity increases with pressure, and the elas-
tic deformation of concentrated contacts results in a friction coefficient of 0.001–0.005 (Jedzurski and
Moyer, 1997).
Fretting wear is characterized by short (<1 mm) reciprocating strokes or by strokes that are signifi-
cantly shorter than the contact zone, effectively preventing wear debris from escaping from between the
wear surfaces. This can lead to an experimental result with very little weight loss and the conclusion of
no wear. If, however, the actual application experiences significant vibration between the surfaces, the
wear could be orders of magnitude above that predicted by the experiment.

42.7  Developments
Recent developments related to some main tribological components in mechanical systems are as follows:
• Bearings: Bearing technology, from sliding, rolling element to hydrodynamic and hydrostatic,
has been analyzed to a high degree. This is mature technology; contact the vendors for the latest
design principles. Advancements are mainly due to the use of friction and wear coatings.
• Earthmoving equipment: Abrasion by soil and rock is mitigated using replaceable components or
application of cermets and coatings.
• Gears: Gear technology is evolving with the use of advanced PVD coatings to mitigate adhesion
and friction.
• Pumps: Vane and gear pump technology is evolving using new materials and advanced coatings
to mitigate wear.
• Seals: Seal technology is being advanced by new seal designs like brush and plate seals and the use
of advanced wear materials and coatings. Ferrofluidic magnetic particle seals have proven very
effective in high-speed vacuum spindles.
• O-rings: O-ring manufacture is a complex process, which can lead to quality variation. Many
vendors participate in the market based on a limited number of sources for starting materials. No
significant improvements are reported.
Design for Lubrication and Tribology 42-13

• Transmissions: Modeling has improved the efficiency and durability of geared transmissions, as
has the continuous development of lubricant additives.
• Others: Chain drives, continuously variable transmissions (CVT), couplings, CV joints, rope/wire
drives, splines.

Answers to Frequently Asked Questions

• Adhesive wear of rubbing surfaces is generally worse when a metal is rubbing against itself.
• Simulating friction/wear tests must include the correct materials and surface treatments as well as
lubricants and chemical environment.
• Erosion by solid particle is very sensitive to particulate material. As per Peterson (1980), never
substitute materials when performing simulating tests.
• Flash temperature theory calculates the instantaneous temperature increase at the surface inter-
face due to dissipation of frictional energy in the tiny real area of contact zones. Rub your clean
hand over a squeaky-clean countertop as fast as you can, and locally, nerves in your fingertips will
experience 100°C, causing a burning sensation.
• Friction coefficients greater than 1 can occur as adhesion between two very conformal surfaces,
say two flat and polished surfaces, can be substantial due to van der Waals’ attraction between
molecules, thereby exerting a significant traction force while the normal force may be zero, result-
ing in a coefficient of friction of ∞.
• Friction coefficients are a function of both contacting materials; the friction coefficient between a
steel skate and ice is substantially different from that between steel and wood, or steel and rubber.
• Graphite does not work as a solid lubricant in the absence of water or some other vapors.
• Rolling friction coefficients can be an order of magnitude lower than sliding.
• Static friction or breakaway friction is usually greater than the dynamic friction measured when
surfaces are in motion.
• Static friction can be a function of how long surfaces have been pressed together due to diffusion.
• Very smooth surfaces, <0.5 μm Ra, can lead to increased friction and galling due to van der Waals’
forces.
• Proof of the real area of contact size occurred when tests were conducted with a steel pin rubbing
on a glass disk. An infrared camera looking through the glass disk picked up the small areas of
real contact as hot spots moving over the contact surface as the frictional heat generated in the
real contact areas was dissipated. So the real area of contact equals the load divided by the yield
strength of the softer material, probably 1/10,000 of the apparent contact surface area.
• Increasing the hardness of a metal does not always result in reduced wear, as it generally also
increases brittleness.
• Stainless steel may be nitrided to improve wear resistance, but nitriding will tie up the chrome
in the stainless as chrome nitride, hindering the formation of the protective corrosion-resistant
chrome oxide layer, thus causing the stainless steel to rust.
• Some stainless steels can corrode in the presence of seawater or salty air.
• Phosphorous- and sulfur-based additives in lubricating oil can be very effective on steel surfaces
due to the formation of iron phosphates on the metal surface but generally do not reduce friction
of aluminum-, titanium-, or nickel-based alloys.
• Variability of wear rate measurements is usually high. Maximum values are often 3× the mini-
mum values, higher for poor performing material combinations.
• High-velocity contacts can generate enough frictional energy to substantially heat the surface
layers of the contact to hundreds of degrees above the bulk component temperature (Block, 1963).
• Water (>1000 ppm) in lubricating oil can reduce fatigue life and increase wear.
Wear volume is a function of load; thickness loss, a function of pressure.
42-14 Design for Lubrication and Tribology

Acknowledgment
The author acknowledges Tim Oertley for providing useful information.

References
Aluminum Standards and Data (2000), The Aluminum Association, Inc., Arlington, VA 22209.
ASM Handbook, ASM International, Materials Park, OH 44073-0002.
ASME Wear Control Handbook, ASME, Fairfield, NJ 07007.
Blok H. (1963) The flash temperature concept, Wear, 6, 483–494.
Bosma R. and Moes H. (1970) Design charts for optimum bearing configurations 2: The pivoted pad
thrust bearing, ASME Journal of Lubrication Technology, 92, 572–577.
DellaCorte C., Radil, K.C., Bruckner, R.J., and Howard, S.A. (2008) Design, fabrication, and performance of
open source generation I and II compliant hydrodynamic gas foil bearings, Tribology Transactions,
51(3), 254–264.
Dunaevsky V.V. (1997) Friction temperatures, Tribology Data Handbook, CRC Press, Boca Raton, FL,
pp. 462–473.
Engel P.A. and Yang Q. (1995) Impact wear of multi-plated electrical contacts, Wear, 181–183, 730–742.
Green I. and Bushin B. (1991) Measurements of asperity temperatures of read/write head slider bearing in
hard magnetic recording disk, Transactions of the ASME, Journal of Tribology, 113(3), 547.
Griffioen J.A. and Winer W.O. (1985) Infrared surface temperature measurements in a sliding ceramic-
ceramic contact, Proceedings of Leeds/Lyon Conference, IPC Publishing, London, U.K.
Jahanmir S. and Fischer T.E. (1994) Friction and wear of ceramics, Handbook of Lubrication and Tribology,
Vol. 3, E.R. Booser (Ed.), CRC Press, Boca Raton, FL, pp. 103–120.
Jedzurski T. and Moyer C.A. (1997) Rolling bearing performance and design data, Tribology Data
Handbook, CRC Press, Boca Raton, FL, pp. 645–668.
Lubrecht A.A., Ten Napel W.E., and Bosma R. (1988) Design charts for three lobe bearings, ASME Journal
of Lubrication Technology, 29–35.
Moes H. and Bosma R. (1971) Design charts for optimum bearing configurations 1: The full journal
bearing, ASME Journal of Lubrication Technology, 93, 302–306.
Neale, M.J. (1996) Frictional properties of materials C-8, Tribology Handbook, John Wiley & Sons,
New York.
Neale M.J. (1995) (Ed.) Tribology Handbook, John Wiley & Sons, New York.
Paxton R.R. (1979) Manufactured Carbon: A Self Lubricating Material for Mechanical Devices, CRC Press,
Boca Raton, FL.
Peterson M. (1980) Design considerations for effective wear control, Wear Control Handbook, ASME,
New York, pp. 438–439.
Peterson M.B. and Ramalingam S. (1980) Effect of temperature on friction for soft solids. Coatings for
tribological applications, Fundamentals of Friction and Wear of Materials, ASM, Metals Park, OH,
p. 341.
Rabinowicz E. (1965) Compatibility pairs for elemental materials, Friction and Wear of Materials,
John Wiley & Sons, New York, Fig 2.22, p. 39.
Rabonowicz E. (1980) Wear coefficients—Metal, Wear Control Handbook, ASME, New York, p. 491.
Rowe W.B. and O’Donoghue J.P. (1971) Design Procedures for Hydrostatic Bearings. The Machinery
Publishing Co., Ltd, London, U.K.
Ten Napel, W.E. and Bosma, R. (1980) Sinusoidal three-lobe bearings—Optimization and stability charts,
Transactions of the ASME, Journal of Lubrication Technology, 102, 416–424.
Vingsbo O. and Soderburg. Fretting map in which amplitude is plotted against wear rate.
Winer W.O. and Cheng H.S.(1980) Film thickness, contact stress and surface temperatures, Wear Control
Handbook, ASME, New York, pp. 137–139.
43
Fluid Film (Hydrodynamic)
Lubrication
Nomenclature (Dimensional, Nondimensional)....................................43-1
43.1 Introduction.....................................................................................43-2
43.2 Reynolds Equation...........................................................................43-3
43.3 Boundary Conditions.....................................................................43-4
Short-Bearing Theory  •  Long-Bearing Theory
43.4 Finite Journal Bearings...................................................................43-7
43.5 Dynamic Properties of Lubricant Films.................................... 43-11
43.6 Linearized Force Coefficients......................................................43-12
Andras Z. Szeri 43.7 Stability of a Flexible Rotor.......................................................... 43-16
University of Delaware References...................................................................................................43-20

Nomenclature
(Dimensional, Nondimensional)
b,B bearing damping matrix
C radical clearance
D bearing diameter
f,F oil-film force
fμ friction force
h,H film thickness
h n minimum film thickness
J(·), R(·) imaginary, real
k,K bearing stiffness matrix
k s,Ks shaft stiffness
L bearing length
M rotor mass
N rotational speed
OB center of bearing
OJ center of journal
p,P fluid pressure
℘ viscous dissipation
Q rotation
R journal radius
(·)r,t,(·)R,T radial, tangential component
(r,t),(R,T) coordinate directions
Re Reynolds number

43-1
43-2 Design for Lubrication and Tribology

S Sommerfeld number
(x,z),(X,Z) Cartesian coordinates
(x1 , y1 ),( X1 ,Y1 ) amplitude of motion (journal center)
(x2 , y2 ),( X2 ,Y2 ) amplitude of motion (rotor mass center)
α angular distance, pad—load line
β pad angular extent
μ lubricant viscosity
υ, Υ eigenvalue
φ attitude angle
ψ angular distance from load line
τ shear stress
θ angular distance from line of centers
θcaν position of film-cavity interface
θ1 pad leading edge
ω, Ω angular velocity
(·)0 pertaining to journal equilibrium position
.
(·) time derivative

43.1  Introduction
Hydrodynamic bearings are machine elements that utilize hydrodynamic lubrication principles in the
task of providing for effortless relative motion of adjacent components. There are several different types
of hydrodynamic bearings in use today, dictated by the precise nature of the motion that needs to be
controlled. If the center of curvature of both contacting surfaces is located on the same side of the con-
tact, we speak of a conformal bearing; when on opposite side, the bearing is counterformal (Figure 43.1).
Counterformal bearings, say ball bearings or gears, make contact over only a small area; in conse-
quence, the local pressure in the film separating the surfaces is large and the film is thin. Under these
conditions, the viscosity might be orders of magnitude larger than in the reference state. Furthermore,
the surfaces might deform elastically. We refer to counterformal bearings as “thin-film” bearings.
Conformal bearings, in contrast, have large contact area and their film is thick; “thick-film” bearings are
usually called upon to support rotating shafts. If the bearing needs to be placed at the end of the shaft to
support axial load, the bearing type employed is a thrust bearing (Figure 43.2). When the support of a
radially loaded rotating shaft is required, journal bearings are employed (Figure 43.3).
Although either a liquid or a gas might be used to lubricate a hydrodynamic bearing, in this chapter,
we discuss only liquid-lubricated bearings; because of the lubricant’s compressibility, the equation gov-
erning the distribution of film pressure in gas bearings is nonlinear, making the theoretical treatment of
gas bearings different, and more involved, than that of liquid bearings. Gas bearings will be dealt with
elsewhere within this volume.

r2 r2 U2
U2

r1
U1 r1

(a) U1 (b)

FIGURE 43.1  Bearing types: (a) conformal and (b) counterformal.


Fluid Film (Hydrodynamic) Lubrication 43-3

Sliding surface
WT or runner

R2 R1 N
L

W
B

Pads

FIGURE 43.2  Geometry of fixed-pad thrust bearing.

W/2 Q W/2
W
Q
Journal Bearing
e Lubricant film
2RJ N
OB O
J C
N
Q/2 Q/2
L 2RJ
2RB

FIGURE 43.3  Geometry of journal bearing.

If the geometry of the lubricant film of a hydrodynamic bearing is convergent–divergent in the direc-
tion of relative motion, as in a journal bearing, the film pressure will drop below ambient somewhere
past the position of minimum film thickness and the film will cavitate. In contrast, the geometry of
(traditional) thrust bearings is purely convergent (Figure 43.2), precluding cavitation of the film within
the clearance space. The problems normally associated with specifying pressure boundary conditions
at film-cavity interface are, therefore, nonexistent in thrust bearings. In addition, the relative motion
between bearing and runner is rigid-body translation without rotation, thus simplifying the right-hand
side of the governing equation. The mathematical treatment of thrust bearings is, therefore, simpler than
that of journal bearings. As journal-bearing analysis provides a richer variety of design considerations
than that of thrust bearings, the discussion here is restricted to journal bearings. For a more detailed
treatment of thick-film bearings than the one included here, the reader is referred to Szeri (2011).

43.2  Reynolds Equation


The distribution of pressure in the incompressible lubricant of a journal bearing is governed by the
Reynolds equation (Reynolds, 1886). In its normalized form, this equation is (Szeri, 2011)

2 . .
∂ ⎛ 3 ∂P ⎞ ⎛ D ⎞ ∂ ⎛ 3 ∂P ⎞ ⎡ ∂H ⎛e j ⎞⎤
⎜⎝ H +
⎟⎠ ⎜⎝ ⎟⎠ ⎜⎝ H ⎟⎠ = 12p ⎢ + 2 ⎜ cos q + e sin q⎟ ⎥ . (43.1)
∂q ∂q L ∂Z ∂Z ⎢⎣ ∂q ⎝w w ⎠ ⎥⎦

The nondimensional variables introduced here have the definition

x 2z h p
q= ; Z = ; H = = 1 + ecos q; P = 2 .
R L c mN ( R/C )

43-4 Design for Lubrication and Tribology

Equation 43.1 is a statement of the across-the-film averaged conservation of mass for the lubricant at the
point (θ, Z) of the bearing (reference) surface; the left-hand side is proportional to the rate of change of
pressure-induced flow, which is balanced by the rate of change of shear flow on the right. The pressure
flow is of two kinds, the first term on the left pertains to flow around the circumference of the bearing
while the second term relates to flow along its axis. That these orthogonal flows are uncoupled follows
from the thin-film approximation, which, in essence, linearizes the Navier–Stokes equation. Weak cou-
pling occurs only through a cavitation boundary condition.
Equation 43.1 holds for all similar journal bearings; bearing geometry is specified by the single similar-
ity parameter (D/L). For industrial bearings, this parameter has a value not too different from unity, and
in these bearings, the circumferential pressure flow is of the same order of magnitude as the axial pressure
flow. However, on occasion, a bearing is “long” and could be approximated by (D/L) → 0 in Equation 43.1;
in such a bearing, circumferential pressure flow dominates axial flow. Its geometric opposite, the so-called
short bearing, is approximated in Equation 43.1 by (L/D) → 0. In short bearings, axial flow balances circum-
ferential shear flow, and circumferential pressure flow is nonexistent. Even though both asymptotic cases
represent gross approximations to practical bearings, at least for a first estimate the long-bearing approxi-
mation might suffice when (D/L) < 1/4; in contrast, the short-bearing approximation might be useful when
(D/L) > 4. The great advantage of working with either the long-bearing or the short-bearing approximation
is that these possess simple analytical solutions but, of course, give only approximate answers.

43.3  Boundary Conditions


For liquid-lubricated journal bearings, specification of the trailing edge boundary conditions is made
difficult because of cavitation of the lubricant film somewhere in the diverging portion of the clearance
space. The exact position of this film-cavity interface is determined by the solution itself, making the
cavitation boundary condition nonlinear. Thus, though the Reynolds equation is a linear equation, the
existence of a cavitation boundary spoils the linearity of the finite-bearing problem. The long-bearing
and short-bearing approximations, in contrast, typically gloss over cavitation and employ linear bound-
ary conditions, thus maintaining the linearity of the problem.

Linear conditions:
1. Sommerfeld condition: P(0) = P(2π) = 0.
This condition can be applied to both long and short bearings (Sommerfeld, 1904). The resulting
pressure profile is skew-symmetric P(π + α) = − P(π − α) and calculates negative pressures of the
same magnitude as positive pressures. Only rarely is this arrangement found in practice and has
little physical significance.
2. Gümbel condition: P(0) = P(2π) = 0, P ≥ 0.
The half-Sommerfeld or Gümbel condition (Gümbel and Everling, 1925) disregards the negative
pressure loop of the Sommerfeld solution and, therefore, is applicable to β = π partial arc bearings.

Nonlinear conditions:
3. Swift–Stieber condition: P = ∂P/∂θ = 0 at θ = θcaν(Z).
This boundary condition (Swift, 1931; Stieber, 1933) satisfies mass conservation at the film-cavity
interface and yields pressure profiles that agree well with experiments. However, it is unable to
reproduce the subcavity pressure often manifested by experiments, downstream of the film-­cavity
interface. This boundary condition is not available to “short” bearings, and when applied to the
“long” bearing, θcaν (as all other quantities) is independent of the axial coordinate.
4. Coyne–Elrod and Floberg conditions (Floberg, 1965; Coyne and Elrod, 1971) are also based on flow
continuity considerations across the film-cavity interface. They are able to predict subcavity pres-
sures but are difficult to incorporate into a numerical scheme and, therefore, are not often used.
Fluid Film (Hydrodynamic) Lubrication 43-5

43.3.1  Short-Bearing Theory


The applicable forms of the differential equation and boundary conditions are (DuBois and Ocvirk, 1955)

2 .
d ⎛ 3 dP ⎞ ⎛ L ⎞ ⎡ ∂H ⎛e ⎞⎤
⎜H ⎟ = 12p ⎜⎝ ⎟⎠ ⎢ + 2 ⎜ cos q + ej sin q⎟ ⎥ ,
dZ ⎝ dZ ⎠ D ⎢⎣ ∂q ⎝w ⎠ ⎥⎦

P = 0 at Z = ±1, P (0) = P (2p) = 0, P ≥ 0. (43.2)


Solution of Equation 43.2 is obtained by double integration with respect to Z:

2 . .
⎛ L⎞ (1 − Z 2 ) ⎡⎛ j⎞ ⎛e⎞ ⎤
P = 6p ⎜ ⎟ ⎢⎜ 1 − 2 ⎟ esin q − 2 ⎜ ⎟ cos q ⎥ . (43.3)
⎝ D ⎠ (1 + ecos q)3 ⎢⎣ ⎝ w ⎠ ⎝ w ⎠ ⎥⎦

Relative to a right-handed coordinate system with unit vectors R and T, directed along and perpendicu-
lar to the line of centers (Figure 43.9), respectively, the components of the force exerted on the shaft by
the lubricant are given by the integrals

2 1 p
f R /LD ⎛ L⎞
FR ≡ =⎜ ⎟
mN (R/C)2 ⎝ D ⎠ ∫ ∫ P cosq dq dZ
−1 0

2 . .
⎛ L ⎞ ⎡ 4pe ⎛ j ⎞ 2p 2 (1 + 2e2 ) ⎛ e ⎞ ⎤
2
= −⎜ ⎟ ⎢ 2 2 ⎜
1−2 ⎟ + ⎥, (43.4a)
⎝ D ⎠ ⎢⎣ (11 − e ) ⎝ w⎠ (1 − e2 )5/2 ⎜⎝ w ⎟⎠ ⎥⎦

2 1 p
fT /LD ⎛ L⎞
FT ≡ =⎜ ⎟
mN (R/C ) ⎝ D ⎠
2 ∫ ∫ P cosq dq dZ
−1 0

2 . .
⎛ L⎞ ⎡ p e ⎛ j ⎞ 8pe ⎛ e ⎞ ⎤
2
=⎜ ⎟ ⎢ 2 3/2 ⎜
1−2 ⎟ + ⎥. (43.4b)
⎝ D ⎠ ⎢⎣ (1 − e ) ⎝ w ⎠ (1 − e2 )2 ⎜⎝ w ⎟⎠ ⎥⎦

When ε̇ = φ̇ = 0, the lubricant force f equilibrates the constant external load w, that is, w + f = 0. The jour-
nal center now occupies its “static equilibrium position” ε = ε0, φ = φ 0. Conversely, when ε̇ ≠ φ̇ ≠ 0, the
excess lubricant force propels the journal along an orbit about its static equilibrium position.
The attitude angle, the angular distance from load line to line of centers in the direction of shaft rota-
tion, can be obtained from j = tan −1 fT /f R ; in the equilibrium steady-state position (ε̇ = φ̇ = 0), it has the
value

⎛ p 1 − e2 ⎞
j 0 = tan −1 ⎜ ⎟ . (43.5)
⎝4 e ⎠

The friction variable cμ is calculated from the surface integral of the shear stress τxy = μRω/h:

⎛ R⎞ f 2p 2S
cm ≡ ⎜ ⎟ m = . (43.6a)
⎝ C⎠ w 1 − e2

43-6 Design for Lubrication and Tribology

TABLE 43.1  Performance Prediction for a Full Journal


Bearing, (L/D) = 1/4
Parameter Short-Bearing Approximation (Exact Solutiona)
ε 0.1 0.6 0.9
S 15.84 (16.20) 1.00 (1.07) 0.053 (0.074)
φ 82.71 (82.31) 46.32 (46.72) 20.83 (21.85)
cμ 314.2 (322.1) 24.67 (26.73) 2.40 (3.50)
qS 0.628 (0.621) 3.77 (3.72) 5.65 (5.59)
Computer solution of finite bearing obtained with the Swift–
a

Stieber boundary condition.

The friction variable is related to the viscous dissipation in the bearing ℘ = Rw f m through

℘ = (2pwNC)c m . (43.6b)

In Equation 43.6a,

2 2
mN ⎛ R ⎞ ⎛ D⎞ (1 − e2 )2
S≡ ⎜⎝ ⎟⎠ = ⎜⎝ ⎟⎠ , (43.7)
w /LD C L pe p 2 (1 − e2 ) + 16e2

is the Sommerfeld number, defined for the equilibrium state ε̇ = φ̇ = 0.


Predictions from short-bearing theory are compared to accurate two-dimensional numerical solu-
tions at (L/D) = 1/4 in Table 43.1 (Szeri, 2011).

43.3.2  Long-Bearing Theory


Although the differential equation for long bearings supports the Sommerfeld boundary condition,
solutions obtained under this condition show the journal moving sideways upon loading, perpendicu-
lar to the load line. As this condition occurs only in very special circumstances, Sommerfeld conditions
will not be considered here.
The Reynolds equation and Gümbel boundary conditions for the “long” journal bearing are

. .
d ⎛ 3 dP ⎞ ⎡ ∂H ⎛e j ⎞⎤
⎜⎝ H ⎟⎠ = 12p ⎢ + 2 ⎜ cos q + e sin q⎟ ⎥ ,
dq dq ⎢⎣ ∂q ⎝w w ⎠ ⎥⎦

P (0) = P (2p ) = 0, P ≥ 0. (43.8)


Under these conditions, bearing performance is characterized by

. .
⎧⎪ esin q(2 + cos q) ⎛ j ⎞ 1⎡ 1 1 ⎤ ⎛ e ⎞ ⎫⎪
P = 12p ⎨ 2 ⎜
1− 2 ⎟ + ⎢ − ⎥ ⎬, (43.9)
⎪⎩ (2 + e )(1 + ecos q) ⎝
2
w ⎠ e ⎣⎢ (1 + ecos q)2 (1 + e)2 ⎥⎦ ⎜⎝ w ⎟⎠ ⎪⎭

. .
⎪⎧ e2 ⎛ j⎞ 1 ⎡p 8 ⎤ ⎛ e ⎞ ⎪⎫
FR = −12p ⎨ ⎜ 1 − 2 ⎟ + − ⎬, (43.10)
⎩⎪ (2 + e )(1 − e ) ⎝
2 2
w ⎠ (1 − e2 )3/2 ⎢⎣ 2 p (2+e2 ) ⎥⎦ ⎜⎝ w ⎟⎠ ⎭⎪

Fluid Film (Hydrodynamic) Lubrication 43-7

TABLE 43.2  Long-Bearing Solutions


Gümbel Cond. Swift–Stieber Cond.
ε S φ S φ
0.1 0.33704 86.339 0.24144 69.032
0.2 0.16736 82.596 0.12373 66.900
0.3 0.11004 78.679 0.08376 64.464
0.4 0.08053 74.472 0.06289 61.638
0.5 0.06177 69.819 0.04931 58.296
0.6 0.04795 64.477 0.03895 54.234
0.7 0.03639 58.035 0.02993 49.098
0.8 0.02549 49.675 0.02110 42.181
0.9 0.01392 37.263 0.01151 31.666

. .
6pe ⎡ p ⎛ j ⎞ 4 ⎛ e ⎞⎤
FT = ⎢ 1 − 2 − ⎥, (43.11)
(2 + e2 ) ⎢⎣ (1 − e2 )1/2 ⎜⎝ w ⎟⎠ (1 − e2 ) ⎜⎝ w ⎟⎠ ⎥⎦

⎛ p 1 − e2 ⎞
j 0 = tan −1 ⎜ ⎟. (43.12)
⎝2 e ⎠

The friction variable for the steady case *ε = φ̇ = 0 is given by

⎛ R ⎞ f m esin j 2p 2S (2 + e2 )(1 − e2 )
⎜⎝ ⎟⎠ = + ,S= . (43.13)
C W 2 1 − e2 6pe p 2 (1 − e2 ) + 4e2

Table 43.2 compares long-bearing performance for two distinct boundary conditions (Szeri, 2011).

43.4  Finite Journal Bearings


The aspect ratio of industrial bearings is in the range 0.25 < L/D < 1.5; neither the short-bearing nor the
long-bearing approximation applies to these bearings. Furthermore, the angular extent is rarely 180°
or 360°; common ranges are 30° < β < 60° in pivoted-pad bearings and 120° < β < 160° in the “viscosity
pump” bearings of large rotating machinery. If the pad is centrally loaded, the ratio α/β = 1/2, where the
angle α is measured from the pad leading edge to the load line. For “offset” loading, α/β ≠ 1/2.
Over a finite-bearing pad of aspect ratio L/D and angular dimension β, the lubricant pressure satis-
fies Equation 43.1; solutions of this equation are usually obtained subject to the Swift–Stieber boundary
conditions:

P = 0 at Z ± 1 ⎫

⎪ (43.14)
P = 0 at q = q1 ,q1 + b ⎬.

P = ∂P /∂q = 0 at q = qcav ⎪⎭

The new parameter θ1 represents the angular distance between the line of centers and the leading edge
of the pad; this parameter reoccurs in the definition of the film thickness:
43-8 Design for Lubrication and Tribology

⎛ x⎞
H = 1 + ecos q = 1 + ecos ⎜ q1 + ⎟ , (43.15)
⎝ R⎠

Equation 43.15 also introduces the eccentricity ratio as the fourth defining parameter. The pressure
differential equation and its boundary conditions thus contain four dimensionless parameters in all:

{L/D, b, e,q1}. (43.16)



Two of these, L/D and β, define bearing geometry and are a given for a particular bearing. The other
two parameters, ε and θ1, specify the position of the rotating shaft within the bearing and are, therefore,
dependent on loading conditions and the viscosity of the lubricant. For an isothermal bearing, loading
conditions and lubricant viscosity can be characterized with just two parameters, the Sommerfeld num-
ber, S, and the ratio α/β (Figure 43.4).
On specifying the couple (ε, θ1), Equation 43.1 yields a pair of values (S, α/β) via the formulas

−1/ 2
(
S = FR2 + FT2 ) ,

=
⎡ (
a ⎣p − q1 + tan FT /FR ⎦
−1 ⎤
,
)
b b
(43.17)

where the force components are computed from

1 q1 + b 1 q1 + b
1 1
FR =
2 ∫∫ P cos q dq dZ ; FT =
2 ∫ ∫ P sinq dq dZ (43.18)
0 q1 0 q1

Thus, to each ordered pair (ε, θ1), there corresponds another ordered pair (S, α/β). Equations 43.1,
43.14, and 43.17 specify the complicated relationship between these two sets of parameters, which is
best portrayed in graphical form. This is illustrated in Figure 43.5 for an L/D = 1, β = 160°, centrally
loaded (α = β/2), fixed journal bearing. Figure 43.5a displays the relationship between ε = 1 − hn/C and
the Sommerfeld number, while Figure 43.5b shows the dependence of θ1 = π − (φ + β/2) on S. Note that
in preparing these solutions, the analyst varied θ1 on input at each eccentricity ratio until α = β/2 was
attained. Thus, α/β is not an unknown in Figure 43.5.

C
W W W
A C
B A B α
α
β β
(a) (b) (c)

FIGURE 43.4  Fixed-type journal bearings: (a) full bearing, β = 360° ; (b) partial bearing, centrally loaded, α = β/2;
and (c) partial bearing, eccentrically loaded (offset), α > β/2. (Reprinted with permission from Raimondi, A.A. and
Szeri, A.Z., Journal and thrust bearings, in Handbook of Lubrication, ed., E.R. Booser, Vol. II, CRC Press, Boca
Raton, FL. © 1984.)
Fluid Film (Hydrodynamic) Lubrication 43-9

To compute plots similar to Figure 43.5, the analyst specifies the couple (ε, θ1) and lets the numerics
determine the corresponding (S, αβ). During the course of the design, in contrast, the engineer specifies
the magnitude and the line of action of the load, the speed of rotation, and the viscosity of the lubricant,
that is, specifies the set (S, αβ) and permits the journal to select its own equilibrium position (ρθ1). Plots
such as Figure 43.5 reconcile these two viewpoints.

1.0
20,000
0.9
(dimensionless)

0.8
w 10,000
0.7 5,000
hn θ
2,000
C
n

0.6
h

β = 2.79 (160°)
Minimum-film-thickness variable,

0.5 L =1 Re = RωC =1 (Laminar)


D V
0.4

0.3

0.2

0.1

0
0 0.01 0.02 0.05 0.10 0.20 0.50 1.0 2.0 5.0 10
R 2
(a) Bearing characteristic no., S = C ( ) µN
P (dimensionless)

100

90
20,000
80
Position of min. film thickness, φ (deg)

10,000
70

60

50
5,000
40
2,000 w
30
Re=1 hn φ
20
β = 2.79 (160°)
10 L =1
D
0
0 0.01 0.02 0.05 0.10 0.20 0.50 1.0 2.0 5.0 10

( )
2 µN
R
Bearing characteristic no., S = (dimensionless)
(b) C P

FIGURE 43.5  Performance curves for a centrally loaded, fixed partial journal bearing: (a) minimum film thick-
ness, (b) attitude angle, (c) friction (power loss) variable, and (d) inlet flow variable. (Reprinted with permission
from Raimondi, A.A. and Szeri, A.Z., Journal and thrust bearings, in Handbook of Lubrication, ed., E.R. Booser,
Vol. II, CRC Press, Boca Raton, FL. © 1984.)
(continued)
43-10 Design for Lubrication and Tribology

103

5
β = 2.79 (160°)
L =1
D
2

102
(dimensionless)

50

0
,00
20

0
,00
10
2πWNC

00
20
H

5,0
00
2,0
Power loss variable.

10

Re= RωC = 1
2 V

0
0 0.01 0.02 0.05 0.10 0.20 0.50 1.0 2.0 5.0 10

( )
2 µN
R
Bearing characteristic no., S = (dimensionless)
(c) C P

5.0

20,000
10,000
4.0
(dimensionless)

00
3.0 5,0 0
0
2,0
RCNL
Q

β = 2.79 (160°)
R =1 L =1
2.0 e
Inlet flow variable,

D Q–QS

Q Q
1.0 S
2

0
0 0.01 0.02 0.05 0.10 0.20 0.50 1.0 2.0 5.0 10

( )
2 µN
Bearing characteristic no., S = R (dimensionless)
(d) C P

FIGURE 43.5 (continued)  Performance curves for a centrally loaded, fixed partial journal bearing: (a) minimum
film thickness, (b) attitude angle, (c) friction (power loss) variable, and (d) inlet flow variable. (Reprinted with
permission from Raimondi, A.A. and Szeri, A.Z., Journal and thrust bearings, in Handbook of Lubrication, ed.,
E.R. Booser, Vol. II, CRC Press, Boca Raton, FL. © 1984.)
Fluid Film (Hydrodynamic) Lubrication 43-11

Raimondi and Boyd (1958) published an extensive set of solutions of centrally loaded, isothermal
bearings. Pinkus and Sternlicht (1961) compiled solutions for eccentrically loaded partial arc bearings.
The performance curves in Figure 43.5 are taken from Raimondi and Szeri (1984); they are computed at
various values of the Reynolds number Re = ρCRω/μ.

43.5  Dynamic Properties of Lubricant Films


In Figure 43.6, we schematize a rotor of weight 2ω and its support. Under steady load, the journal center
is displaced from the bearing center to the steady operating position OJs.
Assuming, initially, rigid supports, rotor response to small excitation, say an imbalance, will be
as shown by the solid curve in Figure 43.7. This type of response would be expected were the rotor
running on rolling contact bearings. Rigidly supported rotors cannot be operated at a critical speed
and can become “hung” on the critical when attempting to drive through. On replacing the rigid
supports with hydrodynamic bearings, however, a lubricant film adds another spring and consid-
erable damping to the system. Such changes are expected to (1) lower the critical speed below that
calculated for rigid supports and (2) reduce the amplitude of vibration, as indicated by the dashed
curve in Figure 43.7.
In this example, the excitation is mass imbalance and occurs at running speed. In practice, excitation
frequencies can be different from the shaft speed. In addition, lubricant films themselves can originate
destructive self-excited vibrations. A classical case is “subharmonic resonance,” occurring at one-half
of the running speed.

Rotor
Oil film
damping
Shaft

ω
ks
Journal
Oil film
spring
2w

FIGURE 43.6  Schematics of rotor-shaft configuration.


Vibration amplitude, (X-direction)

No oil film or
" rigid " oil film

With oil
film

Speed, ω

FIGURE 43.7  Effect of oil film on shaft response. (Reprinted with permission from Raimondi, A.A. and
Szeri, A.Z., Journal and thrust bearings, in Handbook of Lubrication, ed., E.R. Booser, Vol. II, CRC Press, Boca
Raton, FL. © 1984.)
43-12 Design for Lubrication and Tribology

Ob
y
kxx bxx

byy
OJ
df
ωt
kyy
Elliptical orbit
x OJS

FIGURE 43.8  Representation of oil film as a simple dynamical system of springs and dampers (cross-film springs
k xy, kyx and dampers byx, bxy are not shown). (Reprinted with permission from Raimondi, A.A. and Szeri, A.Z.,
Journal and thrust bearings, in Handbook of Lubrication, ed., E.R. Booser, Vol. II, CRC Press, Boca Raton, FL.
© 1984.)

The rotor-shaft configuration of Figure 43.6 is reduced to a simple dynamic system of springs and
dashpots in Figure 43.8. A mass ω/g (one-half the rotor mass) is concentrated at Ojs(xs, ys), the steady
running position of the journal. If some excitation df occurs at frequency ω, the mass center will respond
by orbiting about Ojs as indicated. It is assumed here that the dynamic displacement from the static
equilibrium position is small (when the dynamic displacement is large, the present discussion no longer
applies). Displacements in the x and y direction, respectively, are opposed by the oil-film force:

⎛ −df x ⎞ ⎛ kxx kxy ⎞ ⎛ dx ⎞ ⎛ bxx bxy ⎞ ⎛ dx ⎞


⎜ ⎟= ⎜ ⎟+ ⎜ ⎟. (43.19)
⎜⎝ −df y ⎟⎠ ⎜⎝ k yx k yy ⎟⎠ ⎜⎝ dy ⎟⎠ ⎜⎝ byx byy ⎟⎠ ⎜⎝ dy ⎟⎠

The spring coefficients k xx, k xy, kyx, kyy and the damping coefficients bxx, bxy, byx, byy enable linear repre-
sentation of the oil-film force. For example, if the vibrating system consisted solely of a mass M running
on hydrodynamic oil film and excited by a force df at frequency ω, the motion of the mass center would
be described by

⎛ dx⎞ ⎛ kxx kxy ⎞ ⎛ dx ⎞ ⎛ bxx bxy ⎞ ⎛ dx ⎞ ⎛ sin w t ⎞


M⎜ ⎟ +⎜ ⎜ ⎟ +⎜ ⎜ ⎟ = df ⎜ ⎟.
⎜⎝ dy ⎟⎠ ⎝ k yx k yy ⎟⎠ ⎜⎝ dy ⎟⎠ ⎝ byx byy ⎟⎠ ⎜⎝ dy ⎟⎠ ⎜⎝ cos w t ⎟⎠

A more thorough examination of the problem reveals that the force on an orbiting journal is dependent
on acceleration as well as on position and velocity (see, e.g., Szeri et al., 1983). For simplicity, however,
we do not include acceleration, that is, fluid inertia, effects here.
When the bearing constitutes an element in a complex dynamical system, it is usually incorporated
in the system equations of motion through its spring and damping coefficients. All eight oil-film coef-
ficients are required in order to make accurate dynamical analyses of rotor-shaft configurations.

43.6  Linearized Force Coefficients


Let OB represent the center of the bearing as in Figure 43.9. The static equilibrium position of the rotat-
ing shaft is Ojs; here, the eccentricity ratio is e 0 and the attitude angle φ 0. The components of the lubricant
force, resolved along the fixed directions (R,T), are (f0)R,(f0)T.
Fluid Film (Hydrodynamic) Lubrication 43-13

OB

∆φ e0+∆e t
φ0 ft fT

l in e
e0

d
OJ

Loa ( f0 )T f
OJs T
( f0 )x
( f 0 ) y fR fr r
y
x
(fo)R f0

FIGURE 43.9  Force decomposition in journal bearings (for illustration only, not to scale).

If there is a small unbalanced force on the journal, the journal will orbit about its static equilibrium
position (e 0,φ 0). At a particular instant, it will occupy position OJ, a generic point of the orbit. In OJ, the
eccentricity ratio is e = e 0 + Δe, the attitude angle is φ = φ 0 + Δφ, and the journal possesses instantaneous
velocity (e·,eφ·). The instantaneous lubricant force is now (fr, f t) relative to the instantaneous radial and
tangential coordinates (r,t).
The instantaneous force is now referred to the fixed (R, T) coordinate system:

⎛ fR ⎞ ⎛ fr ⎞ ⎛ cos Δj −sin Δj ⎞
⎜ ⎟ =Q⎜ ⎟;Q = ⎜ . (43.20a)
⎜⎝ fT ⎟⎠ ⎜⎝ ft ⎟⎠ ⎝ sin Δj cos Δj ⎟⎠

For small departures from equilibrium, we have the approximation

cos Δj ≅ 1; sin Δj ≅ Δj ,

and Equation 43.20a can be written as

f R = f r − Δj ft ; fT = Δj f r + ft . (43.20b)

We are interested in evaluating the increase in the force components over their equilibrium values:

Δf R = f R − ( f 0 )R ,

ΔfT = fT − ( f 0 )T .

Keeping in mind (43.20b), the force excess can be evaluated from

Δf R = f r − Δj ft − ( f 0 )R ,
(43.21)
ΔfT = Δj f r + ft − ( f 0 )T .

43-14 Design for Lubrication and Tribology

The scalar functions fr, f t are now approximated via Taylor series expansion about the equilibrium posi-
tion. Neglecting higher order terms, for fr we have

⎛ ∂f ⎞ ⎛ ∂f ⎞ ⎛ ∂f ⎞ ⎛ ∂f ⎞ .
f r = ( f 0 )R + ⎜ R ⎟ de + ⎜ R ⎟ dj + ⎜ R ⎟ de + ⎜ .R ⎟ dj , (43.22)
⎝ ∂e ⎠ ⎝ ∂j ⎠ ⎝ ∂e ⎠ ⎝ ∂j ⎠

.
where all partial derivatives are evaluated in the equilibrium position e = j = 0. The expansion is similar
for f t.
Substituting the Taylor expansions into the nondimensional form of Equation 43.21, we obtain

⎛ ∂FR ∂FR ⎞ ⎛ ∂FR ∂FR ⎞


− FT ⎛ de⎞ .
⎛ dFR ⎞ ⎜ ∂e ∂(j /w ) ⎟ ⎛ d(e /w ) ⎞
.
∂j ⎟ ⎜ ∂(e/w )
⎜ ⎟ =⎜ ⎟⎜ ⎟ +⎜ ⎟⎜ ⎟. (43.23)
⎜⎝ dFT ⎟⎠ ⎜ ∂FT ∂FT ⎜ ⎟ ∂FT ∂FT ⎟ ⎜ d(j. /w )⎟
⎜⎝ ∂e + FR ⎟⎟ ⎝ d∂⎠ ⎜⎜ ⎝ ⎠
∂j ⎠ ⎝ ∂(e/w ) ∂(j /w ) ⎟⎠

To calculate the force derivatives required in Equation 43.23, we have to go to the Reynolds equation
(43.1). However, under dynamic loading conditions, it is expedient to employ another nondimensional
pressure, defined by

P
Pˆ= . . (43.24)

(
1 − 2j /w )
. 
Note that Pˆ → P as j → 0. In terms of the “dynamic” pressure P the Reynolds equation has the form

∂ ⎛ 3 ∂ Pˆ ⎞ ⎛ D ⎞ ∂ ⎛ 3 ∂ Pˆ ⎞
2 .
e /w
⎜ H +
⎟ ⎜ ⎟ ⎜ H ⎟ = −12pesin q + 24p . . (43.25)
∂q ⎝ ∂q ⎠ ⎝ L ⎠ ∂Z ⎝ ∂Z ⎠ (1 − 2j /w )


It should be noted from Equation 43.25 that the dynamic pressure P is a function of four variables
ε,φ,ε̇ /ω and φ̇ /ω; dependence on the attitude angle φ enters via the film thickness formula:

H = 1 + cosq = 1 + e(y − j ),

where
ψ is the angle measured from the (fixed) load line
φ is the (variable) attitude angle

Here, for simplicity, we solve Equation 43.25 subject to the Gümbel boundary condition to obtain the
force coefficients. The nondimensional lubricant-force components are

. 1 π
⎛ j ⎞ ⎡1 ⎤
Pˆ cos q dq dZ ⎥ ,
FR = ⎜ 1 − 2 ⎟ ⎢
⎝ w ⎠ ⎢2

∫∫
0 0


. 1 π
⎛ j ⎞ ⎡1 ⎤
FT = ⎜ 1 − 2 ⎟ ⎢
⎝ w ⎠ ⎢2

∫ ∫ Pˆ sinq dq dZ ⎥⎥⎦ . (43.26)
0 0
Fluid Film (Hydrodynamic) Lubrication 43-15

Employing the notation


1 π
1
2 ∫ ∫ Pˆ cosq dq dZ = Fˆ ,
0 0
R

1 π
1
2 ∫ ∫ Pˆ sinq dq dZ = Fˆ , T

0 0

we write Equation 43.26 in the symbolic form


. .
⎛ j ⎞ ⎧⎪ e /w ⎫⎪
FR = ⎜ 1 − 2 ⎟⎟ FˆR ⎨e, j , . ⎬,

⎝ w ⎠ ⎩⎪ (1 − 2j /w ) ⎭⎪
. .
⎛ j ⎞ ⎪⎧ e /w ⎪⎫
FT = ⎜ 1 − 2 ⎟ FˆT ⎨e, j , . ⎬. (43.27)
⎝ w ⎠ ⎪⎩ (1 − 2j /w ) ⎭⎪

.
The partial derivatives of the force components, Equation 43.27, are evaluated at ε = ε0, φ = φ 0, ε  = φ̇ = 0
(Szeri, 2011)

⎛ ∂FˆR ∂FˆR Fˆ ⎞ ⎛ ∂FˆR 2FˆR



− T − .
⎛ dFR ⎞ ⎜ ∂e e∂j e ⎟ ⎛ de ⎞ .
⎜ ∂e /w ⎟ e ⎛ d(e /w ) ⎞
⎜ ⎟ =⎜ ⎟⎜ ⎟ +⎜ ⎟ ⎜ ⎟. (43.28)
⎜⎝ dFT ⎟⎠ ⎜ ∂ FˆT ∂ FˆT FˆR ⎟ ⎜⎝ edj ⎟⎠ ⎜ ∂ FˆT ˆ
FT
⎟ ⎜⎝ ed(j. /w )⎟⎠
⎜ − ⎟ ⎜⎝ . − ⎟
⎝ ∂e e∂j e ⎠ ∂e /w e ⎠

The first matrix on the right describes the oil-film response to shaft displacement and is called the (non-
.
dimensional) stiffness matrix K. The second matrix describes response to velocities (ε /ω), εd(φ̇ /ω) and
is called the (nondimensional) damping matrix B.
The nondimensional stiffness and damping matrices are related to their dimensional counterparts
k and b, respectively, through

C Cw
K= 2 k; B = 2 b. (43.29a)
LD mN ( R/C ) LD mN ( R/C )

The elements of k and b have dimensions force/length and force/velocity, respectively. The alternative
definition

C Cw
K= k and B = b (43.29b)
W W

is often employed in the literature. Then K‾= SK and B‾ = SB, where S is the Sommerfeld number.
For the short bearing under half-Sommerfeld conditions, the stiffness and damping matrices (relative
to the R,T coordinate system) are (Szeri, 2011)

⎛ −8pe(1 + e)2 −p 2 ⎞ ⎛ −2p 2 (1 + 2e2 ) 8pe ⎞


2 2⎜
⎛ L ⎞ ⎜ (1 − e2 )3 2 3/2 ⎟
(1 − e ) ⎛ L⎞ (1 − e2 )5/2 (1 − e2 )2 ⎟
K =⎜ ⎟ ⎜ 2 ⎟; B=⎜ ⎟ ⎜ ⎟ . (43.30)
⎝ D ⎠ ⎜ p (1 + 2e2 ) −4pe ⎟ ⎝ D⎠ ⎜ 8pe −2p 2 ⎟
⎜⎝ (1 − e2 )5/2 (1 − e2 )2 ⎟⎠ ⎜⎝ (1 − e2 )2 (1 − e2 )3/2 ⎟⎠

43-16 Design for Lubrication and Tribology

In most cases, the static load acts along the vertical or through the pivot point for a pivoted pad. It is
convenient then to refer the stiffness and damping matrices to the (X = x/C, Y = y/C) coordinate system
where x is along the load line and y points to the right (counterclockwise shaft rotation), as depicted in
Figure 43.9. The components K XX,…, BYY are then obtained by simple clockwise rotation through the
static attitude angle φ 0:

⎛ K XX K XY ⎞ ⎛ K RR K RT ⎞ T ⎛ BXX BXY ⎞ ⎛ BRR BRT ⎞ T


⎜⎜ ⎟ = Q ⎜⎜ ⎟Q ; ⎜⎜ ⎟ = Q ⎜⎜ ⎟Q . (43.31)
⎝ K YX K YY ⎟⎠ ⎝ K TR K TT ⎟⎠ ⎝ BYX BYY ⎟⎠ ⎝ BTR BTT ⎟⎠

The results of Figure 43.10 pertain to a single, fixed pad. Industrial journal bearings are routinely com-
prised of several pivoted pads; that being the case, the results obtained for fixed pads must be con-
verted to represent pivoted pads (Lund, 1964), which then must be assembled to obtain bearing response
(Nicholas et al., 1979).

43.7  Stability of a Flexible Rotor


Consider a weightless elastic shaft supporting a disk of mass 2M at its midpoint. The shaft, in its turn,
is supported by identical, single-pad journal bearings at its ends. In Figure 43.11, the geometric center
of a bearing pad is designated by OB . Under the steady load w = Mg, the center of the rotating journal
moves out of OB and into its static equilibrium position Ojs. As long as the rotor-bearing system is left
undisturbed, the rotor will remain in this position. If, however, disturbed and the disturbances are
small, the journal will leave its equilibrium position and proceed along a closed orbit around it. Now,
at time t, the journal center is in OJ (x1, y1) and the mass center of the rotor has moved to OM(x2,y 2). This
is what occurs in well-designed, stable rotor-bearing systems. Under other, unstable conditions, the
journal is unable to find a limit cycle, it spirals outward until metal-to-metal contact occurs between
journal and bearing. There is great physical importance attached, therefore, to knowing the criteria
that demarcates stable from unstable operation. To find these criteria, we begin with the equations of
motion (Figure 43.11):
Rotor:

−ks (x2 − x1 ) = Mx2


(43.32)
−ks ( y2 − y1 ) = My2 .

Bearing:

2df x + ks (x2 − x1 ) = 0
(43.33)
2df y + ks ( y2 − y1 ) = 0.

where
k s is the shaft stiffness
dfx, dfy are the components of the lubricant force on the shaft, in excess of the equilibrium force
fx = w, fy = 0

The force increments dfx, dfy are given by Equation 43.19.


It is assumed here that OJ and OM undergo harmonic motion of the type
Fluid Film (Hydrodynamic) Lubrication 43-17

2
10
R+C
β = 2.79 (160°)
50 L =1
U D
y
R
e
20 W R =1.0
e
K
xy
Stiffness coefficient. C K (dimensionless)

h R =20,000
10 n e
R =1.0 x
e K
xx
5.0
20,000
W

2.0 K
yy

1.0 1.0
1.0
20,000 K
yx
0.5 R =20,000
-K e
yx
R =1.0
e
0.2

0.1 1.0

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
(a) Bearing eccentricity ratio. E = e/c (dimensionless)

102
R+C
β = 2.79 (160°)
L =1
50 D U

R y
e
W
20
(dimensionless)

h B
10 n xx
x

5.0
R =1.0
R =1.0 e
UB

e
R =20,000
e
W
Damping coefficient. C

2.0

1.0
B =B
xy yx B
yy
0.5 R =1.0
e
R =20,000
e
R =20,000
0.2 e R =1.0
e

0.1

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
(b) Bearing eccentricity ratio. E = e/c (dimensionless)

FIGURE 43.10  Performance curves for a centrally loaded, fixed partial journal bearing: (a) bearing stiffness and
(b) bearing damping. (Reprinted with permission from Raimondi, A.A. and Szeri, A.Z., Journal and thrust bear-
ings, in Handbook of Lubrication, ed., E.R. Booser, Vol. II, CRC Press, Boca Raton, FL. © 1984.)
43-18 Design for Lubrication and Tribology

OM
ω(N)
OB Ks

OJs y
2Mg
OJ (x1 ,y1 ) OM (x2 ,y2 )

x
OJs(0,0 ) y
dfy
OJ (x1 ,y1 )
k3(y2–y1)

dfx
OM (x2 ,y2 ) k3(y2–y1)
k3(x2–x1)
k3(x2–x1)

FIGURE 43.11  Schematic of flexible rotor supported on two identical journal bearings.

⎛ x1 ⎞ ⎛ x1 ⎞ ⎛ x2 ⎞ ⎛ x2 ⎞
ut ut
⎜ ⎟ =⎜ ⎟e ; ⎜ ⎟ =⎜ ⎟ e . (43.34)
⎜⎝ y1 ⎟⎠ ⎜⎝ y1 ⎟⎠ ⎜⎝ y2 ⎟⎠ ⎜⎝ y2 ⎟⎠

The frequency of this motion is, in general, complex:

u = R(u) + iJ (u). (43.35)

eut = e R (u )t (cos J (u)t + i sin J (u)t ).



The real part (υ) is the damping exponent; if (υ) < 0, the motion is stable and is unstable if (υ) > 0.
The neutral state of stability (threshold of stability) is characterized by (υ) = 0. The imaginary part (υ)
is the damped natural frequency (the orbiting frequency) of the motion.
Substituting Equation 43.34 into Equations 43.32 and 43.33 and taking into account Equation 43.19,
we obtain

⎛ a + 2K xx + 2 ϒBxx 2K xy + 2 ϒBxy ⎞ ⎛ X1 ⎞ ⎛ 0⎞
⎜⎝ 2K yx + 2 ϒB yx ⎜ ⎟ = ⎜ ⎟. (43.36)
a + 2K yy + 2 ϒB yy ⎟⎠ ⎜ Y ⎟ ⎜⎝ 0⎟⎠
⎝ 1⎠

Here we applied the notation

C ⎧x y ⎫
{K ij , Bij , K s } = {kij , w bij , ks }; { X1 ,Y1} = ⎨ 1 , 1 ⎬ ;
W ⎩C C ⎭
u w K ϒ 2Ω2 (43.37)
ϒ= ; Ω= ; a = 2s 2 .
w wN ϒ Ω +1

where w N = ks /M is the natural frequency of the rotor.


Fluid Film (Hydrodynamic) Lubrication 43-19

Equation 43.36 possesses a nontrivial solution if and only if the system determinant vanishes. By
expanding the determinant and equating it to zero, we would obtain a fourth-order polynomial in Y.
Solving this so-called frequency equation would yield Y. Instead of solving directly for Y, however, we
seek conditions for marginal stability, a state that is characterized by the vanishing of the real part of Y.
To this end, we separate the determinant in Equation 43.36 into its real and imaginary parts, which are
then individually equated to zero. In this manner we obtain

2(K xy B yx + K yx Bxy − K yy Bxx − K xx B yy )


a= .
(Bxx + B yy )

Using the definition for α, we find the instability threshold value of the relative frequency as

a
Ω2c = , (43.38)
ϒ (K S − a )
2

in terms of the yet undetermined whirl ratio Y. If Ω < Ωc, the system is stable; at Ωc = 1, instability sets in
at the system natural frequency ωN.
To render Equation 43.38 useful, we need to determine the whirl ratio Y = ν/ω; this can be accom-
plished by expanding the imaginary part of Equation 43.36 to obtain (Szeri, 2011)

a 2 + 2(K xx + K yy )a + 4(K xx K yy − K xy K yx )
ϒ2 = − . (43.39)
4(Bxx B yy − Bxy B yx )

In the state of neutral stability, Υ is purely imaginary, Υ = i(Υ); thus,

uWhirl
= −ϒ 2 , (43.40)
w

Note that both Ωc and Υ are functions of the (dimensionless) shaft stiffness KS and the elements of the
stiffness and damping matrices K and B; the latter two being evaluated at, and depending on, the static

30
Stability parameter c/w Mω2
( )

20 Unstable
c k = 100
w s

10
10 4
2

Stable 1

0
0 0.01 0.1 1 10
( )
S L
D
2

FIGURE 43.12  Stability of single mass rotor on full journal bearings, mounted on a rigid support. (Reprinted
with permission from Raimondi, A.A. and Szeri, A.Z., Journal and thrust bearings, in Handbook of Lubrication,
ed., E.R. Booser, Vol. II, CRC Press, Boca Raton, FL. © 1984.)
43-20 Design for Lubrication and Tribology

equilibrium position of the journal (ε 0, φ 0). Thus, for a given system, a stability chart can be prepared as
a sole function of the static position of the journal or, alternatively, as a sole function of the Sommerfeld
number.
The stability characteristics of a flexible rotor carried on full journal bearings supported on a rigid
foundation can be estimated from Figure 43.12 (Raimondi and Szeri, 1984).

References
Cameron, A. 1966. The Principles of Lubrication. John Wiley & Sons, New-York.
Coyne, J. C. and Elrod, H. G. 1971. Conditions for the rupture of a lubricating film, Part II, New boundary
conditions for Reynolds’ equation. ASME J. Lubr. Technol., 93, 156–167.
DuBois, G. B. and Ocvirk, F. W. 1955. The short-bearing approximation for plain journal bearings. Trans.
ASME, 77, 1173–1178.
Floberg, L. 1965. On hydrodynamic lubrication with special reference to sub-cavity pressures and number
of streamers in cavitation regions. Acta Polytech. Scand. Mech. Eng. Ser., 19, 3–35.
Gümbel, L. and Everling, E. 1925. Reibung und Schmierung in Machinenbau. Krayn, Berlin, Germany.
Lund, J. W. 1964. Spring and damping coefficients for the tilting-pad journal bearing. ASLE Trans.,
7, 342–352.
Nicholas, J. C., Gunter, E. J., and Allaire, P. E. 1979. Stiffness and damping properties for the five-pad tilt-
ing pad bearing. ASLE Trans., 22, 113–124.
Pinkus, O. and Sternlicht, B. 1961. Theory of Hydrodynamic Lubrication. McGraw-Hill, New York.
Raimondi, A. A. and Boyd, J. 1958. A solution for the finite journal bearing and its application to analysis
and design, Trans. ASLE, 1, 159–209.
Raimondi, A. A. and Szeri, A. Z. 1984. Journal and thrust bearings. In Handbook of Lubrication, ed.,
E. R. Booser, Vol. II, CRC Press, Boca Raton, FL.
Reynolds, O. 1886. On the theory of lubrication and its application to Mr. Beauchamp Tower’s experi-
ments. Philos. Trans. R. Soc. Lond., 177, 157–234.
Sommerfeld, A. 1904. Zur hydrodynamischen Theorie der Schmiermittelreibung. Zeit. f. Math. Phys.
50, 97–155.
Stieber, W. 1933. Das Schwimmlager. Krayn, Berlin, Germany.
Swift, H. W. 1931. The stability of lubricating films in journal bearings. Proc. Inst. Civil Eng., 233, 267–322.
Szeri, A. Z. 2011. Fluid Film Lubrication. 2nd edn. Cambridge University Press, Cambridge, U.K.
Szeri, A. Z., Raimondi, A. A., and Giron, A. 1983. Linear force coefficients for squeeze-film damper. ASME
J. Lubr. Technol., 105, 326–334.
44
Journal Bearings
Nomenclature..............................................................................................44-1
44.1 Introduction.....................................................................................44-2
44.2 Fixed Geometry Journal Bearing Stability..................................44-2
44.3 Tilting Pad Journal Bearings........................................................ 44-6
Pivot Loading  •  Pad Pivot Offset  •  Pad Preload  •  Pad Length-to-
Diameter Ratio  •  Frequency Dependency
44.4 Tilting Pad Pivots..........................................................................44-10
44.5 Tilting Pad Bearing Lubricating Flow........................................44-12
John C. Nicholas 44.6 Advanced Tilting Pad Journal Bearings....................................44-12
Lufkin–Rotating 44.7 Conclusions....................................................................................44-14
Machinery Technology References...................................................................................................44-15

Nomenclature
A area, mm2 (in.2)
c, cb bearing radial clearance, mm (in.)
Cd, Cb bearing diametral clearance, mm (in.)
Cp, cp pad diametral, radial clearance, mm (in.)
d inlet orifice diameter (in.)
D journal diameter, mm (in.)
e bearing eccentricity, mm (in.)
F x , F y X, Y direction forces, N (lbf)
L bearing axial length, pad axial length, mm (in.)
Lu = W/(L*D) bearing unit loading, MPa (lbf/in.2)
m = 1 − (Cb/Cp) pad preload (dim)
n number of inlet orifices
np number of pads
N rotor speed, rpm
Ob, Oj, Op bearing, journal, pad center
P pressure, MPa (lbf/in.2)
Pin, Ph inlet, housing pressure, MPa (lbf/in.2)
Q flow Lpm (gpm)
R, Rp journal, pad radius, mm (in.)
Rb tilting pad bearing radius, mm (in.)
S Sommerfeld number (dim)
Tmax maximum Babbitt temperature, °C (°F)
Us = (π*N*D)/720 journal surface velocity, m/s (f/s)
W bearing load, N (lbf)

44-1
44-2 Design for Lubrication and Tribology

X horizontal coordinate (dim)


Y vertical coordinate (dim)
α = ϕp/χ pad pivot offset (dim)
χ pad arc length (deg)
ε = e/c bearing eccentricity ratio (dim)
μ lubricant viscosity, N s/mm2 (lbf s/in.2)
ϕ angle from the +X-axis, with rotation, to the unloaded pad pivot
ϕp angle from pad leading edge to pivot (deg)
ψ journal attitude angle (deg)
ω journal rotational speed (1/s)

44.1  Introduction
The journal bearing designs described in this chapter are common to many types of industrial turboma-
chinery, including, but not limited to, steam turbines, gas turbines, centrifugal compressors, axial com-
pressors, gas expanders, motor/generators, and pumps. The types of bearings most commonly found in
turbomachinery include plain cylindrical, axial groove, pressure dam, elliptical or lemon bore, offset
half, multilobe, and tilting pad.
The fundamental geometric parameters for all journal bearings are journal diameter, pad arc angle,
length-to-diameter ratio (L/D ratio), and operating clearance. For the fixed geometry bearing types con-
sisting of multiple pads and for tilting pad bearings, there are also variations in the number of pads, pad
preload, pad pivot offset, and load orientation (bearing load on or between pads for a fixed geometry bear-
ing and bearing load on or between pivots for a tilting pad bearing). In addition to the geometric param-
eters, there are several important operating parameters including oil type (oil viscosity as a function of
temperature), oil inlet temperature, rotating speed, bearing gravity load, and applied external loads. Volute
loads in pumps, steam loads in steam turbines, and mesh loads in gear boxes are examples of external loads.
The vibration attenuation for any rotor/bearing system is significantly influenced by the journal bear-
ing design. A good bearing design will provide sufficient bearing damping for vibration suppression.
Additionally, the bearing must operate at reasonable metal temperatures for long life. A poor design
will lead to vibration problems or high-temperature operation, both resulting in reduced bearing life
and increased downtimes.
Journal bearing design has evolved over the years from primarily fixed geometry bearings to tilting
pad bearings. Fixed geometry or sleeve bearings have the annoying property of creating an excitation
force that can drive the rotor unstable by creating a subsynchronous vibration. This phenomenon usu-
ally occurs at relatively high rotor speeds and/or light bearing loads. This inherent bearing-induced
instability can be completely eliminated with the tilting pad bearing.

44.2  Fixed Geometry Journal Bearing Stability


Fixed geometry bearings create an excitation force that can drive the rotor unstable by creating a sub-
synchronous vibration. This phenomenon usually occurs at relatively high rotor speeds and/or light
bearing loads or, more generally, at high Sommerfeld numbers. The Sommerfeld number, S, a dimen-
sionless value, is defined in the following:

m × N × L × D3
S= (44.1)
60 × W × C 2d

The problem with fixed geometry bearings (i.e., all journal bearings excluding tilting pad bearings)
is that they support a resultant load with a displacement that is not directly in line with the resultant
Journal Bearings 44-3

load vector but at some angle with rotation from the load vector. This angle can approach 90° for a high
Sommerfeld number bearing that is lightly loaded and running at high speed [1].
Bearings with Sommerfeld numbers that range over 1.0 definitely fall into this category. As an exam-
ple, the Sommerfeld number for a high-speed, lightly loaded bearing is calculated as follows:
μ = 2.0 × 10−6 lbf s/in., 1.38 × 10−8 N s/mm2 (average oil viscosity)
N = 10,000 rpm (rotor speed)
L = 2.5 in., 63.5 mm (bearing axial length)
D = 5.0 in., 127 mm (journal diameter)
W = 625 lbf, 2780 N (bearing load)
Cd = 0.008 in., 0.203 mm (bearing diametral clearance)
S = 2.6
The specific case of a vertically downward gravity load is illustrated in Figure 44.1 for a two-axial
groove bearing. This sleeve bearing supports the vertically downward load with a displacement that is
not directly downward but at some angle with rotation from bottom dead center that typically ranges
from 0° to 90°. This angle is defined as the attitude angle, ψ, as shown in Figure 44.1.
Figure 44.2 illustrates the hydrodynamic pressure distribution for a high-speed, lightly loaded,
unstable journal bearing. An example would be the S = 2.6 bearing described earlier. Note that the bear-
ing eccentricity ratio, ε = e/c (Figure 44.1), is very small, typically less than 0.3% or 30% of the bearing
clearance, and the attitude angle, ψ, approaches 90°. In this manner, a light −Y direction load is sup-
ported by a +X displacement. This occurs since the resulting pressure profile supporting the light load
becomes very small with very little change from the maximum film to the minimum film locations. For
equilibrium, the summation of all vertical components of the hydrodynamic forces times the area must
be equal and opposite to the external load, W. Likewise, the sum of all horizontal forces must be zero.
This can only occur for attitude angles that approach 90°.
Since a downward vertical load is supported by a horizontal displacement, any downward force pertur-
bation will result in a horizontal displacement which results in a horizontal force which in turn produces
a vertical displacement, etc. Thus, the bearing generates unstable cross-coupling forces that actually drive
the rotor and cause it to vibrate at a frequency that is normally in the range of 40%–50% of running speed.

Y
Eccentricity ratio
ε = e/c, 0 ≤ ε ≤ 1.0

Attitude angle, Ψ

R+c R

Ob D
X
ω e Oj

Ψ
W

Oil at
viscosity µ

FIGURE 44.1  Two-axial groove fixed geometry bearing.


44-4 Design for Lubrication and Tribology

Unstable bearing

ω
Ob
Oj
X

Light –Y load Light load/hight speed


supported by hight cross-coupling
+X displacement unstable

FIGURE 44.2  High-speed, lightly loaded, unstable bearing.

Conversely, Figure 44.3 illustrates a relatively low-speed, heavily loaded (i.e., low Sommerfeld num-
ber), stable journal bearing. Note that the bearing eccentricity ratio, ε, is very large, typically greater
than 0.6% or 60% of the bearing clearance, and the attitude angle, ψ, approaches 0°. In this manner,
a heavy −Y direction load is supported by a −Y displacement. This occurs since the resulting pressure
profile supporting the heavy load becomes very large with steep gradients from the maximum film to

Stable bearing
Y Heavy –Y load
supported by
–Y displacement

Ob
ω X
Oj

W
Heavy load/ low speed
low cross-coupling
stable

FIGURE 44.3  Low-speed, heavily loaded, stable bearing.


Journal Bearings 44-5

. 5× 1×
12

Machine natural
11

Shaft whip
Frequency = 3800

10

9
Machine speed (rpm × 1000)

us
no
o
l
whir

hr

nc
6

Sy
Oil

4 3×

2
0 1 2 3 4 5 6 7 8 9 10 11 12
Frequency (events/min × 1000)

FIGURE 44.4  Bearing-induced shaft whip and oil whirl.

the minimum film locations. From a force summation, ΣFx = 0 and ΣFy = W. This can only occur for
attitude angles that approach 0°. Since a downward load is now supported by a vertical displacement,
cross-coupling forces are at a minimum and the bearing is stable.
Sleeve journal bearing–induced oil whirl and shaft whip are illustrated in Figure 44.4. The 1× or
synchronous vibration line is clearly indicated on the plot. Oil whirl, caused by the destabilizing cross-
coupling forces produced by high-speed, lightly loaded sleeve bearings (i.e., high Sommerfeld number),
manifests itself as approximately 50% of running speed frequency (shaft vibrates approximately once
per two shaft revolutions). This can be seen in Figure 44.4 at speeds below about 7500 rpm (below twice
the rotor’s first critical).
Above 7500 rpm, the instability frequency locks onto the rotor’s first fundamental natural frequency,
which is about 3800 cpm. This reexcitation of the rotor’s first natural frequency is sleeve bearing–induced
shaft whip which shows up as a vibration component that is below 50% of running speed and occurs at
speeds that are above twice the rotor’s first critical [2].
Usually, sleeve bearings are designed not to go unstable until the rotor speed exceeds twice the rotor’s
first critical speed. Thus, an approximate 0.5× is a rare occurrence and bearing-induced instabilities
usually show up as shaft whip at frequencies less than 50% of synchronous speed.
Axial groove bearings have a cylindrical bore with typically two to four axial oil feed grooves. These
bearings have been very popular in relatively low-speed turbomachinery. For a given bearing load mag-
nitude and orientation, the stability characteristics of axial groove bearings are primarily controlled by
the bearing clearance. Tight clearances produce higher instability thresholds, but tight clearance bear-
ings present other problems that make them undesirable. For example, as clearance decreases, the bear-
ing’s operating oil film temperature increases. Furthermore, Babbitt wear during repeated start-ups will
increase the bearing’s clearance, thereby degrading stability. In fact, many bearing-induced instabilities
44-6 Design for Lubrication and Tribology

in the field are caused by bearing clearances that have increased due to wear from oil contamination,
repeated starts, or slow rolling with boundary lubrication.
Because of these limitations, other fixed-bore bearing designs have evolved to counteract some of the
poor stability characteristics of axial groove bearings. Some antiwhirl sleeve bearing examples include
pressure dam bearings [3], offset half bearings [4], elliptical bearings [5], and multilobe bearings [6,7].
These bearing designs have been successful in increasing the instability threshold speed compared to
axial groove bearings. Pressure dam, elliptical, and offset half bearings remain very popular with gear-
box bearing designers and are still being used today in some lower speed applications. They are also
often found in motor/generator applications.

44.3  Tilting Pad Journal Bearings


Even though they are costlier and more complex than fixed geometry bearings, tilting pad bearings
have gained popularity with turbomachinery designers because of their superior stability performance.
Unlike fixed geometry bearings, tilt pad bearings generate very little destabilizing cross-coupled stiff-
ness regardless of geometry, speed, load, or operating eccentricity [8]. However, rotors supported on
tilting pad bearings are still susceptible to instabilities due to other components within the machine
such as labyrinth seals on centrifugal compressors.
A schematic of a five-pad tilting pad bearing loaded between pivots is illustrated in Figure 44.5 [9,10].
Note that the journal center, Oj, is directly below the bearing center, Ob, and the downward load is sup-
ported by a vertically downward displacement. Because of the pad’s ability to tilt, the attitude angle is
zero, and thus, the cross-coupled destabilizing forces are zero.
Figure 44.6 shows a typical shaft centerline plot for a tilting pad bearing during an actual run-up to
high speed. At very low speed, the journal is just starting to lift off of the bottom of the bearing. This
point is located at the very bottom of the plot. As speed increases, the journal center rises in the bearing

Pivot

ω
X
+Ob
+Oj
cb
R+

φp W
Load
between
pivots
χ

Offset, α= φp/χ
α = 0.5 (centrally pivoted)

FIGURE 44.5  Tilting pad bearing schematic.


Journal Bearings 44-7

User invalidated data High speed

Low speed

FIGURE 44.6  Tilting pad bearing shaft centerline position.

in almost a straight upward trajectory with only a slight bend to the left. The maximum speed point is
at the top of the plot. The attitude angle is very small, about 7° with rotation from the vertical. Thus, the
cross coupling produced by the bearing is essentially zero.

44.3.1  Pivot Loading


Another advantage of tilting pad bearings is the many design parameters that are available for optimiza-
tion. Two examples include the number of pads and pad pivot loading. A typical tilting pad bearing is
shown in Figure 44.7 with five tilting pads loaded between pivots. Between-pivot loading provides more
symmetric stiffness and damping coefficients compared to on-pivot loading. Furthermore, a four-pad
tilting pad bearing loaded between pivots produces stiffness and damping properties that are equal in
the horizontal and the vertical directions [11]. Symmetric support properties provide circular orbits,
whereas asymmetric supports produce highly elliptical orbits. Circular orbits are preferable since their
vibration amplitudes are smaller going through the rotor critical compared to the major axis of an
equivalent elliptical orbit.
For this reason, steam turbine bearings are almost always four-pad bearings loaded between pivots
for superior unbalance response attenuation. Conversely, centrifugal compressor bearings are predomi-
nately five-pad bearings loaded on pivot which produce asymmetric stiffness and damping properties
that promote stable operation.
The dynamic symmetry of a load-between-pivot four-pad tilting pad bearing is also useful in moving
interfering critical speeds out of the operating speed range. Several examples are discussed in Reference
11 for axial compressors originally designed to operate on dynamically asymmetric fixed geometry
axial groove bearings. The axial groove bearing’s asymmetric stiffness and damping properties often
result in split first critical speeds. Since the horizontal stiffness and damping are much softer than the
vertical, a horizontal first critical speed may appear several hundred to several thousand revolutions
per minute lower than the vertical first critical. Conversely, the load-between-pivot four-pad bearing’s
symmetric properties produce a single first critical peak that is approximately midway between the axial
groove bearing’s split peaks.
Moving critical speeds out of the operating range may also be accomplished by switching from tilting
pad bearings to fixed geometry bearings [12]. Reference 12 discusses two case histories for an induction
44-8 Design for Lubrication and Tribology

FIGURE 44.7  Five-pad, load-between-pivot tilting pad journal bearing.

motor and a steam turbine. In both cases, the original configuration with tilting pad bearings produced
a single first critical speed peak within the operating speed range. Switching to a fixed geometry bearing
splits the first critical, resulting in the lower peak below operating speed, and the upper critical above
operating speed.

44.3.2  Pad Pivot Offset


Pad pivot offset is another parameter available to a tilting pad bearing designer. Referring to Figure 44.5,
pad pivot offset is defined as

a = fp /c (44.2)

The pad pivot is offset when it is moved in the direction of rotation by some angle from the centered
position. Centrally pivoted pads have always been more popular with turbomachinery designers, but
recently, more bearings are designed with offset pivots, typically around 60%. Offset pivots increase load
capacity and thereby reduce bearing operating temperatures. Stiffness and damping are also increased
as pad pivot offset increases.
For example, for a five-pad tilting pad bearing with a pad arc length of 60°, if the pivot is located 30°
from the pad’s leading edge, the offset is 30/60 = 0.5% or 50% which is centrally pivoted. If the pivot is
located 36° from the pad’s leading edge, the offset is 36/60 = 0.6% or a 60% offset pivot.

44.3.3  Pad Preload


Another important tilting pad bearing parameter available to the bearing designer is pad preload.
Referring to Figure 44.8, tilting pad bearing preload is defined as

⎛C ⎞
m =1− ⎜ b ⎟ (44.3)
⎝ Cp ⎠

Journal Bearings 44-9

Journal
Op

cp – cb

Oj
Rp
R
Rb
Tilting
pad
cb

Rb = R + cb Preload, m = 1 – cb/cp
Rp = R + cp
Typical m = 0.2 – 0.6
(20% – 60%)
cb = Assembled bearing
clearance
cp = Pad clearance

FIGURE 44.8  Preloaded pad.

The pad clearance is twice the pad’s radius of curvature minus the journal diameter. The bearing
clearance is the diameter of the largest cylinder that will fit into the assembled bearing minus the
journal diameter. If the pad preload is zero, then the pad clearance and the bearing clearance are
equal. For preloaded pads, the pad clearance is greater than the bearing clearance. A preloaded pad
is illustrated in Figure 44.8. Typical preload values range from 0.2 to 0.6. When a pad is preloaded,
a converging film section always exists and the pad will produce a hydrodynamic force even if the
bearing load approaches zero.
For example, for a bearing diametral clearance of 0.006 in., 0.152 mm, and a pad diametral clearance
of 0.012 in., 0.305 mm, the pad preload is 1 − (0.006/0.012) = 1 − (0.152/0.305) = 0.5% or 50% preload.
One advantage of light pad preload is that, in many cases, bearing damping tends to increase while
bearing stiffness remains approximately constant. Both of these trends help to increase bearing effective
damping (the amount of bearing damping that is effective in rotor vibration suppression) and decrease
rotor vibration levels. One disadvantage to decreased pad preload is that the top pads become unloaded
and may be susceptible to pad flutter. Fluttering pads may cause Babbitt damage near the pad’s leading
edge corners.
If the pad preload is known, the eccentricity ratio above which the top pads become unloaded can be
calculated by [9]

m
e= (44.4)
(1 − m) ⋅ sin f

Conversely, if the journal eccentricity is known, then the pad preload below which the top pads become
unloaded can be calculated by [9]

(esin f)
m= (44.5)
(1 + esin f)

44-10 Design for Lubrication and Tribology

The angle ϕ is measured with rotation (counterclockwise, Figure 44.5) from the +X-axis to the
unloaded pad pivot. For example, for a five-pad bearing with between-pivot loading as in Figure 44.5,
ϕ = 18°. For a five-pad bearing with on-pivot loading, ϕ = 54°. For a four-pad bearing with between-
pivot loading, ϕ = 45°.
As an example, calculate the amount of pad preload necessary to insure that the top pads are loaded
for a four and a five-pad tilting pad bearing, both loaded between pivots, with an operating eccentricity
ratio of 0.5:
For five pads loaded between pivots, ϕ = 18°. From Equation 44.5, m = 0.13.
For four pads loaded between pivots, ϕ = 45°. From Equation 44.5, m = 0.26.
Thus, for the five-pad bearing, the pad preload must be >0.13 for the top pads to be loaded at an oper-
ating eccentricity of 0.5. More preload is necessary for the four-pad bearing with m >0.26 to insure top
pad loading.
For another example, calculate the eccentricity above which the top pads become unloaded for a
five-pad bearing loaded on pivot with a pad preload of 0.3. From Equation 44.4 with m = 0.3 and ϕ = 54°,
ε = 0.53. Thus, for this bearing, the top pads become unloaded and susceptible to flutter for ε > 0.53.
Unloaded pad flutter may be minimized by chamfering the unloaded pad’s leading edge and/or
increasing the pad preload on the unloaded pads above and beyond the loaded pad preload.

44.3.4  Pad Length-to-Diameter Ratio


Another design parameter available to the tilting pad bearing designer is pad axial length. In many
applications, increasing pad axial length increases effective damping by increasing bearing damping
and decreasing bearing stiffness [11]. For this reason, tilting pad bearing L/D ratios have increased in
recent years for many applications from 0.5 to 0.75.

44.3.5  Frequency Dependency


The tilting pad’s ability to tilt adds np more degrees of freedom compared to a fixed geometry bearing
where np = number of pads. These additional degrees of freedom create additional stiffness and damping
coefficients referred to as “full coefficients” or “pad coefficients.” The number of coefficients for a fixed
geometry bearing is eight, while the number of coefficients for a tilting pad bearings is 2(5np + 4). Thus,
for a five-pad tilting pad bearing, the number of coefficients is 58.
For computational efficiency and physical interpretation, the full coefficients may be reduced to
the familiar eight coefficients by assuming a pad frequency [8]. These eight coefficients are referred
to as the reduced coefficients [13–17]. As the pad frequency decreases from synchronous, tilting pad
bearing stiffness increases and damping decreases [13]. This frequency dependency effect is ampli-
fied by low preloaded bearings operating in the high Sommerfeld number region, i.e., high speed,
light load [13].

44.4  Tilting Pad Pivots


One important and often overlooked feature of a tilting pad bearing is the pad pivot. Pivot wear during
operation leads to an increase in the bearing clearance with a subsequent decrease in bearing damping.
This results in increased rotor vibration during operation [18]. Pivot stiffness is also a consideration.
The pivot flexibility is in series with the stiffness and damping produced by the bearing’s oil film. As the
pivot stiffness decreases, the effective stiffness and effective damping produced by the bearing decrease
[18]. Equations for calculating the pivot stiffness may be found in References 18 and 19. The equations
for equivalent stiffness and damping are derived in Reference 18.
Journal Bearings 44-11

The two most popular types of tilting pad pivots are cylindrical or line contact pivots and spherical or
ball-in-socket pivots. The cylindrical pivot does not provide a misalignment capability. Consequently,
pads with cylindrical pivots may experience edge loading, resulting in decreased bearing damping.
Additionally, the pivots may wear during operation, causing the bearing clearance to increase, again
resulting in decreased bearing damping. Typical cylindrical pivot wear is shown in Figure 44.9.
Spherical pivots provide a misalignment capability, thus eliminating edge loading. If designed cor-
rectly, the pivot stiffness is very high, resulting in high effective damping, and the pivot stresses are
relatively low, greatly reducing pivot wear during operation [18]. A typical spherical pivot is shown in
Figure 44.10.

FIGURE 44.9  Cylindrical pivot wear on pads and housing.

FIGURE 44.10  Spherical pivot.


44-12 Design for Lubrication and Tribology

44.5  Tilting Pad Bearing Lubricating Flow


Lubricating flow calculations for a tilting pad bearing are relatively simple since, unlike fixed geometry
bearings, the flow is essentially speed and eccentricity independent. The flow depends only on inlet ori-
fice geometry, discharge orifice geometry, bearing end seal clearances, and, to a lesser extent, lubricant
viscosity.
Assuming that oil is the lubricant and assuming typical values for oil density plus a typical value of
0.61 for the flow coefficient through a short tube, the oil flow through an inlet orifice hole in U.S. cus-
tomary units is [9]

Q = 25A ΔP (44.6)

For n number of inlet holes,

Q = (19.64)(nd 2 ) ΔP (44.7)

where

ΔP = Pin − Ph (44.8)

For a tilting pad bearing with an evacuated housing, the housing pressure, Ph, is zero gage, and
Equations 44.6 and 44.7 may be used to calculate the oil flow through a tilting pad bearing with ΔP = Pin,
the oil inlet pressure.
For a tilting pad bearing with a flooded housing, Equations 44.6 through 44.8 may be used to cal-
culate the oil flow through a tilting pad bearing. However, in this case, Ph, the housing pressure, is
unknown, making the solution iterative.
As an example, calculate the oil flow for the following 5 in. tilting pad bearing:
Pin = 20 psig
Five inlet holes with a diameter of 0.125 in.
Case 1—Evacuated without end seals
Case 2—Flooded with end seal diametral clearances of 0.006 in.
Case 1: Using Equation 44.7 with ΔP = Pin = 20 psig,
Q = 6.9 gpm.
Case 2: Assume that Ph = 5 psig.
Thus, the inlet ΔP = 15 psig.
From Equation 44.7, the inlet Q = 5.9 gpm.
The discharge area with two end seals is A = 2(π/4)(5.0062 − 5.02) = 0.094 in2.
From Equation 44.6, the discharge ΔP = 5 psig and the discharge Q = 5.3 gpm.
Since the inlet and discharge Q are nearly equal, we have a solution of an oil flow of 5.3­–5.9 gpm.
The solution may be refined by slightly increasing Ph and recalculating.

44.6  Advanced Tilting Pad Journal Bearings


In recent years, the focus of tilting pad bearing designers has been on reduced temperature designs. This
has been driven partly by the stringent rotordynamics specifications that turbomachinery manufactur-
ers must adhere to in order to sell their rotating equipment. Large journal diameters result in less severe
critical speeds. However, larger diameter journals result in high surface velocities, often exceeding
Journal Bearings 44-13

90 m/s (300 f/s). Furthermore, for increased efficiencies, compressors and turbines are required to run
at even higher speeds, again increasing surface velocity. At these high surface speeds, special care must
be taken to cool the bearing.
One way to reduce operating temperatures at very high speeds is to increase bearing clearance above
and beyond the standard clearance range of 1.5–2.0 mils (μm) of diametral bearing clearance per inch
(mm) of journal diameter. Open clearances in the range of 3.0–4.0 mils/in. (μm/mm) have been used
strictly for reduced operating temperatures. Unfortunately, at these high clearance values, bearing
damping often decreases, resulting in higher vibration levels.
Another way to reduce operating temperatures at very high speeds is to increase the bearing oil flow.
This, however, often produces only marginal decreases in bearing operating temperature due to hot oil
carryover and inlet oil mixing. These two phenomena are illustrated in Figure 44.11 [20]. For a con-
ventional tilting pad bearing, a substantial percentage of hot oil is carried over by the journal from the
trailing edge of the upstream pad into the leading edge of the next downstream pad. This is called hot oil
carryover. The amount of carryover is unknown, but a good estimation used by many bearing designers
is 60% hot oil carryover.
Additionally, the cool inlet oil mixes with the discarded preceding pad’s hot oil before it enters the
downstream pad’s leading edge. This is called inlet oil mixing. Thus, for this conventional design, oil
enters the pad’s leading edge at a substantially higher temperature compared to the oil inlet temperature
(i.e., the temperature of the oil as it exits the oil cooler).
An example of the diminishing effects of increasing oil flow on operating temperature is illustrated
in Figure 44.12. This example is for a typical four-pad tilting pad gearbox pinion bearing whose journal
diameter is 7.0 in., 178 mm, and is operating at 305 f/s, 93 m/s, with a unit load of 452 psi, 3.1 MPa. For
a relatively low flow rate of 75 lpm, the Babbitt temperature is 120°C. Normal temperature alarms are
set at 110°C with normal trip levels set at 120°C. Initially, increasing flow decreases temperature sub-
stantially until about 150 lpm. Flow rates above 150 lpm produce only a marginal effect on Tmax for the
reasons discussed earlier.

Hot oil
carry-over Warm oil

Mixing

Cool inlet oil

Hot oil

FIGURE 44.11  Pad-to-pad hot oil carryover and hot-cool oil mixing.
44-14 Design for Lubrication and Tribology

178 mm Four pad tilting pad pinion bearing — Q vs. Tmax

120
Maximum Babbitt temperature, Tmax (°C)

115

110

105

100
50 75 100 125 150 175 200 225 250 275 300
Oil flow, Q (Lpm)

FIGURE 44.12  Tilting pad pinion bearing operating temperature vs. oil flow.

A very effective way to reduce operating temperatures for high-speed operation is to utilize the cool
inlet oil more effectively by forcing it directly into the pad’s leading edge with special spray bar blockers
[20], leading edge feed grooves [21], or inlet oil feed slots [22,23]. With these flow configurations, hot oil
carryover may be significantly reduced and inlet oil mixing essentially eliminated.
In 1991, Tanaka [24] published results showing the cooling effect of removing the tilting pad bear-
ing’s end seals leading to an evacuated bearing housing design. The oil between the pads in a conven-
tional flooded housing with end seals produces additional heat and power loss due to the entrapped
oil “churning” within the housing. This churning is eliminated with an evacuated housing, thereby
decreasing operating temperature and power loss. However, since all of the oil that is required to enter
the pad’s leading edge must be provided by the inlet oil, more inlet oil must be supplied compared to a
flooded housing design.
Other important design features that further reduce operating temperatures include pad pivot offset
[25], babbitted copper pads, and, in some specialized designs, behind-the-pad bypass cooling [20].
As an example, Figure 44.13 shows the effectiveness of an evacuated babbitted steel pad bearing com-
pared to the evacuated babbitted copper pad bearing with spray bar blockers and bypass cooling shown
in Figure 44.14. These test data were from actual load testing on a gearbox 200 mm, 7.87 in., tilting pad
pinion bearing. Note that the surface velocity is extremely high at Us = 112 m/s (369 f/s) with a bearing
unit load of Lu = 2.1 MPa (298 psi). Inlet oil temperature is plotted against the embedded temperature
sensor readings. The bearing with copper pads reduced operating temperatures by 14°C–18°C depend-
ing on the particular sensor. Data from only one sensor were available for the steel pad bearing, while
data from four sensors were available for the copper pad bearing.

44.7  Conclusions
Oil flow, operating temperature, and power loss are all important considerations for bearing design-
ers. The reduced temperature features discussed herein will greatly aid the bearing designer. However,
the main purpose of the journal bearings is to suppress rotor vibration. Again, the geometric features
discussed herein will aid bearing designers in adopting journal bearings with high effective damping
for rotor vibration suppression. Care must be taken to analyze the entire rotor-bearing system prior to
Journal Bearings 44-15

200 mm tilting pad journal bearing


120
10,580 rpm, 369 f/s, 112 m/s
Unit Load, 298 psi, 2.1 MPa

Temperature sensor (°C)


110 Babbitted steel pads

100

Babbitted copper pads


90

80
45 50 55 60 65 70
Oil inlet temperature (°C)

FIGURE 44.13  Babbitted steel pads vs. babbitted copper pads, tilting pad journal bearing test stand data.

FIGURE 44.14  Advanced tilting pad bearing design with spray bar blockers and bypass cooling. (From Nicholas,
J.C., Tilting pad journal bearings with spray-bar blockers and by-pass cooling for high speed, high load applica-
tions, Proceedings of the Thirty-Second Turbomachinery Symposium, Texas A&M University, College Station, TX,
p. 27, 2003. With permission.)

finalizing the bearing design. That is, a rotordynamics analysis of the rotor with the optimized bearings
must be undertaken to verify the validity of the bearing optimization. Both a forced response analysis
and a stability analysis should be considered.

References
1. Nicholas, J. C., Hydrodynamic journal bearings—Types, characteristics and applications, mini
course notes, 20th Annual Meeting, The Vibration Institute, Willowbrook, IL, p. 79, 1996.
2. Newkirk, B. L., Taylor, H. D., Shaft whipping due to oil action in journal bearing, General Electric
Review, 28, 559, 1925.
3. Nicholas, J. C., Allaire, P. E., Analysis of step journal bearings—Finite length, stability, ASLE
Transactions, 23, 197, 1980.
44-16 Design for Lubrication and Tribology

4. Mehta, N. P., Singh, A., Stability analysis of finite offset-halves pressure dam bearings, ASME Journal
of Tribology, 108, 270, 1986.
5. Mehta, N. P., Singh, A., Gupta, B. K., Stability of finite elliptical pressure dam bearings with rotor flex-
ibility effects, ASLE Transactions, 24, 269, 1981.
6. Flack, R. D., Lanes, R. F., Effects of three-lobe bearing geometries on rigid-rotor stability, ASLE
Transactions, 25, 221, 1982.
7. Lanes, R. F., Flack, R. D., Effects of three-lobe bearing geometries on flexible rotor stability, ASLE
Transactions, 25, 377, 1982.
8. Lund, J. W., Spring and damping coefficients for the tilting-pad journal bearing, ASLE Transactions,
7, 342, 1964.
9. Nicholas, J. C., Tilting pad bearing design, Proceedings of the Twenty-Third Turbomachinery
Symposium, The Turbomachinery Laboratory, Texas A&M University, College Station, TX, p. 179,
1994.
10. Nicholas, J. C., Gunter, E. J., Allaire, P. E., Stiffness and damping coefficients for the five pad tilting pad
bearing, ASLE Transactions, 22, 112, 1979.
11. Nicholas, J. C., Kirk R. G., Four pad tilting pad bearing design and application for multi-stage axial
compressors, ASME Journal of Lubrication Technology, 104, 523, 1982.
12. Nicholas, J. C., Moll, R. W., Shifting critical speeds out of the operating range by changing from tilting
pad to sleeve bearings, Proceedings of the Twenty-Second Turbomachinery Symposium, Texas A&M
University, College Station, TX, p. 25, 1993.
13. Parsell, J. K., Allaire, P. E., Barrett, L. E., Frequency effects in tilting-pad journal bearing dynamic
coefficients, ASLE Transactions, 26, 222, 1983.
14. Barrett, L. E., Allaire, P. E., Wilson, B. W., The eigenvalue dependence of reduced tilting pad bearing
stiffness and damping coefficients, ASLE Transactions, 31, 411, 1988.
15. Ha, H. C., Yang, S. H., Excitation frequency effects on the stiffness and damping coefficients of a
five-pad tilting pad journal bearing, ASME Journal of Tribology, 121, 517, 1999.
16. Brockett, T. S., Barrett, L. E., Exact dynamic reduction of tilting-pad bearing models for stability
analyses, STLE Tribology Transactions, 36, 581, 1993.
17. Cloud, C.H., Stability of rotors supported by tilting pad journal bearings, PhD dissertation,
University of Virginia, Charlottesville, VA, 2007.
18. Nicholas, J. C., Wygant, K. D., Tilting pad journal bearing pivot design for high load applications,
Proceedings of the Twenty-Fourth Turbomachinery Symposium, Texas A&M University, College
Station, TX, p. 33, 1995.
19. Kirk, R. G., Reedy, S. W., Evaluation of pivot stiffness for typical tilting-pad journal bearing designs,
ASME Journal of Vibration, Acoustics, Stress and Reliability in Design, 110, 165, 1988.
20. Nicholas, J. C., Tilting pad journal bearings with spray-bar blockers and by-pass cooling for high
speed, high load applications, Proceedings of the Thirty-Second Turbomachinery Symposium, Texas
A&M University, College Station, TX, p. 27, 2003.
21. Dmochowski, W., Brockwell, K., DeCamillo, S., Mikula, A., A study of the thermal characteristics of
the leading edge groove and conventional tilting pad journal bearings, ASME Journal of Tribology,
115, 219, 1993.
22. Gardner, W. W., Tilting pad journal bearing using directed lubrication, U.S. Patent No. 5,288,153,
Waukesha Bearing Corporation, Waukesha, WI, 1994.
23. Ball, J. H., Byrne, T. R., Tilting pad hydrodynamic bearing for rotating machinery, U.S. Patent
No. 5,795,076, Orion Corporation, Grafton, WI, 1998.
24. Tanaka, M., Thermohydrodynamic performance of a tilting pad journal bearing with spot lubrica-
tion, ASME Journal of Tribology, 113, 615, 1991.
25. Simmons, J. E. L., Lawrence, C. D., Performance experiments with a 200 mm, offset pivot journal pad
bearing, STLE Tribology Transactions, 39, 969, 1996.
45
Thrust Bearings
45.1 Introduction.....................................................................................45-1
45.2 Thrust Bearing Types......................................................................45-3
45.3 Alignment and Equalization.........................................................45-3
45.4 Design Criteria.................................................................................45-4
45.5 Low-Speed Performance Equations..............................................45-6
Slack-Side Performance  •  Guidelines for Performance Equations
45.6 Industrial Applications................................................................. 45-11
Low-Speed Industrial Applications  •  High-Speed Industrial
Applications  •  Temperature Limitations  •  Temperature
Charts  •  High-Speed/Low-Load Symptoms
Scan M. DeCamillo 45.7 Direct Lube Bearings.....................................................................45-19
Kingsbury, Inc.
45.8 Hydroelectric Applications..........................................................45-19
Bruce R. Fabijonas Acknowledgments.....................................................................................45-20
Kingsbury, Inc. References...................................................................................................45-20

45.1  Introduction
Hydrodynamic thrust bearings hold the axial position of rotating shafts, transmitting the axial force
to the stationary casing by means of a thin film of fluid (Figure 45.1). The bearings are used in a wide
variety of rotating machinery including turbines, compressors, pumps, motors, generators, gear boxes,
and many custom applications. Bearing sizes range from 5.2 m in diameter, which can carry loads on the
order of 50 million newtons, down to 50 mm and smaller diameter bearings that support rotors spin-
ning tens of thousands of revolutions per minute.
The principle of hydrodynamic operation was presented in 1886 by Osborne Reynolds in his theoreti-
cal paper on oil-lubricated journal bearings.1 Reynolds concluded that oil, because of its adhesion to the
journal, is dragged between the journal and the journal bearing, forming a wedge-shaped film which
develops a pressure distribution that supports the load (Figure 45.2). This wedge-shaped film is essential
for high load capacity.
Reynolds’ theory also applies to hydrodynamic thrust bearings, although the wedge shape, inherently
formed by the journal bearing geometry, has to be machined into the flat surface of a thrust bearing.
A different method was demonstrated in 1898 by Albert Kingsbury2 who conceived that an oil wedge
would automatically take shape if individual “shoes” were supported on pivots (Figure 45.3). Pivoted
shoe thrust bearings are often referred to as Kingsbury bearings or Michell bearings3 based on patents
established in the early 1900s.
The separation of the rotating and stationary surfaces by a thin film of fluid results in very low
friction. The coefficient of friction in operation is one to two orders of magnitude lower than the
sliding coefficient of friction of the materials. The separation of surfaces also results in negligible
wear, giving the hydrodynamic bearing an exceptionally long life, in many cases equivalent to the
life of the machine. A high degree of precision is required for surface flatness, surface finish, and

45-1
45-2 Design for Lubrication and Tribology

Stationary
casing

Rotating
shaft

Axial force

Shaft collar

Hydrodynamic
thrust bearing

Fluid film

FIGURE 45.1  Hydrodynamic thrust bearing schematic.

Oil adheres
to joumal

Is dragged Rotation
between
surfaces
+
Load
Forms a
wedge-
shaped film

Develops a
pressure
distribution

FIGURE 45.2  Illustration of Reynolds’ hydrodynamic principle.

“ Collar”
Motion
Oil

Pivot Bearing block


or “shoe”
Resultant
Center

(stationary)
Pmax

FIGURE 45.3  Albert Kingsbury’s sketch of the pivoted shoe.


Thrust Bearings 45-3

manufacturing tolerances to accommodate the thin fluid films which, depending on the application,
are on the order of 5–50 μm.

45.2  Thrust Bearing Types


A flat plate is the simplest fixed geometry thrust bearing, the basic shape being that of a washer. The
segmented plate in Figure 45.4a is a common configuration, with radial grooves dividing the surface
into segments. These grooves improve performance by facilitating the entrance of fluid between the sur-
faces. Flat-faced bearings have minimal load capacity and relatively high friction. They are mostly used
as positioning bumpers to handle light and occasional axial loads.
Taper land thrust bearings (Figure 45.4b) are segmented plates with a taper machined on the surface
of each segment to form the wedge-shaped film geometry necessary for high load capacity. The taper
needs to be on the order of magnitude of the film thickness to be effective and is typically sized to
provide optimum performance at the machine’s normal running condition. It is important to note that
the tapers are directional and so have effective load carrying capability in only one direction of rota-
tion. The flat portion of the surface following the taper provides support for the rotor during start-up
and shutdown.
Pivoted shoe thrust bearings (Figure 45.4c) incorporate individual segments called shoes or pads.
Each shoe is supported on a pivot which allows the shoe to tilt and form a wedge-shaped film. Unlike
the fixed geometry of the taper land bearing, the pivoted shoes adjust to changes in operating conditions
and provide optimum capacity over a broad range of loads and speeds. Operation in either direction of
rotation is possible when the pivot is centrally positioned.
Three common pivot designs are shown in Figure 45.5. A radial rib on the back of the pad creates
a line contact pivot that allows the pad to tilt in the direction of rotation (Figure 45.5a). Pivots with a
spherical contact surface (Figure 45.5b), and ball and socket pivot designs (Figure 45.5c), have the capa-
bility to pivot tangential and radial to the direction of rotation.

45.3  Alignment and Equalization


Proper casing and bearing manufacturing tolerances are necessary for the bearings to develop maxi-
mum load capacity. Basically, variations between the bearing surfaces and collar face must be less than

(a) (b) (c)

FIGURE 45.4  Basic types of hydrodynamic thrust bearings and corresponding film thickness profiles for
(a) segmented plate, (b) tapered land, and (c) pivoted shoe bearings.
45-4 Design for Lubrication and Tribology

(a) (b) (c)

FIGURE 45.5  Three common thrust bearing designs: (a) line contact, (b) spherical contact, and (c) ball and
socket.

Equalized

Edge Radial
load tilt

(a) (b) (c)

FIGURE 45.6  Misalignment, edge load and equalizing schematic for (a) tapered land and line contact pivots,
(b) spherical pivoted shoes, and (c) equalizing feature.

the film thickness in order for all of the bearing segments to support a fairly equal portion of the applied
load. If pad height variations are too great, or in the case of misalignment illustrated in Figure 45.6, the
few pads closest to the collar face will be overloaded.
The fixed geometry of the taper land bearing is particularly sensitive to misalignment and susceptible
to damage from edge load at the outer diameter (Figure 45.6a). Alignments better than 0.0001 mm/mm
across the diameter are required for maximum load capacity, which is unpractical. In practice, taper
land load ratings are reduced to account for misalignment, and it is common to incorporate a feature to
adjust for alignment during assembly in larger machines. Pad tilt and pivot deflection allow pivoted shoe
bearings to tolerate more misalignment and accommodate higher loads. The line contact pivot remains
susceptible to edge load where the radial tilt capability of the spherical pivot eliminates edge load issues
(Figure 45.6b).
Shoes for nonequalizing pivoted shoe bearings are often provided as a set with controlled heights.
Equalizing pivoted shoe bearings incorporate a mechanism that equalizes the load among shoes (Figure
45.6c). This improves bearing load capacity, adjusts for misalignment, and relaxes bearing and casing
manufacturing tolerances. The pivoted shoe thrust bearing in Figure 45.7 uses a system of levers called
leveling plates that accommodate misalignments on the order of 0.002 mm/mm.
Film profile, misalignment, and equalization considerations are taken into account in Table 45.1,
which provides a comparison of bearing types under similar application conditions. The unit load rep-
resents the axial load divided by the bearing surface area. Load capacity also depends on other param-
eters, which are covered in the following sections.

45.4  Design Criteria


Hydrodynamic thrust bearings are made from a variety of materials which, depending on the applica-
tion, may include metals, carbon graphite, polymers, and ceramics. Bearings have been designed for
applications using fluids such as air, mercury, water, liquid refrigerants, alcohol, gasoline, hydraulic
Thrust Bearings 45-5

Thrust
shoes

Leveling
plates
Base
ring

FIGURE 45.7  Kingsbury equalizing thrust bearing.

TABLE 45.1  Example Guidelines for Babbitted Steel Bearings in Oil


Hydrodynamic Thrust Bearing Type Maximum Unit Load (MPa)
Flat plate 0.20
Segmented plate 0.52
Taper land (one direction of rotation) 1.7
Nonequalizing pivoted shoe 2.8
Equalizing pivoted shoe 3.5

fluid, mineral oil, and synthetic lubricants. Higher viscosity fluids generate thicker films and can accom-
modate higher loads. Capacity also depends on the surface speed, the bearing type and geometry, and
the strength and integrity of the materials and fluid. While load capacities vary dramatically over this
extensive range of applications, the criteria governing load capacity are fairly straightforward.
Bearing design or selection typically begins by establishing the bearing inner diameter, which must
fit around the shaft that passes through the bearing. The outer diameter is determined by the applica-
tion’s thrust load and desired unit pressure on the bearing surface area, which is based on the bear-
ing’s load capacity. Outer diameters range 1.5–2.5 times the inner diameter for most applications.
The number of segments or shoes is based on obtaining a segment length-to-width ratio between 0.7
and 1.3, the optimum ratio being 1.0, allowing for lubricant passages that occupy 10%–15% of the
circumference.
Film thickness is the most important parameter for low-speed applications. There are two criteria
commonly considered. The first is to design for loads that provide a sufficient minimum film thickness
such that high points of the manufactured bearing and collar surfaces do not come in contact during
operation. Minimum film thickness in this case can be on the order of 0.005–0.010 mm. In applications
susceptible to particle contamination, a second consideration is to design for loads that give a minimum
film thickness larger than the filter size. In this way, abrasive particles that pass through the filter can
also pass through the film without scoring and damaging the bearing surfaces. For example, one should
design for a load that gives a minimum film thickness greater than or equal to 0.015 mm for a 15 μm
filter.
As surface speeds increase, load capacity becomes more dependent on the integrity of the bearing
materials and fluid. The hydrodynamic film thickness increases with speed and the bearing can sup-
port more load. At the same time, frictional shear of the fluid in the film increases the operating tem-
peratures. The combination of higher temperature and increased load causes more elastic and thermal
distortion of the bearing surfaces, which limits the amount of load that can be applied. The mechanical
and thermal limitations of the bearing materials or fluid may also be limiting factors.
45-6 Design for Lubrication and Tribology

45.5  Low-Speed Performance Equations


The goal in low-speed design is to size the bearing to support an axial load with a reasonable minimum
film thickness. Film thickness is related to load by Reynolds’ partial differential equation. Although
there is no general analytic solution to the equation, numerical solutions are relatively straightforward
using finite element software and techniques. Approximate formula and empirical factors are used for
purposes of this section to allow a quick, first approximation of bearing performance.
The following equations can be derived from Reynolds’ theory applied to a simple rectangular tapered
wedge of length ℓ and width w with some basic assumptions including constant viscosity, incompressible
fluid, and laminar flow.4 The equations are used to approximate performance of a sector-shaped thrust
bearing segment of similar length and width, as illustrated in Figure 45.8. Frictional power loss is used
by system designers to determine lube system requirements and machine efficiency:
Mean film temperature (°C):

Tm = Tin +
( K t Pavg ) (45.1)

( rCh L )
Minimum film thickness (m):

6muK p L
h0 = (45.2)
Pavg

Bearing power loss (W):

⎛ mu 2 wK f n ⎞
F =⎜
⎜ ⎟⎟ (45.3)
⎝ h0 ⎠

Taper-to-film-thickness ratio:

m=
( h1 − h0 ) (45.4)
h0

Pressure
distribution
Load
X
U
Fluid flow
h0 h1

FIGURE 45.8  Sector approximation based on a rectangular wedge.


Thrust Bearings 45-7

Here
ℓ is the length of segment in the direction of motion (m)
w is the radial width of segment (m)
n is the number of segments
u is the surface speed at the mean diameter (m/s)
μ is the fluid absolute viscosity at Tm (Pa s, N s/m2)
Paνg is the average film pressure (Pa)
ρ is the fluid density (kg/m3)
Ch is the fluid specific heat (J/(kg °C))
h1 is the leading edge film thickness (m)
Tin is the fluid inlet temperature (°C)

L is a “side leakage factor” based on empirical data,5 which accounts for fluid exiting at the outer and
inner diameters. A quadratic fit of the data for ℓ/w ratios between 0.4 and 1.6 is provided here for
convenience:

2
⎛⎞ ⎛⎞ ⎛⎞
L = 1.0091 − 0.7201⎜ ⎟ + 0.1557 ⎜ ⎟ , 0.4 ≤ ⎜ ⎟ ≤ 1.6
w
⎝ ⎠ w
⎝ ⎠ ⎝w⎠

The influence of the wedge-shaped film is contained in the dimensionless integration factors Kp, Kf, and
Kt, which are functions of the taper-to-film-thickness ratio m. For taper land bearings, the term h1 − h0
in Equation 45.4 represents the rise of the taper from the leading to trailing edge of the segment. For
pivoted shoe design, the ideal pivot location is at the center of pressure x‾ which is also a function of m.
The factors are

1 2
Kp = ln (1 + m ) −
m2 m(2 + m)

4 6
Kf = ln (1 + m ) −
m 2+m

⎡ K f (2 + m) ⎤
Kt = ⎢ ⎥
⎢⎣ 12K p (1 + m ) ⎥⎦

⎛ m2 + 6m − ( 4m + 6 ) ln ( m + 1) ⎞
x = ⎜1 − ⎟


⎝ (
2m ( m + 2 ) ln ( m + 1) − 2m ⎟
⎠)
The radial position of the center of pressure of the sector is not available from the derivation, which was
based on a rectangle. The pivot is placed 52%–55% of the pad width, toward the outer diameter, to bal-
ance the sector-shaped area.
The formula for Kp has a maximum value when m = 1.2 which, by Equation 45.2, provides an opti-
mum minimum film thickness. It is therefore desirable to design for this condition in many applica-
tions. The corresponding factors are

m = 1.2, K p = 0.0267, K f = 0.753, K t = 3.419, and x = 0.58



45-8 Design for Lubrication and Tribology

TABLE 45.2  Approximate Equations for Fluid Properties vs. Temperature T (°C)
Density (kg/m3) Specific Heat (J/(kg K)) Viscosity (Pa s, N s/m2)
ρ = A + BT + CT2 Ch = A + BT + CT2 μ = A exp(B/(T + C))
A B C A B C A B C
Water 1001.7 −0.0673 −0.0036 4201 −0.9836 0.0116 1.88E−5 669.67 147.84
ISO VG 32 889.5 −0.6097 0 1836 3.7303 0 5.58E−5 911.03 109.15
ISO VG 68 896.0 −0.6140 0 1821 3.7593 0 4.98E−5 959.79 98.62

Density, specific heat, and viscosity data for fluids are available from many sources. It is useful to have charts
or equations vs. temperature for the calculations. Approximate equations are given in Table 45.2 for conve-
nience. Density and specific heat are curve-fitted data, and a variation of Vogel’s model is used for viscosity.

Example 45.1
Consider a hydrodynamic pivoted shoe thrust bearing that has an outer diameter of 250 mm,
an inner diameter of 125 mm, and six 51° pad segments. Operating conditions are for a load of
50,000 N at 400 rpm. The lubricant is ISO VG 68 oil with an inlet temperature of 50°C.

Solution
The parameters for Equations 45.1 through 45.4 are obtained from the given information and
simple geometric formulas, being mindful of unit conversions:

Number of pads n 6
Mean diameter (m) dm 1
(0.250 + 0.125) = 0.1875
2

Pad arc length (m) ℓ 1


2 (0.1875) × 51 × p / 180 = 0.0834
Pad width (m) w 2 (0.250 − 0.125) = 0.0625
1

Length-to-width ratio ℓ/w 0.0834/0.0625 = 1.33


Bearing area (m2) A 0.0834 × 0.0625 × 6 = 3.13 × 10−2
Average pressure (Pa) Paνg (50,000)/(3.13 × 10−2) = 1.60 × 106
Surface speed (m/s) u π × (0.1875) × 400/60 = 3.93
Inlet temperature (°C) Tin 50

Designing for an optimum film thickness:

m = 1.2, K p = 0.0267, K f = 0.753, K t = 3.419, and x = 0.58



The side leakage factor is L = 1.0091 − 0.7201(1.33)+0.1557(1.33)2 = 0.327.
The mean film temperature in Equation 45.1 can be approximated using the density and
specific heat at the inlet temperature. From Table 45.2 for ISO VG 68,

r = 896.0 − 0.6140 ( 50 ) = 865 kg/m 3 , and Ch = 1821 + 3.7593 ( 50 ) = 2009 J/(kg °C)

Substituting in Equation 45.1,

(
⎛ 3.419 × 1.60 × 106 ⎞
Tm = 50 + ⎜
)
⎟ = 59.6°C
⎜⎝ (865 × 2009 × 0.327 ) ⎟⎠

Thrust Bearings 45-9

Using Tm in the viscosity equation for ISO VG 68 from Table 45.2,

⎛ 959.79 ⎞
( )
m = 4.98 × 10 −5 exp ⎜ ⎟ = 0.0215 N s/m
⎝ (59.6 + 98.62) ⎠
2

Minimum film thickness and frictional power loss are obtained from Equations 45.2 and 45.3:

6 × 0.0215 × 3.93 × (0.0834 ) × 0.0267 × 0.327


h0 = = 0.015 × 10 −3 m = 0.015 mm
1.60 × 106

2
0.0215 × (3.93) × (0.0834 ) × (0.0625) × 0.753 × 6
F= = 521 W
0.015 × 10 −3

This completes the example calculation for the pivoted shoe thrust bearing. The same perfor-
mance would be obtained if the example were for a taper land bearing, in which case the rise of
the taper, h1 − h0, would be calculated. The film thickness at the entrance to the film is obtained
from Equation 45.4:

m = 1. 2 =
(h1 − 0.015)
0.015

from which h1 = 0.033 mm. The rise of the taper is h1 − h0 = 0.033 − 0.015 = 0.018 mm.

45.5.1  Slack-Side Performance


Many applications use a double thrust bearing configuration to accommodate axial loads in either direc-
tion. A common arrangement is shown in Figure 45.9. The two bearings may be the same size, different
sizes, and even different types as required for the machine’s forward and reverse load conditions. The
bearing for the main direction of load is referred to as the main bearing, and the other as the slack-side
bearing. An axial clearance is set between the bearings and collar to allow for misalignment, thermal
expansion, and development of fluid films when the machine is in operation. Filler plates or shims are
used to set the clearance and the axial position of the rotor in the casing during assembly.

Outlet oil

Filler plate

Base ring

Shaft
BC
Collar

Seal ring

Inlet oil

FIGURE 45.9  Double equalizing thrust bearing configuration.


45-10 Design for Lubrication and Tribology

The slack-side power loss can be calculated using Equation 45.3 by considering that when the main
bearing is loaded, the remainder of the axial clearance is the fluid film thickness of the slack-side
bearing.

Example 45.2
Following the conditions of Example 45.1, suppose the there is a double-acting thrust bearing of
equal design and size and that design axial clearance is 0.38 mm:
The loaded-side mean film thickness is (0.033 + 0.015)/2 = 0.024 mm.
The mean film thickness on the slack side is hm = 0.38 − 0.024 = 0.356 mm.
Equation 45.4 can be expressed in terms of mean film thickness, hm, as h0 = 2hm/(m + 2).
The slack-side minimum film thickness is h0 = 2 × 0.356/(1.2 + 2) = 0.223 mm = 0.223 × 10−3 m.
In an unloaded state, the mean film temperature of the slack side is much cooler than the
loaded side, and the viscosity at the inlet temperature can be used to obtain a conservative esti-
mate of the slack-side bearing friction loss. From the viscosity equation from Table 45.2 for ISO
VG 68 at 50°C,

⎛ 959.79 ⎞
( )
m = 4.98 × 10 −5 exp ⎜ ⎟ = 0.0318 N s/m
⎝ (50 + 98.62) ⎠
2

and from Equation 45.3,

F=
( 0.0318 ) × ( 3.93) × ( 0.0834 ) × ( 0.0625) × 0.753 × 6 = 52 W
0.223 × 10−3

The total power loss for the double thrust bearing is the sum of the loaded- and slack-side losses:
521 + 52 = 573 W.

45.5.2  Guidelines for Performance Equations


Equations 45.1 through 45.4 can be used to estimate performance for a very broad range of hydrody-
namic bearing designs, sizes, fluids, and applications. The lower limit of the applicable speed range is
0.05–0.15 m/s, depending on many factors. Below these speeds, the film enters into boundary lubrica-
tion conditions where the equations no longer apply.6 The upper end of the applicable speed range is
approximately 10 m/s for low-viscosity fluids like water and up to 35 m/s for higher viscosity fluids such
as oil. Additional factors start to influence performance at higher speeds, and the equations begin to
underestimate temperature and power loss and overestimate minimum film thickness.
A centrally pivoted thrust shoe provides an advantage for operation in either direction of rotation.
Although the equations calculate a pivot offset, results can be used as a reasonable approximation for
center-pivoted shoe bearings in oil-lubricated applications. Successful operation of center pivot designs
is attributed to small-scale elastic and thermal crown of the shoe surface in high-viscosity-fluid applica-
tions, which results in close to optimum performance.7 This is not the case with low loads and friction
in low-viscosity fluids like water, and it may be necessary to manufacture a precisely crowned surface8
to achieve suitable load capacity for center pivot operation.
Equations in this section are appropriate for taper land and nonequalizing pivoted pad bearings
under conditions of good alignment. The flat portion of the taper land’s segment, illustrated in Figure
45.4b, occupies 10%–20% of the arc length. The total arc length, including the flat portion, can be used
Thrust Bearings 45-11

in the calculations. The film thickness profile in a sector is more complex than the simple rectangular
wedge of the derivation, and steeper than optimum tapers are sometimes used to improve cooling.
References 8 and 9 contain associated design information.
Hydrodynamic equations are based on a wedge-shaped film and are not applicable for flat-faced bear-
ings like the segmented plate in Figure 45.4a. Design and performance are typically based on experi-
ence and empirical methods.9,10 The equations in this section can be used to obtain rough estimates of
segmented plate performance by selecting a very poor taper ratio of m = 0.20. This gives a relatively high
frictional loss typical of experimental results with oil for loads ranging 0.3–1.0 MPa.

45.6  Industrial Applications


Hydrodynamic bearings are common in the turbines, generators, gearboxes, motors, pumps, and
compressors used for power generation, oil and chemical processing, and shipboard applications. The
bearings are mostly taper land and pivoted shoe designs manufactured from steel with a thin layer of
tin-based babbitt, also called white metal, bonded to the bearing surface. Mineral oil or synthetic lubri-
cants are used for lubrication and cooling.
Babbitt is not required for hydrodynamic operation but serves to minimize damage to the collar and
shaft under adverse conditions. The thrust collar is an important component of the bearing. It may be
an integral part of the shaft or a separate component attached by a key and nut or shrink fit. Thrust faces
within 0.012 mm parallel and square with the shaft are customary for most industrial applications, with
surfaces flat within 0.008 mm and finished 0.2–0.4 μm Ra. The thrust bearing babbitt faces are finished
0.4–0.8 μm Ra.
Load capacities for the thrust bearings are provided by the bearing manufacturers. Rated loads rep-
resent the bearing’s capability under continuous operation and are usually provided in tables or plots
as a function of speed. Other terminologies are load capacity, allowable load, and safe load. The terms
maximum load, maximum allowable load, or ultimate load imply a critical limit, in which case the bear-
ings would be designed for continuous operation below the specified values.
The type of application must also be taken into account. Bearings for machines with fairly consistent
operating conditions can be designed close to the rated load. In the case of adverse conditions or critical
applications, an appropriate safety factor should be used such that worse-case conditions do not exceed
the capacity of the bearing. Machine and application requirements are governed by original equipment
manufacturer and industry specifications.

45.6.1  Low-Speed Industrial Applications


Table 45.3 contains rated loads for a commercial range of babbitted steel, equalizing pivoted shoe thrust
bearings for speeds from 100 to 3600 rpm, which is a common range for a wide variety of industrial
applications. The size column represents the outer diameter of the bearing surface in inches, the inner
diameter is one-half of the outer diameter, and the bearings have six 51° shoes. An ISO VG 68 lubricant
is customary for the speed range. The ratings are based on recommended film thicknesses, which follow
a trend of thicker films for larger bearing sizes and higher surface speeds. The resulting unit loads range
from 1.2 to 3.4 MPa. Equations in Section 45.5 can be used to calculate the minimum film thickness and
power loss.
The sizes, ratings, and calculated performance values are applicable for both vertical and horizontal
machines. A typical vertical arrangement is to have the bearings submerged in a pot of oil as in Figure
45.10 along with cooling coils to remove heat and maintain the desired pot temperature. The runner in
vertical applications serves the same purpose as the thrust collar. In a typical horizontal arrangement
(Figure 45.9), pressurized oil is delivered by an external lube system. The oil is provided through ports
in the casing to an annulus around the back of each thrust bearing base ring. Seal rings on either side of
the bearing allow the bearing cavity to flood with oil that then exits through a discharge port at the top.
45-12 Design for Lubrication and Tribology

TABLE 45.3  Rated Load (kN) for Kingsbury Type J Equalizing Pivoted Shoe Thrust
Bearings Based on ISO VG 68 Oil Supplied at 50°C
Shaft rpm
Size Area (mm ) 2 200 400 800 1,500 1,800 3,000 3,600
5 8,000 9 11 12 13 14 15 15
6 12,000 15 17 19 21 22 24 25
7 16,000 23 25 28 32 33 36 37
8 21,000 32 36 40 44 46 50 51
9 26,000 43 48 54 60 62 67 69
10.5 36,000 63 71 79 88 91 99 102
12 47,000 88 98 110 123 126 138 142
13.5 59,000 117 132 148 164 169 184 190
15 73,000 152 171 192 213 219 239 —
17 93,000 207 232 261 290 298 — —
Shaft rpm
Size Area (mm2) 100 200 300 400 500 600 700
19 117,000 242 272 291 305 317 326 335
21 142,000 309 347 371 390 404 417 428
23 171,000 386 433 464 486 505 520 534
25 202,000 473 531 568 596 619 638 654
27 235,000 570 640 685 719 746 769 789
29 272,000 679 762 815 855 887 915 —
31 310,000 798 896 958 1005 1043 — —
33 352,000 929 1042 1115 1170 — — —
37 442,000 1226 1376 1472 — — — —
41 543,000 1571 1764 — — — — —

Journal
bearing

Runner

Cooling
coils

Thrust
bearing

FIGURE 45.10  Vertical thrust bearing arrangement.

Flow rates on the order of 5 L/min/kW of power loss provide sufficient oil for the films and adequate
capacity to remove heat, giving a temperature rise of approximately 7°C exiting the bearing. It is impor-
tant to note that the slack-side power loss in Example 45.2 was for conditions with the main bearing
under load. A separate calculation of the slack-side bearing power loss under load needs to be performed
to establish the required slack-side flow. The calculation follows the method, for Example, 45.1 using the
slack-side geometry and slack-side design loads.
A reduction in temperature, power loss, and oil flow can be obtained by using a lower viscosity, ISO
VG 32 oil for cases where the actual load is less than the rating.11 Film thickness and power loss can be
compared using the equations in Section 45.5 and appropriate lubricant properties from Table 45.2.
Thrust Bearings 45-13

Self-lubricating bearing systems can provide an advantage in certain horizontal applications. The
systems use an oil circulator, fitted around the outer diameter of the collar, as a viscous pump to deliver
pressurized oil through the bearings and cooler.12 The self-lubricating system eliminates the need for
external bearing lubrication and backup emergency systems. Lower speed and less critical applications
may also be internally lubricated by nonpressurized systems using ring and disk features that spin with
the shaft, dip into a sump, and distribute the lubricant to the bearings.

45.6.2  High-Speed Industrial Applications


High-speed applications are mostly horizontal units that operate with bearing surface speeds in the
range of 35–110 m/s. A sophisticated computer model, incorporating features such as thermal and
mechanical deformations, is required for accurate analysis at high speeds.13 Turbulence in the film14–16
and churning in nonfilm areas17 dramatically increase power loss. Radial discharge ports, common in
low-speed applications, generate excess heat. Proper collar clearance and a suitable tangential oil outlet
are customary in high-speed applications (Figure 45.11).
The following methods and equations allow an approximation of high-speed performance for equal-
izing pivoted shoe thrust bearing sizes 5 through 17 of Table 45.3, which are common sizes used in
high-speed rotating machinery. Equations 45.5 through 45.7 provide the rated load, loaded-side rec-
ommended flow, and oil temperature rise based on using ISO VG 32 turbine oil supplied at 50°C. The
parameter D corresponds to the babbitt outer diameter in mm, u is the surface speed at the mean diam-
eter in meters per second, and kW is the bearing power loss in kilowatts.
Bearing rated load (kN):

kN = 6.13 × 10−5 × D 2.407 × u 0.186 (45.5)

Flow (liters/minute):

q = 6.95 × 10−6 × D 2.030 × u1.16 (45.6)



Temperature rise (°C):
ΔT = 36.0 × kW / q (45.7)

The equations in Section 45.5 can be used to estimate loaded- and slack-side power loss for surface
speeds up to about 70 m/s. Equation 45.8 is an empirical curve-fit of power loss data for speeds between

Tangential
outlet
Ro
tat
ion

FIGURE 45.11  Oil control ring and tangential outlet discharge configuration for high-speed industrial
applications.
45-14 Design for Lubrication and Tribology

70 and 110 m/s based on the equalizing bearing sizes and oil viscosity stated above, as well as appropriate
control ring clearances and tangential port sizes.
Bearing power loss (kW):
(45.8)
−7 1.714 2.22
kW = 2.18 × 10 × D ×u , 70 < u 110m/s

Many high-speed industrial applications require equal load capacity in both directions, in which case
an equal portion of flow is provided to an equivalent size slack-side bearing. An extra 50% increase in
power loss is a conservative approximation for the slack-side losses in this case. Equation 45.7 can be
used to calculate temperature rise of the oil for the individual loaded and slack sides and for the total
flow through the bearing.

Example 45.3
Calculate the rated load, power loss, oil flow, and lubricant temperature rise for a Table 45.3,
size 9 equalizing pivoted shoe thrust bearing at a shaft speed of 10,500 rpm, using an ISO VG 32
oil with an inlet temperature of 50°C. Consider a double bearing configuration with equal load
capacity in either direction.

Solution
The size, 9, is the babbitt outer diameter in inches and the inner diameter is half the outer diam-
eter. This corresponds to an outer diameter of 0.2286 m and an inner diameter of 0.1143 m. The
surface speed is calculated from the bearing mean diameter:

Mean diameter (m) dm 1


2(0.2286 + 0.1143) = 0.1715
Surface speed (m/s) u p × (0.1715) × 10, 500 / 60 = 94.3

The surface speed is above 70 m/s, and so Equation 45.8 is used for power loss. The babbitt outer
diameter D is 228.6 mm.
Bearing rated load: kN = (6.13 × 10−5) × (228.6)2.407 ×(94.3)0.186 = 68 kN.
Recommended oil flow: q = (6.95 × 10−6) × (228.6)2.030 × (94.3)1.16 = 83 L/min.
Loaded-side power loss: kW = (2.18 × 10−7) × (228.6)1.714 × (94.3)2.22 = 58 kW.
For equal capacity, the equivalent size slack bearing also requires a flow of 83 L/min for a
total of 166 L/min to the double bearing. Corresponding power loss is 50% higher for the slack
side, giving a total power loss of 87 kW. The temperature rise of the lubricant through the bear-
ing from Equation 45.7 is
87
ΔT = 36.0 × = 18.9°C
166

45.6.3  Temperature Limitations


Film thickness is fairly substantial at high speeds, and load capacity becomes dependent on the integrity
of the bearing materials and lubricants. A temperature detector embedded in the pad, close to the sur-
face, is the best method for monitoring the integrity of the bearings. The recommended position is the
Thrust Bearings 45-15

75/75
Location

Maximum
temperature Isotherms

Rotation

Peak
pressure Pressure
profile

Temperature
detector

FIGURE 45.12  Temperature, pressure, and critical location.

75/75 location, which represents 75% of the arc length in the direction of rotation and 75% of the radial
width from inner to outer diameter (Figure 45.12).
Temperature limits for lubricants vary among the types and brands. Most turbine mineral oils can
operate to 140°C before breakdown becomes a concern. A 75/75 temperature of 120°C is a reasonable
limit. Lubricants containing antiwear or extreme pressure additives typically are not recommended
for hydrodynamic bearings. The additives are beneficial for gears and roller bearings but tend to form
deposits on hydrodynamic bearing surfaces which cause excessive heat and failure. If the additives are
used, 75/75 temperatures may need to be restricted well below 95°C.
Babbitt is an exceptional material for protecting the shaft against damage. It has excellent resistance
to seizure and galling, embeds hard debris, and has good corrosion resistance. The same properties leave
it relatively soft, and its strength is limited by temperature and pressure. Temperatures near 135°C are
typically imposed as a critical limit. Allowing 15°C for alarm and trip settings place maximum levels for
continuous operation at 120°C.
Temperature limits, however, also depend on load. Pad surface temperatures increase in the direction
of rotation with maximum temperatures toward the trailing edge while pressures peak closer to the
center (Figure 45.12). There is an intermediate location, roughly in the area of the 75/75 position, where
the temperature/pressure combination is closest to the yield point of the babbitt.
Yield strength, calculated peak pressure, and 75/75 temperature can be used to assess the integrity
of the babbitt. Peak pressures are obtained from Reynolds’ equation and calculated by most computer
codes. A general rule of thumb is that the peak pressure is 2.5 times the average pressure. Yield points
from ASTM specification B23 Grade 2, a common grade used in industry, are plotted in Figure 45.13. As
an example, assume a bearing is running with an average film pressure of 3 MPa and the 75/75 detector
reading is 120°C. The peak pressure is approximately 2.5 × 3 MPa = 7.5 MPa. The yield point at 120°C,
from Figure 45.13, is approximately 15 MPa. The bearing is therefore running with a safety factor of 2
with respect to the yield point of the babbitt.

45.6.4  Temperature Charts


The rated loads for high-speed applications in Equation 45.5 are based on film thickness. Figures 45.14
through 45.17 contain 75/75 pad temperature data that can be used to assess if the application may be
45-16 Design for Lubrication and Tribology

ASTM B-23 grade 2 Babbitt


yield point (0.125% of gage length)
45
40
35

Yield point (MPa)


30
25
20
15
10
5
0
0 20 40 60 80 100 120 140 160 180
Temperature (°C)

FIGURE 45.13  High-speed criteria = temperature.

160

150

140 4 Mpa

3 Mpa
130
75/75 Temperature (°C)

2 Mpa
120

1 Mpa
110

100

90

80

70

60
0 20 40 60 80 100 120 140
Surface speed (m/s)

FIGURE 45.14  Center pivot steel 75/75 pad temperatures.

limited by temperature. The figures are based on babbitted pivoted shoe bearings, sizes, and designs as
described in Table 45.3, recommended flows from Equation 45.6, appropriate control ring clearances
and tangential port sizes, and ISO VG 32 oil supplied at 50°C.
Example 45.3 is used to explain the use of the figures, assuming 110°C is specified as an application
limit. Example 45.3’s rated load at 94.3 m/s is 68 kN. The corresponding unit load is 2.6 MPa, obtained
by dividing the rated load by the bearing area from Table 45.2. The 75/75 pad temperature for a center
pivot steel pad from Figure 45.14 at 2.6 MPa and 94.3 m/s is approximately 125°C, which exceeds the
110°C specification.
Thrust Bearings 45-17

160

150

140

130
4 Mpa
75/75 Temperature (°C)

120
3 Mpa
110
2 Mpa
1 Mpa
100

90

80

70

60
0 20 40 60 80 100 120 140
Surface speed (m/s)

FIGURE 45.15  Sixty percent offset pivot steel 75/75 pad temperatures.

160

150

140

130
75/75 Temperature (°C)

4 Mpa
120

3 Mpa
110
2 Mpa
1 Mpa
100

90

80

70

60
0 20 40 60 80 100 120 140
Surface speed (m/s)

FIGURE 45.16  Center pivot Cr–Cu 75/75 pad temperatures.


45-18 Design for Lubrication and Tribology

160

150

140

130
75/75 Temperature (°C)

120

110
4 Mpa
3 Mpa
100
2 Mpa
1 Mpa
90

80

70

60
0 20 40 60 80 100 120 140
Surface speed (m/s)

FIGURE 45.17  Sixty percent offset pivot Cr–Cu 75/75 pad temperatures.

Two options exist in high-speed applications to lower the operating temperatures. Pivot offset has a
significant influence in lowering the temperatures at high speeds, and 60% offset pads are typical for high-
speed applications. The 75/75 location temperatures from Figure 45.15 are below the 110°C criteria, and so
60% offset steel pads are suitable for the application. A chrome-copper backing material is a second option
for reducing operating temperatures. The higher heat transfer capability of copper draws heat away from
the babbitt surface. Figure 45.16 indicates that center pivot chrome-copper pads also operate below 110°C.

45.6.5  High-Speed/Low-Load Symptoms


With proper bearing design and manufacturing, and with attention to cleanliness of the lubricant,
hydrodynamic thrust bearings can provide many years of trouble-free service. There are, however,
external influences that can cause troublesome symptoms.18 Shaft currents and certain chemicals attack
babbitt. The result is typically an increase in temperature over time that can eventually lead to a failure
if the situation is not corrected. Babbitt will also show symptoms of fatigue in applications with frequent
starts and stops or cyclic loading. Reference 18 provides information on recognizing and troubleshoot-
ing symptoms.
Two unusual symptoms have been investigated relating to high-speed applications under conditions
of low load. The first is a situation where frictional drag on the pad surface exceeds the hydrodynamic
pressure resulting in a moment that acts to close the leading edge of the pad. This high-speed/low-load
phenomenon causes very high pad temperatures that are unexpected at low loads. A hydrodynamic
taper at the leading edge of the pad has been successful in resolving this situation.19
The second symptom is damage of the pad pivot contact areas. Under unloaded conditions, the thrust
pads can experience a vibratory motion called pad-flutter. Little or no axial load can also result in a
subsynchronous axial vibration of the entire rotor. In either case, the resulting vibrations can cause
significant fretting damage to the pivots. This results in a shift in rotor position and an increase in axial
Thrust Bearings 45-19

clearance. Vibration levels escalate as the damage progresses, typically resulting in a forced shutdown of
the machine. Adjustments to endplay or oil flow can change the dynamics of the machine and shift the
vibration out of the operating range. An axial damping mechanism can also be incorporated.20

45.7  Direct Lube Bearings


Relative motion between rotating and stationary components in a flooded bearing cavity, other than the
oil films, generates additional power loss and heat. The losses are called churning or parasitic losses as
they rob power from the machine while providing no benefit.21 Churning losses increase rapidly above
70 m/s and approach film friction levels above 110 m/s. The rapid increase in power loss requires sub-
stantial quantities of oil to remove the associated heat and cool the bearings.
Direct lube bearings are designed to eliminate churning losses, which correspondingly reduce pad
temperatures and oil flow requirements. The bearings channel flow directly to the collar surface. Some
direct lube bearing designs incorporate nozzles that spray oil directly on the collar surface between the
pads.22 The leading-edge-groove design (Figure 45.18) introduces cool oil directly into the leading edge
of the oil film.23
Channeling flow direct to the films bypasses nonfilm areas of relative motion such as the gap between
the inner base ring diameter and shaft illustrated in Figure 45.9. The full benefit of directed lubrication
is obtained by increasing the gap around the collar outer diameter, where surface speeds are highest,
and opening up the oil discharge port to allow the oil to exit freely. This assures the evacuation of oil
from areas of relative motion that do not support the load, and is referred to as an evacuated housing
configuration.
Direct lube bearings are most suitable for applications above 70 m/s. A 30% reduction in power loss
and a 15°C reduction in pad temperature are possible with the oil flow kept at conventional flooded
bearing levels. The lower pad temperatures allow an option to further reduce flow and power loss on the
order of 50% at higher speeds.

45.8  Hydroelectric Applications


In June of 1987, the American Society of Mechanical Engineers designated the first Kingsbury thrust
bearing application as an International Historic Mechanical Engineering Landmark. The bearing was

FIGURE 45.18  Direct lube leading edge groove.


45-20 Design for Lubrication and Tribology

FIGURE 45.19  Hydroelectric thrust shoe, hand scraped and with high-pressure lift pocket.

installed in June of 1912 in a hydroelectric turbine generator at the Holtwood Hydroelectric Station,
located on the Susquehanna River in Holtwood, Pennsylvania, and operated for 75 years with original
parts and negligible wear.24
Hydroelectric applications incorporate the largest hydrodynamic bearings. While sizes range from
0.8 to 5.2 m in diameter, the bearings are governed by the same hydrodynamic principles. Elastic and
thermal deformations are more of a design consideration due to the large shoes tremendous loads. Shoe
thickness, transient operation, and start-up and shutdown are important in design.25,26
As in any hydrodynamic bearing, care and precision are required for the runner and bearing sur-
face finish and flatness. Runners are typically ground and then lapped to obtain the proper surface
flatness, finish, and perpendicularity with the shaft axis. Babbitt surfaces are lapped or hand scraped
(Figure 45.19) to a surface plate or to the runner surface. Various hydroelectric thrust bearing designs
have evolved over time. Designs include springs and elastic supports, thermal insulation, surface
modifications, and different surface materials such as PTFE, introduced in the former Soviet Union
in the early 1970s.27
Vertical hydroelectric thrust bearings support the weight of the rotor and generator during start-up
and shutdown. Breakaway torque can be calculated by simple torque formula using the load at start-
up, the mean radius of the bearing surface, and static coefficient of friction which ranges 0.15–0.25
for babbitt on steel. Although hydroelectric units have sufficient torque to break the static friction,
high-pressure lift systems are sometimes used to ease start-up and prevent wear of the bearing surface
when the machine comes to rest at shutdown.25 A lift pocket is machined in the surface of each thrust
pad (Figure 45.19), and lubricant is provided under pressure by a high-pressure lubrication system,
which separates the surfaces. The high-pressure lift system is turned on prior to start-up and until the
unit develops a sufficient hydrodynamic film. The system is again energized during shutdown.

Acknowledgments
The authors would like to thank Kingsbury, Inc., for permission to publish the information in this
article as well as for all of the tables and figures containing data that have been accumulated from the
company’s engineering and research and development laboratories.

References
1. Reynolds, O. 1886. On the theory of lubrication and its application to Mr. Beauchamp Tower’s
experiments. Philos. Trans. R. Soc. Lond., Part A 177:157–234.
2. Kingsbury, A. 1950. Development of the Kingsbury thrust bearing. Mech. Eng. 72:957–962.
Thrust Bearings 45-21

3. Michell, A.G.M. 1905. The lubrication of plane surfaces. Z. Angew. Math. Phys. 52:123–137.
4. Fuller, D.D. 1956. Theory and Practice of Lubrication for Engineers. New York: John Wiley & Sons.
5. Kingsbury, A. 1931. On problems in the theory of fluid film lubrication, with an experimental
method of solution. Trans. ASME 53:59–74.
6. Elwell, R.C. and Booser, E.R. 1972. Low-speed limit of lubrication. Mach. Des. June:129–133.
7. Raimondi, A.A. 1960. The influence of longitudinal and transverse profile on the load capacity of
pivoted pad bearings. ASLE Trans. 3:265.
8. Raimondi, A.A. and Szeri, A.Z. 1987. Journal and thrust bearings. In CRC Handbook of Lubrication,
Vol. II, ed. E.R. Booser, pp. 426–434. Boca Raton, FL: CRC Press, Inc.
9. Wilcock, D.F. and Booser, E.R. 1957. Bearing Design and Application. New York: McGraw Hill Book
Co.
10. Oberg, E., Jones, F.D., Ryffel, H.H., McCauley, C.J., Heald, R.M., and Hussin, M.I, eds. 2008. Machinery’s
Handbook, 28th edn., pp. 2245–2257. New York: Industrial Press Inc.
11. Glavatskih, S. and DeCamillo, S. 2004. Influence of oil viscosity grade on thrust pad bearing opera-
tion. J. Eng. Tribol. 218:401–412.
12. Marchione, M. 2002. Operation and current developments of self-contained, self-lubricating thrust
and journal bearing systems. In 19th International Pump Users Symposium Proceedings, Houston,
TX, pp. 43–57.
13. Pinkus, O. 1990. Thermal Aspects of Fluid Film Tribology. New York: ASME Press.
14. Abramovitz, S. 1956. Turbulence in a tilting-pad thrust bearing. Trans. ASME 78:7.
15. Elrod, H.G. and Ng, C.W. 1967. A theory for turbulent fluid films and its application to bearings.
ASME J. Lubr. Technol. 89:346–362.
16. Gregory, R.S. 1974. Performance of thrust bearings at high operating speeds. ASME J. Lubr. Technol.
96:7–14.
17. Gregory, R.S. 1979. Factors influencing power loss of tilting-pad thrust bearings. ASME J. Lubr.
Technol. 101:154–163.
18. Kingsbury, Inc. A General Guide to the Principles, Operation and Troubleshooting of Hydrodynamic
Bearings. Philadelphia, PA: Publication HB, Kingsbury, Inc., 1997.
19. Wilkes, J., DeCamillo, S., and Kuzdale, M.J. 2000. Evaluation of a high speed, light load phenomena
in tilting pad thrust bearings. Proceedings of the 29th Turbomachinery Symposium, Turbomachinery
Laboratory, September, Texas A&M University, College Station, TX, pp. 177–185.
20. DeCamillo, S. 2006. Current issues regarding unusual conditions in high-speed turbomachinery.
Fifth EDF/LMS Poitiers Workshop Proceedings, Futuroscope, France, pp A.1–A.10.
21. Booser, E.R. 1990. Parasitic power losses in turbine bearings. STLE Tribol. Trans. 33:157–162.
22. Bielec, M.K. and Leopard, A.J. 1969. Tilting pad thrust bearings: Factors affecting performance and
improvements with directed lubrication. Proc. I. Mech. E. 184:93–102.
23. Mikula, A.M. 1985. The leading-edge-groove tilting-pad thrust bearing: Recent developments.
ASME J. Tribol. 107:423–430.
24. The Kingsbury bearing at Holtwood, ASME Identification Number HH 0587, ASME, New York,
1987.
25. Chambers, W.S. and Mikula, A.M. 1988. Operational data for a large vertical thrust bearing in a
pumped storage application. Trans. Soc. Tribol. Lubr. Eng. 31:61–65.
26. Ettles, C.M., Seyler, J., and Bottenschein, M. 2003. Some effects of start-up and shut-down on thrust
bearing assemblies in hydro-generators. ASME J. Tribol. 125:824–832.
27. Glavatskih, S.B. 2003. Tilting pad thrust bearings. In Tribological Research and Design for Engineering
Systems, ed. Dowson, D. et al., pp. 379–390. Amsterdam, the Netherlands: Elsevier.
46
Hydrodynamic Step
and Wedge Bearings
Theo G. Keith
University of Toledo

Sorin Cioc Nomenclature..............................................................................................46-1


University of Toledo
46.1 Design Variables..............................................................................46-2
L. Moraru 46.2 Step and Wedge Bearings...............................................................46-2
Politehnica University Geometry  •  Lubricant  •  Bearing Surface
of Bucharest References...................................................................................................46-21

Nomenclature
a ratio of the inlet and outlet film thicknesses (a = h1/h2)
B bearing width (m)
f friction coefficient
F friction force per unit width (N/m)
h film thickness (m)
h1 inlet film thickness (m)
h2 outlet film thickness (m)
L length of the bearing (m)
L1 length of the inlet region in the step bearing (m)
L2 length of the outlet region in the step bearing (m)
n ratio of the inlet length and the total length in the step bearing (n = L1/L)
p pressure (Pa)
p1 inlet pressure (Pa)
p2 outlet pressure (Pa)
pS pressure at the step (Pa)
p∗S nondimensional film pressure at the step location (in the step bearing)
Q volume flow rate, per unit width (m2/s)
Q * nondimensional volume flow rate
u flow velocity in the x direction (m/s)
U velocity of the bearing (m/s)
W load carried per unit width (N/m)
W* nondimensional load capacity
x coordinate along the bearing length (m)
y transverse coordinate (m)
z coordinate along the bearing width (m)
μ viscosity (Pa s)
ρ density (kg/m3)
τ shear stress (Pa)

46-1
46-2 Design for Lubrication and Tribology

46.1  Design Variables


The design of a step or wedge (or slider) bearing must take into account a multitude of variables, such
as the following:
1. Bearing type
a. Self-acting, hydrostatic, or combined
2. Geometry
a. Geometrical dimensions
b. Ideal geometry or with misalignment and/or wear
c. Bearing general shape: linear, curved, step, pivot, or elastic
3. Lubricant
a. Liquid (incompressible) or gas (compressible)
b. Newtonian or non-Newtonian (such as Bingham plastic)
c. Flow type: laminar or turbulent
d. Lubricant properties, including dependence on temperature
4. Lubricant supply
a. Supply conditions: pressure, flow rate, and temperature
b. Supply configuration/geometry
c. Supply restrictors
5. Bearing surface
a. Bearing surfaces: solid or porous
b. Surface geometry (roughness profiles)
6. Thermal conditions
a. Required operating temperatures
b. Cooling system
7. External conditions
a. Constant or variable load (such as vibration, shocks): magnitudes
b. Constant or variable sliding speed (such as starting or stopping)
8. Other issues
a. Cavitation
b. Bearing stability
Because of the large number of variables, no simple design procedure exists, and therefore, the user
must consider the influence of each variable and decide on what options are best suited for the given
application.
In the following, some of the classical results are presented for both step and wedge bearings.
References are provided for those cases that are too complicated for simple exposition. It should be
noted that most of such cases require numerical evaluation that cannot be easily described herein.

46.2  Step and Wedge Bearings


In 1918, Lord Rayleigh (William Strutt) wrote a seminal paper entitled “Notes on the Theory of
Lubrication” [1]. Rayleigh provided a detailed analysis of infinitely wide slider bearings, that is, bear-
ings that had no side leakage. His purpose was to examine the load carrying capacity of slider bearings.
Rayleigh discovered that the load carrying capacity of curved sliders was only slightly better than that of
linear sliders. He employed variational calculus and found that a slider in the form of a step could carry
the largest load. Over the years, an untold number of papers have appeared that deal with various effects
associated with the use of these bearings. In honor of its discoverer, the step bearing is now generally
referred to as the Rayleigh step slider bearing.
Hydrodynamic Step and Wedge Bearings 46-3

In what follows the geometry design parameter will first be considered. This will enable the classi-
cal cases of wedge and step bearings with and without side leakage to be described. Expressions for the
pressure, the peak pressure, the load carrying capacity, the frictional force, the friction coefficient, and
the volume flow rate will be presented. Performance optimization will also be discussed. Geometric
variations of both wedge and step bearings will be described.
Many papers involving the wedge and step bearings lubricated with various non-Newtonian fluids
have appeared. Results of several of these will be considered. Research associated with gas wedge and
step bearings will also be explored. Results for the porous step bearing will be presented. And results
concerned with the alteration of the bearing surface to improve performance will be described.

46.2.1  Geometry
46.2.1.1  Infinitely Wide Bearings
Because of the geometric simplicity of these bearings, one-dimensional analyses can be found in virtu-
ally every textbook containing fundamentals of hydrodynamic lubrication. Figure 46.1 presents the
geometry, coordinate systems, and notation for these bearings.
Classical one-dimensional analysis of these bearings is generally based on a number of simplify-
ing assumptions: steady flow; constant density; constant viscosity; no inertia effects; infinite width
(B →∞), that is, no side leakage; bearing surface moving at constant speed +U; and a stationary slider;
alternately the slider moves with −U over a stationary bearing, and zero or constant pressure at the
inlet and outlet.

46.2.1.1.1 Infinitely Wide Linear Slider Bearing


The simple linear slider consists of two smooth, flat surfaces, separated by a small angle to form a self-
acting lubricating wedge. The pressure distribution for the set of assumptions cited earlier and for this
geometry is governed by Reynolds equation:

d ⎛ 3 dp ⎞ dh
h = 6mU (46.1)
dx ⎜⎝ dx ⎟⎠ dx

z z

B B

x x
L
L L1 = nL L2 = (1–n)L

Slider Slider
p1 = 0 p2 = 0
Inlet Outlet Inlet region
h1 h1
h2 ps Outlet region h2
y y
x Bearing x
U Bearing U
Linear wedge bearing Rayleigh step bearing

FIGURE 46.1  Geometry, coordinate systems, and notation for linear wedge and step bearings.
46-4 Design for Lubrication and Tribology

Since h varies linearly with x, Equation 46.1 may be integrated twice and the boundary conditions
applied to obtain

6mUL ⎡ 1 1 h1h 2 ⎛ 1 1 ⎞⎤
p= ⎢ − − ⎜ 2 − 2 ⎟⎥ (46.2)
h1 − h 2 ⎣ h h1 h1 + h 2 ⎝ h h1 ⎠ ⎦

where it is also assumed that the inlet is flooded (the liquid film fills the entire height).
The load per unit width supported by the wedge bearing is obtained by integrating the pressure distribu-
tion over the bearing surface (of length L and width B = 1):

L
6mUL2 ⎛ h1 h − h2 ⎞
W = pdx =
∫ 2 ⎜
ln − 2 1
(h1 − h 2 ) ⎝ h 2

h1 + h 2 ⎠
(46.3)
0

The nondimensional load capacity, which is the inverse of the Sommerfeld number, S, is

1 Wh 22 6 ⎛ a −1 ⎞
W* = = = ln a − 2 (46.4)
S mUL2 (a − 1)2 ⎜⎝ a + 1 ⎟⎠

where a = h1/h2 is the film height ratio. It should be mentioned that in practical applications the bearing
film thickness ranges from 0.005 to 0.1 mm [2].
The maximum pressure, which occurs at the location where dp/dx = 0, is

⎛ mUL ⎞ ⎡ a − 1 ⎤ ⎛ mUL ⎞ *
pmax = ⎜ 2 ⎟ ⎢ ⎥ = ⎜ 2 ⎟ pmax (a) (46.5)
⎝ h 2 ⎠ ⎣ 4a(a + 1) ⎦ ⎝ h 2 ⎠

The frictional force is found by integrating the wall shear stress: τw = μ(∂u/∂y)y = 0, where the velocity in
the x direction is, u = U(1 − y/h) + (1/12μ)(dp/dx)(y2 − hy), and is a linear combination of Couette and
Poiseuille flows through an opening of height h with the lower surface moving at a speed of +U and the
upper surface stationary:

L L
⎛ du ⎞ 4 mUL ⎛ 3(a − 1) ⎞ ⎛ mUL ⎞ *
∫ ∫
Fy = 0 = (tw )dx = ⎜ m ⎟ dx = −
⎝ dy ⎠ y = 0

(a − 1)h 2 ⎝
ln a − ⎟
2(a + 1) ⎠
= −⎜
⎝ h2 ⎠
⎟ F (a) (46.6)
0 0

The friction coefficient can be calculated by dividing the frictional force by the load:

F h2 ⎡ ( 4 ln a − 6(a − 1) / (a + 1)) ⎤ h
f= = (a − 1) ⎢ ⎥ = 2 f* (46.7)
W 6L ⎢⎣ ( ln a − 2(a − 1) / (a + 1)) ⎥⎦ L

Because there is no side leakage, the rate of lubricant flow per unit width is

h1 h2
a
∫ ∫
Q = u(0, y )dy = u(L, y )dy = Uh 2
a +1
= Uh 2Q * (a) (46.8)
0 0
Hydrodynamic Step and Wedge Bearings 46-5

TABLE 46.1  Nondimensional Linear Slider


Performance Parameter Values

a Q* W* F* f* p*max
1 0.5 0 1 ∞ 0
1.2 0.54545 0.07551 0.91916 12.17328 0.01894
1.4 0.58333 0.11771 0.86472 7.34628 0.02976
1.6 0.61538 0.14109 0.82567 5.85221 0.03606
1.8 0.64286 0.15336 0.79608 5.19100 0.03968
2 0.66667 0.15888 0.77259 4.86262 0.04167
2.5 0.71429 0.15773 0.72916 4.62288 0.04286
3 0.75000 0.14792 0.69722 4.71357 0.04167
3.5 0.77778 0.13599 0.67109 4.93498 0.03968
4 0.80000 0.12420 0.64839 5.22071 0.03750

Table 46.1 presents numerical values for the nondimensional performance parameters of the linear
slider bearing as a function of the film height ratio.

46.2.1.1.2 Infinitely Wide Rayleigh Step Bearing


The step bearing may be thought of two regions of constant film thickness joined at the step. The volume
flow rate per unit width, (B = 1), through each region of the bearing is

h1 h2
Uh1 h13 ⎛ dp ⎞ Uh 2 h 3 ⎛ dp ⎞

Q = u1dy =
2
− ⎜ ⎟ =
12 m ⎝ dx ⎠ 1 ∫ u 2dy =
2
− 2 ⎜ ⎟
12m ⎝ dx ⎠ 2
(46.9)
0 0

The pressure distribution for this simple two domain geometry is governed by Reynolds equation,
Equation 46.1, which for the step bearing is d2p/dx 2 = 0. This expression is integrated twice, and the
boundary conditions are applied for each region. The first integration reveals that the pressure gradient
in each section of the bearing is constant. The second integration reveals that the pressure distribu-
tion in each region is linear with the peak pressure occurring at the step. Using (dp/dx)1 = ps/nL and
(dp/dx)2 = −ps/(1 − n)L in Equation 46.9 yields the pressure at the step

6 mU(h1 − h 2 ) 6 mULn(1 − n)(a − 1) ⎛ mUL ⎞ *


ps = = = ⎜ 2 ⎟ ps (a,n) (46.10)

( ) (
⎡ h13 / nL + h 32 / (1 − n)L ⎤
⎣ ⎦ ) ⎡(1 − n)a 3 + n ⎤ h 22
⎣ ⎦ ⎝ h2 ⎠

where
a = h1/h2 ≥ 1
n = L1/L ≤ 1
ps* is the nondimensional film step pressure

The linear pressure distributions in both regions can be conveniently written as

⎛ mUL ⎞ 6(1 − n)(a − 1) ⎛ x ⎞


p=⎜ 2 ⎟ , 0 ≤ x ≤ nL
⎝ h 2 ⎠ ⎡(1 − n) a 3 + n ⎤ ⎜⎝ L ⎟⎠
⎣ ⎦
(46.11)
⎛ mUL ⎞ 6n(a − 1) ⎛ x⎞
p=⎜ 2 ⎟ 1− , nL ≤ x ≤ L
⎝ h 2 ⎠ ⎡(1 − n)a 3 + n ⎤ ⎜⎝ L ⎟⎠
⎣ ⎦
46-6 Design for Lubrication and Tribology

The volume flow rate per unit width is

Uah 2 a 3h 2 U(1 − n)(a − 1) Uh 2 ⎧⎪ (1 − n)a 3 + na ⎫⎪ *


Q= − = ⎨ ⎬ = Uh 2Q (a,n) (46.12)
2 2 ⎡(1 − n)a 3 + n ⎤ 2 ⎪ ⎡(1 − n)a 3 + n ⎤ ⎪
⎣ ⎦ ⎩⎣ ⎦⎭

The load carried per unit width is obtained by integrating the pressure distributions

nL (1− n )L
ps L ⎛ mUL2 ⎞ 3n(1 − n)(a − 1) ⎛ mUL2 ⎞ *
W=
∫ pdx +
∫ pdx =
2
=⎜ 2
⎝ h2
⎟ =⎜ 2
⎠ ⎡⎣(1 − n)a + n ⎤⎦ ⎝ h 2
3 ⎟ W (a,n)

(46.13)
0 nL

The shear stresses acting on the lower moving surface (y = 0) in the two regions of the step bearing are
given respectively as

⎛ ∂u ⎞ mU h1 ⎛ dp ⎞ ⎡ mU ah 2 ⎛ dp ⎞ ⎤
(tw )0 − nL = m ⎜ ∂y ⎟ =−
h1
− ⎜ ⎟ = −⎢
2 ⎝ dx ⎠ 1
+ ⎜ ⎟ ⎥
⎝ ⎠ y =0 ⎣ ah 2 2 ⎝ dx ⎠ 1 ⎦

⎛ ∂u ⎞ ⎡ mU h 2 ⎛ dp ⎞ ⎤
(tw )nL − L = m ⎜ ⎟ = −⎢ + ⎜ ⎟ ⎥ (46.14)
⎝ ∂y ⎠ y = 0 ⎣ h2 2 ⎝ dx ⎠ 2 ⎦

The frictional force per unit width on the lower moving surface is then determined by the integration of
the shear stress acting on the respective region length:

nL L
⎛ mUL ⎞ ⎧⎪ 3n(1 − n)(a − 1)2 n + (1 − n)a ⎫⎪ ⎛ mUL ⎞ *
∫ 0 nL

Fy = 0 = (tw )0 − nL dx + (tw )nL − L dx = − ⎜
⎝ h
⎟⎨ 3
2 ⎠ ⎪ (1 − n)a + n
⎩ ⎣
⎡ ⎤

+
a
⎬ = −⎜

⎭ ⎝
⎟ F (a,n)
h2 ⎠

(46.15)

The friction coefficient is given by

Fy =0 ( mUL/h2 ) F* (a,n) h2 ⎧
⎪ ⎡(1 − n)a 3 + n ⎤ ⎡⎣(1 − n)a + n ⎤⎦ ⎫⎪ h
⎣ ⎦ 2 *
f= = = ⎨(a − 1) + ⎬ = f (a,n) (46.16)

W ( mUL /h ) W (a,n)
2 2
2
*
L
⎩⎪
3na (1 − n )(a − 1)
⎭⎪
L

Table 46.2 provides values for the nondimensional design parameters of an infinitely wide Rayleigh step
bearing for values of step height ratio when n = 0.5, that is, L1 = L2.
Rahmani et al. [3] considered the effect of hydrostatic pressure on the performance of an infinitely
wide Rayleigh step bearings, that is, p(0) = pin and p(L) = pout. Results were presented in terms of a non-
dimensional parameter Ψ = (pin − pout )h 22 / (mUL), which can be positive, negative, or zero. They found
( )
that the nondimensional load, W* = Wh 22 / (mUL2 ), was always negative for Ψ < −0.295; was positive
or negative depending upon the values of n and a for −0.295 < Ψ ≤ 0; and was always positive for Ψ > 0.

46.2.1.2  Optimum Performance


46.2.1.2.1 Linear Slider Bearing
For an imposed minimum film thickness, h2, the maximum load that can be carried by the bearing is
obtained when dW*/da = 0. Table 46.3 gives bearing performance values in this case.
Hydrodynamic Step and Wedge Bearings 46-7

TABLE 46.2  Nondimensional Rayleigh Step Slider


Performance Parameter Values for n = 0.5

a p*s Q* W* F* f*
1.0 0.0000 0.5000 0.0000 1.0000 ∞
1.2 0.2199 0.5367 0.1100 0.9387 8.5356
1.4 0.3205 0.5534 0.1603 0.9212 5.7486
1.6 0.3532 0.5589 0.1766 0.9185 5.2006
1.8 0.3513 0.5585 0.1756 0.9183 5.2281
2.0 0.3333 0.5556 0.1667 0.9167 5.5000
2.5 0.2707 0.5451 0.1353 0.9030 6.6722
3.0 0.2143 0.5357 0.1071 0.8810 8.2222
3.5 0.1709 0.5285 0.0855 0.8565 10.0214
4.0 0.1385 0.5231 0.0692 0.8327 12.0278

TABLE 46.3  Optimum Nondimensional Values for an Infinitely Wide Linear


Wedge Bearing

h1 Qh 2 Wmax h 22 1 Fh 2 fL
a= Q* =
*
Wmax = Smax = F* = f* =
h2 U mUL2 *
Wmax mUL h2
2.1887 0.6864 0.1602 6.2406 −0.7542 4.7066

TABLE 46.4  Optimum Nondimensional Values for an Infinitely Wide Rayleigh Slider
Bearing

h1 L1 ps h 22 Qh 2 * Wmax h 22 Fh 2 fL
a= n= ps* = Q* = Wmax = F* = f* =
h2 L mUL U mUL2 mUL h2

1.8660 0.7182 0.4125 0.6220 0.2063 −0.8453 4.0981

46.2.1.2.2 Rayleigh Step Bearing


Since W* depends on two parameters, a and n, the maximum nondimensional load is determined
when both (∂W*/∂a)n = 0 and (∂W*/∂n)a = 0 are satisfied. The first differentiation produces L1/L2 = n/
(1 − n) = 2a3 − 3a 2, and the second yields L1/L2 = n/(1 − n) = a3/2. By equating the two expressions, a cubic
equation is obtained: 4a3 − 12a2 + 9a − 1 = 0. The roots of this equation are a = 1, 1 ± 3 /2. The only root
that is meaningful is a = 1 + 3 /2 = 1.8660. With this value, the optimum inlet length ratio is found to be
( )
n = 3 + 2 3 / 9 = 0.7182. Hence, the optimum position of the step is nearer to the outlet than the inlet.
Table 46.4 contains optimum nondimensional results using these values.
Rahmani et al. [3] studied the effect of hydrostatic pressure on the optimum performance of infinitely
wide Rayleigh step bearings. Optimization in terms of the frictional force and the friction factor were
also considered.

46.2.1.3  Finite Width Bearings


To provide more realistic analysis, the effects of side leakage must be taken into account. Fortunately,
some analytical results are available and will be briefly described in the following. Because of the form of
the two-dimensional Reynolds equation, the analysis of the Rayleigh step bearing involves less complex
functions than does the linear slider and therefore will be considered first.
46-8 Design for Lubrication and Tribology

46.2.1.3.1 Step Bearing of Finite Width


Reynolds equation reduces to Laplace’s equation for the bearing pressure in two dimensions:

∂2p ∂2p
+ =0 (46.17)
∂x 2 ∂z 2

As shown in Figure 46.1, the pressure is zero on all boundaries and ps(z) at the step. The step pressure is
symmetrically distributed across the slider, that is, ∂p/∂z = 0 at z = B/2 and p(0) = p(B) = 0.
Solution to this problem can be accomplished by applying the separation of variables method. The
solution was originally presented by Archibald in References 4 and 5. It can also be found in Hamrock
[6] and Pinkus and Sternlicht [7]. Like the infinitely wide Rayleigh step bearing, the two-dimensional
problem in x and z is solved separately in the inlet and outlet regions, and then connected at the step.
Separate coordinate axes for each region are located at the inlet and outlet boundaries. The pressures
within the inlet and outlet regions are found to be, respectively,

∞ ∞
⎧ sin(lk z)sinh(lk x) ⎫ ⎧ sin(lk z)sinh(lk x) ⎫
p= ∑ ⎨Pk
k =0 ⎩
sinh ( l L
k 1 )
⎬ and p = −

⎨Pk
k =0 ⎩

sinh(lk L2 ) ⎭
⎬ (46.18)

where
λk = (2k + 1) π/B, k = 0,1,2,…
Pk are Fourier coefficients

The step pressure is evaluated in a similar manner to that used to determine the step pressure for the
infinitely wide bearing, that is, the volume flow rate for each region at the step is set equal to each other.
Performing this operation and making use of several identities yields the Fourier coefficients appearing
in Equation 46.18:

24 mU(h1 − h 2 )
Pk = (46.19)

k {
l B h coth(lk L1 ) + h 32 coth(lk L2 )
2 3
1 }
The load carried by the inlet and outlet regions is determined by integrating the pressure distributions
over the respective regions and summing the results to obtain

48mU(a − 1)
∞ ⎧ ⎡ tanh ( lk nL / 2 ) + tanh ⎡⎣(lk (n − 1)L) / 2 ⎤⎦ ⎤ ⎫⎪
W=
Bh 22 ∑ ⎪⎨⎩⎪ l1 ⎢⎢⎣
k =0
4
k a3 coth(lk nL) + coth ⎡⎣ lk (n − 1)L ⎤⎦ ⎥⎦ ⎪
⎥⎬

(46.20)

The frictional force is found to be

mUBL ⎡ n + (1 − n)a ⎤ 24 mU(a − 1)2


∞ ⎧ ⎡ ⎤⎫
F=
h 2 ⎢⎣ a ⎥+
⎦ h2B ∑ ⎪⎨⎩⎪ l1 ⎢⎢⎣ a coth(l nL) + cot
k =0
3
k
3
k
1
k

⎥⎬
h ⎡⎣ l (n − 1)L ⎤⎦ ⎥⎦ ⎪

(46.21)

The volume flow rate entering the inlet region is

⎧ ∞ ⎧⎪ ⎡⎡ ⎤ ⎤ ⎫⎪⎫⎪
⎪ a 4a 3 (a − 1) 1 1
Q1 = Uh 2B ⎨ −
⎪⎩ 2 B2 ∑k =0 ⎩
⎨ 2 ⎢⎢ 3 ⎥ ⎥ ⎬⎬ (46.22)
⎪ lk sinh(lk nL) ⎢⎣ ⎢⎣ a coth(lk nL) + coth ⎡⎣ lk (n − 1)L ⎤⎦ ⎥⎦ ⎥⎦ ⎭⎪⎪⎭

Hydrodynamic Step and Wedge Bearings 46-9

The volume flow rate leaving the outlet region is

⎧ ∞⎧⎪ ⎡⎡ ⎤ ⎤ ⎫⎪⎫⎪
⎪ 1 4(a − 1) 1 1
Q2 = Uh 2B ⎨ +
⎪⎩ 2 B2 ∑
k =0 ⎩
⎨ 2 ⎢ ⎢ ⎥ ⎥ ⎬⎬ (46.23)
⎪ lk sinh(lk nL) ⎢⎣ ⎢⎣ a coth(lk nL) + coth ⎡⎣ lk (n − 1)L ⎤⎦ ⎥⎦ ⎥⎦ ⎭⎪⎪⎭
3

The side leakage is simply Q1 − Q2.

46.2.1.3.2 Linear Slider Bearings of Finite Width


The separation of variables method has been used to obtain analytical expressions for a finite linear
slider bearing lubricated by a Newtonian incompressible lubricant, for example, see Muskat et al. [8]
and Hays [9]. However, the series solutions involve Bessel functions which are best summed on a
computer. As an alternate procedure, the load capacity of a finite size slider can be estimated by mul-
tiplying the load capacity of the infinite bearing by a factor hW = Wfinite * * *
length / Winfinite , where Winfinite is
calculated from Equation 46.4. Example values for the factor η W are provided in Table 46.5; see Frene
et al. [10].
Similarly, the friction force can be evaluated by multiplying the friction force of the infinite bearing
by a factor ηF = Ffinite length/Finfinite, where the denominator is computed from Equation 46.6. Example val-
ues for the factor ηF are provided in Table 46.6 [10].

46.2.1.4  Optimum Performance


46.2.1.4.1 Step Bearing of Finite Width
Archibald [4] and [5] found for a square step bearing (L = B) that the maximum load capacity occurred
when h1 = 1.7 h2 and L1 = 1.2 L2, that is, a = 1.7 and n = 0.545. Kettleborough [11] confirmed the load capac-
ity value of 0.07256 using numerical methods. Table 46.7 presents results of a spreadsheet computation
for a square step bearing at this length and height ratio showing the load capacity and the frictional force

TABLE 46.5  Values for the Factor η W


B/L
a 0.25 0.5 1 2 4
1.5 0.055 0.17 0.42 0.69 0.835
2 0.06 0.18 0.44 0.69 0.84
5 0.85 0.23 0.48 0.72 0.87
Source: Frene, J. et al., Hydrodynamic
Lubrication Bearings and Thrust Bearings,
Tribology Series 33, Ed. D. Dowson, Elsevier,
London, U.K., pp. 92–93, 1997.

TABLE 46.6  Values for the Factor ηF


B/L
a 0.25 0.5 1 2
1.5 0.096 0.97 0.98 0.99
2 0.90 0.92 0.94 0.97
5 0.69 0.74 0.83 0.95
Source: Frene, J. et al., Hydrodynamic
Lubrication Bearings and Thrust Bearings,
Tribology Series 33, Ed. D. Dowson,
Elsevier, London, U.K., pp. 92–93, 1997.
46-10 Design for Lubrication and Tribology

TABLE 46.7  First Five Terms in Series Solution for a Square Step Slider, a = 1.7 and n = 0.545

k 2k + 1 λk Wk* W* Fk* F* f*


0 1 3.14159 0.07087 0.07087 0.83517 0.83517 0.08486
1 3 9.42478 0.00141 0.07228 0.00238 0.83755 0.08630
2 5 15.708 0.00019 0.07247 0.00051 0.83806 0.08647
3 7 21.9911 0.00005 0.07252 0.00019 0.83825 0.08651
4 9 28.2743 0.00002 0.07254 0.00009 0.83834 0.08652
5 11 34.5575 0.00001 0.07254 0.00005 0.83838 0.08653

TABLE 46.8  W* Values for Various Bearing Length to Width Ratios, n = 0.5
a╲L/B 0.1 0.25 0.5 0.75 1 1.5 2 4
1.2 0.10272 0.09209 0.07437 0.05800 0.04450 0.02655 0.01677 0.00458
1.4 0.14968 0.13420 0.10838 0.08452 0.06485 0.03869 0.02444 0.00668
1.5 0.16012 0.14355 0.11594 0.09041 0.06937 0.04139 0.02615 0.00714
1.6 0.16496 0.14789 0.11944 0.09314 0.07147 0.04264 0.02694 0.00736
1.7 0.16586 0.14870 0.12010 0.09365 0.07186 0.04288 0.02709 0.00740
1.8 0.16406 0.14708 0.11879 0.09263 0.07108 0.04241 0.02679 0.00732
1.9 0.16044 0.14385 0.11617 0.09059 0.06951 0.04148 0.02620 0.00716
2 0.15567 0.13957 0.11272 0.08790 0.06745 0.04024 0.02542 0.00695
3 0.10007 0.08972 0.07246 0.05651 0.04336 0.02587 0.01634 0.00447
4 0.06466 0.05797 0.04682 0.03651 0.02802 0.01672 0.01056 0.00289
5 0.04448 0.03988 0.03221 0.02511 0.01927 0.01150 0.00726 0.00198

calculations for the first five terms of the series in Equations 46.20 and 46.21. As may be observed, the series
for the load capacity converges quickly, and the first term is within a few percent of the converged result.
It should be mentioned that the first few terms in the load capacity series, Equation 46.20, for a square
bearing when n = 0.5 are especially simple and convenient for hand calculation.
Table 46.8 presents load capacity values for various step bearing length to width ratios. All calcula-
tions were made with n = 0.5. It can be seen that the load capacity increases as L/B decreases, that is, the
bearing becomes wider, and W* approaches values for an infinitely wide bearing.
Table 46.8 also reveals that a maximum value occurs around the height ratio of 1.7. Optimum design
load values for a step bearing were obtained by Rohde [12] and presented in Table 46.9.

46.2.1.5  Other Geometries


46.2.1.5.1 Infinitely Wide Curved Slider Bearings
Over the years, various shapes of infinitely wide sliders have been examined. Arguably the most com-
prehensive analysis was presented by Bagci and Singh [13]. Expressions were shown for the pressure,
load, frictional force, volume flow rate, location, and maximum pressure. Expressions for the optimized
load results were also presented. Table 46.10 provides some of the film shapes examined.

46.2.1.5.2 Step Bearing of Finite Width


46.2.1.5.2.1 Step Designs  Table 46.11 presents volume flow rate data for step bearings at three L/B ratios
that represent wide, square, and narrow bearings. The computations were performed using Equations
46.22 and 46.23 for a film ratio, a = 1.7 and for n = 0.5. As may be seen, nearly one-third of the flow
leaks out the sides of the square bearing. For a bearing with an aspect ratio above 3, over 40% of its
flow leaks out of the sides. The effect of the side leakage greatly erodes the load-carrying capacity of
Hydrodynamic Step and Wedge Bearings 46-11

TABLE 46.9  Optimum W* Values for


Various Bearing Length to Width Ratios
a n L/B W*
1.8660 0.7182 0 0.20627
1.81 0.69 0.25 0.16542
1.74 0.64 0.5 0.12745
1.69 0.55 1 0.07259
1.68 0.51 2 0.02710
Source: Rohde, S.M., ASLE Trans., 17(2),
105, 1974.

TABLE 46.10  Optimum Film Ratio and Load Capacity for Various
Curved Sliders
*
Title Film Shape h(x) aopt Wmax
Linear (tapered) h2[1 + (a − 1)x/L] 2.188 0.16025
Exponential h2emx = h2ax/L 2.3 0.16518
Catenoidal h2cosh(mx) 2.4 0.175
Polynomial h2[1 + (a − 1)(x/L)n] 2.4 0.177
Cycloidal h2[1 + (a − 1)(x/L) − sin(2πx/L)/2π] 2.0 0.192
Truncated cycloidal h2[1 + 2(a − 1)(x/2L) − sin(πx/L)/2π] 2.4 0.175
Source: Bagci, C. and Singh, A.P., ASME J. Tribol., 105, 48, 1983.

TABLE 46.11  Volume Flow Rates for Three Step Bearing


Configurations
L/B Q1 Q2 Q1 − Q2 % Leakage
Wide 0.1 0.58085 0.55478 0.02606 4.5
Square 1 0.75558 0.51922 0.23636 31.3
Narrow 10 0.85000 0.50000 0.35000 41.2

the step bearing. This occurs as the length to width ratio, L/B, becomes large and causes the pressure
distribution to be diminished. Comparing the optimum value of W* = 0.2060 for an infinitely wide
Rayleigh step bearing to the value W* = 0.2063 for a square step bearing with a = 1.7 and n = 0.545 clearly
illustrates this significant reduction. The data in Table 46.8 reveal that there is an order of magnitude
reduction in the load-carrying capacity between a step bearing with L/B =1.0 and one with L/B =4.0.
To reduce the deleterious side leakage effect, Archibald [4] suggested that the step should be curved,
as illustrated in Figure 46.2. Wilcock [14] proposed a depressed pocket area in the surface of the bear-
ing. The pocket helped to reduce high surface shear stress and thereby enabled a reduction in power
consumption. Subsequently other redesigns took the form of shrouds and/or pockets of various shapes.

FIGURE 46.2  Archibald’s curved step.


46-12 Design for Lubrication and Tribology

FIGURE 46.3  Concentric Rayleigh step journal bearing with a single step and groove.

Kettleborough [15] used an electrolytic tank to experimentally investigate six different step designs of
square Rayleigh step bearings. Rohde [12] used finite elements to analyze five step shapes.
46.2.1.5.2.2 Rayleigh Step Journal Bearings  Archibald [4] modified his series solution for the finite
width step bearing and applied it to a concentric journal bearing with a single step, Figure 46.3. A single
groove was also included in order to maintain the constant clearance. Two cases were considered: the
half bearing and the full bearing.
Archibald and Hamrock [16] extended the analysis to include a journal bearing with more steps.
However, it was assumed that the film thickness within each step and ridge region was constant, that is,
they neglected any curvature effects. For cases in which there were numerous steps, the approximation
was found to be acceptable; however, for a small number of steps, for example, 4, the results were found
to be inaccurate.
Comprehensive analyses were performed by Hamrock [17] and Hamrock and Anderson [18–20]. In
[17], Hamrock employed a linearization to solve Reynolds equation for a compressible lubricant and
obtained the pressure distribution within an infinitely long journal bearing containing N evenly dis-
tributed steps and grooves. It was found that the total load capacity of a Rayleigh step journal bearing is
larger than that of a plain journal bearing provided there were a small number of steps and provided it
operated at a high compressibility number, Λ = 6μUR/paC2, where pa is the ambient pressure and C is the
radial clearance of the journal bearing. A search for optimum conditions of the Rayleigh step journal
bearing revealed that

• As the compressibility number, Λ, increases, the optimal step film thickness ratio increases, and
the optimal ridge-to-step length ratio increases, although very modestly.
• The number of pads placed around the journal should be few in number.

Hamrock and Anderson [18] described two theoretical analyses (1) of a single-step, concentric, finite
length (B is finite) Rayleigh journal bearing and (2) of a multistep, eccentric, infinite length (B → ∞)
Rayleigh journal bearing. It should be noted that this article is a condensed version of two earlier NASA
Technical Notes [19] and [20], which contain considerably more information and data.

46.2.1.5.2.3 Radial Rayleigh Step Bearings  Archibald [5] also analyzed a sector step slider by solving
Laplace’s equation for the pressure in the r, θ plane using the separation of variables method. This step
bearing with a semicircular step was also treated numerically by Kettleborough [15]. He found that the
load supported was W = 0.946mw R o4 /h 22 .
Hydrodynamic Step and Wedge Bearings 46-13

FIGURE 46.4  Fan-shaped step bearings.

In a brief note, Hamrock [21] modified Archibald’s series solution [4] and determined the optimum
number of sectors and the parallel step configuration for a step-sector thrust bearing lubricated with
an incompressible fluid. Optimization of the load capacity revealed that aopt = 1.668 and nopt = 0.558. For
maximum stiffness, it was found that aopt = 1.467 and nopt = 0.558.
Figure 46.4 is an illustration of a fan-shaped step bearing. This bearing has been analytically inves-
tigated by Liu et al. [22]. The separation of variables method was used to produce a set of analytical
expressions for the pressure, load capacity, flow rate, and torque loss.

46.2.2  Lubricant
46.2.2.1  Effects of Viscosity
Thus far, all solutions have been obtained assuming that the lubricant is Newtonian, that is, the shear
stress τ is linearly related to the strain rate ∂u/∂y and the viscosity, μ, is constant. However, many com-
mon lubricants exhibit non-Newtonian behavior, and the viscosity is not constant. Numerous viscosity
relations are available, for example, see Briant et al. [23]. A relation based on kinetic theory of liquids
which connects viscosity to temperature and pressure is μ = μoe αpe −βT, where α and β are empirical con-
stants. The more important case is that in which μ(T). However, this case involves the energy equation,
and solutions are generally obtained by numerical methods. On the other hand, a few variable viscosity
cases are accessible and can offer some useful conclusions.

46.2.2.2  Effects of Pressure


Following the same variational approach used by Rayleigh in [1], Charnes et al. [24] demonstrated that
the stepped slider was the optimum profile for an infinitely wide slider bearing lubricated by a fluid with
viscosity that varied with pressure according to the Barus formula [25], μ = μoeαp. To determine the opti-
mum configuration and load capacity, they used a transformation described by Muskat and Evinger [26]:

p
dp
pˆ = mo
∫m (46.24)
0

where p̂ may be interpreted as the pressure. With the transformation and the Barus formula, it is
easy to show that p = − (1/α)ln(1 − αp̂). This expression enables the load capacity to be determined:
L


W = −(a −1 ) ln(1 − ap)dx
0
oped for the case
ˆ . The x distribution of p̂ can be obtained from the pressure distributions devel-
of an infinitely wide Rayleigh step bearing in which the viscosity is constant; however, p̂
46-14 Design for Lubrication and Tribology

replaces p and μo replaces μ. It was found that the loading is increased for the optimum step slider when
viscosity increases with pressure. This implies that the true pressure is always higher than that found assum-
ing the viscosity is constant.

46.2.2.3  Non-Newtonian Fluid Models


46.2.2.3.1 Power-Law Fluids
A lubricant that obeys the power-law fluid model is governed by

n n −1
⎛ ∂u ⎞ ⎡ ⎛ ∂u ⎞ ⎤ ⎛ ∂u ⎞
t = m⎜ ⎟ = ⎢ m ⎜ ⎟ ⎥⎜ ⎟ (46.25)
⎝ ∂y ⎠ ⎢⎣ ⎝ ∂y ⎠ ⎥⎦ ⎝ ∂y ⎠

where m and n̄ are viscometric experimental constants. n̄ is referred to as the rheological characteristic
index. This index is a fluid property, and its deviation from unity determines the non-Newtonian behav-
ior of the fluid. When n̄ = 1, the fluid is Newtonian; when n̄ < 1, the fluid is a pseudo-plastic or shear thin-
ning fluid; and when n̄ > 1, the fluid is a dilatant or shear thickening fluid. For most fluids, the rheological
index ranges somewhere between 0.4 and 2.5. The effective viscosity in a power-law model fluid is seen
to vary with strain rate, ∂u/∂y.
Wang and Jin [27] analyzed flow of a power-law lubricant through a one-dimensional Rayleigh step
bearing. They used a first-order small parameter perturbation method to obtain an approximate solu-
tion of the modified Reynolds equation developed by Dien and Elrod [28] for Couette dominated flows.
The results revealed that the nondimensional load capacity increased with an increase of the index, n̄.
They also concluded that dilatant lubricants may be superior to pseudo-plastic lubricants.
Elkouh and Yang [29] obtained an exact solution of the same problem described in Reference 27. To
accomplish this, an exact solution for generalized Couette flow of a power-law fluid was first developed.
The results of this analysis were used to facilitate the solution of a power-law fluid flow in the Rayleigh
step bearing. The computed results were compared to the values obtained by Bourgin and Gay in [30]
for a range of indices between 0.6 and 1.0. In general, good agreement was obtained. Comparisons were
also made with the results of Wang and Jin [27], which revealed substantial differences particularly for
low values of the rheological index.

46.2.2.4  Other Viscosity Models


46.2.2.4.1 Rabinowitsch Fluid
Auloge et al. [31] demonstrated that the Rayleigh step bearing is the optimum shape of the one-dimensional
slider bearing that can support maximum load for given width of contact zone, the sliding ratio, and
the fluid rheology. It was assumed that the contact was lubricated by polymer-thickened oil that was
described by the constitutive equation of the so-called Rabinowitsch fluid: τ + kτ3 = μ∂u/∂y, where k and
μ are constants determined by experiment. When k = 0, the fluid is Newtonian; when k > 0, the fluid is a
pseudo-plastic or a shear thinning fluid; and when k > 0, the fluid is a dilatant or shear thickening fluid.
The maximum load capacity and optimal step configuration were determined numerically for this par-
ticular non-Newtonian lubricant model.

46.2.2.4.2 Micropolar Fluids


It is well known that lubricant’s characteristics can be altered because of additives. The additives can
take various forms. In one model, they can be regarded as a suspension of tiny solid particles within the
base lubricant. The resulting mixture belongs to a special class of fluids termed micropolar fluids. For
low concentrations of additives, the lubricant can be treated as Newtonian, but for higher concentra-
tions, the mixture must be characterized by non-Newtonian constitutive relations. Eringen [32] pre-
sented a theory for micropolar fluids. The micropolar fluid model has found applicability in numerous
Hydrodynamic Step and Wedge Bearings 46-15

lubrication studies, for example, [33] and [34]. Maiti [35] numerically determined the load capacity and
frictional resistance of a step slider bearing with a micropolar lubricant. It was found that both were
larger than those of a Newtonian fluid. Ramanaiah and Dubey [36] demonstrated that the slider step
bearing configuration provided the maximum load capacity when run with a micropolar lubricant.

46.2.2.5  Second-Order Fluids


Multigrade lubricants contain additives of long-chain polymer molecules that affect the behavior of the
base oil. This form of lubricant was created in order to reduce the viscosity’s dependence on tempera-
ture. However, it was subsequently revealed that the additives produced complex rheological effects.
Harnoy [37] provided a list of five fluid characteristics affected by polymer additives:
1. The viscosity of the base oil is increased.
2. The reduction of viscosity with temperature is moderated.
3. The viscosity becomes a decreasing function of the strain rate.
4. Normal stresses are introduced.
5. Stress-relaxation characteristics are present, and the lubricant is termed viscoelastic (viscous
effects are of first order, whereas elasticity effects are of second order).
Various models have been developed to analyze viscoelastic fluid flows. In the case of multigrade lubri-
cants, problems may be analyzed in terms of differential constitutive equations, in which the stress
components are written directly in terms of the strain rate components [37]. However, this type of con-
stitutive relation is restricted to flow problems in which the Deborah number (the ratio of the relaxation
time of the fluid to the characteristic time of the flow) is small, that is, De ≪ 1.
Hydrodynamic lubrication theory of viscoelastic fluids is based on the second-order fluid equation
of Rivlin and Erickson [38]. The constitutive relations for a slow second-order fluid were developed
by Coleman and Noll [39]. Numerous lubrication investigations involving second-order fluids have
appeared over the years, for example, [40,41]. Bujurke et al. [42] obtained compact expressions for the
pressure, load capacity, and frictional coefficient for a Rayleigh step bearing using the Coleman–Noll
constitutive equation for a second-order fluid. It was found that the load increased and the coefficient
of friction decreased compared to results for a Newtonian lubricant. The maximum dimensionless load
was found to occur at a slightly smaller step ratio than for the Newtonian case.

46.2.2.5.1 Bingham Fluid


The Bingham fluid model is relatively simple and is characterized by the existence of a yield stress at zero
strain rate which must be exceeded in order to cause fluid motion. It is found that the Bingham fluid
model can be effectively used to represent grease. Greases are made of mineral or synthetic oils and are
a suspension of soaps and thickeners. The latter act as sponges that contain the oil which, when within
a bearing film, gradually release the oil at a very slow rate.
Radulescu and Vasiliu [43] developed a simplified method to analyze grease-lubricated Rayleigh step
bearings. The geometric model is similar to that for a Newtonian fluid except that in the inlet region, the
effective film thickness is reduced because of the presence of a stagnant core. Expressions were obtained
for the flow capacity, pressure, load capacity, and friction coefficient. The theoretical results were com-
pared with data obtained experimentally and with results from the Newtonian fluid model.
Hong et al. [44] used a finite volume computational fluid dynamics (CFD) method (SIMPLEC) to
numerically solve the Navier–Stokes equations for a two-dimensional Rayleigh step bearing with a
Bingham fluid. The results are more accurate than those obtained from the solution of Reynolds equation.
An electrorheological (ER) fluid contains a suspension of extremely fine nonconducting particles sus-
pended within an electrically insulating fluid. The viscosity of an ER fluid changes dramatically when
an electric field is applied to the fluid. When activated, an ER fluid can be regarded as a Bingham plastic
with a yield stress that is dependent upon the strength of the electric field. Leel et al. [45] used Bingham
plastic constitutive properties in an experimental investigation of the flow of an ER fluid within the
46-16 Design for Lubrication and Tribology

Rayleigh step bearing. Similarly, magnetorheological (MR) fluids experience an increased viscosity
when a magnetic field is applied.

46.2.2.6  Effects of Density


46.2.2.6.1 Gas Step Bearings
Gas bearings have been employed in a wide variety of applications. For example, they are used in dental
drills, high-precision instruments, machine tools, computer peripheral devices, high-speed turboma-
chines, and navigational instruments. A review of gas bearings research and development from 1915 to
1990 was performed by Pan [46].
Gas bearings produce very little friction even at high speeds (at ambient temperatures, the viscosity of
air is roughly 1/10,000th that of a light oil); hence, they have low frictional losses and generate minimal
amounts of heat. Gas lubricants are chemically very stable over a wide range of temperatures; they nei-
ther freeze nor boil; they are nonflammable; they do not cavitate; and they do not contaminate bearing
surfaces. Gas bearings also have low-noise characteristics.
On the negative side, gas bearings do not have much load-carrying capacity and have large startup
wear. Also, because the gas film thickness is quite small, gas bearings require superior surface finish and
manufacture. Further, thin gas films offer very little cushion or damping capacity; consequently, gas
bearings are prone to certain vibration instabilities.
Hamrock [47] has pointed out three important points regarding gas bearings that are illustrated in
Figure 46.5: (1) At very low velocities, gas bearing behavior is approximately that of liquid lubricated
bearings; (2) at very high velocities, the product of ph is constant, and therefore, load is independent of
speed as well as viscosity; and (3) at very high velocities, side leakage effects are negligible, and finite and
infinite width-bearing solutions approach ph = constant.
There are basically two classes of gas bearing problems: thin film lubrication and ultrathin film
lubrication. A key nondimensional parameter that delineates these classes is the Knudsen number, Kn,
which is the ratio of the molecular mean free path, λ, to the characteristic length, which in this case
is the film thickness, h. The Knudsen number can also be shown to be proportional to the ratio of the
Mach number, M, and the Reynolds number, Re. Table 46.12 identifies various flow regimes according
to the magnitude of the Kn; see Faria and San Andrés [48].
It should be mentioned that the modified Reynolds equation for a rarefied gas has been derived by two
different approaches: one includes slip velocity boundary conditions within the Reynolds equation and
the other uses the Boltzmann equation to derive the lubrication equation.

High-speed asymptote
Incompressible low-
speed asymptote

Infinite width
Load, W Incompressible low-
speed asymptote

Finite width

Speed, U

FIGURE 46.5  Effect of speed on load capacity for self-acting gas bearings. (From Hamrock, B.J., Fundamentals of
Fluid Film Lubrication, McGraw-Hill, New York, pp. 319–325, 1994. With permission.)
Hydrodynamic Step and Wedge Bearings 46-17

TABLE 46.12  Gas Bearing Flow Regimes


Kn < 10−3 Fluid is a continuum; Reynolds equation for compressible fluids may be used
10−3 ≤ Kn ≤ 10 Fluid is a rarefied gas; a modified Reynolds equation must be used
Kn > 10 Fluid is a free molecular flow
Source: Faria, M.T.C. and San Andrés, L., ASME J. Tribol., 122, 124, 2000.

The bearing number, Λ = 6mUL / pa h 22 , plays an important role in gas bearing problems. A large value
of Λ indicates that compressible effects are significant. On the other hand, a value of Λ = 0 applies to an
incompressible fluid. Thus, for small values of the bearing number, the behavior of gas lubricants may
be approximated by liquid lubricants.
Because the governing equation in gas bearing problems is nonlinear, no exact analytical solutions
are available for a finite gas step bearing. Consequently, limiting asymptotic approaches, perturbation
methods, linearizations, and numerical methods have been used to obtain solutions for gas step bearings.

46.2.2.6.2 Infinitely Long Step Slider


Assuming no side leakage, the mass rate of flow through a step slider may be written as

Urh h 3 r ⎛ ∂p ⎞
 = rQ =
m − (46.26)
2 12 m ⎜⎝ ∂x ⎟⎠

For the following set of assumptions: (1) p(x), (2) gas is a perfect gas, p = ρRT, and (3) gas flow is isother-
mal (p/ρ = constant), Equation 46.26 can be written

d ⎛ Uph h 3p dp ⎞
⎜ − ⎟=0 (46.27)
dx ⎝ 2 12m dx ⎠

Nondimensionalizing Equation 46.27 using X = x/L, P = p/pa, H = h/h2, and Λ = 6mUL / pa h 22 yields

d ⎛ dP ⎞
ΛPH − H3P =0 (46.28)
dX ⎜⎝ dX ⎟⎠

Equation 46.28 is integrated, and a constant of integration, C, is set equal to PH, evaluated at the
unknown location X0 where dP/dX = 0. Equation 46.28 is integrated a second time to obtain P + D + (C/H)
ln(P − C/H) = ΛX/H2, where D is a second constant of integration. Utilizing the boundary condition
P(0) = 1, the pressure within the inlet region can be written as

⎛ PH1 − C ⎞ Λ
H1(P − 1) + Cln ⎜ ⎟= X (46.29)
⎝ H1 − C ⎠ H1

Utilizing the boundary condition P(1) = 1, the pressure within the outlet region is (note H2 = 1 within
the outlet region)

⎛ P−C ⎞
(P − 1) + C ln ⎜ ⎟ = Λ(X − 1) (46.30)
⎝ 1− C ⎠

These two pressure distributions, Equations 46.29 and 46.30, are evaluated at the step, that is, at X = Xs,
P(Xs) = Ps to produce
46-18 Design for Lubrication and Tribology

⎛ PH −C ⎞ Λ ⎛ P −C ⎞
H1(Ps − 1) + C ln ⎜ s 1 ⎟= X s and (Ps − 1) + C ln ⎜ s ⎟ = Λ(X s − 1) (46.31)
⎝ H1 − C ⎠ H1 ⎝ 1− C ⎠

For given values of H1, Xs, and Λ, these equations may be solved to determine the unknowns C and Ps.
However, the pair are nonlinear algebraic equations which do not yield analytical solutions and there-
fore must be solved numerically. Szeri [49] has presented the following asymptotic solutions:
For the asymptotic case of Λ → 0,

⎡ X (1 − X s )(H1 − 1) ⎤ ⎡ X (1 − X s )H12 + X s ⎤
Ps = 1 + Λ ⎢ s 3 ⎥ + O(Λ 2 ) C = ⎢ s 3 ⎥ H1 (46.32)
⎣ (1 − X s )H1 + X s ⎦ ⎣ (1 − X s )H1 + X s ⎦

P −1 X P −1 1− X
= , 0 ≤ X ≤ Xs = , Xs ≤ X ≤ 1 (46.33)
Ps − 1 X s Ps − 1 1 − X s

For the asymptotic case of Λ → ∞,

(H1 − 1)(1 − b) ⎪⎧ ⎡ Λ(X s − X) ⎤ ⎪⎫


P =1+ ⎨exp ⎢ − ⎥⎬, 0 ≤ X ≤ Xs
1+ e ⎩⎪ ⎣ H12 ⎦ ⎭⎪

(H1 − 1)(1 − a ) ⎪⎧ ⎡ Λ(1 − X) ⎤ ⎪⎫


P =1+ ⎨1 − exp ⎢ − ⎬, Xs ≤ X ≤ 1 (46.34)
1+ e ⎩⎪ ⎣ H12 ⎥⎦ ⎭⎪

where
(
a = exp − ΛX s /H12 )
β = exp[− Λ (1 − Xs)/H1]
ε = α (H1 − 1 − H1β)

These limiting cases may be plotted for given values of Λ, a = h1/h2, and n = L1/L.
Faria and San Andres [48] have used modern finite element and finite difference methods to numeri-
cally investigate one-dimensional, isothermal, gas-lubricated slider linear, and step bearings. Figure 46.6 is
a plot of the pressure distribution within a Rayleigh step slider bearing for three bearing numbers.

2
Λ = 1000

Λ = 100
p/pa
1.5
Λ = 10

1
0 0.2 0.4 0.6 0.8 1
x/L

FIGURE 46.6  Computed nondimensional pressure distributed within an isothermal gas lubricated Rayleigh step
bearing for three bearing numbers. (From Faria, M.T.C. and San Andrés, L., ASME J. Tribol., 122, 124, 2000. With
permission.)
Hydrodynamic Step and Wedge Bearings 46-19

Gross [50] used a numerical approach to determine the effect of the bearing number on the load
capacity for a square step bearing with the step located at n = 0.5 for two step ratios a = 2 and 3. The gas
lubricant was assumed to be isothermal. It was found that the load-carrying capacity of the gas bearing
is considerably smaller than a step bearing with a liquid lubricant for bearing numbers above 40 for a = 2
and above 100 for a = 3.

46.2.2.6.3 Head-Disk Interface in Hard Disk Drives


One of the most important direct applications of the concept of finite length slider bearings is the head-
disk interface found in modern magnetic hard disk drives. An example of this geometry is shown in
Figure 46.7, Wu and Bogy [51].
The slider is supported by an elastic system, which balances the pressure distribution developed in the
slider bearing, such that the read-write element is maintained at a prescribed distance from the rotating
disk surface. The minimum film thickness can be on the order of 10 nm or smaller (below 5 nm in cur-
rent high-density disk drives), while the gas mean free path is about 65 nm [51]. As such, the Knudsen
numbers can be as high as 100, and therefore, the governing equation used to describe the flow of air
between the slider and the disk was modified from the Reynolds equation to account for the rarefaction
and slip effects. The steady-flow equation can be written as [52,53]

∂ ⎛ dP ⎞ ∂ ⎛ dP ⎞ 6 mUL ∂(PH)
⎜ Q PPH3 ⎟ + ⎜ Q PPH3 = (46.35)
∂X ⎝ dX ⎠ ∂Z ⎝ dZ ⎟⎠ pa h 22 ∂X

where
X and Z are the x and z coordinates normalized by the reference length L
P is the pressure normalized by the ambient pressure pa
H is the film thickness divided by the reference thickness taken here as the thickness at the exit
QP is the flow factor accounting for the rarefaction and slip effects

The flow factor can be calculated as a function of the momentum accommodation coefficient and the
Knudsen number Kn = 2l/ p h, where λ is the mean free path [54].
The governing equation can be integrated numerically (coupled with the energy equation since the
temperature affects the viscosity and the mean free path) using the appropriate boundary conditions.
The slider is then optimally designed to insure the desired flying height, including the dynamic effects,
Bogy et al. [55].

1.00 Air bearing surface Read-write


element
0.88

0.75
Read-write element 0.8
0.63
Rails
0.50 256 – 6m

0.38

0.25
Fully recessed region 0m
1.0

0.13 0.0 Wall profile region


0.0

0.00
(a) 0.00 0.16 0.31 0.47 0.63 0.78 0.94 1.09 1.25 (b) 2.5 E–6m

FIGURE 46.7  Example of a National Storage Industry Consortium (NSIC) bearing surface design; dimensions
are in millimeters; the leading edge is on the left; on (a), darker color corresponds to smaller distance between the
disk and the slider surface; (b) is an isometric representation of the slider. (From Wun, L. and Bogy, D.B., J. Comput.
Phys., 172, 640, 2001. With permission.)
46-20 Design for Lubrication and Tribology

46.2.3  Bearing Surface


46.2.3.1  Porous Step Bearings
A porous step slider bearing can have either or both slider and bearing surfaces made of a porous mate-
rial. The porous material is often sintered bronze (90% copper, 10% tin), but iron–copper mixtures have
also been used. The lubricant can be liquid, solid, or gas. The porous surface is often backed by a solid
surface to provide strength. The porous matrix can be pressurized to force lubricant through the matrix
pores, or it can be impregnated with lubricant which can be slowly released during operation. Porous
bearings are relatively simple, and they are inexpensive. They find application in situations where non-
porous bearings are impractical, for example, in harsh environments and where direct lubrication is
not possible. Self-lubricated applications involve bearings for small fractional horsepower motors that
drive fans found in hair driers, sewing machines, record players, tape recorders, vacuum cleaners, and
so forth.
Although the idea of utilizing porous bearings in self-lubricating bearings was advanced as early as
the 1920s, serious research on porous bearings has been underway for more than a half century. Sneck
[56] provides a thorough review of porous gas bearings, and Kumar [57] provides a critical review of
porous metal bearings.
In order to analyze porous bearings, Reynolds equation is modified, using Darcy’s law to account for
the lubricant flow at the surface. Darcy’s law is similar to Fourier’s heat conduction law, for example,
qx = −k∂T/∂x which compares to u = − (Φ/μ)∂p/∂x, where Φ is the porosity of the material. The pressure
within the porous matrix satisfies the Laplace equation provided that the lubricant is a Newtonian fluid.
Prakash and Vij [58] obtained closed form expressions for the pressure, load capacity, and coefficient
of friction for an infinitely wide porous plane slider. It was found that as porosity increased, the load
capacity decreased and the coefficient of friction increased.
Verma et al. [59] repeated the variational approach of Rayleigh and demonstrated that the porous step
was the optimum shape to carry the maximum load. They also determined the optimum step height and
length ratios to carry the maximum load in terms of the porosity. It was found that the porous step slider
could support a greater load than the porous inclined slider. Kumar [60] extended the calculations by
computing the coefficient of friction for this geometry.
Srinivasan [61] determined closed form expressions of the performance characteristics for an infi-
nitely wide double-layered porous slider bearing. It was found that the effect of the second layer was to
increase load capacity and decrease the frictional coefficient. Naduvinamani [62] used an approximate
method to analyze a double-layered Rayleigh step slider bearing for the case when the lubricant was
a second-order fluid. The maximum load capacity was determined to be close to 9% larger compared
to the load for a conventional porous Rayleigh step bearing. Also an increase of more than 11% in the
maximum load was realized over that of a Newtonian fluid with the same viscosity.

46.2.3.2  Surface Effects


46.2.3.2.1 Surface Roughness Profile
Numerous studies have been performed to determine the influence of the surface roughness on the
hydrodynamic lubrication, as well as to investigate the influence of texturing the surfaces at various
scales. The most often used method to account for the surface roughness was first introduced by Patir
and Cheng [63]. This method modifies the Reynolds equation by including three nondimensional func-
tions, or “flow factors,” Φ depending on the roughness statistics. The first two factors, also called “pres-
sure flow factors,” represent the ratio between the average flow in the rough case and the flow in the
perfectly smooth case generated by the pressure variations in x and z directions (Poiseuille component);
the third factor, called the “shear flow factor,” represents the corresponding ratio for the sliding-induced
flow (Couette component). Compared to perfectly smooth surfaces, roughness can significantly affect
the pressure distribution inside the fluid film, especially when the minimum film thickness is close to
Hydrodynamic Step and Wedge Bearings 46-21

the roughness size. The effect of the roughness can be both detrimental or advantageous, depending on
the slider geometry roughness orientation and magnitude, Patir and Cheng [64]. For example, in case of
wide bearings, they found that the load capacity was increased when the roughness was oriented trans-
versally to the flow; for narrow bearings, the load capacity increased for longitudinal roughness, while for
square bearings, isotropic surfaces were best. In general, roughness should oppose (be perpendicular to)
the dominant flow in the slider to increase load capacity. However, these observations cannot be gener-
alized for all roughness profiles, as they were determined for surfaces having the same roughness. When
different roughness profiles were used, the shear flow factor became predominant, and the load capacity
increased appreciably if the moving surface was smooth, and decreased appreciably if it was rough [64].
In addition, stability problems occurred in certain working conditions due to the decrease of the stiff-
ness in the case of the “smooth moving case.”

46.2.3.2.2 Surface Texturing


Compared to random roughness profiles, one can introduce deterministic, engineered profiles (called
surface texturing) with the purpose of improving the slider performance: load capacity, friction coef-
ficient, and stiffness and damping coefficients, as well as general reliability. Texturing can be manufac-
tured using lasers, chemical etching, plastic deformation, etc. As expected, a complete theoretical study
of the fluid flow between textured surfaces is far more complex than can be handled by the classical
Reynolds equation, and thus is performed numerically, by solving the Navier–Stokes and energy equa-
tions. However, some analytical methods based on modifications of the Reynolds equation have also been
proposed. For example, Tonder [65] considered the case of very high bearing numbers Λ = 6mUL/p0h 2r ,
where hr is the reference film thickness, while more recently, de Kraker et al. [66] developed a texture
averaged Reynolds equation, similar to Patir and Cheng’s method for rough surfaces, but where the flow
factors were calculated by integrating the full Navier–Stokes equations on local unit cells; the local (cell)
and global (slider) flows were coupled by the mass conservation imposed “in an averaged sense.” They
found that the flow factors depend on the working conditions (sliding velocity and pressure gradient);
they also showed that inertia (disregarded in the Reynolds equation) plays an important role.
The effect of the texture on the slider performance depends on the size, shape, density, pattern, and
orientation, as well as on the operating conditions. Due to the multidimensionality of the texturing
optimization, a large range of optimum parameters have been attained in different studies [67]. Etsion
[68] confirmed that a trial and error procedure is generally used for this purpose.

References
1. Rayleigh, L., Notes on the theory of lubrication, Philosophical Magazine, Series 6, 35(205), 1918,
1–12.
2. Frene, J., Nicolas, D., Degueurce, B., Berthe, D., and Godet, M., Hydrodynamic Lubrication
Bearings and Thrust Bearings, Tribology Series 33, Ed. D. Dowson, Elsevier, London, U.K., 1997,
pp. 87–88.
3. Rahmani, R., Shirvani, A., and Shirvani, H., Analytical analysis and optimization of the Rayleigh
step slider bearing, Tribology International, 42, 2009, 666–674.
4. Archibald, F.R., Simple hydrodynamic thrust bearing, Transactions of the ASME, 72(4), 1950,
393–400.
5. Archibald, F.R., The stepped shape film applied to a journal bearing, Journal of the Franklin Institute,
253(1), 1952, 21–27.
6. Hamrock, B.J., Fundamentals of Fluid Film Lubrication, McGraw-Hill, New York, 1994, pp. 198–205.
7. Pinkus, O. and Sternlicht, B., Theory of Hydrodynamic Lubrication, McGraw-Hill Book Co., New York,
1961, pp. 118–120.
8. Muskat, M., Morgan, F., and Meres, M.W., Studies in lubrication VII the lubrication of plane sliders
of finite width, Journal of Applied Physics, 11, 1940, 739–748.
46-22 Design for Lubrication and Tribology

9. Hays, D.F., Plane sliders of finite width, ASLE, 1(2), 1958, 233–238.
10. Frene, J., Nicolas, D., Degueurce, B., Berthe, D., and Godet, M., Hydrodynamic Lubrication
Bearings and Thrust Bearings, Tribology Series 33, Ed. D. Dowson, Elsevier, London, U.K., 1997,
pp. 92–93.
11. Kettleborough, C.F., The stepped thrust bearing—A solution by relaxation methods, Transactions of
the ASME Journal of Applied Mechanics, 76, 1954, 19–24.
12. Rohde, S.M., Finite element optimization of finite stepped slider bearing profiles, ASLE Transactions,
17(2), 1974, 105–110.
13. Bagci, C. and Singh, A.P., Hydrodynamic lubrication of finite slider bearings: Effect of one dimen-
sional film shape, and their computer aided designs, ASME Journal of Tribology, 105, 1983, 48–66.
14. Wilcock, D.F., The hydrodynamic pocket bearing, Transactions of the ASME, 77, 1955, 311–319.
15. Kettleborough, C.F., An electrolytic tank investigation into stepped thrust-bearings, Proceedings of
the Institution of Mechanical Engineers, 169, 1955, 679–688.
16. Archibald, F.R. and Hamrock, B.J., The Rayleigh step bearing applied to a gas-lubricated journal
of finite length, Transactions of the ASME, Series F, Journal of Lubrication Technology, 89(1), 1967,
38–47.
17. Hamrock, B.J., Rayleigh step journal bearing part I—Compressible fluid, Transactions of the ASME,
Series F, Journal of Lubrication Technology, 89, 1968, 271–280.
18. Hamrock, B.J. and Anderson, W.J., Rayleigh step journal bearing part II—Incompressible fluid,
Transactions of the ASME, Series F, Journal of Lubrication Technology, 90, 1969, 641–650.
19. Hamrock, B.J. and Anderson, W.J., Incompressibly lubricated Rayleigh step journal bearing I—
Zero-order perturbation solution, NASA TN D-4839, October 1968.
20. Hamrock, B.J. and Anderson, W.J., Incompressibly lubricated Rayleigh step journal bearing II—
Infinite-length solution, NASA TN D-4873, November 1968.
21. Hamrock, B.J., Optimum parallel step-sector lubricated with an incompressible fluid, ASME Journal
of Tribology, 107, 1985, 282–283.
22. Liu, S., Chen, W.W., and Hua, D.Y., Analytical solution to the hydrodynamic lubrication of fan
shaped thrust step bearings, ASME Journal of Tribology, 132, 2010, 024504.
23. Briant, J., Denis, J., and Parc, G., Rheological Properties of Lubricants, Editions Technip, Paris, France,
1989, pp. 23–63.
24. Charnes, A., Osterle, F., and Saibel, E., On the solution of the Reynolds equation for slider-bearing
lubrication—VIII the optimum slider profile for viscosity a function of the pressure, Transactions of
the ASME, 77, 1955, 33–36.
25. Barus, C., Isothermals, isopiestics, and isometrics relative to viscosity, American Journal of Science,
45, 1893, 87–96.
26. Muskat, M. and Evinger, H.H., Studies in lubrication IX the effect of the pressure variation of viscos-
ity on the lubrication of plane sliders, Journal of Applied Physics, 11, 1940, 739–748.
27. Wang, J. and Jin, G., The optimum design of the Rayleigh slider bearing with a power law fluid,
Wear, 129, 1989, 1–11.
28. Dien, I.K. and Elrod, H.G., A generalized steady-state Reynolds equation for non-Newtonian fluid
with application to journal bearings, Transactions of the ASME, Journal of Tribology, 105, 1983,
385–390.
29. Elkouh, A.F. and Yang, D., Flow of a power-law fluid in a Rayleigh step, Transactions of the ASME,
Journal of Tribology, 113, 1991, 428–433.
30. Bourgin, P. and Gay, B., The optimum Rayleigh bearing in terms of load capacity or friction for non-
Newtonian lubricants, Transactions of the ASME, Journal of Tribology, 107, 1985, 59–67.
31. Auloge, J.Y., Bourgin, P., and Gay, B., The optimum design of one-dimensional bearings with non-
Newtonian lubricants, Transactions of the ASME, Series F, Journal of Lubrication Technology, 105,
1983, 391–395.
32. Eringen, A.C., Simple microfluids, International Journal of Engineering Science, 2, 1964, 205–217.
Hydrodynamic Step and Wedge Bearings 46-23

33. Shukla, J.B. and Isa, M., Externally pressurized optimum bearing with micropolar fluid as lubricant,
Japanese Journal of Applied Physics, 14(2), 1975, 275–279.
34. Khonsari, M.M. and Brewe, D.E., On the performance of finite journal bearings lubricated with
micropolar fluids, STLE Tribology Transactions, 32(2), 1989, 155–160.
35. Maiti, G., Composite and step slider bearings in micropolar fluid, Japanese Journal of Applied Physics,
12(7), 1973, 1058–1064.
36. Ramanaiah, G. and Dubey, J.N., Optimum slider profile of a slider bearing lubricated with a micro-
polar fluid, Wear, 42, 1977, 1–7.
37. Harnoy A., Non-Newtonian viscoelastic effects, in Bearing Design in Machinery, Marcel Decker, New
York, 2003, Chapter 19, pp. 582–595.
38. Rivlin, R.S. and Erickson, J.L., Stress deformation relations for isotropic materials, Journal of Rational
Mechanics Analysis, 4, 1955, 323–425.
39. Coleman, B.D. and Noll, W., An approximation theorem for functionals with applications in con-
tinuum mechanics, Archive for Rational Mechanics Analysis, 6, 1960, 355–370.
40. Rout, B. and Roy, J.S., Step slider bearing with viscoelastic lubricant, Wear, 66(3), 1982, 333–344.
41. Sawyer, W.G. and Tichy, J.A., Non-Newtonian lubrication with second-order fluid, ASME, Journal of
Tribology, 120(3), 1998, 622–628.
42. Bujurke, N.M., Jagadeeswar, M., and Mulimani, B.G., Rayleigh step bearing with second-order fluid,
Japanese Journal of Applied Physics, 26(12), 1987, 2121–2126.
43. Radulescu, A.V. and Vasiliu, F., A simplified calculus method for a grease-lubricated Rayleigh step
bearing, Lubrication Science, 11(3), 1999, 271–283.
44. Hong, Y., Chen, D., Wang, J., and Li, X., Numerical modeling of a Newtonian and Bingham fluid in
a Rayleigh step bearing, ASME, Journal of Tribology, 125, 2003, 206–210.
45. Leel, T.H., Lingard, S., Atkin, R.J., and Bullough, W.A., An experimental investigation of the flow
of an electro-rheological fluid in a Rayleigh step bearing, Journal of Physics, D. Applied Physics, 26,
1993, 1592–1600.
46. Pan, C.H.T., Gas bearings (1915–1990), in Achievements in Tribology, Eds. L.E. Sibley and F.E.
Kennedy, ASME TRIB-Vol. 1, 1990, pp. 31–55.
47. Hamrock, B.J., Fundamentals of Fluid Film Lubrication, McGraw-Hill, New York, 1994, pp. 319–325.
48. Faria, M.T.C. and San Andrés, L., On the numerical modeling of high-speed hydrodynamic gas
bearings, ASME, Journal of Tribology, 122, 2000, 124–130.
49. Szeri, A.Z., Fluid Film Lubrication Theory & Design, Cambridge University Press, Cambridge, U.K.,
1998, pp. 392–407.
50. Gross, W.A., Gas Film Lubrication, John Wiley & Sons, Inc., New York, 1962, pp. 188–197.
51. Wun, L. and Bogy, D.B., Numerical simulation of the slider air bearing problem of hard disk drives
by two multidimensional upwind residual distribution schemes over unstructured triangular
meshes, Journal of Computational Physics, 172, 2001, 640–657.
52. Fukui, S. and Kaneko, R., Analysis of ultra-thin gas film lubrication based on linearized Boltzmann
equation: First report–derivation of a generalized lubrication equation including thermal creep flow,
ASME, Journal of Tribology, 110, 1988, 253–261.
53. Liu, N. and Bogy, D.B., Temperature effect on a HDD slider’s flying performance at steady state,
Tribology Letters, 35, 2009, 105–112.
54. Chen, D. and Bogy, D.B., Simulation of static flying attitudes with different heat transfer models for
a flying-height control slider with thermal protrusion, Tribology Letters, 40, 2010, 31–39.
55. Bogy, D.B., Fong, W., Thornton, B.H., Zhu, H., Gross, H.M., and Bhatia, C.S., Some tribology and
mechanics issues for 100-Gb/in2 hard disk drive, IEEE Transactions on Magnetics, 38(5), 2002,
1879–1885.
56. Sneck, H.J., A survey of gas-lubricated porous bearings, Transactions of the ASME, Series F, Journal
of Lubrication Technology, 89, 1968, 804–809.
57. Kumar, V., Critical overview of porous metal bearings, Wear, 63(2), 1980, 271–287.
46-24 Design for Lubrication and Tribology

58. Prakash, J. and Vij, S.K., Hydrodynamic lubrication of a porous slider, Journal of Mechanical
Engineering Science, 15, 1973, 232–234.
59. Verma, P.D.S., Agrawal, V.K., and Bhatt, S.B., The optimum profile for a porous slider bearing, Wear,
48(1), 1978, 9–14.
60. Kumar, V., Friction of a plane porous slider of optimum profile, Wear, 62, 1980, 417–418.
61. Srinivasan, U., The analysis of a double-layered porous slider bearing, Wear, 42(2), 1977, 205–215.
62. Naduvinamani, N.B., Non-Newtonian effects of second-order fluids on double-layered porous
Rayleigh-step bearing, Fluid Dynamics Research, 21, 1997, 495–507.
63. Patir, N. and Cheng, H.S., An average flow model for determining effects of three-dimensional
roughness on partial hydrodynamic lubrication, Transactions of the ASME, Journal of Tribology,
100, 1978, 12–17.
64. Patir, N. and Cheng, H.S., Application of average flow model to lubrication between rough sliding
surfaces, Transactions of the ASME, Journal of Tribology, 101, 1979, 220–229.
65. Tonder, K., A simplified assessment of the performance of differentially textured hard disk sliders,
Tribology International, 38, 2005, 641–645.
66. de Kraker, A., van Ostayen, R.A.J., and Rixen, D.J., Development of a texture averaged Reynolds
equation, Tribology International, 43, 2010, 2100–2109.
67. Rahmani, R., Mirzaee, I., Shirvani, A., and Shirvani, H., An analytical approach for analysis and
optimization of slider bearings with infinite width parallel textures, Tribology International, 43, 2010,
1551–1565.
68. Etsion, I., State of the art in laser surface texturing, ASME, Journal of Tribology, 127, 2005, 248–253.
47
Compliant Foil Bearing
Technology: An Overview
Nomenclature.............................................................................................. 47-1
47.1 Introduction..................................................................................... 47-3
47.2 Analysis of Gas-Lubricated Foil Journal Bearings..................... 47-4
Compliant Foil Bearing  •  Analysis  •  Nature of
Solution  •  Performance Characteristics
47.3 Advancement in Foil Bearing Design Analysis........................ 47-20
Compliant Foil Bearing Analysis Including Smooth Top
Foil in the Bump-Type Bearings  •  Theoretical Basis and
Program Logic  •  Numerical Prediction of Foil Bearing
Performance  •  Boundary Conditions  •  Method of
Solution  •  Performance Parameters  •  Bearing Stiffness
Coefficient  •  Sample Solutions  •  Parametric Study:
Performance Characteristics  •  Concluding Remarks
Hooshang Heshmat Acknowledgment.......................................................................................47-40
Mohawk Innovative References...................................................................................................47-40
Technology, Inc Bibliography...............................................................................................47-40

Nomenclature
A area (for hybrid foil magnetic bearings)
Aij structural rigidity foil bearing
AMB active magnetic bearing
A1 constant
A1 constant
A 2 constant
B constant
c damping (lbf-s/in.)
c r critical damping (lbf-s/in.)
C radial clearance; sway space
C g gas constant
Cij two-dimensional influence coefficient
Cr bearing radial clearance
CSFB compliant surface foil bearing
D diameter of shaft or bearing
DB effective flexural rigidity of foil
DT flexural rigidity of top smooth foil
e shaft displacement; eccentricity coefficient

47-1
47-2 Design for Lubrication and Tribology

E elastic modulus
EB modulus of elasticity of bump foil (lb/in.2)
F frictional force; FC/(πμωLR 2), dimensionless bearing friction
Fy static force in y direction
gij bearing geometric function
h film thickness; (h/C)
hB bump height (in.)
hN∼ nominal film thickness parameter
H horizontal reaction load (lb/in.); power loss
k stiffness (lbf/in.)
K bearing stiffness; spring coefficient
K B bump foil stiffness (lb/in./in.)
L width of bearing in the z direction; bearing axial length
m mass (lbm)
M, Mo bending moments, lb
N revolutions per unit time
P pressure, unit load (W/LD) in journal bearing
Pa ambient pressure
Po uniform pressure (psi)
ℜij influence factor
r radial coordinate; (r/R 2)
R radius of bearing or journal
R B bump radius; radius of bearing housing
Rj radius of journal
Re Reynolds number, (ρRωh/μ) in journal bearings
s distance between (pitch) backing spring
t time, foil thickness; t/C
tB bump foil thickness (in.)
tS thickness of spring backing
tT top foil thickness (in.)
t1 thickness of top leaf
T bearing torque
T T/PaCR 2
Ta ambient temperature
U linear velocity
vB tangential deflection
w local deflection of top smooth foil (in./in.)
wB local deflection of bump under load WB (in./in.)
W load; W/(PaR 2)
WB unit load at center of bump (lb/in.)
x coordinate along foil (in.); horizontal coordinate
x displacement (in.)
ẋ velocity (in./s)
ẋ˙ acceleration(in./s2)
x,y rectangular coordinates
y vertical coordinate
z axial coordinate
α dimensionless parameter; structural compliancy
αij compliance coefficient of bearing backing spring (two-dimensional)
β angular extent of bearing pad, (θE − θ S)
Compliant Foil Bearing Technology 47-3

δ maximum surface deflection at the midplane of CSFB; logarithmic decrement


δij surface deflection
δx bearing surface deflection in horizontal direction
δy bearing surface deflection in vertical direction
∈, ε shaft displacement to clearance ratio, e/C; eccentricity ratio (e/C)
∈θ tangential strain
γ gamma; root of characteristic equation
Λ [6 μω/Pa](R/C)2, speed parameter
μ viscosity lb-s/in.2; lubricant viscosity; absolute viscosity
η coefficient of friction
νB Poisson’s ratio for bump foil
ϕ attitude angle (θ 0 − π)
ϕ L load angle
σθ tangential stress, lb/in.2
θ angular coordinate
θa attitude angle
θE end of bearing pad
θfs angular position at start of hydrodynamic pressure (deg)
θie end of pad
θis start of pad
θ 0 half angle subtended by bump, radians; angular position of hmin
θs start of bearing pad
θ1 angular position of start of foil; start of hydrodynamic film
θ 2 end of hydrodynamic film
ξ normalized angular extent of bearing pad, (θ/β); nondimensional damping ratio (c/cr)
ω angular velocity
ζ (δ/C)(D/L)

Subscripts
a ambient
E end
max maximum
min minimum
N nominal
r radial
S start

47.1  Introduction
Air-lubricated foil bearings have demonstrated excellent promise for several applications such as air-
craft air cycle machines, compressors, blowers, high-speed motors, gas turbines, turbochargers, and
other high-speed turbomachinery. The foil bearing will deform as hydrodynamic pressure is generated.
They can be constructed to deform in an advantageous manner with respect to the application, whether
it be for high load capability or for high-speed operation. It is also possible to introduce frictional damp-
ing by causing rubbing as deformation takes place and by selecting material combinations that will
optimize the Coulomb damping effect.
Computer codes have been developed and upgraded over the past decade to evaluate the performance
of a gas journal bearing using a spring-supported compliant foil as the bearing surface. The analytical
model considers a journal bearing, consisting of a hollow cylindrical sleeve, the inside of which contains
47-4 Design for Lubrication and Tribology

springs of corrugated or bump foils lined on top with a very thin foil. The foil is anchored at its leading
edge but is free at the trailing end. The bumps act as a spring, and the foil, when loaded, deflects, produc-
ing a film thickness higher than in an equivalent rigid surface.
Since stiffness of the bearing will depend upon both the stiffness of the gas film and the stiffness of
bump foils, the deflection of the bump foil is particularly important when the gas film is stiff compared
with the foil, which is a situation that makes optimum use of the compliance of the bending-dominated
foil bearing. Additional computer-based analyses have been developed to calculate the elastic deforma-
tion (structural stiffness) and performance of bending-dominated compliant foils (bump and smooth
top foils) to the hydrodynamic computer design analysis. The output of these structural codes provides
the appropriate spatial stiffness input.
The computer codes developed by one of the authors, when property coupled and applied, predict the
performance characteristics of gas-lubricated compliant surface foil journal bearings. Performance char-
acteristics include load capacity, film thickness, pressure and surface deformation profile, power require-
ments, and dynamic characteristics in the form of stiffness in two degree of freedom. These performance
characteristics are computed as functions of compliant foils’ structural stiffness and bearing geometry,
loads and load direction, running speed, and fluid viscosity. The following assumptions have been made:
1. The flow is assumed to be laminar, compressible, and isothermal.
2. The gas is assumed to be ideal.
3. The bearing operating condition and flow are steady state, and the bearing stiffness coefficients
are computed with use of small perturbation on a primary steady-state journal rotating frequency
and position.
The aforementioned assumptions still leave the codes applicable to a broad range of applications. A
derivation of the equations governing the performance of gas-lubricated foil journal bearing, along with
description of their solution, is given in the next section. This will be followed by a description of the
computer codes including an input description, sample cases, and comparisons with results of experi-
ments. For more detailed treatment of the subject matter and the design methods employed to integrate
compliant foil bearings into function rotor-bearing systems, the reader is referred to the upcoming
ASME book publication entitled “Foil Bearing Technology—Applications and Integrations.”

47.2  Analysis of Gas-Lubricated Foil Journal Bearings


This chapter is concerned with an evaluation of the performance of a gas journal bearing using a spring-
supported compliant foil as the bearing surface. The wear analysis, conducted for both single and multi-
pad configurations, is concerned with the effects that the various structural, geometric, and operational
variables have on bearing behavior. Following the solution of the relevant differential equation, tabular
or graphical solutions are provided for a range of relevant geometric and operational parameters. The
solutions include values of the collinear and cross-coupled spring coefficients due to both structural and
hydrodynamic stiffness. Desirable design features with regard to start of bearing arc, selection of load
angle, number of pads, and degree of compliance are discussed.

47.2.1  Compliant Foil Bearing


The generalized configuration of a compliant foil journal bearing, shown in Figure 47.1a, consists of a
hollow cylinder, the inside of which contains strips of corrugated or bump foils lined on top with a very
thin foil; the basic element of such a strip is shown in Figure 47.1b. The foil is anchored at its leading edge
but is free at the trailing end. The bumps act as springs, and the foil, when loaded, deflects, producing
(as in any elastohydrodynamic bearing) a film thickness higher than that in an equivalent rigid surface.
The actual appearance of such a bearing is shown in Figure 47.2.
Compliant Foil Bearing Technology 47-5

Bhydro Kstruct

Bstruct
Khydro

Compliant
Hydrodynamic component of
component of stiffness
damping
Shaft
Compliant
component, Hydrodynamic
frictional component of
damping stiffness

(a)

Casing

(b) Telescoped
bump foil

FIGURE 47.1  The compliant foil bearing. (a) Operational concept; (b) foil bearing.

As compared to a conventional gas bearing, the compliant foil bearing has the following advantages:
• Higher load capacity. In terms of the usual load capacity–film thickness relation, the foil bearing
can, for a given nominal film thickness (this term will be discussed more fully later on), carry a
higher load.
• Lower power loss. The existence of larger films is also responsible for yielding lower power losses.
• Journal/runner changes. Due to the ability of the top and bump foils to yield and deflect, the
bearing can tolerate excursions of the runner or misalignment of the mating surfaces. The ability
of the foils to deflect stability and stiffness permits the maintenance of a preselected stiffness by
proper choice and distribution of the bumps, and provides high damping.
• High temperature capability. The bearing exhibits superior qualities in operating at high tempera-
tures, primarily in that it is not subject to thermal distortion.
• Endurance of foreign matter. Due to its higher film thickness and ability to conform, the present
bearing can tolerate the incursion of foreign matter without undue harm to the mating surfaces.
Foil bearings are an attractive alternative to conventional rigid gas bearings or rolling element bearings
due to their capacity to accommodate misalignment, rotor vibrations, shock loading, centrifugal growth,
and elastic and thermal distortions. Because of the aforementioned features and benefits, foil bearings
have been built in various configurations [1–3] and introduced into a wide array of advanced high-speed
47-6 Design for Lubrication and Tribology

Casing ring

Smooth top foil Bump foil

(a)

(b)

FIGURE 47.2  Photographic presentation of basics of foil journal bearing. (a) Basic bearing components; (b)
assembled bearing.

machinery, particularly in the aerospace and energy-related industries. Sample experimental results for
both journal and thrust bearings over a range of operating parameters are given in [4–6]. Some of the
machine experience and experimental results gained with air-lubricated journal and thrust bearings
over a range of operating parameters have provided insight into the structural features of these bearings.

47.2.2  Analysis
47.2.2.1  Assumptions
In formulating the equations governing the elastohydrodynamics of the foil bearing, the following
assumptions are made:
1. The stiffness of the foil is taken to be uniformly distributed and constant throughout the bearing
surface. The stiffness, K B , is linear and is, thus, independent of the amount of bump deflection.
2. The foil is assumed not to “sag” between bumps but rather to follow the deflection of the bumps
themselves.
3. The deflection of the foil in its response to the acting forces is dependent on the local effect only,
that is, on the force acting directly over the particular point.
4. The fluid in the film is isothermal and behaves like a perfect gas.
47.2.2.2  Differential Equation
With the nomenclature of the journal bearing given in Figure 47.3, the relevant Reynolds equation can
be written [7] as
1 ∂ ⎡ 3 ∂p ⎤ ∂ ⎡ 3 ∂p ⎤ 6 mU ∂
R2 ∂q ⎢⎣
ph +
∂q ⎥⎦ ∂z ⎢⎣
ph
∂z ⎥⎦
=
R ∂q
( ph ) (47.1)

Compliant Foil Bearing Technology 47-7

φL= 0
W W

φL φL

θ
θB

θE

β θ
0

C

φ0

θ0

FIGURE 47.3  Nomenclature for foil journal bearing.

Writing

⎛z⎞ ⎛ p ⎞ ⎛h⎞
z =⎜ ⎟ p =⎜ ⎟h =⎜ ⎟
⎝R⎠ p
⎝ ⎠a ⎝C ⎠

and using for Λ the constant


2
6 mUR 6 mw ⎛ R ⎞
Λ= = ⎜ ⎟
PaC 2 Pa ⎝ C ⎠

we obtain

∂ ⎡ 3 ∂p ⎤ ∂ ⎡ 3 ∂p ⎤ ∂
∂q ⎢⎣
ph +
∂q ⎥⎦ ∂z ⎢⎣ ph ∂z ⎥⎦ = Λ ∂q ph ( ) (47.2)

The film thickness variation h(θ) is that due both to eccentricity e and to the deflection of the foil
under the imposed hydrodynamic pressures. Since the latter is proportional to the local pressure,
we have

h = C + e cos(q − fo ) + K1( p − pa )
47-8 Design for Lubrication and Tribology

fF fF
lO

FIGURE 47.4  Configuration of bump foil.

where K1 is constant reflecting the structural rigidity of the bumps. It was shown [8] that this K1 is
given by

⎛ aC ⎞
K1 = ⎜ ⎟
⎝ Pa ⎠

where
3
2 pa S ⎛ l0 ⎞
a=
CE ⎝ t ⎠
(
⎜ ⎟ 1−n
2
) (47.3)

is the compliance of the bearing. The quantities s, l0, and t are defined in Figure 47.4. Consequently, the
normalized film thickness is given by

⎛h⎞
h = ⎜ ⎟ = 1+ ∈cos (q − fo ) + a ( p − 1) (47.4)
⎝C ⎠

47.2.2.3  Boundary Conditions


The construction of foil journal bearings, essentially, does not permit the generation of subambient
pressures. Whenever diverging portions of the film tend to produce subambient pressures in the fluid
film, that is, on top of the foil, the prevailing ambient pressure pa underneath the foil will lift it up until
the pressures on both sides of the foil are equalized. This fact has different implications for the start
and end of the hydrodynamic film. Figure 47.5 shows two relevant cases of the pad having a diverging
film thickness at either the start, case (a), or at the end of the arc, case (b). In the first case, the tendency
of the diverging film to generate negative pressures on top of the foil will lift it off the bumps. The
foil will continue to be lifted off even when, nominally, with respect to the bumps, the film thickness
starts to decrease; the foil will then merely start to approach the bumps again, maintaining a constant
film thickness and constant ambient pressure. When the foil again contacts the bumps, hydrodynamic
pressures will start to form again. The location of recontact θ1 is that point where the film thickness
h1 equals the film thickness at θs, the start of the pad. Thus, the entire portion θs θ1 is ineffective as a
bearing surface.
For case (b), with a diverging film occurring at the trailing portion of the pad, at some point where
negative pressures would have occurred if the surface were rigid, the foil lifts off and aligns itself parallel
Compliant Foil Bearing Technology 47-9

(a)
Shaft U
h = h3 = const.
foil

θs Bumps θmax θ1 θ s
θmax θs

θ1
+ θE
+
θ2
θmin U Foil
(b) Shaft st.
= con
h = h2

Bumps
θmin θ2 θE

FIGURE 47.5  Boundary conditions in foil journal bearing. (a) Commencement of hydrodynamic film; (b) end of
hydrodynamic film.

to the shaft, with h = h2 = constant. In essence, this situation is similar to the trailing edge condition in
cavitating, liquid-lubricated journal bearings. Here, as with cavitation, the film ends at an unknown
angular position θ 2 that, from continuity requirements, must fulfill both the zero pressure and zero
pressure gradient boundary conditions. Thus, the boundary conditions for the solution of Equation
47.2 are

⎛ p⎞
at q = qs p = ⎜ ⎟ =1 (47.5a)
⎝ pa ⎠

⎧ ⎛ p ⎞
⎪p = ⎜ ⎟ = 1

at q = q2 ⎨ ⎝ pa ⎠ (47.5b,c)
⎪ ∂p
⎪ =0
⎩ ∂q

⎛L⎞ ⎛ p ⎞
at z = ± ⎜ ⎟ p = ⎜ ⎟ = 1 (47.5d)
⎝D⎠ ⎝ pa ⎠

47.2.2.4  Performance Parameters


With the solution p(θ, z) accomplished, the performance quantities of the bearing can then be obtained
by proper integration. The load capacity is from

( L /2) q2
− cos q
Fy , x =
∫ ∫ (p − p )
− L / 2 qs
( )
a
sin q
Rdqdz

47-10 Design for Lubrication and Tribology

given in dimensionless form as follows:

+ (L /D ) q2
Fy , x − cos q
Fy , x =
pa R2 ∫( ) ∫ ( p − 1)
− L /D qs
sin q
dqdz (47.6)

The dimensionless load is then given by

w ⎛ L⎞ ⎛ p ⎞
W= 2
= 4 ⎜ ⎟ ⎜ ⎟ = Fy2 + Fx2 (47.7)
pa R ⎝ D ⎠ ⎝ pa ⎠

and the load angle

⎛F ⎞
tan fL = ⎜ x ⎟ (47.8)
⎝ Fy ⎠

The torque on the journal is


(L / 2) qE
⎡ Rh ⎛ ∂p ⎞ mR3w ⎤
T=
∫( ) ∫ ⎢⎣ 2 ⎜⎝ ∂q ⎟⎠ +
− L / 2 qS
h
⎥ dq dz

and in normalized form

(L /D ) qE
T ⎧ h ⎛ ∂p ⎞ Λ 1⎫
T=
paCR2
=
∫( ) ∫ ⎨⎩ 2 ⎜⎝ ∂q ⎟⎠ + 6 h ⎬⎭ dq dz
− L /D qS
(47.9)

Earlier

h = Equation 46.4 for qS 〈q〈q2


h = h2 = constant for q2 〈q〈qE


47.2.2.5  Bearing Stiffness Coefficients


For small displacements from the bearing equilibrium position (e, θ 0), the spring coefficient is given by
the general term

∂F 1 ∂F
K= sin q0 + cos q0
∂e e ∂q0

Properly normalized, the collinear and cross-coupled spring coefficients are thus given by

∂Fx 1 ∂Fx
K xx = sin q0 + cos q0 (47.10)
∂∈ ∈ ∂q0

∂Fx 1 ∂Fx
K xy = − cos q0 + sin q0 (47.11)
∂∈ ∈ ∂q0
Compliant Foil Bearing Technology 47-11

∂Fy 1 ∂Fy
K yx = sin q0 + cos q0 (47.12)
∂∈ ∈ ∂q0

∂Fy 1 ∂Fy
K yy = − cos q0 + sin q0 (47.13)
∂∈ ∈ ∂q0

where
KC
K=
pa R2

47.2.2.6  Method of Solution


For the numerical solution of the Reynolds equation, the dependent variable was represented by a finite
number of points located at intersections of a grid mesh. Particular attention was devoted to regions of
the bearing film in which boundary-layer phenomena may occur at high values of Λ (trailing edges of
pad). Thus, a variable grid in the j direction was employed.
With the grid network, as shown in Figure 47.6, the Reynolds equation (47.2) can be written in finite
difference form as

⎧⎪⎛
3 ∂p ⎞
⎛ ( Δq j + Δq j −1 ) ⎞ ⎫⎪
⎨⎜⎝ ph ⎟⎠ ⎜ ⎟⎬
⎪⎩ ∂z i +1/2 j ⎝ 2 ⎠ ⎪⎭

⎧⎪⎛ ∂p ⎞ ⎫⎪
+ ⎨⎜ ph 3 − Λph ⎟ ( Δz )⎬
⎩⎪
⎝ ∂q ⎠ ij +1/2
⎭⎪

⎪⎧⎛ ∂p ⎞ ⎛ Δq j + Δq j −1 ⎞ ⎪⎫
+ ⎨⎜ − ph 3 ⎟ ⎟⎠ ⎬
⎪⎩ ⎝ ∂z i −1/2 j ⎜⎝
⎠ 2 ⎪⎭

⎧⎪⎛ ∂p ⎞ ⎫⎪
+ ⎨⎜ − ph 3 + Λph ⎟ ( Δz ) = 0 ⎬ (47.14)
⎩⎪
⎝ ∂q ⎠ ij −1/2 ⎪⎭

ri

1 +(L/D) N
M

o i+1,j

o 2 o4 o Ѳj
Z=0 i,j
i,j–1 1 3 i,j+1
o
i–1,j

1 0
–(L/D) β

FIGURE 47.6  Grid network for finite difference solution.


47-12 Design for Lubrication and Tribology

A typical term on the left-hand side of Equation 47.14 is then approximated by central difference for-
mulas as

⎧⎪ 1 ⎛ Δq j + Δq j −1 ⎞ ⎫⎪
⎨ pi +1/2 jhi +1/2 j ( pi +1 j − pij ) ⎜
3
⎟⎬
⎪⎩ Δz ⎝ 2 ⎠ ⎪⎭
Δz
+ pij +1/2hij3+1/2 ( pij +1 − pij )
Δq j

− Λhij +1/2 Δz pij1/2 +  (47.15)


where

hij = C2 + g ij + kij ( pij − 1) (47.16)


Equation 47.16 was then substituted into Equation 47.15, and the difference equation about ij point is

fij ( pi −1 j , pij , pi +1 j , pij −1 , pij +1 ) = 0


which is linearized by the Newton–Raphson method [9] as

∂fi(, j ) (n +1) ∂fi(, j )


n n
∂fi(, nj ) (n +1)
fi(, j ) +
n

∂Pi +1 j
( n
)
pi −1, j − pi(−1), j +
∂Pi , j
(
pi , j − pi(, j) +
n
)
∂Pi +1, j
( )
pi(+1, j) − pi(+1), j +  = 0
n +1 n
(47.17)

where
n is the old value
n+1 is the new value

The Reynolds equation (47.17) can be represented in matrix notation as follows:

[C j ]{Pj } + [E j ]{Pj −1} + [D j ]{Pj +1} = {R j }


j = 1, 2,…, N

and solved by the column method. Earlier, the symbols denote

[C j ] is the tri-diagonal coefficient column of order M

[D j ],[E j ] are the diagonal coefficient matrices of order M

N is the number of columns

M is the number of rows


{Pj } is the vector of jth column of unknowns, [C j ],[E j ],[D j ]

{R j } is the vector of jth column of right-hand sides



Compliant Foil Bearing Technology 47-13

The foil lift off (the boundary conditions given by Equation 47.5) was satisfied by setting negative pres-
sures to zero during the back sweep in the column method. This is the same technique used for satisfy-
ing the cavitation boundary conditions in liquid-lubricated bearings.

47.2.3  Nature of Solution


Several peculiar features of the compliant foil journal bearing should be brought to light because, to
some degree, they are a matter of definition. This pertains first to the notion of film thickness, in terms
of which load capacity is usually defined, as well as the relation between the effective and physical arcs
of a journal-bearing pad.

47.2.3.1  Nominal Film Thickness


In rigid journal bearings, the minimum film thickness is a clear and fixed quantity. It occurs at the line
of centers, and its value is constant across the axial width of the bearing. Also, generally, the film thick-
ness anywhere is constant in the z direction. Since, in our case, pressures cause proportional deflections
of the bearing surface, the film thickness in the interior of the bearing, where pressures are highest, will
be larger than at the edges (z = ± L/2); also, since the maximum larger pressures occur near the line of
centers, the film thickness in the interior of the θ = θ 0 line will not necessarily be the smallest. The situa-
tion is represented in Figure 47.7. The axial film thickness at θ = θ 0 has a value of hmin at the edges (where
p = pa), but only there; elsewhere along θ = θ 0, the film thicknesses are larger than along another angular
position θ = θN, where, because the pressures are lower, the film thickness, on the average, is smaller than
at the line of centers. Figure 47.8 shows a three-dimensional (3D) film thickness plot for a 120° pad in
which, while film thickness at the edge (z = 1/2) is small over most of the pad area, the surface has been
deflected into much larger values of h.
For the purposes of the present discussion, a nominal film thickness hN will be defined as the
minimum film thickness that occurs along the bearing center line, that is, at z = 0. In Figure 47.9, this
central film thickness is plotted for a center line z = 0 at various values of α. While h min for the rigid
case occurs at θ = 180° with increasing values of α, the value of this h min, or our hN, shifts downstream
and increases in value; at α = 5, it is twice the value of the rigid case and has shifted downstream by
nearly 100°. This should be kept in mind later on, when load capacity, that is, the W − hN relation,
is plotted; an increase in load while increasing ∈ may also produce an increase in the nominal film
thickness.

–(L/2) Z=0 Journal (L/2)

h
hθN
hnominal

o

θ=θN

θ=θ0

FIGURE 47.7  Minimum and nominal film thickness.


47-14 Design for Lubrication and Tribology

β = 120°, Є = 0.6, θ0 = 220°


(L/D) = =1, α = 5

(L/2)
Z=
θE

0
Z=

θS

FIGURE 47.8  Film thicknesses in a 120° bearing pad.

1.4

1.2
10
hN
5
1.0
hN
1
0.8
h hN a=0
0.6
hmin
0.4

L/D = 1 Λ =1.0
0.2 θ0 = 180 β =360
z=0 = 0.6

0
40 80 120 160 200 240 280 320 360
θ, degrees

FIGURE 47.9  Location of nominal film thickness.

47.2.3.2  Actual and Effective Bearing Arc


As discussed in Section 47.2.3.1, compliant foil bearings suffer a penalty in their ability to generate
hydrodynamic pressures wherever the pad arc commences in a diverging region. Thus, for example, in
a full bearing that starts at θs = 0, the following two typical cases may arise:
1. θ 0 = 180°: The film is convergent from the start, and pressures commence at θ = 0.
2. θ 0 = 220°: The foil lifts off and remains parallel to the shaft; film convergence and hydrodynamic
pressures do not commence until θ = 80° (twice the value of θ1 in a rigid bearing). This is shown
graphically in Figure 47.10, where there are no pressures over the first 80° of bearing arc.
Compliant Foil Bearing Technology 47-15

L/D = Λ = a = 1
= 0.6 θ0 = 220°

FIGURE 47.10  Pressure profile in a 360° journal bearing.

L/D = Λ = a = 1
= 0.6 θ0 = 220°

FIGURE 47.11  Pressure profile in a three-pad (120°) journal bearing.

From a hydrodynamic standpoint, case (2) is equivalent to a bearing with θs = 80°. Not only is its
physical start at θ = 0 not adding much in the way of extending the pressure profile, but it actually
imposes a penalty. Had the bearing arc started at θ = 40° instead of θ = 0, the pressures would have
started there instead of θ = 80°. A similar situation for the case of a three-pad design is shown in
Figure 47.11, where, due to the ineffectiveness of 80° of the first 120° pad, it contributes little to the
hydrodynamics of the bearings. This effect can be seen in Table 47.1, which shows that by shifting, in
a 360° bearing, the line of centers from 180° to 270°, there was a loss in load capacity of nearly 30%
as well as a reduction in hN.
In designing a foil bearing, if the eccentricity is fixed for the particular application, it is best to start
the bearing at θs = ϕ (for a vertical load); if the eccentricities are liable to vary, some compromise value
of θs > 0 can be chosen.

47.2.4  Performance Characteristics


There are six geometric, structural, and operational parameters relevant to a foil journal bearing. These
are pad angular extent (β), structural compliancy (α), length-to-diameter ratio (L/D), speed param-
eter (Λ), load angle (ϕL), and number of pads. There are also the eccentricity ratio (E) and the attitude
47-16 Design for Lubrication and Tribology

TABLE 47.1  Effect of Load Angle in 360° Bearing (∈ = 0.6;


(L/D) = Λ = α = 1; θE = 360°)
θo ϕL ϕ θ2 P‾max ‾
h ‾   × 102
W T‾ × 10
N

0 — — — 0 0.40 0 —
20° 163.1° 4° 32° 1.018 0.41 1.04 43.9
45° 141.6° 11.6° 74° 1.087 0.48 11.5 38.1
60° 129.8° 20.2° 95° 1.133 0.52 21.4 33.9
90° 107.2° 27.2° 142° 1.201 0.58 40.0 27.2
180° 33.1° 33.1° 240° 1.253 0.62 56.8 24.6
220° −11.9° 28.1° 272° 1.245 0.62 54.9 23.7
245° −41.9° 23.1° 300° 1.228 0.61 50.0 23.7
270° −72.9° 17.1° 325° 1.201 0.58 40.1 25.5

angle ϕ, the latter tied to the load angle ϕL . In the parametric study to follow, a set of standard condi-
tions consisting of

⎛L⎞
⎜ D ⎟ = Λ = a = 1; ∈ = 0.6
⎝ ⎠

will be used, and any parametric variation will commence from this set of reference values.

47.2.4.1  Full Bearing


Table 47.2 gives a detailed listing of the performance of a vertically loaded (ϕL = 0) full 360° bearing as
a function of (L/D), α, and ∈. Note should be taken of the fact that the start of the bearing, that is, θs, is
so chosen as to avoid idle (p = pa) regions at the upstream portion of the bearing. In effect, this requires
θs = ϕ. The case of nonvertically loaded foil bearings ϕL ≠ 0 is given in Table 47.1. Some of the noteworthy
points emerging from these tabulations are the following:
• Effect of α. While in terms of ∈, there is a drastic drop in load capacity with a more compliant
bearing, in terms of hN is actually an increase in load capacity. This is illustrated in Figure 47.12.
At large values of α, α > 10, the load the bearing can support is low due to the fact that the flexible
foil deflects sufficiently to maintain high film thicknesses, even at large eccentricities. Thus, from
a design standpoint, it may be advisable to use high compliance bearings at low loads; high loads,
however, can be supported only with bearings of low values of α. The relation between ∈ and hN
is given in Figure 47.13 where again we see that in highly compliant bearings (particularly at high
(L/D) ratios), an increase in eccentricity may produce an increase in hN, a phenomenon opposite
to rigid bearings, where hmin is the inverse of ∈.
• The variation of torque with α is shown in Figure 47.14, where we see that at high values of α, there
may be a decrease in torque with eccentricity. This, of course, is tied to the fact that, as shown in
Figure 47.12, at high α’s, an increase in ∈ also produces an increase in film thickness.
• Effect of Λ. The performance of a foil bearing as a function of Λ conforms to the familiar pattern of
compressible lubrication. After an initial rise in W ‾ with an increase in Λ, the load capacity, both in
‾ ‾
terms of an increase in W as well as a rise in hN, tends to flatten off and approach an asymptotic value,
as shown in Figure 47.15. The torque, however, rises almost as a linear function of the increase in Λ.
The more compliant bearing shows lower power losses due to the prevailing higher film thicknesses.

47.2.4.2  Multipad Bearing


Two multipad bearings are examined in this section. The three-pad design consists of three 120° arcs;
the five-pad design has five 72° arcs. In each case, the vertical line of symmetry dissects the bottom pad
Compliant Foil Bearing Technology 47-17

TABLE 47.2  Performance of a 360° Journal Bearing (Λ = 1.0; ϕL = 0)


∈ α ϕ = θs θ2 − γ P‾max ‾
h ‾ × 102
W T‾ × 10
N

L/D = 0.5
0.3 0 63.5 81.5 1.046 0.70 4.4 9.38
1 59.0 87.0 1.043 0.72 4.2 9.15
5 48.5 100.5 1.037 0.80 3.2 8.53
10 40.0 114.0 1.025 0.85 2.9 8.09
20 34.0 114.0 1.017 0.94 2.1 7.54
0.6 0 40.0 62.0 1.23 0.40 17.9 15.85
1 36.0 76.0 1.144 0.51 13.7 13.93
5 32.0 97.0 1.073 0.68 8.3 8.19
10 30.0 105.0 1.05 0.795 6.1 7.33
20 27.0 114.0 1.033 0.90 4.2 6.54
0.9 0 12.0 52.0 3.73 0.10 157.3 26.1
1 19.0 71.0 1.33 0.41 34.7 13.8
5 21.0 91.0 1.12 0.64 14.8 9.8
10 21.0 99.0 1.077 0.76 9.8 8.5
20 21.0 108.0 1.048 0.91 6.3 7.3
L/D = 1.0
0.3 0 57.0 97.0 1.137 0.70 27.9 22.7
1 49.0 104.0 1.107 0.77 23.7 21.2
5 36.0 117.0 1.061 0.94 14.8 18.6
10 28.0 120.0 1.041 1.04 10.3 17.5
20 20.0 132.0 1.025 1.14 6.37 16.5
0.6 0 36.0 77.0 1.539 0.40 94.9 31.1
1 33.0 95.0 1.253 0.62 56.8 24.6
5 28.5 112.0 1.114 0.90 28.8 19.1
10 25.0 117.0 1.074 1.055 19.4 16.9
20 20.0 130.0 1.046 1.22 12.2 15.0
0.9 0 13.0 59.0 4.850 0.10 504.5 58.6
1 21.0 86.0 1.434 0.52 102.8 28.1
5 23.0 98.0 1.164 0.86 42.9 19.5
10 21.5 108.5 1.103 1.05 27.8 16.7
20 19.0 127.0 1.063 1.26 17.2 14.4
L/D = 1.5
0.3 0 52.0 103.0 1.218 0.70 70.0 33.4
1 43.0 113.0 1.152 0.82 53.2 30.4
5 29.0 119.0 1.076 1.03 28.5 26.4
10 21.0 141.0 1.048 1.13 18.3 24.8
0.6 0 35.0 88.0 1.731 0.40 208.9 45.6
1 32.0 104.0 1.311 0.68 112.0 34.3
5 26.0 120.0 1.135 1.00 52.8 26.3
10 22.0 137.0 1.086 1.18 34.1 23.1
0.9 0 14.0 68.0 5.300 0.10 898.9 85.1
1 23.0 95.0 1.485 0.56 179.7 39.2
5 23.0 112.0 1.184 0.96 74.2 26.7
10 21.0 122.0 1.116 1.17 47.5 22.8
47-18 Design for Lubrication and Tribology

10 0
0
=1
5 θ0 =180
0
Nos. refer to a
0.5
1 (L/D) = 1.0
2
1.5
1
1.0 5
1
W 0.5
5

5
0.2

0.1

0.05

0.02
0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
hN

FIGURE 47.12  Load capacity of full journal bearing.

1.2

1.1

1.0 5

0.9 5

0.8

0.7
5
hN 1
0.6
1
0.5
α
= 1
0.4 0
=1.0 θ0 =180°
0.3 0.5
(L/D) = 1.0
0.2 1.5

0.1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

‾n for a 360° bearing.


FIGURE 47.13  Relation between ∈ and h
Compliant Foil Bearing Technology 47-19

35
=1.0 θ0 =180°
(L/D) = 1
Nos. refer to a a=0
30

1
25
T × 10

20
5

15

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

FIGURE 47.14  Variation of torque in a full journal bearing.

1
1.8 θ0 =180° 5
(L/D) = 1 = 0.6
W
1.6
hN
T 1
1.4 Nos. refer to a
5
1.2
1
W, hN T × 10–

1.0
1

0.8

0.6
5
0.4

0.2

0
0 5 10
Λ

FIGURE 47.15  Effect of Λ in full bearings.


47-20 Design for Lubrication and Tribology

so that ϕL = 0 represents a load passing through the midpoint of the bottom pad. Tables 47.3 and 47.4
give a spectrum of solutions for the performance of the three-pad bearing, and these results show the
following:
• Variation with load angle. Because of the cyclic nature of this bearing (symmetry for each 120°),
there is much less variation with either W ‾ or T‾ with a shift in load angle. In particular, there is
no acute loss of load capacity when the line of centers passes between pads. The optimum load
angle for α = 1 is ϕL = − 10°; for α = 5, it is ϕL = − 14°. The improvement in load capacity over that of
central loading (ϕL = 0) is of the order of 10%–15%.
• Variation with number of pads. Figure 47.16 shows the variation of one-, two-, and three-pad bear-
ings as a function of load angle. The plot shows clearly a drop in load capacity with the number
of pads, that is, with a drop in extend of bearing arc β. As seen, the optimum for the 360° bearing
occurs at ϕL = 0, at which point the torque also reaches its minimum value. The three-pad bearing,
as was said previously, reaches an optimum at ϕL = − 10°, whereas the five-pad bearing reaches an
optimum at ϕL = − 15°. In Figure 47.17 where the loads for the three bearings are all plotted for a
fixed shaft position, the effect of a shorter β is seen to be less at high values of α than at low ones.
The torque seems to be at a minimum for the three-pad configuration.

47.2.4.3  Stiffness
Table 47.5 gives the values of the four spring coefficients for two values of compliance, the limiting case
of α = 0 and α = 1. The α = 0 case differs from a rigid gas bearing in that the subambient pressures are
eliminated from the pressure profile. A comparative evaluation of the stability characteristics of the
one- and three-pad bearings is, of course, best done in a study of a rotordynamic system, particularly
when the cross-coupling components vary not only in magnitude but also in sign. However, the follow-
ing items can be deduced from the tabulated K data:
• When plotted against W ‾, the Kyy’s are about the same for both designs, whereas the K xx’s are lower
for the three-pad configuration.
• With the more compliant case, the K’s tend to level off with a rise in eccentricity, the values of the
coefficients approaching the structural stiffness of the system.
In general, the advantage of the compliant bearings in the area of stability lies in that levels of stiffness
can be selected by the designer via a proper combination of structural and hydrodynamic stiffnesses.
Thus, instead of making its inertias and supports suit the inherent stiffness of purely hydrodynamic
bearings, the designer may try to tailor and adjust bearing stiffness to the demands of his rotordynamic
system.

47.3  Advancement in Foil Bearing Design Analysis


An essential step toward better understanding of the inner workings of gas-lubricated elastohydrody-
namic foil bearings is the modeling of coupled smooth top foil with the compliant structural elements
in a novel or state-of-the-art foil bearing concept. We have attempted here to show a method of predict-
ing bearing performance and its dynamic behavior by combining the entire compliant foil elements
within the bearings and exploiting the interlocking relationship within the deformable media and fric-
tional elements. Elastic deformation of smooth top foils and bump foils (backing springs) is inextricably
tangled via the interface layers; such layers are either introduce deliberately (solid lubricants such as
coatings) or self-generated layers of fine debris into the contacts. Examples of tribological contacts are
a smooth top foil rubbing against the surface of bump foils, and contact between the bump foils under-
neath surfaces and inner surface of the bearing shell. Accurate modeling is required to understand and
design foil journal bearings to provide not only desired operating stiffnesses but also frictional damping
characteristics of the bearings under various static and dynamic loads. Since, during the rotor/bearing
Compliant Foil Bearing Technology 47-21

TABLE 47.3  Performance of a Three-Pad Bearing


((L/D) = Λ = 1; B = 120°)
c θ0 ϕL ϕ P‾max ‾   × 102
W T‾ × 10
α = 1
0.3 40 −140.0 79.2 1.073 12.1 21.4
210 7.2 37.2 1.072 12.7 21.8
217 −2.3 37.0 1.075 13.7 21.9
220 −2.3 37.7 1.079 14.1 22.0
225 −5.2 39.8 1.082 14.6 22.0
245 −10.0 55.0 1.088 15.0 21.5
275 −17.4 77.6 1.077 12.6 21.4
0.6 29 −145.0 69.2 1.188 24.0 27.8
180 44.1 44.1 1.133 25.2 18.6
208 −2.6 30.6 1.197 37.2 27.7
220 −10.6 29.4 1.215 39.8 28.8
245 −17.0 48.0 1.284 36.9 29.0
270 −25.5 64.5 1.185 28.7 27.4
0.9 15 −145.0 50.9 1.497 59.3 62.9
196 0.0 16.3 1.340 69.5 36.7
210 −10.8 19.2 1.375 76.9 45.7
230 −16.1 33.9 1.572 74.3 71.8
245 −14.7 50.3 1.412 52.4 77.3
260 −26.2 53.8 1.463 55.3 56.4
α = 5
0.3 38 143.0 74.9 1.049 8.01 20.7
210 5.1 35.1 1.038 7.52 21.1
214 0.0 34.3 1.040 7.87 21.1
220 −5.3 34.7 1.043 8.31 21.2
225 −8.6 36.4 1.045 8.60 21.2
245 −13.3 51.7 1.052 9.07 20.8
275 −21.1 73.9 1.050 8.18 20.7
0.6 32 145.0 67.4 1.104 15.7 25.4
180 55.0 55.5 1.048 12.6 16.5
205 3.6 28.6 1.077 16.1 25.3
220 −9.7 30.3 1.103 18.3 26.7
245 −16.3 48.7 1.121 18.4 27.2
270 −23.1 66.9 1.106 15.9 25.4
0.9 25 144.0 61.0 1.178 24.4 46.5
200 3.5 23.5 1.122 25.3 34.4
203 0.0 23.0 1.119 26.3 36.2
210 −6.3 23.7 1.158 28.1 41.6
230 −16.3 33.7 1.198 29.7 67.1
245 −18.9 46.1 1.202 28.2 72.1
260 −21.5 58.5 1.186 25.2 53.7
47-22 Design for Lubrication and Tribology

TABLE 47.4  Mode of Loading of Three-Pad


Bearing ((L/D) = Λ = 1; β = 120° each)
α ∈ ϕ ‾
W ϕL ϕ ‾
W
1 0.3 37.5 13.8 −10 55.0 15.0
1 0.6 28.5 36.0 −10 29.0 39.8
1 0.9 16.0 68.0 −10 18.5 78.0
5 0.3 35.0 7.8 −14 53.0 9.0
5 0.6 30.0 17.0 −14 41.0 18.6
5 0.9 23.5 26.2 −14 30.0 29.8

0.80 L/D = 1 =1
a=1 = 0.6
One pad — 360°
0.70 Three pads — 120°
Five pads — 72°

0.60
hN
0.50

0.40
W
0.30 W

T×10–1
0.20

W
0.10

0
180 140 100 60 20 –20 –60 –80
θL

FIGURE 47.16  Performance of multipad bearings.

operation at various speeds, there exists microslip between the foil elements, the frictional aspect of all
elastic contact regions and elastohydrodynamic components is at work at all times, making this prob-
lem highly nonlinear in several levels. For the purpose of system design, it is, therefore, the designers
attempt to effectively predict their respective values (elastofrictional behavior). It is important to note
here that determination of bearing dynamic performance (K’s and B’s) is far more important than the
determination of its permissible load-carrying capacity. The dynamic characterization of foil bearings
is, therefore, an essential requirement for the system design integration, which will be explored later in
this chapter.

47.3.1 Compliant Foil Bearing Analysis Including Smooth


Top Foil in the Bump-Type Bearings
The foil bearing simulation requires accurate prediction of the foil bearing operating performance and
dynamic characteristics. Therefore, this nonlinear elastohydrodynamic solution is presented here sepa-
rately from the earlier analytical treatments of bearing in this chapter.
In this section, we are concerned with an evaluation of the performance of a gas foil journal bearing
using a backing spring (corrugated strip bump foil) supported, compliant, smooth top foil as the bearing
Compliant Foil Bearing Technology 47-23

(L/D) = 1 Λ=1
θ0 = 180°
θ0 = 220°
0.5 Nos. refer to a
= 0.6

1
0.4

0.3 1
5

0.2 1
1
5 W

0.1
1 3 5
No. of pads

FIGURE 47.17  Relative performance of multipad bearings.

TABLE 47.5  Values of Spring Coefficients ((L/D) = Λ = 1; ϕL = 0)


∈ α ϕ ‾
W K‾xx K‾xy K‾yx K‾yy
β = 360°
0.6 0 35.7 0.951 1.920 −0.125 −2.345 3.237
0.75 0 24.1 1.894 3.416 −1.166 −3.989 8.981
0.9 0 12.8 5.055 7.202 −6.024 −10.151 44.593
0.6 1 32.1 0.568 1.129 0.174 −0.693 1.130
0.75 1 26.3 0.7833 1.231 0.0254 −0.686 1.378
0.9 1 21.4 1.028 1.268 −0.098 −0.627 1.602
Three-Pad-120° design
0.6 0 26.0 0.635 1.123 −0.092 −2.05 2.635
0.75 0 17.4 1.321 2.102 −0.752 −3.710 7.432
0.9 0 8.6 3.695 4.728 −3.344 −8.768 37.103
0.6 1 25.5 0.359 0.5702 0.0451 −0.758 0.801
0.75 1 20.5 0.511 0.673 −0.017 −0.821 1.051
0.9 1 16.3 0.689 0.759 −0.057 −0.855 1.274

surface. The configuration of such foil journal bearings, shown in Figure 47.18, consists of a bearing cas-
ing, the inside of which contains strips of corrugated bump foils or backing springs lined on top with a
thin foil (smooth top foil). The basic element of such a strip is shown in Figure 47.19. The foil is anchored
at its trailing edge but is free at the leading end. The corrugated strip foil acts as backing springs, and
the top foil, when loaded, deflects, producing a conformal foil as in any elastohydrodynamic bearing.
47-24 Design for Lubrication and Tribology

Journal foil bearing nomenclature

θ
θs θe

Backing spring Y Corrugated


corrugated strip strip fixed
free end end

θ1

ω
X
O e

O
φ0 θ2
Smooth top foil

Bearing
casing

FIGURE 47.18  Foil bearing assembly.

Foil bearing compliant support nomenclature

Foil loading

Smooth top foil


Undeflected
backing
spring W1 W2 W3 Deflected
backing spring

S Segment

Bump 1 Bump 2 Bump 3

FIGURE 47.19  Foil bump strips.


Compliant Foil Bearing Technology 47-25

It must be stressed that in analyzing foil bearings, static force displacements play an important
role since the compliance of the bearing surface governs the operational characteristics. The efforts
to analyze foil journal bearings required the formation of specialized nonlinear elasticity solutions.
Techniques that aimed to find the best groupings, in a global sense, were conceptualized with the aid
and use of prior knowledge. Hence, a novel program has been developed and forms the basis of a coupled
elastohydrodynamic solution to compliant surface foil bearing (CSFB).

47.3.2  Theoretical Basis and Program Logic


Unlike conventional approaches, the solution of the governing hydrodynamic equations dealing with
compressible fluid is coupled with the structural resiliency of the bearing surfaces. The solution to the
structural resiliency is provided in two stages, as presented in the flow diagram of Figure 47.20.
The first-level elasticity solutions deal with compliant elements (bump foil/backing spring) system
stiffness matrices assuming (a) the smooth foil follows the deflection of the backing springs, (b) the
smooth top foil surface will not follow the indentation of the corrugated backing spring, and (c) the
deflection of the top foil and backing springs in responding to the acting hydrodynamic pressure is
dependent on local effects only.
Based on past investigation of corrugated bump foil strips, it is shown that the bumps near the fixed
end have a higher stiffness. In some cases, the stiffness near the fixed end can be two to three times
higher than near the free end. This phenomenon is related to interacting forces that exist in a corrugated
foil strip. A comprehensive computer program was developed to compute the stiffness, deflection, dis-
placement, and reacting forces, including friction forces for each bump for various geometries and load
distribution.
A sample solution for a corrugated foil consists of 10 bumps given in Table 47.6. The compliance of an
individual bump is represented in dimensionless form

C
an =

( )
Pa [wn /dn ]

where
α is the compliancy factor
Subscript n refers to the bump number from the fixed end
w designates load at the bump per unit transverse
δ is respective bump deflection

It is apparent from Table 47.6 that this nonuniform stiffness distribution (in θ direction) contributes
greatly to the nonlinearity of stiffness characteristic of foil bearings.
Given the characteristics of this type of bearing and because bearing surface displacement (defor-
mation) is usually much greater than the operating fluid film thickness, the direct iteration technique
between governing equations leads to an inherent divergence problem. Here we have refined the numer-
ical iterative scheme and provided solutions for a number of operational parameters. A simplified flow
diagram representing program logic is presented in Figure 47.20.
In this scheme, the bearing’s backing springs were modeled, and a resiliency matrix αij (i in θ and j in
z direction) was determined. The film thickness in line with the deflection of the bearing surface, uti-
lizing αij, which was substituted into the Reynolds equation, was then solved by the finite difference
method. The first-level computation provided hydrodynamic pressures, which were obtained from the
solution of the modified Reynolds equation. Hydrodynamic pressures were then substituted into the
second-level elasticity equation to solve for the bearing surface (smooth top foil) deformation and film
thickness. The second-level elasticity analysis was via a finite element (FE) elasticity routine modeling
47-26 Design for Lubrication and Tribology

Compliant surface foil bearing program logic

1 Initial foil bearing Boundary and


geometry: [gij]int operating conditions

2 Calculate structural
element compliance:
backing spring [αij]

Apply local pressure-deflection relationship


3
hij  [gij]Int + [αij] [Pij]

4 Hydrodynamic analysis:
linearized finite difference with variable grid

Pressure field 5

[Pij]Old
9 6
Elasticity program
New pressure distribution Combine top foil and backing springs
[Pij]New Compute bearing surface deflections
[αij] as a function of [Pij]Old
Generate influence coefficients [Cij]
10 A
7
Accuracy check Modify bearing geometry:
[gij]Int [gij]New
12
8
Compliancy Determine film thickness:
check g New + [Cij][Pij]
hij  [ ij]
10 B
Pressure convergence check 11 B
Top foil influence factor:
[Pij]New – [Pij]Old
[Âij] = [αij]/[Cij] Final

Final film thickness


W, Є, and φ 11 A

FIGURE 47.20  CSFB program logic.

TABLE 47.6  Compliance as a Function of Number of


Bumps under Uniform Load Distribution
Bump Number
α‾n 1 2 3 4 5 6 7 8 9 10
0.2 0.6 1.0 2 5 8 11 14 17 20
Compliant Foil Bearing Technology 47-27

the smooth top foil along with backing springs as a thin elastic foil on an elastic foundation. The previ-
ously determined backing spring coefficients (i.e., 1/αij) were included in the model to suit requirements
for boundary conditions of FE analysis. A direct iterative scheme was used to solve these equations, and
the approach was extremely successful since elastic contribution of top smooth foil compared with the
backing spring, in a global sense, was found to be relatively small but not ineffectual. Bearing perfor-
mance was chiefly dominated by backing springs and an intermediate solution, as portrayed in Figure
47.20, boxes 1 through 4. However, the effect of top foil deformation on bearing dynamic coefficients, K’s
and B’s, was not negligible in order to not account for total effort (i.e., the top smooth foil contribution
was about 10%–30%, respectively).

47.3.3  Numerical Prediction of Foil Bearing Performance


With the nomenclature of CSFBs, as given in Figures 47.18 and 47.19, the two-dimensional (2D) com-
pressible Reynolds equation can be written as

1 ∂ ⎡ 3 ∂p ⎤ ∂ ⎡ 3 ∂p ⎤ ∂rh
hr + h r ⎥ = 6 mw (47.18)
R2 ∂q ⎢⎣ ∂q ⎥⎦ ∂z ⎢⎣ ∂z ⎦ ∂q

However, the density is related to the pressure by an equation of state, generally of the form

p = C g rn (47.19)

where
Cg is a constant
n = 1 for isothermal conditions
n = γ = Cp/Cv for adiabatic conditions

The usual situation with a gas bearing is approximated by the isothermal case, so it will be assumed in
further analysis that p = ρCg and considering the following nondimensional parameters:
2
⎛ Z⎞ ⎛ p⎞ ⎛ h⎞ 6mw ⎛ R⎞
Z =⎜ ⎟; p =⎜ ⎟; h =⎜ ⎟; Λ = ⎜⎝ ⎟⎠
⎝ R⎠ ⎝ pa ⎠ ⎝ C ⎠ Pa C

Normalizing the Reynolds equation (47.18), we obtain

∂ ⎡ ∂p ⎤ ∂ ⎡ 3 ∂p ⎤ ∂
∂q ⎢⎣
ph3 +
∂q ⎥⎦ ∂z
( )
⎢⎣ ph ∂z ⎥⎦ + Λ ∂q ph (47.20)

Formulating the equations governing the elastohydrodynamics of the CSFB, the following assumptions
are made for the reason of (a) rapid numerical convergence and (b) coupled auxiliary elasticity equa-
tions dealing with the structural resiliency of the bearing’s backing springs. Therefore, as a first step
in computation, the effect of top smooth foil supported on an elastic foundation is neglected. In other
words, bending and membrane stresses on the foil at the first-level numerical scheme (see Figure 47.20)
are neglected. Thus, the deflection of the foils, in their responses to the acting forces, is dependent on the
local effect only, that is, on the force acting directly over the particular point. Furthermore, the stiffness
of the backing spring is taken to be nonuniformly distributed and dependent on the angular location of
the bump foil strip, as discussed in the preceding section.
47-28 Design for Lubrication and Tribology

The film thickness variation hij is due both to bearing initial geometry [gij]int (eccentricity e and bear-
ing initial surface shape) and to the deflection of the compliant surface under imposed hydrodynamic
pressure. Since the latter is proportional to the local pressure, we have

hij = [ g ij ]Int + [a ij ][ pij − pa ] (47.21)


Considering the initial configuration of CSFB with a single cylindrical top foil but disrupted at θ = 0
(Figure 47.19), then

[ g ij ]Int = C + e cos(q − Φ0 ) (47.22)


where normalized αij

a ijC
a ij = (47.23)
pa

reflects the structural compliancy coefficient of the backing spring. Consequently, the normalized film
thickness is given by

⎛ hij ⎞
hij = ⎜ ⎟ = 1 + ∈cos(q − f0) + a ij[ pij − 1] (47.24)
⎝C⎠

47.3.4  Boundary Conditions


Regarding the boundary conditions for the solution of Equation 47.20, the first-level elasticity solu-
tions are similar to the aforementioned assumptions and equations, as was stated in Section 47.2.2.3 via
Equations 47.5a and 47.5b.

47.3.5  Method of Solution


The range of analytical parameters and design variables involved in the present method of solution is
fairly wide. The overall program, therefore, involved a number of computer codes, some of which had to
be iterated on in order to arrive at integrated results for the bearing system. A flow diagram, shown in
Figure 47.20, presents the CSFB program logic diagram and iteration loop.
For the numerical solution of Reynolds equation (47.20), the dependent variable was represented by
a finite number of points located at intersections of a grid mesh. By substituting Equation 47.24 into
Equation 47.20 and with the grid network, the Reynolds equation (47.20) was written in finite differ-
ence form. The resultant form was linearized by the Newton–Raphson method, represented in matrix
notation and solved by the column method (Section 47.2.2.6) and [10,11]. The boundary conditions were
satisfied by setting negative pressures to 1 during the backsweep in the column method.
The resultant pressure field [pij] was then taken to the next level of elasticity analysis, box #6 in Figure
47.20, to include the effect of the top foil in the CSFB analysis. In this higher level of elasticity analysis,
the mathematical model of the foil bearing consisted of a staggered thin top foil supported via backing
springs (αij) subjected to the resultant pressure field, pij.
The combined top foil and backing spring deflection was computed for the given pij and resulted in a
deflection matrix, δij. The overall structural stiffness is given by

[Pij ]
[ Aij ] = (47.25)
[dij ]

Compliant Foil Bearing Technology 47-29

where Aij is reflecting the structural rigidity of the top foil and backing spring. The new normalized
influence coefficient can be written as

C
Cij = (47.26)
Pa[ Aij ]

The deflected shape of the foil (Figure 47.27), along with the C‾ij, was used for the film thickness relation-
ship; we have

[hij ]new = [ g ij ]new + [Cij ][Pij −1] (47.27)


where

1+ ∈cos(q − f0) + [dij ]


[ g ij ]new =
C

The new film thickness relationship, Equation 47.27, was substituted in Reynolds equation (47.20).
Referring to the diagram of Figure 47.20, iteration between elasticity analysis and governing hydrody-
namic analysis required few iterations to achieve final film thickness. The performance parameters as
well as membrane stress contours were obtained when numerical convergence based on error criterions
was satisfied.

47.3.6  Performance Parameters


With the solution of p(θ, z) accomplished, the performance quantities of the bearing can then be
obtained by proper integration, as was demonstrated in Section 47.2.2.4.

47.3.7  Bearing Stiffness Coefficient


For small displacements from the bearing equilibrium position (e, θ 0), the stiffness coefficient is given by
the general term, as was demonstrated in Section 47.2.2.5.

47.3.8  Sample Solutions


To demonstrate the aforementioned analytical model and numerical procedure, the following example
problem is used. The journal foil bearing parameters are D = 100 mm, L = 75 mm, C = 0.1 mm, single-
pad configuration with an operating condition: Pa = 100 kPa, μ = 0.02 cp, and ∈ = 0.6, n = 10,000 rpm,
t = 0.15 mm. For simplicity, αij matrix is developed from data given in Table 47.6. The results of the
intermediate solution, which is neglecting bending and membrane stresses on the smooth top foil, are
presented in Figures 47.21 through 47.23. Figure 47.21 is composite plots of hydrodynamic pressure, film
thickness, and deflections taken at the bearing center line (z‾ = 0) along the normalized angular extent
of bearing pad, (ξ = θ/β), with an additional film thickness profile at the edge of the bearing (z‾ = L/D).
Corresponding pressure and film thickness profiles are given in Figures 47.22 and 47.23, respectively.
The resultant film and pressure profiles are the inputs to our second-level elasticity equation, as called
for in box #6 of the flow diagram (Figure 47.20). For an aligned journal-bearing condition, it is sufficient
to model only one-half of the bearing (symmetric about the midplane, ‾z  = 0). The top foil (membrane)
grid network was prepared for elasticity analysis via FE, as shown in Figure 47.24. The top foil is con-
sidered symmetric about the bearing center line (i.e., ‾z  = 0), and relevant boundary conditions, such
47-30 Design for Lubrication and Tribology

First order elasticity analysis


1.5

1.2 Pressure at Z = 0

0.9
Film thickness

h, δ, P
at Z = 0
0.6

at ± (L/D)
0.3
Deflection Z = 0

0.0
0.0 0.2 0.4 0.6 0.8 1.0
Extent of bearing pad (ξ)

FIGURE 47.21  Film thickness, deflection, and pressure.

First order elasticity analysis


1.2
L/D = 0.75
= 0.6
1.0 Pressure at Z = 0
Λ = 1.25
0.2
0.7
0.3
P/Pmax

0.5
0.4

0.2

Z = L/D
0.0
0.0 0.2 0.4 0.6 0.8 1.0
Extent of bearing pad (ξ)

FIGURE 47.22  Foil journal-bearing pressure profile.

First order elasticity analysis


1.2
L/D = 0.75
= 0.6
1.0 Λ = 1.75
Film thickness, (h/C)

@ Z=0
0.8

@ Z = 0.2
0.6 @ Z = 0.3
@ Z = 0.4
@ Z = L/D
0.4
0.1 0.3 0.5 0.7 0.9
Extent of bearing pad (ξ )

FIGURE 47.23  Film thickness profile.


Compliant Foil Bearing Technology 47-31

Top foil
fixed end θE
θs

Symmetry
boundary
ω
Z=0
Z = L/D Mid plane of
top foil

FIGURE 47.24  Top foil FEA mesh and boundary conditions.

Trailing
Leading end
end
ω

Z=0

Z = L/D

FIGURE 47.25  Pressure profile applied to top foil.

as leading edge of the foil, are free, and its trailing end that is anchored to the bearing housing (fixed
end) was also applied. Not shown in this figure are distributed backing springs αij at the nodal points.
Figure 47.25 depicts the top foil model with applied hydrodynamic pressure vectors taken from the
computed pressure profile, shown in Figure 47.22. The FE analysis of the compliant structural elements
(top smooth and backing spring system) was conducted with the input variables mentioned earlier.
Intermediate results are plotted in Figures 47.26 through 47.30. Figures 47.26 and 47.27 provide 3D
relative deflection of the top foil. Figure 47.26 shows the top foil displacement/deflection contour, while
Figure 47.27 presents an end view of the initial and new geometry of the top foils. Note that the object
is viewed from the midplane of the foil bearing in Figure 47.27. Figures 47.28 through 47.30 provide
highlighted stress contours in the top smooth foil.

Displacement
0.001310
Trailing 0.001180
Leading end 0.001050
0.000919
end 0.000788
0.000656
ω 0.000525
0.001310 1.0E–016 0.000394
0.000263
0.000131
Z=0 1.0E–016

Z = L/D

0.000919

FIGURE 47.26  Top foil displacement contour under hydrodynamic pressure.


47-32 Design for Lubrication and Tribology

Initial surface
profile ω
[gij] int

Smooth foil
deflection
[gij]new
FIGURE 47.27  Top foil deflections.

σx
Trailing 728.000000
403.000000
Leading end 250.000000
10.8000000
–228.00000
end –487.00000
y ω –706.00000
–945.00000
–1.18E+003
–1.42E+003
Z=0 –1.66E+003

Z = L/D

FIGURE 47.28  Stress contour in top foil, σx .

σz
2.10E + 003
Trailing 1.83E + 003
1.56E + 003
Leading end 1.29E + 003
1.02E + 003
end 755.000
486.000
ω 217.000
–52.300
–321.000
–590.000
Z=0
σZ = –321.000
Z = L/D

σZ = 1.83E + 003

FIGURE 47.29  Stress contour in top foil, σz .

Utilizing Equations 47.25 through 47.27, and iterating and following numerical scheme sequences,
as numbered in Figure 47.20, after several passes, pressure convergence was achieved. Using the
relationship

[a ij ]
ℜij = (47.28)
[Cij ] final

Compliant Foil Bearing Technology 47-33

Von mises
3.29E + 003
Trailing 2.96E + 003
2.63E + 003
Leading end 2.30E + 003
1.98E + 003
end 1.65E + 003
1.32E + 003
ω 991.00000
662.00000
5.0000000 334.00000
Z=0 5.0000000

Z = L/D

3.29E + 003

FIGURE 47.30  Top foil von Mises stress contour.

Second order elasticity analysis


2.0
L/D = 0.75
= 0.6
1.6
Influence factor Âij = αij/Cij

Λ = 1.25
@Z=0
1.2
@ Z = 0.3
0.8 @ Z = 0.4

0.4
@ Z = L/D

0.0
0.0 0.2 0.4 0.6 0.8 1.0
Extent of bearing pad (ξ)

FIGURE 47.31  Top smooth foil influence factors.

at the end of successful completion of the computation, ℜij , an influence factor was obtained (Figure
47.20, box 11B), which is a measure of the top foil’s membrane effect. The magnitude of the influence fac-
tor, ℜij , can be interpreted as a stiffening or softening effect of the top smooth foil due to the membrane
effect on the overall structural compliancy of the bearing. Figure 47.31 shows a composite plot of ℜij as
a function extent of the large bearing pad for four different values of ‾z . As can be seen from the plots of
Figure 47.31, the stiffening effect of the top foil is greater at the middle of the bearing than at its edge by
at least 30%. In fact, the top foil appears to be having a softening effect at the bearing surface near the
edge (about 40%–50%).
Furthermore, in the top foil near the trailing end where hydrodynamic pressure approaches zero, ℜij
also approaches zero. These data (ℜij ) were stored and utilized to reduce a computation time via bypass-
ing second-level elasticity analysis for similar top foil thickness and dimensions.

47.3.9  Parametric Study: Performance Characteristics


In the parametric study to follow, a set of standard conditions of L/D = 0.75 and β = 350° with variable α
given in the Table 47.6 and t‾ = 3 will be used, and any parametric variation will commence from this set
of reference values. Due to the nature of CSFB, unlike rigid journal bearings, the film thickness varies in
47-34 Design for Lubrication and Tribology

Film thickness distribution


1.5
L/D = 0.75
= 0.9
1.2

Dimensionaless film, h
0.9

0.6 hN
Λ = 1.25

0.3 Λ = 0.25
Λ = 0.125
0.0
70 90 110 130 150 170 190 210 230 250 270
Effective bearing arc, θ (deg)

FIGURE 47.32  Location of minimum film thickness.

the z direction. Since pressures cause proportional deflections of the bearing surface, the film thickness
in the interior of the bearing, where pressures are highest, will be larger than at the edges (z = ±L/D);
also, since the maximum pressures occur near the line of center (z = 0), the film thickness in the inte-
rior of the θ = ϕ 0 line will not necessarily be the smallest. The situation is represented in Figure 47.32.
Therefore, a nominal film thickness, hN, will be defined as the minimum film thickness that occurs
along the bearing center line, that is, at z = 0. In Figure 47.32, this central film thickness is plotted for
a center line z = 0 at various values of Λ. This value of hN shifts downstream and increases in value by
increasing Λ. This should be kept in mind when hN values are plotted as a function of ∈ and Λ. Similar
plots were generated for pressure profiles, Figure 47.33, where the apex of the pressures slightly shifts
downstream as a function of Λ.
Table 47.7 gives a detailed listing of the performance of a vertically loaded, single-pad bearing as a
function of Λ and ∈. Some of the noteworthy points emerging from these tabulations are plotted in
Figures 47.34 through 47.43. The operating W ‾ parameters, ϕ, h‾ N, f‾ , ΔPmax, ζ, and K’s are given as a func-
tion of ∈ and Λ in Figures 47.34 through 47.43. Note that the rate of deflection ζ is taken at the center line
of the bearing where hN occurs. In comparing plots of Figure 47.36 with Figure 47.37, it can be concluded
that ζ is inversely proportional to hN.

47.3.10  Concluding Remarks


An analytical model of bending-dominated, compliant-surface, hydrodynamic foil gas journal bear-
ings was developed to account for the 3D variation of structural stiffness in r, θ, and the z direction,
and to provide a solution to investigate the effects of top smooth foil on the bearing performance. The
3D distribution of the compressible fluid film thickness and 2D pressure, as well as the shear losses in
foil bearings of finite length, was computed in line with structural elasticity of the bearing compliant
surfaces. Structural stiffness characteristics of the foil bearing’s backing spring (bump foils) and mem-
brane effects of the top foil were included and integrated in the numerical scheme and were dealt in two
levels of iteration. As for the first-level numerical analysis, deformation of the bearing surface without
the effect of the top foil was formulated in relationship to the pressure. In line with the deflection of the
bearing top foil surface, the film thickness was substituted into the Reynolds equation and solved using
the finite difference method. This technique provided the advantages of utilizing variable structural
Compliant Foil Bearing Technology 47-35

10
8

Length
6
4
2
0
15
10
5
Pressure (psi)

0
–5
–10
–15
–20
–25
–30
0 5 10 15 20
(a) –20 –15 –10 –5

Foil bearing pressure profiles


2.0
L/D = 0.75
= 0.9
1.8
Dimensionless pressure, ρ

Λ = 1.25
1.6
0.63

1.4
Λ = 0.25

1.2 Λ = 0.125

1.0
50 75 100 125 150 175 200 225 250 275
(b) Effective bearing arc, θ (deg)

FIGURE 47.33  Pressure versus bearing number (Λ). (a) Third iteration, fully developed EHD pressure profile
view from axial groove; (b) foil bearings pressure at journal-bearing midplane (Λ).

stiffness characteristics of the backing spring, numerical stability, and rapid convergence in obtained
intermediate solutions.
Using the FE method, the membrane effect of the elastic top foil was included in the overall analysis.
Further, the combined smooth top foil and elastic foundation was formulated as an influence coefficient
to accelerate the numerical solution. The overall program logic proved to be an efficient and sound tech-
nique to handle the complex structural compliances of various foil bearings.
Parametric analysis was conducted to establish tabulated data for use in the design and applica-
tion of a foil bearing/rotor system. The technique reported herein can readily be adapted to deal
with various foil bearing configurations and can evaluate the performance of innovative foil bearing
designs.
47-36 Design for Lubrication and Tribology

TABLE 47.7  Nondimensional; Parametric Study of IR&D Foil


Magnetic Bearing: L × D × C = 3 in. × 4 in.
∈ ‾  × 103
W ϕ F‾ hN ζ P‾max
Λ = 0.125 0.200 0.200 69 1.889 0.800 0.007 0.095
0.400 0.400 59 2.262 0.605 0.019 0.266
0.600 0.600 46 3.092 0.410 0.021 0.707
0.800 0.800 28 4.928 0.245 0.057 2.210
0.950 0.950 16 6.907 0.176 0.123 4.750
Λ = 0.25 0.200 0.200 68 1.889 0.800 0.008 0.190
0.400 0.400 58 2.259 0.606 0.021 0.530
0.600 0.600 43 3.009 0.421 0.035 1.360
0.800 0.800 26 4.360 0.283 0.091 3.550
0.950 0.950 16 5.375 0.221 0.163 6.310
Λ = 0.63 0.200 0.200 65 1.881 0.805 0.013 0.4730
0.400 0.400 53 2.219 0.618 0.033 1.290
0.600 0.600 37 2.821 0.461 0.075 2.950
0.800 0.800 23 3.500 0.354 0.151 5.910
0.950 0.950 15 4.084 0.296 0.229 8.870
Λ = 0.88 0.200 0.200 63 1.875 0.808 0.018 0.660
0.400 0.400 50 2.195 0.629 0.045 1.750
0.600 0.600 35 2.742 0.485 0.096 3.760
0.800 0.800 22 3.280 0.385 0.178 6.950
0.950 0.950 15 3.679 0.329 0.257 0.960
Λ = 1.25 0.200 0.200 60 1.865 0.813 0.025 0.920
0.400 0.400 46 2.158 0.645 0.060 0.236
0.600 0.600 32 2.585 0.515 0.121 4.740
0.800 0.800 20 3.032 0.421 0.209 8.140
0.950 0.950 15 3.297 0.365 0.289 11.200

1.5
L/D = 0.75
Nos.refer to Λ 1.25
1.2
0.88
Dimensionless load, W

0.63
0.9
0.38
0.6 0.25
0.125
0.3

0.0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Eccentricity ratio ( )

FIGURE 47.34  Load capacity versus eccentricity ratio.


Compliant Foil Bearing Technology 47-37

80
L/D = 0.75
70 Nos. refer to Λ

60

Attitude angle, (φ0)


1.25
50 0.88
40 0.63
0.25
0.38
30 0.125

20

10
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Eccentricity ratio ( )

FIGURE 47.35  Attitude angle versus eccentricity ratio.

1.0
L/D = 0.75
0.9
Nos. refer to Λ
0.8
0.7
Nominal film, hN

0.6
0.63
0.5
0.88
0.4 1.25

0.3
0.38
0.2 0.25
0.1 0.125

0.0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Eccentricity ratio ( )

FIGURE 47.36  Nominal film versus eccentricity ratio.

0.4
L/D = 0.75
Nos. refer to Λ
Deflection rate, ξ = δD/CL

0.3 1.25
0.88
0.63
0.2 0.38
0.25
0.125
0.1

0.0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Eccentricity ratio ( )

FIGURE 47.37  Relative deflection versus eccentricity ratio.


47-38 Design for Lubrication and Tribology

8
L/D = 0.75
7 1.25
Nos. refer to Λ

Dimensionless friction, F
6
0.88
5
0.63
5 0.38
4 0.25

3 0.125

1
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Eccentricity ratio ( )

FIGURE 47.38  Friction force versus eccentricity ratio.

12
L/D = 0.75
1.25
10 Nos. refer to Λ
0.88
Pressure, ∆Pmax (Psi)

8 0.63
0.38
6
0.25
4
0.125

0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Eccentricity ratio ( )

FIGURE 47.39  Peak pressure versus eccentricity ratio.

1.6

1.4 L/D = 0.75 1.25


Dimensionless stiffness, Kxx

Nos. refer to Λ 0.88


1.2
0.63
1.0
0.38
0.8
0.25
0.6
0.125
0.4

0.2

0.0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Eccentricity ratio ( )

FIGURE 47.40  Stiffness K xx versus eccentricity ratio.


Compliant Foil Bearing Technology 47-39

–0.0
1.25 L/D = 0.75
Nos. refer to Λ

Dimensionless stiffness, Kxx


–0.2 0.88

0.63
–0.4

0.125 0.38
0.25
–0.6

–0.8

–1.0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Eccentricity ratio ( )

FIGURE 47.41  Stiffness Kyx versus eccentricity ratio.

0.4
1.25 L/D = 0.75
0.3 Nos. refer to Λ
Dimensionless stiffness, Kxx

0.88
0.2 0.63
0.38
0.1 0.25
0.125
0.0

–0.1

–0.2

–0.3

–0.4
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Eccentricity ratio ( )

FIGURE 47.42  Stiffness K xy versus eccentricity ratio.

3.0
L/D = 0.75
Nos. refer to Λ
Dimensionless stiffness, Kyy

2.5

2.0
1.25
1.5 0.88
0.38
1.0 0.63

0.25
0.5
0.125
0.0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Eccentricity ratio ( )

FIGURE 47.43  Stiffness Kyy versus eccentricity ratio.


47-40 Design for Lubrication and Tribology

Acknowledgment
The author is grateful to Mohawk Innovative Technology, Inc., for its support of the work referenced in
this chapter.

References
1. Heshmat, H., Multi-stage support element compliant hydrodynamic bearing, U.S. Patent Number
4,300,806, November 17, 1981.
2. Silver, A. and Friedericy, J.A., Foil bearing arrangements, U.S. Patent Number 3,893,733, July 8,
1975.
3. Silver, A. and Wenban, J.R., Foil bearing, U.S. Patent Number 4,178,046, December 11, 1979.
4. Heshmat, H., Advancements in the performance of aerodynamic foil journal bearings: High speed
and load capability, ASME Paper No. 93-Trib-32, Accepted for Publication by ASME Trans. J. Tribol.,
and Presentation at the STLE/ASME Tribology Conference, October 24–27, 1993, New Orleans, LA.
5. Heshmat, H. and Ku, C.-P.R., Structural damping of self-acting compliant foil journal bearings,
ASME Trans. J. Tribiol., 116(1), 1994, 76–82.
6. Chen, S.H. and Arora, G., Development of a 1000F, 10,000 start/stop cycle foil journal bearing for
a gas turbine, Presented at AIAA/ASME/SAE/ASEE, 22nd Joint Propulsion Conference, June 16–18,
1986, Huntsville, AL, Paper No. 86–1457.
7. Walowit, J.A. and Arno, J.N., Modern Developments in Lubrication Mechanics, Applied Science
Publishers, Ltd., London, U.K., 1975.
8. Heshmat, H., Walowit, J.A., and Pinkus, O., Analysis of compliant foil gas thrust bearings, to be
published.
9. Castelli, V., and Pirvics, J., Review of numerical methods in gas bearing film analysis, Trans. ASME J.
Lubr. Technol, 90, 1968, 777–792. 1968.
10. Heshmat, H., Analysis of compliant foil bearings with spatially variable stiffness, Presented at AIAA/
SAE/ASME/ASEE 27th Joint Propulsion Conference, June 24–26, 1991, Sacramento, CA, Paper No.
AIAA-91-2102.
11. Heshmat, C.A. and Heshmat H., An analysis of gas lubricated, multi-leaf foil journal bearings with
backing springs, ASME Paper No. 94-Trib-61, Accepted for Publication in the ASME Trans. J.
Tribol., and Presented at the ASME/STLE Tribology Conference, October 16–19, 1994, Lahaina, HI.
12. Heshmat, H., High load capacity compliant foil hydrodynamic thrust bearing, U.S. Patent Number
5,961,217, July 30, 1998.

Bibliography
Heshmat, H., Walowit, J., and Pinkus, O., Analysis of gas-lubricated compliant journal bearings, J. Lubr.
Technol. Trans. ASME, 105(4), 1983, 647–655.
Agarwal, G.L., 1990, Foil gas bearing for turbomachinery, SAE Paper No. 901236.
Heshmat, H., Role of compliant foil bearings in advancement and development of high-speed tur-
bomachinery, Presented at the Second ASME Pumping Machinery Symposium, Fluid Engineering
Conference, June 20–24, 1993, Publication of ASME Pumping Machinery, Ed. P. Cooper, FED—Vol.
154, pp. 359–377.
Suriano, F.J., Gas-lubricated turbine end foil bearing development program, Technical Report,
AFWAL-TR-81-2119, AiResearch Manufacturing Co. of Phoenix, Phoenix, AZ, December 1981.
Suriano, F.J., Dayton, R., and Woessner, F., 1993, Test experience with turbine-end foil bearing equipped
gas turbine engines, ASME Paper 83-GT-73.
Meacham, W.L., Klaas, R.M.F., Dayton, R., and Durkin, E., 1993, Dynamic stiffness and damping of foil
bearings for gas turbine engines, SAE Paper No. 931449.
Compliant Foil Bearing Technology 47-41

Gu, A.L., Process fluid foil bearing liquid hydrogen turbopump, Presented at 24th Joint Propulsion
Conference, July 1988, Boston, MA, Paper No. AIAA-88-3130.
Saville, M., Gu, A., and Capaldi, R., Liquid hydrogen turbopump foil bearing, Presented at AIAA/SAE/
ASME/ASEE, 27th Joint Propulsion Conference, June 24–26, 1991, Sacramento, CA, Paper No.
AIAA-91-2108.
Heshmat, H., A feasibility study on the use of foil bearings in cryogenic turbopumps, Presented at AIAA/
SAE/ASME/ASEE, 27th Joint Propulsion Conference, June 24–26, 1991, Sacramento, CA, Paper No.
AIAA-91-2103.
Zorzi, E.S., Gas lubricated foil bearing development for advanced turbomachines, Tech. Report
AFAPL-TR-76-114, Vols. I and II, AiResearch Manufacturing Co. of Arizona, Phoenix, AZ, November
1976.
Arakere, N.K. and Nelson, H.D., An analysis of gas-lubricated foil-journal bearings, STLE Tribol. Trans.,
35(1), 1992, 1–10.
Agarwal, G.L., Stability of rotors in compressible fluid film bearings, J. Propul., 4(5), 1988, 437–444.
Timoshenko, S.P. and Gere, J.M., Theory of Elastic Stability, 2nd edn., Engineering Societies Monographs,
McGraw-Hill Book Co., New York, 1961.
Timoshenko, S.P., Strength of Materials, Part II Advanced Theory and Problems, 3rd edn., R.E. Krieger
Publishing Co., Huntington, NY, 1976.
48
Components with
Nonconforming Contacts
48.1 Scope..................................................................................................48-1
48.2 Design Process.................................................................................48-1
Andrew V. Olver 48.3 Lubrication Analysis.......................................................................48-2
Imperial College London References.....................................................................................................48-3

48.1  Scope
The contact between two solid bodies is said to be nonconforming if the shape of the surfaces is such
that they initially touch only at a point or along a line [1]. When load is applied, the pressure becomes
high and a small area of contact is developed due to elastic deformation. Examples of mechanical com-
ponents in which nonconforming contact occurs are gears, rolling element (ball or roller) bearings,
cams and tappets, automotive CV joints, roller-workpiece contacts in metal cold-rolling, and wheel-rail
contacts. In addition, it is often observed that rough surfaces, even those of nominally conforming
geometry, are arguably nonconforming on a local scale where initial contact occurs only on the crests
of the roughness.
The size and shape of the area of contact and the distribution of pressure for many—but not all—of
these components are adequately described by Hertz theory (Chapter 4) and refinements or develop-
ments of it. Equally, when lubricated, nonconforming contacts typically operate in the elastohydrody-
namic (piezo-viscous-elastic) regime (Chapter 18).
For this reason, the relevant engineering analysis for many of these components includes deter-
mination of the Hertz (maximum) contact pressure and of the (minimum) elastohydrodynamic film
thickness. These, in turn, may be used in the determination of performance, strength, or durability
(life). Here and in the chapters that follow, we describe typical design procedures and analysis meth-
ods for nonconforming mechanical components, including detailed treatments of rolling bearings
(Chapter 49) and gears (Chapter 50).

48.2  Design Process


Figure 48.1 shows a typical design procedure for an engineering machine element with a nonconform-
ing contact or contacts. Details vary with the exact type of component. Initial design is usually on the
basis of strength: sizing so as to limit the stresses, for example, the Hertz stress or maximum contact
pressure to less than a permitted value.

48-1
48-2 Design for Lubrication and Tribology

Form, fit and function,


Choice of machine element (gears, bearings, cam, CV joint etc.)
Preliminary sizing, choice of material, manufacturing method
Choice or definition of lubricant

Tribological analysis (see figure 48.2)


(contact conditions, stresses, lubrication)

Design refinement
Critical design review
Prototype manufacture
Development (testing, analysis of results, design changes)

Series production

FIGURE 48.1  Typical design procedure for components with nonconforming contacts.

48.3  Lubrication Analysis


Central to the process is a lubrication or, more generally, a tribological analysis. For nonconform-
ing contacts, this is usually an estimation of the sliding speed and film thickness, as illustrated in
Figure 48.2.
In most applications, both the solid materials and the lubricant (oil, grease, and working fluid) are
selected before this is attempted. For example, nearly all ball bearings are made from a single (1% C,
1¼% Cr) grade of steel. Likewise, oil type and viscosity grade cannot be chosen or optimized for each
component; the oil system is common to the whole machine. In helicopter transmissions, the same
oil lubricates the 20,000 rev/min input bearing and the 200 rev/min output gear; often, this is also the
engine lubricant. The temperature range is often more than 200°C. It is therefore not possible to select
the lubricant to ensure, for example, a particular film thickness; rather, we must design the component
(shape, size, and treatment) to accept the lubricant and materials specified or available.

Load, speed history, spectrum, surface roughness , materials


Contact conditions, radii, geometry (R1x, R1y, R2x, R2y, etc.)
Kinematics: u1, u2, ubar , ∆u
Contact pressure, dimensions: p0, a, b
EHL calculations (isothermal, smooth film thickness, hmin, hc)
Lambda value

Sliding traction, thermal analysis, film temperature


Revise inlet temperature, TB
Wear modeling, revise roughness

Evaluate contact pressures, traction, film thickness,


sliding speed, temperature
Calculate life, durability
Revise design, change lubricant

FIGURE 48.2  Tribological analysis for nonconforming contacts.


Components with Nonconforming Contacts 48-3

It is necessary to decide whether or not an isothermal analysis is sufficient. In gears, for example,
sliding friction between the teeth raises the temperature of the contact inlet above that of the bulk
component or above that of the oil supply. In addition, the entire oil system may vary in temperature
according to the duty cycle. This necessitates the use of a coupled thermal model of lubrication [2,3] in
which the lubricant film thickness, traction, and heat transfer equations are solved simultaneously, per-
haps by an iterative method, as suggested by Figure 48.2. An alternative may be the use of design points
(see 4.0) where a number of separate isothermal analyses are performed at extreme or representative
temperatures.
A second problem is the evolution of roughness with time. Most lubricated nonconforming compo-
nents have a surface roughness which varies during initial running to a degree which itself depends on
the tribological conditions. This is usually beneficial (“running-in”) but is often highly dependent on
lubricant properties and chemistry and is not susceptible to rigorous analysis, arising from a combina-
tion of wear, plastic deformation, and boundary film formation and removal. This is a feature of CV
joint raceways and gear teeth but does not usually arise in well-designed rolling bearings where materi-
als are hard, surfaces are smooth, and sliding speeds are usually very low [4,5].
Finally, it is observed that there is no completely general method for assessing the life of components
analyzed. (However, an explicit method is available for rolling bearings and is described in the chapter
Rolling Bearings.) It is necessary to evaluate the significance of the calculated parameters (Hertz pres-
sure, sliding speed, temperature, film thickness etc.) against established allowable, against material
limits, and in the light of development testing, service experience, or known failure modes.
The next chapters concern the design and analysis of rolling bearings, gears, and cams and tappets.

References
1. Johnson, K L, Contact Mechanics, Cambridge University Press, Cambridge, U.K., 1985, p. 1.
2. Olver, A V, Spikes, H A, Prediction of traction in elastohydrodynamic lubrication, Proc. Inst. Mech.
Eng. Part J: J. Eng. Tribol., 212, 1998, 321–332.
3. Olver, A V, Testing transmission lubricants: The importance of thermal response, Proc. Inst. Mech.
Eng. Part G: J. Aerospace Eng., 205, 1991, 35–44.
4. Blau, P J, On the nature of running in, Tribol. Int., 38, 2005, 1007–1012.
5. Benyajati, C, Olver, A V, The effect of a ZnDTP anti-wear additive on micropitting resistance of car-
burised steel rollers, AGMA Technical Paper, 04FTM06, American Gear Manufacturers Association,
Alexandria, VA, 2004, 1–10.
49
Lubrication of Rolling
Element Bearings
Nomenclature..............................................................................................49-1
49.1 Introduction.....................................................................................49-3
49.2 Rolling Element Bearing Types.....................................................49-3
49.3 Elements of Bearing Theory and Performance...........................49-4
Introduction  •  Fatigue Criterion and Fatigue Limit
Stress  •  Selection of Fatigue Limit Stress  •  Internal
Stresses  •  Standardized Methodology
49.4 Bearing Lubrication Methods........................................................49-6
Circulating Oil Lubrication  •  Oil Bath Lubrication  •  Oil
Mist Lubrication  •  Oil-Spot Lubrication  •  Solid
Lubrication  •  Grease Lubrication
49.5 Calculations of Effective Bearing Lubrication............................49-9
Introduction  •  Local Film Thickness and Pressure Buildup
E. Ioannides 49.6 Calculation of the Lubrication Effect in Bearing Life..............49-13
S Ioannides Tribology 49.7 Lubrication Contamination Effects............................................ 49-16
and Engineering Introduction  •  Calculation of the Contamination Factor ηc
Consultants Ltd
49.8 Bearing Friction Torque Calculation..........................................49-22
Guillermo E. Simple Model for Bearing Friction  •  4S for Bearing
Frictional Torque
Morales-Espejel
SKF Engineering and 49.9 Closure.............................................................................................49-26
Research Centre References...................................................................................................49-28

Nomenclature
a Hertzian semiwidth, rolling direction (m)
αISO, aSLF stress-life modification factor (−)
A material constant (life equation) (−)
A area (m2)
Ad deformed amplitude, sinusoidal roughness (m)
Ai initial amplitude, sinusoidal roughness (m)
b Hertzian semiwidth, transverse direction (m)
c exponent in the life equation (−)
C bearing basic dynamic load rating (N)
dm mean diameter of the bearing (m)
D Dawson lubrication parameter D = Λ−1 = Rq / hmin (−)
e exponent in the life equation
E′ reduced Young modulus, EHL (Pa)
F normal contact force (N)

49-1
49-2 Design for Lubrication and Tribology

Fa axial force in the bearing (N)


Fr radial force in the bearing (N)
h film thickness (m)
h exponent in the life equation
I volume-related stress integral (−)
L Moes parameter, L = αE′(η 0u – / (2E′R ))1/4 (−)
x
L10m modified rating life (million of revs.)
M total frictional moment in the bearing (Nm)
M ( ) 3/ 4
Moes parameter, M = ⎡ F / EʹRx2 ⎤ ( 2 EʹRx /(h0u )) (−)
⎣ ⎦
Mdrag frictional moment component from oil drag (Nm)
Mrr frictional moment component from rolling (Nm)
Mseal frictional moment component from the seals (Nm)
Msl frictional moment component from sliding (Nm)
n number of over-rollings per revolution (life equation) (−)
n number of points in Simpson’s rule (−)
N number of load cycles (−)
p pressure (Pa)
p exponent in the life equation (−)
po maximum Hertzian pressure (Pa)
P equivalent load in the bearing (N)
Pu fatigue load limit (N)
R radius of curvature (m)
Rq r.m.s. of the surface roughness (m)
S probability of survival (−)

u mean velocity, u – = (u + u ) / 2 (m/s)
1 2
V volume (m3)
VR volume at risk (m3)
w exponent in the life equation (−)
x rolling direction coordinate (m)
y transverse direction coordinate (m)
z distance below the surface (stresses) (m)
z′ stress-weighted average depth (m)
α viscosity-pressure coefficient (Pa−1)
Λ lubrication quality parameter, Λ = h/Rq (−)
κ lubrication quality parameter, κ = ν/ν1 (−)
λ wavelength of the surface waviness (m)
λ′ modified wavelength of a wave component (m)
λ‾ minimum wavelength in directions x and y, λ‾ = min(λx, λy) (m)
μ friction coefficient (−)
∇ ( )
generalized wavelength for waviness, ∇ = l / a M /L (−)
η penalty factor of aSLF for total added stress (−)
η 0 lubricant viscosity at atmospheric conditions (Pa s)
ηb lubrication factor of aSLF, aISO (−)
ηc, cu contamination factor of aSLF, aISO (−)
ν lubricant viscosity (cSt)
ν1 lubricant reference viscosity (cSt)
φish factor for inlet shear heating effects (−)
φrs factor for replenishment-starvation (−)
Ψbrg bearing-type characteristic number (for the lubrication factor) (−)
Lubrication of Rolling Element Bearings 49-3

σ general subsurface stress (Pa)


σ limit general representation of fatigue limit (Pa)
τ shear stress component (Pa)
τu fatigue limit shear stress (Pa)
τvM shear von Mises stress (Pa)

49.1  Introduction
Rolling bearings are mechanical components that transmit loads from rotating to stationary machine
elements with high precision and extremely low energy loss. Linear motion rolling bearings serve the
same purpose in the transmission of linear movements. Rolling bearings are one of the most com-
monly used machine elements second only to nuts and bolts. It is estimated that approximately 60
billion rolling element bearings are in use today. The rolling element bearings are a relative newcomer,
introduced at the end of the nineteenth century, and since then they have become the prominent
method of supporting rotating shafts under load. Their continuing success is due to improvements in
the quality of steel, the quality of the contacting surfaces with very low roughness that usually allows
for effective lubrication, and the improvements in design, all of which are essential for the longevity
of the bearings.
The main purpose of the lubrication is to separate the surfaces of the rolling element and the bearing
rings with a medium, the lubricant, which has low shear strength, and thus, this is first to yield without
causing damage to the rolling element or the bearing ring in spite of the very high contact pressures. In
addition, the lubricant may act as a coolant and also carries additive packages with antioxidant, anti-
foam, extreme pressure (EP), and other additives which improve the performance and the longevity of
the rolling bearings.
It is therefore important when designing rolling element bearings in machinery to consider
carefully the operating conditions of the bearings and select the appropriate lubricant and lubrica-
tion method to obtain the required life of the bearing that allows a trouble-free operation of the
machinery.
In the following sections, the fundamentals of lubricant and lubrication method selection will be
described together with the available calculation methods for obtaining the required performance from
the rolling element bearings.

49.2  Rolling Element Bearing Types


There are different types of rolling bearings which display different characteristic properties that can
be used in different applications. A comprehensive description of the different types of bearings can
be found in the bearing manufacturers’ catalogs [1–4]. In general, rolling element bearings can be
subdivided into two main categories: ball bearings (with balls as rolling elements) and roller bearings
(with rollers as rolling elements). Furthermore, different types of bearings are suited for radial loads,
thrust or combined loads, and angular misalignment (self-aligned bearings). In Figure 49.1, some of the
basic types are shown. Many different grades of steel are used in the manufacturing of rolling element
bearings, depending on intended applications and bearing size, for example, steels for case, induction,
and through hardening; stainless steels; and high-temperature steels. The most common is the 52100
steel (through hardening) containing approximately 1% carbon and 1.5% chromium according to ISO
683-17:1999. In addition to steel, ceramics are also used as bearing material. Again, different ceramics
are suitable, for example, Si3N4, ZrO2, and Al2O3, but the most common ceramic used is Si3N4, silicon
nitride, which offers the best overall performance in rolling contact. Typical properties of bearing steel
and the silicon nitride are given later [3], in Table 49.1.
49-4 Design for Lubrication and Tribology

FIGURE 49.1  Examples of bearing types.

TABLE 49.1  Material Properties


Material Properties Bearing Steel Bearing Grade Silicon Nitride
Mechanical properties
Density, g/cm3 7.9 3.2
Hardness 700 HV10 1600 HV10
Modulus of elasticity, kN/mm2 210 310
Thermal expansion, 10−6/K 12 3
Electrical properties (at 1 MHz)
Electrical resistivity, Ω m 0.4 × 10−6 (conductor) 1012 (insulator)
Dielectric strength, kV/mm — 15
Relative dielectric constant — 8

49.3  Elements of Bearing Theory and Performance


49.3.1  Introduction
One of the earlier models for the calculation of the dynamic capacity of rolling bearings, and the
expected life in an application, was based on the Weibull weakest-link strength theory (Weibull [5],
Lundberg and Palmgren [6]). That model focused on the most appropriate failure mode of those times,
the subsurface-originated fatigue. Since then bearings have continuously improved in design, manu-
facturing precision, steel integrity, and heat treatment. Moreover, cleanliness of modern steels has sig-
nificantly reduced the occurrence of subsurface-initiated bearing failures, and failure now frequently
commences at the surface.
The fatigue model of Lundberg and Palmgren [6] was reformulated by Ioannides and Harris [7] and
Ioannides et al. [8] with the introduction of an additional fatigue strength parameter to provide a bet-
ter representation of experimental findings regarding rolling bearings load-life relationship (Lorösch [9],
Lubrication of Rolling Element Bearings 49-5

Voskamp [10], Furumura et al. [11]). Central to all bearing life models is the fatigue life modeling. The crack
initiation models invoke a stress-power law to account for the portion of the life spent in the initiation of
a crack, and this part dominates the lifetime in the fatigue of brittle bearing steels. This method evolved
from using a single stress value (the maximum orthogonal shear stress) to assess the risk of fatigue, which
is appropriate for strictly Hertzian stress fields, to models which use general three-dimensional (3D) con-
tact stress fields in the material with a multiaxial fatigue criterion. The methods of [7] and [8] are appro-
priate for application to realistic “rough” surface, lubricated contacts encountered in rolling bearings. It is
therefore possible now to calculate the lives of rolling bearings under more realistic application conditions
where the site of failure shifts to the surface. Furthermore, the effects of roughness, contamination, edge
stresses, and internal stresses (hoop and residual) can be included in the life calculations.
The expanded version of the original life formula of LP introduced in Reference 6 which accounts for
local stresses and a fatigue limit is expressed in Equation 49.1:

1 (s − s limit )cdV
ln ∝ N e ∫V h
s > s limit (49.1)
S z

As shown by Harris [12], this can be numerically approximated using Simpson’s rule as follows:

⎡ c⎤
j =n i =n (s vm,ij − s vm,limit ) ⎥
e ⎢ p ab
1− h
1 dr
ln = N ⎢
S ⎢ 9n
2 Σ Σ
j =1
ℑj ℑi
i =1 r ih


(49.2)

⎣ ⎦

49.3.2  Fatigue Criterion and Fatigue Limit Stress


In multiaxial fatigue, various criteria have been proposed, for example., see Kakuno and Kawada [13].
Each failure criterion is usually specific to a particular mode of failure and efforts to find a universal
criterion that may represent different modes had only partial success. It may be noted, however, that in
contact fatigue when normal contact stresses alone are involved, the various shear stresses which have
been used as criteria are all proportional in magnitude to one another. Such criteria proposed in the lit-
erature are maximum Hertzian pressure, maximum orthogonal shear stress [6], maximum shear stress,
maximum surface traction, von Mises stress, and maximum shear stress amplitude regardless of plane
and direction. Further to this maximum shear stress amplitude, modifications which add the normal
stress to the plane of maximum shear stress amplitude or which add the hydrostatic pressure effect have
also been proposed (see Crossland [14], Dang Van et al. [15], Ioannides et al. [8]). The ISO 281:2007 [16]
uses the maximum orthogonal shear stress criterion which is simple but cannot formally account for the
effect of other part of the stress tensor on life.

49.3.3  Selection of Fatigue Limit Stress


In an ideal contact where there is only the Hertzian stress field, the fatigue limit values appear to be as high
as 2200 MPa. In typical real bearing applications, additional stresses exist due to manufacturing process
tolerances and operating conditions that are not always detailed in the calculations. For this reason, in the
daily use of the calculation method, it is prudent to select a lower fatigue limit value of 1500 MPa Hertzian
stress. Furthermore, if the manufacturing quality of the bearing or the cleanliness of the steel is not ideal,
the fatigue limit stress can decrease to a level of 1100 MPa Hertzian stress, Barnsby [17].

49.3.4  Internal Stresses


Fatigue is the result of the full stress tensor applied to the material. This tensor includes stresses gener-
ated during bearing manufacturing (i.e., residual stresses), fitting or other ring deformations (i.e., hoop
49-6 Design for Lubrication and Tribology

stresses), and the load over-rolling stresses. Residual stresses are mostly determined by heat treatment
and manufacturing processes, but they also change during operation. Hoop stress is caused by the inter-
ference fit between inner ring bore and shaft to prevent the inner ring from rotating. Press fitting of the
outer ring in the housing creates similar, but compressive, hoop stresses. Lubricant film effects can be
also included in the calculations; see Sections 49.4 and 49.5.

49.3.5  Standardized Methodology


The model described by Equation 49.3 has been standardized (ISO 281:2007 [16]). A set of experimen-
tally adjustable constants are required so that it may be used as an engineering predictive model. For
continuity purposes, the concepts of basic dynamic capacity C and equivalent load P, established by
Lundberg and Palmgren [6], are preserved in Reference 16. The effects of localized stresses and the pres-
ence of a fatigue limit stress are introduced via a life adjustment factor aISO in the following equation:
p
⎛C⎞
Lnm = a1aISO ⎜ ⎟ (49.3)
⎝P⎠

where the exponent p = 3 for ball bearings and 10/3 for roller bearings. The rationale for developing this
equation as an extension of Reference 6 is given in Reference 8.
The actual stress state of the rolling contact may be the result of many surface-induced interactions
from local macro- and micro-surface features. This, in turn, can be related to the amount of the surface
separation, that is, lubrication conditions, and the particulate contamination of the contact. Following
Tallian et al. [18], a simple rating parameter of the lubrication condition of the bearing is the viscosity
ratio, κ. The κ parameter is defined as the ratio between the actual lubricant viscosity at the operating
temperature of the bearing (ν) and a standardized reference viscosity (ν1) rated as adequate for the
lubrication of the bearing (ISO 281:2007 [16]). The lubrication parameter κ can also be related to the
well-known specific film thickness using the approximation: κ ≈ Λ1.3.
In order to form an adequate lubricant film between the rolling contact surfaces, the lubricant must
retain a certain minimum viscosity when the lubricant is at operating temperature. The bearing life
may be extended by increasing the operating viscosity ν. The reference kinematic viscosity, ν1, can be
estimated by means of the diagram in Figure 49.2, depending on bearing speed and pitch diameter, Dpw,
ISO 281:2007 [16].
When the viscosity ratio, κ, is estimated, the effects of lubrication on life can be calculated using the
ISO methodology [16]. In the case of radial roller bearings, Figure 49.3 indicates how the κ-dependent
life-modifying factor aISO is obtained, and Reference 16 contains similar information for other types of
bearings.
The life modification factor aISO is the more recent development of originally individual life modi-
fication factors developed for different lubrication regimes and different materials. These were further
extended to account for further conditions that influence the life of bearings, for example., hoop and
residual stresses (Zaretsky [19]). The aISO is different from such factors in that it stems directly from
estimates of all accountable stresses and the fatigue limit instead of being constructed as the product of
factors derived for each condition independently.

49.4  Bearing Lubrication Methods


49.4.1  Circulating Oil Lubrication
Lubrication of rolling-element bearings with jet or under-race techniques requires the use of recir-
culating oil systems. This is necessary especially in high-speed cases when the temperature of the oil
increases inducing aging of the oil. These recirculating oil systems can include external pumps, filters,
Lubrication of Rolling Element Bearings 49-7

Vv mm/s"

1000 2

500 5
10
20
200
50

in
m
100

/
100

r ev
n,
200
50
500

20 1,00
1,50 0
3,00 0
0
10 2,00
10,0 0
00 5,00
0
5 50,0 20,0
10 5 00 00
3
10 20 50 100 200 500 1,000 2,000 Dpw mm

FIGURE 49.2  Reference kinematic viscosity.

aISO
x = 4 2 1 0.8
50

20
0.6

10

5
0.5

1 0.4

0.3
0.5

0.2

0.2 0.15

0.1 0.1

0.005 0.01 0.02 0.05 0.1 0.2 0.5 1 2 5 ecCu/P

FIGURE 49.3  Life modification factor, aISO, for radial roller bearings.
49-8 Design for Lubrication and Tribology

heat exchangers for oil cooling, scavenge pumps, and flow control devices, or they may be simple built-in
pumps merely to circulate lubricant within a gearbox. Filtering improves the cleanliness of the oil and
thus the life of the bearing.
For rolling-element bearing applications where speeds are too high, jet lubrication is frequently used
to lubricate and control the bearing temperature by removing generated heat. In jet lubrication, the
placement of the nozzles, number of nozzles, jet velocity, lubricant flow rates, and removal of lubricant
from the bearing and its immediate vicinity are all very important for satisfactory operation. Even the
internal bearing design is a factor to be considered.
As speeds of the main shaft of turbojet engines are pushed upward, the lubricant is progressively
thrown centrifugally outward. Increased flow rates will only add to heat generation from the churning
of the oil. Brown [20] originally described an “under-race oiling system” used in a turbofan engine for
both ball and cylindrical roller bearings. In this technique, the lubricant is directed under and through
holes into the inner race to cool and lubricate the bearing. High flow rates may be employed so that
substantial amounts of lubricant may pass only under the bearing for cooling. This method is widely
employed in jet engines.

49.4.2  Oil Bath Lubrication


For many applications where shaft speeds are low and cooling requirements are not demanding, splash
or bath lubrication is sufficient. Then part of the bearing or an adjacent component is immersed in
lubricant. Since a finite and generally small quantity of lubricant is used, the system should be enclosed
and well sealed. This is also necessary to keep foreign debris and dirt out of the system since the lubri-
cant does not pass through filters as it would in a circulating system. If contamination is present but
improved cleanliness is required, a kidney-loop filtering system can be installed which circulates the oil
and filters it.
The bath lubrication is generally inexpensive and has low maintenance costs. It is fairly common
in industrial machinery where speed and temperature conditions are not severe. When speeds and
operating temperature are somewhat higher, then increased and reliable oil lubrication is necessary,
and oil pickup rings can be used. The pickup ring hangs loosely on a sleeve on the shaft and transports
oil from the bottom to a collecting trough. The oil then flows through the bearing back into a reservoir
at the bottom.

49.4.3  Oil Mist Lubrication


This method uses a suspension of fine oil particles in air as a fog or mist to transport oil to the bearing.
The fog is condensed at the bearing so that the oil particles will wet the bearing surfaces.
Very low oil flow rates are used for such lubrication, and the airflow has cooling function. Air-oil
mist lubrication is nonrecirculating; the oil is passed through the bearing once and then is discarded.
This type of lubrication has not been recommended for some time due to possible negative environ-
mental effects.
Recent developments of this method use extremely small quantities of oil and limit the amount of
stray mist to the minimum. Oil mist lubrication systems are used today in very specific applications like
military, oil, and gas.

49.4.4  Oil-Spot Lubrication


Accurately measured very small quantities of oil are directed to each bearing by compressed air. This
minimum quantity enables bearings to attain low temperatures at high speeds. The oil is supplied
through a metering unit and a nozzle. It is projected onto the bearing, or it just flows to the bearing
raceways driven by surface tension. When using circulating oil and the spot oil, the oil return ducts
Lubrication of Rolling Element Bearings 49-9

should be adequately dimensioned to avoid the building up of oil in the bearing. In order to increase
bearing service life, both oil and compressed air should be free of contaminants.

49.4.5  Solid Lubrication


For unusual environmental conditions, extreme temperatures, or low pressures (vacuum), rolling-element
bearings can be lubricated by solid films. The use of solid film lubrication generally limits both bearing
life and speeds to considerably less than those attained with grease or oil lubrication. Solid lubricants
may be bonded films, transfer films, or loose powder applications.
Transfer film lubrication is employed in cryogenic systems such as rocket engine turbopumps or
high-temperature applications such as furnaces and x-ray tubes. Organic and inorganic materials such
as PTFE, graphite, hexagonal boron nitride, molybdenum disulfide, and tungsten disulfide, as well as
cadmium, silver, and gold, are used for this purpose. More recently PVD methods are used to produce
MoST (titanium-reinforced molybdenum disulfide) and DLC. Some of their properties and tribological
behavior can be found in Kadiric [21].

49.4.6  Grease Lubrication


Grease lubrication is the simplest and most common method of lubrication for rolling-element bear-
ings and is gaining ground as the number of bearings sealed and greased for life is constantly increas-
ing, especially for smaller sizes. Lubricating greases consist of a fluid phase of oil and a solid phase of
a thickener. Additives are used, both as add-on in the grease or as chemically bound to the thickener.
Starvation (oil films thinner than the case where there is an abundance of oil) is a common character-
istic of grease lubrication. When the starvation becomes severe, bearing failure will occur. The grease,
however, has a self-adjusting capability, and when starvation pushes up the temperature of the bearing,
more oil is released or fresh grease is entrained and the starvation is reduced. This can be a repeated
several times until such a release is not possible any more. The grease has then reached its life, and the
bearing requires relubrication with fresh grease, otherwise, the bearing failure is imminent. Bearing
manufacturers’ catalogs provide information on the relubrication intervals depending on the bearing
type, temperature, speed, and load. A comprehensive review of the grease lubrication mechanisms can
be found in Lugt [22]. For the calculation of the κ value, most bearing manufacturers recommend the
use of the viscosity of the base oil of the grease. It is possible, however, to account in some simple way for
the reduction of the lubricant film, because of starvation. In Reference 17, it is suggested that for grease
lubrication, 0.7 of the calculated base oil film thickness should be used.
One of the disadvantages of grease lubrication is the inability to dissipate heat, unlike oil lubrication
where the oil can convect heat away from the bearing and the bearing temperature can be thus regu-
lated. This implies that the operating temperature is more difficult to estimate and requires thermal
calculations or measurements.
More details for the various types of lubrication can be found in the bearing manufacturers’ catalogs.

49.5  Calculations of Effective Bearing Lubrication


49.5.1  Introduction
49.5.1.1  Early Developments
The concentrated nature of the stress makes the fatigue performance of rolling bearings more dependent
on the quality of the lubricant film separating the mating surfaces. After the discovery of the physical
mechanism allowing the generation of an elastohydrodynamic lubrication (EHL) film in highly loaded
nonconformal contacts (e.g., Grubin [23]), a number of studies were directed to the use of this theory
to improve the reliability and performance of rolling bearing applications. Dawson [24] was among the
49-10 Design for Lubrication and Tribology

first ones to notice the relationship between an EHL parameter, namely, D (D = Rq/h), and the micro-
pitting life of lubricated rolling disks. He observed that the lower this parameter (to a certain limit),
the higher the micro-pitting resistance of the disks. Using a two-disk and a four-ball test rig, Tallian
et al. [25] apply the ratio Λ = h/R q (reciprocal of D) to relate lubrication to the performance of rolling
contacts. This initial work was followed by further detailed investigations, Tallian et al. [26] and Tallian
[27], in order to relate the measured percentage of metal-to-metal contact to the calculated penetration
of asperities into the EHL film and performance of the rolling contact. The new findings were also con-
firmed by the work of Skurka [28] and Danner [29] using endurance testing of tapered roller bearings.
This early work resulted in the prompt adoption of the EHL lubrication ratio (Λ) by the bearing industry
and by the engineering community with the publication of the ASME Engineering Design Guide [30].
This shows a first example of the practical use of the newly discovered EHL theory in the design of roll-
ing bearing applications. Note also that the ASME relationships between the specific film thickness (Λ)
and relative fatigue life of rolling contacts are commonly used at present for general machine design.
Later, Christensen [31], Christensen and Tønder [32], and later Patir and Cheng [33–34] also investigated
the theoretical aspects of roughness effects in partial hydrodynamic lubrication and the significance of
the Λ-ratio to the mechanism involved in the lubricant film buildup. However, they treated roughness in
a statistical way and did not consider its elastic deformation. The effect of roughness was modeled with
the use of flow factors, which were introduced in the coefficients of the Reynolds equation. Following a
more pragmatic approach, Andreason and Lund [35] also reviewed several ball bearing endurance tests
performed at various speeds and temperatures and lubricated either with oil circulation or grease. In
the follow-up to this work, Andreason and Snare [36] introduced the new concept of the viscosity ratio
(κ), the actual viscosity used in the application divided by a reference viscosity considered necessary for
good lubrication, in order to simplify the estimation of the amount of EHL separation of the rolling con-
tact. Therefore, viscosity ratios equal or larger than one were considered as good quality in lubrication
and lower than one as poorer quality in lubrication. Andreason and Snare [36] also developed a lubrica-
tion factor (a2) to adjust the basic rated life of rolling bearings to the lubrication conditions of the appli-
cation. The new model based on the viscosity ratio (κ) rather than the specific film thickness (Λ) was
further illustrated by Bolton [37] and, some years later, by Heemskerk [38]. In the work of Heemskerk,
endurance test results of rolling bearings are also presented.

49.5.1.2  New Developments in Micro-EHL


With the development of refined numerical solutions and the increase of computer power, researchers
started to focus on deterministic ways to model the effect of roughness in EHL contacts. Some early
works are due to Goglia et al. [39,40], Kweh et al. [41], Lubrecht [42], Venner [43], and Sadeghi and Sui
[44]. Later works show the contribution of many other researchers who have applied efficient numeri-
cal methods to evaluate pressure and stresses resulting from surface asperities interacting with the
EHL film, among many others, Ioannides and Kuijpers [45], Ai and Cheng [46], Ai and Lee [47], and
more recently Chapkov et al. [48]. The use of numerical simulations for the solution of the EHL of
interacting surfaces has also shown the flattening of the surface roughness and increase of pressure
fluctuations. This effect was also investigated by Tripp and Ioannides [49], who performed exten-
sive parametric calculations to relate roughness geometry to the theoretical fatigue life of the con-
tact. However, in the later years, a different approach based on semianalytical solutions for the study
of deterministic micro-EHL has been gaining strength among researches, mainly because it allows
the engineer to understand the basic mechanisms involved in this complex problem. For instance, it
showed that long-wavelength roughness components deform, while short wavelengths do not; all this
is captured in a simple relationship based on a generalized wavelength parameter involving wavi-
ness geometry and operating conditions. This approach is known as “amplitude reduction,” initiated
from the works on sinusoidal roughness of Greenwood and Johnson [50], Venner [43], Greenwood
and Morales-Espejel [51], Venner and Lubrecht [52], Venner et al. [53], and the mathematical for-
malism due to Hooke [54]. Finally, the introduction of sliding and complete 3D surface features and
Lubrication of Rolling Element Bearings 49-11

roughness is due to Hooke et al. [55,56]. Experimental evidence showing good agreement with predic-
tions of the amplitude reduction methodology has been given by several authors, for example., Choo
et al. [57] and Félix-Quiñonez et al. [58], and the detailed comparisons between theory and experi-
ments from Hooke et al. [56] and Gabelli et al. [59]. The use of these novel techniques has proved to
be effective in the analysis of real bearing roughness textures, allowing for a better estimation of the
degree of surface separation and expected reliability of the rolling contact. Recently Gabelli et al. [59]
have included the calculation of subsurface stresses, in the analysis in rolling bearing life effects from
particle indentations and roughness; see sub-section 6. The “amplitude reduction” technique repre-
sents an important differentiation on the previous work done by Tallian in the development of the
widely used parameter Λ and the effects on bearing life. The effect of roughness and lubrication on the
life rating of rolling bearings must be based on the use of the whole topography rather than only on
the Rq parameter. However, for specific surface finishing processes like that used in rolling bearings
(e.g., grinding and honing), it is still possible to reduce the complexity of the approach to semianalyti-
cal equations based on Rq. The detailed micro-EHL analysis for pressures, deformations, and stresses
used in this work has already been published elsewhere (e.g., [59]). It is stressed that although the
general theory used here can consider sliding (e.g., [55,56]), the following description focuses on pure
rolling conditions.

49.5.2  Local Film Thickness and Pressure Buildup


The “amplitude reduction” methodology for micro-EHL can be applied to real bearing surfaces and real
operating conditions, as described in Reference 60. Thus, the deformed clearances, pressure fluctua-
tions, and stresses can be found. Certainly, this methodology has some limitations.

49.5.2.1  Micro-EHL Model Limitations


The “amplitude reduction” scheme used, as discussed here (e.g., [60]), for the calculation of deformed
clearances and pressure fluctuations from micro-EHL effects has two basic limitations:
1. The scheme is based on the assumption that the micro-geometry amplitudes are small and the
mean viscosity is high.
2. It is also based on the idea that full-film conditions apply.
Therefore,
1. Small amplitudes: Morales-Espejel et al. [60] suggest that for typical EHL conditions, the min-
imum pressure should not be lower than 0.5 GPa for the linearization to be applicable. Large
micro-geometry features would tend to produce large pressure fluctuations, whereas in valleys,
the pressure drops below this limit.
2. Full film: In general, some metal-to-metal contact can be tolerated, as long as it does not
affect considerably (say no more than 5%) the load balance between lubricated areas and solid
contacts. The amount of metal-to-metal contact would depend on the way the roughness is
deformed and the wall-slip conditions between lubricant and wall. Thus, it depends as well on
the geometry of the whole topography and operating conditions. It seems difficult to determine
a single limiting condition. However, as found from the comparison with dry contact calcula-
tions, the use of the model in the present examples for “nominal” κ ≥ 0.5 seems reasonable. For
lower κ values, it appears risky to use this approach, and care should be taken in the absence of
a mixed lubrication model. Here a convenient solution is the use of interpolated results from
the last point in the micro-EHL model and dry contact model results, which will always be the
limiting case in pressure and deformation.
The only limitation of the micro-stress model is the elastic behavior of the material and the half-space
assumption.
49-12 Design for Lubrication and Tribology

Micro-geometry effects in bearing life depend on the balance of two effects: the local increase in
stresses due to deformation and the local increase in film thickness. On the one hand, the elastic defor-
mation of the micro-geometry produces an increase in the surface separation; thus, the actual lubrica-
tion quality factor (κ or Λ) will increase compared to the nominal value (undeformed micro-geometry).
In general, an increase in local film thickness is beneficial in lubrication (it reduces surface interaction).
On the other hand, the pressure buildup increases the state of stresses. This will reduce the fatigue life of
the surface. Therefore, for a particular operating conditions and surface topography, the two effects will
combine to give the final result. A real physically based lubrication factor should combine the effects of
the two mechanisms; therefore, a model consistent with the physics of the problem should account for
this coupling in the relationship between the lubrication quality factor and bearing life.

49.5.2.2  Micro-Geometry Deformation


Consider the case of pure rolling; Venner and Lubrecht [52] describe in a simple curve the amplitude
ratio (deformed/undeformed) of the surface waviness for circular, pure rolling EHL contacts with any
nonisotropic harmonic pattern. The Venner and Lubrecht amplitude reduction equation, found via
curve fitting from numerical simulations, reads

Ad 1
=
Ai 1 + 0.15 f (r )∇ + 0.015[ f (r )∇]2

⎧⎪e1−1/ r , " r >1 l


f (r ) = ⎨ with r = x (49.4)
⎪⎩ 1, " r ≤1 ly

defining ∇ = (λ‾/a)M0.5L−0.5, where Ad is the deformed amplitude of the waviness, Ai is the initial unde-
formed amplitude, and λ‾ = min(λx, λy). The variable ∇ was initially proposed by Venner et al. [53] for
line contacts and later extended for point contacts [52]. The problem to relate a deformed real micro-­
geometry to ∇ is that there is no unique representative wavelength λ (like in waviness), except for a fixed
cutoff length used in metrology or for a physically based length related, for instance, to the Hertzian
contact semiwidth. Morales-Espejel et al. [61] propose a methodology to overcome this problem and
include the effect of elastic deformation in the lubrication quality factor κ.

49.5.2.3  Micro-Stress Effect


Micro-geometry effects in bearing life depend on the balance of two effects: the local rise of stresses due
to deformation vs. the local increase of film thickness. Deformation is induced by pressure fluctuations
which can have an effect on life. To account for the micro-stresses, Ioannides et al. [8] introduce a stress
concentration factor involving the lubrication quality of rolling bearings (named ηb); this factor is fur-
ther discussed and linked to micro-EHL behavior in Gabelli et al. [59]. Figure 49.4 shows the parameter
ηb as a function of κ, as adopted by ISO [16].
Following Reference 16, the rolling bearing life is calculated from Equation 49.3:

p
⎛ ⎛C⎞ ⎞
⎜ L10m = aISO ⎜ ⎟ ⎟
⎜ ⎝ P ⎠ ⎟⎠

In Equation 49.3, the stress-life modification factor aISO, as defined in Reference 8, has the following form:

−c /e
w
1 ⎛ P ⎞
aISO = aSLF = 1 − ⎜h u ⎟ (49.5)
10 ⎝ P ⎠

Lubrication of Rolling Element Bearings 49-13

0.9

0.8

Lubrication factor, ηb 0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
10–1 100 101
Lubrication conditions, κ

FIGURE 49.4  Lubrication factor for rolling bearing life (thick line), as a function of lubrication quality param-
eter κ, as given in ISO. (From International standard: Rolling bearings—Dynamic load rating and rating life, ISO
281:2007. With permission.)

The penalty factor 0 ≤ η ≤ 1 used in aSLF gives an average account for the actual stress-status of the rolling
contact in addition to the idealized smooth Hertzian stress. For an analytical rating of the η factor, it is
required to quantify the real stress condition developed in the rolling contact during bearing operation.
The actual stress state of the rolling contact may be the result of many surface-induced interactions from
local macro- and micro-surface features. This, in turn, can be related to the amount of surface separa-
tion, that is, lubrication conditions, and particulate contamination of the contact. Following Reference
16 a simple rating parameter of the lubrication condition of the bearing is the viscosity ratio κ. Detailed
explanations about the derivation of this engineering lubrication parameter can be found in Reference
8. The penalty factor 0 ≤ η ≤ 1 used in Equation 49.5 can be described as a product of two concurrent
and interrelated quantities: the lubrication factor ηb and the contamination factor ηc as

h ( bcc , dm ,k ) = y brg hb (k )nom hc (bcc , dm ,k ) = hb (k )brg hc (bcc , dm ,k ) (49.6)


49.6  Calculation of the Lubrication Effect in Bearing Life


The theoretical relationship between the lubrication quality, characterized by the viscosity ratio κ, and
the corresponding reduction to the life and fatigue load limit of the bearing will be discussed hereafter.
For this purpose, one needs to quantify the fatigue life reduction resulting from an actual rolling bear-
ing, with standard surface roughness, compared to the one characterized by an ideally smooth raceway,
as from a purely Hertzian, friction-free, stress distribution hypothesis. This can be performed by com-
paring the theoretical fatigue life between a real bearing (with standard roughness) and the fatigue life
of a hypothetical bearing with ideally smooth and friction-free surfaces. Thus, the following life ratio
has to be quantified:

L10.r aSLF .rough


= (49.7)
L10. s aSLF . smooth
49-14 Design for Lubrication and Tribology

The previous ratio can be evaluated numerically using the fatigue life stress integral applied to an actual
rolling contact; see examples of use of this methodology in Reference 59,

c
1 ti − tu
ln ≈ AN e
S ∫
Vr
z ʹh
dV (49.8)

In Equation 49.8, the relevant quantity affecting the life ratio (Equation 49.7) is the volume-related stress
integral which reads
c
ti − tu

I=
∫ z ʹh
dV (49.9)

Using the previous notation, the life equation (Equation 49.3) can now be expressed as

1/ e
N 1 ⎛ ln(1/S) ⎞
L10 = ≈ (49.10)
106 u u ⎜⎝ AI ⎟⎠

In this equation, the stress integral (I) can be computed for both standard roughness and for an ideally
smooth contact; thus, it can be used for the estimation of the expected effect on the bearing life as given
by the life ratio of Equation 49.7. In other words, the following applies:

1/ e
⎛ L10.r ⎞ ⎛ I smooth ⎞ ⎛ aSLF .rough ⎞
⎜ ⎟ = ⎜⎜ ⎟⎟ =⎜ ⎟ (49.11)
⎝ L10. s ⎠ ⎝ I rough ⎠ (m,n) ⎝ aSLF . smooth ⎠ (m,n)

In general the ratio (Equation 49.11) depends on the surface topography (index m) and amount of sur-
face separation or amount of interposed lubricant film (index n).
The lubrication factor ηb can now be directly derived from Equation 49.11 by introducing the
stress-life factor of Equation 49.6. Under the hypothesis of ideally clean lubricant, the contamination
factor ηc can be set to unity. Thus, for standard rolling bearing roughness, the stress-life factor can
be written as

−c /e
w
1 ⎛ P ⎞
aSLF .rough = 1 − ⎜ hb u ⎟ (49.12)
10 ⎝ P ⎠

Similarly, in case of a hypothetical bearing with an ideally smooth raceway, the ηb factor can be set to
unity, and the stress-life factor then becomes

−c / e
w
1 ⎛P ⎞
aSLF . smooth = 1− ⎜ u ⎟ (49.13)
10 ⎝P ⎠

Inserting Equations 49.11 and 49.12 into Equation 49.10, the following is derived:

1/ w
w −1/ c
P ⎛P ⎞ ⎛ I smooth ⎞
hb(m,n) = 1− 1− ⎜ u ⎟ ⎜⎜ ⎟⎟ (49.14)
Pu ⎝P ⎠ ⎝ I rough ⎠( m , n )

Lubrication of Rolling Element Bearings 49-15

Equation 49.14 shows that a (m × n) matrix of numerically derived values can be constructed starting
from the calculation of the fatigue life and related stress-volume integral of standard rough bearing
surfaces. This calculation must be extended to include different amount of surface separation (film
thickness), from thin films up to full separation of the rolling contact. The calculation procedure illus-
trated in Reference 59 was applied for the numerical evaluation of the ηb(m,n) considering a representative
sample of actual rolling bearing surfaces.
Following the methods described earlier, a set of ηb(m,n) values was obtained. The resulting data points
and interpolation curves are plotted in the κ − ηb diagram shown in Figure 49.4. For clarity, only a
representative group of typical bearing surfaces is presented. The numerically generated curves of ηb(κ)
consistently show a typical trend with a rapid decline of ηb for a reduction of the nominal lubrication
conditions κ of the contact.
It must be pointed out, as discussed in Reference 48, that the dimensionless wavelength parameter
would possibly provide a better correlation with the pressure buildup at the asperities and thus with
ηb. A lubrication parameter, which would include, in addition to the film thickness, a measure of the
asperities wavelength, as discussed in Reference 48, would probably provide a better differentiation
between the different roughness textures and the corresponding ηb compared to the one shown in
Figure 49.4.
However, the requirement of standardized dynamic life ratings is basically to ensure a lower bound-
ary, or safe limit, to the performance of rolling bearings. This with the many possible types of surface
roughness textures presently used in typical, good-quality rolling bearing products. For this purpose,
the simple viscosity ratio approach seems sufficient and indeed convenient, considering that κ is a very
well-established lubrication parameter in bearing engineering practice. Regarding the general shape of
the ηb(κ) curves shown in Figure 49.4, it is also to be noticed that there is a close resemblance to the life-
ratio (Λ) curves obtained by Tallian et al. [26], thus indicating that a single basic physical phenomenon
is indeed being observed.
A limitation of the numerical calculations is the requirement of a minimum amount of oil film pres-
ent in the contact (to preserve continuity condition of the fluid flow equations that are used). Therefore,
κ values lower than ∼0.2 could not be easily evaluated.
Purely dry conditions could also be used to estimate the fatigue stress integral for conditions in
which the film can be considered negligible. However, as shown by the general trend of the curves of
Figure 49.5, the tendency of ηb is toward the origin of the diagram for conditions approaching the nomi-
nal lower bound of the κ range.
For comparison, in Figure 49.5 is also shown the lubrication factor equation described in Reference 8:

5/2
hb (k )brg ⎛ b (k ) ⎞
hb (k )nom = = ⎜ 3.387 − 1b2 (k ) ⎟ (49.15)
y brg ⎝ k ⎠

In Equation 49.15, the constants b1 and b2 are assigned for three intervals of the κ range, and ψ brg is a con-
stant characterizing each of the four main types of rolling bearings: radial ball, radial roller, thrust ball,
and thrust roller bearings. In the present assessment, we compare the numerically obtained ηb to the nor-
malized standard form of the lubrication factor ηb(κ)nom of Equation 49.15. It is shown that the relation-
ship (Equation 49.15), indicated in Figure 49.5 with a solid thick line, has a good safety setting compared
to the numerically calculated ηb results. Indeed Figure 5.2 shows that nearly all analyzed surface rough-
ness samples are well above the standard limit line. This suggests that Equation 49.15 is a reasonable safe
choice for rating the effect of lubrication conditions of the bearing and the expected endurance life.
Some of the roughness textures used in the numerical evaluation belonged to bearings that were also
endurance tested. Comparisons between endurance testing lives and lives obtained using the lubrica-
tion factor (Equation 49.15) are further discussed in the experimental results in Reference 59, which
deals with endurance testing of rolling bearing under various lubrication conditions.
49-16 Design for Lubrication and Tribology

0.75

Lubrication factor ηb
0.5
Surface 1
Surface 2
Surface 3
Surface 4
Surface 5
Surface 6
Surface 7
0.25 Surface 8
Surface 9
Surface 10
Standard
Surface 11
Surface 12
Surface 13
Surface 14
0
0.1 1 10
Viscosity ratio κ

FIGURE 49.5  Summary of numerically calculated lubrication factor for typical surface textures of bearing race-
ways. The lubrication factor used by bearing standard (thick solid line) is also plotted, as from Equation 49.14.
(From Gabelli, A. et al., Tribol. Trans., 51, 428, 2008. With permission.)

This approach is basically embodied in Reference 16 and can be used to calculate κ (Figure 49.2) and
the effects of lubrication on bearing life (Figure 49.3). A similar approach, but further refined, can be
found in Reference 3.

49.7  Lubrication Contamination Effects


49.7.1  Introduction
Surface damage can arise from a number of different sources such as damage occurring during trans-
port, damage from incorrect mounting procedures, and damage from the over-rolling of large con-
taminant particles in the lubricant. Of all these possible sources, the last is probably the most common
source of surface damage. When contamination is present in the lubricant in the form of solid particles,
the solid particles are entrained in the contacts, and damage of the mating surfaces will occur. The
most prominent sources of particles in mechanical systems are manufacturing and assembly, ingres-
sion from external sources, and internal generation. Manufacture and assembly may include particu-
late matter in new oil for the initial fill, particles remaining from manufacturing and assembly in
a new system, or particles from new components installed into an existing system. These particles
may include manufacturing swarf, grinding and polishing compounds, or fibers from wipers. Initial
contamination from manufacturing and assembly is transient and tends to be removed after the first
0%–5% of system operation. Ingression from external sources may include particulate matter in refill
oil, ingression through filler caps, breathers, seals or open reservoirs, or particles introduced during
maintenance activity. Environmental particles may include mineral grains, such as sand, dust, soot,
or emissions from other local manufacturing processes. Internal generated particles may include oil
additive precipitates or wear debris.
Particle denting, and contamination marks found on bearing raceways, can induce stress concentra-
tions and facilitate surface-initiated fatigue. The lubricant film developed at the dent and related local
surface stresses are also significant to the crack initiation mechanism. When an assessment of some typ-
ical equations used in rolling bearing dynamic ratings is carried out, it is found that the degree of lubri-
cation of the rolling contact and the cleanliness conditions of the oil are significant to the prediction of
Lubrication of Rolling Element Bearings 49-17

the life expectancy of the bearing. The micro-EHL analysis to qualify the micro-contact stress condition
and its effect on the bearing endurance ratings has been discussed in the previous section.
The contamination damage is in the form of dents, and this change of topography results in film
variations, additional pressure fluctuations, and consequently additional local stresses, which, in turn,
have a detrimental effect on life. An example of this pressure fluctuation and film variation is shown in
Figures 49.6 and 49.7.
Figure 49.7 shows the von Mises stress fields (normalized respect to the maximum Hertzian pressure)
in two planes (at y = 0 and x = 0) in the center of the contact corresponding to the distribution of pres-
sures shown in Figure 49.6. The maximum von Mises stress calculated for this case is 0.42 p0, which is
increased (23%) from the one given in Reference 47 due to the presence of a realistic shoulder height at
the edge of the dent.
The use of a unified numerical approach, based on FFT for the calculation of EHL micro-contact
stress, allows the application of the same basic methodology, as used in the assessment of the lubrica-
tion factor, for the contamination factor ηc also. In case of the lubrication factor, the stress assessment
was performed with reference to typical surface roughness textures. For the contamination factor,
a similar procedure can be applied for the stress resulting from particle over-rolling and resulting
denting of the bearing raceway. This is performed assuming that the initial shape of the dent, after the
indentation process, remains unchanged [59]. Several researchers, for example, [62,63], have shown
that plastic deformations during the dent over-rolling may change the initial geometry of the dent
imposing additional residual stresses. Furthermore, under very low lubricant film, wear may also be
a factor. Despite all this complexity, a conservative starting point is herewith assumed considering
that no changes will occur to the initial conditions of the dent after indentation. Furthermore, the
mechanism of particle entrapment, particle over-rolling, and dent formation has also been reported
[62,64–69]. The methodology that culminates in a practical method used in the ISO [16] was intro-
duced in Reference 68, that is, a measure is given to the fatigue damage induced by surface denting

Dimensionless pressures

1.5
1
p/po

0.5
0
1
0.5 1
0 0.5
–0.5 0
–1 –0.5
–1.5 –1
–1.5
y/a x/a
Deformed geometry and clearances
zd /hc

–5

–10
1
1
0 0 0.5
–1 –0.5
–1.5 –1
y/a x/a

FIGURE 49.6  Normalized pressures and clearances as calculated with the present technique for the dent example
shown in Reference 59 and with a dent geometry including dent-shoulder height of 0.15 μm. Pure rolling conditions
when the dent is at the center of the contact.
49-18 Design for Lubrication and Tribology

Von mises, τvM/p0


1

0.9
0.25
0.8

0.7 0.2

0.6
0.15
z/a

0.5

0.4
0.1
0.3

0.2
0.05
0.1

0
–2 –1.5 –1 –0.5 0 0.5 1 1.5
(a) x/a

Von mises, τvM/p0


1
0.3
0.9

0.8
0.25
0.7

0.6 0.2
z/a

0.5
0.15
0.4

0.3 0.1

0.2
0.05
0.1

0
–1.5 –1 –0.5 0 0.5 1 1.5
(b) y/a

FIGURE 49.7  Normalized von Mises subsurface stress distribution for the x and y planes at the center of the
contact. Contact pressure distribution corresponding to Figure 49.6, for example, [59]. (a) von Mises stresses at the
center of the contact for y = 0. (b) von Mises stresses at the center of the contact for x = 0.

by evaluating the stress added to the contact from the over-rolling of dents that is the dented stress
integral of the rolling contact.

49.7.2  Calculation of the Contamination Factor ηc


Very much as in the case of rough lubricated contacts, the dented stress integral of the rolling contact is
compared to the stress conditions resulting from an ideally smooth raceway, that is, the smooth stress
integral. Thus, the following life ratio is quantified [59]:
Lubrication of Rolling Element Bearings 49-19

1/ e
⎛ L10.d ⎞ ⎛I ⎞ ⎛a ⎞ (49.16)
⎜ ⎟ = ⎜ smooth ⎟ = ⎜ SLF .dented ⎟
⎝ L10.s ⎠ (m,n,i ) ⎝ I dented ⎠ (m,n,i ) ⎝ aSLF . smooth ⎠ (m,n,i )

As in the analysis of Section 49.5, this ratio can be evaluated numerically using the fatigue integral
(Equation 49.10) and the FFT methodology developed in Reference 60. Following the hypothesis
adopted in Reference 8 (see Equation 49.7), the ratio (Equation 49.16) is assumed to be dependent of the
amount of surface denting (i.e., particle contamination density of the oil indicated with the index m), the
Hertzian size (i.e., proportional to the bearing mean diameter and indicated with the index n), and the
amount of interposed oil film of the rolling contact (indicated with the index i). For simplicity, the effect
of localized stresses from the overall roughness is decoupled from the stress concentration of dents
(assumed as the dominating effect): thus allowing to set ηb = 1. The stress-life factor (Equation 49.6) for a
standard bearing under the hypothesis of predominant effect of contamination particle induced denting
can then be written as

−c /e
w
1 ⎛ P ⎞
aSLF .dented = 1 − ⎜ hc u ⎟ (49.17)
10 ⎝ P ⎠

Similarly, in case of a bearing without surface denting, the contamination factor can be set to unity:
ηc = 1, and the stress-life factor then becomes

−c / e
w
1 ⎛P ⎞
aSLF . smooth = 1− ⎜ u ⎟ (49.18)
10 ⎝P ⎠

Inserting Equations 49.17 and 49.18 into Equation 49.16, the following is derived:

1/w
w −1/ c
P ⎛P ⎞ ⎛ I smooth ⎞
hc (m, n, i) = 1− 1− ⎜ u ⎟ ⎜⎝ I ⎟ (49.19)
Pu ⎝P⎠ dented ⎠ (m,n,i )

Equation 49.19 shows that a matrix (m,n,i) of numerically derived values of ηc can be constructed start-
ing from the numerical calculation of the volume-related fatigue-stress integral computed considering
different amount of contamination denting on a number of different bearing races. Also for this evalua-
tion, this can be accomplished using the rolling contact fatigue stress integral of Equation 49.10 applied
to different bearing contact conditions.
Following this approach [8,59], the life ratio of Equation 49.16 is evaluated to represent bearings
exposed to lubricants with different amount of contamination particles. In order to perform this calcu-
lation, it is required to have a measure of the population of dents that are found on typical raceways of
bearings exposed to lubricants with various degree of particle contamination. An example of statistical
measurement of the dent population found on the bearing raceway, Figure 49.8, can also provide a direct
representation of the effect of a given oil cleanliness and related operating conditions.
The evaluation of the stress conditions resulting from several types of dent distributions can be per-
formed in different ways: (1) using an explicit direct method, starting from the 3D mapping of actual
dented regions of bearing raceways and proceeding as explained in case of the lubrication factor stress
integral calculation or (2) implicitly, by computing the stress integral for different dent geometries (ref-
erence dents) and related lubrication conditions. These basic data can then be used to compute actual
49-20 Design for Lubrication and Tribology

100

70.8%

No. of dents per mm2


10 14.8%

6.5%

2.0% 2.1%
1
1.2%
1.0%
0.7%
0.5%
0.4%

0.1
5–15 25–35 45–55 65–75 85–95
15–25 35–45 55–65 75–85 >95

Equivalent dent diameter (µm)

FIGURE 49.8  Example of dent population statistics obtained from 3D sample maps of a rolling bearing raceway.
The bearing had operated under severe contamination conditions; comparable to an ISO 4406 cleanliness code clas-
sification within the range of -/19/16 to -/21/17. (From Gabelli, A. et al., Tribol. Trans., 51, 428, 2008. With permission.)

dented surface by multiplying the volume-related stress integral of each specific type of dent with the
number of dents (dent distribution) found on the dented region of the bearing. The application of this
method requires the automatic counting and categorization of dent population from 3D surface samples
of the bearing raceway. A dent counting and classification system can be applied to characterize differ-
ent dent patterns found in bearing applications, and given dent populations were qualified in relation to
the operating condition of the bearing, for example, in Figure 49.8.
For the use of the aforementioned methodology, the increase volume-related stress integral of a spe-
cific set of reference dent geometries can be carried out. This calculation is extended to include the effect
of the oil film thickness (lubrication conditions) on the resulting stress risers at the dent.
In this process, it is important to allow for the local oil film effects in the assessment of the con-
tamination factor. The stress reduction effect induced by the lubricant film, as shown in the calculation
example of Figure 49.9, can be significant and must be included in the analysis albeit with simplifica-
tions [59]. This method is suited to rate the contamination dent population found on an actual bearing
race, as, for instance, for the dent population shown in Figure 49.8, and relate this to a corresponding
range of cleanliness classification of the lubricant.
The numerically calculated stress integral of the dented region (additional volume-related stress inte-
gral of the dent population) is the parameter that characterizes the contamination factor of Equation
49.19. As illustrated in Figure 49.9, the magnitude and distribution of the stress rise at the dent from
a given dent geometry is strongly affected by the lubricant film present in the rolling contact. Thicker
lubricant films will result in a reduction (damping) and redistribution of contact stress developed at the
dent, while a negligible film thickness will sharpen the stress and raise the stress concentration to its
maximum.
To include this effect, stress damping related to the lubricant film was applied in the parametri-
cal evaluation of the dent contact pressure distribution, in accordance with the numerically calculated
results [60]. In this approach, the stress attenuation is related to the viscosity ratio κ of the bearing, under
the assumption that the expected averaged lubricant film of the rolling contact will be proportional to
Lubrication of Rolling Element Bearings 49-21

MPa MPa
3500 3500
3000 3000
2500 2500
2000 2000
1500 1500
1000 1000
500 500
0 0
0.2 0.2
0.1 0.2 0.1 0.2
0 0.1 0 0.1
0 0
–0.1 –0.1 –0.1 –0.1
y(mm) –0.2 –0.2 x(mm) y(mm) –0.2 –0.2
x(mm)

FIGURE 49.9  Example of contact pressure calculation of a typical dent (150 μm diameter and 5 μm depth) under
dry and lubricated conditions. Left: contact stress under dry condition (no lubricant film). Right: same dent show-
ing the contact stress attenuation induced by 0.55 μm oil film present in the rolling contact. In both cases, the
nominal Hertzian pressure is p 0 = 1.255 GPa. (From Gabelli, A. et al., Tribol. Trans., 51, 428, 2008. With permission.)

this lubrication parameter. Note also that bearing size variation will also affect the life ratio of Equation
49.16 and the corresponding ηc of Equation 49.19. Large bearings will have a large smooth stress inte-
gral, which will have a dominant effect over the dent stress integral. Furthermore, the maximum, dent-
related, stress concentration has a natural upper limit due to the maximum particle size that may be
transported in a lubricant stream; thus, it will not be dependent on the bearing size but rather only
affected by the cleanliness grade of the lubricant. Large diameter bearings have therefore an advantage
in terms of sensitivity to the effect of contamination, compared to bearings of smaller diameter.
Equation 49.19 provides a theoretical basis for the estimation of the contamination factor ηc, and this
can be used to establish an engineering model of ηc for the ranges of the cleanliness classification rat-
ing of the lubricant. In Reference 70, the contamination factor ηc is shown as function of κ for different
values of the bearing mean diameter, dm:


( ⎣ ) (
hc (k , dm )bcc = min c1(bcc )k 0.68dm0.55 ,1 ⎡1 − c2 (bcc )dm−1/3 ⎤
⎦ ) (49.20)

where c1 and c2 are constants assigned according to the oil cleanliness classification rating. This classi-
fication is based on the ISO 4406 cleanliness scale (or to a corresponding filtration quality grading, i.e.,
ISO 4372). Oil cleanliness levels may be assessed quickly and cost-effectively by means of particle coun-
ters, either in the laboratory or on-line directly in the lubrication circulating system of the machine,
according to NAS 1638 or ISO 4406:1999 [71]; see also Reference 17. It is possible, in many cases, to
correlate cleanliness classes with filter ratings. Thus, it is possible to use the filter efficiency, the beta
value βX, counting the number of particles upstream and downstream of the filter, as the contamination
characteristic required for the estimation of ηc:

Number of Particles > X micron − Upstream


bX ≥ (49.21)
Number of Particles > X miccron − Downstream

This possibility is included in ISO [1–4,8,16,17], and drawing on the experience of many applications, a
compilation of filter ratings and fluid cleanliness is provided by Needelman [72]:

b6 = 200

49-22 Design for Lubrication and Tribology

1.0
2000
0.9
1000
0.8
500
0.7
200
0.6
100
0.5
50
0.4
25
0.3
dm
0.2 mm
0.1

κ
0.1 0.3 0.5 0.7 0.9 1.1 1.3 1.5 1.7 1.9 2.1 2.3 2.5 2.7 2.9 3.1 3.3 3.5 3.7 3.9

FIGURE 49.10  Typical set of curves for ηc vs. lubrication parameter κ, bearing size (pitch diameter), and a mea-
sured level of contamination, for example, ISO –/13/10. (From Barnsby, R., Ed., Life Ratings for Modern Rolling
Bearings—A Design Guide for the Application of International Standard ISO 281/2, ASME STLE, ASME Tribology
Division Technical Committee, 2003. With permission.)

Figure 49.10 shows a set of curves for ηc vs. lubrication parameter κ, bearing size (pitch diameter), and a
measured level of contamination, for example, ISO 4406:1999 code-/13/10.
Similar diagrams are available by Ioannides et al. [8] and in ISO 281:2007 [16] for other cleanliness
classes (and filter ratings) and other types of lubrication, such as circulating, oil bath, and grease lubri-
cation. The differentiation between these types of lubrication systems follows from assumptions made
concerning the change of lubricant cleanliness due to abrasive particles during operation [16].

49.8  Bearing Friction Torque Calculation


The main function of rolling bearings is to transmit movement with low friction. In order to do this, roll-
ing bearings rely mainly on concentrated contacts working under the rolling mechanism itself, which
is more energy-efficient than sliding but which also generates friction. Rolling implies low-interaction
areas (“contacts”) and high concentrations of loads. Due to this and to the elastic deformation of the
contacting surfaces, some sliding is unavoidable. Other sources of energy loss appear in rolling bearings
because they are lubricated with either grease, which generally requires contact (sliding) seals, or oil,
which in many cases requires a lubricating bath. Apart from the economic impact, energy losses also
have an important effect in the bearing operation and life, since they increase the operating temperature
and reduce the viscosity.
The amount of friction in the bearing depends on the load and other factors. The most important are
the bearing type and size, the operating speed, the properties of the lubricant, and the quantity of the
lubricant. Friction occurs in many parts of the bearing where rolling and sliding occur, like the rolling
element and race contacts, flange contacts, and cage contacts. The friction in the bearing is very impor-
tant, because the generated heat may affect the operating temperature of the bearing and consequently
the viscosity of the lubricant. This, in turn, affects the separation of the contacts in the bearing which
finally affects the life of the bearing.
For the calculation of the friction torque, a number of models exist at different sophistication levels.
These can be found in the bearing manufacturers’ catalogs, for example, [1–4]. The simplest model con-
sists of a constant coefficient of friction which can be applied directly to the load and can be read in a
table for different types of bearings.
Lubrication of Rolling Element Bearings 49-23

49.8.1  Simple Model for Bearing Friction


The next level of friction torque is the two-term model. The two terms, M0 and M1, are the load-indepen-
dent and the load-dependent friction torque terms. The total friction torque in the bearing M is given by
the addition of these two components:

M = M0 + M1 (49.22)

1. The load-independent term M0 accounts for the following friction sources:


a. Shearing of oil while passing through the bearing contacts
b. Drag losses (oil lubrication) and ball/pocket friction (ball bearings)
2. The load-dependent term M1 accounts for the following friction sources:
a. Rolling friction, effect of lubrication
b. In some cases, steel hysteresis is considered

49.8.2  4S for Bearing Frictional Torque


A higher-level model of bearing friction is possible by separating out the fundamental sources of fric-
tion, rather than their load dependency [3]. This approach (so-called four sources model, 4S), besides
providing a more accurate calculation of friction, enables engineers to understand better the lubrication
process and quantification of the main sources of friction in the bearing for given application condi-
tions. The true physical friction sources in rolling bearings are rolling, sliding, seals, and drag losses.

49.8.2.1  Rolling Friction


Rolling friction energy losses are always present in rolling contacts (either dry or lubricated). There
are several sources of rolling friction. Energy is spent to introduce the lubricant into the contact
(Figure 49.11) and to reject the excess (EHL process). Elastic hysteresis losses in the steel (energy dis-
sipation in the deformation process) and even adhesion forces between surfaces are mechanisms that
can generate rolling friction. In order to account for the total rolling frictional moment in the bear-
ing, Mrr, first the load distribution in the different rolling element contacts must be established. This
depends on the bearing external loads, radial Fr and axial Fa, and on the bearing geometry (bearing

Rolling element

Rolling velocity

Rolling
resistance Pressure
Lubricant

Inner ring

FIGURE 49.11  Lubricant film thickness and pressure distribution in an elastohydrodynamically lubricated con-
tact. The large tail in the pressure at the inlet of the contact produces a resultant moment opposite to the rolling
direction. (Courtesy of SKF, Gothenburg, Sweden.)
49-24 Design for Lubrication and Tribology

type and size, number and size of rolling elements). The contributions of each contact are then added
together. Rolling resistance is also affected by two other factors—inlet shear heating and kinematic
replenishment/starvation.
Inlet shear heating occurs because not all the lubricant present at the inlet is entrained in the contact
and some of it will be rejected causing the reverse flow. This process produces heat, and the viscosity of
the lubricant is highly reduced by the temperature; lower viscosity at the inlet of the contact means lower
film thickness and, therefore, lower rolling resistance. This effect is taken into account in the 4S friction
model by means of the multiplication factor ϕish.
Kinematic replenishment/starvation occurs when high speeds or high lubricant viscosities hamper
the replenishment of lubricant in the raceway after a rolling element has passed. In this case, the lubri-
cant will not have sufficient time to flow back from the sides to the center of the raceway. This is kine-
matic starvation, which will produce a reduction of the lubricant availability in the inlet of the contact
and reduce the film thickness and the rolling resistance. The replenishment/starvation effect is consid-
ered in the 4S friction model by means of the multiplication factor ϕrs. This factor is also a function of
the lubricant supply mechanism. The inlet shear heating and kinematic replenishment/starvation fac-
tors may interact, but the model has been calibrated to take this interaction into account.

49.8.2.2  Sliding Friction


Sliding friction in rolling contacts is always present. There are two important sources of sliding in a roll-
ing contact: macro-sliding caused by contact conformity due to macro-geometry features, for example,
the contact between balls and curved raceways in ball bearings (osculation) and spinning, which is slid-
ing with angular velocity, and micro-sliding caused by the geometrical distortion from elastic deforma-
tion (Figure 49.12). The slip profile in the contact area will produce friction losses by means of lubricant
shearing and/or asperity contact, depending on the film thickness/roughness ratio.

1. Lubricant shearing: The friction coefficient due to lubricant shearing in one contact is

1
mEHL =
F∫t dA (49.23)
A

where
F is the normal load in the contact
τ is the shear stress in the lubricant
A is the contact area

The shear stress in the contact will depend on the slip profile (sliding speed) and the lubricant
rheology.
2. Asperity contacts: When the film thickness is not sufficient to completely separate the surfaces,
some asperity interaction will take place. This will increase the friction losses due to sliding, since
the friction coefficient of the asperity contacts is larger than the one from shearing the oil. The fol-
lowing equation gives the total friction coefficient coming from shearing the oil and from asperity
contacts:

msl = j bl mbl + (1 − j bl )mEHL (49.24)


where μbl is the friction coefficient between the interacting asperities; it is influenced by the lubricant
additive package. The function ϕbl is a weighting factor for the influence of asperity and lubricant shear-
ing mechanisms. Modeling and experimentation have been used to obtain the equation and diagram
provided, for example, in Reference 3.
Lubrication of Rolling Element Bearings 49-25

Rolling
element axis
Original raceway form

Contact area
A A1 Original rolling
element form
Direction of rotation

Sliding velocity

A A1
Contact area

Lines of rolling motion

FIGURE 49.12  Reverse lubricant rolling element raceway contact with a curved contact surface and the influ-
ence of elastic deformation. This depicts how sliding friction is produced. (Courtesy of SKF, Gothenburg, Sweden.)

49.8.2.3  Seal Friction


Seal friction is caused by the sliding between the lip of the seal and the moving steel counterface. For
contact seals, for example, indicated by designation suffixes RSH, RS1, LS, CS, CS2, and CS5 in bearing
manufacturers’ product catalogs, the contribution of the seal to the frictional moment represents a large
percentage of the total bearing friction. The 4S model provides a method to calculate the seal contribu-
tion to friction [3].

49.8.2.4  Drag Losses


The energy losses due to oil drag in an infinite lubricant bath (Figure 49.13). Drag losses represent the
energy loss due to the agitation of the oil caused by the rotation of the bearing. The oil being a viscous
fluid will exert a braking force in all moving surfaces in contact with it.

49.8.2.5  Complete Model


The total frictional moment in the bearing from the 4S model is

M = f ishf rs Mrr + M sl + M seal + Mdrag (49.25)



The model can reproduce typical Stribeck curves, as seen in friction measurements, with relatively
high friction in the low-speed (or low-viscosity) region, followed by an area of lower friction because
of the film buildup for higher speeds. As the oil film increases with even higher speeds, rolling friction
becomes dominant and will increase the frictional moment, until kinematic starvation and inlet shear
heating reduce the film thickness. The starting torque of the bearing can be calculated by setting the
speed equal to zero. Then friction will only come from sliding and seals.
49-26 Design for Lubrication and Tribology

D
d

Oil level H

FIGURE 49.13  Schematics of a bearing in a lubricating oil bath. The oil level H, as measured for the model, is
depicted. (Courtesy of SKF, Gothenburg, Sweden.)

49.8.2.6  4S Friction Model as an Engineering Tool


Due to the separation of the physical sources of friction, the 4S friction model can be used as an engi-
neering tool, since more information than only friction values is obtained from the model. Two exam-
ples will be given where the model is used to explain the functioning of the bearing.

49.8.2.7  Selection of Lubricant Viscosity


Using the results from Equation 49.25, it is possible to plot separately the different friction sources
for an open (unsealed) spherical roller bearing lubricated in an oil bath with very high viscosity
(Figure 49.14a). For very low speeds, there is a small area of high sliding losses, due to the asper-
ity interaction; very quickly the sliding losses decrease (due to the film buildup) to a steady value.
However, rolling losses grow from zero (at zero speed) and become dominant very quickly, with a
maximum value at about 500 rev/min. As velocity continues increasing, rolling friction decreases
because of kinematic starvation and/or inlet shear heating; this is no doubt a sign of excessively
high viscosity in the application. The factors ϕish and ϕ rs are now taking their toll. By looking at this
behavior, lower-viscosity oil might be recommended for this application.
The opposite behavior can also be demonstrated, where the same bearing and operating conditions
apply, but now the model has been run using very-low-viscosity lubricant (Figure 49.14b). In this case,
sliding losses are dominant in almost all the velocity spectrum. Certainly when increasing speed, slid-
ing losses decrease, but not quickly enough. In contrast, rolling losses are very small in comparison
with sliding. This is because the film thickness is so thin that the asperity interaction in sliding remains
dominant. The bearing could fail because of insufficient lubricant film. In this case, higher-viscosity oil
might be recommended for this application.
The required constants and information needed to apply this model can be found in Reference 3.

49.9  Closure
The basic methodology for the derivation of the lubrication and contamination factor embedded in
bearing life rating of ISO [16] and the bearing manufacturer’s catalogs [1–4] is based on a simple basic
theory of micro-EHL that can be applied for the evaluation of both factors. The theory is based on the
application of micro-EHL pressures and the parametrical evaluation of the Ioannides-Harris volume
stress integral related to smooth/rough and smooth/dented conditions of real bearing surfaces.
Lubrication of Rolling Element Bearings 49-27

50

Frictional moment, Nm
40

30

20

10

0
0 200 400 600 800 1000 1200 1400 1600 1800 2000
(a) Rotational speed, r/min
Total fridtional moment
Rolling frictional moment
Sliding frictional moment
Drag frictional moment

40
Frictional moment, Nm

30

20

10

0
0 200 400 600 800 1000 1200 1400 1600 1800 2000
(b) Rotational speed, r/min
Total fridtional moment
Rolling frictional moment
Sliding frictional moment
Drag frictional moment

FIGURE 49.14  Frictional moment vs. rolling speed at constant temperature, plotting the different moment com-
ponents for an open spherical roller bearing lubricated with oil bath. (a) Very-high-viscosity oil and (b) very-low-
viscosity oil. (Courtesy of SKF, Gothenburg, Sweden.)

The influence of contamination on bearing fatigue depends on a number of parameters including


bearing size, relative lubricant film thickness, size distribution of solid contaminant particles, and
type of contamination (e.g., soft or hard materials). The influence of these parameters on bearing life is
not only complex but it is further complicated by the fact that many of the parameters are difficult to
quantify with sufficient precision even in case of applying direct measurements. This has hindered past
attempts to address this problem by adopting purely empirical methods in the determination of rolling
bearing service life factors. The guidance obtained from comparisons with the experimental results is
invaluable because it provides guidance to the model so that its results are consistent with the observa-
tions of endurance life for the upper, lower, and intermediate level of contamination and lubrication
conditions used. The basic theory of the lubrication and the contamination factor clearly shows that the
quality of oil film and the cleanliness condition of the lubricant are important. This is essential for a
realistic prediction of the life expectancy of rolling bearings.
Thus, the lubrication and contamination factors are bearing life service factors designed to improve the
ability of the life model to represent the complexity of rolling contact fatigue life environment, under more
49-28 Design for Lubrication and Tribology

realistic service-like application conditions. It is clear from the previous sections that lubrication influences
the life of bearings and that good lubrication large κ values, with clean lubricant (little or no contamination),
extend the life of the bearings. Alternatively, these thicker films and clean lubricant allow downsizing, that
is, smaller bearings can carry the same load without loss of life. The separation of the working surfaces in
the bearing indicated by the lubrication parameter (κ) is related to the thickness of a lubricating film and the
combined roughness of the mating surfaces. The lubricating film, everything else being the same, depends
on the viscosity of the lubricant, which, in turn, is affected by the temperature of the bearing. The tempera-
ture of the bearing is the result of the equilibrium of the heat generated by the friction torque and the heat
transferred in and away from the bearing. This interdependence of the parameters describing lubrication
(and contamination) in a bearing, the operating temperature, and the bearing endurance emphasizes some
of the possibilities that a designer has in the design of bearings and their lubrication. It is important, how-
ever, to consider all the possible failure modes of bearings, ISO [73], as all the specifics of the application
for which the bearing is designed, or operates in, for example, load, lubrication, speed, and surrounding
atmosphere, can affect the performance and the duration of the bearing. The successful design will deliver
the required duration, with sufficient certainty and reasonable cost, taking into account all these factors,
and lubrication is a very important aspect of this.

References
1. FAG General Catalogue.
2. NSK General Catalogue.
3. SKF General Catalogue, 6000 EN, 2005.
4. TIMKEN General Catalogue.
5. Weibull, W., A statistical theory of the strength of materials, Proceedings of the Royal Swedish Institute
for Engineering Research, No. 151, 1939, pp. 4–45.
6. Lundberg, G. and Palmgren, A., Dynamic capacity of rolling bearings, Acta Polytech. Mech. Eng.
Series, 1(3), 1947.
7. Ioannides, E. and Harris, T.A., A new fatigue life model for rolling bearings, Trans. ASME J. Tribol.,
107, 1985, 367–378.
8. Ioannides, E., Berling, G., and Gabelli, A., An analytical formulation for the life of rolling bearings,
Acta Polytechnica Scandinavica, Mech. Eng. Series, (137), 1999.
9. Lorösch, H-K., Research on longer life for rolling-element bearings, ASLE Trans., 41(1), 1985, 37–43.
10. Voskamp, A.P., Material response to rolling contact loading, Trans. ASME J. Tribol., 107, 1985,
359–366.
11. Furumura, K., Murakami, Y., and Abe, T., The development of bearing steels for long life roll-
ing bearings under clean lubrication and contaminated lubrication, ASTM STP 1195, J. Hoo, Ed.,
Philadelphia, 1993, pp. 199–210.
12. Harris, T., Establishment of a new rolling contact bearing life calculation method, Final Report, U.S.
Navy Contract N68335-93-C-0111, 1994.
13. Kakuno, E. and Kawada, Y., A new criterion of fatigue strength of a round bar subjected to com-
bined static and repeated bending and torsion, Fatigue of Engineering Materials and Structures, 2,
1979, 229.
14. Crossland, B., The effect of pressure on the fatigue of metals, Mechanical Behaviour of Materials
Under Pressure, H. Pugh, Ed., Elsevier, London, U.K., 1970, pp. 299–354.
15. Dang Van, K., Griveau, B., and Message, O., Biaxial and multiaxial fatigue, Proceedings of Second
International Conference on Multiaxial Fatigue, December, Sheffield, U.K., 1985, pp. 479–496.
16. International standard: Rolling bearings—Dynamic load rating and rating life, ISO 281:2007.
17. Barnsby, R., Ed., Life ratings for modern rolling bearings—A design guide for the application of
International Standard ISO 281/2, ASME STLE, ASME Tribology Division Technical Committee,
2003.
Lubrication of Rolling Element Bearings 49-29

18. Tallian, T., Simplified contact fatigue life prediction model—Part II: New model, ASME Trans.
J. Tribol., 114, 1992, 214–222.
19. Zaretsky, E. V., Ed., STLE Life Factors for Rolling Bearings, STLE SP-34, STLE, Park Ridge, IL, 1992.
20. Brown, P.F., Bearings and dampers for advanced jet engines, SAE Paper No. 700318, April 1970.
21. Kadiric, A., Thermo-mechanical analysis of layered rough surface contacts, PhD thesis, University of
London, London, U.K., 2004.
22. Lugt, P.M., A review on grease lubrication in rolling bearings, Tribol. Trans., 52, 2009, 470–480.
23. Grubin, A.N., Investigation of the Contact of Machine Components, Kh.F. Ketova, Ed., Central
Scientific Research Institute for Technology and Mechanical Engineering, Moscow, Russia (DSIR
translation 337), Vol. 30.
24. Dawson, P.H., Effect of metallic contact on the pitting of lubricated rolling surfaces, J. Mech. Eng. Sci.,
4(1), 1962, 16–21.
25. Tallian, T.E., Chiu, Y.P., Huttenlocher, D.F., Kamenshine, J.A., Sibley, L.B., and Sindlinger, N.E.,
Lubricant films in rolling contact of rough surfaces, ASLE Trans., 7, 1964, 109–126.
26. Tallian, T.E., Sibley, L.B., and Valori, R.R., Elastohydrodynamic film effects on the load-life behaviour
of rolling contacts, ASME paper No. 65-LUBS-11, ASME Spring Lubrication Symposium, June 8, New
York, 1965.
27. Tallian, T.E., On competing failure modes in rolling contacts, ASLE Trans., 10, 1967, 418–439.
28. Skurka, J., Elastohydrodynamic lubrication of roller bearings, J. Lubr. Technol., 92(2), 1970, 281–291.
29. Danner, C.H., Fatigue life of tapered roller bearings under minimum lubricant film, ASLE Trans.,
13(4), 1970, 241–250.
30. Bamberg, E.N., Ed., Life adjustment factors for ball and roller bearings, ASME Engineering Design
Guide, ASME, New York, 1971.
31. Christensen, H., Stochastic models for hydrodynamic lubrication of rough surfaces, Proc. Inst.
Mech. Eng., 184(55), 1969–1970, 1013–1022.
32. Christensen, H. and Tønder, K., The hydrodynamic lubrication of rough bearing surfaces of finite
width, ASME J. Lubr. Technol., 93(3), 1971, 324–330.
33. Patir, N. and Cheng, H.S., An average flow model for determining effects of three-dimensional
roughness on partial hydrodynamic lubrication, ASME J. Lubr. Technol., 100, 1978, 12–17.
34. Patir, N. and Cheng, H.S., Application of average flow model to lubrication between rough sliding
surfaces, ASME J. Lubr. Technol., 101, 1979, 220–230.
35. Andreason, S. and Lund, T., Ball bearing endurance testing considering EHD lubrication, Proceedings
of Institution of Mechanical Engineers, Leeds EHL Symposium C36, Leeds, U.K., 1972, pp. 138–141.
36. Andreason, S. and Snare, B., Adjusted rating life of rolling bearings, Ball Bearing J., (184), 1975, 1–6.
37. Bolton, W.K., Elastohydrodynamics in practice, Rolling Contact Fatigue: Performance Testing of
Lubricants, R. Tourret and E.P. Wright, Eds., The Institute of Petroleum, London, U.K., 1977,
pp. 17–25.
38. Heemskerk, R., EHD lubrication in rolling bearings—Review of theory and influence on fatigue life,
Stratto da Tribologia e Lubrificazione, 4, 1980, 3–7.
39. Goglia, P.R., Cusano, C., and Conry, T.F., Effects of irregularities on the EHL of sliding line contacts,
Part 1—Single irregularities, Trans. ASME J. Tribol., 106, 1984, 104–112.
40. Goglia, P.R., Cusano, C., and Conry, T.F., Effects of irregularities on the EHL of sliding line contacts,
Part 2—Wavy surfaces, Trans. ASME J. Tribol., 106, 1984, 113–119.
41. Kweh, C.C., Evans, H.P., and Snidle, R.W., Micro-EHL of an elliptical contact with transverse and
three-dimensional sinusoidal roughness, Trans. ASME J. Tribol., 111, 1989, 577–583.
42. Lubrecht, A.A., The numerical solution of the elastohydrodynamically lubricated line and point
contact problem using multigrid techniques, PhD dissertation (ISBN 90-9001583-3), University of
Twente, Enschede, the Netherlands, 1987.
43. Venner, C.H., Multilevel solutions of the line and point contact problems, PhD dissertation (ISBN
90-9003974-0), University of Twente, Enschede, the Netherlands, 1991.
49-30 Design for Lubrication and Tribology

44. Sadeghi, F. and Sui, P.C., Compressible elastohydrodynamic lubrication of rough surfaces, Trans.
ASME J. Tribol., 111, 1989, 56–62.
45. Ioannides, E. and Kuijpers, H.C., Elastic stresses below asperities in lubricated contacts, Trans.
ASME J. Tribol., 108, 1986, 394–402.
46. Ai, A. and Cheng, H.S., Influence of moving dent on point EHL contact problems, STLE Tribol.
Trans., 37(2), 1994, 323–333.
47. Ai, A. and Lee, S.C., Effect of slide-to-roll ratio on interior stresses around a dent in EHL contacts,
STLE Tribol. Trans., 39(4), 1996, 881–889.
48. Chapkov, A.D., Colin, F., and Lubrecht, A.A., Influence on harmonic surface roughness on the
fatigue life of elastohydrodynamic lubricated contacts, Proceedings of the Institution Mechanical
Engineers., Part A, Journal of Power and Energy, 220, 2006, 287–294.
49. Tripp, J.H. and Ioannides, E., Effects of surface roughness on rolling bearing life, Proceedings of
Japan International Tribology Conference, October, Nagoya, Japan, 1990, pp. 797–802.
50. Greenwood, J.A. and Johnson, K.L., The behaviour of transverse roughness in sliding elastohydro-
dynamically lubricated contact, Wear, 153, 1992, 107–117.
51. Greenwood, J.A. and Morales-Espejel, G.E., The behaviour of transverse roughness in EHL contacts,
Proc. Inst. Mech. Eng. Part J J. Eng. Tribol., 208, 1994, 121–132.
52. Venner, C.H. and Lubrecht, A.A., Amplitude reduction of non-isotropic harmonic patterns in cir-
cular EHL contacts, under pure rolling, Proceedings of the 25th Leeds-Lyon Symposium on Tribology
1998, Elsevier Tribology Series, 34, D. Dowson et al., Eds., Elsevier, Amsterdam, the Netherlands,
1999, pp. 151–162.
53. Venner, C.H., Couhier, F., Lubrecht, A.A., and Greenwood, J.A., Amplitude reduction of waviness
in transient EHL line contacts, Proceedings of the 23rd Leeds-Lyon Symposium on Tribology 1996,
Elsevier Tribology Series, 32, D. Dowson et al., Eds., Elsevier, Amsterdam, the Netherlands, 1997,
pp. 103–112.
54. Hooke, C.J., Surface roughness modification in elastohydrodynamic contacts operating in the elastic
piezoviscous regime, Proc. IMechE, Part J J. Eng. Tribol., 212, 1998, 145–162.
55. Hooke, C.J., Li, K.Y., and Morales-Espejel, G.E., Rapid calculation of pressures and clearances in
rough, rolling-sliding elastohydrodynamically lubricated contacts. Part 1: Low-amplitude, sinusoi-
dal roughness, Proc. IMechE, Part C J. Mech. Eng. Sci., 221, 2007, 535–550.
56. Hooke, C.J., Li, K.Y., and Morales-Espejel, G.E., Rapid calculation of pressures and clearances in
rough, rolling-sliding elastohydrodynamically lubricated contacts. Part 2: General, non-sinusoidal
roughness, Proc. IMechE, Part C J. Mech. Eng. Sci., 221, 2007, 551–564.
57. Choo, J.W., Olver, A.V., and Spikes, H.A., The influence of transverse roughness in thin film, mixed
elastohydrodynamic lubrication, Tribol. Int., 40, 2006, 220–232.
58. Félix-Quiñonez, A., Ehret, P., and Summers, J.L., On three-dimensional flat-top defects passing
through an EHL point contact: A comparison of modeling with experiments, Trans. ASME J. Tribol.,
127, 2005, 51–60.
59. Gabelli, A., Morales-Espejel, G.E., and Ioannides, E., Particle damage in Hertzian contacts and life
ratings in rolling bearings, Tribol. Trans., 51, 2008, 428–445.
60. Morales-Espejel, G.E., Lugt, P.M., van Kuilenburg, J., and Tripp, J.H., Effects of surface
micro-geometry on the pressures and internal stresses of pure rolling EHL contacts, Tribol. Trans.,
46, 2003, 260–272.
61. Morales-Espejel, G.E., Gabelli, A., and Ioannides, E., Micro-geometry lubrication and life ratings of
rolling bearings. Proc. IMechE Part C J. Mech. Eng. Sci., 224, 2010, 2610–2626.
62. Xu, G., Sadeghi, F., and Hoeprich, M.R., Residual stresses due to debris effects in EHL contacts,
Tribol. Trans., 40(4), 1997, 613–620.
63. Dommarco, R.C., Bastias, P.C., Hahn, G.T., and Rubin C.A., The use of artificial defects in
the 5-ball-rod rolling contact, Wear, 252, 2002, 430–437.
Lubrication of Rolling Element Bearings 49-31

64. Xu, G., Sadeghi, F., and Cogdell, J.D., Debris denting effects on elastohydrodynamic lubricated con-
tacts, ASME J. Tribol., 119, 1997, 579–587.
65. Nelias, D. and Ville, F., Detrimental effects of debris dents on rolling contact fatigue, ASME J. Tribol.,
122, 2000, 55–64.
66. Ville, F., Coulon, S., and Lubrecht A.A., Influence of solid contaminants on the fatigue life of lubri-
cated machine elements, Proc. IMechE Part J J. Eng. Tribol., 220, 2006, 441–445.
67. Sayles, R.S. and Ioannides, E., Debris damage in rolling bearings and its effect on fatigue life, ASME
J. Tribol., 110, 1988, 26–31.
68. Webster, M.N., Ioannides, E., and Sayles R.S., The effect of topographical defects on the contact
stress and fatigue life in rolling element bearings, Proceedings of 12th Leeds-Lyon Symposium on
Tribology, September 3–4, Lyon, France, 1985.
69. Kang, Y.S., Sadeghi, F., and Hoeprich, M., A finite element model for spherical debris denting
in heavily loaded contacts, ASME J. Tribol., 126, 2004, 71–80.
70. Harris, T., Rolling Bearing Analysis, 4th edn., Wiley Interscience Publication, New York, 2001,
pp. 915–921.
71. Hydraulic fluid power—Fluids—Method for coding the level of contamination by solid particles
ISO 4406:1999.
72. Needelman, W.M. and Zaretsky E.V., New equation show oil filtration effects on bearing life, Power
Transm. Des., 33(8), 1991, 63–70.
73. International standard: Rolling bearings—Damage and failures—Terms, characteristics and causes
ISO 15243:2004.
50
Gear Lubrication
50.1 Introduction.....................................................................................50-1
50.2 Gear Failure Modes.........................................................................50-2
50.3 Gear Types........................................................................................50-4
50.4 Lubricant Selection and Delivery Method...................................50-8
Liquid Lubricant Delivery Methods
50.5 Liquid Lubricant Flow Rate Requirements................................50-12
50.6 Film Thickness Calculation.........................................................50-13
Robert F. 50.7 Flash Temperature Calculation................................................... 50-14
Handschuh 50.8 Lubrication Starvation.................................................................. 50-16
National Aeronautics and 50.9 Summary Comments on Gear Lubrication...............................50-18
Space Administration References...................................................................................................50-18

50.1  Introduction
Gear lubrication is usually the third consideration after the bending strength and contact stress have
been satisfied on a particular design. Gear lubrication provides two principal functions: lubrication and
cooling. Without these two lubrication-related functions, a successful design from the bending strength
and load capacity view will result in early failure.
After the life of component is determined based on expected load and number of cycles, other
­application-specific variables now must be evaluated. The two most prevalent are the gear pitch line
velocity and the expected operational and environmental conditions of the device. These important
design parameters will dictate how the lubrication requirement is satisfied, what type of lubricant will
be used, and what, if any, other systems are required to make the design successful.
Success and failure mechanisms can be shown on this rather simple graph. In Figure 50.1, the effect
of speed and load on successful designs and many gear failure types are shown.
As far as the mechanisms shown on this figure, scoring and wear can be caused by excessive load, excessive
rotational speed, and/or the combination of both. It is desirable to stay within the “no failure” zone, but usu-
ally, design constraints make the end configuration a compromise of satisfying various gearing parameters
to avoid short operational life of the system at hand. Gear speed is usually quantified via pitch line velocity.
Scoring and wear are the two predominant lubrication-related failure mechanisms as shown in Figure 50.1.
Pitch line velocity for a given gear—this quantity is found via the following equation:

Vpl = (RPM) ⋅ (pd p )


where
Vpl is the pitch line velocity in ft/min
RPM is the gear revolutions per minute
dp is the pitch diameter of the gear under study in feet

50-1
50-2 Design for Lubrication and Tribology

Breakage

Pitting
Load
Wear
(lacks oil film
due to slow
speed)
Scoring
(too much heat
due to high speed)
No
failure
(good oil film)

Speed

FIGURE 50.1  Role of rotational speed and applied load to various gear failure mechanisms. (From Dudley, D.,
Handbook of Practical Gear Design, McGraw-Hill Book Co., New York, 1984. With permission.)

In metric units, the equation is the following:

⎛ pd p ⎞
Vpl = (RPM) ⋅ ⎜ ⎟
⎝ 60 ⎠

where
Vpl is the pitch line velocity in m/s
RPM is the gear rev/min
dp is the pitch diameter of the gear under study in meters

The pitch diameter of a gear is given by the following equation:

Pitch diameter = (number of teeth) / (diametral pitch) … English units


Pitch diameter = (number of teeth) * (module)… SI units


50.2  Gear Failure Modes


Possible gear failure modes were shown in Figure 50.1. Usually, the gear bending and contact stress
calculations are conducted initially. If the gear mesh will not meet the load capacity (bending and con-
tact stress) and life requirements, the design is altered until these requirements are met. The standards
proposed by the American Gear Manufacturers Association and ISO (ISO 6336) are a way to satisfy
these basic requirements. Once the design is satisfactory for the bending and contact stress standpoint,
then other gear-related issues are addressed. Bending and contact fatigue examples are shown later.
Bending fatigue usually results in crack initiation and propagation from the location of highest stress—­
typically in the fillet-root region. Examples are shown for spur and spiral bevel gears in Figures 50.2 and
50.3, respectively. Also, more details of gear failure types and appearance can be found in the following
Gear Lubrication 50-3

FIGURE 50.2  Bending fatigue of a spur gear.

FIGURE 50.3  Bending fatigue of a spiral bevel pinion.

reference material: Drago (1988), Bartz (1993), Alban (1993), ANSI/AGMA 1010-E95 (2007), and AGMA
912-A04 (2004).
Contact fatigue is initiated typically along the active profile of the surface where load sharing is mini-
mized (highest point of single tooth contact) and there is relative sliding among the two meshing gear
surfaces. The actual contact stress in many gear systems is in excess of 1.7 GPa (250 ksi). The material
can crack subsurface and propagate until the gear surface can no longer carry the applied load. This type
of failure is shown from test hardware in Figure 50.4. When the surface starts to pit as shown in Figure
50.4, the smooth transfer of load between teeth becomes interrupted. This surface failure will then tend
to spread to neighbor teeth due to the increase in dynamic loading.
These two types of gear failures were shown to present the physical differences from these to the ones
caused by lubrication issues. Figure 50.5 is an example of scoring of the gear surface. When scoring
occurs, the lubricating film was insufficient to keep the meshing surfaces separated. This can occur from
many causes such as rough tooth surface profile, insufficient lubricant viscosity, lack of extreme pressure
(EP) additives, high overall operating temperature, high or very low pitch line velocity or sliding veloc-
ity between the meshing surfaces, and applied load. When scoring occurs, locally high temperature of
50-4 Design for Lubrication and Tribology

FIGURE 50.4  Contact fatigue of a spur gear.

FIGURE 50.5  Spiral bevel gear that scored then surface pitting followed.

the surface can be reached instantaneously. This locally high temperature can result in surface hardness
reduction (annealing). The end result will be a gear that had scoring damage that will then lead to short
operation life due to contact or bending failures.

50.3  Gear Types


Different types of gearing can be put into three general categories. There are parallel axis gearing that
includes spur, helical, and double helical gear meshes. This first type of gearing is shown in Figures 50.6
through 50.8.
The three gear meshes are shown, having both members with external teeth. Each of these gear
meshes can also be manufactured with an external–internal gear mesh. Detailed information of
external and internal gear spur gears is shown in Figures 50.9 and 50.10, respectively. Single or
Gear Lubrication 50-5

FIGURE 50.6  Spur gear mesh.

Single-helical teeth

FIGURE 50.7  Single-helical gear mesh.

Double-helical teeth

FIGURE 50.8  Double helical gear mesh (herringbone shown earlier).

­ ouble helical gears have the same basic detailed information as shown later in the normal plane to
d
the helix angle.
The second type of gear mesh is that described as intersecting axis gearing. These types of gears are
generally called bevel, spiral bevel, and crown (face) gears and are shown in Figure 50.11. These gears can
intersect at various angles as shown in the following.
50-6 Design for Lubrication and Tribology

Pinion
Outside rb , 1 r
p,1
Line of action diameter ra , 1
(pressure line)
Base j Pitch circle
circle
Pressure
angle,j A
S Tooth profile
Pitch circle Distance
B Whole depth between
Addedum centers, C
Base Line of Root (tooth)
circle Working centers fillet
depth
Clearance Top
Dedendum land
j rb , 2
Circular tooth
rp , 2
thickness, t

Chordal tooth
ra , 2
thickness

Circular
pitch, p Gear

FIGURE 50.9  External–external gear mesh.

Pitch point
Clearance

Line of action
(pressure line)
Circular pitch
Pressure angle, j
Base circle of Base circle
internal gear diameter
F Distance between
Circular centers, C
thickness Pitch diameter
Outside diameter

Internal diameter

Root diameter

Tooth fillet Dedendum

Addendum
Working depth

Whole depth

FIGURE 50.10  External–internal gear mesh.

Nomenclature for this type of gearing is shown in Figure 50.12. Many of the same gear nomenclature
as shown earlier for parallel axis gears are similar. Addition of the pitch angles of a pinion and gear of
this type of gear mesh results in the shaft angle of the gear mesh. Many of the gear geometry parameters
are calculated at the furthest point along the tooth from the apex of the pitch cone. The axes of the pitch,
root, and face angles are shown intersecting at a common apex in Figure 50.12; however, this is not a
requirement. Sometimes, depending on the design requirements, these intersection locations are differ-
ent for each of the pitch, root, and face angles to increase strength, load capacity, or overall contact ratio.
Gear Lubrication 50-7

45°

Usual form Mitre gear

Mitre gear Mitre gear

Crown gear Internal bevel gear

FIGURE 50.11  Intersecting axis gear mesh arrangements.

Apex of pitch cone

Face cone
Pitch cone
Root cone Root Cone distance
angle Addendum angle
Pitch angle
Pitch Face Dedendum angle
Face width angle angle

Addendum
Dedendum
Pitch
angle

Outside pitch diameter

Inside pitch diameter


Pitch radius of
Back cone equivalent spur gear
distance

Outside diameter

FIGURE 50.12  Bevel gear nomenclature.

The last type of gearing is designated as nonintersecting axis gears. Gears of this type include hypoid,
crossed helical, and worm gears. The general arrangements of these gear types are provided in Figures
50.13 through 50.15.
Other important aspects for a gear arrangement are the ratio range, maximum pitch line velocity, and
anticipated efficiency. Most gearing arrangements are very efficient. Only worm gears, crossed helical
gears, and very-high-ratio hypoid gears have less efficient operation. Table 50.1 (Coy et al., 1985) pro-
vides a basic comparison of various types of gear efficiency and maximum pitch line velocity.
50-8 Design for Lubrication and Tribology

FIGURE 50.13  Hypoid gear mesh (typically used in auto or truck rear axle).

FIGURE 50.14  Crossed helical gear mesh.

Worm

ω
W

WX

λW

j jn
Wr Wt WN

ωG
λW

Worm wheel

FIGURE 50.15  Worm gear mesh.

50.4  Lubricant Selection and Delivery Method


Lubrication system delivery and lubricant selection are important considerations that can determine
the successful operation of a drive system design. Low-speed gearing may have an open gear mesh
that transfer very high loads at relatively low rotational speed. In this case, grease lubrication may be
Gear Lubrication 50-9

TABLE 50.1  Gear Type, Pitch Line Velocity, and Expected Efficiency
Nominal Gear Maximum Pitch Line Nominal Efficiency
Type Ratio Range Velocity, m/s (ft/min) at Rated Power (%) Remarks
Parallel shafts
Spur 1–5 100 (20,000)a 97–99.5
1–5 20 (4,000)b 97–99.5
Helical 1–10 200 (40,000)a 97–99.5 15°–35° helix angle
1–10 20 (4,000)b 97–99.5
Double helical 1–20 200 (40,000)a 97–99.5
1–20 20 (4,000)b 97–99.5
Intersecting shafts
Straight bevel 1–8 50 (10,000)a 97–99.5
1–8 5 (1,000)b 97–99.5
Spiral bevel 1–8 125 (25,000)a 97–99.5 35° spiral angle most used
1–8 10 (12,000)b 97–99.5
Zerol bevelc 1–8 50 (10,000)a 97–99.5
1–8 5 (1,000)b 97–99.5
Nonintersecting, nonparallel shafts
Worm 3–100 50 (10,000)a 50–90
3–100 25 (5,000)b 50–90
Double-enveloping 3–100 50 (10,000)a 50–98
worm
3–100 20 (4,000)b 50–98
Crossed helical 1–100 50 (10,000)a 50–95
1–100 20 (4,000)b 50–95
Hypoid 1–50 50 (10,000)a 50–98 Higher ratios have low efficiency
1–50 20 (4,000)b 50–98
Source: Coy, J. et al., Gearing, NASA RP 1152, AVSCOM TR-84-C-15, 1985.
a Aircraft high precision.

b Commercial.

c Registered trademark, The Gleason Works, Rochester, New York.

applied via a brush or sprayed at intervals to avoid excessive tooth wear. On open gearing, the grease
will be pushed out of the contacting region of the teeth just due to meshing, and the gears rely on
residual grease, maintaining some level of lubrication. Some typical gear and bearing greases are shown
in Table 50.2 for aerospace applications. More information on lubrication selection can be found in
AGMA (2004).
Additional information can be found in the American Gear Manufacturers Association standards on
industrial gear lubrication in ANSA/AGMA 9005-E02 (2008) and Bartz (1993).

50.4.1  Liquid Lubricant Delivery Methods


In moderate- to high-speed gear applications, liquid lubrication is typically utilized. Table 50.3 provides
some of the more common liquid lubricants and their operating temperature range.
Also, some basic flow rate and jet location information is contained in Tables 50.4 and 50.5.

50.4.1.1  Splash Lubrication


There are a couple of common ways to design these systems. One option is to use the gears themselves
to distribute the lubricant. This type of gearbox is called splash lubrication. The gears dip down into the
50-10 Design for Lubrication and Tribology

TABLE 50.2  Greases Used in the Lubrication of Gears and Bearings


Recommended
Type Specification Temperature Range °C (°F) General Composition
Aircraft: high-speed ball and MIL-G-38220 −40 to 204 (−40 to 400) Thickening agent and fluorocarbon
roller bearing
Aircraft: synthetic EP MIL-G-23827 −73 to 121 (−100 to 250) Thickening agent, low-temperature
synthetic oils or mixtures of EP additives
Aircraft: synthetic MIL-G-21164 −73 to 121 (−100 to 250) Similar to MIL-G-23827 plus MoS2
molybdenum disulfide
Aircraft: general purpose, MIL-G-81322 −54 to 177 (−65 to 350) Thickening agent and synthetic
wide operating temperature hydrocarbon
range
Helicopter: oscillating MIL-G-25537 −54 to 71 (−65 to 160) Thickening agent and mineral oil
bearing
Plug valve: gasoline and oil MIL-G-6032 0 to 95 (32 to 200) Thickening agent, vegetable oils, glycerols,
resistant or polyesters
Aircraft: fuel and oil resistant MIL-G-27617 −1 to 204 (D930 to 400) Thickening agent and fluorocarbon and
fluorosilicone
Ball and roller bearing: MIL-G-25013 −73 to 232 (−100 to 450) Thickening agent and silicone fluid
extreme high temperature
Source: Coy, J. et al., Gearing, NASA RP 1152, AVSCOM TR-84-C-15, 1985.

TABLE 50.3  Typical Liquid Lubricant Operational Considerations


Type Specification Bulk Temperature Limit in Air, °C (°F)
Mineral oil MIL-L-6081 93 + (200+)
Diester MIL-L-7808 149 + (300+)
Type II ester MIL-L-23699 218 (425)
Triester MIL-L-9236B 218 (425)
Super-refined and 218 (425)
synthetic mineral oils
Fluorocarbona 288 (550)
Polyphenyl ethera 315 (600)
Source: Coy, J. et al., Gearing, NASA RP 1152, AVSCOM TR-84-C-15, 1985.
a Not recommended for gears.

TABLE 50.4  Power Level in hp per Flow Rate in gpm for the Following Lubricant
Temperature Rise (°F)
hp/gpm Flow Condition Typical Temperature Rise Comments
400 Copious <50°F General industrial
800 Adequate <75°F Typical aviation
1200 Lean <100°F Lightweight, high-efficiency aviation
1600 Starvation >100°F Only use for unusual applications
Source: Drago, R., Fundamentals of Gear Design, Butterworths Publishers, Oxford, U.K., 1988.

lubricant and distribute by the rotation of one of the gear members through the lubricant. This type of
system is very simple since there are no pumps, reservoirs, filters, etc. The use of housings, lubricant
slingers, and covers must be carefully designed so that all dynamic components are properly lubricated.
The system, when done correctly, works very well, with long life and little maintenance ever needed. This
system is shown in Figure 50.16.
Gear Lubrication 50-11

TABLE 50.5  Liquid Lubrication Suggested Delivery Method and Jet Locations
Pitch Line Velocity m/s (ft/min) Jet Location Comments
15.2 (3,000) None Splash lubrication is adequate
15.2–25.4 (3,000–5,000) Into mesh Lubrication is the primary function, with cooling
secondary; splash lubrication with baffles and channels
can be sufficient
25.4–61.0 (5,000–12,000) Out of mesh or into mesh Cooling is primary function; sufficient lubricant will
adhere to teeth for lubrication
61.0–101.6 (12,000–20,000) Out of mesh Cooling is dominant need; large out-of-mesh flow needed
> 101.6 (20,000) Into mesh and out of mesh Low flow into mesh for lubrication, large flow out of
mesh for cooling
Source: Drago, R., Fundamentals of Gear Design, Butterworths Publishers, Oxford, U.K., 1988.

Hypoid splash lubricated gearbox

Gearbox

Output
gear

Input
pinion

Lubricant
level

FIGURE 50.16  Example of hypoid splash-lubricated gearbox.

There are some drawbacks from a system such as this. First of all, since the system is closed and there
is no filtration of the lubricant, debris and wear particles can find their way through the meshing inter-
faces of the gears and bearings. This in turn will result in more wear. Also, the gears not only transmit
power, but they must also distribute the lubricant to the rest of the dynamic components within the
gearbox. This results in power being expended by driving at least one of the gear members through the
contained lubricant. Therefore, the efficiency is not as high as it could be. The heat generated within
the gearbox has to be absorbed by the lubricant and convected away from the housing. There is no heat
exchanger to help dissipate the heat generated so the lubricant chosen for this type of system typically
has high room temperature viscosity to be able to operate at the steady-state condition where the heat
generated within the gearbox balances with the heat convected away.

50.4.1.2  Dry Sump Lubrication


If gearbox efficiency is of most concern and the components are operating above 15 m/s (3000 ft/min)
pitch line velocity, then an alternative and more complex system will be necessary. An example of a sys-
tem that would be needed is shown in Figure 50.17. Lubricant is minimized in the gearbox.
This system is much more complex, but the gear system will operate at its best efficiency because of the
following reasons: lubricant is jet fed to where it is needed; lubricant is scavenged away from the gears,
thus reducing gear drag parasitic losses (windage and pumping losses); the lubricant is filtered prior to
being jet fed to the gears and bearings; and the heat removed from the gearbox is dissipated at the heat
exchanger. All of these reasons make this the most efficient system; however, the complexity has been
increased significantly over a splash liquid lubrication.
50-12 Design for Lubrication and Tribology

Lubricating jet

Gearbox

Gearbox
Scavenge
drain
pump

Relief/
pressure
Control
value Heat
Reservoir exchanger

Supply
Filter
pump

FIGURE 50.17  Liquid lubrication system.

50.5  Liquid Lubricant Flow Rate Requirements


Based on the power that the gear mesh must transfer and the expected efficiency, the amount of lubri-
cant flow rate for an expected power loss can be found for a liquid lubricant that is jet fed and scavenged
away (assuming all heat will be absorbed by the lubricant):

´
Q = m Υ C p (Toutlet − Tinlet )

where
Q is the power that will be absorbed
mÝ is the lubricant flow rate
Cp is the heat capacity of the lubricant
(Toutlet − Tinlet) is the temperature change that will occur for a given flow rate to absorb the power
expended to the lubricant

The amount of heat needed to be rejected is based on the efficiency of the gears, bearings, and seals of
the drive system. The flow rate of lubricant is based on the lubricant’s heat capacity and the amount of
temperature rise of the lubricant that is permissible for the design application.
A rough sizing equation for jet flow can be found as follows (Drago, 1988):

m Υ´ = 22 ⋅ d 2 ⋅ ΔP
Gear Lubrication 50-13

where
mY is the flow rate (gpm)
d is the jet diameter (orifice) (in.)
ΔP is the pressure drop of the jet (psi)

50.6  Film Thickness Calculation


When assessing lubrication requirements, one of the important calculations to be made is the film
thickness. The film thickness, when combined with the surface roughness, called the “lambda
ratio,” is found. The lambda ratio is found from the lubricant film thickness divided by the surface
roughness:

hmin
l= 1/ 2


(Δ 2
a + Δ b2 )
where
λ is the lambda ratio
hmin is the minimum lubricant film thickness
Δa and Δb are the rms gear surface finishes of surface a and b, respectively

Figure 50.18 indicates the various lubrication regimes for gear contacts.
If this ratio is greater than 1, the gear contact is in the full elastohydrodynamic regime. In this case,
the surface roughness is less than the film thickness, and asperities on the meshing gear surfaces should
not interact.
If this ratio is much less than 1 (less than about 0.3), then the contact is called lubrication-starved
regime. In this case, the asperities will interact, and wear (scuffing) and pitting will result.
In between these two regimes is the mixed elastohydrodynamic region. Here, there will be some
asperity contact, but lubrication film will minimize this effect. Many drive system components that
operate at low rotational speeds or low pitch line velocity are classified into this mixed lubrication
regime.
As part of the calculation procedure, the film thickness between the meshing gears requires calculation.
This calculation depends on the curvatures of the meshing gears at the point of contact where the com-
bined effects of load, sliding and rolling velocity of the contact, and the assumed lubrication conditions
are the most extreme. If the current discussion is limited to spur gears (two dimensional), the procedure

Vp

Vg
Vp

Vp Vg

Vg
Elasto-
Metal Boundary Mixed hydrodynamic
contact lubrication lubrication lubrication

(Minimum lubricant film thickness)/(Sum of the square of roughness height of contacting bodies)

FIGURE 50.18  Film thickness and surface roughness effect on lubrication regime.
50-14 Design for Lubrication and Tribology

to do the calculation follows. The procedure can be found in many texts and reports, but the procedure
found in Coy et al. (1985) and Hamrock (1984) will be followed. In these references, the film thickness
is based on the calculation of several nondimensional parameters. These parameters include material,
geometry, speed, and lubricant related. The nondimensional equation is the following (Anderson and
Lowenthal, 1980):

h
H min = = 2069U 0.67G 0.53W −0.67 (1 − 0.61e −0.73k )
Rx

where U is the speed parameter:

umo
U=
E ʹR x

where

Vg + V p
u=
2

u is the average rolling velocity, μo is the lubricant viscosity, E′ is the equivalent modulus of elasticity,
and R x is the effective radius in the direction of rolling. E′ is given by the following:

2
Eʹ =

((1 − g ) /E ) + ((1 − g ) /E )
2
g g
2
p p

G is the material parameter:

G = Eʹa

α is the pressure viscosity coefficient of the lubricant.


Finally, W is the load parameter:

FH
W=
EʹRx2

where FH is the normal applied load.


If the surface roughness is known either from the specification provided or from actual measure-
ment, then the lambda ratio can be found and the lubrication regime identified.
A way to describe film thickness effects on surface life is shown in Figure 50.19. The ratio
of film thickness to surface roughness can have a surface life benefit if the ratio is in excess of 2
(Townsend, 1992).
Another way to examine the lambda ratio effects on possibility of surface wear-related effects is
shown in Figure 50.20. In this figure, the probability of wear-related problems is proved as a function of
pitch line velocity.

50.7  Flash Temperature Calculation


Another way to assess a design for lubrication-related issues is to conduct a “flash temperature” evalua-
tion. In Dudley (1984), the equation to make this calculation is the following:
Gear Lubrication 50-15

Usual operating
region
100

80

Percent film 60
Region of Possible Region of
lubrication- surface increased
40 related sur- distress life
face distress with
20 severe
sliding

0
0.3 0.4 0.6 0.8 1.0 2.0 4.0 6.0 8.0 10
Film parameter, Λ

FIGURE 50.19  Lambda ratio and effects on surface performance.

3.0

2.0
5%

40%
1.0
80%

0.5
Probability of
Specific film thickness, λ

wear-related distress

0.2

0.1

0.03

0.02

0.01
50 100 500 1,000 5,000 10,000 50,000
Pitch line velocity, fpm

FIGURE 50.20  Film thickness ratio lambda versus pitch line velocity with probable wear-related surface distress.
50-16 Design for Lubrication and Tribology

3/ 4
⎛W ⎞ n1p/2
T f = Tb + Zt ⎜ t ⎟
⎝ Fe ⎠ Pd3/ 4

⎛ r − r /m ⎞ Pd1/ 4
1 2 G
Zt = 0.0175 ⎜ ⎟
⎜⎝ ( rr
1 2/( r1+r2))
1/ 4
⎟⎠ (cos j t )3/ 4

where
Tf is the flash temperature (°F)
Tb is the blank temperature and usually taken to be the inlet oil temperature (°F)
Wt is the tangential driving load (lb)
Fe is the face width in contact (in.)
np is the rotation speed of the pinion (rev/min)
Pd is the diametral pitch (1/in.)
ρ1 and ρ2 are the radius of curvatures for the pinion and gear, respectively, at the point of contact
under consideration
mG is the gear ratio
ϕt is the transverse pressure angle (deg)

Radius of curvatures ρ1 and ρ2 are found from the tangent from the base circles of each gear to the point
of contact under consideration.
Normally, the flash temperature will be calculated at the pinion tip where the sliding velocity is the
greatest, but if there is tip modifications applied, then the highest flash temperature can be calculated
at the highest and lowest point of single tooth contact for the pinion. Table 50.6 provides guidance on
maximum flash temperature values for some common lubricants.
In high-speed gearing applications, one other consideration must be made. As mentioned in Dudley
(1984), spur gears and straight bevel gears are not very efficient at expelling air-lubricant efficiently. If
pitch line velocity is greater than 50 m/s (10,000 ft/min), this phenomenon can become an issue. At pitch
line velocities above this level, face width to diameter ratio should be low.
Axial meshing velocity can be found from the following equation:

pdn p
Vaxial = (ft/ min) English units
12 tan Ψ

pdn p
Vaxial = (m/s) SI units
60, 000 tan Ψ

where
d is the pitch diameter of the pinion in inches or mm
np is the pinion speed in revolutions per minute
Ψ is the helix angle.

A guide to determine the severity of the axial meshing velocity is provided in Table 50.7.

50.8  Lubrication Starvation


Lubrication starvation is an extreme condition where the combination of high speed, load, and other
operational parameters can exist such that destruction of the gear meshing components can occur
Gear Lubrication 50-17

TABLE 50.6  Flash Temperature Maximum to Prevent Scoring in Spur Gears


Type of Lubricant Specification Flash Temperature (°F)
Petroleum SAE 10 250
SAE 30 375
SAE 60 500
SAE 90 (gear lubricant) 600
Diester, compounded 75 SUS at 100°F 330
Petroleum SAE 30 plus mild EP additive 425
Source: Dudley, D., Handbook of Practical Gear Design, McGraw-Hill Book Co.,
New York, 1984, Table 2.9.

TABLE 50.7  Helical Gear Meshing Velocity Evaluation


Axial Meshing Velocity
m/s (ft/min) Severity
400 (80,000) No trouble except in very large units where thermal distortion may be enough to
require correction (needs good oil-jet system and generous size casing)
500 (100,000) Probably no serious trouble (needs good oil-jet system and generous size casing)
700 (140,000) Probably have some trouble. May be manageable if gears are not too large and
thermal distortions are handled by compensations in tooth fit
850 (170,000) Usually difficult to handle. Much skill in tooth compensations needed plus
special quality of lubricant
1,000 (200,000) Probably impractical to handle even with utmost design skill
Source: Dudley, D., Handbook of Practical Gear Design, McGraw-Hill Book Co., New York, 1984, Table 7.11.

FIGURE 50.21  Lubrication-starved gear.

rather rapidly and, in some cases, instantaneously. In aviation applications, this can occur due to the
complete failure of the primary lubrication system from pump failure and large leakage rates of the
lubrication system or in military applications, from ballistic damage to the gearbox housing or lubrica-
tion system. An example of this type of failure is shown in Figure 50.21. The gear shown was run nearly
to destruction after the primary lubrication system was shut off to simulate an aerospace gearbox expe-
riencing loss of primary lubrication.
50-18 Design for Lubrication and Tribology

50.9  Summary Comments on Gear Lubrication


Gear lubrication considerations in the design of any drive system are of extreme importance. While
initial design challenges need to be satisfied for gear load capacity (bending and contact), failure to
adequately access lubrication requirements can make a good structural design fail miserably from a
lubrication-related failure mode. Careful consideration of the ambient conditions, lubricants chosen,
gear pitch line velocities, film thickness, and flash temperature are necessary. The information provided
in this short overview is intended to provide an awareness of the potential catastrophic issues that lubri-
cation failures can cause.

References
AGMA ISO 14179-1, Gear reducers—Thermal capacity based on ISO/TR 14179-1. March 2004.
AGMA 912-A04, Mechanisms of gear tooth failure, American Gear Manufacturers Association,
Alexandria, VA October 2004.
Alban, L., Systematic Analysis of Gear Failures, 4th edn., American Society of Metals, Materials Park, OH,
July 1993.
Anderson, N. and Lowenthal, S., Spur Gear System Efficiency at part and Full Load, NASA TP-1622,
AVRADCOM TR-79-46, February, 1980.
ANSI/AGMA 1010-E95, Appearance of gear teeth—Terminology of wear and failure, American Gear
Manufacturers Association, Alexandria, VA April 2007.
ANSI/AGMA 9005-E02, Industrial gear lubrication, American Gear Manufacturers Association, Alexandria,
VA January 2008.
Bartz, W. J., Lubrication of Gearing, Mechanical Engineering Publications Ltd., London, U.K., 1993.
Coy, J., Townsend, D., and Zaretsky, E., Gearing, NASA RP 1152, AVSCOM TR-84-C-15, 1985.
Drago, R., Fundamentals of Gear Design, Butterworths Publishers, Oxford, U.K., 1988.
Dudley, D., Handbook of Practical Gear Design, McGraw-Hill Book Co., New York, 1984.
Hamrock, B., Lubrication of Bearings, NASA RP-1126, August 1984.
ISO 6336, Calculation of load capacity of spur and helical gears, International organization for standard-
ization, Geneva, Switzerland, 1996.
Townsend, D., Dudley’s Gear Handbook, 2nd edn., McGraw-Hill, Inc., New York, 1992.
51
Cams
Nomenclature.............................................................................................. 51-1
51.1 Introduction..................................................................................... 51-2
51.2 Unsteady Lubrication...................................................................... 51-2
51.3 Cams and Tappets........................................................................... 51-3
Andrew V. Olver 51.4 Discussion......................................................................................... 51-5
Imperial College London References..................................................................................................... 51-6

Nomenclature
a acceleration, a = du‾/dt
b0 hertz semiwidth
h film thickness
hmin minimum film thickness
p pressure
t time
D1 instantaneous cam dimension D2 − 2R (Figure 51.1)
D2 instantaneous cam lift (Figure 51.1)
E' reduced elastic modulus (~220 GPa for steel on steel)
F contact load
G1 dimensionless group (Equation 51.7)
G3 dimensionless group (Equation 51.8)
H distance from center of rotation to center of curvature (Figure 51.1)
H* dimensionless film thickness (Equation 51.6)
R radius of curvature of the cam at the point of contact
u1 velocity of cam surface in the tangent plane
u2 velocity of follower surface in the tangent plane
u 0 velocity of contact point in the tangent plane
u– entrainment velocity (Equation 51.9)
W load per unit cam width
X coordinate in the rolling direction
η viscosity
η viscosity at zero pressure
ω angular speed dθ/dt
θ angle of rotation

51-1
51-2 Design for Lubrication and Tribology

51.1  Introduction
In principle, the contact behavior of all nonconforming mechanical components can be analyzed in the
same fashion, as has been described for gears or rolling bearings. However, particular problems arise in
cases where the conditions are unsteady, that is, when the load, geometry, or speed changes rapidly with
time. This is notably the case for the contact between cams and tappets (cam followers) in automotive
valve gear but also occurs in many other types of machine element, including, for example, ball-type
CV joints (where there is reciprocating motion between ball and raceway) and to nonconjugate action
in gears (where there is a reversal of entrainment and rapidly changing geometry associated with the
contact of the tooth corner). Piston ring-liner contacts and some rolling bearings (e.g., helicopter and
wind turbine feathering bearings) are also subject to reciprocating action. The steady-state treatment
of lubrication is not necessarily valid under these conditions. In this section, we describe an alternative
approach based primarily on the work of Hooke [1–6] and give some examples of analysis of some cam-
tappet and other problems involving reciprocating action.

51.2  Unsteady Lubrication


For a line contact with time-varying conditions, the problem involves solving the one-dimensional
combined squeeze-entrainment Reynolds equation:

∂ ⎛ h3 ∂p ⎞ ∂h ∂h
⎜ ⎟=u + (51.1)
∂x ⎝ 12h ∂x ⎠ ∂x ∂t

The viscosity, η, varies with pressure, p, according to the Barus relation:

h = h0 eap (51.2)

The film thickness, h, is given by

x2
h = h0 + +d (51.3)
2R

where
h0 is the undeformed separation of the surfaces
R is the reduced radius
δ is the elastic deformation

Hooke [1] suggests that the film thickness is adequately predicted by the steady-state relationship
(∂h/∂t = 0) except in the immediate vicinity of a reversal of entrainment (u– = 0). Here, we may write

u = at (51.4)

where a is the acceleration at the time of reversal (when u– = 0) which may be assumed constant.
Hooke provides regression equations, representing the solution of Equations 51.1 through 51.4, for
the minimum film thickness hmin during the reversal.
As for steady-state elastohydrodynamic lubrication (EHL), separate equations are necessary to
describe the film thickness for each of the fluid film regimes: rigid-isoviscous (hydrodynamic), elastic-
isoviscous (“soft EHL”), rigid-piezoviscous, and elastic-piezoviscous (“hard EHL”).
Cams 51-3

For the elastic-piezoviscous case [2],

H * = 1.97 G3−0.14 G10.856 (51.5)


Here

2/3
hmin ⎛ W 2 ⎞
H* = (51.6)
R ⎜⎝ h02aR ⎟⎠

1/ 6
⎛ a 3W 5 ⎞
G1 = ⎜ 2 4 ⎟ (51.7)
⎝ h0aR ⎠

and

2/3
⎛ W7 ⎞
G3 = ⎜ 3 4 2 5 ⎟ (51.8)
⎝ E h0 a R ⎠

For a real component, we may use this system of equations to estimate the actual minimum film thick-
ness near the time at which the reversal of entrainment occurs. The film thickness depends on one
parameter not present in the steady state, namely, a, the acceleration. Equation 51.5 through 51.7 show
that the film thickness rises in proportion to a0.71. At all other times, the steady-state equations provide
the film thickness with sufficient accuracy, and the film thickness varies approximately as u– 0.7.
Several authors have provided solutions to the time-dependent elastohydrodynamic cam-tappet
problems [5–8] in point as well as line contact. Corresponding experimental measurements have been
performed by Glovnea and Spikes [9]. Results show that the minimum film thickness does not occur at
exactly the time at which the velocity is zero. Rather, the minimum film thickness occurs at a slightly
later time. The minimum film thickness at the contact center occurs when the contact has swept a dis-
tance equal to the semiwidth b0:

2b0
t min =
a

An example of this approach is now given for a cam-tappet contact.

51.3  Cams and Tappets


Determining the correct radius of curvature, entrainment velocity, and acceleration for the contact
between a cam and a follower is not always straightforward. However, the following simple example
serves to illustrate the method.
A simplified (constant radius, flat follower) cam and tappet is illustrated in Figure 51.1.
In general, the follower has a kinematically indeterminate tangential velocity (u2). In some cases, the
follower slides against the cam (u2 = 0—a long finger follower approximates to this), but in others, the
follower is completely free to displace laterally, and the velocity is approximately that of the cam, u2 = u1.
More commonly, the follower is in tractive sliding so that its velocity depends in part on the friction in
the lubricated cam-follower contact; under such circumstances, we might expect 0 < u2 < u1.
51-4 Design for Lubrication and Tribology

F O is centre of rotation
Q is centre of curvature
u2
H = OQ
u2 = tangential speed of follower
u1 ω (R–H cos ωt) + u2
u=
Q 2 2
w

D2
R
ωt H If u2 = 0 (sliding): u = –ωD1/2

O If u2 = ωD1 : u=0
D1 (D2 – D1)
If u2 = u1 (no slip) u=ω = ωR
2

FIGURE 51.1  Geometry and entrainment speed of a constant radius cam in contact with a plane follower. Note
that Q lies on the common normal to both surfaces at the point of contact.

The entrainment speed is, by definition, the average of the speeds at which each of the surfaces sweeps
into the contact:

(u1 − u0 ) + (u2 − u0 ) u1 + u2
u= = − u0 (51.9)
2 2

Here u0 is the time derivative of the horizontal position of the actual contact point:

d
u0 = (H sin w t ) = Hw cos w t (51.10)
dt

So

w u
u= (R − H cos w t ) + 2 (51.11)
2 2

where H is the distance of the center of curvature from the center of rotation. This may conveniently be
written in terms of the dimensions D1 and D2 (Figure 51.1) as follows:
For a slider follower, u2 = 0:

−w D1
u= (51.12)
2

For a free follower, u2 = u1:

D2 − D1
u =w = wR (51.13)
2

The minus sign in Equation 51.12 shows that the entrainment speed directed toward the left of
Figure 51.1 for a slider follower—and hence that the inlet is to the right of the contact. In contrast,
Equation 51.13 shows that the inlet is to the left for a free follower. Since the entrainment velocities
predicted for the two cases are opposite in direction, the entrainment speed is zero at an intermediate
value of u2. From Equation 51.11,
Cams 51-5

u = 0 if u2 = w D1 (51.14)

Any follower that operates under tractive sliding conditions (for which 0 < u2 < u1) is therefore at risk
of developing a very low or zero entrainment speed (and hence will have low film thickness and cor-
respondingly poor durability). Thus, bearing drag or excessive inertia in a free follower that may cause
it to lag behind its theoretical no-slip speed is potentially damaging. A more widely recognized [2]
condition leading to zero entrainment speed is the case of a slider follower where D1 = 0: or D 2 = 2R. In
this case, the acceleration, a, at the reversal point is given by differentiating Equation 51.11 with u2 = 0,
with the result

w2
a= H 2 − R2 (constant radius cam, plane slider follower) (51.15)
2

This quantity may be used in conjunction with Hooke’s analysis (e.g., Equations 51.5 through 51.8) to
determine the lubricant film thickness. For cams of variable nose radius, R and H vary with time, and
the acceleration may be written

⎛ dR 1 dD2 ⎞ (51.16)
a = −w 2 ⎜ −
⎝ dq 2 dq ⎟⎠

where θ = ωt is the cam angle. Also, since the local values of R and H must lead to a continuous curve,
D2 and R are related by [10]

d 2 D2
R= + D2 (plane follower) (51.17)
dq

Returning to the constant radius cam, we now take some example values: H = 40 mm, R = 15 mm,
D 2 = 30 mm, (D1 = 0), and ω = 1000 rev/min, leading (from Equations 51.12 and 51.15) to u– = 0 and
a = 200 m/s2 . For η 0 = 10 cP and E' = 220 GPa (steel/steel), we find from Equations 51.7 and 51.8 that
G1 = 260, G 3 = 25, and the minimum film thickness from Equations 51.5 and 51.6 is 25 nm. The sliding
speed (equal to u1 = ω D 2 since u2 = 0) is 3.2 m/s. On the other hand, a free (no slip) follower gives a
steady entrainment speed under the same conditions of ωR = 1.6 m/s and a film thickness of around
150 nm.
A particularly disadvantageous situation arises if H = R = D1/2 when both entrainment speed and
acceleration are identically zero across the cam nose. Although strictly outside the range of validity
of Hooke’s analysis, it is clear that a very low film thickness and poor durability is likely (Figure 51.2).

51.4  Discussion
Only the simplest type of cam-tappet contact has been considered here; nevertheless, the following find-
ings are of practical importance.
Cam-follower contacts frequently display reversals of entrainment direction, leading to minima in
the thickness of the lubricant film. The values of the minima may be estimated, for line contacts, using
the approach due to Hooke and involving the determination of the local acceleration, a.
In most practical cases, the follower has a kinematically indeterminate tangential velocity (u2).
Certain velocities of the follower are damaging because they lead to stationary and zero values of the
51-6 Design for Lubrication and Tribology

R
H=R

FIGURE 51.2  Cam and follower with an entrainment speed and acceleration of identically zero at the nose. This
design would be expected to show poor durability.

entrainment speed and thus a very low film thickness. These conditions can occur if a notionally free
follower, such as a roller, is slowed below its theoretical no-slip speed.
A design of cam in which the curvature of the nose radius passes through the cam center is also
disadvantageous (Figure 51.2) because an elastohydrodynamic film cannot be formed. Designs like this
should be avoided.
Finally, it has been observed [1] that “because the rate of change of entrainment velocity a is related
to the third derivative of lift, d3D2/dθ3, very small changes in the lift profile can significantly alter the
minimum film thickness at entrainment reversal.” This follows from Equation 51.17, but the observa-
tion needs treating with caution since large values of dR/dθ invalidate Equation 51.3 upon which the
geometrical analysis has been established. The analysis will be incorrect if there are significant changes
of curvature within the region comprising the contact and the adjacent pressurized inlet. In contrast, it
seems likely that similar cam profiles will give similar behavior.

References
1. Hooke, C J, The minimum film thickness in lubricated line contacts during a reversal of entrain-
ment-general solution and the development of a design chart, Proc. Inst. Mech. Eng. Part J: J. Eng.
Tribol., 208 (1994) 53–64.
2. Hooke, C J, The minimum film thickness in line contacts during reversal of entrainment, J. Tribol.,
115 (1993) 191–199.
3. Hooke, C J, The minimum film thickness in lubricated line contacts during a reversal of
­entrainment-isoviscous behaviour, Proc. Inst. Mech. Eng. Part J: J. Eng. Tribol., 206 (1992) 337–345.
4. Hooke, C J, The minimum film thickness in lubricated line contacts during a reversal of entrain-
ment-piezoviscous behaviour, Proc. Inst. Mech. Eng. Part J: J. Eng. Tribol., 206 (1992) 417–424.
5. Dowson, D, Taylor, C M, and Zhu, G, A transient elastohydrodynamic lubrication analysis of a cam
and follower, J. Phys. D: Appl. Phys., 25 (1992) A313–A320.
6. Messé, S and Lubrecht, A A, Transient elastohydrodynamic analysis of an overhead cam/tappet
contact, Proc. Inst. Mech Eng. Part J: J. Eng. Tribol., 214 (2000) 415–425.
7. Bell, J C, Reproducing the kinematic conditions for automotive valve train wear in a laboratory test
machine, Proc. Inst. Mech. Eng. Part J: J. Eng. Tribol., 210 (1996) 135–144.
8. Kushwaha, M and Rahnejat, H, Transient elastohydrodynamic lubrication of finite line conjunction
of cam to follower concentrated contact, J. Phys. D: Appl. Phys., 35 (2002) 2872–2890.
Cams 51-7

9. Glovnea, R P and Spikes, H A, The influence of cam-follower motion on elastohydrodynamic film


thickness, Tribology Research: From Model Experiment to Industrial Problem—A Century of
Efforts in Mechanics, Materials Science and Physico-Chemistry, Proceedings of the 27th Leeds-Lyon
Symposium on Tribology, Tribology and Interface Engineering Series, Vol. 39, 2001, pp. 485–493.
10. Stone, R and Leonard, H J, Determination of the instantaneous radius of curvature of a cam at the
contact point with a flat follower moving orthogonally, Proc. Inst. Mech. Eng. Part J: J. Eng. Tribol.,
208 (1994) 147–149.
52
Lubrication Oil Systems
Jan Ploszaj
Howard Marten
Company Ltd.
52.1 Noncirculation Oil System............................................................. 52-1
Hooshang Heshmat Oil Bath Applications  •  Oil Ring Lubrication  •  Lubricant Supply by
Mohawk Innovative
Fixed Disks
Technology, Inc. 52.2 Circulating Oil System.................................................................... 52-7
Introduction  •  Sizing the Lubrication System  •  Main System
George J.W. Components  •  Typical Upgrades  •  Typical Oil Lubrication System
Staniewski Process and Instrumentation Diagram and Layout  •  Troubleshooting
Ontario Power Generation References...................................................................................................52-22

52.1  Noncirculation Oil System


52.1.1  Oil Bath Applications
For small rotating machines, a local oil bath lubrication system is frequently selected due to its simple
design, relatively good reliability, and low operation cost. Typical examples of oil bath lubrication are
horizontal rotating machines such as electric motors, centrifugal overhang pumps, or small gearboxes.
The contact surfaces are lubricated by oil mist formed by the continuous action of splashing oil caused
by rotating parts in the bearing or gearbox housings.
One of the critical design considerations is a proper balance between oil mist generation to provide
adequate supply of oil to the lubricated contact zone and minimization of friction losses by rotating
parts moving through oil. If the oil level is lower than the optimum, the amount of oil being splashed is
reduced, causing a potential for lubricant starvation in the lubrication zone. As a result, contact at asperi-
ties may occur and the contact zone temperature may rise. The temperature increase would cause reduc-
tion in the oil viscosity, which in turn will further reduce oil film thickness, as well as cause thermal
expansion of the bearing components. All of these are precursors of catastrophic bearing failure. If the
oil level is higher than the optimum, oil will create resistance for rotating element movement, and will
churn and aerate, producing air bubbles that can trap heat. These will increase power losses and increase
lube oil temperature, accelerating the oil oxidation process due to elevated temperature and excessive air
presence. In some applications, a high oil level may also promote leakage through the seals.
Equipment manufacturers recommend the preferred oil level, which usually depends on the equip-
ment rotation speed, shaft size, and applied load. For typical horizontal shaft applications with ball
bearings, the preferable oil level in the stationary condition (not running) is at the middle of the ball
located at the lowest position in the rolling element bearing.
Maintaining the oil level within a proper range is important to extend equipment reliability and
minimize drag forces. Equipment manufacturers may provide a sight glass or window to visually set and
monitor the oil level in bearing housings. Constant level oilers were introduced, allowing continuous
replenishment of oil lost by leakage. Shaft seal improvements may minimize problems associated with
leakage and contamination.

52-1
52-2 Design for Lubrication and Tribology

Direction of rotation

FIGURE 52.1  Vented constant level oiler with side connection. (Courtesy of Trico Corporation, Pewaukee, WI.)

The majority of existing constant level oilers allow for level adjustment, facilitating installation.
Typically, they are mounted on the side of equipment facing the direction of shaft rotation; see Figure 52.1.
This can protect against oil sump surges during start-up operation and overfeed of the bearing housing.
Some models apply a bottom mount configuration to minimize this occurrence.
For splash lubrication in a horizontal shaft application with rolling element bearings, it is a com-
mon practice to allow the oil to freely flow into each side of the rolling element bearing. Most bear-
ing housings should have a bypass opening beneath the bearing to allow the oil to return freely
to the sump.
A typical example of a lubrication method for vertical shaft applications with a plain bearing is shown
in Figure 52.2. Thrust bearing parts are submerged in oil while the guide bearing is only partially cov-
ered. Shaft sealing in these applications is usually provided by a thin cylindrical sleeve inside the bear-
ing inner ring support.
A common problem in these applications is the potential for oil leakage over the cylindrical sleeve
into the rotor winding area due to a siphon effect if oil exceeds the maximum allowable level. In some
applications, additional vent holes are applied in the guide bearing to equalize pressures between areas

FIGURE 52.2  Simplified typical shaft seal arrangement with plain bearings for vertical applications.
Lubrication Oil Systems 52-3

FIGURE 52.3  Simplified typical shaft seal arrangement with rolling element bearings for vertical applications.

above the cylindrical tube and bearing housing to minimize this siphoning effect. In all cases, oil must
be added with caution and never exceed the maximum allowable level.
In vertical shaft applications with rolling element bearings (particularly spherical roller or angular
contact bearings), oil bath lubrication may produce a high drag force if the oil flow to the bearing com-
partment is not controlled. Typically, a special metering device or orifice is provided on the internal
bearing housing to regulate the oil flow rate and prevent power losses. Oil level is usually set at the center
of the lower bearing. See Figure 52.3.

52.1.2  Oil Ring Lubrication


Oil ring bearings have been in wide use for over a century in electric motors, pumps, fans, and other
large machines. The calculation procedure for oil ring lubrication is presented in the Tribology Data
Handbook [1].
Typical design characteristics are the following:
• Bearing length/bearing diameter ratio is between 0.6 and 1.2.
• Oil ring ID/bearing diameter ratio is from 1.5 to 1.9.
• Bearing bore/shaft journal clearance ratio is from 0.05 to 0.07 mm (0.0020 to 0.00275 in.).
There are two different applications of oil rings. In the first case, the oil ring is a part of the journal bear-
ing and the shaft is not modified (see Figure 52.4). The axial travel restriction of the ring is generated
within the upper bearing housing. The design function of the oil ring is to deliver the lubricating oil
directly to the inlet of the journal-bearing packet.
In the second case, the ring is located in a special groove on the shaft located between bearings
(see Figure 52.5). The design function of the ring is to generate oil mist, which is the main lubrication
medium for the rolling element bearings.
Experimental studies suggest that oil ring bearings operate under starved hydrodynamic conditions.
For that reason, the oil delivery rate by the ring is of primary concern. According to Heshmat, there are
four characteristic regimes of oil ring behavior. The regimes are shown in Figure 52.6 [2].
Regime I: At the low end of journal rotation, there is contact between the ring’s inner surface and the
journal, and the linear speeds of mating surfaces are about the same. As a result, there is a linear rise in
ring rotation (NR) with the journal. At the upper end of Regime I, ring speed (NR) and oil delivery (QR)
reach a local maximum.
52-4 Design for Lubrication and Tribology

Oil ring Oil ring

Thrust end Drive end

FIGURE 52.4  Plain bearing lubricated by oil ring. (Courtesy of Sulzer Pumps Inc., Burnaby, BC.)

Oil ring

FIGURE 52.5  Rolling element bearings lubricated with oil rings. (Courtesy of Sulzer Pumps Inc., Burnaby, BC.)

Regime II: At the beginning of this regime, direct frictional drag yields to a state of boundary lubrica-
tion between the ring and the journal. As a result, slippage occurs and ring speed (NR) drops. Since the
speed has decreased, so does the amount of oil delivery (QR) by the ring. However, with a further rise in
journal speed, a full hydrodynamic film is established between the journal and the ring. This causes a
rapid increase in both ring speed (NR) and oil delivery (QR). Once again, at the upper end, a local maxi-
mum in ring speed is achieved. Oil flow, however, at the end of this regime is an absolute maximum and
represents the highest possible oil delivery by the ring.
Lubrication Oil Systems 52-5

NR

Regime Regime Regime Regime


I II III IV
QR

σz

σθ

Journal speed

FIGURE 52.6  Oil ring behavior as a function of journal speed.

Regime III: The drop in ring speed and oil delivery is associated with the onset of ring oscillation in
the plane of rotation. While small oscillations already appear during the trailing portion of Regime II, in
Regime III, the values (б𝜃) become very large within a short span. This causes a drastic reduction in oil deliv-
ery with only a slight reduction of ring speed. At the end of this regime, the oil ring delivery approaches zero.
Regime IV: This regime is characterized by further instability in ring behavior. Although the oscilla-
tory vibration is reduced, the conical and axial vibrations of the ring are significantly increased. While
the ring speed once again tends to increase with journal speed, oil delivery is essentially negligible. A
typical pattern of transporting oil by the ring in Regime I and II is shown in Figure 52.7.

Hot oil layer


Cold oil
ω layer
0+

0R
Bearing
Hydrodynamic φ
R
oil film
Oil level at ωR = 0

Oil tail Oil surface level Oil tail


A ωR A
Oil sheet Oil sheet
Sump

View A–A

FIGURE 52.7  Pattern of oil transport by oil ring in Regime I and II.
52-6 Design for Lubrication and Tribology

Typical journal speed limits for such lubrication are 10–14 m/s. Journal surface finish should be less
than about 1/10 to 1/3 of the calculated minimum oil film thickness. Self-lubricated journal bearings
need axial grooves in the bore to distribute oil from the feeding device.
Typical oil ring immersion is between 5% and 25% of the oil ring bore diameter, with the most com-
mon immersion being 15%.
A typical ring material is bronze SAE-660. Usually, bronze rings deliver approximately 10% more
oil than similar rings made from brass or Muntz metal. The increase in ring unit weight (ring weight/
circumferential length), up to its optimum of approximately 24 N/m, shows a substantial increase in its
oil delivery. Further increase in ring unit weight decreases oil delivery drastically throughout the speed
range. The profile of the ring is one of the most important factors in ring performance.
For applications shown on Figure 52.4, one of the important elements of the ring profile is its trap-
ezoid configuration. Because of high side drag, either side of the ring can stick to a wall of the bearing
slot, causing immediate reduction in ring speed. To reduce the side drag, the side length of the ring
(perpendicular to the shaft) should be small (1–2 mm for larger rings) and the rest is usually cut at an
angle of 30°.

52.1.3  Lubricant Supply by Fixed Disks


Disks usually require more axial space than oil rings but offer the following advantages:
• They operate better on tilted shafts.
• They are a more positive supply method at very low speeds.
• They can be arranged to deliver oil at high speeds and under substantial pressure.
An example of a disk-oiled bearing is shown in Figure 52.8.

FIGURE 52.8  Typical disk-oiled bearing lubrication arrangement. (Courtesy of Sulzer Pumps Inc., Burnaby, BC.)
Lubrication Oil Systems 52-7

52.2  Circulating Oil System


52.2.1  Introduction
In some industrial applications, a large oil circulation system continuously supplies lubricating oil or
hydraulic fluid to various system components to perform different design functions. Such systems can
be very reliable and cost efficient.
There are specific requirements for different applications, which must be incorporated into the sys-
tem design configuration. Practically, lubrication means a separation of the moving surfaces and, thus,
reduction of friction and associated heat generation. This will reduce formation of wear particles and
improve machine efficiency. In addition, surface separation could reduce surface fatigue, operating
noise, and vibration.
Carrying away of external contaminants and internally generated debris is a design function of the
circulation oil system. This is primarily achieved by application of particulate filtration and purification
units and partially through the settling on the bottom in the oil reservoir. Solid particulates are of con-
cern due to their abrasive action on critical system components and their negative impact on component
performance (i.e., potential loss of hydrostatic pressure on the scored thrust pads or journal bearings).
Moisture may cause rusting of the lubricated metal surfaces or initiate the hydrolysis process of some
fluids (i.e., phosphate esters) or oil additives.
Another vital design function of the circulation oil system is removal of unwanted heat from the
system’s various lubricated components to maintain a stable lubrication condition, preventing surface
deterioration and eliminating thermal stress on the oil. In practice, the stable lubrication condition is
achieved by maintaining the oil viscosity in a narrow range and, thus, minimizing change in oil film
thickness and its load carrying capacity. Excessive heat could change the metallurgy of the surface layer
and promote oxidation or thermal breakdown of the lubricant chemical structure. Accurate estima-
tion of heat transfer in large lubricating systems allows design engineers to determine the heat removal
requirement and select an appropriate oil flow rate. High flow rates usually require larger reservoirs and
bigger coolers to maintain proper heat transfer. This may cause a potential problem during shutdown,
particularly if oil circulation would be terminated too fast. In such circumstances, oil would be imme-
diately oxidized or thermally broken down, generating degraded compounds, which can precipitate and
potentially block the oil lines.
It should be noted that in many applications, users initially warm up the oil and the lubricated system
components during initial start-up operations by recirculating the oil in a small circuit or using res-
ervoir mounted heaters, bypassing coolers. Some systems use external heaters to raise oil temperature
before it will be circulated through the system.
From a safety and performance perspective, adequate control of oil aeration as well as foaming or
vapor concentration in different portions of the circulating oil system is an important design consider-
ation. Local environmental conditions and safety requirements may significantly influence the design
of specific components of the circulating oil systems.

52.2.2  Sizing the Lubrication System


The main factor dictating the flow and, as a result, the size of oil lubrication systems is the amount of
heat generated by the machine. As a general rule, the temperature difference between oil entering the
machine and oil leaving the machine should be between 6°C and 20°C (10°F and 36°F). Depending
on the type/grade of oil and machine requirements, the oil supply temperature should be between
40°C and 60°C (104°F and 140°F). Most equipment manufacturers provide sizing calculations which
include recommendations concerning oil flow and oil grade. If such data are not available (e.g.,
on existing old machinery), heat generated by the machine can be determined from the following
formula:
52-8 Design for Lubrication and Tribology

H = Q ∗ 7.5 ∗ 0.5 ∗ ΔT ∗ 60

where
H is the heat load in Btu/h
Q is the oil flow in GPM
ΔT is the temperature difference between oil exiting and entering the machine in °F

or by

H
HL =
3412

where HL is the heat load in kW


After heat load is calculated, proper temperature can be achieved by increasing or lowering oil flow.
Based on required flow, all other components of the lubrication system can be sized.
The size of the reservoir has to be large enough to allow time for
1. Gases entrapped in the oil to escape
2. Particles in oil to settle down
Depending on the amount of gases and particles, oil grade, and oil temperature, the size of the reservoir
can vary greatly. In heavy oil applications (ISO VG220 or VG320), reservoir working volume is typi-
cally 20–40 times the required oil flow (e.g., if the required oil flow to the machine is 38 L/m [10 GPM],
reservoir size would be between 750 and 1500 L [200 and 400 USGAL]). In light oil applications such as
ISO VG32, the working volume can be as low at 3–5 times the required flow. In the majority of typical
applications with oils between ISO VG32 and ISO VG100, the reservoir volume should be between 6 and
10 times the required oil flow.
There are many other factors that either increase the required reservoir size (such as water ingress) or
help to reduce its size (deaerating trays, baffles, etc.).
Note: Reservoir working volume is defined as the amount of oil between minimum level and suction
loss level.
The next step in sizing a lubrication oil system is to select proper components. The following are gen-
eral rules based on industry standards for selecting major equipment:

1. Pumps: typically, for positive displacement pumps (gear, screw, etc.), flow should be approxi-
mately 120% ± 5% of required oil flow (e.g., required oil flow is 38 L/m [10 GPM], pump flow should
be approximately 45 ± 2.3 L/m [12 GPM ± 0.6 GPM]). Centrifugal pumps should be sized for exact
flow required without any margin due to a different principle of flow regulation.
2. Filters: filter sizing is typically based on manufacturers’ tables or software. Filter pressure drop for
a new element should be below 0.35 bar (5 psi) (at operating temperature), and the element should
be replaced when pressure drop reaches 1 bar (15 psi). In some cases where design of the system
permits, elements can be changed at higher differential pressure.
3. Coolers: standard (catalogue items) cooler sizing is based on manufacturers’ tables or software. It
is important to limit pressure drops on both oil and water sides (typically between 0.7 and 1 bar
[10 and 15 psi]). In many cases, to increase the interval between maintenance and cleaning, it is
also desirable to include excess cooler surface (for shell and tube coolers, it is called the fouling
factor).
4. Heaters: heaters are sized to warm up the entire reservoir oil volume from minimum ambient
temperature to minimum allowable start temperature in a certain time. In most cases, minimum
start-up temperature is between 10°C and 21°C (50°F and 70°F). Typical heating time is 12 h;
Lubrication Oil Systems 52-9

however, in some cases, it may be reduced to allow for shorter downtime. To prevent oil burning,
low-density heaters (below 2 W/cm2) are preferable.
5. Piping: Table 52.1 shows recommended line sizes in Imperial units based on oil flow and line type
(pressure, suction, return). Line sizes are determined based on permissible oil velocity in the pip-
ing. Data in the table follow accepted industry standards of
a. 3 m/s (10 ft/s) for pressurized lines
b. 1.2 m/s (4 ft/s) for pump suction lines
c. 0.6 m/s (2 ft/s) for pressure return lines (from relief valves, control valves, vents, etc.)
d. 0.3 m/s (1 ft/s) for gravity return lines
6. If tubes or hoses are used for oil lines, their sizing should also be based on the velocity as listed
earlier. The following formula can be used to determine velocity in the oil lines in SI system:

V = Q*21.237/ID2

where
V is the velocity in line (m/s)
Q is the flow (L/m)
ID is the line inside diameter (mm)

Similarly, in Imperial units,

V = Q * 0.4085/ID2

where
V is the velocity in line (ft/s)
Q is the flow (GPM)
ID is the line inside diameter (in.)

7. Manual valve sizes should follow pipe sizing.


8. Control valves should be sized per manufacturer recommendations.
9. If instruments are mounted on a gauge panel or instrument stands and sensing lines are run using
tubing, it is recommended to use 12 mm (1/2 in.) OD or 10 mm (3/8 in.) OD tube sizes. In some
cases, for distances shorter than 3 m or higher pressures, 6 mm (1/4 in.) OD tube can be used as well.

52.2.3  Main System Components


The main components and basic construction of circulating oil systems have changed little in 30 years.
There was, however, a significant improvement in oil filtration and instrumentation. Also, in modern
hydraulic systems, design engineers more frequently apply variable speed pumps, which operate at opti-
mum energy consumption, providing the exact amount of oil required by the system. This reduces
energy consumption and improve oil life as fluids are less stressed.

52.2.3.1  Traditional Reservoir Design


Typical industrial oil reservoirs have a rectangular shape with length approximately twice the width.
The height usually is equal to the width, and the bottom plate has a slope of 1:15–1:30 for appropriate
drainage of water and impurities. Some reservoirs may have circular bottoms for better moisture sepa-
ration and more efficient fluid circulation. From a maintenance perspective, it is desirable to have the
entire inside of the reservoir accessible for inspection or cleaning.
52-10

TABLE 52.1  Pipe Sizing Table for Carbon and Stainless Steel Piping Material (Flow Rate in USGPM)
NPS (in.) 1/2 3/4 1 1 1/4 1 1/2 2 2 1/2 3 4 6 8 10 12 14
1. Gravity return lines at 2 ft/s (used for external return lines based on one-half full pipe)
Pipe schedule 10 1 2 3 5 7 12 17 26 44 98 168 264 373 451
40 1 2 3 5 7 10 15 23 40 90 155 242 346 427
80 1 2 3 4 6 10 13 21 36 81 142 231 336 411
2. Gravity return lines at 2 ft/s (used for internal return lines based on full pipe)
Pipe schedule 10 2 4 6 10 14 23 33 52 88 197 337 528 746 905
40 2 4 6 10 13 21 30 46 79 179 310 483 693 854
80 2 3 5 8 11 19 26 41 71 162 283 462 671 822
3. Suction lines at 4 ft/s
Pipe schedule 10 4 8 12 20 28 45 67 103 176 393 674 1056 1492 1804
40 4 7 11 19 25 42 59 92 158 358 619 976 1385 1707
80 3 5 9 16 22 37 52 82 142 323 565 924 1342 1643
4. Pressurized lines at 10 ft/s
Pipe schedule 10 11 19 29 51 69 113 169 258 441 982 1686 2639 3731 4511
40 9 17 27 46 63 104 148 229 394 894 1548 2440 3464 4267
80 7 13 22 40 55 91 131 204 356 807 1416 2310 3355 4107
Return Lines and Common Headers Sizing Table Notes
5. Pressurized, nonpressurized, and gravity return lines
Individual pressurized returns, except (PRV) outlet Pressurized line (4) Sized for max. flow rate
Individual nonpressurized returns outlet Gravity return line (1) Sized for max. flow rate, one-half full
Individual PRV outlet Pressurized line (4) But no smaller that PRV outlet
Common header for two relief valves, on a two-pump Pressurized line (4) Header sized for a single pump flow rate, but
system, one running and one standby no smaller then PRV outlet
Common header for more than two relief valves, on a Gravity return line (2) Header sized for a single pump flow rate, but
system where more than one pump running and at no smaller that PRV outlet
least one standby
Common header for pressurized and nonpressurized Gravity return line (2) Header sized for total flow rate
returns and vents
Values in tables are for oil and water applications
Design for Lubrication and Tribology
Lubrication Oil Systems 52-11

Reservoir structure is usually reinforced with stiffeners on the outside and baffle plates inside,
which divide the internal space to increase oil retention time and promote air bubble foaming and
moisture precipitation. The typical material of construction is carbon steel usually painted with spe-
cial epoxy paint. In some applications, stainless steel is used particularly if water concentration or
fluid cleanliness is of concern (i.e., steam turbine control systems with phosphate ester fire-resistant
fluids).
Main oil return pipes should be located at the end of the reservoir, opposite to the pump suction
location. In small reservoirs, return pipes are usually submerged in the oil to minimize fluid splash and
foam formation. In large reservoirs, however, return pipes are partially above the oil level to promote
the air release process upstream of oil reservoir and minimize foaming. Usually, the slope of the return
pipes is very small to minimize turbulence. At the entrance to the oil reservoir, there are special screen-
ing baskets for collecting large solid impurities originating from the lubrication system or maintenance
activities. Any lines returning from pressure relief valves (PRVs) and system purification should be
submerged in the oil to minimize splashing.
To minimize moisture ingress, some reservoirs are separated from the surrounding environment
by desiccant-type air breathers. Other reservoirs are purged with dried air or neutral gases to control
moisture in the oil. There are also applications with air extraction fans connected to the highest point
of the reservoir for removing oil vapor and maintaining subatmospheric pressure inside the tank. To
support the air bubble removal process from the oil, reduce foam, and control oil vapor concentration,
internal baffles should have openings above oil level to equalize the subatmospheric pressure within the
reservoir. However, such systems require oil separators to reduce the volume of oil droplets and vapor
carried out to the environment.
Reservoirs must also have level gauges and inspection ports as well as a low point drain, usually con-
nected to the additional bypass purification system, and access plates for periodic cleanout. The supply
to the purification system should prevent siphoning of the reservoir below a safe level.
Pump suction is usually located below the lowest oil level, and efforts must be made to minimize
formation of air-entraining vortices. In some applications, pump suction is located slightly above the
bottom of the reservoir to avoid pickup of water and dirt.
No copper, cadmium, zinc, and lead should be used in circulating oil systems due to poor resistance
to corrosion and to minimize their catalytic effect on the oil oxidation process. In addition, all other
materials such as gaskets, seals, diaphragms, hoses, and surface coatings should be resistant to the cir-
culating fluid type and maintain adequate fluid properties at the maximum operating conditions during
extended services. If journal bearings are used in such systems, then bearing linings made of tin base
Babbitt should be considered.

52.2.3.2  Pumps and Drivers


Pumps are very important components of the oil circulating systems, and therefore, their reliability and
efficiency are critical. For most industrial circulating oil systems, gear pumps are a cost-effective solu-
tion. They are capable of generating over 69 bar (1000 psi) and a flow rate of up to 950 L/m (250 GPM); see
Table 52.2. Screw-type pumps are selected for critical applications, and they can deliver up to 3800 L/m
(1000 GPM) flow at up to 138 bar (2000 psi) discharge pressure. Centrifugal pumps are usually selected
for applications where large-volume flow is required at moderate pressures up to 12 bar (175 psi) and for
applications with oil viscosity below 250 cSt.
Most pumps are driven by electric motors. In some applications, however, the pump shaft can be
driven directly by the low-speed side of the gearbox, or in the case of steam turbines, the main pump
can be driven by the turbine’s main shaft. It is a common practice to have an auxiliary pump to permit
continuous operation if the main pump fails. There are some applications which require an emergency
pump of reduced capacity just to support safe start-up and rundown operation to prevent bearing dam-
age. To avoid conditions where single incident or equipment failure can cause loss of pumping, the rec-
ommended practice is that all pumps should be driven from independent and different power sources.
52-12 Design for Lubrication and Tribology

TABLE 52.2  Oil Circulating Systems


Oil Oil Reservoir
Viscosity at Feed Retention Rating
Application Duty 40°C (cSt) (L/m) Pump Type (min) Filter Type (μm)
Electrical Bearings 32–62 7.5 Gear 5 Strainer 50
machinery
General Bearings 68 40 Gear 8 Dual cartridge 100
Paper mill Bearings, gears 150–220 80 Screw or 40 Dual cartridge 120
dryer section gear
Steel mill Bearings, gears 150–460 640 Screw or 30 Dual cartridge 150
gear
68–680 900 Screw or 20 Dual cartridge
gear
Heavy duty Bearings, 32 2250 Screw or 5 Pleated paper 5
gas turbines controls centrifugal
Steam turbine Bearings, seals, 32–46 3800 Centrifugal 5 Bypass 15% 10
generator turning gears,
some control

Also, pump motor overloads should not trip the motor breaker but just provide a warning signal. It is
also important that the auxiliary and emergency pump drivers should be sized for allowing cold start-
up operation.
Some pumps may need an emergency power supply to ensure their availability for duty. Other
industrial applications require that the heavily loaded shaft is lifted prior to rotation. To generate ini-
tial hydrostatic pressure in the main journal bearings, small additional positive displacement pumps
are used.
An important aspect of pump selection is adequate determination of pump suction conditions.
Usually, pump suctions should be below the minimum operating level in the reservoir. The exact level
may be determined by specific pump type and viscosity grade of pumped fluid in the worst operating
condition. Typically, the level is at least 15 cm (6 in.) below the operating fluid level in the reservoir.
Pump suction should not be lower than 50 mm (2 in.) above reservoir floor to minimize ingression of
contaminants.
Electric motors used to power pumps usually have three phases, 4-pole TEFC with case B insula-
tion. For emergency, pump’s motors operated typically on 125/250 V.D.C. and are supplied with TEFC
construction.

52.2.3.3  Filters
Filters remove contaminants from circulating fluid. Depending on the machine that the lubricating
or hydraulic system is servicing, different levels of filtration are required. Some components such as
precise control or servo valves may require very fine filtration (i.e., filters with beta ratio β 3(c) ≥ 1000
or better—filter capable of removing 3 μm and greater particles with efficiency of at least 99.9%),
while other machines may tolerate larger particles. Filtration level recommendations should always
be provided by machine manufacturers to ensure warranty terms are maintained and equipment
functions for a long time. The pressure drop with the start-up viscosity should be within the system’s
capabilities.
In general, full flow filtration is recommended to ensure that the circulating oil is filtered on each pass
through the system components. The online full flow filters should be upstream of critical components
to prevent damage. In some applications, additional filters may also be provided to protect specific criti-
cal components, such as servo valves. The bypass filtration arrangement (i.e., kidney type) may also be
Lubrication Oil Systems 52-13

FIGURE 52.9  Typical duplex filter arrangement. (Courtesy of HYDAC Filtertechnik GmbH, Niederkrüchten,
Germany.)

used particularly in applications where there is a need for high fluid cleanliness before energizing the
main or auxiliary pump.
There are many types of filters and filter media. The simplest filters commonly used on lubrication
units are canister-style spin-on filters. These are found on smaller lubrication units not exceeding
114 L/m (30 USGPM) and almost all mobile lubrication systems due to their simplicity of maintenance
and very low cost of spare parts. Filters are available in most common media (pleated paper, fiberglass,
water absorbing) and a wide range of efficiencies (micron ratings).
Where larger flows (above 114 L/m [30 USGPM]) are involved or duplex filters are required, typically,
cartridge type filters are employed (see Figure 52.9). Cartridge filters can be provided as a single vessel
or assembly of duplex vessel with a switch-over valve.
For any type of filter, wherever possible, it is recommended to install (as a minimum) drain valves (on
the “dirty” side of the filter) to allow for oil to be drained from the housing before element replacement.
It is also necessary to install vent valves that allow for air to be removed from the filter housing after
element replacement.

52.2.3.4  Coolers
The design function of coolers is to maintain the circulating oil within a specified temperature range
under all operating conditions. Usually, a cooler is a device for efficient heat transfer from one medium
to another. These media are separated by a solid wall, so that they never mix. However, the coolant pres-
sure should always be less than the lube oil pressure to avoid oil contamination with cooling fluid. The
location of the coolers should allow for maintenance and replacement. In general, all coolers should
have an adequate vent arrangement to permit air removal.
Typically, in circulating oil lubrication applications, two types of coolers are used: oil to water and
oil to air. In the majority of cases, oil to water coolers are provided. There are two basic designs of oil to
water coolers:
• Shell and tube
• Plate type
The source of cooling water depends on local water availability. If lake or river water is used, the tempera-
ture of that water would most likely change with the season. In such cases, the size of the cooler should be
selected with caution to avoid overcooling during winter time, particularly if the cooler is located upstream
of the online filters. Rapid cooling may cause significant differential pressure increases due to lower oil vis-
cosity and occasionally cause operational problems if filter pressure is associated with system logic.
52-14 Design for Lubrication and Tribology

Straight-tube heat exchanger Shell-side


fluid in
(one pass tube-side)
Tube bundle with Tube sheet
Tube sheet
straight tubes

Shell

Shell-side
Tube-side fluid out Tube-side
fluid in fluid out

FIGURE 52.10  Typical shell and tube heat exchanger. (Courtesy of ITT Standard, Buffalo, NY.)

52.2.3.4.1 Shell and Tube Heat Exchangers


Shell and tube heat exchangers consist of a series of tubes inserted inside a shell. In typical applications,
coolant runs inside the tubes and oil runs inside the shell. This allows for easy maintenance of the tube
interior. Cooling water, especially when not filtered or purified, can carry substantial amounts of foreign
material that eventually settles in the cooling tubes. The tube interior can be cleaned using specially
designed brushes.
Oil runs over the tubes that are being cooled. Baffles inside the shell redirect oil and increase effi-
ciency of the coolers (as shown in Figure 52.10). A set of tubes is called the tube bundle. There are many
different types of tubes used in the cooler that change its performance properties: plain, longitudinally
finned, etc. Shell and tube heat exchangers are used in any size of oil lubrication systems. They are a
very economical choice for smaller systems and are the preferred selection for more severe applications
because of their robust build.

52.2.3.4.2 Plate Heat Exchangers


Plate heat type of exchangers (see Figure 52.11) are composed of multiple, thin, slightly separated
plates that have very large surface areas and fluid flow passages for heat transfer. This stacked-plate

FIGURE 52.11  Plate-type cooler principle. (Courtesy of ITT Standard, Buffalo, NY.)
Lubrication Oil Systems 52-15

arrangement can be more compact than the shell and tube heat exchanger. Advances in gasket and
brazing technology have made the plate-type heat exchanger increasingly practical and popular. There
are two types of plate coolers:
• Gasketed type called plate and frame coolers
• Brazed or welded coolers
Plate and frame coolers allow for periodic disassembly, cleaning, and inspection. They are commonly
used in power plants, steel plants, and water treatment plants.
The second type of plate coolers has permanently bonded plates assembled by brazing (brazed pack
coolers) or welding (welded pack coolers). Brazed plate coolers are typically used on small lube oil sys-
tems. However, they can be manufactured for large flows as well. Welded-type coolers are used either at
very high temperatures where gasketed coolers cannot be applied or where cooling media is aggressive
to the gaskets. Both brazed and welded coolers are not repairable (throw away) but may provide for a
cost-effective solution due to their high efficiency and low cost.

52.2.3.5  Control Valves


Other important parts of any oil lubrication system are control valves. There are different valves used in
circulating oil systems:

• PRVs
• Pressure control valves (back pressure or pressure reducing)
• Temperature control valves

Some of the valves govern the safety and operation of the entire lube oil console. Therefore, proper selec-
tion of these valves is very important for system functionality.

52.2.3.5.1 Pressure Relief Valves


PRVs are installed in various places in the lube oil system in order to protect components and piping
from overpressure. Overpressure can be caused by closed valves or thermal expansion.
The first set of relief valves is found in the discharge of the positive displacement pumps before
any other valve in the system. Typical positive displacement pumps can produce pressure greater than
design pressure of some of the system components. The relief valves are installed to protect piping and
components from damage or rupture which may cause injury and requirement for repairs. Relief valves
are typically set at minimum 1.7 bar (25 psi) above the normal pump operating pressure and below
design pressure of any component in the system. They relieve oil directly back to the oil reservoir. Pump
relief valves are normally sized to allow for maximum 10% pressure accumulation (pressure cannot
rise more than 10% from when the valve starts opening to when it is fully open). After opening, a typi-
cal PRV will close at 80% of the set pressure. Some types of valves called sliding spool valves will close
instantaneously.
Relief valves are also commonly installed on

• Filter vessels (to protect vessels from thermal expansion)


• Cooling vessels (to protect vessels from thermal expansion)

Similar types of relief valves are used in all three applications (sliding spool valves cannot be used for
thermal expansion purposes as they leak minimum amounts of fluid continuously).

52.2.3.5.2 Back Pressure Control Valves


Back pressure control valves are used together with the positive displacement pumps. Their func-
tion is to redirect the excess oil produced by the pump back to the reservoir and maintain oil
52-16 Design for Lubrication and Tribology

pressure at a preset value. These valves are mounted on the pump header or in smaller systems on
the discharge line and are set at the pressure required at system outlet. In some cases, it is desir-
able to maintain oil pressure above cooling water pressure to prevent any coolant from entering
the oil in case of cooler tube or plate damage. In such an application, the back pressure control
valve is set at a pressure approximately 5% higher than maximum cooling water pressure and
another valve (pressure-reducing valve) is used to maintain the oil supply pressure value. It is very
important to properly size pressure control valves. During normal operation, the valve will bypass
excess oil (difference between pump capacity and system oil consumption) to the reservoir. In dual
pump applications, where both pumps could run simultaneously for short period (pumps switch
over), the back pressure control valve must be capable of handling the extra oil flow without causing
large overpressure.

52.2.3.5.3 Pressure-Reducing Control Valves


Pressure-reducing control valves are used together with either positive displacement (always with back
pressure control valve) or centrifugal pumps. These values reduce the pressure of the pump to the pres-
sure required by the machine. Pressure reducing valves are mounted at the discharge of the oil system
just before outlet connection.
For systems with centrifugal pumps, typically, only pressure-reducing valves are needed (back pres-
sure control valve is not required). In many cases involving centrifugal pumps, basic manual control
valves (globe type, butterfly type, custom adjustable orifices) are successfully applied for pupose of pres-
sure control.

52.2.3.5.4 Temperature Control Valves


Temperature control valves are installed on the lube oil system in order to control oil supply tempera-
ture. Either three-way mixing-type or two-way bypass-type valves are used on the oil side. TCV valves
may be self-operating or air-operated type. TCV mixes cold oil coming from the cooler with hot oil
bypassing the cooler to provide the desired supply temperature.
Water flow–modulating valves can be installed on the water supply to the cooler for the purpose of
temperature control. In that case, valves modulate the amount of water flowing through the cooler to
maintain oil temperature.
In low demanding applications and where coolant temperature and bearing heat loads are steady, it
is also possible to apply manual-type valves either on oil or water sides for the purpose of temperature
control.

52.2.3.6  Auxiliary Elements


Lubrication units depending on location and application will contain a number of auxiliary compo-
nents completing the system. The most common auxiliary elements found in oil lubrication systems are
the following:
1. Reservoir heaters: if oil lubrication units are located outdoors in a cold climate or in unheated
indoor locations, it is necessary to equip the reservoir with a heater. In most cases, it is not advis-
able to start pumps if the oil temperature is below 10°C–21°C (50°F–70°F) due to limitations in the
pumps’ capability to handle viscous oil. The most common type of heater is a direct immersion
electric unit mounted through the wall of the reservoir. In many cases, especially for larger vol-
ume tanks or where the application does not allow for the system to shut down, to replace failed
heaters, either pipe insert or reservoir top–mounted heaters are used. Pipe insert heaters have
heating elements located inside a pipe or tube. The complete heater or individual elements can
be then removed and replaced without draining the tank. Top-mounted heaters can be straight
or “L”-shaped in design, and similar to pipe insert heaters, they can be removed during system
Lubrication Oil Systems 52-17

operation. It is important to properly design a top-mounted heater as its large portion is exposed
to air—this part of the heater is known as a “cold section.” The cold section must be long enough
to reach below the minimum oil level in the reservoir. In some cases, angle-mounted heaters are
also used. They are similar to top-mounted units but are mounted on an angle and enter the oil
reservoir above maximum level through the side wall.
2. Accumulators: accumulators are used in the oil lubrication system for two purposes. The first and
most common reason is for pump switch over. Many severe applications cannot tolerate even a
brief loss of oil supply. In such cases, should a working pump fail, it is necessary to provide suf-
ficient oil supply before a standby pump takes over. Accumulators serve very well for that purpose
and will reliably provide oil until standby pump builds the required pressure. Accumulators are
sized typically for between 2 and 4 seconds of oil supply for pump switch-over applications. The
second and less common purpose of using accumulators is to provide emergency oil supply in
case of complete oil lube unit failure.
3. Valves: every oil lubrication unit contains a number of different valves. These will include check
valves and manual valves used for isolation, venting, and fluid sampling or maintenance pur-
poses. To avoid accidental draining of the oil, the drain valve should be locked closed.
4. Sampling Values: the location of primary and secondary sampling valves is critical to provide
appropriate condition monitoring of oil and some components, specifically for determining oil
cleanliness level or moisture contamination. Typically, two locations are selected for primary
sampling valves: on the main return line or on the main supply line upstream of the system filters.
In addition, several secondary sampling valves may be required for additional diagnostics, prefer-
ably immediately downstream of critical components.
5. Another desirable feature is proper connections for performing an external flushing of the circu-
lating oil system.
6. Instrumentation: to properly monitor operation of the oil lubrication system, it is necessary to
install some indicators or other monitoring instruments. As a minimum, it is recommended to
provide the following instruments:
a. Pressure gauge on oil supply
b. Temperature gauges before and after cooler
c. Differential pressure gauge for filters
d. Oil reservoir level gauge
e. High and low oil reservoir level switches
f. Oil flow monitors
A central instrument panel can provide coordinated control and monitoring information for
operators. If remote indication or alarms are required, a number of switches of transmitters should
be installed. For more sophisticated systems, a modern computer-based control system may be used.
7. Oil-conditioning sensors: in the last decade, oil-conditioning sensors have become more popu-
lar to enhance end-user predictive maintenance programs. Several different types of sensors are
available primarily to detect wear debris, oil viscosity, and moisture concentration. Sensors may
indicate alarms or trends if they are associated with special software.

52.2.3.7  Piping Arrangements


Proper sizing of piping for required flow and pressures is necessary for the oil lubrication systems to
operate safely and properly. In general, the piping system should be designed to minimize local heating
from external sources, allow for proper thermal expansion, and withstand vibration. If required, at least
a reflector plate should be used to minimize heat transfer from local sources.
Guidance is provided in Table 52.1 on determining proper line sizes of pipes, tubes, and hoses. Sizing
piping wall thickness to resist operating pressure is equally important. ASME standards provide formu-
las on determining pipe wall thickness, and manufacturers’ catalogues offer the same information for
52-18 Design for Lubrication and Tribology

tubing and hoses [3]. If threaded pipe joints are used, it is necessary to remember that cutting threads
removes a substantial amount of material. It is common practice to use heavier walls for small diameter
pipes just for that reason (schedule 80 pipes for sizes up to 2 NPS—nominal pipe size for carbon steel
pipe [50 DN—nominal diameter—with 5.54 mm wall thickness] and sizes up to 1.25 NPS for stainless
steel pipes [32 DN with 4.85 mm wall thickness]).
Material selection of the piping plays a significant role in many applications. Where most lubrica-
tion systems can accept carbon steel pipes, there are many cases where stainless steel lines are required
or preferred. Stainless steel lines are commonly used in refineries, offshore applications, areas with

21
PDIS

26 23 21
18 6
20 PS 20 PI 20 Filter 20
FIS
TI TS 17
A C

B
Lube oil
supply
Filter
Cooler
19

16

24

15 15
14 14
13
J Cooling
water
13 supply

10 11 10 11
9 9 Water
Motor Motor return
7 cooling
3
LSL
12 12
Lube oil
return

LG
2 TI
HTR
8 8
E 5

19 25
4 TS
6
1 Cooler Cooler

Components list: Components list: 19


1 – Reservoir 14 – Check valve
2 – Heater 15 – Isolation valve
3 – Filler/breather 16 – Cooler
4 – Drain valve 17 – Temperature control valve H H
5 – Level/temperature indicator 18 – Temperature indicator Cooling J J Cooling
water water
6 – Temperature switch 19 – Filter supply Cooling Cooling
supply
7 – Level switch 20 – Isolation valve water return water return
8 – Suction strainer 21 – Differential pressure indicator
9 – Pump and switch
22 – Pressure indicator
10 – Motor
11 – Pump/motor adapter 23 – Pressure switch
24 – Pressure control valve Dual cooler option
12 – Drive coupling
25 – Transfer valve
13 – Relief valve 26 – Flow meter/switch

FIGURE 52.12  Typical P&ID (process and information diagram) of a small oil lubrication oil system.
Lubrication Oil Systems 52-19

Back pressure control value Duplex filter


Pump relief valve
Differential pressure
gauge
Pressure
switch Temperature
gauge

Pressure
gauge
Motor

Oil supply
Temperature
switch
Fill conn. / breather

Oil return
Pump

Cooler
Heater
Oil reservoir
Level indicator
Drain value

FIGURE 52.13  Layout of typical small oil lubrication system (per P&ID shown in Figure 52.12). (Courtesy of
Howard Marten Company Ltd., Pickering, Ontario, Canada.)

Accumulators

Oil reservoir Relief value

Duplex Cooler
filter

Electrical box Heaters Level gauge


Instrument panel Emergency pump

AC motors AC pumps Emergency DC motor

FIGURE 52.14  Layout of a large oil lubrication system. (Courtesy of Howard Marten Company Ltd., Pickering,
Ontario, Canada.)
52-20 Design for Lubrication and Tribology

TABLE 52.3  Troubleshooting Chart


Problem Description
Possible Causes Remedy

Noisy Pump
Cavitations Check suction line for obstruction or dirty suction strainer
Viscosity too high Check oil recommended by machine supplier for given temperature
and service
Pump picking up air
Low oil level in reservoir Add oil to reservoir
Loose or damaged suction line Tighten or replace line
Worn or damaged shaft seal Replace
Incorrect rotation Check and correct
Worn or damaged internal parts Repair or replace pump
Worn bearings Repair or replace pump

Low or Erratic Pressure


Low oil level Add oil to reservoir and ensure pump suction line is submerged
Relief valve not functioning properly:
Valve sticking and leaking Check for broken spring, check piston for dirt or scoring and replace
parts as necessary, reset
Valve setting too low Increase pressure setting to operating pressure ± 25 psi (1.72 bar)
(higher setting may be needed)

No Pressure in the System


Low oil level Add oil to reservoir and ensure pump suction line is submerged
Pump not running or running in reverse Check if pump runs and check rotation
Pump’s shaft or drive coupling broken Repair or replace
Relief valve stuck open Check valve for dirt or piston scoring and repair
Faulty gauge Check gauge or replace
Outlet check valve of the standby pump Check seat of the valve for dirt or scoring; replace valve if needed
leaking
Pressure control valve setting incorrect Adjust valves setting

Erratic Flow Rate


Excessive foaming Check for air leaks in pump suction
Check oil returning from bearings
Low oil level Add oil to the reservoir and check for leaks
Faulty pump operation See “noisy pump” given earlier
Overheating of the System
No cooling water supply Check cooling water isolation valves
Clogged cooler Check cooler condition; clean with air, water, or brushes; and repair/
replace as necessary
Cooler fan motor wrong rotation (on air Check rotation and correct
coolers)
Cooler fan motor not operating (on air Check motor power supply
coolers) Check motor bearings
Repair or replace as needed
Heater thermostat set too high Reset thermostat
Continuous operation of relief valve Check relief valve and pressure control valves
Check for system blockage and reset valves as needed
Viscosity too high Check recommended type of oil
Lubrication Oil Systems 52-21

TABLE 52.3 (continued)  Troubleshooting Chart


Problem Description
Possible Causes Remedy
Blocked component beyond pump, Inspect all components
restricting oil flow
Malfunction of temperature control Replace thermal assembly (for self-operating valves)
valve Check valve components and setting for air-operated control valves
Cooler undersized Check cooler sizing sheet and actual heat generated by machine
Cooling water temperature too high Check design cooling water temperature

Reservoir Heating Problem


Heater not operating Check power supply
Check for loose leads
Check heater power consumption
Burned heater Replace
Long heating time Check design heating time (in most cases, design heat-up time is 12 h)
Check location—high wind will extend heating time

corrosive environments, and systems requiring high cleanliness. Many applications also use mixed pip-
ing with carbon steel before the filter and stainless steel after the filter.
The type of fittings used on the oil lubrication unit plays a significant role in how clean the system is.
Where typical applications will accept threaded or socked welded joints, butt welds are used where high
cleanliness is desired. Butt welds are also used in pipe sizes over 2 NPS (50 DN) (threaded or socket weld
fittings are not common is sizes above 2 NPS).

52.2.4  Typical Upgrades


Many oil lubrication units are fitted with additional equipment, enhancing their performance and func-
tionality. Typical upgrades will include the following:

• Oil purification units (applicable for large-volume tanks or systems with substantial water
ingress). Typically, either a centrifuge or vacuum dehydration unit is used. In some applications,
coalescing filters are also applied
• Kidney purification/filtering units (small flow applications)
• Emergency DC pumps (used for safe equipment shutdown)
• Gravity rundown tanks (used for safe equipment shutdown)
• Duplex components such as pumps, filters, coolers
• Isolation valves used for system maintenance
• Various types of instrumentation adding in monitoring and troubleshooting
• VFD drives for air coolers (used in place of temperature control valves)

52.2.5 Typical Oil Lubrication System Process and


Instrumentation Diagram and Layout
A diagram (see Figure 52.12) and layout (see Figure 52.13) show, typical small oil lubrication system
with flow range between 8 and 152 L/m (2 and 40 GPM). 3D layouts provide details of piping and tank
arrangement.
Figure 52.14 shows a sample of a large oil lubrication system designed to meet American Petroleum
Institute Specification (API 614 Chapter 2) [4].
52-22 Design for Lubrication and Tribology

52.2.6  Troubleshooting
General guidance for trouble shooting and maintenance of the oil lubrication units is provided in Table
52.3. Detailed information on maintenance of individual components such as pumps, coolers, filters,
etc., is typically found in documentation provided by system manufacturers (Table 52.3).

References
1. Booser, E.R., Tribology Data Handbook, The Society of Tribologists and Lubrication Engineers,
CRC Press, Boca Raton, FL, 1997.
2. Heshmat, H., Starved bearing technology: Theory and experiment, PhD thesis, Renselaer Polytechnic
Institute, Troy, New York, 1988.
3. American Society of Mechanical Engineers Standard B31.1, Power Piping.
4. American Petroleum Institute Standard 614, Lubrication, Shaft-sealing and Oil-control Systems and
Auxiliaries for Petroleum, Chemical and Gas Industry Services.
53
Surface Texturing
53.1 Introduction..................................................................................... 53-1
53.2 Background....................................................................................... 53-3
53.3 Basic Modeling.................................................................................53-5
53.4 Optimization....................................................................................53-8
53.5 Microreservoirs and Microtraps...................................................53-8
Izhak Etsion 53.6 Recent Studies..................................................................................53-9
Technion—Israel 53.7 Summary...........................................................................................53-9
Institute of Technology References................................................................................................... 53-10

53.1  Introduction
Surface texturing as a means for enhancing tribological properties of mechanical components is well
known for many years. Perhaps the most familiar and the earliest commercial application of surface tex-
turing is that of cylinder liner honing. For many years, the landing zone of thin-film magnetic storage
disks was laser textured, and surface texturing is also being considered as a means for overcoming adhe-
sion and stiction in micro-electro mechanical systems (MEMS) devices. Fundamental research work
on various forms and shapes of surface texturing for tribological applications is carried out worldwide
and various texturing techniques are employed in these studies, including abrasive blasting, ultrasonic
machining, electrical discharge machining, ion beam texturing, vibrorolling, various etching tech-
niques, and laser texturing. Of all the practical microsurface patterning methods, it seems that laser
surface texturing (LST) offers the most promising concept. This is because the laser is extremely fast and
allows short processing times; it is clean to the environment and provides excellent control of the shape
and size of the texture, which allows realization of optimum designs. By controlling energy density,
the laser can safely process hardened steels, ceramics, and polymers as well as crystalline structures.
Indeed, LST is gaining more and more attention in the tribology community as is evident from the
growing number of publications on this subject. LST produces a very large number of microdimples on
the surface (see Figure 53.1), and each of these microdimples can serve either as a microhydrodynamic
bearing in cases of full or mixed lubrication, a microreservoir for lubricant in cases of starved lubrica-
tion conditions, or a microtrap for wear debris in either lubricated or dry sliding.
The pioneering work on LST started at Technion in Israel as early as 1996 [1,2]. At about the same
time, work on LST was done in Germany but, unfortunately, most of it is published in the German
language and, hence, is not even referenced in English archive journals. A few exceptions are papers
from the group led by Geiger at the University of Erlangen-Nuremberg, for example [3,4]. This group
used an excimer laser with a mask projection technique; a mask is illuminated with the laser beam
and its geometrical information is projected onto the textured surface. This method was applied to a
punch, used in a backward cup extrusion process for the production of rivets, and showed a substantial
increase of up to 169% in cold forging tool life. These as well as many other papers on LST are described
in a review of the state of the art of LST covering this subject until 2005 [5]. A recent review on LST and

53-1
53-2 Design for Lubrication and Tribology

FIGURE 53.1  LST regular microsurface structure in the form of circular microdimples, each having a diameter
of the order of 100 μm and a depth of the order of 1–10 μm.

Publication numbers
70

60

50

40

30

20

10

0
97

99

01

02

03

06

07

08

10
04
96

98

00

05

09
19

19

20

20

20

20

20

20

20
19

19

20

20

20

20
to
Up

FIGURE 53.2  Number of papers on surface texturing published (in English alone) each year from 1996 to 2010.

applications covers the period through 2007 [6], and another recent review focuses on surface textur-
ing for in-cylinder friction reduction [7]. The evolution of surface texturing, at least from the research
aspect, shows a dramatic growth over the last 3 years as presented in Figure 53.2. Today, the benefits of
surface texturing have been successfully demonstrated in numerous tribological applications including;
automotive, bearings and seals, Elastohydrodynamic (EHD) lubrication, magnetic storage, etc. While
the microhydrodynamic bearing function of textured features in cases of full lubrication has been stud-
ied extensively, both theoretically and experimentally, and is well understood, the other two functions,
namely, microreservoirs for lubricant in cases of starved lubrication conditions and microtraps for wear
debris are still far from completion and much research work is still needed in these directions.
In the following, the current state of the art in surface texturing will be described, focusing mainly on
hydrodynamic and hydrostatic lubrication applications. A brief review of the research work done so far
on the microreservoir and microtrap functions will also be given, and some thoughts on future needs
will be presented.
Surface Texturing 53-3

53.2  Background
Surface texturing is a powerful means of enhancing hydrodynamic lubrication between parallel surfaces
in relative sliding, which otherwise when untextured cannot provide any significant hydrodynamic
load-carrying capacity. This is demonstrated in Figure 53.3, which shows two parallel surfaces, the lower
one moving at a relative sliding velocity U with respect to the upper surface. The surface texturing is
represented in Figure 53.3a by a single protruding asperity attached to the upper surface. The ambient
pressure surrounding the two surfaces is Pa. In the absence of any texturing, the relative sliding velocity
results in viscous shear only without any effect on the pressure between the two flat surfaces and, hence,
zero load-carrying capacity F = 0. The introduction of the protruding asperity changes the local film
thickness into a converging diverging one, which generates a hydrodynamic pressure distribution as
shown in Figure 53.3b. Due to the relative velocity U, the pressure increases above Pa in the converging
portion and decreases below Pa in the diverging portion of the clearance. At low velocities, the maxi-
mum pressure is smaller than the absolute value of the cavitation pressure Pc and the pressure distribu-
tions is antisymmetric about Pa. Integrating the pressure distribution along its axial span results in zero
load capacity since the above and below ambient pressures cancel each other. As the velocity U increases,
the pressure distribution becomes asymmetric about Pa since the minimum pressure value is bounded
from below by the cavitation pressure Pc, while the maximum pressure is not limited. The integration of
the pressure now results in a net positive value F > 0. When the textured surface contains a large number
of protruding asperities, the total load-carrying capacity is the sum of their individual contributions.
Exactly the same effect with an asymmetric pressure distribution, as shown in Figure 53.3b, can be
obtained with indented dimples instead of protruding asperities, only in this case the diverging portion
of the clearance precedes the converging one. It should be noted here that because of the microscale of
the asperities or dimples (see Figure 53.1), the total load capacity of textured parallel surfaces is rela-
tively small compared to the load capacity that can be generated in conventional hydrodynamic slider
bearings. Hence, surface texturing is mostly beneficial in cases where conformal mating surfaces at very
small uniform clearances are required as, for example, in various sealing applications. Also, as shown
in Figure 53.4, the dimple configuration is a better choice for surface texturing compared to that of pro-
trusions. This is mainly because when the surfaces are brought into contact the real contact area with
dimples is much larger than that with protrusions. Hence, the average contact pressure in the dimples
case is much smaller and the wear is much lower. Another advantage of the dimple configuration is the
smaller separation that can be obtained between the mating surfaces, which allows better sealing and
much smaller leakage. Indeed, LST has been proven extremely beneficial in various sealing applications
for liquids and gas (see for example [8–12]). In some of these applications where high-pressure liquid

Pa Pa
U

(a)
P
Low U; F = 0
High U; F > 0

Pa

Pc
(b)

FIGURE 53.3  Schematic description of parallel sliding surfaces: a single protrusion (a) and the hydrodynamic
pressure distribution over the single protrusion (b).
53-4 Design for Lubrication and Tribology

Protrusions Dimples
Complicated etching technology Simple and cheap laser technology

High wear Low wear

High leakage (seals) Low leakage/spacing

FIGURE 53.4  Advantages of a dimple configuration in surface texturing compared to that of protrusions.

po
po hmax c pa heq c pa
hp
x x

(a) (b)

FIGURE 53.5  Comparison of two equivalent sets of parallel surfaces: a step change in the clearance (a) and par-
tial surface texturing (b).

has to be sealed, such high pressure may eliminate the cavitation in the individual dimples and, thus,
hamper the generation of hydrodynamic load-carrying capacity. The solution for this problem is a par-
tial surface texturing adjacent to the high-pressure side as shown, for example, in Figure 53.5. When
the clearance between two parallel surfaces has a step change (see Figure 53.5a) and the larger clearance
hmax is facing the higher pressure p0, a hydrostatic load-carrying capacity can be generated due to the
restriction in the Poiseuille flow caused by the smaller clearance c facing the lower pressure pa. A similar
effect is obtained by “partial” surface texturing, where high-density dimples at the high-pressure end
form an equivalent “step” of height heq as shown in Figure 53.5b. This solution has been used successfully
in high-pressure seals as described in [13]. Both the original “full” texturing and its modification form
of “partial” texturing, as well as textured microgrooves, were successfully applied to piston rings and
cylinder liners [14–17]. The partial texturing concept was also found very useful in generating substan-
tial hydrodynamic load capacity in hydrodynamic slider bearings (see Figure 53.6) and thrust bearings
of the simplest form of parallel sliding disks [18,19]. Here, the effect of the equivalent step is similar to
the very efficient Raleigh step in a slider bearing. Through extensive theoretical modeling it was found
that the most important parameter in surface texturing for full fluid film lubrication is the aspect ratio
of the dimples (depth over diameter ratio). It is this parameter that can be optimized in order to provide
maximum load-carrying capacity, maximum film stiffness, and minimum friction coefficient. The area
density of the texturing also affects the efficiency of surface texturing and this parameter too can be
optimized. Yet another important parameter that can be optimized in cases of partial texturing is the
textured portion ratio Bp/B (see Figure 53.6).

2rp hp
h0

z x
2r1
Bp

FIGURE 53.6  Cross section of a partially laser-surface-textured parallel slider bearing.


Surface Texturing 53-5

53.3  Basic Modeling


We shall use a typical textured seal application [8] to demonstrate the basic modeling of surfaces textur-
ing. The geometrical model of the textured surface is displayed in Figures 53.7 and 53.8. Each dimple
is modeled by an axisymmetric spherical segment with a base radius r p and depth hp (see Figure 53.8).
The dimples are distributed uniformly over the annular surface with an area density Sp. Each dimple is
located in the center of an imaginary square cell of sides 2r1 × 2r1 (see Figure 53.7c) where

rp p
r1 = (53.1)
2 Sp

It is assumed that the clearance between the nominally parallel mating surfaces (see Figure 53.8) is
fully filled with an incompressible viscous (Newtonian) fluid having a constant viscosity μ. The ratio
Ri between the inner and outer radii, r i and ro, of the seal ring under consideration is larger than 0.7.
This allows neglect of curvature effects, and, consequently, a circular sector containing one dimple
column in the radial direction (see Figure 53.7a and b) is assumed to be rectangular, subjected in
the lateral x direction to a relative sliding velocity U, corresponding to the tangential velocity at the
mean radius of the seal ring.
The 2D, steady-state form of the Reynolds equation for an incompressible Newtonian fluid in a lami-
nar flow is given by

∂ ⎛ 3 ∂p ⎞ ∂ ⎛ 3 ∂p ⎞ ∂h
⎜ h ⎟ + ⎜ h ⎟ = 6 mU (53.2)
∂x ⎝ ∂x ⎠ ∂z ⎝ ∂z ⎠ ∂x

x
pout

p(–r1 ; z) p(–r1 ; z)
2r1
2rp

U
2r1

(c)
pin
(b)

ri r0

(a)

FIGURE 53.7  Schematic description of a textured seal ring: (a) dimple distribution, (b) a single dimple column
with its coordinate system and boundary conditions, and (c) individual dimple cell.
53-6 Design for Lubrication and Tribology

C 2rp
h(x,z)

hp

2r1 2r1

x
z

FIGURE 53.8  Film thickness and geometry of dimples in a textured seal.

where
z and x are the radial and circumferential directions’ Cartesian coordinates, respectively
h and p are the local film thickness and pressure, respectively, at a specific point of the seal

In order to reduce Equation 53.2 to a dimensionless form, the dimensionless Cartesian coordinates
X and Z, dimensionless local film thickness H, and dimensionless pressure P are defined as

x z h p
X= ; Z= ; H= ; P= (53.3)
rp rp C pa

where
pa is the ambient pressure
C is the seal clearance

After substitution of Equation 53.3 into Equation 53.2 the Reynolds equation in its dimensionless form
becomes

∂ ⎛ 3 ∂P ⎞ ∂ ⎛ 3 ∂P ⎞ ∂H
⎜H ⎟+ ⎜H ⎟ =Λ (53.4)
∂X ⎝ ∂X ⎠ ∂Z ⎝ ∂Z ⎠ ∂X

where the dimensionless parameter Λ is given by

6 mUrp
Λ= (53.5)
paC 2

By specifying the film thickness distribution H(x,z), the dimensionless parameter Λ, and the relevant
boundary conditions, Equation 53.4 can be solved for the pressure distribution in the seal clearance.
Integrating the pressure over the seal area gives the opening force acting in the axial direction to prevent
contact between the rings for reliable operation of the mechanical seal.
Since the microdimples are evenly distributed, it is assumed that the pressure distribution is periodic
in the circumferential direction with a period equal to the imaginary square cell size 2r1. Because of this,
it is sufficient to consider only one radial dimple column as shown in Figure 53.2b, with the following
boundary conditions:

p(x , z = ri ) = pin ; p(x , z = ro ) = pout (53.6)



Surface Texturing 53-7

Dimensionless values of the pressure at the two edges of one dimple column are obtained by normaliza-
tion according to Equations 53.3. In the circumferential direction, periodicity condition of the pressure
is applied so that

p(x = −r1 , z ) = p(x = r1 , z ) (53.7)


The boundary conditions in dimensionless form are given as follows:

⎛ r ⎞ p ⎛ r ⎞ p
P ⎜⎜ X , Z = i ⎟⎟ = in ; P ⎜⎜ X , Z = o ⎟⎟ = out ;
⎝ rp ⎠ pa ⎝ rp ⎠ pa

⎛ r ⎞ ⎛ r ⎞
P ⎜⎜ X = − 1 , Z ⎟⎟ = P ⎜⎜ X = 1 , Z ⎟⎟ (53.8)
⎝ rp ⎠ ⎝ rp ⎠

The boundary conditions at the inner and outer radii of the seal, Equation 53.6, and the periodicity condi-
tion, Equation 53.7, should be complemented by the conditions at the boundaries of possible cavitation
regions associated with each individual dimple. As explained in Section 53.2, these cavitation regions are
responsible for the asymmetric hydrodynamic pressure distribution and, hence, are the only source for
load-carrying capacity in parallel surface sliding. The relatively simple Reynolds boundary condition,
also known as the Swift–Stieber condition, (see, e.g., Reference 20) or any other more advanced cavitation
boundary condition (e.g., Reference 21) can be assumed. The Reynolds condition implies that on the cavita-
tion boundary the pressure gradient normal to the boundary is zero, and the pressure inside the cavitation
region is retained constant close to zero.
The Reynolds equation, Equation 53.4, with its appropriate boundary conditions, can be solved by
a finite difference method using a nonuniform grid over the radial dimple column, shown in Figure
53.2b, where a denser grid is applied within the area of the dimples. Numerical tests show that, for
dimple density values Sp in the range between 10% and 50%, the best accuracy of pressure calculation
is obtained when the grid applied in the area of the dimples is about five times denser than that outside
of the dimples.
The finite difference method leads to a set of linear algebraic equations for the nodal values of the
pressure, which should be solved with the boundary conditions at the inner and outer radii of the seal,
Equation 53.6, and the periodicity condition, Equation 53.7. This linear equations set may be solved by
various standard methods. The successive overrelaxation Gauss–Seidel iterative method [22] is one pos-
sible method. It requires an initial approximation of the solution and, in the case of a mechanical seal,
for example, a linear hydrostatic pressure distribution may be used for this purpose if a more precise
solution is unknown. Although the iteration algorithm of Gauss–Seidel is not always the most effective
one, this method is convenient for the determination of the previously unknown cavitation regions.
The basic modeling described earlier can be easily modified to fit different applications. These modifi-
cations include, for example, solving a nonlinear Reynolds equation for compressible fluids [11], solving
the Reynolds equation simultaneously with a dynamic equation for piston rings [14] or with the equation
of elasticity for elastomeric seals and bearings [23,24]. Although the theoretical models based on solving
the Reynolds equation showed good agreement with experimental results, for example, [8,13,15,19], it
was argued on several occasions, for example [25], that the Reynolds equation may not be valid when
applied to textured features that have a large aspect ratio (the ratio of depth over diameter or width) and
that the full Navier–Stokes (NS) equations should be employed. In order to clarify this issue, both the
full NS equations and the Reynolds equation were solved for the case of a compressible fluid at no sliding
but with a pressure differential to simulate a hydrostatic gas seal [26]. A comparison between the two
solution methods illustrated that in spite of potential large differences in local pressures, the differences
in load-carrying capacity are small for realistic geometrical parameters of LST. Hence, the Reynolds
53-8 Design for Lubrication and Tribology

equation can be safely used for most LST applications. It should be noted here that, as discussed in [6],
surface texturing was also attempted in nonparallel sliding with full hydrodynamic lubrication applica-
tions such as thrust bearings and journal bearings. However, in these cases the texturing is beneficial
only when the global film convergence is small enough as was shown, for example, in Reference 27 that
deals with textured magnetic recording sliders. In yet another nonconformal contact application such
as ball bearings, it was found that surface texturing can be beneficial only with very shallow dimples
[28]. These limitations underline again the most beneficial use of surface texturing in parallel surface
hydrodynamic and hydrostatic applications.

53.4  Optimization
In order to fully benefit from surface texturing, a proper optimization of the geometrical parameters
must be performed in accordance with the application in hand. This includes the aspect ratio of the
dimples, their area density, and, in cases of partial surface texturing, the textured portion. The pre-
ferred and most efficient way to optimize surface texturing is by parametric analysis in a theoretical
model. This was done in several models for bearings [18,27], various seals [8,11,13,23,24,29], piston
rings [14,30], and magnetic recording tapes [31] to obtain maximum load capacity, minimum friction,
maximum film stiffness, and minimum leakage. Many experimental studies attempting optimiza-
tion by trial-and-error approach can also be found in the literature (see, e.g. [6]). These include, for
example, different texturing geometries and dimple shapes like squares, triangles, ellipses, grooves,
etc. Unfortunately, quite often wrong conclusions are arrived at in theses studies due to insufficient
experimental data. Another typical mistake, which should be avoided in optimizing surface texturing,
concerns a comparison of different texturing shapes based on a common parameter such as dimple
size or area density. Here too, wrong conclusions are usually made regarding the optimum shape for
best performance. The correct procedure for finding an optimum texturing among different shapes is
first to optimize each shape individually in terms of its own optimum aspect ratio and area density and
only then compare the individual optimums to find the ultimate one. Such a procedure is described
in Reference 17, where an optimum conventional untextured barrel-shaped piston ring was compared
with an optimum surface textured cylindrical piston ring in a firing diesel engine resulting in about
4% improvement in fuel consumption with the optimum flat face textured ring. In previous stud-
ies, when the texturing was applied to the conventional barrel-shaped piston ring, no difference was
observed between the textured and untextured cases, leading to the wrong conclusion that surface
texturing has no benefit in piston rings.

53.5  Microreservoirs and Microtraps


As early as 1984, Suh and coworkers [32] used an etching technique to produce a grooved surface for
removing oxide wear debris from the interface of electrical contacts. This technique was later replaced
by abrasive machining to form grooves that the authors termed “undulated” surfaces, which were used
in boundary lubrication [33] and as reservoirs for a solid lubricant [34].
The LST technique provides similar benefits when the dimples function as microtraps and, thus,
improve endurance of electrical conductivity [35] and enhance life under fretting fatigue [36]. LST was
also found beneficial as microreservoirs in boundary lubrication [37,38] and for solid lubricants [39–41].
Differently from the microhydrodynamic function of dimples, the microreservoir and microtrap func-
tions do not as yet have theoretical models and, hence, optimization in these cases still has to be done
experimentally by trial-and-error approach [42].
Some of the future challenges in attempting to get better physical insight on the microtrap and micro-
reservoir functions include, for example, the mechanism by which dimples are filled with wear debris
[35] and what should be the best microreservoir volume compared to the available volume of lubricant
in case of starvation [15]. In Reference 15, the effect of surface texturing geometry on piston ring friction
Surface Texturing 53-9

under starved lubrication condition was studied. A high friction, even higher than that for an untex-
tured case, was observed with the deepest dimples as soon as the oil feed rate dropped. It was expected
that the “microreservoir” feature of the texturing will be best achieved with deep dimples, but it became
clear that this is not the case. A possible explanation for this surprising behavior is provided in [15],
assuming that as the amount of lubricant present between the sliding surfaces diminishes, its distribu-
tion over the surface is different in the untextured and textured cases. In the former case, the lubricant
may be evenly distributed over the entire surface, whereas in the latter it starts to become more concen-
trated in the vicinity of the dimples. In this case, the hydrodynamic effect of the “microhydrodynamic
bearings” is restricted to a smaller surface area than with full lubrication and the clearance between the
sliding surfaces rapidly diminishes. With further reduction of the lubricant feed rate, the contact zone
may become highly starved, and, under certain conditions, the surface texturing may in fact become
detrimental rather than beneficial. It was speculated in [15] that, with very little amount of lubricant,
the deepest dimples are partly empty and, thus, can act as “microtraps” and remove from their vicin-
ity whatever little amount of lubricant that would otherwise stay on an untextured or shallow-textured
surface. Similar effect of increased friction under lubricant starvation can be observed with increasing
area density of surface texturing. Here again, the increasing volume of the microreservoirs may become
detrimental due to “trapping” of the already small amount of available lubricant at the bottom of the
dimples far away from the contact interface where it is most needed.

53.6  Recent Studies


As indicated in Section 53.1, several reviews [5–7] can be found in the literature covering surface textur-
ing studies in the period until 2007. In this section, more recent studies published in the years 2008–
2010 on various aspects of surface texturing will be listed for the benefit of the interested reader. This
will be done without going into further details other than grouping these studies according to their
relevance to a common subject. The list of these subjects includes typical tribological components such
as various bearings, seals, and various engine components. It also includes bio-tribology-related aspects,
elasto-hydrodynamic lubrication (EHL), coating, solid lubricants, cavitation, inertia effects, etc. A large
group of miscellaneous studies on various subjects is also included.
Different aspects of surface texturing in journal bearings can be found in References 43–47, while
slider bearings, thrust bearings, and seals are dealt with in References 48–54. A few studies related to
slide ways [55,56] and a number of studies related to internal combustion engine components [17,57–59]
close the list of conventional tribological mechanical components. Bio-inspired and bio-related subjects
can be found in References 60–66. EHL and soft EHL problems including surface texturing of elasto-
mers and polymers are dealt with in References 23–24 and 67–78. Coating and surface-texturing-related
issues are reported in References 79–84. Studies on surface texturing as a storage for solid lubrication
can be found in References 40–41 and 85–87. Cavitation and inertia effects associated with surface
texturing were studied in References 88–92. Within the large group of miscellaneous studies, many are
devoted to various methods of surface texturing. These include laser processing [93–98], various grind-
ing and cutting methods [99–103], along with various other techniques [104–108]. Another subgroup of
miscellaneous studies contains experimental investigations on the effect of geometry and orientation of
various textured shapes on the tribological performance of textured surfaces [109–120]. Other different
miscellaneous aspects of surface texturing can be found in References 121–132.

53.7  Summary
Surface texturing, and more specifically LST technology, has great potential in improving tribological
performance of various mechanical components over a wide range of different operating conditions.
The microdimples produced on the surface can act as microhydrodynamic bearings in cases of full or
mixed lubrication with either incompressible or compressible lubricants. These dimples can serve as
53-10 Design for Lubrication and Tribology

microreservoirs for lubricant in cases of starved lubrication conditions, in EHL, and for solid lubricants,
and they can also provide microtraps for wear debris in either lubricated or dry sliding.
Surface texturing is most beneficial in cases of parallel sliding with full fluid films. In these cases, the
effect of texturing geometry on the tribological performance can be easily modeled and optimized for
best required performance. In nonparallel/nonconformal lubricated contacts, the benefit of surface tex-
turing is rather limited and may even become detrimental [133]. The functions of microreservoirs and
microtraps are not yet fully understood and this is where extensive research is still needed.

References
1. Etsion, I. and Burstein, L., 1996, A model for mechanical seals with regular microsurface structure.
Tribology Transactions, 39, 677–683.
2. Etsion, I., Halperin G., and Greenberg, Y., 1997, Increasing mechanical seal life with laser-textured
seal faces. Proceedings of the 15th International Conference on Fluid Sealing, BHR Group, Maastricht,
the Netherlands, pp. 3–11.
3. Geiger, M., Roth, S., and Becker, W., 1998, Influence of laser-produced microstructures on the tribo-
logical behavior of ceramics. Surface and Coatings Technology, 100–101, 17–22.
4. Geiger, M., Popp, U., and Engel, U., 2002, Eximer laser micro texturing of cold forging tool surface—
Influence on tool life. Annals of the CIRP, 51, 231–234.
5. Etsion, I., 2005, State of the art in laser surface texturing. Journal of Tribology Transaction ASME, 127,
248–253.
6. Etsion, I., 2010, Laser surface texturing and applications, in: Recent Developments in Wear Prevention,
Friction and Lubrication, Ed. G.K. Nikas, Research Signpost, Kerala, India, Chapter 3.
7. Etsion, I., 2010, Surface texturing for in-cylinder friction reduction, in: Tribology and Dynamics of
Engine and Powertrain: Fundamentals, Applications and Future Trends, Ed. H. Rahnejat, Woodhead
Publishing Ltd., New Delhi, India, Chapter 13.
8. Etsion, I., Kligerman, Y., and Halperin, G., 1999, Analytical and experimental investigation of laser-
textured mechanical seal faces. Tribology Transactions, 42, 511–516.
9. Etsion, I., 2000, Improving tribological performance of mechanical seals by laser surface texturing.
Proceedings of the 17th International Pump Users Symposium, Houston, TX, pp. 17–22.
10. Yu, X.Q., He, S., and Cai, R.L., 2002, Frictional characteristics of mechanical seals with a laser-textured
seal face. Journal of Materials Processing Technology, 129, 463–466.
11. Kligerman, Y. and Etsion, I., 2001, Analysis of the hydrodynamic effects in a surface textured circum-
ferential gas seal. Tribology Transactions, 44, 472–478.
12. McNikel, A. and Etsion, I., 2004, Near-contact laser surface textured dry gas seals. Journal of Tribology
Transaction ASME, 126, 788–794.
13. Etsion, I. and Halperin, G., 2002, A laser surface textured hydrostatic mechanical seal. Tribology
Transactions, 45, 430–434.
14. Ronen, A., Etsion, I., and Kligerman, Y., 2001, Friction-reducing surface texturing in reciprocating
automotive components. Tribology Transactions, 44, 359–366.
15. Ryk, G., Kligerman, Y., and Etsion, I., 2002, Experimental investigation of laser surface texturing for
reciprocating automotive components. Tribology Transactions, 45, 444–449.
16. Golloch, R., Merker, G.P., Kessen, U., and Brinkmann, S., 2004, Benefits of laser-structured cylinder
liners for internal combustion engines. Proceedings of the 14th International Colloquium Tribology,
January 13–15 Esslingen, Germany, pp. 321–328.
17. Etsion, I. and Sher, E., 2009, Improving fuel efficiency with laser surface textured piston rings.
Tribology International, 42, 542–547.
18. Brizmer, V., Kligerman, Y., and Etsion, I., 2003, A laser surface textured parallel thrust bearing.
Tribology Transactions, 46, 397–403.
Surface Texturing 53-11

19. Etsion, I., Halperin, G., Brizmer, V., and Kligerman, Y., 2004, Experimental investigation of laser sur-
face textured parallel thrust bearings. Tribology Letters, 17, 295–300.
20. Hayashi H. and Taylor C.M., 1980, A determination of cavitation interfaces in fluid film bearings
using finite element analysis. Journal Mechanical Engineering Science IMechE, 22, 277–285.
21. Elrod, H.G., 1981, A cavitation algorithm. Journal of Lubrication Technology Transaction ASME, 103,
350–354.
22. Gerald, C.F. and Wheately P.O., Applied Numerical Analysis, 5th edn., Addison-Wesley Publishing
Co., New York, 1994.
23. Shinkarenko, A., Kligerman, Y., and Etsion, I., 2009, The effect of elastomer surface texturing in soft
elasto-hydrodynamic lubrication. Tribology Letters, 36, 95–103.
24. Shinkarenko, A., Kligerman, Y., and Etsion, I., 2010, Theoretical analysis of surface textured elasto-
mer sleeve in lubricated rotary sliding. Tribology Transactions, 53, 376–385.
25. Sahlin, F., Glavatskih, S.B., Almkvist, T., and Larsson, R., 2005, Two-dimensional CFD analysis of
micro-patterned surfaces in hydrodynamic lubrication. Journal of Tribology Transaction ASME, 127,
96–102.
26. Feldman, Y., Kligerman, Y., Etsion, I., and Haber, S., 2006, The validity of the Reynolds equation
in modeling hydrostatic effects in gas lubricated textured parallel surfaces. Journal of Tribology
Transaction ASME, 128, 345–350.
27. Murthy, A.N., Etsion, I., and Talke, F.E., 2007, Analysis of textured air bearing sliders with rarefaction
effects. Tribology Letters, 28, 251–261.
28. Mourier, L., Mazuyer, D., Lubrecht, A.A., and Donnet, C., 2006, Transient increase of film thickness
in micro-textured EHL contacts. Tribology International, 39, 1745–1756.
29. Feldman, Y., Kligerman Y., and Etsion I., 2007, Stiffness and efficiency optimization of a hydrostatic
laser surface textured gas seal. Journal of Tribology Transaction ASME, 127, 407–410.
30. Kligerman, Y., Etsion, I., and Shinkarenko, A., 2005, Improving tribological performance of piston
rings by partial surface texturing. Journal of Tribology Transaction ASME, 127, 632–638.
31. Raeymaekers, B., Etsion, I., and Talke F.E., 2007, A model for magnetic tape/guide friction reduction
by laser surface texturing. Tribology Letters, 28, 9–17.
32. Saka, A., Liou, M.J., and Suh, N.P., 1984, The role of tribology in electrical contact phenomena. Wear,
100, 77–105.
33. Tian, H., Saka, N., and Suh, N. P., 1989, Boundary lubrication studies on undulated titanium surfaces.
Tribology Transactions, 32, 289–296.
34. Mosleh, M., Laube, S.J.P., and Suh, N.P., 1999, Friction of undulated surfaces coated with MoS2 by
pulsed laser deposition. Tribology Transactions, 42, 495–502.
35. Varenberg, M., Halperin, G., and Etsion, I., 2002, Different aspects of the role of wear debris in fret-
ting wear. Wear, 252, 902–910.
36. Volchok, A., Halperin, G., and Etsion, I., 2002, The effect of surface regular micro-topography on
fretting fatigue life. Wear, 253, 509–515.
37. Kovalchenko, A., Ajayi, O., Erdemir, A., Fenske, G., and Etsion, I., 2004, The effect of laser texturing
of steel surfaces and speed-load parameters on the transition of lubrication regime from boundary
to hydrodynamic. Tribology Transactions, 47, 299–307.
38. Kovalchenko, A., Ajayi, O., Erdemir, A., Fenske, G., and Etsion, I., 2005, The effect of laser surface
texturing on transition in lubrication regimes during unidirectional sliding contact. Tribology
International, 38, 219–225.
39. Voevodin, A.A. and Zabinski, J.S., 2006, Laser surface texturing for adaptive solid lubrication. Wear,
261, 1285–1292.
40. Rapoport, L., Moshkovich, A., Perfilyev, V., Lapsker, I., Halperin, G., Itovich, Y., and Etsion, I., 2008,
Friction and wear of MoS2 films on laser textured steel surfaces. Surface and Coatings Technology,
202, 3332–3340.
53-12 Design for Lubrication and Tribology

41. Rapoport, L., Moshkovich, A., Perfilyev, V., Gedanken, A., Koltypin, Yu., Sominski, E., Halperin,
G., and Etsion, I., 2009, Wear life and adhesion of solid lubricant films on laser-textured steel sur-
faces. Wear, 267, 1203–1207.
42. Gualtieri, E., Menabue, C., Rettighieri, L., Borghi, A., and Valery, S., 2009, Surface micro-texturing for
improving tribological behavior of steel. Metallurgia Italiana, 10, 49–53.
43. Sinanoglu, C., Nair, F., and Goksen, E., 2008, Effect of micro and macro pits of journal surface on
radial pressure distribution of journal bearing. Indian Journal of Engineering and Materials Sciences,
15, 300–310.
44. Korzynski, M., 2009, Relief making on bearing sleeve surface by eccentric burnishing. Materials
Processing, 209, 131–138.
45. Sinanoglu, C., 2009, Design of neural model for analysing journal bearings considering effects of
transverse and longitudinal profile. Industrial Lubrication and Tribology, 61, 132–139.
46. Kango, S. and Sharma, R.K., 2010, Studies on the influence of surface texture on the performance of
hydrodynamic journal bearing using power law model. International Journal of Surface Science and
Engineering, 4, 505–524.
47. Pyoun, Y.S., Park, J.H., Suh, C.M., Cho, I., Lee, C.S., Park, I.G., Amanov, A., Gafurov, A., and Park,
J., 2010, Tribological characteristics of radial journal bearings by ultrasonic nanocrystal surface
modification technology. International Journal of Modern Physics B, 24, 3011–3016.
48. Cupillard, S., Cervantes, M.J., and Glavatskih, S., 2008, Pressure buildup mechanism in a textured
inlet of a hydrodynamic contact. Journal of Tribology Transaction ASME, 130, 021701.
49. Li, H., Zheng, H., and Yoon, Y.C., 2009, Air bearing simulation for bit patterned media. Tribology
Letters, 33, 199–204.
50. Cupillard, S., Glavatskih, S., and Cervantes, M.J., 2009, 3D thermohydrodynamic analysis of a tex-
tured slider. Tribology International, 42, 1487–1495.
51. Wan, Y. and Xiong, D.S., 2008, The effect of laser surface texturing on frictional performance of face
seal. Journal of Materials Processing Technology, 197, 96–100.
52. Bai, S.X., Peng, X.D., Li, Y.F., and Sheng, S., 2010, A hydrodynamic laser surface-textured gas mechan-
ical face seal. Tribology Letters, 38, 187–194.
53. Rahmani, R., Mirzaee, I., Shirvani, A., and Shirvani, H., 2010, An analytical approach for analysis
and optimisation of slider bearings with infinite width parallel textures. Tribology International,
43, 1551–1565.
54. Tonder, K., 2010, Dimpled pivoted plane bearings: Modified coefficients. Tribology International,
43, 2303–2307.
55. Nakano, M., Miyake, K., Korenaga, A., Sasaki, S., and Ando, Y., 2009, Tribological properties of pat-
terned NiFe-covered Si surfaces. Tribology Letters, 35, 133–139.
56. Ogawa, H., Sasaki, S., Korenaga, A., Miyake, K., Nakano, M., and Murakami, T., 2010, Effects of surface
texture size on the tribological properties of slideways. Proceedings of the Institution of Mechanical
Engineers Part J—Journal of Engineering Tribology, 224, 885–890.
57. Gangopadhyay, A. and McWatt, D.G., 2008, The effect of novel surface textures on tappet shims on
valvetrain friction and wear. Tribology Transactions, 51, 221–230.
58. Caciu, C., Decenciere, E., and Jeulin, D., 2008, Numerical analysis of the consequences of roughness
modifications in 3D hydrodynamic contacts. Tribology Transactions, 51, 483–493.
59. Caciu, C., Decenciere, E., and Jeulin, D., 2008, Parametric optimization of periodic textured surfaces
for friction reduction in combustion engines. Tribology Transactions, 51, 533–541.
60. Shafiei, M. and Alpas, A.T., 2008, Fabrication of biotextured nanocrystalline nickel films for the reduc-
tion and control of friction. Materials Science and Engineering C—Biomimetic and Supramolecular
Systems, 28, 1340–1346.
61. Lee, Y.C., Thompson, H.M., and Gaskell, P.H., 2009, Thin film flow over flexible membranes contain-
ing surface texturing: Bio-inspired solutions. Proceedings of the Institution of Mechanical Engineers
Part J—Journal of Engineering Tribology, 223, 337–345.
Surface Texturing 53-13

62. Gao, L.M., Yang, P.R., Dymond, I., Fisher, J., and Zhongmin, J., 2010, Effect of surface texturing on the
elastohydrodynamic lubrication analysis of metal-on-metal hip implants. Tribology International,
43, 1851–1860.
63. Man, H.C., Chiu, K.Y., and Guo, X., 2010, Laser surface micro-drilling and texturing of metals for
improvement of adhesion joint strength. Applied Surface Science, 256, 3166–3169.
64. Kustandi, T.S., Choo, J.H., Low, H.Y., and Sinha, S.K., 2010, Texturing of UHMWPE surface via NIL
for low friction and wear properties. Journal of Physics D—Applied Physics, 43, Art. 015301.
65. Gao, L.M., Meng, Q.E., Liu, F., Fisher, J., and Jin, Z.M., 2010, The effect of aspherical geometry
and surface texturing on the elastohydrodynamic lubrication of metal-on-metal hip prostheses
under physiological loading and motions. Proceedings of the Institution of Mechanical Engineers
Part C—Journal of Mechanical Engineering Science, 224, 2627–2636.
66. Zhao, W.J., Wang, L.P., and Xue, Q.J., 2010, Fabrication of low and high adhesion hydrophobic Au sur-
faces with micro/nano-biomimetic structures. Journal of Physical Chemistry C, 114, 11509–11514.
67. He, B., Chen, W., and Wang, Q.J., 2008, Surface texture effect on friction of a microtextured
poly(dimethylsiloxane) (PDMS). Tribology Letters, 31, 187–197.
68. Vaverka, M., Vrbka, M., Poliscuk, R., Krupka, I., and Hartl, M., 2008, Numerical evaluation of pressure
from experimentally measured film thickness in EHL point contact. Lubrication Science, 20, 47–59.
69. Krupka, I., Vrbka, M., and Hartl, M., 2008, Effect of surface texturing on mixed lubricated non-
conformal contacts. Tribology International, 41, 1063–1073.
70. Mourier, L., Mazuyer, D., Lubrecht, A.A., Donnet, C., and Audouard, E., 2008, Action of a femtosec-
ond laser generated micro-cavity passing through EHL contact. Wear, 264, 450–456.
71. Shinkarenko, A., Kligerman, Y., and Etsion, I., 2009, The validity of linear elasticity in analyzing sur-
face texturing effect for elastohydrodynamic lubrication. Journal of Tribology Transaction ASME,
131, 021503.
72. Krupka, I., Poliscuk, R., and Hartl, M., 2009, Behavior of thin viscous boundary films in lubricated
contacts between micro-textured surfaces. Tribology International, 42, 535–541.
73. Shinkarenko, A., Kligerman, Y., and Etsion, I., 2009, The effect of surface texturing in soft elasto-
hydrodynamic lubrication. Tribology International, 42, 284–292.
74. Wang, X.L., Liu, W., Zhou, F., and Zhu, D., 2009, Preliminary investigation of the effect of dimple size
on friction in line contacts. Tribology International, 42, 1118–1123.
75. Vrbka, M., Samanek, O., Sperka, P., Navrat, I., Krupka, I., and Hartl, M., 2010, Effect of surface tex-
turing on rolling contact fatigue within mixed lubricated non-conformal rolling/sliding contacts.
Tribology International, 43, 1457–1465.
76. Mourier, L., Mazuyer, D., Ninove, F.P., and Lubrecht, A.A., 2010, Lubrication mechanisms with laser-
surface-textured surfaces in elastohydrodynamic regime. Proceedings of the Institution of Mechanical
Engineers Part J—Journal of Engineering Tribology, 224, 697–711.
77. Zhu, D., Nanbu, T., Ren, N., Yasuda, Y., and Wang, Q.J., 2010, Model-based virtual surface textur-
ing for concentrated conformal-contact lubrication. Proceedings of the Institution of Mechanical
Engineers Part J—Journal of Engineering Tribology, 224, 685–696.
78. Kanakasabai, V., Warren, K.H., and Stephens, L.S., 2010, Surface analysis of the elastomer in lip
seals run against shafts manufactured with micro-cavity patterns. Proceedings of the Institution of
Mechanical Engineers Part J—Journal of Engineering Tribology, 224, 723–736.
79. Wan, D.P., Chen, B.K., Shao, Y.M., Wang, S.L., and Hu, D.J., 2008, Microstructure and mechanical
characteristics of laser coating-texturing alloying dimples. Applied Surface Science, 255, 3251–3256.
80. Yasumaru, N., Miazaki, K., and Kiuchi, J., 2008, Control of tribological properties of diamond-
like carbon films with femtosecond-laser-induced nanostructuring. Applied Surface Science, 254,
2364–2368.
81. Wan, Y., Xiong, D.S., and Li, J.L., 2010, Cooperative effect of surface alloying and laser texturing on
tribological performance of lubricated surfaces. Journal of Central South University of Technology, 17,
906–910.
53-14 Design for Lubrication and Tribology

82. Chouquet, C., Gavillet, J., Ducros, C., and Sanchette, F., 2010, Effect of DLC surface texturing on fric-
tion and wear during lubricated sliding. Materials Chemistry and Physics, 123, 367–371.
83. Qian, S.Q., Zhu, D., Qu, N.S., Li, H., and Yan, D., 2010, Generating micro-dimples array on the hard
chrome-coated surface by modified through mask electrochemical micromachining. International
Journal of Advanced Manufacturing Technology, 47, 1121–1127.
84. Lamraoui, A., Costil, S., Langlade, C., and Coddet, C., 2010, Laser surface texturing (LST) treatment
before thermal spraying: A new process to improve the substrate-coating adherence. Surface and
Coatings Technology, 205, s164–s167.
85. Basnyat, R., Luster, B., Muratore, C., Voevodin, A.A., Haasch, R., Zakeri, R., Kohli, P., and Aouadi, S.M.,
2008, Surface texturing for adaptive solid lubrication. Surface and Coatings Technology, 203, 73–79.
86. Li, J.L., Xiong, D.S., Dai, J.H., Huang, Z., and Tyagi, R., 2010, Effect of surface laser texture on friction
properties of nickel-based composite. Tribology International, 43, 1193–1199.
87. Li, J.L., Xiong, D.S., Wu, H.Y., Huang, Z., Dai, J., and Tyagi, R., 2010, Tribological properties of laser sur-
face texturing and molybdenizing duplex-treated Ni-base alloy. Tribology Transactions, 53, 195–202.
88. Qiu, Y. and Khonsari, M.M., 2009, On the prediction of cavitation in dimples using a
mass-conservative algorithm. Journal of Tribology Transaction ASME, 131, 041702.
89. Fesanghary, M. and Khonsari, M.M., 2010, On self-adaptive surface grooves. Tribology Transactions,
53, 871–880.
90. Dobrica, M.B., Fillon, M., Pascovici, M.D., and Cicone, T., 2010, Optimizing surface texture for
hydrodynamic lubricated contacts using a mass-conserving numerical approach. Proceedings of the
Institution of Mechanical Engineers Part J—Journal of Engineering Tribology, 224, 737–750.
91. Han, J., Fang, L.A., Sun, J.P., and Ge, S., 2010, Hydrodynamic lubrication of microdimple textured
surface using three-dimensional CFD. Tribology Transactions, 53, 860–870.
92. Cupillard, S., Glavatskih, S., and Cervantes, M.J., 2010, Inertia effects in textured hydrodynamic con-
tacts. Proceedings of the Institution of Mechanical Engineers Part J—Journal of Engineering Tribology,
224, 751–756.
93. Harimkar, S.P. and Dahotre, N.B., 2009, Rapid surface microstructuring of porous alumina ceramic
using continuous wave Nd:YAG laser. Materials Processing, 209, 4744–4749.
94. Zhou, Y.Q., Shao, T.M., and Yin, L., 2009, A method of micro laser surface texturing based on optical
fiber focusing. Laser Physics, 19, 1061–1066.
95. Vilhena, L.M., Sedlacek, M., Podgornik, B., Vizintin, J., Babnik, A., and Mozina, J., 2009, Surface tex-
turing by pulsed Nd:YAG laser. Tribology International, 42, 1496–1504.
96. Semaltianos, N.G., Perrie, W., Cheng, J., French, P., Sharp, M., Dearden, G., and Watkins, K.G., 2010,
Picosecond laser ablation of nickel-based superalloy C263. Applied Physics A, 98, 345–355.
97. Huang, J., Beckemper, S., Gillner, A., and Wang, K., 2010, Tunable surface texturing by polariza-
tion-controlled three-beam interference. Journal of Micromechanics and Microengineering, 20, Art.
No. 095004.
98. Shin, H.M., Choi, H.W., and Kim, S.D., 2010, Hybrid (LASER Plus CNC) process for lubricant groove
on linear guides. International Journal of Advanced Manufacturing Technology, 46, 1001–1008.
99. Stepien, P., 2009, Regular surface texture generated by special grinding process. Journal of
Manufacturing Science and Engineering Transaction ASME, 131, 011015.
100. Greco, A., Raphaelson, S., Ehmann, K., Wang, J.Q., and Lin, C., 2009, Surface texturing of tribologi-
cal interfaces using the vibromechanical texturing method. Journal of Manufacturing Science and
Engineering Transaction ASME, 131, 061005.
101. Denkena, B., Kastner, J., and Wang, B., 2010, Advanced microstructures and its production through
cutting and grinding. CIRP Annals—Manufacturing Technology, 59, 67–72.
102. Yan, J., Zhang, Z., Kuriyagawa, T., and Gonda, H., 2010, Fabricating micro-structured surface by using
single-crystalline diamond endmill. International Journal of Advanced Manufacturing Technology,
51, 957–964.
Surface Texturing 53-15

103. Greco, A., Martini, A., Liu, Y.C., Lin, C., and Wang, Q.J., 2010, Rolling contact fatigue performance of
vibro-mechanical textured surfaces. Tribology Transactions, 53, 610–620.
104. Nair, R.P. and Zou, M., 2008, Surface-nano-texturing by aluminum-induced crystallization of amor-
phous silicon. Surface and Coatings Technology, 203, 675–679.
105. Cannon, A.H. and King, W.P., 2010, Microstructured metal molds fabricated via investment casting.
Journal of Micromechanics and Microengineering, 20, 025025.
106. Parkansky, N., Beilis, I.I., Gindin, D., Alterkop, B., Boxman, R.L., Moshkovich, A., Perfilyev,
V., Rapoport, L., and Rosenberg, Yu., 2010, Steel surface modification by pulsed air arc treatment.
Surface and Coatings Technology, 205, 287–293.
107. Byun, J.W., Shin, H.S., Kwon, M.H., Kim, B.H., and Chu, C.N., 2010, Surface texturing by micro
ECM for friction reduction. International Journal of Precision Engineering and Manufacturing, 11,
747–753.
108. Jucius, D., Guobiene, A., and Grigaliunas, V., 2010, Surface texturing of polytetrafluoroethylene by
hot embossing. Applied Surface Science, 256, 2164–2169.
109. Zum Gahr, K.H., Wahl, R., and Wauthier, K., 2009, Experimental study of the effect of microtexturing
on oil lubricated ceramic/steel friction pairs. Wear, 267, 1241–1251.
110. Galda, L., Pawlus, P., and Sep, J., 2009, Dimples shape and distribution effect on characteristics of
stribeck curve. Tribology International, 42, 1505–1512.
111. Pawlus, P., Galda, L., Dzierwa, A., and Koszela, W., 2009, Abrasive wear resistance of textured steel
rings. Wear, 267, 1873–1882.
112. Suh, M.S., Chae, Y.H., Kim, S.S., Hinoki, T., and Kohyama, A., 2010, Effect of geometrical parameters
in micro-grooved crosshatch pattern under lubricated sliding friction. Tribology International, 43,
1508–1517.
113. Predescu, A., Pascovici, M.D., Cicone, T., Popescu, C.S., Grigoriu, C., and Dragulinescu, D., 2010,
Friction evaluation of lubricated laser-textured surfaces. Lubrication Science, 22, 431–442.
114. Yu, H.W., Wang, X.L., and Zhou, F., 2010, Geometric shape effects of surface texture on the generation
of hydrodynamic pressure between conformal contacting surfaces. Tribology Letters, 37, 123–130.
115. Wang, J.D., Chen, H.S., Han, Z.L., and Chen, D., 2010, Investigation of the effect of milli-scale dim-
ples on planar contact lubrication. Tribology Transactions, 53, 564–572.
116. Yan, D.S., Qu, N.S., Li, H.S., and Wang, X., 2010, Significance of dimple parameters on the friction of
sliding surfaces investigated by orthogonal experiments. Tribology Transactions, 53, 703–712.
117. Meng, F.M., Zhou, R., Davis, T., Cao, J., Wang, Q.J., Hua, D., and Liu, J., 2010, Study on effect of
dimples on friction of parallel surfaces under different sliding conditions. Applied Surface Science,
256, 2863–2875.
118. Galda, L., Dzierwa, A., Sep, J., and Pawlus, P., 2010, The effect of oil pockets shape and distribution on
seizure resistance in lubricated sliding. Tribology Letters, 37, 301–311.
119. Koszela, W., Galda, L., Dzierwa, A., and Pawlus, P., 2010, The effect of surface texturing on seizure
resistance of a steel-bronze assembly. Tribology International, 43, 1933–1942.
120. Miyake, K., Nakano, M., Korenaga, A., Mano, H., and Ando, Y., 2010, Tribological properties of
nanostripe surface structures—A design concept for improving tribological properties. Journal of
Physics D—Applied Physics, 43, Art. 465302.
121. Vincent, C., Monteil, G., Barriere, T., and Gelin, J.C., 2008, Control of the quality of laser surface
texturing. Microsystem Technology, 14, 1553–1557.
122. Gualtieri, E., Borghi, A., Calabri, L., Pugno, N., and Valeri, S., 2009, Increasing nanohardness and
reducing friction of nitride steel by laser surface texturing. Tribology International, 42, 699–705.
123. Shen, C., Huang, W., Ma, G.L., and Wang, X., 2009, A novel surface texture for magnetic fluid lubrica-
tion. Surface and Coatings Technology, 204, 433–439.
124. Zhang, H.S. and Komvopoulos, K., 2009, Scale-dependent nanomechanical behavior and anisotropic
friction of nanotextured silicon surfaces. Journal of Materials Research, 24, 3038–3043.
53-16 Design for Lubrication and Tribology

125. Jackson, R.L., 2010, A scale dependent simulation of liquid lubricated textured surfaces. Journal of
Tribology Transaction ASME, 132, 022001.
126. Luo, B.H., Shum, P.W., Zhou, Z.F., and Li, K.Y., 2010, Preparation of hydrophobic surface on steel by
patterning using laser ablation process. Surface and Coatings Technology, 204, 1180–1185.
127. Marchetto, D., Rota, A., Calabri, L., Gazzadi, G.C., Menozzi, C., and Valeri, S., 2010, Hydrophobic
effect of surface patterning on Si surface. Wear, 268, 488–492.
128. Zhao, W.J., Wang, L.P., and Xue, Q.J., 2010, Design and fabrication of nanopillar patterned Au tex-
tures for improving nanotribological performance. ACS Applied Materials and Interfaces, 2, 788–794.
129. de Kraker, A., van Ostayen, R.A.J., and Rixen, D.J., 2010, Development of a texture averaged Reynolds
equation. Tribology International, 43, 2100–2109.
130. Zhao, W.J., Wang, L.P., and Xue, Q.J., 2010, Influence of micro/nano-textures and chemical modifica-
tion on the nanotribological property of Au surface. Colloids and Surfaces A: Physicochemical and
Engineering Aspects, 366, 191–196
131. Salamon, D., Lammertink, R.G.H., and Wessling, M., 2010, Surface texturing inside ceramic macro/
micro channels. Journal of the European Ceramic Society, 30, 1345–1350.
132. Marian, V.G., Predescu, A., and Pascovici, M.D., 2010, Theoretical analysis of an infinitely wide rigid
cylinder rotating over a grooved surface in hydrodynamic conditions. Proceedings of the Institution
Mechanical Engineers Part J—Journal of Engineering Tribology, 224, 757–763.
133. Kovalchenko, A., Ajayi, O., Erdemir, A., and Fenske, G., 2011, Friction and wear behavior of laser
textured surface under lubricated initial point contact. Wear, 271, 1719–1725.
54
Sliding Bearings
54.1 Introduction....................................................................................54-1
54.2 Types of Sliding Bearings.............................................................54-1
Metallic and Bimetal Bearings  •  Plastic Bearings  •  Multilayer
Bearings  •  Fiber Wound, or Composite Bearings
54.3 Bearing Types and Example Ratings..........................................54-3
54.4 Sliding Bearing Design.................................................................54-4
54.5 Load and Pressure..........................................................................54-4
54.6 Sliding Speed..................................................................................54-7
54.7 PV Values........................................................................................54-8
54.8 Design Parameters and Measurement......................................54-10
Timothy Alan
54.9 Bearing Life................................................................................... 54-11
Parsons
Oiles America Corporation
54.10 Failure Modes and Robustness..................................................54-12
Running Clearance  •  Thermal  •  Fatigue  •  Fluid-Based
Jianpeng Feng Damage  •  Corrosion and Other Chemical Mechanisms
Oiles America Corporation References...................................................................................................54-13

54.1  Introduction
Most sliding bearings are used in applications with relatively high static and dynamic loads, low speeds,
and aggressive environments where roller bearing durability may be limited. Most sliding bearing solu-
tions are either self-lubricated or lubricated with oils or greases. Industries, such as food contact, medi-
cal, water treatment and distribution, clean rooms, and image processing applications, often enjoy the
benefits of grease-free self-lubricated bushing solutions. Self-lubricating sliding bearing technologies
are often selected for use in near-permanent installations where regular greasing maintenance could
be very difficult and expensive. Examples would include bridges, hydroelectric turbine installations,
and difficult to access locations found in some nuclear power installations. While there are many self-
lubricated design methodologies, the primary mechanism of self-lubricated sliding bearings involves
leaching chemistries within the bearing material. Motion causes these chemicals to develop lubricating
films on running surfaces that provide the desired friction and wear characteristics.

54.2  Types of Sliding Bearings


The sliding bearing market consists of an almost endless number of branded products and design types.
This chapter presents some of the most common and reliable types of sliding bearings.

54.2.1  Metallic and Bimetal Bearings


Metallic bearing styles vary greatly. Metallic sliding bearings are fabricated from copper alloys, cast
iron, numerous sintered metal materials, and other metal alloys. Figure 54.1 represents some of the most

54-1
54-2 Design for Lubrication and Tribology

Bronze polka-dot bearings Cross section of oil Bimetal and sintered


impregnated sintered bearing powdered metal bearings

FIGURE 54.1  Common sliding bearings.

common metallic bearing solutions. A very recognizable metallic sliding bearing is the “polka-dot”
bearing. It consists of various alloy substrate metals (usually bronze-based) and a symmetric pattern
of cross-drilled holes through the bushing wall in which lubricating plugs are installed. Most sintered
bearing production methods consist of forming a green shape from powdered metal, then sintering the
green shape with conventional sintering technologies. The most popular sintered powder metal bearings
are subjected to an oil impregnation process following the sintering process to provide some level of
self-lubrication. Many different bimetallic designs are offered, but some of the most popular bimetallic
bearings involve a CuSnPb layer sintered on a metal backing. Most bimetallic bearings are used in well-
lubricated applications.

54.2.2  Plastic Bearings


Base plastics used for plastic bearings include polyacetal (POM), polyamide (Nylon), polyester,
polytetrafluoroethylene (PTFE, or DuPont’s Teflon®), polybutylene terephthalate (PBT), polyphen-
ylene sulfide (PPS), and polyether ether ketone (PEEK). Custom formulations include proprietary
lubricants and chemicals along with numerous readily available chemicals. Plastic-bearing material
developers are constantly trying to develop unique combinations of these chemicals to improve char-
acteristics such as wear resistance, friction, continuous upper temperature limit (CUTL), low tempera-
ture performance, and other desired characteristics. Polymerization and polymer science innovation
is typically driven by the well-known large capital polymer manufacturers. Plastic-bearing material
developers usually innovate using chemistry beyond these readily available polymers.

54.2.3  Multilayer Bearings


Various multilayer-bearing architectures exist in industry. The images in Figure 54.2 display some of
the common multilayer cross-sectional architecture. Two of the most common architectures are lami-
nates of (1) steel backing (e.g., ASTM A109 type1), usually copper-plated, sintered bronze powder, and
a plastic layer (wear surface) and (2) traditional metal alloy backing attached to numerous metal wear
surfaces. The metal wear surface often seen with this second type of multilayer bearing is a powdered
metal sintered to the metal backing where the powdered metal is impregnated with numerous types of
solid lubricants and oils.
These materials are typically produced in a strip steel shape through a calendaring process. The strip
is slit to desired widths and formed into various shapes. The steel strip backing material is acting as a
carrier for the wear surface (Figure 54.3).

Sintered powdered metal with Plastic (usually PTFE)


oils, solid lubricants, etc. filled layer

Bronze powder

Metallic backing

FIGURE 54.2  Multilayer and sintered metal backed bearings.


Sliding Bearings 54-3

FIGURE 54.3  Typical steel backed plastic lined bearings.

Consumers often mistakenly conclude that multilayer bearings have worn excessively when the
bronze layer can be seen. In most cases, visibility of the bronze layer is desired. The bronze helps to carry
load and establish a desired surface that optimizes friction and wear performance.

54.2.4  Fiber Wound, or Composite Bearings


Fiber wound bearings are of interest primarily because of ease and speed of manufacture, good
mechanical properties, and overall durability. Many fiber wound bushings have tensile properties in
excess of 350 MPa. As with plastic plain bearings, there are numerous material technologies avail-
able. However, some of the most common fiber wound bearing technologies are polyester or PTFE
fibers wound with various epoxy resins and possessing lubricating fillers on the running surfaces.
Usually, the lubricating technology is applied to only the first few millimeters of the running sur-
face. The static surface is usually the backing material consisting of the wound fiber and epoxy. The
traditional manufacturing method involves winding epoxy soaked fibers on a rotating mandrel.
The rough wound and cured product is removed from the winding spindle, and then the bushing is
machined to size.

54.3  Bearing Types and Example Ratings


Note that Tables 54.1 and 54.2 represent ranges and broad bearing types. Ratings can vary greatly
within a single bearing type. For example, ratings within a plastic-bearing family can differ by 3×
to 4× when comparing filled and unfilled material formulations. The use of periodic lubrication can
greatly increase certain ratings. It is very common to find wildly varying ratings throughout the
literature for seemingly identical applications. This is often influenced by test conditions differences

TABLE 54.1  Common Sliding Bearing Properties


Max P Range Max V Max PV Range Recommended Running Temperature
Bearing Type (N/mm2) Range (m/s) (N/mm2)(m/s) Surface Condition Range (°C)
Plastic 5–20 <1–2.5 0.5–3 0.8 Ra max −40 to 120
Plugged bronze 15–75 {150} 0.2–0.5 {1} 0.75–1.6 {3.25} 1.6 Ra max −40 to 300
Multilayer 20–50 0.3–2.5 1.5–3.5 0.8 Ra max −40 to 220
Fiber wound 100+ 0.2 1.2 0.8 Ra max −40 to 120
54-4 Design for Lubrication and Tribology

TABLE 54.2  Common Running Surface Material Properties


Bearing Pressure Running Surface Running Surface
Type (N/mm2) Running Surface Material Hardness Roughness Ra
Up to 24.5 C steel (S45C, SNC415, SCM435). Corrosion- HB 150 1.6 max
resistant steel (SUS304, SUS403)
Metallic 24.5–49 Materials above plus surface harden—for HB 250
example, induction and carburizing
49–98 Surface hardening as above plus additional HRC 50
nitriding or hard chrome
Plastic and Up to 49 C steel (S45C, SNC415, SCM435). Corrosion- HB 120 0.8 max
multilayer resistant steel (SUS304, SUS403)
49–98 Surface hardening, carburizing, hard chrome HRC 45
plating, etc.

used to determine the ratings. As a result of all of this, the reader is encouraged to use this as a guide
and consult closely with the manufacturer to optimize proper bearing type, material types, and
application.
The sliding bearing industry routinely specifies Ra running surface condition requirements.
The detailed influences of complex running surface conditions (e.g., peak orientation, peak counts,
Abbott Firestone2 curves etc.) related to sliding bearing performance cannot be adequately described
in this chapter, but the reader is encouraged to research this topic further to better understand these
dependencies.

54.4  Sliding Bearing Design


The product of pressure and velocity is the PV value, an important sliding bearing design factor. Contact
pressure and sliding velocity are specific to each application.

54.5  Load and Pressure


Two types of loads are considered when designing sliding bearings: static load (which remains con-
stant during operation) and dynamic load (which varies in amplitude during operation) (Table 54.3).
Dynamic load types can include the following:
• Impact load
• Periodically repeated load
• Random load
For repeated loads and alternate loads, the average load can be used to evaluate the pressure. When an
impact load is applied, factors of safety should be applied. The following formula can be used to estimate
bearing load:

TABLE 54.3  Coefficient of Impact Load


Degree of Impact fω Example
Few impact loads 1.0–1.2 Electric machine, machine tool, measurements, etc.
Medium impact load 1.2–1.5 Railway, automobile, rolling mill, metal-cutting machine, etc.
Strong impact load 1.5–3.0 Crusher, agriculture machine, construction machine, etc.
Sliding Bearings 54-5

F = fw × Fc

where
F is the bearing load (N)
Fc is the estimated load (N)
fω is the coefficient of load (coefficient of impact)

Note that the aforementioned calculation method cannot be applied for hydrostatic applications. To
better understand hydrostatic application physics, please refer to Chapter 43.
Following the previous calculations, a straightforward solution for applied pressure on sliding bear-
ing surfaces results. If the pressure is beyond the allowable pressure limit, special care should be paid to
ensure the safety of the materials by studying contact pressure.
The nominal pressure can be estimated as follows:

F F
Pmean = =
D×L A

where
Pmean is the bearing mean pressure
F is the bearing load
A is the bearing projected area
D is the bearing inside diameter
L is the bearing length

The actual contact pressure does not distribute evenly on bearing inside surfaces. It is greatest at the
bottom of the bearing (Figure 54.4). If we assume the pressure level changes from zero to a maximum
at the bottom and pressure follows a cosine distribution, mathematical models for both Pmax and Pφ can
be set up to describe the pressure zone:

⎛j ⎞
Pj = Pmax × cos ⎜ ⎟
⎝q ⎠

P = Pmax.cos( . π)
θ
P max

FIGURE 54.4  Pressure graphic.


54-6 Design for Lubrication and Tribology

F
Pmax = q /2
= K c × Pmean
cos (j × p /q ) cos(j )dj

D ×L ×
∫ 0

where the clearance constant Kc is given by

1 p 2 − q2
Kc = = (q<p )
q /2
cos (q / 2) × p × q
cos (j × p / q ) cos(j )dj

0

4
Kc = (q = p )
q + sin q

The contact areas are determined by different aspects, such as clearance, material properties, and loads.
The equation provided here is for metallic shafts and metallic bearings. However, a more compliant
pressure pattern and contact area will result if composite materials with much lower moduli are used.
In these cases, either 3D numerical methods or actual physical tests have to be carried out for contact
pressure pattern and contact area.
The formula for contact angle θ was derived as follows3,4:

⎛ 2b ⎞
q = sin −1 ⎜ ⎟
⎝ Ds ⎠

where

b=
( )
2F 1 − ns /Es + 1 − nb /Eb
2 2
( )
p ×L 1/Ds − 1/Db

F is the applied force


νs and νb are the Poisson’s ratios for shaft and bearing
Es and Eb are the elastic moduli for shaft and bearing
Ds and Db are the diameters of shaft and bearing
L is the length of bearing

The Pmax /Pmean ratios for several central angles are listed in Table 54.4.

TABLE 54.4  Pmax /Pmean Ratio


for Different Contact Angles
Central Angle (°) Pmax/Pmean
30 8.69
60 3.08
90 2.12
120 1.67
150 1.42
180 1.27
Sliding Bearings 54-7

54.6  Sliding Speed


Straightforward methods, described as follows, are applied to determine sliding speed (Figures 54.5
through 54.7).
Bushing type—rotational motion
Rotation:

pdN
Vrotation = (m/s)
103

Oscillation:

pdqC
Voscillation = (m/s)
180 × 103

Washer type—rotational motion


Oscillation:

pDqC
V= (m/s)
180 × 103

Rotation:

pDN
V= (m/s)
103

D D

(a) (b)

FIGURE 54.5  Continuous and oscillating graphic. (a) Continuous rotation. (b) Oscillating rotation.

FIGURE 54.6  Washer geometry.


54-8 Design for Lubrication and Tribology

P P
Bearing
Plate

ød
St

FIGURE 54.7  Axial sliding motion.

2St × C
V=
103

where
V is the sliding velocity (m/s)
N is the revolution (Hz)
C is the cycle (Hz)
d is the shaft diameter (mm)
D is the washer outer diameter (mm)
θ is the oscillating angle (°)
St is the stroke (mm)

54.7  PV Values
The PV value, the product of load (P) and sliding speed (V), is a critical bearing design factor. PV
values represent power loss by a bearing as μ × P × V. Such power loss creates frictional heat. Bearing
temperature under a continuous operation condition can increase when the sliding surface con-
dition changes due to influences such as contamination of foreign particles, material fatigue, and
resultant generation of foreign particles. When this happens, the sliding surface can be damaged and
seizure can occur. To help design for this, calculated PV values are balanced against allowable PV
values for the products of interest. Product catalogs normally include allowable loading capability
(allowable PV value, allowable pressure, and allowable velocity) for each bearing material. However
the PV limit value does not match for all cases. In other words, P, V, and PV are constrained, but
allowable PV values are influenced by various bearing characteristics and lubrication mechanisms,
including designed lubricant type (e.g., solid lubricant plugged bearing or oil retaining bearing)
and plastics bearing material technologies. These other factors are significant influences on allow-
able PV limits and prevent the use of simple PV calculations based on published standard material
yield limits. Additionally, the allowable PV limit is dependent on temperature (again, the balance of
frictional heat and heat transfer out of the bearing), bearing shape, friction conditions, and various
environmental conditions. Generally, the allowable maximum PV value will be smaller than the
product of the allowable maximum contact pressure (Pmax) and the allowable maximum velocity
(Vmax) (Figure 54.8).
The limiting PV values for a material can be determined using a journal bearing test machine. Few
standard test methods exist for journal bearing test machines, but the architecture is typically a speci-
fied shaft rotating in the test bearing. Loads and various environments can be applied through the bear-
ing to the shaft depending on the particular test plan (Figure 54.9).
The method for determining the allowable PV for a particular bearing material is to develop a test
plan consisting of many different velocities and loads. At each velocity, a load-stepping test is conducted.
Sliding Bearings 54-9

Pmax

Allowable maximum. PV value

Service range

Vmax

FIGURE 54.8  Allowable max PV.

Oscillation
Torque
Test Shaf
Couplin

Hydraulic

Motor Hydraulic

FIGURE 54.9  Journal bearing test equipment example.

Each condition is run and the friction torque and contact temperature are monitored and plotted during
the entire test process. At each loading, the friction torque and bearing temperature are allowed to reach
equilibrium and remain there for ∼30 min before the load is increased. When the total load reaches a
certain level, the friction torque and temperature will not be able to stabilize and the slope of the curve
will increase quickly. The sharply increasing friction torque and temperature will cause bearing failure.
The same test procedure is repeated for each velocity and the PV curve is thus generated. The test process
is illustrated in Figure 54.10.
The test methods for tribological behavior (friction and wear) of metal- and thermoplastics-based
bearing materials have been included in ISO5 7148-1 and 7148-26. Those two standards describe and
compare different test methods, such as pin-on-disk, block-on-ring, and plain bearing-on-shaft.
Any selected test method should simulate the actual application as closely as possible. ASTM7 D3702
describes wear rate and coefficient of friction testing using a thrust washer approach. This test method
uses the rotary specimen (thrust washer shape) to run against stationary steel washer (generic tests) or
customized mating materials for recording friction and wear under different test conditions such as
54-10 Design for Lubrication and Tribology

600 6

500 5

400 4

Friction torque (N.m)


Temperature (°F)

300 3
Load-step

200 2

Temperature
100 1
Friction torque

0 0
0 1 2 3 4 5
Time (h)

FIGURE 54.10  PV limitation test procedure.

speed, load, lubrication, and temperature. While journal bearing testing can involve significant varia-
tion, the thrust washer test method produces resolution that can be used to understand statistical differ-
ences for standardized test requirements. This can be important when testing and reviewing data that
needs to contain as little application-induced error as possible. Additional ISO standards define general
materials properties and requirements for sliding bearings. Examples of these include ISO 3547-48, ISO
43819, ISO 4382-110, and-211, ISO 438312, and ISO 4384-113, and -214.

54.8  Design Parameters and Measurement


Although cylindrical bushing designs appear simple, measurement methods are a topic of debate. For
example, the diameters (or more accurately, the cylindricities and concentricities) of many cylindrical
bushing products appear unacceptable when the bushing is measured in the free state with conventional
measurement methods. However, many of these seemingly unacceptable bushings measure well within
specification when they are measured pressed into appropriate housings. To help address these challenges,
quality assurance standards have been introduced by ISO/TC123. As an example, ISO 3547-115 lists the
preferred normal dimensions for wrapped bushings, such as inside and outside diameters, recommended
lengths, and chamfers. Different procedures for testing the outside and inside diameters of wrapped bush-
ings were introduced in ISO 3547-216. Test A procedure is shown in Figure 54.11. The wrapped bushing
is installed between mounting blocks while maintaining the split parallel-to-vertical load direction. The
predefined load is applied while the shaft is inserted only into the wrapped bushings. The vertical dis-
placement difference is used to calculate the bushing outside diameters.
As with wrapped bushings, understanding plastic and composite bushing measurement methods is
critical to successful implementation. As a general rule, deformable bushings should be installed and
measured in specific ring gauges. Attribute data can be obtained with go-no-go pins when the deform-
able bushing is installed in a ring gauge. If deformable bushing variable data is desired, care must be
taken to ensure Measurement System Analysis (MSA) thresholds are satisfied. Often, traditional mea-
surement tools (bore mics, touch probe CMM, etc.) appear to be reliable methods, but they actually
generate excessive measurement error.
Sliding Bearings 54-11

Force
1

2 a

1. Wrap bearing
2. Mounting fixture

FIGURE 54.11  ISO Test A setup for wrapped bushings.

54.9  Bearing Life


Many of the predictive methods show good results when they are associated with well-understood
homogeneous materials and stable environments. The challenges for developing universal wear and life
predictive models for sliding bearings are how to accurately represent the numerous types of materials
and environmental conditions. Additionally, by design, most plain bearing solutions change signifi-
cantly during their life—especially critical elements such as area and pressure. Good examples of this
are the run-in and the steady-state wear and friction differences found in traditional plugged bronze
bushing solutions. The basic proposal in this section is that plain bearing life predictive models are
dependent on wear factors for each material and application family. These wear factors need to be deter-
mined by actual wear tests of the particular bearing product under the conditions that you are most
interested in. With that said, a traditional sliding bearing predictive wear function is

W = K 1 × (P × V × T )

where
W (wear) is the wear dimension (mm)
P is the bearing pressure (kg/cm2)
V is the sliding velocity (m/min)
T is the working time (h)
K1 is the specific wear rate (mm/[kg/cm2·m/min·h])

Wear rates have strong dependencies on many factors, but the Table 54.5 demonstrates how signifi-
cant the impact lubrication can be.
Another design tool is a PV–K function that can be developed for sliding bearing concepts (again,
where K is the wear factor [mm/((kg/cm2)(m/min)h)]). Because temperature can have such a strong
influence on wear and life, this PV–K chart should be used in combination with a T–K chart to under-
stand the strength of the temperature dependency. Combined, these two charts are good tools capable
of producing specific wear rate functions that can be used in the aforementioned bearing life equation to
estimate plain bearing life. Plain bearing manufacturers have usually performed PV–K and T–K testing
on their materials and products. While most sliding bearing manufacturers can clearly explain life pre-
dictions for popular products and applications, more obscure situations may require testing to develop
meaningful wear rate functions. An example of a PV–K study is shown as a reference in Figure 54.12.
54-12 Design for Lubrication and Tribology

TABLE 54.5  Typical Sliding Bearing Specific Wear Rates


Condition Wear Rate
No lubricants 1 × 10−3 to 1 × 10−5
Boundary lubrication by low speed, etc. 1 × 10−5 to 1 × 10−7
Relatively good lubrication by greasing 1 × 10−6 to 1 × 10−8
Hydrodynamic lubrication 1 × 10−8 to 1 × 10−10

1 × 10–5
Specific wear rate K(mm/(kgf/cm2. m/min. h))

5 × 10–6

1 × 10–6
20 50 100 500 1000
PV value (kgf/cm2. m/min)

FIGURE 54.12  PV–K study example.

54.10  Failure Modes and Robustness


The following sections provide some information regarding common reasons for bearing failure.

54.10.1  Running Clearance


Establishing and maintaining proper running clearance are critical for sliding bearing life. While a
number of conditions can cause improper running clearance, the principal causes include incorrect use
of design standards; machining mistakes; and unplanned application conditions of temperature, loads,
or speeds. Too little clearance causes the chain reaction of excessive heat generation, continuous reduc-
tion in clearance due to thermal growth, high wear rates, high friction, and possibly eventual seizure
or catastrophic damage of the assembly. Likewise, excessive clearance can result in poor distribution of
load, chatter, noise, aggressive wear, and eventual failure. Plain bearing literature often identifies general
running clearance recommendations. General recommendations will often be successful, but because
actual clearance is heavily dependent on the particular application, the reader is encouraged to look at
clearance design recommendations at a case-by-case basis.

54.10.2  Thermal
If an application performs outside thermal design limits, failure can occur quickly. The two principal
mechanisms involved in thermal-related failures are running clearance change due to excessive material
expansion or contraction and property effects related to material loss. The effect of running clearance
Sliding Bearings 54-13

change was described earlier and the effect of loss properties is directly related to critical bearing char-
acteristics such as maximum P and PV limitations. In other words, when temperatures increase, usually,
critical material properties decrease. This change in material properties affects bearing capacities and
durabilities. Understanding the thermal environment is especially important when nonmetallic bearing
solutions are being considered.

54.10.3  Fatigue
Fatigue failures are obviously most applicable to relatively low ductility sliding bearing solutions such
as metallic bearings. Fatigue failures are characterized by typical stress fractures and, usually, severe
abrasive wear from dislodged particles. The typical reason for fatigue failures is high cyclic loading.

54.10.4  Fluid-Based Damage


Many sliding bearing applications involve lubricants. Various destructive mechanisms can manifest
themselves if certain clearance and fluid properties occur between the running and the static surfaces,
such as cavitation. Significant pressure gradients across a liquid layer can drive vapor out of the liq-
uid. When this vapor exists in large quantities between the running and the static surfaces, significant
­damage can occur from the loss in lubricating properties and the erosion-type damage that the vapor
can cause.

54.10.5  Corrosion and Other Chemical Mechanisms


While the combinations of possible chemical mechanisms are nearly endless, oxidation-related chem-
istries are some of the most common in sliding bearing applications. The user of sliding bearings
should always understand material compatibility between the bearing material and the environments
to which the bearing is exposed. The user should also be aware that the chemistry of the environment
often changes through the life of the application. One of the classic examples demonstrating a chang-
ing environment through the life of an application is gearbox lubricant drain and replace actions. In
many cases, lubricant corrosion inhibitor chemistries can accelerate unwanted bushing loss property
effects—especially in nonmetallic bushing applications. The regular removal and replacement of new
gearbox lubricants can expose bushings to freshened active chemistries that can significantly acceler-
ate loss property effects that compatibility test results may not simulate. The plain bearing user should
remember the importance of temperature on chemical reactions.

References
1. E.J. Abbott and F.A. Firestone, Specifying surface quality: a method based on accurate measurement
and comparision, Mechanical Engineering 55 (1993): 569–572.
2. R. Budynas and J. Nisbett, Shigley’s Mechanical Engineering Design, 8th Edn., McGraw-Hill Primis,
New York, 2006, p. 123.
3. V. Popov, Contact Mechanics and Friction—Physical Principles and Application, Springer, Heidelberg,
Germany, 2010, p. 61.
4. International Organization for Standardization. www.iso.org
5. ASTM International formerly known as American Society for Testing and Materials. www.astm.org
6. ASTM A109-08, Standard Specification for Steel, Strip, Carbon (0.25 Maximum Percent), Cold-Rolled,
ASTM International, West Conshohocken, PA, 2008.
7. ISO 7148-1, Plain bearings—Testing of the tribological behaviour of bearing materials—Part 1:
Testing of bearing metals, ISO, Geneva, Switzerland, 1999
54-14 Design for Lubrication and Tribology

8. ISO 7148-2, Plain bearings—Testing of the tribological behaviour of bearing materials—Part 2:


Testing of polymer-based bearing materials, ISO, Geneva, Switzerland, 1999.
9. ASTM D3702 94, Standard Test Method for Wear Rate and Coefficient of Friction of Materials in
Self-Lubricated Rubbing Contact Using a Thrust Washer Testing Machine, ASTM International, West
Conshohocken, PA, 2009.
10. ISO 3547-4, Plain bearings—Wrapped bushes—Part 4: Materials, ISO, Geneva, Switzerland, 2006.
11. ISO 4381, Plain bearings—Lead and tin casting alloys for multilayer plain bearings, ISO, Geneva,
Switzerland, 2000.
12. ISO 4382-1, Plain bearings—Copper alloys—Part 1: Cast copper alloys for solid and multilayer
thick-walled plain bearings, ISO, Geneva, Switzerland, 1991.
13. ISO 4382-2, Plain bearings—Copper alloys—Part 2: Wrought copper alloys for solid plain bearings,
ISO, Geneva, Switzerland, 1991.
14. ISO 4383, Plain bearings—Multilayer materials for thin-walled plain bearings, ISO, Geneva,
Switzerland, 2000.
15. ISO 4384-1, Plain bearings—Hardness testing of bearing metals—Part 1: Compound materials, ISO,
Geneva, Switzerland, 2000.
16. ISO 4384-2, Plain bearings—Hardness testing of bearing metals—Part 2: Solid materials, ISO,
Geneva, Switzerland, 1982.
17. ISO 3547-1, Plain bearings—Wrapped bushes—Part 1: Dimensions, ISO, Geneva, Switzerland,
2006.
18. ISO 3547-2, Plain bearings—Wrapped bushes—Part 2: Test data for outside and inside diameters,
ISO, Geneva, Switzerland, 2006.
55
Magnetic Bearings
Alan B. Palazzolo
Texas A&M University 55.1 Overview........................................................................................... 55-1
55.2 Actuators........................................................................................... 55-2
Zhiyang Wang Heteropolar Magnetic Bearings  •  Homopolar Radial Magnetic
Texas A&M University Bearings  •  Thrust (Axial) Magnetic Bearing Actuators  •  Materials
55.3 Sensors...............................................................................................55-8
Jung Gu Lee Inductance Sensors  •  Eddy Current Sensors  •  Optical
Texas A&M University
Sensors  •  Capacitive Sensors  •  Self-Sensing
Albert F. Kascak 55.4 Control............................................................................................ 55-11
US Army Research 55.5 Power Electronics.......................................................................... 55-14
Laboratory 55.6 Finite Element Magnetic Field Simulation................................ 55-16
Andrew J. Provenza 55.7 Auxiliary/Backup/Catcher Bearings.......................................... 55-18
National Aeronautics and Acknowledgments..................................................................................... 55-21
Space Administration References................................................................................................... 55-21

55.1  Overview
Design of magnetic bearings (MBs) requires a good understanding of rotordynamics, controls, elec-
tronics, and electromagnetics. MBs were developed in the 1940s by Jesse Beams [1,2] of the University
of Virginia for ultracentrifuge application. These devices have since been applied to an ever widening
scope of commercial applications such as vacuum pumps, industrial compressors, pumps and turbines,
energy storage flywheels, machine tools, and even artificial hearts [3]. Other potential applications
include ultrahigh-temperature gas turbines [4] and satellite attitude control flywheels [5]. MBs offer
some key advantages over oil or gas film passive bearings such as (a) environmentally friendly, (b) oil-
free operation, (c) elimination of process contamination, (d) reduced footprint from eliminating lube
system, (e) reduced maintenance, (f) reduced power losses, (g) no minimum speed requirement, (h) DN
values (D is the diameter in mm and N is the rotation rate in RPM) greater than 3,000,000, and (i) adap-
tive adjustment of stiffness and damping. Some disadvantages include lower unit load capacity relative
to oil film bearings: (150–200 psi) vs. 600–1000 psi, and generally higher initial cost.
The major components of an active MB (AMB) system are shown in Figure 55.1. The heart of the
AMB system is the MB actuators (MB model), which exerts forces on the journals (rotor model). These
forces pull the journals toward target positions along the shaft near where relative position sensors are
located. The levels and directions of the pull forces are determined by the “brain” of the AMB system,
the controller, which compares the instantaneous positions of the shaft at the sensors to target values
set by the machine operator or manufacturer. The major objectives of the controller are similar to any
other feedback-controlled system: robust stability and disturbance rejection, where disturbances may
include support motions, imbalance, noise, hydraulic or gas loads, etc. The outputs of the controller
are typically small power signals that are amplified to drive currents through coils, which, along with
permanent magnets (PMs), produce the required MB actuator forces. This is accomplished with power
(servo) amplifiers (PA), which are operated in a voltage or current control mode. The servo amplifier

55-1
55-2 Design for Lubrication and Tribology

Disturbance
X Position sensor

– PA
X Position + PA
Decoupling
target Σ Controller IC LPF PA
PA
+ Choke
Y Position Σ Controller CDM PA MB Rotor
IC LPF
target PA
– Disturbance
Y Position sensor

FIGURE 55.1  Major components of an active magnetic bearing system.

TABLE 55.1  Magnetic Bearing Advantages and Disadvantages


Advantages Disadvantages
Very low and predictable friction losses Bearing envelope larger than conventional
No lubrication and associated pumps, piping, Higher initial cost
filters, heaters, etc.
Can operate in a vacuum Requires power, control, and sensing electronics
Speed limited only by journal stress due to A backup bearing is needed for start-up, shutdown, and in case of
centrifugal load power or control failure
Contamination-free operation Cooling is sometimes necessary to remove heat from electrical
losses
Reduced maintenance Most rotordynamics design codes are incapable of performing the
Active (adaptive) control of stiffness, damping, electromechanical simulations required for magnetic bearing and
and alignment machine design
Inherent force control useful for in situ fault
monitoring and parameter identification
Long life and little wear
Extreme temperature operation: cryogenic to 1000°F

(SA) converts the low-power reference signals into high-power signals that provide the AMB actuator
voltage and current. A certain class of AMBs provide fail-safe operation through redundancy in the
flux-driving components [6,7]. The coil currents in the AMB coils are adjusted according with the coil
failure status in order to continue to provide the desired x-, y-, and z-direction forces, even with one
or more faults in the AMB coils or SAs. This is implemented with a current distribution matrix (CDM)
that is updated in response to a coil integrity status change. The redundancy built into the actuator may
result in electrical instability and significant mutual inductance coupling. This is corrected by employ-
ing a decoupling choke that typically consists of a separate magnetic flux circuit device that contains
windings from each of the AMB poles. Filter stages are shown between the controller and CDM blocks
in Figure 55.1. This figure depicts a decentralized control with a single input–single output (SISO) archi-
tecture. Some machines may require a more complex multiple input–multiple output (MIMO) architec-
ture control for which many sensors may be utilized to determine the corrective forces at a given journal
location [8].
Table 55.1 lists some advantages and disadvantages of AMBs compared with conventional bearings.

55.2  Actuators
MBs operate in accordance with Ampere’s law, Gauss’s law, Faraday’s law, and constitutive relations
(Ohm’s law, B–H curves, etc.). Ampere’s law may be stated as follows:
Magnetic Bearings 55-3

 

∫ H ⋅ dl = NI (55.1)

where
H⃗ is the magnetic field intensity
dl ⃗ is a differential of path length along the closed magnetic circuit
N is the number of turns of the enclosed current I

One-dimensional (1D) magnetic circuit analysis makes the assumption that H⃗ is parallel to the path and
is constant in a given material.

55.2.1  Heteropolar Magnetic Bearings


Application to the upper pole flux path in Figure 55.2 yields

H S LS + H J LJ + 2H g L g = NI (55.2)

The magnetic intensity and flux density for a linear material are related by the constitutive law:

B = mH (55.3)

where μ is the magnetic permeability. Leakage is ignored so that flux is conserved around the loop
yielding:

f = BS AS = BJ AJ = B g Ag (55.4)

where A represents the cross-sectional area of the flux path. The flux is obtained by substituting
Equations 55.3 and 55.4 into Equation 55.1:

(RS + RJ + 2R g )f = NI (55.5)

LJ = flux path length in journal Stator

LS = flux path length in stator IB + iCq (t)

g = air gap with journal centered S N


Rotor
µ0 = permeability in free space

Fq q ω
– = relative permeability of journal material
µJ
Journal
– = relative permeability of stator material
µS

Ag = flux cross-sectional area in air gap


N S

(per pole)
IB – iCq (t)
AJ = flux cross-sectional area in journal

N turn coil

FIGURE 55.2  Typical magnetic flux loops (dashed lines) in a heteropolar magnetic bearing actuator.
55-4 Design for Lubrication and Tribology

where the reluctances are defined by

LS L Lg
RS = , RJ = J , R g = (55.6)
mS AS mJ AJ m0 Ag

The air gap flux density becomes

NI
Bg = dB (55.7)
Ag (RS + RJ + 2Rg )

where the flux derate factor dB accounts for leakage and fringing effects and is typically taken as about
0.85. The Maxwell stress tensor formula determines the force exerted by the magnetic field and has the
1D form:

B 2g Ag
F= (55.8)
2m0

From this formula, the magnetic pressure is limited to

2
FMAX BMAX
PMAX = = (55.9)
A 2m0

where
B has the unit of tesla (1 T = 1 N/(A·m))
A has the unit of m2
μ 0 = 4π × 10−7 N/A2

The best magnetic conducting medium (iron cobalt alloy) has a saturation flux density (SFD) of about
2.3 T, which from Equation 55.9 yields a maximum magnetic pressure of 2.1 MPa (294 lb/in.2). The total
force on the journal in Figure 55.2 is obtained from Equations 55.7 and 55.8 as

dB2 N 2 ⎛ (I B + iCq )2 (I B − iCq )2 ⎞


Fq = ⎜ − 2⎟
(55.10)
m0 Ag ⎝ (RS + RJ + 2( g − q)/m0 Ag ) (RS + RJ + 2( g + q)/m0 Ag ) ⎠
2

where
LJ is the flux path length in journal
LS is the flux path length in stator
g is the air gap with journal centered
μ 0 is the permeability in free space
μ̄J is the relative permeability of journal material
μ̄S is the relative permeability of stator material
Ag is flux cross-sectional area in air gap (per pole)
AJ is the flux cross-sectional area in journal
AS is the flux cross-sectional area in stator
IB is the bias current in pole pair q
Magnetic Bearings 55-5

iCq is the control current in pole pair q


q is the displacement of the shaft away from its centered position in the q-direction
N is the number of turns

Linearization of Equation 55.10 about the centered position q = 0 and iCq = 0 yields

Fq = K iqiCq + K Pq q (55.11)

where

2
d2 N 2 ⎛ 4 ∗ I B2 ⎞ ⎛d N⎞ ⎛ 8 ∗ I B2 ⎞
K = B ⎜
q
i 2⎟
and K Pq = ⎜ B ⎟ ⎜ 3⎟
(55.12)
m0 Ag ⎝ (RS + RJ + 2 g /m0 Ag ) ⎠ ⎝ m0 Ag ⎠ ⎝ (RS + RJ + 2 g /m0 Ag ) ⎠

which are the current and position stiffness coefficients for a single, opposing pole pair aligned along
some arbitrary direction q. The number of pole pairs range from 2 (x- and y-directions) to a large num-
ber. A reason for the large number of pole pairs is that as the individual pole width becomes smaller the
flux path width decreases, which permits the diameter of the bearing to decrease.

55.2.2  Homopolar Radial Magnetic Bearings


These actuators have lower hysteresis and eddy current losses as a result of preventing full flux rever-
sals by maintaining one polarity of air gap flux at each end of the bearing. Although the flux drops to
near zero between the poles, an electron on the journal only experiences one polarity of flux as the
shaft rotates. The homopolar radial MBs (HOMBs) generally employ PMs, which provide an additional
benefit of providing bias flux without I 2R ohmic losses. The PM flux flows axially along the stator and
rotor and must traverse the laminations perpendicular to their plane, which increase the material path
reluctance. The PMs exhibit the demagnetization behavior shown in Figure 55.3 and expressed by the
equation

H PM
BPM = BSAT + BSAT (55.13)
HC

Consider Ampere’s law equations (55.1 and 55.13) applied to the PM in Figure 55.3:

⎡ H ⎤
H PM LPM = LPM ⎢(BPM − BSAT ) C ⎥ (55.14)
⎣ BSAT ⎦

LPM
B
BSAT
1 1

H BSAT , HC , APM
–HC

FIGURE 55.3  Generic demagnetization curve.


55-6 Design for Lubrication and Tribology

Utilize conservation of flux and define the PM permeability as

BSAT
mPM = (55.15)
HC

to obtain

H PM LPM = RPMf1 − LPM H C (55.16)


where the PM reluctance and equivalent source term are defined by

LPM
RPM = VPM = (Ni)PM = LPM H C (55.17)
mPM APM

Figure 55.4 shows a type of HOMB with four poles in the active plane: PMs causing an axial flow of
bias flux, coils on each of the four active poles, a back iron ring that conducts the bias flux axially, and a
dead plane without coils that conducts bias flux radially.
Ampere’s law (Equations 55.1 and 55.2) possesses an analog with Kirchoff’s voltage law, as also does
flux conservation with Kirchoff’s current law. These observations justify the representation of the
magnetic circuit of the homopolar bearing in Figure 55.4 with the equivalent electric circuit shown in
Figure 55.5.
The reluctances of the material paths in the active poles are neglected in the model, which is consis-
tent with assumption that these poles are made from a high permeability alloy such as cobalt–iron. The
other reluctances are defined by
Pole air gaps:

g −x g +x g−y g+y
RX + = , RX − = , RY + = , RY − = (55.18)
m0 AP m0 AP m0 AP m0 AP

FIGURE 55.4  Solid model of a single active plane, four-pole HOMB.

RBI RPM VPM


+–

RX+ RY+ RX– RY–

Φ2 Φ3 Φ4 Φ1 RY– Φ5 RX– Φ6 RY+ Φ7 RX+


– –
VX+ +
– VY+ +
– VX– + VY– +

RR

FIGURE 55.5  1D magnetic circuit model of a single active plane, four-pole HOMB.
Magnetic Bearings 55-7

Axial paths (back iron, PM, rotor):

LBI L LPM
RBI = , RR = R , RPM = (55.19)
mBI ABI mR AR mPM APM

where x and y are the displacements of the journal with respect to the bearing. The equivalent sources
are defined by

VX + = VX − = NiCX , VY + = VY − = NiCY , VPM = LPM H C (55.20)


Solution of the magnetic circuit equations yields Bias flux density in pole air gaps with rotor centered
(x = y = 0):

dB LPM H C
BBIAS =
4 AP (( g /2 m0 AP ) + RBI + RPM + RR )

The linearized force expression for the centered rotor case is

⎧⎪FX ⎫⎪ ⎧⎪iCx ⎫⎪ ⎧⎪ x ⎫⎪
⎨ ⎬ = KI ⎨ ⎬ + KP ⎨ ⎬ (55.21)
⎩⎪FY ⎭⎪ ⎩⎪iCy ⎭⎪ ⎩⎪ y ⎭⎪

where the position and current stiffness are defined by

(dB LPM H C )2
KP = (55.22)
4 AP m0 g (( g /2m0 AP ) + RBI + RPM + RR )2

NdB2 LPM H C
KI = (55.23)
AP m0[( g /m0 AP ) + 2( g /m0 AP )(RBI + RPM + RR )]
2

55.2.3  Thrust (Axial) Magnetic Bearing Actuators


These actuators are available with both PM or electromagnet supplied bias flux and may be combined
into a single combined radial-thrust actuator sharing the same bias flux. Figure 55.6 shows an electro-
magnetic, bidirectional actuator consisting of a thrust disk that is secured to the rotor, a stator plate on
either side of the thrust disk, and coils embedded in each stator plate. Coil currents drive flux that cir-
culates axially through the air gaps. The relevant, axial-related control poles and the axial disturbances
generally occur at much lower frequencies than in the radial direction, which allows the stator and rotor
components of the bearing to have unlaminated (solid) forms.

55.2.4  Materials
The selection of materials for the MB actuators should consider the relative importance of cost,
weight, reliability, availability, and efficiency. In general, there exists a trade-off between magnetic
strength (as measured by permeability and SFD) and mechanical strength (as measured by yield
strength [YS] and toughness). Mechanical strength is improved by heat treating the alloy to smaller
grain sizes, since the YS is inversely proportional to the square root of the grain size. However,
the permeability will decrease, and the power loss will increase, with finer grain sizes. Mechanical
55-8 Design for Lubrication and Tribology

Stator plate

Shaft Thrust

disk

Coil

FIGURE 55.6  Basic components of an electromagnet thrust magnetic bearing actuator.

strength is especially important when selecting rotor components due to the large centrifugally
induced stresses. Cobalt–iron alloys (Hiperco, permendur) have high YS (up to 100 ksi, 690 MPa),
SFD (up to 2.3 T), and cost. Although silicon–iron alloys have lower YS (∼50 ksi, 345 MPa) and SFD
(∼1.5 T) than cobalt–iron alloys, they have higher resistivity (∼500 μΩ mm vs. 400 μΩ mm), which
reduces heating and losses. Quenched and tempered 4340 steel provides very high strength (YS >
200 ksi, 1380 MPa), but with a limited SFD of about 1 T. The lower resistivity of 4340 (∼300 μΩ mm)
relative to silicon–iron (∼500 μΩ mm) makes it more susceptible to eddy current–related heat gen-
eration and losses.

55.3  Sensors
Position sensors in an MB system are utilized to provide both control and monitoring functions. From
a control standpoint, the measured positions of the journals are compared with the target (reference)
values to form errors that the controller is designed to attenuate in a reliable manner. It is important to
recognize that although the aim of sensing is to measure the position of the shaft’s geometric centerline,
the translational motions of the shaft’s surface are what are actually measured. The two motions are
identical only for a perfectly circular shaft with homogeneous electrical or optical properties. The false
motion introduced by the spinning shaft’s out-of-roundness and electrical or optical inhomogeneity is
called runout. The different types of sensors have varying degrees of sensitivity to runout and electro-
magnetic interference—noise EMI, which is particularly prevalent in systems that employ switching
SAs. Important factors for selecting a MB sensor system include cost, bandwidth, resolution, and runout
and EMI susceptibility.

55.3.1  Inductance Sensors


This type of position sensor employs a miniature version of the C-core magnet circuit shown for the
actuator in Figure 55.7. However, in this application, the variation of the inductance of the C-core
with changes in shaft position coordinate q is utilized to produce a voltage that is proportional to q. By
Equation 55.7 and Figure 55.2, the inductance of the C-core is given by

N SfgS N S B gS AgS N S2 N S2 m0 AgS


LS = = = dBS ≈ dBS (55.24)
IS IS (RSS + RJS + (( g S − qS )/m0 AgS )) g S − qS

Magnetic Bearings 55-9

Light position detector

Drive Laser
circuit
Amplification
Lens
Lens

Displacement

FIGURE 55.7  Simplified diagram of an optical position sensor.

where the S subscript is utilized to distinguish the sensor C-core, which is typically much smaller than
the actuator C-core, and it is assumed that the shaft and sensor body reluctances may be neglected if
they are made from very high permeability material. Further assuming that qS ≪ gS yields

N S2 m0 AgS N S2 m0 AgS
LS ≈ L0 + ΔL where L0 = dBS and ΔL(t ) = dBS qS (t ) (55.25)
gS g S2

An inductive AC bridge circuit is utilized to produce a voltage that is amplitude modulated by ΔL(t),
which in turn is proportional to qS (t). The carrier frequency is typically near 20 kHz. Demodulation of
this AC signal produces a DC output signal vqS (t) that is proportional to qS (t):

vqS (t ) = kqSq(t ) (55.26)


Improvement in the signal-to-noise ratio is obtained by utilizing multiple sensor C-cores arranged
around the circumference of a sensor ring.

55.3.2  Eddy Current Sensors


The eddy current displacement coil consists of a flat coil whose longitudinal axis is pointed at the shaft
surface, a sensor drive circuit providing a radio frequency (RF) carrier current in the 0.5–5.0 MHz
range, and a conditioning circuit that delivers a direct current DC output voltage proportional to instan-
taneous changes in the shaft position. The magnetic field generated by the AC sets up eddy currents
in a “virtual coil” on the shaft surface that circulate in the opposite direction of the sensor coil cur-
rents, thereby reducing the flux in the sensor coil and in its corresponding inductance. This “trans-
former” effect changes the effective impedance at the terminals of the coil with the inductive component
decreasing and resistive component increasing with decreasing air gap between the sensor coil and the
shaft. Adding a capacitor to the sensor creates a resonance circuit that greatly amplifies the sensitivity of
the sensor coil’s impedance to changes in the gap between the coil and the shaft. A common approach
for converting the shaft’s motions to a linearly related output voltage is to drive the resonant circuit with
a fixed current source at a fixed frequency and then demodulate the amplitude or the phase of the volt-
age that appears across the sensor terminals. This typically requires an independent oscillator, phase
detector, low-pass filter, and voltage-conditioning circuitry.

55.3.3  Optical Sensors


Optical position sensors use a light source and detector system to measure the shaft’s lateral or axial dis-
placements. Laser triangulation position sensors are employed for this purpose and contain a solid-state
laser light source and a position-sensitive detector PSD or CMOS/CCD detector.
55-10 Design for Lubrication and Tribology

For the PSD detector, a PIN diode is exposed to a tiny spot of light that causes a change in local resis-
tance causing electron flow in four electrodes. The location of the light spot is computed using

I2 − I 4 I −I
x = kX , y = kY 1 3 (55.27)
I2 + I 4 I1 + I 3

where k X and kY are scale factors. An active pixel sensor (APS) is employed in the complementary metal–
oxide–semiconductor/charge-coupled device (CMOS/CCD) position detector. This is an image sensor
consisting of an integrated circuit containing an array of pixel sensors, each pixel containing a photo-
detector and an active amplifier. The CMOS APS is commonly used in cell phone cameras, web cam-
eras, and some digital single-lens reflex camera (DSLRs). Such an image sensor is produced by a CMOS
process (and is hence also known as a CMOS sensor) and has emerged as an alternative to CCD imager
sensors. For serial processing, the sensor is partitioned into individual pixels whose exposure value can
be read out sequentially. The position of the light spot can be computed with the methods of photogram-
metry directly from the brightness distribution. For parallel processing, matrix sensors with parallel
processing were developed. The density of light of each pixel is compared with a global threshold value
both line by line and in columns. From all columns and all lines the one element that is brighter than a
given threshold value is the average value of the coordinates computed of the light spot. A second type
of optical position sensor utilizes two fiber bundles arranged side by side. Light exits one side, reflects
off the target, and returns to the sensor through both sides. A ratiometric calculation of those two sig-
nals provides the distance measurement, which is independent of target reflectance variations, that is,
reflectance compensated.

55.3.4  Capacitive Sensors


The basic equations for a capacitor are
Capacitance:

Q dQ dV 1
C= , Current i = =C , Stored energy E = CV 2 (55.28)
V dt dt 2

The simplest capacitor consists of two parallel conductive plates separated by a dielectric with permit-
tivity ε (such as air). The capacitance for this device is given by

eA
C= (55.29)
g

where
A is the plate area
g is the gap between the plates

The sensor electrode and the target plane electrode (shaft surface) form the two parallel plates in a
capacitive displacement sensor. Equation 55.29 assumes a uniform electric field exists over the entire
plate area. In practice, fringing will occur around the perimeter of the electrodes so Equation 55.29
provides an approximation. Fringing is greatly reduced by installation of a guard ring as suggested by
Lord Kelvin. The guard ring is an annular-shaped electrode that surrounds the sensor electrode. The
two electrodes are separated by an insulator and are actively kept at the same electric potential. This
contains the sensor’s electric field and thereby improves linearity.
Magnetic Bearings 55-11

An AC bridge circuit is employed to measure the instantaneous change in the reactance

1
XC = (55.30)
jC

of the sensor, relative to its value at some reference Cref corresponding to a nominal value of the gap
between the target (shaft surface) and the sensor. Note from Equations 55.29 and 55.30, the reactance
is proportional to the gap and not its inverse. Therefore, the electronics output voltage (after demodula-
tion) will have the form:

V = GC (C −1 − Cref
−1
), V = SC Δg , Δg = g − g nom (55.31)

The carrier (excitation) frequency of the capacitive sensor driver electronics is a major factor in deter-
mining bandwidth. A meaningful change in output voltage requires several cycles of the changing elec-
tric field in the gap. Some vendors utilize 10–20 kHz; however, others use frequencies as high as 1 MHz
in order to increase bandwidth. The higher excitation frequency yields a bandwidth of 15 kHz as com-
pared to about 5 kHz for the lower carrier frequency devices.

55.3.5  Self-Sensing
This approach is similar to the inductance sensor approach described earlier that utilizes a dedicated
self-powered coil to measure the variation of inductance as the rotor-sensor gap changes. The impor-
tant difference is that for self-sensing, the coils of the MB actuator serve the dual purpose of actuation
and shaft position sensing [9], either in a parallel or in a multiplexing mode. The high-frequency ripple
current produced by a switching power amplifier can be utilized for the sensing purpose. Alternatively,
a high-frequency signal can be added to the input of a linear power amplifier for the same purpose. The
corresponding component of the current waveform is extracted and the gap (displacement) is inferred
from it. A possible drawback of self-sensing is that the ratio of the material path reluctances to the air
gap reluctances increases and may exceed unity at the high frequencies utilized for sensing, which may
lead to diminished sensitivity, bandwidth, and linearity relative to dedicated position sensors. Still self-
sensing may be preferred due to reduced costs and cabling.

55.4  Control
The central objective of MB controllers is to maintain the positions of the journals and thrust collar
at some predetermined target positions relative to the radial and thrust bearing stators, respectively.
This objective must be met under all operating conditions and in the presence of static and dynamic
disturbances such as side loads, thrust load, imbalance, runout, electrical noise EMI, etc. The errors
in the control objective are obtained from the relative position sensors that provide the instantaneous
positions of the journals and thrust collar. The error is minimized with the MB forces that pull the shaft
toward the target positions. Control is typically implemented in one of three forms: analog circuitry,
digital (digital signal processor [DSP] or field programmable gate array [FPGA]), or a hybrid combina-
tion of analog and digital.
The controller’s architecture is typically divided into two levels. The hierarchal level varies control
gains, filter settings, feed-forward signals, effective stiffness and damping, journal position, etc. as the
operating conditions of the machine, such as speed, pressure ratio, temperature, etc., varies. This pro-
vides a distinct “adaptive” advantage of MBs over conventional passive bearings. The core level of the
controller operates within the hierarchal level and provides the function of minimizing the shaft position
55-12 Design for Lubrication and Tribology

errors via a classical control stage, or plant-based control law approach. The latter relies heavily on an
accurate model of the plant, which may be updated using neural networks, and typically couples all
controller outputs with all controller inputs (MIMO control). Some examples include H∞, sliding mode,
LQR, etc. The classical approach typically cascades control stages that may include, low-pass or notch
filters, lead and lag, differentiator, proportional, integral, etc. These may be utilized in a SISO strategy
that develops a control force in one axis of control based solely on a single position error measurement
(typically adjacent to and in the direction of the actuator). The classical control stages may also be used in
a MIMO approach, which is the case for strongly gyroscopic systems, characterized by natural frequen-
cies that vary significantly with speed [8]. It is well known that stability is enhanced by increasing phase
lead and decreasing amplitude in the feedback loop. The phase angle increase may be accomplished with
a differentiator or lead stage. The differentiator stage provides much phase lead (90°) but also has an
infinite gain as frequency goes to infinity. This characteristic of a differentiator decreases gain stability
margin and amplifies high-frequency EMI and runout. Therefore, a phase lead compensation stage may
be utilized for improving phase stability margin. A form of the transfer function for a phase lead stage is

p s+z
TFLEAD (s) = * (55.32)
z s+ p

The zero and infinite frequency gains of this stage are 1 and GL∞ = p/z. Denote the desired peak phase
lead to be ϕpeak and the desired frequency (in Hz) at which it occurs f peak. Then ϕpeak, f peak, GL∞, z, and p are
related by the equations

1 + sin(fpeak ) 2p
GL∞ = , z= f peak , p = z * GL∞ (55.33)
1 − sin(fpeak ) GL ¶

Figure 55.8 clearly illustrates the trade between increasing phase lead and obtaining excessive feedback
gain. The ratio GL∞ of the infinite frequency gain to the zero frequency gain is shown to significantly
increase as the desired peak phase lead ϕpeak increases, as is illustrated in Figure 55.8a and b. The peak
phase lead frequency is typically set close to the frequency of the pole to be stabilized. These poles may
shift with spin speed in rotordynamic systems due to gyroscopic moments. The peak frequency can
track a pole as speed varies, which is one aspect of the patent in Reference 10.

8
14
6

12 4

2
10
0
8 0 10 20 30 40 50 60 70 80 90 100

50
6
40
4 30
20
2
10

0 0
0 10 20 30 40 50 60 0 10 20 30 40 50 60 70 80 90 100
(a) (b)

FIGURE 55.8  (a) Phase lead stage GL∞ vs. ϕpeak (in degrees) and (b) |TFLEAD | and 〈TFLEAD (in degrees) vs. frequency
(in Hz) for f peak = 20 Hz and ϕpeak = 10°, 20°, 30°, 40°, 50°.
Magnetic Bearings 55-13

A phase lag stage is very useful for providing low-frequency (DC) gain to reject errors due to static,
typically weight loading, acting on the MB. A form of the phase lag stage is

s+z z
TFLAG (s) = with TFLAG (s=0) = >1 (55.34)
s+p p

The high gain and the phase lag of this stage are confined to very low frequencies, which reduce their
phase and gain penalties on high-frequency poles. Unlike an integrator, the lag stage’s gain at zero (DC)
frequency is bounded, which eliminates the need for anti-windup compensation to prevent DC insta-
bility (saturation). Figure 55.9 shows the gain and phase lag characteristics of a lag compensation stage
with z = 2 and z/p = 1, 2, …, 10.
Other compensation stages such as notch, low pass, differentiator, proportional are typically cascaded
or summed to form feedback control paths that provide stable levitation while rejecting imbalance, sup-
port motions, noise, runout, etc. Feed-forward control for imbalance rejection (active balancing) is also
employed for some applications [11]. Much work has also been performed on controllers that incorpo-
rate a model of the plant, such as H infinity, LQR, etc. [12]. These approaches often use adaptive means
for model identification in order to improve the effectiveness of the plant-based control law.
Compensation stages have a continuous time form if they are realized with analog circuits and have
a “difference equation” form if they are realized with a digital controller. The difference equation for the
controller output u with the controller input (error) e is

N M
un = − ∑ ak un − k + ∑b e k n−k (55.35)
k =1 k =0

The “z” domain transfer function corresponding to Equation 55.35 is given by

U (z ) b0 + b1z −1 +  + bn z − N
= (55.36)
E(z ) 1 + a1z −1 +  + am z − M

10
8

6
4
2
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2

–20

–40

–60
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2

FIGURE 55.9  Phase lag stage amplitude and phase angle (degrees) vs. frequency (Hz).
55-14 Design for Lubrication and Tribology

The coefficients in Equation 55.35 may be obtained from a continuous “s” domain form of the control
stages, such as is given in Equation 55.32 for the lead stage, or may be obtained more directly from a
digital control model of the system. The former approach may utilize a difference equation, for example,
Euler method, approximation for derivatives in the control stages. An alternative approach is to find an
approximate “z” domain transfer function corresponding with the desired “s” domain transfer function
and then to utilize the relations in Equations 55.35 and 55.36. An approximation for this approach is the
use of the Tustin or “bilinear” formula

2 ⎛ 1 − z −1 ⎞
s= ⎜ ⎟ (55.37)
T ⎝ 1 + z −1 ⎠

for all occurrences of “s” in the “s” domain controller transfer function. This typically requires an inordi-
nate amount of algebraic manipulation, so use of a computer-aided math package, such as MATLAB ®’s
c2dm command, is recommended.

55.5  Power Electronics


The force produced by a MB actuator results from the flux generated in the air gaps as discussed in
Section 55.2. This flux results from either PMs or current-carrying coils inserted on stator poles of the
AMB. The latter source of flux requires an efficient means to drive currents by application of voltage
across the coils. These voltages are obtained from the fixed bus voltage of a power supply. Although
this could be accomplished with a linear power amplifier with a feedback-controlled variable resistance
transistor, the continuous ohmic loss may be prohibitive. The power sources for most AMBs are pulse
width modulated (PWM), SAs, which apply the DC power supply voltage in a pulsed manner utiliz-
ing high-speed switches (metal–oxide–semiconductor field-effect transistor [MOSFET], insulated gate
bipolar transistor [IGBT], etc.). When a switch is closed its voltage drop is ideally zero, and when it is
open its current is ideally zero. Therefore the power dissipated in the switch, being the product of volt-
age and current, is ideally zero both in its open and closed states. Power is dissipated while switching
between the open and closed states; however, this occurs over a very short time. Typical efficiencies for
PWM-SAs are generally greater than 90%, while linear amplifier efficiencies may be as low as 30%. This
is a form of a voltage source inverter that converts energy from a DC form (DC power supply voltage) to
an AC form (PWM-SA output voltage). At the heart of the SA is a current control feedback loop where
the instantaneous commanded current is compared with the measured current in the MB coil(s). As
in any servo device this allows the controlled quantity (current) to change much faster than its open
loop time constant (L/R), which is illustrated in the simplified models of Figure 55.10. Equation 55.21
shows that the feedback loop on current also allows the control force to be able to change faster since it
is ­proportional to the control current.

R V iref + e V 1 i
K
– Ls + R
V +
– L
i
γ
i
i L i L
τ= τ=
(a) V R (b) V iref R + Kγ

FIGURE 55.10  Simplified model of a MB coil as an open loop system and with current feedback control.
Magnetic Bearings 55-15

High side High side


(left) (right) High High Low Low Quadrant

Left Right Left Right

Low side Low side


on off off on 1
(left) (right)
off on on off 2
Power supply voltage –
on on off off 3

off off on on 4

FIGURE 55.11  H bridge circuit and associated quadrants of operation.

The PWM-SA applies a bidirectional voltage across the AMB utilizing an electronic circuit called
an H bridge as illustrated in Figure 55.11. A three-level PWM algorithm varies the length of time that
the +supply voltage, −supply voltage, and 0 voltage are turned on or off at a very high rate (typically
20–50 kHz). The average voltage or current supplied to the load (MB coil) is controlled by changing
the duty cycle of the rectangular voltage pulses. A 100% duty cycle corresponds to the full power sup-
ply voltage being always applied across the coil. The current flows from left to right through the coil
if the high left and low right switches are closed in Figure 55.11. The current flows from right to left
through the coil if the high right and low left switches are closed. The switching power amplifier yields
the desired current waveform plus a high-frequency ripple current at the switching frequency. The ripple
current can be reduced by installing an edge filter at the output of the power amplifier.
The duty cycle is varied by one of a number of PWM approaches. A common approach utilizes a tri-
angular carrier waveform (TCW) at the switching frequency [13]. An electronic comparator is employed
to indicate when an internal voltage representing the desired continuous output voltage of the PWM-SA
is above or below the TCW voltage. The comparator output is a switching function, being either a high
or a low indicating for instance when a two-level PWM-SA output voltage should be set at either the
+ power supply voltage or at the −power supply voltage.
Assume that a high permeability material is utilized so that the material path reluctances may be
ignored in the C-core arrangement shown in Figure 55.7. Then the inductive load across the amplifiers
for either the top or the bottom coil is expressed as

N 2 AdB m0
L= (55.38)
2g

The force produced by either C-core has a maximum allowable slew rate given by

dF dF di
= (55.39)
dt MAX di dt MAX

Substitution of the force expressions from Equations 55.7 and 55.8 and use of Equation 55.38 yields

dF ⎛ VI ⎞
= dB ⎜ ⎟ (55.40)
dt MAX ⎝ g ⎠
MAX
55-16 Design for Lubrication and Tribology

where

di
V =L = inductive voltage drop across the coil (55.41)
dt

This illustrates that the maximum slew rate of the force produced by the C-core is determined solely by
the available power from the power supply and the MBs air gap. For a harmonic force the slew rate is
directly proportional to the frequency, so that the aforementioned limitation on the force slew rate also
becomes a limitation on the bandwidth of the actuator.

55.6  Finite Element Magnetic Field Simulation


1D magnetic circuit modeling may not provide sufficient accuracy for design of some MB actuators;
therefore, a 2D or 3D finite element method (FEM) simulation is required as shown in Figure 55.12.
Commercial codes typically utilize a magnetic scalar potential, node-based vector potential, or edge-
based vector potential method for this purpose. The scalar potential method can be easily used with fast
and accurate results for static analysis. The following discussion provides some modeling suggestions
for building an accurate model for MB, 3D magnetic field simulation.
A set of simulations are performed for the static load case with only bias current or PM sources
and the journal positioned at a possible operating location. A second set of simulations are performed
including perturbations of current and position and the results subtracted from the first to determine
the position stiffness (change in force per change in journal position) and the current stiffness (change
in force per change in control current). Nonlinear static load FEM is utilized to determine the load
capacity based on the nonlinear flux density (B) vs. magnetic field strength (H), BH curve.
Eddy current effects on the dynamic stiffness (transfer function, frequency response) may need to
be considered for high-speed applications or when some of the magnetic circuit is unlaminated. Eddy
currents cause the magnetic field to become more localized nearer to the surface of the flux path as the
field frequency increases. This “skin depth” effect weakens the MB’s load capacity at higher frequencies.
The skin depth is defined by

1
d= (55.42)
pf ms

where
f is the excitation frequency in Hz
μ is the absolute permeability in N/A 2
σ is the conductivity in Ω−1m−1

Radial bearing Combo bearing

Backiron

Magnet

Stator

Rotor

FIGURE 55.12  Three-dimensional magnetic field model of MB actuators.


Magnetic Bearings 55-17

From one [14] to three [15] layers of elements are recommended to be placed within the skin depth for
FEM harmonic analysis simulation. Reference 14 shows that the mesh density can be much coarser out-
side of the skin depth. Linear harmonic analysis (LHA) is unaffected by the bias (DC) flux caused by bias
currents or PMs as long as the properties used in the LHA correspond to the static flux state. A nonlin-
ear transient should be performed out to harmonic steady state if the flux variations cause a significant
change in the permeability. This tends to occur when simulating high-frequency sources with a small
skin depth that causes the flux density to approach saturation levels. The transient analysis may require
considerable execution time and modeling effort, relative to a LHA. Both node-based vector potential
method and edge-based vector potential method can be used in harmonic and transient analyses. The
edge-based method is recommended over the node-based vector potential method for models where
materials have different permeability [16]. Motion-induced eddy currents are important for high-speed
operation characterized by the Peclet number:

g = msVh (55.43)

where
V is the velocity of the motion
h is the element size in the direction of the motion

Upwind-based finite elements are sometimes utilized for Peclet numbers exceeding a value of 2 in order
to insure the stability of the FEM numerical solution.
The FEM model should include an air layer surrounding the metal magnetic path in order to pre-
dict the effects of fringing and leakage. Normal magnetic and tangential magnetic conditions need to
be correctly defined on boundaries and symmetric planes. In some software, tangential electric and
normal electric conditions can also be specified and serves the equivalent purpose. For example, the
surrounding air’s outer surfaces can be specified as a tangential magnetic condition, assuming enough
air was included and no magnetic flux escapes from these boundaries. The boundary conditions for the
symmetric planes can also be specified as tangential since no magnetic flux will enter the other side
due to symmetry. The boundary conditions for the antisymmetric planes need be specified as normal
magnetic conditions.
Hexahedral-shaped elements are always preferred over the tetrahedral-shaped elements since they
will give better results with the same number of elements. The hex mesh may require a somewhat
smooth model shape generally utilizing geometry simplifications. Although voltage-fed or current-fed
solid conductors can be modeled in FEM analysis, elements with current density defined and current
primitive elements (e.g., SOURC36 in ANSYS and Parameterized Conductors in Opera-3d) are most
commonly used to define the excitation coils.
Individual lamination sheets can be modeled, however; this requires extensive work and computer
resources since the aspect ratio requirements of FEM programs will generate a large number of elements
due to the shape of the thin laminate. Another approach is to treat the laminate stack as a material with
anisotropic relative permeability. The relative permeability in directions normal and tangential to the
laminate plane is given by [17,18]

mr llam
(mlam )n = , (mlam )t = (1 − f lam ) + f lam mr , f lam = (55.44)
(1 − f lam )mr + f lam ltotal

Hollaus and Biro [19] suggested and verified that using an anisotropic conductivity with zero or very
low value in the direction normal to the 3D lamination model will produce a good approximation of
the real case. Silva et al. [20] chose a value of zero in the normal direction. Taking account of the eddy
55-18 Design for Lubrication and Tribology

current loops within layers of lamination, Xu et al. [21] studied the 2D case with magnetic flux parallel
to the lamination and advocated using

s
sn = (55.45)
n2

where n is the total lamination count. Wang et al. [22] suggested that a conductivity in the form of

2
s ⎛t ⎞
sn = (55.46)
f lam ⎜⎝ w ⎟⎠

should be used in the direction normal to the plane of the laminates where t is the lamination thickness
and w is the lamination width.

55.7  Auxiliary/Backup/Catcher Bearings


Catcher bearings (CBs), or often referred to as auxiliary or backup bearings, are employed to protect
the AMB and other components in a machine in the event that power to the AMB’s fail or an AMB
becomes overloaded. Since Ishii and Kirk’s [23] initial paper, many researchers have strived to develop
accurate models of CBs and to optimize their performance. In general, plain sleeve, rolling element,
and planetary type bearings are used as CBs. The plain sleeve bearing is typically made of a bronze,
Babbitt lined, or graphalloy material. The soft material minimizes the risk of scoring or damaging
a shaft during a rotor drop event. The bronze material is more suitable for heavier shafts and higher
speeds than the Babbitt. The graphalloy material provides relatively lower friction coefficients than
the bronze or Babbitt and is suitable for higher temperatures. Rolling element bearings (REBs) are the
most common type of backup bearing. The most common REB–CB is a dual full complement angular
contact ball bearing, which supports both radial and axial loads, and is mounted in a face-to-face pre-
loaded configuration in order to prevent windage-induced motion under standby conditions [24]. The
use of a planetary auxiliary bearing may be considered when the DN of the CB becomes a concern, as
in large diameter, high-speed shaft applications [24]. The zero clearance auxiliary bearing (ZCAB) is a
specialized design of the planetary auxiliary bearing. Reference 25 provides a detailed design concept
for the ZCAB. Table 55.2 shows some advantages and disadvantages for each of the CB types. The CB
may be mounted within a damping device such as a preloaded ribbon damper [26] or squeeze film
damper (SFD) [27]. The preloaded ribbon damper generally is made of carbon steel or 17-7 stainless
steel. It provides a sufficient stiffness and damping to absorb mild shock during initial rotor drop. The
SFD is suitable for higher shock. The dynamic characteristics of rotor drop depend on the stiffness and
damping of the support; thus, a careful design is required. The rotor drop simulation model shown in
Figure 55.13 consists of a finite element—Timoshenko beam, CB, support, and AMB (for considering
side loads from the PMs) component models. The CB models are classified into two categories: (1) sim-
plified linear bearing model, which consists of linear springs and dampers, and (2) a detailed nonlinear
REB bearing model, which considers internal thermal growth and forces. The support is typically mod-
eled with linear springs and dampers except when a SFD is included. Details of these model types may
be viewed in References 23, 26, and 27. Factors that affect the vibratory loads and resulting life (number
of drops) of a CB include: (1) surface contact friction value, (2) imbalance level, (3) amount of clearance
(gap) between the rotor and the inner race, (4) stiffness and damping of the support, and (5) the bear-
ing’s internal preload and side loads.
Simulation and test results show that a destructive forward or backward precession whirl motion
may develop after the drop of the spinning rotor on the CB. Forward whirl typically occurs with large
imbalance and low coefficient of friction and backward whirl motion may occur for smaller levels
Magnetic Bearings 55-19

TABLE 55.2  Advantages and Disadvantages of Auxiliary Bearing Types


Type Advantages Disadvantages
Plain auxiliary • Low cost
bearing • Passive, no moving parts in bearing
• Reduced potential for deterioration in • Higher friction coefficients
standby mode
• Condition, wear may be assessed by • Higher heat generation during rundown
measuring clearance with AMBs
• Low cost • Potential for bearing/cage damage during
• Low friction coefficients acceleration
Rolling element • Low heat generation during rundown • Potential for deterioration in standby
auxiliary bearing mode
• Minimum volume with combined • Windage-induced rotation must be
radial/thrust bearing prevented in standby mode
• Greater complexity and cost
• Reduced DN for given rotor diameter • Contamination must be avoided
and speed
Planetary auxiliary • Low friction coefficients • Windage-induced rotation must be
bearings prevented in standby mode
• Low heat generation during rundown • Potential damage during acceleration
(reduced relative to rolling element
auxiliary bearing)
Source: Penfield, S.R. and Rodwell, E., Auxiliary bearing design considerations for gas-cooled reactors, IAEA
Technical Committee Meeting, Palo Alto, CA, 2000.

Rotor model
0.1
0.08
0.06
Shaft radius (m)

0.04
CB CB 0.02
0
A A –0.02
M M –0.04
B B –0.06
–0.08
–0.1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
(a) (b) Axial distance (m)

FIGURE 55.13  Rotor drop model: (a) side view and (b) FEM rotor drop model (X: AMB, Δ: CB).

of imbalance and higher friction. The bearing life is rapidly reduced by these whirl motions, which
create large fatigue stresses or wear in the CB. Decreasing the clearance between the rotor and CB
mitigates violent whirl and conversely increasing the clearance exacerbates the vibration and dam-
age. Softening and integrating a damper into the CB support structure will assist in mitigating whirl;
however, a careful design is required to select optimal values. Increasing the internal bearing preload
may degrade life due to larger mean stresses in components of the REB and due to the increased likeli-
hood of whirl.
Whirl is also mitigated by applying a side load at the bearings either from electromagnetic force (if the
bearing is still powered) or by the natural forces of attraction in a PM AMB.
Figure 55.14 shows simulation results for a horizontal imbalanced rotor undergoing severe reverse
whirl. The nominal values of the design parameters are initial rotor drop speed: 20,000 rpm, clearance:
300 μm, friction coefficients (static, dynamic): μs = 0.2 and μd = 0.4, side load: 0 N, support stiffness and
damping: 100,000 N/m and 5,000 Ns/m, and imbalance: 0.1 kg·mm.
55-20 Design for Lubrication and Tribology

× 103 12,000
1.5
10,000
1
8,000
0.5

Force (N)
z(m)

0 6,000

–0.5 4,000
–1
2,000
–1.5
0
–1.5 –1 –0.5 0 0.5 1 1.5 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
(a) y(m) × 103 (b) Time (s)

FIGURE 55.14  Nominal system simulation results: (a) orbit plot and (b) contact force vs. time.

× 10–4
1800
8
1600
6
1400
4
1200
2
Force (N)

1000
z(m)

0
800
–2
600
–4
400
–6
200
–8
0
–8 –6 –4 –2 0 2 4 6 8 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
(a) y(m) × 10–4 (b) Time (s)

FIGURE 55.15  Orbit and force responses with friction coefficients μs = 0.1 and μd = 0.2: (a) Orbit plot and (b)
contact force vs. time.

× 10–4
2000
8
1800
6
1600
4 1400
2 1200
Force (N)
z(m)

0 1000
–2 800
–4 600

–6 400
200
–8
0
–8 –6 –4 –2 0 2 4 6 8 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
(a) y(m) × 10–4 (b) Time (s)

FIGURE 55.16  Effect of side load, side load = 500 N: (a) Orbit plot and (b) contact force vs. time.
Magnetic Bearings 55-21

× 10–3
14,000
2.5

2 12,000

1.5
10,000
1

0.5 8,000

Force (N)
z(m)

0
6,000
–0.5

–1 4,000

–1.5
2,000
–2

–2.5 0
–2 –1 0 1 2 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
(a) y(m) × 10–3 (b) Time (s)

FIGURE 55.17  Forward whirl motion of rotor: μs = 0.02 and μd = 0.01, imbalance: 0.01 kg·m. (a) Orbit plot and (b)
contact force vs. time.

Figure 55.15 shows that the whirl is eliminated by lowering the coefficients of friction by 50%.
Figure 55.16 shows similar results with the addition of a side load at each bearing. Forward whirl motion
shown in Figure 55.17 occurs at lower friction coefficient and large imbalance.

Acknowledgments
The authors acknowledge the excellent work of Dr. Andrew Kenny who performed R&D in the area of
magnetic field simulation and actuator design and Mr. Randy Tucker who designed and built critical
MB-related test apparatuses for the Vibration Control and Electromechanics lab at Texas A&M and thus
indirectly contributed to this manuscript.

References
1. Beams, J., Production and use of high centrifugal fields, Science, 120 (1954), 619–625.
2. Beams, J., Magnetic bearings, Paper 810A, Automotive Engineering Conference, January, Detroit, MI,
SAE, 1964.
3. Allaire, P. E., Maslen, E. et al., Design of a magnetic bearing-supported prototype centrifugal artificial
heart pump, STLE Tribology Transactions, 39 (1996), 663–669.
4. Provenza, A., Montague, G., Jansen, M., Palazzolo, A., and Jansen, R., High temperature character-
ization of a radial magnetic bearing for turbomachinery, Journal of Engineering for Gas Turbines and
Power, 127(2) (2005), 437–444.
5. Park, J. and Palazzolo, A., Magnetically suspended VSCMGs for simultaneous attitude control
and power transfer IPAC service, Journal of Dynamic Systems, Measurement, and Control, 132(5)
(2010).
6. Li, M. H., Palazzolo, A., Kenny, A., Provenza, A., Beach, R., and Kascak, A., Fault tolerant homopolar
magnetic bearings, IEEE Transactions on Magnetics, 40(5) (2004), 3308–3318.
7. Fault tolerant homopolar magnetic bearings, US Patent 7,429,811, Date of Patent: September 30,
2008, Co-inventors: Palazzolo, A., Li, M.-H., Kenny, A., Beach, R., Kascak, A., Provenza, A. J., and
Tucker, R. P.
8. Lei, S. and Palazzolo, A., Control of flexible rotor systems with active magnetic bearings, Journal of
Sound and Vibrations, 314 (2008), 19–38.
55-22 Design for Lubrication and Tribology

9. Schweitzer, G. and Maslen, E. (eds.), Magnetic Bearings, Theory, Design and Application to Rotating
Machinery, Springer, Zurich, Switzerland, 2009.
10. System and method for controlling suspension using a magnetic field, U.S. Patent 6,323,614B1, Date
of Patent: November 27, 2001, Inventors: Palazzolo, A., Li, M., Na, U. J., and Thomas, E.
11. Knospe, C. R., Humphris, R. R., Maslen, E. H., and Allaire, P. E., Active balancing of a high speed
rotor in magnetic bearings, Proceedings of the International Conference on Rotating Machinery
Dynamics, Springer Verlag, Venice, Italy, April 28–30, 1992.
12. Lin, Z., Zhang, H., Allaire, P. E., Bartlett, R. O., and Zorzi, E. S., Explicit construction of an H-infinity
controller for a rigid rotor on magnetic bearings, Proceedings of 35th Conference on Information
Sciences and Systems, The Johns Hopkins University, Baltimore, MD, pp. 708–713, March 2001.
13. Krein, P. T., Elements of Power Electronics, Oxford University Press, Oxford, U.K., 1998.
14. Kim, C., Magnetic bearing eddy current effects on rotordynamic system response, PhD dissertation,
May 1995.
15. Opera 3d: Training course notes, Vector Fields, May 1998.
16. ANSYS 12.1, Theory Reference, 2010.
17. Barton, M. L., Loss calculation in laminated steel utilizing anisotropic magnetic permeability, IEEE
Transactions on Power Apparatus and Systems, 99(3) (1980), 1280–1287.
18. Kenny, A., Nonlinear electromagnetic effects on magnetic bearing performance and power loss, PhD
dissertation, Texas A&M University, Houston, TX. December 2001.
19. Hollaus, K. and Biro, O., Estimation of 3-D eddy currents in conducting laminations by an anisotro-
pic conductivity and a 1-D analytical model, COMPEL—The International Journal for Computation
and Mathematics in Electrical and Electronic Engineering, 18(3) (1999), 494–503.
20. Silva, V. C., Meunier, G., and Foggia, A., A 3-D finite-element computation of eddy currents and
losses in laminated iron cores allowing for electric and magnetic-anisotropy, IEEE Transactions on
Magnetics, 31(3) (1995), 2139–2141.
21. Xu, J., Lakhsasi, A., Yao, Z., and Rajagopalan, V., A practical modeling method for eddy-current
losses computation in laminated magnetic cores, The Thirty-First IAS Annual Meeting, IAS’96,
October 6–10, San Diego, CA, Vol. 3, p. 1532.
22. Wang, J., Lin, H. Y., Huang, Y. K., and Sun, X. K., A new formulation of anisotropic equivalent conduc-
tivity in laminations, 14th Biennial IEEE Conference on Electromagnetic Field Computation (CEFC),
May 9–12, Miami, FL, p. 1, 2010.
23. Ishii, T. and Kirk, R. G., Transient response technique applied to active magnetic bearing machinery
during rotor drop, ASME Journal of Vibration and Acoustics, 118 (1996), 154–163.
24. Penfield, S. R. and Rodwell, E., Auxiliary bearing design considerations for gas-cooled reactors, IAEA
Technical Committee Meeting, Palo Alto, CA, 2000.
25. Chen, H. M., Walton, J., and Heshmat, H., Test of a zero clearance auxiliary bearing, Proceedings of
MAG’97 industrial Conference and Exhibition on Magnetic Bearings, August, MITI, New York, 1997.
26. Ransom, D., Masala, A., Moore J., Vannini, G., and Camatti, M., Numerical and experimental simu-
lation of a vertical high speed motor-compressor rotor drop onto catcher bearings, Journal of System
Design and Dynamics, 3(4) (2009), 596–606.
27. Sun, G. and Palazzolo, A. B., Rotor drop and following thermal growth simulations using detailed
bearing and damper models, Journal of Sound and Vibration, 289 (2006), 334–359.
56
Face Seals
56.1 Introduction.....................................................................................56-1
56.2 Face Seals...........................................................................................56-1
Face Seal Design  •  Seal Face Materials  •  Seal Arrangements  •  Seal
Support System  •  Seal Design Parameters  •  Other Face Seal
Tom W. Lai Performance Limits
John Crane Inc. References...................................................................................................56-17

56.1  Introduction
Sealing devices are used to stop the fluid from leaking. Figure 56.1 shows the main categories of the
sealing devices. Static seals such as O-rings, gaskets, and packings of various forms and materials are
used when the contacting component surfaces have no relative motion to each other. Readers can find
comprehensive description and discussion in Reference 1. If there is relative motion, the seal is called a
dynamic seal. Dynamic seals can be grouped into two categories, based on the mode of relative motion:
(1) rotating seals and (2) reciprocating seals. The major classes of rotating seal designs are face seals, lip
seals, circumferential seals, and labyrinth seals. Packings were also used for rotating service in earlier
days, but are being replaced by face seals for lower leakage and less power consumption. Rotating seals
are the topics of discussion in this chapter.

56.2  Face Seals


56.2.1  Face Seal Design
The face seal, also commonly called a mechanical end-face seal or a mechanical seal, has been used
extensively in rotating equipment to seal between the stationary housing and the rotating shaft.
Figure 56.2 shows one common configuration for a face seal. A face seal consists of the following major
components: (1) primary ring, which floats axially, (2) mating ring (also called seat), which is fixed to
either the housing or the shaft, (3) secondary seals (O-rings), which fill the gaps between the seal ring
(primary ring or mating ring) and the neighboring components, and (4) spring device, which forces the
primary ring to move forward to maintain a close contact with the mating ring. Other components are
also needed to support the functioning of the face seals. An installation spacer may be used to position
the seal assembly to achieve the target initial face loading during the installation and should be removed
afterward. The drive collar locks and drives the rotation of the rotating seal assembly. A safety bushing
restricts fluid leakage in case the face seal fails.
The sealing faces of the primary ring and mating ring are each lapped and polished to specific
surface flatness and finish conditions. Face flatness is checked with the optical flat or surface pro-
filometer to be typically within two to four helium light bands across the seal face, depending on the
size and application requirements. One helium light band is equal to 11.6 μin. (0.295 μm). When these
nominally flat faces are forced into contact, the interfacial gap is the result of asperity contact and fluid

56-1
56-2 Design for Lubrication and Tribology

Sealing devices

Static Dynamic

Gasket Reciprocating Rotating

O-ring Packing

Labyrinth Circumferential Lip Face

FIGURE 56.1  Main categories of sealing devices.

Gland

Housing

Static
Retaining clip O-ring
Spring disk
Installation
spacer

Spring Primary ring Mating ring Drive


Retainer Safety
Dynamic bushing collar
O-ring Rotating shaft

FIGURE 56.2  Face seal design.

interaction, which is strongly influenced by the surface roughness of both surfaces. The prevailing
gaps between two contacting seal faces are in the order of two to three times the composite surface
roughness (square root of the sum of squares of rms roughness from both surfaces). Typical surface
finishes for various seal face materials are shown in Table 56.1. A polished surface finish is generally
used except for hard face running against hard face in which case one face is matte finished by special
lapping to reduce the severity of the localized contact. This tiny gap, between the rotating and station-
ary end faces, is the primary leakage path of the face seal. This is the most critical leakage path and
it is difficult to control due to the dynamic running conditions. Too small a running gap will lead to
severe face contact, excessive frictional heating, and face damage. Too big a running gap will lead to
high leakage. There is also a secondary leakage path around the dynamic O-ring that seals between the
primary ring bore diameter and the retainer balance diameter. This O-ring is called dynamic because
it has to slide and/or deform to accommodate the axial motion of the primary ring. Seal designs using
dynamic O-rings are called pusher seals. Flexible bellows may be used as the secondary seal instead
of an O-ring. The tail end of the bellows remains fixed while the front end moves with the primary
ring. Face seals using bellows are called non-pusher seals. Non-pusher seal designs minimize the seal
hang-up risk but are more involved in design and manufacturing. Special attention is required in
analyzing the hydraulic force transmission through the bellows and the resultant seal face distortions.
Face Seals 56-3

TABLE 56.1  Typical Surface Finishes for Seal


Face Materials

Surface RMS Roughness


Finish μin. μm
Carbon graphite Polish 3.6 0.09
Silicon carbide Polish 2.4 0.06
Matte 8.4 0.21
Tungsten carbide Polish 2.0 0.05
Matte 8.4 0.21
Alumina Polish 4.0 0.10
Dairy bronze Polish 4.8 0.12
Ni-resist Polish 4.8 0.12

In comparison to seal face leakage, leakage through the static seal (O-ring) leakage paths tends to be
small and is usually negligible.

56.2.2  Seal Face Materials


Typical physical and mechanical properties of common seal face materials are summarized in
Table 56.2. The most popular seal face materials are carbon graphite for the primary ring and silicon
carbide (SiC) for the mating ring. Both have low thermal expansion coefficients and thus thermal
distortions of the seal faces are minimized. Carbon graphite is preferred because of its lubricity and
its ability to wear to conform to the mating face. Carbon graphite lubricity does require the presence
of water or hydrocarbon vapor. To manufacture carbon graphite, compression molded tubes con-
taining a mixture of petroleum coke, carbon black, graphite, binder, and additives are baked at high
temperature over 2000°F (1000°C) to pyrolyze the binder and are further heat treated at about 5000°F
(2800°C) for graphitizing. The tube is then machined to the primary ring shape and impregnated with
thermoplastic resin to fill the through porosity. To increase the material strength, Young’s modulus,
and thermal conductivity, metal such as antimony is sometimes used instead of resin. For contacting
dry (no water vapor) running seals, special additives may be added to help generate lubricious film to
reduce friction.
SiC provides good wear resistance and high thermal conductivity that helps dissipate the heat gen-
erated at the seal face. To manufacture SiC, a mixture of SiC powder, binder, carbon, and some other
sintering aids is pressed into tube form. This “green” material is then machined to near designed shape.
The rough shape is next sintered at high temperature over 3000°F (1700°C). After sintering, the rough
shape is ground and surface finished to meet the design specs. SiC, especially self-sintered SiC grade, is
resistant to corrosion attack. Reaction-bonded SiC, with silicon as the binder, is sometimes chosen for
better lubricity and better chipping resistance. However, reaction-bonded SiC has significant amounts
of silicon that can react with some chemicals, especially caustics, and fail. There are many variations
of SiC products that are gaining acceptance. One popular example is a composite of graphite and SiC.
Another variation is SiC with controlled porosity. These variations were developed to improve seal face
lubrication, either by graphite particulate or residual oil in the pores, during accidental dry running.
It should be noted that some of the advantages are achieved at the cost of a weaker structure and thus
lower material strength.
The next popular hard face material is tungsten carbide. Tungsten carbide is typically made with
either cobalt or nickel binder. Tungsten carbide is a popular seal face material because of its higher
toughness and therefore less chance of breaking or edge chipping during seal installation. However,
tungsten carbide is susceptible to chemical attacks, especially from acids and thus is gradually being
56-4

TABLE 56.2  Typical Seal Face Material Properties


Compressive Thermal Thermal Expansion
Density Young’s Modulus Tensile Strength Strength Conductivity Coefficient Specific Heat
Poisson’s Weibull (BTU/ (W/ (10 −6 (10 −6 (BTU/ (J/
Material Condition (g/cc) Ratio (106 psi) (109 Pa) (103 psi) (106 Pa) (103 psi) (106 Pa) Modulus (ft·h·°R)) (m·°K)) 1/°R) 1/°K) (lb·°R)) (kg·°K))
Carbon Resin 1.8 0.22 3.5 24 7.1 49 34 230 12 5 9 2.7 4.9 0.22 920
graphite filled
Carbon Sb filled 2.2 0.20 3.8 26 9.0 62 40 280 8 14 2.3 4.1 0.15 630
graphite
SiC Self- 3.1 0.14 59 410 34 240 560 3900 9 59 100 2.2 4.0 0.16 670
sintered
SiC Reaction 3.1 0.20 57 390 45 310 390 2700 12 72 120 2.4 4.3 0.17 700
bonded
SiC Self- 2.9 0.20 30 210 11 76 110 740 90 150 2.2 4.0 0.20 820
sintered
10%
graphite
SiC 6% 3.0 0.14 58 400 27 190 360 2500 19 63 110 2.3 4.2 0.14 590
Porosity
WC 6% Ni 14.9 0.25 89 610 130 900 620 4300 24 41 70 2.4 4.3 0.05 210
binder
Al2O3 99% 3.9 0.22 54 370 38 260 380 2600 9.5 17 30 4.6 8.3 0.21 880
Purity
Dairy 8.9 0.33 22 152 77 530 77 530 17 30 9.0 16 0.09 380
bronze
Ni-resist 7.3 0.28 12 83 25 170 99 680 23 40 11 19 0.12 490
Sb: Antimony
Design for Lubrication and Tribology
Face Seals 56-5

replaced by SiC. For some applications in nuclear power industry, the nickel binder grade is preferred to
avoid the risk of radioactive contamination by leached cobalt isotopes.
Alumina (aluminum oxide, frequently referred to as ceramic) is a cost-effective hard face material for
low duty applications. In water, a lubricious boundary film of aluminum hydroxide is formed, which
helps sealing and smooth running. Alumina is corrosion resistant except that the lower purity alumina
(<95%) using silicate as the sintering aid may be susceptible to caustic attack. Other than that, alumina
does not have the thermal shock resistance of SiC or tungsten carbide.
Occasionally, metals such as dairy bronze and Ni-Resist are used as seal ring materials. They can be
successfully applied in low duty oil or water sealing. However, the high thermal expansion coefficient
must be taken into account seriously with these materials.
For limited niche applications, seal faces are sometimes coated with another material, such as hard
chrome, diamond, diamond-like coating, or carbide-derived carbon, to either increase the wear resis-
tance or lubricity. It is imperative to understand the risk of seal face thermal distortion due to material
mismatch and possible coating delamination due to inadequate adhesion. More discussion on seal face
materials can be found in Reference 2.
Brittle materials have significant scatter in material strengths and prediction of brittle material failure
is best described in terms of probability. The two parameter Weibull distribution (Weibull modulus and
characteristic strength) is sometimes used to help quantify the material strength distribution. Weibull
modulus data in Table 56.2 are either reported by the manufacturers or estimated based on limited data
on the strength scatter. Higher Weibull modulus indicates less scatter in the material strength. Material
manufacturers typically report the mean three-point, sometimes four-point, flexural strength test data.
To be used as conservative limits in stress analyses and for fair comparison, these data are converted to
tensile strength estimates by using the following formula based on effective volumes3 for Weibull distri-
bution with volume defect controlling failure:

1/ m 1/ m
St ⎛ 1 ⎞ St ⎛ 3+m ⎞
= , =
S f 3 ⎜⎝ 2(m + 1)2 ⎟⎠ S f 4 ⎜⎝ 6(m + 1)2 ⎟⎠

where
St is tensile strength estimate
Sf 3 and Sf4 are three-point and four-point flexural strength test data, respectively
m is Weibull modulus

The characteristic strength4 So can also be estimated by the following equation:

St ⎛ 1⎞
= Γ ⎜1 + ⎟
So ⎝ m⎠

where Γ is the gamma function.

56.2.3  Seal Arrangements


When the primary ring is mounted on the rotating shaft, it is called a rotating seal head design.
Rotating seal heads are commonly used because of simple construction. For high-speed applications
(peripheral speed greater than 4500 ft/min, or 23 m/s, at the mean face diameter), it is preferred to
have the mating ring fixed on the shaft, that is, the rotating seat design. This helps minimize the
growth of unbalance force caused by the centrifugal effect. Rotating seat designs also have the ben-
efit that the primary ring will tilt to accommodate the seal chamber mounting face run-out (out of
squareness) to the shaft once for all after the installation, rather than once every revolution. This will
56-6 Design for Lubrication and Tribology

avoid excessive dynamic O-ring movement and thus reduce fretting on the balance diameter. The
following is further brief discussion on the seal designs. Additional details and application guide-
lines can be found in Reference 5.

56.2.3.1  Single Seal


The single seal is typically mounted inside the seal chamber (housing) with the process fluid pressure
acting on the outer diameter (OD) side as shown in Figure 56.2, and is called an OD pressurized seal.
This is desirable because hoop stresses will be compressive rather than tensile. Common seal ring mate-
rials such as carbons and carbides have much higher compressive strength than tensile. The outside
mounted seal is subjected to ID pressure and thus tensile hoop stress, and will have a lower pressure
rating. For outside mounted seals, a safety guard may be used to protect the operator from the rotating
components and to confine leakage.

56.2.3.2  Dual Seal


When the process fluid is hazardous, a pressurized dual seal arrangement is used with barrier fluid in
between to isolate the process from the atmosphere. The seal exposed to the process fluid is called the
inboard seal while the seal exposed to the atmosphere is called the outboard seal. The barrier fluid is set
at pressure higher than the process fluid so that the barrier fluid will leak into the process. In addition,
it is important to minimize the eccentricity and angular misalignment to avoid reverse pumping from
shear induced flow6 at the inboard seal. The pressurized dual seal will provide safer isolation of the pro-
cess fluid than unpressurized dual seal. If the process fluid leakage is acceptable to be contained in the
“barrier” fluid area, then the in-between fluid needs not be pressurized. The unpressurized in-between
fluid is specifically called the buffer fluid. Unpressurized dual seals and systems are less expensive to
build and operate than pressurized dual seals and systems. Sometimes the dual seal or a triple seal
arrangement is used for extreme high pressure application to break down the pressure into stages so that
each seal will have reduced loading. No special buffer fluid is used. The process fluid leakage is the fluid
between the seals. The actual pressure staging will depend on the leakage interaction of the seals, and
can be analyzed based on mass conservation of leakage flow. The outboard seal also serves as the safety
seal in case of inboard seal failure. When these seals are mounted one after the other with faces of the
seal heads oriented in the same direction, it is also referred to as the tandem seal arrangement.

56.2.4  Seal Support System


End users often request the face seal to be designed in conformance to and fit in the rotating equipment
built to their industry standards. The most referred pump/seal-related standards are issued by American
National Standards Institute (ANSI) and American Petroleum Institute (API). ANSI B73.1 standard7
is intended for general purpose chemical process pumps, while API 682 standard8 covers heavy duty
pumps for refinery operation, which typically requires more rugged construction. More often than not,
the face seal needs a supporting system to provide clean cool flush so that it can perform its function.
API 682 has defined the standard piping plans for easy communication between users and manufactur-
ers on the specific requirements. Piping plans such as Plan 11—flush flow from pump discharge through
orifice to seal, Plan 13—flush flow out of seal chamber through orifice and to pump suction, Plan 21—
flush flow from pump discharge through orifice and heat exchanger to seal, Plan 23—flush recirculation
from seal by pumping ring through heat exchanger and back to seal, Plan 52—non-pressurized external
fluid reservoir with forced circulation, and Plan 53—pressurized external fluid reservoir with forced
circulation are the commonly referenced.

56.2.4.1  Pumping Ring


One of the critical components in some support systems is an internal circulating device, named pump-
ing ring, which circulates the fluid through the seal chamber and piping system. It is typically mounted
Face Seals 56-7

on the shaft, in close proximity to the seal, and has vanes to pump the fluid either radially outward or
axially forward. In either way, the exit flow has significant circumferential velocity component and it
is highly recommended that the outlet port be drilled tangentially in the housing to effectively convert
velocity into pressure. The pumping ring performance characteristics, pressure head vs. flow rate curve,
can be established by either testing or computational fluid dynamic (CFD) analyses. Such a pumping
curve can be compared against the flow resistance curve of the piping system to determine the equi-
librium pumping flow rate. It should be noted that the pressure head refers to total head, not the static
head. Such distinction is important when using pump affinity law to predict pumping ring performance
at different speeds since it does not generate high pressure head and the elevation difference in the inlet
and outlet ports will affect prediction. The pumped flow, or sometimes called flush as it is injected into
the seal chamber, is to circulate and transfer away the heat generated at the seal interface, fluid churn-
ing loss, and heat soak 9 from high pump temperature. Based on the pumping flow rate, the designer
can then estimate the temperature rise for the flush. The seal flush tends to pick up heat quickly and
reach exit temperature in the bulk fluid around seal.10 API 682 recommends a minimum temperature
margin of 36°F (20°C) below bubble point to avoid creating vapor pockets that are detrimental to seal
performance.

56.2.5  Seal Design Parameters


56.2.5.1  Seal Balance Ratio
By convention, seal balance ratio B is defined as the ratio of closing force area to opening force area as
shown in Figure 56.3. From force equilibrium, this is equivalent to the ratio of average hydraulic closing
pressure on the seal face to the fluid pressure, where pressure refers to the pressure differential relative
to the reference low pressure at ID or OD. Seal balance ratio is determined by seal face outer and inner
radii, and balance radius. The balance radius (Rb) defines the boundary where the fluid pressure acts on
the primary ring. Balanced seals are seals with balance ratio <1. They are commonly used to reduce seal
face loading, especially under high pressure. However, if the seal balance ratio is too low, the seal may
leak excessively. Typical seal balance ratios for plain face seals are
• B ∼ 0.65 for water applications
• B ∼ 0.7 for oil applications
• B ∼ 0.85 for light hydrocarbon applications

56.2.5.2  Seal Balance Ratio Shift with Bellows Secondary Seal


Hydraulic closing force transmitted through the bellows secondary seal does not have a distinctive pres-
sure boundary as does the balance radius (or diameter) as in the O-ring seal although the mean diam-
eter of the bellows serves as a first approximation. However, the exact equivalent balance diameter can
be calculated once the total force is determined through either testing or finite element analyses. Under

Po Po

Pi Rb Ri Ro Pi
Rb Ri Ro

R2o – R2b R2b – R2i


Po > Pi , B= Po < Pi , B=
R2o – R2i R2o – R2i
OD pressurized seal ID pressurized seal

FIGURE 56.3  Definition of seal balance ratio.


56-8 Design for Lubrication and Tribology

high pressure, elastomeric or metal bellows will deform and the equivalent balance diameter will shift.
This shift can be significant and should be taken into account in the design analyses.

56.2.5.3  Spring Force


In addition to the hydraulic pressure, a spring or bellows compression also contributes to the total clos-
ing force. Sometimes the equivalent balance ratio Beq (=B + Psp/|Po − Pi|) is used, instead of seal balance
ratio B, to represent the ratio of total average closing pressure on the face to the fluid pressure differen-
tial. Average spring pressure (Psp) on the seal face is typically designed to be in the range of 15–30 psi
(1–2 bar). Springs can be one single large coil spring or multiple small springs. Large cross-sectional coil
springs can withstand corrosion attack better but multiple springs provide more uniform axial loading
in the circumferential direction. If the axial space is very tight, a wave spring can be used. Coil springs
usually have closed ground ends to provide flat contact against the spring disk, which then transmits
the force to the primary ring. The spring force should be large enough to overcome the secondary seal
(O-ring) friction at no pressure. But too high a spring force can cause severe face contact and damage
the seal, especially in dry running. Once the fluid is pressurized, there is additional hydraulic force to
move the primary ring forward to make contact with the mating ring. For high-pressure applications,
the seal balance ratio is the dominating factor for the closing force. For low-pressure applications, spring
pressure could be the dominating factor instead.

56.2.5.4  Face Width


Radial face width for a plain face seal is typically about 0.187–0.250 in. (4.75–6.35 mm), depending on
the seal size. Noncontacting seals using face pattern such as spiral groove to generate the hydrody-
namic lift will require wider face widths. Wide plain face seals tend to have less leakage, and fewer seal
performance variations due to manufacturing tolerances. On the other hand, the benefit of the narrow
seal face is less total heat generation, which allows the use of higher seal balance ratio to minimize seal
stability (popping open) problems. However, a narrow face is more susceptible to chipping type face
damage, which could result in higher leakage.

56.2.5.5  O-Ring Squeeze


For the dynamic O-ring, it is desirable to have a minimum squeeze to reduce the radial squeeze force
and axial sliding friction while maintaining sealing contact. With large manufacturing tolerances on
the O-ring cross section, it is hard to achieve the minimum squeeze consistently and sometimes a cus-
tom fit is required. Radial percentage squeeze for the dynamic O-ring can be as low as 6%. Static O-rings
can have higher squeeze, 15% or higher, to compensate for possible O-ring shrinkage or compression
set. If the O-ring is confined in a fixed groove, enough groove volume should be provided to accom-
modate its swelling to avoid pressure buildup. When the O-ring seals high pressure gas such as CO2, it
may crack and split if the gas is depressurized too fast. This is called explosive decompression damage.
Controlled slower rate of depressurization has been found helpful in minimizing such risk. Some spe-
cialty grade of elastomer or polymer seal can withstand explosive decompression better and should be
considered. Chemical compatibility check is important in selecting the O-ring material, in addition to
the temperature limit. Typical use and temperature limits of some common O-ring materials are shown
in Table 56.3.

56.2.5.6  Hydrostatic Pressure Effect—k Factor


The pressure gradient factor, frequently referred to as k factor, represents the ratio of average hydro-
static opening pressure on the face to the fluid pressure. If the seal interfacial gap is becoming large, the
k factor will be approaching 1. More typically, the fluid hydrostatic pressure decreases monotonically
from high pressure to low pressure along the varying seal gap and the k factor is <1. If the seal faces are
perfectly parallel, that is, with a constant gap, then the k factor is about 0.5 for incompressible liquid flow
and about 0.67 for compressible gas flow with high pressure ratio.11 The k factor will be higher if the seal
Face Seals 56-9

TABLE 56.3  Typical Use and Temperature Limits of O-Ring


Materials
Temperature Limits
Material Fluid Sealed (°F) (°C)
Fluorocarbon Fuel, oil, acid −20 to 400 −29 to 200
Nitrile (Buna-N) Oil, water −40 to 250 −40 to 120
Ethylene propylene Water, alkali, alcohol −40 to 300 −40 to 150

gap converges, that is, the fluid flows from a larger gap on the high-pressure side to a smaller gap on the
low-pressure side. It is desirable to have an equivalent balance ratio, Beq, larger than the k factor so that
there is some closing force to overcome the secondary friction to ensure face contact.

56.2.5.7  Parallel Sliding Load Support


It is recognized that some load support (opening force) is generated due to sliding at the seal interface
even when the seal faces are nearly perfect flat and parallel. However, it is hard to quantify the mag-
nitude of this parallel sliding load support. Qualitatively speaking, such load support is higher with
more viscous fluids, smaller gaps, and higher sliding speeds.12 It is important to take parallel sliding
load support into account for oil sealing applications, but not as critical for light hydrocarbon sealing
applications.

56.2.5.8  Hydrodynamic Pressure from Grooved Face Pattern


For difficult sealing applications, designers may choose to use special face pattern, such as spiral
grooves, face waviness, or hydropads, to generate additional hydrodynamic lift to relieve most of the
face contact load if not all. These seals typically operate with a gap as large as 100 μin. (2.5 μm), and
therefore a high leakage rate is expected. To counteract the increase in hydrodynamic and hydrostatic
opening forces due to the pattern, a higher seal balance ratio is used. The spiral groove gas seal has been
applied extensively in dry running applications for compressors. Similar design of spiral groove gas seal
is also gaining acceptance in pumps, especially used in a pressurized dual seal arrangement to isolate
the hazardous process fluid with a barrier gas. It should be noted that the gas leakage into the process
side needs to be vented away to avoid vapor lock in the pump. When the spiral groove seal runs in the
liquid, the seal leakage could be 1000 times more than plain face seals and is typically unacceptable. It is
possible to use part of the face pattern to perform upstream pumping, that is, pumping fluid against the
pressure. Using the combination of downstream pumping (pumping along the pressure) and upstream
pumping, various degrees of operating gap and seal leakage can be achieved. Some of the designs can
achieve noncontacting and nonleaking conditions.13 Because of the criticality of face distortion control,
usually sophisticated computer modeling is used to predict the seal performance at the design stage. The
computer program will calculate the fluid flow through the interfacial gap with fluid property variation,
fluid pressure, seal face temperature, and face distortion from fluid–solid interaction. A full 3D com-
putational analysis will require using a large number of finite elements due to the extreme length scales
existing in the face pattern and is cost prohibitive. As a practical compromise, the Reynolds equation
is used, assuming all variations in the seal gap direction (normal to the face) are negligible, instead of
full Navier–Stokes equation, which renders the analysis task more manageable.14 The approach has been
widely adopted in the seal analyses and leads to successful seal designs and operations in the field.

56.2.5.9  Operating Gap, Contact Pressure, and PV Limit


It is desirable to keep the sealing interfacial gap small in order to minimize the leakage. However, seal
faces are not perfectly flat and smooth. For a typical running contacting plain face seal, the mean oper-
ating gap is about two to three times of the combined surface roughness. When the mean gap is about
56-10 Design for Lubrication and Tribology

TABLE 56.4  PV, PcV Limits for Common Face Material Pairs in
120°F (49°C) Water

Mating PV Limit PcV Limit


Primary Ring Ring psi·fpm MPa·m/s psi·fpm MPa·m/s
Carbon graphite SiC 500,000 18 400,000 14
Carbon graphite WC 500,000 18 370,000 13
Carbon graphite Al2O3 100,000 3.5 80,000 2.8
WC WC 120,000 4.2 110,000 3.9

three times the combined surface roughness, asperities start to contact. If the mean gap is less than
the combined surface roughness, the face contact is too severe. High contact pressure can occur on the
face OD or ID edge of the hard faces under pressure loading. Such high line contact pressure at the
edge can cause face chipping problem and should be avoided, preferably even under static condition.
Furthermore, face contact at high sliding speed will generate high frictional heating, which can cause
surface film desorption, fluid degradation, varnish deposition, and heat checking of the faces and result
in severe wear. For well run-in seals without concentrated face contact, there are still PV ­(pressure–
speed) limits beyond which severe face wear can occur. The PV limits15 were established by seal manu-
facturers through numerous tests to guide designers to choose the appropriate face materials and seal
balance ratio for the applications. Typically, PV limits for running in water around 120°F (49°C) with a
minimum expected seal life of 2 years are shown in Table 56.4. These data were based on testing of 2 in.
(51 mm) and 3–5/8 in. (92 mm) diameter shaft seals with a seal balance ratio of 122%–139% at either
1800 or 3600 rpm and various pressures. The traditional PV limit was calculated in the following equa-
tion by assuming a k factor of 0.5:

PV = [(B − k)ΔP + Psp ] Vm or PV = (Beq − k)ΔPVm


where
PV is the pressure–speed limit
B is the seal balance ratio
Beq is the equivalent seal balance ratio, = B + Psp/ΔP
k is the pressure gradient factor
ΔP is the pressure differential, = |Po − Pi|
Psp is the spring pressure
Vm is the mean face speed

However, in these water tests, the k factor actually varied from 0.5 to 0.7, due to fluid vaporization at
the seal face. To better assess the contact severity, a modern computer modeling tool can be used to
reanalyze the test seal operation. This allows more accurate k factor determination and inclusion of
some minor effect from hydrodynamic load support. The revised PV (denoted as PcV) limits, based on
the calculated average contact pressure, are shown in Table 56.4 for comparison. PcV limits represent
the tribological pair rubbing limits and are more suitable for use as the limiting criterion in more exact
seal analyses:

ΔPh
PcV = Beq − k − ΔPVm
ΔP

where Ph represents the average hydrodynamic pressure generated due to parallel sliding load support.
In designing a face seal, it is a delicate “balancing” of Beq, k factor, and Ph so that PcV stays within the
material limit. Such a seal can perform well even with nominal PV being higher than published limits.
Face Seals 56-11

TABLE 56.5  Representative Contact Friction Coefficients for


Running in Water
Mating Contact Friction
Primary Ring Ring Coefficient Source16
Carbon graphite SiC 0.09 Paxton and Hulbert17
Carbon graphite WC 0.1 Paxton and Hulbert17
SiC SiC 0.1 Lohou18

56.2.5.10  Friction at Seal Interface


Friction at the seal interface affects seal face temperature rise, thermal distortion, and leakage flow. It
is the seal engineers’ hope that the friction can be reasonably calculated without resorting to testing
every time. Friction at seal interface comes from viscous shearing of the fluid film and contact friction.
Viscous shearing can be calculated once the fluid film thickness and fluid viscosity are known in the
analysis. Contribution from viscous shearing is typically small. Contact friction comes from shearing
the boundary film. As in the boundary lubrication regime, the contact friction coefficient is expected
to be independent of load and speed, but will most likely vary according to seal face material combina-
tion and sealed fluid. It is best measured in tests run at low speed to minimize the hydrodynamic effect
and in a non-pressurized seal configuration to avoid the hydrostatic effect. There are wide variations of
“friction coefficients” reported in literature.16–18 However, the published data (shown in Table 56.5) cor-
relate best with laboratory test results19 at low speed, 4.1 in./s (0.1 m/s) at mean face diameter, in 120°F
(49°C) water. It is interesting to note that for various material combinations, the typical contact friction
coefficient running in water is ∼0.1.

56.2.5.11  Heat Transfer


A significant amount of heat is generated at the seal interface and it needs to be effectively transferred
away. Typically, the majority of the heat is conducted through both seal rings and dissipated through
heat convection to the flush fluid at the OD around the seal face area. The convective heat transfer coef-
ficient can be estimated by Becker’s empirical equation20 based on a small diameter cylinder rotating in
a big tank of water:

N u = 0.133R2e /3Pr1/3

where
Nu is the Nusselt number, = hD/kt
Re is the Reynolds number, = ρVD/μ
Pr is the Prandtl number, = μCp/kt
h is the convective heat transfer coefficient, = Q/[A(Ts − Tf)]
D is the location diameter of heat transfer surface
V is the tangential velocity at the heat transfer surface
kt is the fluid thermal conductivity
ρ is the fluid density
μ is the fluid viscosity
Cp is the fluid specific heat
Q is the heat flow
A is the heat transfer area
Ts is the solid surface temperature
Tf is the fluid temperature

For poor convective heat transfer fluid like oil, all seal surface area, not just the area close to the
seal interface, will be at high temperature and contribute to convective heat transfer. Under such
56-12 Design for Lubrication and Tribology

circumstances, heat conduction through other components such as sleeve, gland plate, retainer, etc.,
could be important and may need to be taken into account in the analysis model. In addition to heating
at the interface, the windage (or churning loss) at the rotating surface will increase the heat load and
should be considered. The following empirical equations by Bilgen and Boulos21 can be used to estimate
the shear stress τ:

t = ¼ C M rV 2

where loss coefficient CM is calculated as follows:

C M = 10(c /R)0.3 Rec−1.0 Rec ≤ 64, laminar flow

C M = 2(c /R)0.3 Rec−0.6 64 < Rec ≤ 500, transition flow

C M = 1.03(c /R)0.3 Rec−0.5 500 < Rec ≤ 104, turbulent flow

C M = 0.065(c /R)0.3 Rec−0.2 10  < Rec, turbulent flow


4

The churning loss, P, can be calculated using the following equation:

P = ½ C M ρV 3pLR

where
c is the radial gap
R is the radius of cylinder
Rec is the Reynolds number, = ρVc/μ
L is the length of cylinder

There is an increasing number of studies using CFD to analyze the heat transfer of face seals. Analysis
results appear to indicate different exponent values for the variables in Becker’s equation for different
flow regimes. More comprehensive studies are needed before we can close the chapter on prediction of
convective heat transfer coefficients for face seal analyses.

56.2.6  Other Face Seal Performance Limits


56.2.6.1  Leakage
Allowable leakage rates will depend on customers’ specific requirements and should comply with appli-
cable government regulations. Liquid leakage rates on the order of 1.5 g/h/in. (0.001 g/min/mm)22 are
typically acceptable. Leaks of volatile organic compounds (VOC) are strictly regulated and they can be
classified by measurement in terms of ppmv (parts per million by volume) concentration according to
Method 21 of the U. S. Environmental Protection Agency. As a rule of thumb, a leakage rate of 2.9 g/h of
VOC corresponds to 450 ppmv with Method 21 measurement.23 Theoretical mass leakage rates for the
laminar flow through a constant interfacial gap can be calculated in the following11 using Figure 56.3:
For isothermal incompressible laminar flow,

• 2prRmh3 (Po − Pi )
m= −
12 m (Ro − Ri )

For isothermal compressible perfect gas laminar flow,


Face Seals 56-13

• 2pRmh3 (Po + Pi ) (Po − Pi )


m= −
12m 2RgasT (Ro − Ri )

where
m· is the mass leakage rate, negative value indicates radially inward leakage direction
Rm is the mean face radius
h is the interfacial gap
Rgas is the specific gas constant
T is the absolute temperature of fluid

56.2.6.2  Carbon Blistering


Carbon rings may experience blistering (discrete gentle bulging and uplift of surface material) type face
damage during start-up in sealing high-viscosity oil.11,24 Tests had showed that using a rougher surface
finish or metal impregnation on the carbon will mitigate, but not eliminate, the problem. As a design
guide, be careful with using carbon face when the fluid viscosity is greater than 32 cSt with significant
nominal PV, say greater than 30,000 psi·fpm (1 MPa·m/s).

56.2.6.3  Thermal Cycling and Heat Checking


Sometimes the seal operates in an unstable mode; it exhibits periodic variations in the film thickness
and face temperature, with time periods in the order of seconds. The phenomenon is referred to as ther-
mal cycling.25 It appears to be resulting from delayed thermal coning, due to thermal inertia, in response
to frictional heating from face contact. Continuous operation at this point will cause high wear rate and
excessive leakage and is not recommended. Better material combinations or distortion control could be
utilized to avoid this problem.
Under heavy load and high speed, the tungsten carbide (and sometimes SiC) rings may suffer heat
checking (multiple fine radial cracks) type face damage with high intensity face heating.26 No definitive
criterion has been established to avoid such failure yet. But in general, it is advisable to reduce the face
contact pressure with a better seal design and cross section. Hard face materials with higher thermal
stress resistance factor,27 defined as follows, tend to have better resistance to heat checking:

kSt (1 − n)
Rts =
Ea

where
Rts is the thermal stress resistance factor
k is the thermal conductivity
St is the tensile strength
ν is the Poisson’s ratio
E is the Young’s modulus
α is the thermal expansion coefficient

Figure 56.4 compares the thermal stress resistance factor of common hard face materials and confirms
the experience that SiC and WC are better materials than alumina to withstand stresses induced from
sudden temperature change.

56.2.6.4  Minimum Pressure for High-Speed Sealing


Field experiences with high-speed sealing applications suggested that some minimum OD pressure
needs to be maintained so that seal face can be properly wetted when the sealed liquid is on OD side,
56-14 Design for Lubrication and Tribology

Graphite
loaded
WC
SiC
5800 11500 17300 W/m Thermal
stress
resistance
0 factor
500 1000 1500 BTU/(h-in.)
Alumina Porous Self- Reaction
SiC sintered bonded
SiC SiC

FIGURE 56.4  Thermal stress resistance factor of common hard face materials.

especially with hard faces. Based on Wallace’s derivation,28 the following equation can be used to calcu-
late the required minimum pressure differential across the seal face:

Psp +(rw 2 (Ro2 − Ri2 )/8) − s s


ΔP ≥
1− B

where
σs is an empirical parameter, 13.1 psi (0.9 bar), accounting for some surface tension effect
B is the seal balance ratio

Furthermore, for the incompressible Newtonian fluid flow between an infinitely long rotating cylin-
der inside a stationary cylindrical housing, an exact solution to Navier–Stokes equations exist 29 before
flow becomes unstable and toroidal Taylor vortices occur:

b ∂P V2
U = 0, V = ar + , W = 0, =r
r ∂r r

h2 Ro2
a = −w , b=w
1−h 2
1 − h2

By integrating from rotating cylinder radius to housing bore radius, pressure drop can be obtained as
in the following30:

rw 2 Ro2 1 ⎛ 4h2 ⎞
ΔPh = 2 ⎜
1 + h2 + ln h⎟
2 1−h ⎝ 1 − h2 ⎠

where
U is the radial velocity
V is the tangential velocity
W is the axial velocity
P is the fluid pressure
ΔP is the pressure drop across the seal face
ΔPh is the pressure drop, Phousing − Po
Phousing is the fluid pressure at housing bore
Po is the fluid pressure at rotating cylinder radius
Psp is the spring pressure
ρ is the density
Face Seals 56-15

η is the radii ratio, Ro/Rhousing


Ro is the radius of rotating cylinder, seal face outer radius
Rhousing is the radius of housing bore
Ri is the seal face inner radius
ω is the rotating speed of inner cylinder

It is recommended that seal chamber fluid pressure (at housing entrance) be kept above ΔP + ΔPh to
ensure appropriate wetting on the hard seal faces in an OD pressurized seal application. If estimated
ΔPh from previous equation is too high to be accommodated in the application, then a more exact CFD
analysis is recommended for assessing the risk of dry running.

56.2.6.5  Dynamic Tracking


The primary ring is designed to float axially and tilt diametrically to track the mating ring axial move-
ment and angular misalignments. If the seal does not track properly, it can result in excessive leakage
or/and face contact. In API 682 standard, maximum out of squareness of 0.0005 in./in., or 15 μm/3 cm,
for the shaft to seal chamber mounting face is recommended. Contacting seals rely on the contacting
forces on the faces to compel the primary rings to track and typically do not experience dynamic track-
ing problems except for mixer applications. The mixer application has a special challenge due to radial
run-out and large shaft deflection because of the high slenderness ratio of the shaft. The seal has to be
designed to accommodate such motions. Mixer rotating speed is typically low but the fluid pressure can
be high while the seal size is large. It is not uncommon that it uses hard faces to withstand the mechani-
cal stress. Furthermore, if inert gas is used as the barrier fluid to avoid process contamination, then the
seal could be running dry. In aforementioned conditions, the material PV limit becomes an important
factor to be considered for dynamic tracking of the contacting seal.
Noncontacting seals rely on the fluid forces on the seal face to help dynamic tracking. However, the
seal can run unstably and result in high leakage and possible face contact due to the interaction of cross-
coupled fluid forces, seal rings, and supports (O-ring, spring). For a coned face seal with a rotating mat-
ing ring, Green and Etsion31 derived the following equation to determine the critical speed above which
the seal can become unstable, that is, stability threshold:

2
K f* + K s* ⎛ D* ⎞
w =4
2
cr ⎜ 1+ s ⎟
I * ⎝ D*f ⎠

where
ωcr is the critical angular velocity
K *f is the angular fluid film stiffness coefficient
K *s is the angular support stiffness coefficient
I * is the stator moment of inertia
D*s is the angular support damping coefficient
D *f is the angular fluid film damping coefficient

For seals with special face patterns such as spiral grooves to generate hydrodynamic pressure while run-
ning in compressible fluid, no closed form solution for the dynamic tracking analysis is available. Direct
numerical simulation had been used to study the dynamic responses of such seals during start-up and
shutdown.32

56.2.6.6  Varying Operational Conditions


It is important to assess how the seal will perform under varying operational conditions, especially
when the seal will experience face contact and wear. For example, it is not uncommon to find that some
56-16 Design for Lubrication and Tribology

seals appear to work well during dynamic running but start to leak while being depressurized after shaft
rotation stopped. Such static leakage is usually due to excessive OD face wear. Modifying the seal cross
section to decrease the pressure distortion and reduce the face wear usually eliminates such perfor-
mance variation and improves the seal robustness.

56.2.6.7  Friction Hysteresis Effect in Multicomponent Design or in Mating Ring Support


Some seal designs use multiple components for the primary ring or mating ring assembly to isolate
the face and minimize the face distortion. But they are typically made of different materials and an
extremely large force can be transmitted through component contacts, especially under high pressure.
As the components deform, stick–slip may occur at the contact interface due to relative deflection.33 As
a result, the interfacial friction will evolve and cause further face distortion. The final face distortion will
depend on the friction variation that is affected by loading history, not just the final loads, and is thus
called friction hysteresis effect. To achieve more consistent seal performance, it is desirable to minimize
such effects by reducing the contact load or friction coefficient at the component interface.

56.2.6.8  Material Strength Limit


For simplified analyses, designers are using safety factors, based on Coulomb–Mohr theory,4,34 to
account for the scatter in the brittle material strengths. Typically, a safety factor of 2–3 is used. But a
higher safety factor is recommended for large size rings, say 4 in. (100 mm) diameter or greater, as they
are likely to have more surface or volume defects. For sophisticated components with stress concen-
tration feature or critical application, a comprehensive finite element analysis with failure probability
calculation3 should be considered.

56.2.6.9  Seal Temperature Limit


Component materials have their temperature limits. They are exposed to the process fluid temperature
plus the additional temperature rise due to interfacial frictional heating. Elastomeric materials are espe-
cially vulnerable to the temperature effect and their temperature limits are already shown in Table 56.3.
It also needs to be noted that the process fluid, barrier or buffer fluid such as oil also has its temperature
limit before it starts to degrade, break down, and leave deposits on the face. As a general reference, 35
maximum useful bulk oil temperatures for mineral oils are around 330°F (165°C). It is desirable to keep
seal face temperature below these limits to minimize material degradation and coking deposits.

56.2.6.10  Seal Face Wear—Adhesive, Abrasive, and Erosive Wear


Seal face wear is typically mild adhesive wear unless it exceeds the PV limit and becomes severe adhe-
sive wear. Most of the wear occurs during the initial run-in phase unless the operating conditions keep
on changing. Variations in operating conditions will force the seal to change the deformation and
continue to wear at high rate. If the process fluid contains abrasive particles, then hard face (carbide)
materials should be used for both the primary ring and mating ring. To avoid hard abrasion, they
should be harder than 80% of the abrasive hardness. 36 Erosive wear on the seal face is not common,
but can occur due to excessive leakage, which is more likely with noncontacting seals and should be
watched. It can also occur outside the seal face when the flush fluid is injected at high velocity and
impinges on the seal ring. In practice, a flush velocity limit of 1800 ft/min (9 m/s) is typically used in
the industry.5

56.2.6.11  Dynamic O-Ring Extrusion and Seal Hang-Up


Under high sealing pressure, there will be serious deformations in all components. Designer needs to
check the relative diametral clearance at the dynamic O-ring location. The extrusion pressure limit of
O-ring can then be assessed by using Figure 56.5 either for multiple pressure cycles (Ksieski’s results) or
for single cycle (Ashel’s results) based on the nondimensional clearance.37 It is also important to check
that the primary ring does not interfere with the retainer at the balance diameter. The interference can
Face Seals 56-17

Plastic Elastic d(mm) Shore hardness


Viton: 3.53 90

E=1
Buna-N:

7M
3.53 90 Present

5
3.53 70

Pa
work
5.33 70

(90
103

(70)
2.62 70

(50)

Sho
1.78 70
3.53 70 Parker

re)
3.53 70–90 ksieski
Plastic theory Ssy = 30 MPa
Elastic theory f = 0.2

102
Extrusion pressure, p (MPa)

10

90 Shore
80
1 70

Ksieski’s results
for 105 pressure cycles at 70°C

0.1
50 Shore 60 70 80 90

0 0.2 0.4 0.6 0.8 1.0


Non-dimensional clearance, c/d

FIGURE 56.5  Extrusion pressure limits for O-ring seals. (Reprinted from Ashel, R., ASLE Trans., 27(4), 332, 1984.
With permission.)

compromise seal’s ability for dynamic tracking and even cause hang-up. Seal hang-up will result in
excessive leakage and possible severe face damage.

References
1. Buchter, H.H., Industrial Sealing Technology, Wiley-Interscience, New York, 1979.
2. Nau, B.S., Mechanical seal face materials, Proceedings of Institution of Mechanical Engineers Part J,
3(211), 165–183, 1997.
3. Nemeth, N.N., Powers, L.M., Janosik, L.A., and Gyekenyesi, J.P., Cares/Life Ceramics Analysis and
Reliability Evaluation of Structures Life Prediction Program, NASA/TM-2003-106316, 2003.
4. Munz, D., Failure modes: Performance and service requirements for ceramics, Handbook of Materials
Selection, Kutz, M. (Ed.), John Wiley & Sons, New York, pp. 787–808, 2002.
56-18 Design for Lubrication and Tribology

5. Seal Committee Members (Ed.) Mechanical Seals for Pumps: Application Guidelines, 1st edn.,
Hydraulic Institute Parsippany, NJ, 2006.
6. Findlay, J.A., Measurements of leakage in mechanical face seals, 4th International Conference on
Fluid Sealing, May, ASLE, Philadelphia, PA, 1969.
7. ANSI B73.1, Specification for Horizontal End Suction Centrifugal Pumps for Chemical Process,
American National Standards Institute, New York, 2001.
8. API Standard 682, Shaft Sealing Systems for Centrifugal and Rotary Pumps, 3rd edn., American
Petroleum Institute, Washington, DC, 2004.
9. Buck, G.S. and Chen, T.Y., An improved heat soak calculation for mechanical seals, Proceedings of
the 26th International Pump Users Symposium and Short Courses, March, Texas A&M, Houston, TX,
pp. 33–37, 2010.
10. Shirazi, S.A., Soulisa, R., Lebeck, A.O., and Nygren, M.E., Fluid temperature and film coefficient pre-
diction and measurement in mechanical face seals-numerical results, Tribology Transactions, 4(41),
459–470, 1998.
11. Lebeck, A.O., Principles and Design of Mechanical Face Seals, John Wiley & Sons, New York, 1991.
12. Lebeck, A.O., Parallel sliding load support in the mixed friction regime, Part 1—The experimental
data, Part 2—Evaluation of the mechanisms, Journal of Tribology, January, 189–205, 1987.
13. Lai, T., Development of non-contacting, non-leaking spiral groove liquid face seal, Lubrication
Engineering, 8(50), 625–631, 1994.
14. Meck, K.D. and Zhu, G., Improving mechanical seal reliability with advanced computational engi-
neering tools, Part 2: CFD and application examples, Sealing Technology, 2008(2), 7–10, 2008.
15. Johnson, R.L. and Schoenherr, K., Seal wear, Wear Control Handbook, Peterson, M.B. and Winer,
W.O. (Eds.), American Society of Mechanical Engineers, New York, pp. 727–753, 1980.
16. Lebeck, A.O., Dynamic seals, Tribology Data Handbook, Booser, E.R. (Ed.), STLE, New York,
1997.
17. Paxton, R.R. and Hulbert, H.T., Rubbing friction in radial face seals, Lubrication Engineering, 36(2),
89–95, 1980.
18. Lohou, J., Mechanical seals for water injection pumps—A new hard face material, Lubrication
Engineering, 34(6), 320–326, 1978.
19. Lai, T., Friction Coefficients for CSTEDYSM Analyses, 3rd Memo, John Crane, March 2, 2000.
20. Becker, K.M., Measurements of convective heat transfer from a horizontal cylinder rotating in a tank
of water, International Journal of Heat Mass Transfer, (6), 1053–1062, 1963.
21. Bilgen, E. and Boulos, R., Functional dependence of torque coefficient of coaxial cylinders on gap
width and Reynolds numbers, Journal of Fluids Engineering, March, 122–126, 1973.
22. Nau, B.S., Rotary mechanical seals in process duties: An assessment of the state of the art, Proceedings
of the Institution of Mechanical Engineers, 1(199), 17–31, 1985.
23. Kittleman, T., Pope, M., and Adams, W., CMA/STLE pump seal mass emission study, Proceedings
of the 11th International Pump Users Symposium and Short Courses, Texas A&M, Houston, TX,
pp. 57–62, 1994.
24. Guichelaar, P.J., Wilde, D.A., and Williams, M.W., The detection and characterization of blisters on
carbon–graphite mechanical seal faces, Tribology Transactions, 43(3), 395–402, 2000.
25. Parmar, A., Thermal cycling in mechanical seals-causes, prediction and prevention, 13th International
Conference on Fluid Sealing, April 7–9, BHRA, Brugge, Belgium, pp. 507–526, 1992.
26. Netzel, J.P., Surface disturbances in mechanical face seals from thermoelastic instability, Lubrication
Engineering, 5(37), 272–278, 1981.
27. Kingery, W.D., Factors affecting thermal stress resistance of ceramic materials, Journal of the
American Ceramic Society, 1(38), 3–15, 1955.
28. Wallace, N.M., Predicting and ensuring the performance of hard faced seals, 13th International
Conference on Fluid Sealing, April 7–9, Brugge, Belgium, pp. 369–387, 1992.
29. Lueptow, R., Taylor–Couette flow, Scholarpedia, 4(11), 6389, 2009.
Face Seals 56-19

30. Lücke, M. and Roth, D., Structure and dynamics of Taylor vortex flow and the effect of subcritical
driving ramps, Zeitschrift für Physik B—Condensed Matter, 78(1), 147–158, 1990.
31. Green, I. and Etsion, I., Stability threshold and steady-state response of noncontacting coned-face
seals, ASLE Transactions, 28(4), 449–460, 1985.
32. Ruan, B., Numerical modeling of dynamic sealing behaviors of spiral groove gas face seals, ASME
Journal of Tribology, 124(1), 186–195, 2002.
33. Metcalfe, R., End-face seal deflection effects—The problems of two-component stationary or rotat-
ing assemblies, ASLE Transactions, 23(4), 393–400, 1980.
34. Shigley, J.E. and Mischke, C.R., Standard Handbook of Machine Design, McGraw-Hill, New York,
1986.
35. Loomis, W.R. and Fusaro, R.L., Liquid lubricants for advanced aircraft engines, NASA Technical
Memorandum 104531, 1992.
36. Hutchings, I.M., Tribology—Friction and Wear of Engineering Materials, CRC Press, Boca Raton, FL,
1992.
37. Ashel, R., Prediction of extrusion failures of O-ring seals, ASLE Transactions, 27(4), 332–340, 1984.
57
Lip Seals
57.1 Introduction..................................................................................... 57-1
57.2 Basic Lip Seal Design....................................................................... 57-2
57.3 Dynamic Sealing Mechanism........................................................ 57-4
57.4 Pumping Aids................................................................................... 57-4
57.5 Performance Limits......................................................................... 57-5
57.6 PTFE Seals........................................................................................ 57-6
57.7 Excluders........................................................................................... 57-9
57.8 V-Ring Seals.................................................................................... 57-10
57.9 Bearing Seals.................................................................................. 57-12
57.10 Lip Seals for Pressure.................................................................... 57-12
57.11 Seals for Fluctuating Pressure Applications.............................. 57-13
57.12 Plastic Lip Seals for Pressure Applications................................ 57-14
57.13 Seal Materials................................................................................. 57-15
Nitrile  •  Hydrogenated Nitrile  •  Polyacrylic  •  Ethylene
Acrylic  •  Silicone (MVQ, MPQ, MPVQ)  • ​Fluorocarbon  • ​
Robert K. Flitney
Polytetrafluoroethylene
Sealing Technology
Consultant References................................................................................................... 57-17

57.1  Introduction
Lip seals are widely used for sealing rotary shafts. They are now the most numerous rotary shaft seal
being easily found in everything from domestic equipment through automotive and other transport
applications to heavy-duty industrial equipment. Although primarily designed and used for sealing
lubricants, they are increasingly being specified as a replacement for other seal types in, for instance,
low pressure water applications such as domestic washing machines, automotive water pumps, and large
civil equipment such as water turbines.
The development was largely empirical for some 50 years, but, over the last 20 years, detailed math-
ematical analysis has provided a more established basis for understanding the details of how these
seals operate. The development of new elastomer and plastic materials together with the improved
understanding of how they should be used has also contributed to considerable extensions in potential
performance.
The types of seals in common use can be usefully divided into three broad categories:
• Elastomer lip seals, used for sealing of shafts against exit or ingress of liquids
• Plastic, primarily filled PTFE (polytetrafluoroethylene), used as an alternative to elastomer seals
where fluid resistance or lack of lubrication may be a problem and for sealing gas
• Bearing seals, a derivative of lip seals that usually operate at lower lip loads and have exclusion of
contaminants from the bearing as a prime function

57-1
57-2 Design for Lubrication and Tribology

It is important to remember that the elastomer or plastic component is only one part of the seal. The
shaft that rotates within the seal is equally important to creating the correct tribological conditions to
form a reliable dynamic seal. In addition to the surface texture, the machining method is also an impor-
tant factor affecting seal performance. The shaft is also the primary method of heat dispersion from the
seal so shaft design and material are important considerations.
Standard lip seals are designed to operate in a splash or nominally flooded environment but with
little or zero applied pressure. The maximum pressure rating quoted by major manufacturers is 0.3–0.5
bar. Increasing pressure causes distortion of the lip, which increases the contact area, and a rise in the
contact pressure of the lip. The combination of the two creates increased temperatures and wear. Special
designs of lip seal are available for elevated pressure.

57.2  Basic Lip Seal Design


A cross section of a typical lip seal is shown in Figure 57.1. A number of different methods are used to
describe the components of a lip seal, but a standardized vocabulary is available [1,2]. The seal design
looks deceptively simple, but the detail has been developed by practical experience, and more recently
numerical analysis, to derive the current designs.
The precise detail of individual seal designs will vary, depending upon the intended application and
the preferences of a particular manufacturer, but the overall design principle will be the same. The func-
tion of the seal is derived from the asymmetric geometry. The seal is manufactured to provide a sharp
sealing lip that contacts the shaft. The geometry is such that the angle of the elastomer on the oil side
of the seal is steeper than that on the air side. Typical design angles are in the region of 40°–45° for the
oil side angle β and 25°–30° for the air side angle α when the seal is in the free state. After installation,
the interference on the shaft will cause these angles to change by perhaps 10°, providing angles in the
working condition that will be approximately 50° and 20°, respectively. A further contribution to the
asymmetric stress distribution is governed by the position of the spring. The relative spring position is

Elastomer

Metal case

Air side Lubricant side

H
R
t
+
+
β

Hinge α
point Radial
interference

FIGURE 57.1  Cross section of a basic lip seal showing the main design features.
Lip Seals 57-3

offset from the contact line of the lip, R, by typically 10% of the lip length, H. This offset is toward the
air side of the seal. The spring position is important to provide the correct combination of direct loading
to the lip and a contribution to the asymmetric geometry. The predominant choice for the spring is a
garter spring. This provides a localized circumferential spring force in the location required. Some seals,
generally those manufactured in low volumes and larger sizes, are fitted with a finger spring.
The overall geometry of the seal will vary depending on the application and size of seal. The flexibility
of the seal will be governed by the lip length, H, and thickness, t, at the hinge point. This is normally
close to the metal casing. For a standard seal, the thickness will be something less than 50% of the lip
length H. For increased flexibility, the length of the flex section may be increased, for instance, to cope
with larger-than-usual shaft radial deflections. Alternatively, for seals designed to operate at pressure,
the thickness may be increased and the flex length shortened. This will have the effect of reducing the
run-out capability of the seal.
The seal lip is usually molded to an L section metal casing. This supports the lip and forms the static
location to mount in the seal housing. The metal case may be entirely or partially covered in elastomer.
The outside diameter of the casing is an interference fit in the stationary housing to form the static seal.
Using the standard lip profile, a number of variations are used for the sealing of the casing on the outside
diameter, and some of these are shown in Figure 57.2.
A large range of geometries are available to cater for different applications, very often including facili-
ties for the exclusion of contaminant, some of which are discussed later.

(a) (b)

(c) (d)

FIGURE 57.2  Examples of the different lip seal casing designs available. (a) Elastomer-covered casing. (b) Ribbed
elastomer coating. (c) Metal casing. (d) Metal casing with partial elastomer sleeve on the OD to provide improved
retention.
57-4 Design for Lubrication and Tribology

57.3  Dynamic Sealing Mechanism


The dynamic sealing mechanism of elastomer lip seals has been the subject of extensive research. The
existence of a continuous oil film between the lip and shaft was first demonstrated by Jagger in 1957 [3].
The mechanism by which this lubricated seal will also continuously pump the sealed fluid back into the
system was first explained by Kammüller in 1986 [4].
These mechanisms are extensively discussed together with descriptions of the research involved in
Salant 2010 [5] and Flitney 2009 [6].
A new seal has a sharp trimmed or molded lip. During the first hour or so of running, this lip beds in
on the shaft to create a sealing or contact zone that is typically 0.2–0.3 mm wide. This may increase with
continuing seal wear to be something in excess of 0.5 mm after some hundreds or thousands of hours of
operation. An important factor in achieving an optimum lubricant film is the surface texture of the seal
once it has bedded in, and, hence, specific polymer compounds are formulated to achieve the desired
surface on the seal.
The shaft surface texture is, therefore, also a critical parameter in the operation of this sealing mech-
anism as the lubricant film is of the same order as the shaft texture. In addition to the scale of the tex-
ture, the “lay” of the machining will influence the pumping action. Conventional cylindrical grinding
and other traditional finishing operations create what is effectively a micron-scale screw thread on the
shaft. Underneath the seal, this will act as a micro screw pump and effectively pump oil past the seal.
This action is dependent on the direction of rotation and so may be beneficial or create leakage. The nor-
mal recommendation for lip seals is that the shaft should be plunge ground in the area of seal contact,
so that there is no shaft lead. The introduction of new shaft finishing techniques for volume production
has led to further work in this area and some revised recommendations such as those of Schmuker and
Haas, 2007 [7].
To ensure successful bedding in of the seal, the shaft texture has been shown to be quite critical. The
traditional specification for plunge ground surfaces is 0.25–0.8 μm Ra. If the shaft is too smooth, or has
a polished surface, the seal will not bed in correctly, especially in the presence of a viscous lubricant.
However, these recommendations need to be reviewed if a modern high-volume machining method is
employed [7].

57.4  Pumping Aids


A number of seals are produced with what are variously termed pumping vanes, hydrodynamic aids, or
sealing aids. These comprise small vanes or webs on the air side of the lip. The benefits of these aids are
claimed to include the following: to assist during the bedding in process seals when they may be prone to
leak, to help extend the performance parameters of a seal, to continue sealing when a seal would other-
wise be considered to have worn out, to enable a seal to operate with higher speed or run-out conditions,
or to even overcome otherwise faulty seal manufacture. The actual benefits and method of operation are
still a matter of debate [6]. The majority of aids on modern seals are small raised ribs molded on the air
side of the seal lip, Figure 57.3a. The ribs are in contact with the shaft and merge into the lip. If the ribs do
not contact the shaft, they will not be effective. These designs are for shafts that predominantly have one
direction of rotation. In situations that have intermittent reverse rotation, such as vehicle transmissions,
the seal will still function. They do not pump leakage out when used in reverse, as they only function on
fluid that has already leaked.
Where there is a true bidirectional rotation, or a seal may be used for a variety of applications,
then bidirectional seal designs are available, Figure 57.3b. A sinusoidal molding feature has also
been used. It may appear at first that such seals would not work as the two opposing sets of pumping
aids would work against each other. This is not so as the flutes for the “wrong” direction present a
diverging wedge to any leakage and, therefore, have little influence, as discussed for reverse rotation
earlier.
Lip Seals 57-5

Unidirectional

Unidirectional

Bidirectional

(a) Bidirectional

Shaft rotation

(b) Raised flutes

FIGURE 57.3  Lip seal with pumping aids on the seal lip and examples of the shaft contact pattern of unidirec-
tional and bidirectional pumping aids. (a) Examples of “footprints” of seals with pumping features. (b) Seal with
pumpiing features.

There are some concerns with the use of pumping aids. The most significant is that the very action of
providing a pumping aid externally to the seal will help entrain contaminating liquid and pump it into
the sealed system. This places a more exacting role on the exclusion of contaminant.

57.5  Performance Limits


The performance parameters possible with lip seals depend on a combination of the seal design, flexibil-
ity, lip load, etc., but also the seal material and other factors such as the shaft. Any information provided
is, thus, for general guidance.
A major factor is the heat generation under the lip. As the seal continuously runs on a localized sec-
tion of shaft, dispersion of the heat generated by friction is one of the key limiting considerations in the
performance of the seal. The extent of this problem can be demonstrated by considering that a typical
seal on a shaft of 50 or 60 mm diameter will have a power consumption of anything up to 100 W. This
power is dissipated through the seal contact, which will be of the order of 50 mm2. Selecting a seal mate-
rial capable of operating satisfactorily at the bulk oil temperature will cause potentially serious problems
as the seal lip may experience conditions that are up to 40°C higher. This differential may be even greater
in viscous oil, such as transmission applications.
The seal material selection can therefore be critical and have a direct effect on the performance limits
of the seal.
Shaft surface speed will directly affect the amount of heat generated. The seal material characteristics
will also contribute to the lubrication and, hence, thermal limit of a particular material. Figure 57.4 illus-
trates typical speed limits for the elastomer materials used for lip seals. It is also necessary to consider
57-6 Design for Lubrication and Tribology

30,000 15,000 10,000 9,000 8,000 7,000 6,000 5,000 4,500 4,000
40

3500
35 SiL and FPM

3000
30

2500
Shaft surface speed-m/s

25

Shaft speed - RPM


2000
20
ACM

1500
15

10 1000
9
NBR
8
7
6
5 500
4
3
2
1
0 20 40 60 80 100 120 140 160 180 200 – 500 mm
Shall diameter

FIGURE 57.4  Typical performance limits for elastomer lip seals of different materials when running at zero pres-
sure differential.

the life expectancy and also the fluid, particularly the viscosity [8,9]. These performance limits are those
suggested from experience and are based on the requirements for automotive or industrial applications.
It is possible to use the seals outside these parameters, but the life will be considerably shorter.
If a shaft is subjected to appreciable run-out, then the seal will have to flex dynamically to ensure that
the lip follows the shaft. The ability of an individual seal design will depend on the lip design and mate-
rial, as well as any temperature effects on the resilience of the elastomer. The generally accepted limits
of shaft run-out are shown in Figure 57.5. Again, there is a choice to be made, as there can be a trade-off
of reliability against run-out. Figure 57.6 demonstrates the benefits in extending seal life by decreasing
the amount of eccentricity.
A reduction in seal friction provides a double benefit as not only does it become more attractive
because of reduced power consumption, but the reduction of underlip temperature automatically offers
a potential for extended life under the same operating conditions. The energy-saving seal [10] has these
benefits in mind, providing considerably reduced friction, and is available commercially [11].

57.6  PTFE Seals


Lip seals manufactured from filled PTFE are now an important sector of the market, and a variety of
design styles are produced. The earliest designs have a PTFE washer crimped into a spun or stamped
metal housing, Figure 57.7a. This style of seal is still widely used and is particularly suitable for manu-
facturing small quantities of seals. As the use of PTFE lip seals has spread to high-volume industries,
seals that can be produced more economically in higher volumes have been developed. In the bonded
version, one or more PTFE disks are retained in the main seal case through a chemical bonding
Lip Seals 57-7

Total eccentricity
0.7 mm

0.6 mm
1000 RPM
0.5 mm 2000 RPM

0.4 mm 3000 RPM

0.3 mm
5000 RPM
0.2 mm
7000 RPM
0.1 mm

0 mm
0 mm 50 mm 100 mm 150 mm 200 mm 250 mm 300 mm

FIGURE 57.5  Some typical published run-out limits for lip seals.

0.1 0.2 0.3 0.4 0.5 mm


4.0
Average life at 0.254 (“0.010 T.I.R”)
Average life at dynamic runout

3.0

2.0
95% Confidence band

Average life
1.0

0
0.005 0.010 0.015 0.020
Dynamic run-out Inch T.I.R.

FIGURE 57.6  Example of the effect of run-out on seal life.

processes. This may either be by bonding the PTFE directly to the metal case or, alternatively, by
molding elastomer around the case in such a way that it may be used to attach the PTFE disk securely
to the case, Figure 57.7b. The advantages of this compared to the mechanically retained device are
manufacturing efficiency, cost reduction, and some potential for improvement in performance. The
casing may have an elastomer backing and OD similar to an elastomer lip seal, which will facilitate
fitting and sealing of the OD. This construction also allows for the incorporation of elastomer dust
excluders, etc.
The PTFE washer type seals do not have the same sealing mechanism as elastomer lip seals. The
geometry does not present the same sharp lip edge with asymmetric angles as the elastomer seal. The
belled plastic washer forms a PTFE bush running on the shaft. This does not provide the same inher-
ent sealing mechanism as elastomer seals but really forms a close fitting restrictor bush. A typical seal
design will have a PTFE seal contact width of at least 2 or 3 mm; hence, they are sometimes known as
“lay down” seals. The sealing mechanism is enhanced in many applications by providing a spiral thread
form in the rubbing face of the seal. This acts with the same mechanism as a visco-seal, with the seal
returning fluid leaking under the bush to the oil side of the seal. To ensure that the seal can maintain
57-8 Design for Lubrication and Tribology

Elastomer

PTFE

2
3
1

1. PTFE
2. Elastomer
(a) 3. Housing (b)

FIGURE 57.7  Two common methods of construction for lip seals manufactured from PTFE compounds. (a)
PTFE seal champed in a stainless steel housing. (b) PTFE seal bonded to a molded elastomer component.

some static sealing capability, this spiral is often sliced rather than turned. However, for high-volume
automotive seals, it is now common to stamp this feature into the surface of the PTFE. There have been
attempts to improve the static sealing ability by including small “dams” in the spiral feature, but these
have not proved to be completely effective, showing similar leakage rates to a spiral without the features
[12]. As the spiral feature is an integral part of the sealing mechanism, these seals are essentially for one
direction of rotation. If a PTFE lip seal is used without the spiral features, then it can be expected to leak
to a small extent, quoted leakage being up to 0.4 g/h [13].
The spiral can be an effective method of maintaining leak-free dynamic operation, but it will poten-
tially pump contaminant into the lip contact. Exclusion of contaminants from the seal lip area is there-
fore important. There is also a tendency to pump air into the seal, particularly high speeds. This can
cause premature oil degradation, and oil coke deposits in the seal area can inhibit the action of the spiral.
PTFE lip seals have become popular for two types of application:
• Where a lubricating liquid is not continuously available for the seal
• Where the operating conditions are outside the range of elastomer seals, particularly tempera-
ture, fluid compatibility, or the combination of speed and pressure
These seals are now widely used in the automotive industry and also industrial transmissions. This
overcomes the problem of selecting a suitable elastomer for use with modern lubricants that can have
aggressive additive packages and increasing operating temperatures.
The PTFE does not have the inherent flexibility and resilience of elastomer and will therefore not
follow shaft run-out, etc., as effectively. The permissible run-out for a standard PTFE washer type
of seal is less than half that allowed for elastomer seals. The shaft surface finish required is finer,
0.1–0.4 μm Ra, and a hardened shaft may be required. Best results are obtained if the shaft has a sur-
face hardness of 55–60 HRC, although 45 HRC may be used for low-pressure applications and very
clean conditions.
PTFE seals with a spiral have better flexibility than the simple washer type and, hence, can operate
with more run-out.
Various designs have been tried to improve the flexibility of PTFE seals. This includes an elastomer
flex component [14] or a completely revised geometry [15].
Lip Seals 57-9

As an alternative, PTFE lined seals are manufactured by the bonding of a more flexible PTFE ele-
ment to the contact surface of elastomer oil seals. This is carried out as one molding/bonding process,
and the advantages compared to the PTFE washer seals are considerably improved dynamic seal per-
formance on high-speed shafts and significantly reduced specific lip force, leading to reduced frictional
heating [8]. However, new designs of PTFE seals are being introduced with reduced friction as a prime
objective, [15,16].

57.7  Excluders
The vast majority of lip seals are used to seal a clean lubricant within a system. However, it is also neces-
sary to keep the external environment away from the seal and, ultimately, the lubricant. A very signifi-
cant part of the lip seal overall system can be exclusion of contaminant. For sealed rolling bearings, this
is the overriding concern, with retention of the lubricant a secondary consideration.
There are two factors to consider as a starting point with excluders for lip seals:
1. It is well established that a lip seal will pump into the lubrication system. An inward-facing seal
will therefore potentially entrain dust. It is also more likely to entrain any liquid and pump it into
the bearing or lubrication system.
2. The amount and type of contaminant to be excluded. The potential contaminants can be broadly
classified as follows:
a. Dry: Fine dust, sand, and large particles
b. Liquid: Typically water and aqueous solutions
c. Slurry: Mud, slurry mixtures, etc.
Within any of these categories, the contaminant may be of variable quantity and the types of solid
will have varying degrees of abrasiveness. The exclusion may be required for both static and dynamic
situations. A wide range of excluder types are used with lip seals, with the selection depending on
the severity of the application, operating conditions, and life expectancy. They can be broadly clas-
sified as
a. Radial dust lip
b. Multiple radial dust lips
c. Axial dust lips
d. Multiple radial and axial lips
e. Labyrinth and flinger features with any of the previously listed types
Virtually any form of excluder involves providing a seal external to the lip seal. Because of the inward
pumping action of the lip seal, this creates a potential problem as the volume between the seal and
excluder will be gradually evacuated by the pumping action of the seal.
The simplest form of excluder is to use a radial dust lip, molded into the heel of the seal, Figure 57.8.
These are intended to protect the main seal lip from a light dust environment. It is necessary to ensure
that this lip is not too effective as a seal to avoid problems of a vacuum between it and the seal.
As an alternative, a felt dust lip has been found effective as it will filter dust particles but allow the
space between the seal and dust lip to breath.
Axial dust lips are popular as part of the exclusion system. These will normally be in contact with a
flange section of the shaft, probably as part of a cassette arrangement. With these, the seal manufacturer
supplies a shaft sleeve and often an excluder system designed to be appropriate to the application. Examples
are shown in Figure 57.9. A shaft flange can provide a centrifugal effect to any contaminant in the seal area
and will disperse much of any contaminant present when the shaft is rotating. There is also the tendency to
disperse any wear debris created from the dust lip, which further helps to keep the seal area clear.
It is also possible to fit the wiper lip to the rotating flange, Figure 57.9b. The rotation of the flexible
sealing element will subject it to centrifugal force, which will tend to lift the seal from contact during
57-10 Design for Lubrication and Tribology

FIGURE 57.8  Simplest form of excluder, a simple dust lip on the atmospheric side of a seal.

(a) (b)

FIGURE 57.9  Two examples of more effective excluders: (a) a cassette with a radial metal flange and axial dust lip
on the seal and (b) a cassette design that has a rotating axial seal lip molded to the shaft sleeve.

rotation. The excluder lip can therefore be designed to have a higher contact stress when static, to pre-
vent ingress of contaminant, but will operate primarily as a centrifugal seal when the shaft is rotating
at high speed.
For slow speeds and very contaminated environments, such as agricultural and earthmoving equip-
ment, more extensive exclusion systems are required. They will usually involve a multiple arrangement
of dust lips combined with a labyrinth, to impede contaminant access, and radial flanges to provide
centrifugal expulsion. The interspaces between the lips will be lubricated. The heat generation with this
type of sealing system will be a crucial factor, and, hence, speed is limited.

57.8  V-Ring Seals


The V-ring is essentially an axial lip seal. The seal is a form of elastomeric lip seal that seals on a radial
face rather than axially along the shaft. A typical seal is shown in Figure 57.10.
The seal is normally stretched to fit directly on the shaft with the elastomer tension then providing
sufficient friction to drive the seal. It will rotate with the shaft and seals axially against a radial face.
Lip Seals 57-11

FIGURE 57.10  V-ring used as a bearing seal.

This radial face can be a bearing race, bearing housing, a stamped washer, or even the metal casing of a
lip seal. The seal fitting dimensions are such that a very light lip loading is created. The seal may therefore
run dry in many applications with a low friction, potentially only 25%–30% of a standard lip seal. The
configuration of the lip is such that as speed increases, centrifugal force will tend to reduce the seal lip
interference, which enables the seal to operate at high speeds without excessive heat buildup. Above a
speed of 15–20 m/s, the seal lip has lifted off the counterface and the seal then operates as a clearance seal.
The entire seal is subject to centrifugal force as the shaft rotates. Above certain speeds, which depend
on diameter, some form of radial retention will be required. For general-purpose sizes, the limiting
speed for operation without radial retention is 10–12 m/s. For very large sizes, such as 1–2 m diameter,
radial retention should always be used. This can be achieved by designing a suitable housing to retain
the seal or using a clamping band.
Among the potential advantages of this design of seal compared with a radial lip seal are the following:
• Shaft surface finish less critical
• Will operate with considerable run-out and misalignment
• Running surface is less critical
• High speed capability
• Axial shaft movement accommodated by seal flexing
Typical surface requirements for these seals are
• Radial surface finish: 0.4–0.8 μm Ra
• Flatness: 0.4 mm per 100 mm diameter
In certain high-speed applications, the seal can be mounted statically to operate against a radial face.
In this arrangement, the seal contact stress will not reduce as speed increases, so this would normally be
used in liquid-lubricated application.
57-12 Design for Lubrication and Tribology

V-ring seals are widely used as bearing seals in small high-speed applications such as electric motors
and other bearing seals in conditions of light contamination. For more arduous applications, they are
widely used as an excluder seal on the atmospheric side of a lip seal.

57.9  Bearing Seals


Often, seals are fitted directly within rolling element bearings. They are normally more compact than
separate seal bearings, with integral seals being lubricated with grease on assembly by the bearing manu-
facture. The primary purpose is in fact the exclusion of contaminant. In the majority of cases, the seal will
be designed to seal the bearing from the external environment. It may be expected to leak a small amount
of the grease but, in so doing, lubricate the seal and assist with contaminant exclusion. The selection of a
bearing seal presents a compromise, minimizing friction while providing adequate exclusion.
They may comprise a very-simple-looking design that is an elastomeric washer with a light lip contact
on the inner race or a more complex design that has both sealing lip and a wiper lip. In the vast major-
ity of cases, the seal is designed such that the lip is outward facing. This encourages the bearing grease
under the lip to counteract ingress of contamination.
Additional seal types have been developed for deep groove ball bearings, referred to as the low-fric-
tion seal. These are designed to have very low contact pressure to fulfill the demand for adequate sealing
and low-friction operation.
These seals are used in bearings for arrangements where contamination is moderate and the presence
of moisture or water spray, etc., cannot be ruled out or where a long service life without maintenance is
required.

57.10  Lip Seals for Pressure


As discussed earlier, standard lip seals are only suitable for use for pressures up to 0.3–0.5 bar. Increasing
the pressure will distort the lip and both disturb the orientation of the seal lip angles and increase the
contact area. This, together with increased lip loading from the pressure, will create additional friction
and, hence, heating of the lip and limit seal performance.
Specific designs of lip seal are available for operation up to 10 bar. The reasons for selecting a lip seal
for pressurized applications may include the following:
• Compact seal envelope
• Economic seal assembly compared with alternatives
• Conformability of elastomer can provide improved sealing at large diameters, for example, in
excess of 0.5 m, compared with alternatives
Lip seals are popular for a number of specific applications where the benefits listed earlier are major
considerations. These include the following:
• Shaft seals on fluid power hydraulic pumps and motors where they may seal a casing pressure of
1–7 bar.
• Shaft seals on hydraulic motors in agricultural and mechanical handling where circuit design
may create higher pressures, up to 150–200 bar.
• Vehicle steering column input shaft seal for hydraulic power steering racks. These seals are
required to have low friction to avoid excessive steering effort and may operate at low pressure
during most steering but up to 70 bar at full load. Shaft surface speeds are essentially very low.
• Ship stern shaft seals where a series of seals will seal externally between sea and bearing lubricant
and internally between the lubricant and machinery space. The lubricant is typically pressurized
to counteract the head of sea water on the external seal, with the internal seal running at the pres-
sure of the head of oil.
• Water turbine shafts under a variety of conditions.
Lip Seals 57-13

(a) (b)

FIGURE 57.11  Two examples of lip seals used for higher pressures: (a) a supported lip and (b) a shortened and
strengthened lip flex section.

To resist the distortion of the seal lip, two alternatives can be considered:

• Standard lip seal design but with the use of a lip support, Figure 57.11a
• Revised seal design with a stiffer lip, Figure 57.11b

A number of techniques are used to resist distortion including short thicker elastomer section, exten-
sion of the metal casing into the flex section of the lip, and redesign of the radial flange of the seal to
support the lip.
The use of the support lip will reduce the potential of the seal to accommodate shaft run-out, etc.,
and so it will require more attention to the installation tolerances. The stiffer lip designs will also reduce
flexibility and require improved shaft alignment compared with standard seals.
The lip contact pressure will still be increased, and this will limit the speed capability of the seals. As
pressure is increased, the permissible speed is reduced. The permissible speed will depend on both the
seal design and also the material. Examples of typical recommendations are given in Figure 57.12. There
is some difference in capabilities of the individual designs from the suppliers. At higher pressures, only
very slow or intermittent speeds are permissible.
The ability of the elastomer material to operate at high temperature will also affect the performance
of the seal and so an FKM seal may be expected to operate at up to twice the speed of NBR for a given
pressure condition. It is important to consider the resistance of the arrangement to dislodging of the
seal by pressure and also the potential for static leakage around the seal. Normally, it will be necessary
to have a flange or circlip to retain the seal in the housing. A seal with an elastomer outside diameter is
advisable to ensure an adequate static pressure seal.

57.11  Seals for Fluctuating Pressure Applications


Although the stiffer profile of the lip seals designed for operation at elevated pressure will resist distor-
tion, the lip design itself still promotes a relatively high lip load that will vary significantly with toler-
ances. The effect of pressure is still to increase the tendency of the seal to reverse pump the oil and reduce
lubrication. One approach is to design the lip to encourage lubrication, especially at lower pressure. The
profile of the sealing edge has been designed so that the contact surface angle to the shaft on the air side
is larger than that on the oil side at zero pressure. At the same time, an artificial groove has been added
to the sealing edge profile in the area where the most wear has been noted previously. This addition min-
imized the risk of wear and the buildup of damaging particles. As pressure increases above 3 bar and up
to 7 bar, a more conventional seal profile is presented and lubricant will be returned across the seal [17].
57-14 Design for Lubrication and Tribology

1.0

0.9

0.8 Ø 25

0.7 Ø 50

Ø 100
Pressure, MPa
0.6
Ø 150
0.5

0.4

0.3

0.2

0.1

0
1000 2000 3000 4000 5000 6000 7000
Shaft speed, min–1

FIGURE 57.12  Typical PV limits for lip seals being used at pressure. The actual limits will be very dependent on
seal design and material plus the shaft design.

In a hydraulic pump, the pressure is usually changing constantly between 0 and 7 bar. This means
that the sealing edge pumping direction of the new design of seal is always changing. At low pressure,
the sealing edge transports from the inside to the outside while the opposite is the case at high pressure.
This means that oil is constantly being exchanged in the seal contact zone, preventing the deposit of
carbonized oil and dry running. The typical groove wear is also reduced to a minimum. This design of
seal can therefore be considered for applications where there is a constantly varying pressure profile up
to the pressure limits of the seal.

57.12  Plastic Lip Seals for Pressure Applications


Plastic lip seals can be used for a wide range of applications. The type of seal used for pressure applica-
tions is usually the plain washer design. The spiral groove design will be prone to leakage under pres-
surized conditions. The additional strength of the PTFE material means that it will withstand pressure
without excessive distortion. The washer type lip seal can be used up to a pressure of 30 bar. There is
still the problem that exists with elastomer rotary seals of heat generation under the lip, so as pressure
increases, the speed capability decreases. The maximum combination of speed and pressure recom-
mended for these seals is 100 bar × m/s [13]. So, for instance, at 30 bar, the speed would be limited to
3.3 m/s. A further potential limiting factor is the design of the hardware, particularly the shaft, as this
will influence dispersion of the frictional heat generated. Frequency of stop/starts and fluctuations in
system pressure may also affect the performance as these will inhibit the bedding in of the seal and
accelerate wear.
There are two significant factors that can influence the selection of a PTFE lip seal for pressurized
applications:
• The seal will tend to leak more than an elastomeric seal. Typical leakage rates may be between
0 and 0.4 mL/h [13]. It may be necessary to arrange additional leakage collection measures such
as a backup seal.
Lip Seals 57-15

• As the seal is not dependent on the sealed fluid for lubrication, it may be used to seal nonlubri-
cating fluids such as water and mild chemicals or dry gas. In this case, the boundary lubrication
properties of the seal and shaft combination are critical and selection and finish of the shaft is as
important as the seal.
For nonlubricating liquids such as water, it is important to have a hard shaft surface, 50–60 Rockwell
C, a surface finish of 0.2–0.3 μRa, and a polished texture. For dry operation, a fine surface in the range
0.1–0.2 μRa is preferred. The reduced heat dissipation for dry applications must also be considered.

57.13  Seal Materials


The seal material selection is critical to reliable operation. For elastomer lip seals, as explained in the
dynamic sealing mechanism section, it is crucial that a polymer specifically compounded for lip seals
is used to ensure that the microgeometry is generated on the surface of the elastomer as it beds in. It
is then necessary to select the polymer with respect to operating temperature, speed (and hence heat
generated at the seal lip), and the sealed fluids. It is essential to consider all the fluids to which the
seal may be exposed. The most obvious is the sealed fluid, such as oil or grease, but it may also include
cleaning fluids, pressure washing, road salt, and one of the most aggressive hot air. The development of
synthetic lubricants and long oil change intervals together with environmental constraints on heavy
metal compounds has led to the development of many new additives for lubricants. These can be chemi-
cally aggressive and cause rapid degradation of elastomer compounds. It is therefore essential that the
elastomer compound is selected by testing with the actual lubricant to be used in a system. Conversely,
if a system lubricant is to be respecified, then seal compatibility must be considered.
The most common materials used for lip seals are described later, and this will give a general guide to
which material may be appropriate for a given application. However, the properties of individual com-
pounds from different manufacturers can vary widely, even within the same polymer group, so close
liaison with the seal supplier is required.

57.13.1  Nitrile
Nitrile (NBR) compounds are very widely used for general-purpose sealing because they have good
mechanical properties, reasonable resilience, good wear properties, and resistance to most mineral-
based oils and greases. Seal properties vary quite widely, depending on the acrylonitrile content, and so
the grade of material will have to be selected to achieve the correct compromise of properties. The cure
system will also impact on the temperature range. Peroxide-cured material has a higher temperature
range but generally poorer mechanical properties and these have generally proved to be unsuitable for
lip seal applications. The major limitations of NBR for lip seals is the performance limitation created by
the limited temperature range and lack of resistance to the additives used in many modern lubricants.
Typical temperature range: minimum temperature −20°C to −45°C depending on compound.
Maximum general temperature: +100°C, up to 120°C for short periods.

57.13.2  Hydrogenated Nitrile


Hydrogenated nitrile (HNBR) is a development of conventional nitrile polymer in which the polymer
chain is subjected to a hydrogenation process that has the effect of improving the resistance of the
polymer chain to chemical attack. This provides a material with improved chemical and temperature
resistance and also good mechanical properties. A well-developed HNBR will have both a high strength
and high elongation to break, making it a very useful material. However, it will have a higher cost than
standard NBR materials. It is popular for lip seals in heavy industries such as metal manufacturing
where good fluid and wear resistance are required [18].
57-16 Design for Lubrication and Tribology

Temperature range: −20°C to +150°C. Special compounds can provide flexibility to −40°C usually at
the expense of other properties.

57.13.3  Polyacrylic
These are oil-resistant elastomers that can be used with automotive engine and transmission lubri-
cants at a much lower cost than fluorocarbons. It has good resistance to mineral oils including those
with high-additive packages such as transmission lubricants, including automatic transmission fluids
(ATFs). The high-temperature capabilities also give better air and ozone resistance than NBR. It has
relatively low resilience and poor low-temperature properties. It is used for some rotary shaft lip seals,
but a common use is as the static elastomer cover on seals with an FKM or PTFE lip.
Temperature range: −15°C to +150°C.

57.13.4  Ethylene Acrylic


This has improved low-temperature flexibility compared with polyacrylic material (ACM). The low-
temperature properties are quite good. Again, it may be used as the outer static elastomer on a rotary
shaft lip seal casing where fluorocarbon is used for the dynamic lip.
Temperature range: −30°C to +150°C.

57.13.5  Silicone (MVQ, MPQ, MPVQ)


The main attribute of silicone elastomers is the wide temperature range. They can be used from below
−50°C to in excess of 200°C. The major problem is the relatively poor tensile strength and tear resistance
which in turn makes them susceptible to wear and also more prone to damage on assembly. Care is often
required due to the limited mechanical properties which can be further reduced by a higher swell than
some other elastomers. In lip seals, this can affect the geometry of the lip and, hence, seal performance.
They are used for lip seals because of the superior low-temperature flexibility.
Temperature range: −60°C to +200°C

57.13.6  Fluorocarbon
Fluorocarbons (FKM) are able to provide a high-temperature oil-resistant elastomer. They are capable
of providing a wide chemical compatibility at temperatures up to 200°C. A limiting factor has been the
limited low-temperature performance, but this is progressively improving. There are several different
fluorocarbon groups that have quite different properties and are often not interchangeable. Fluorine in
the polymer chain forms a very strong bond within the chain against fluid attack. The original devel-
opment was a copolymer, also known as a dipolymer, of hexafluoropropylene (HFP) and vinylidene
fluoride (VDF). Terpolymers use tetrafluoroethylene (TFE) in addition to the HFP and VDF, which
increase fluorine content to 68%, providing improved fluid resistance but inferior compression set and
low-temperature performance. These materials are have traditionally not been serviceable below about
−10°C. Development of FKM materials is extremely active, and new materials with improved properties,
especially low-temperature performance, are becoming available [19] and are being specified for lip seals
with a low-temperature duty requirement.
Temperature range: −20°C to −40°C (depending on compound) to +200°C

57.13.7  Polytetrafluoroethylene
Polytetrafluoroethylene (PTFE) has virtually universal chemical resistance and a very low coefficient of
friction with excellent dry running properties. Seals will usually be made from sintered PTFE prior to
Lip Seals 57-17

TABLE 57.1  Examples of the Range of Potential Fillers That May Be Used in PTFE Rotary Shaft Seals
Filler Typical Applications Potential Shaft Material
None, virgin PTFE Low pressure, light duties easily extrudes Steel, chromed steel, cast iron,
at temperature widest fluid resistance stainless steel, nickel alloys
Glass fiber/MoS2 Mineral oils, synthetics Steel, hardened chromed steel
Graphite fiber Medium duty, high wear resistance Steel, hardened
Carbon/graphite Mineral oils, aqueous fluids, steam, dry Steel, chromed steel, stainless
gas, pneumatics
Carbon fiber Mineral oils, aqueous fluids, synthetic Steel, chromed steel, cast iron,
fluids, and phosphate esters stainless, aluminum, bronze alloys
Suitable for soft counterfaces
Good extrusion resistance
Polymer (usually Ekonol) Mineral, synthetic, and aqueous Steel, chromed steel, cast iron,
lubricants stainless steel
Dry gas, relatively low abrasion

TABLE 57.2  Examples of Temperature Limits of Lip Seal Materials


Material Type Minimum Temperature (°C) Maximum Temperature (°C)
Nitrile (NBR) −20 to −45 100
Polyacrylic (ACM) −15 150
Ethylene acrylic (AEM) −30 150
Hydrogenated nitrile (HNBR) −20 to −40 150
Silicone
Fluorosilicone (MVQ, MPQ, MPVQ) −60 200
Fluorocarbon (FKM) −20 to −40 200
PTFE −200 260
Note: It is important to note that these temperature ranges provide the maximum permissible range of operation
including the underlip temperature of the seal; bulk fluid temperatures will be considerably less. Seal life at these
temperatures may be very limited; see References 8 and 9. Upper temperature limits are also dependent on the
seal-fluid compatibility. The lowest temperature capabilities of NBR, HNBR, and FKM elastomers are only available
from a limited number of premium-grade polymers to special order.

machining, but molding is also possible for high-volume applications [20]. A range of fillers are used to
modify the material properties. These are important for both mechanical and tribological properties.
Extreme care is required to ensure the filler used is suitable for rotary seals and the specific fluid and
shaft material as, otherwise, very high seal or shaft wear can occur. Examples of fillers that may be used
are given in Table 57.1.
Temperature range: −200°C to 260°C.
The overall temperature range for each of the material categories above is summarized in Table 57.2.

References
1. ISO 6194-2 Rotary shaft lip type seals—Part 2: Vocabulary.
2. ISO 16589-2: Rotary shaft lip-type seals incorporating thermoplastic sealing elements—Part 2:
Vocabulary.
3. Jagger, E.T., Rotary shaft seals: The sealing mechanism of synthetic rubber lip seals running at atmo-
spheric pressure, Proceedings of the IMechE, 1957, 171, 597–616.
4. Kammüller, M., Zur abdichtwirkung von radial wellendichtringen, Dr-Ing thesis, Universität
Stuttgart, Stuttgart, Germany, 1986.
57-18 Design for Lubrication and Tribology

5. Salant, R.F., Soft elastohydrodynamic analysis of rotary lip seals, Proceedings of IMechE Part C:
Journal of Mechanical Engineering Science, 2010, 224(12), 2637–2647.
6. Flitney, R.K., Advances in our understanding of elastomer dynamic seals, Proceedings of 8th EDF
LMS Workshop, Dynamic Sealing in Severe Operating Conditions, Poitiers, France, October 2009.
7. Schmuker, S. and Haas, W., Effect of the machining method of shaft surface on rotary lip seal,
Proceedings of 19th International Conference on Fluid Sealing, September 25–26, Poitiers, France,
pp. 134–146, 2007.
8. Flitney, R.K., Seals and Sealing Handbook, 5th Edition, Elsevier, the Netherlands, 2007, pp. 103–148.
9. Simrit, Radial Shaft Seal Technical Manual, www.simritna.com
10. Heiland, M. and Bock, E., Simmering ESS—Energy saving seal, Proceedings of 14th International
Sealing Conference, October, Stuttgart, Germany, pp. 78–86, 2006.
11. Maximising long-term protection and minimising energy consumption for agricultural and con-
struction machinery, Sealing Technology, August 2009.
12. Bauer, F. and Haas, W., PTFE lip seals with spiral groove—The penetration behaviour, hydrodynamic
flow and back pumping mechanisms. Proceedings of 18th International Conference on Fluid Sealing,
October 12–14, Antwerp, Belgium, pp. 303–316, 2005.
13. Duhring, B. and Iverson, G., The application of plastics in dynamic seals, Proceedings of IMechE, Part
J: Journal of Engineering Tribology, 1999, 213, 215–226.
14 Toth, D.M. and Tripathy, B.S., Elastomeric hinged seal for improved flexibility, World Patent
2005/069764, August 2005.
15. Low-friction shaft seal improves fuel economy and reduces emissions, Sealing Technology, 2010,
2010(4), 1–16.
16. Bock, E., Raclot, I., and Lutaud, D., POP: Friction reduced PTFE engine oil seals, Proceedings of 16th
International Sealing Conference, October 12–13, Stuttgart, Germany, pp. 553–562, 2010.
17. Bock, E., Vogt, R., and Schreiner, P., New radial shaft seal concepts for sealing hydraulic pumps and
motors, Sealing Technology, 2003, 11, 6–10.
18. Embury, P. and Armour, J., Influence of elastomer compound design on the performance of rotary
shaft lip seals, Sealing Technology, 2004, 2004(8), 7–10.
19. Farrow, P. and Merli, F., New low-temperature FKM for effective sealing in the chemical industry,
Sealing Technology, 2010, 2010(1), 8–12.
20. Stern, C., Maier, M., and Schlipf, M., Innovative seal solutions with New PTFE processing tech-
niques, Proceedings of 16th International Sealing Conference, October 12–13, Stuttgart, Germany,
pp. 282–290, 2010.
58
Brake and Clutch*
58.1 Introduction.....................................................................................58-1
58.2 Brakes as Kinetic Energy Dissipaters...........................................58-2
58.3 Brake Types and Design.................................................................58-3
Friction Relationships  •  Coefficient of Friction
Roberto C. Dante 58.4 Clutch Types and Design................................................................58-7
Universidad Nacional 58.5 Brake Noise and Vibrations...........................................................58-8
Autónoma de México Noise Onset  •  Noise and Vibration Damping
58.6 Friction Materials’ Compositions, Processes,
Carlo Navire and Properties................................................................................ 58-10
Isibond S.a.s
58.7 Friction Couple: Brake Pads and Rotors....................................58-13
Bruno Tron 58.8 Future Trends.................................................................................58-15
NVH Advisor References................................................................................................... 58-16

58.1  Introduction
This chapter deals with brake and clutch materials, which in general are called automotive friction
materials. Although brakes and clutches have opposite purposes—the former dissipates kinetic energy,
while the latter transmits motion—basically, the friction materials used for both applications are con-
ceptually similar so that they can be described together and the same definitions can be used with
certain latitude. For this reason, the brake friction materials are described in detail, and the same basic
concepts can be extended to clutch friction materials. The fundamental brake concepts and the most
general equations related to brakes are treated and presented in order to provide the readers with the
general principles governing any type of brake. Basic principles of both brake and clutch design are
presented and discussed.
Some of the most important challenges concerning brakes are noise, vibrations, and harshness
(NVH); considerable effort was made in the last two decades to determine the origin, the mechanism of
transmission, and damping of brake acoustic waves. The analysis of brake noise and vibration is the first
step to understand this phenomenon and find adequate solutions.
Tribological aspects concerning disc and friction materials are presented for dry friction, illustrat-
ing the main interface characteristics such as surface oxide film formation and material transfer [1–3].
Automotive friction materials are composite materials, and their blends contain five essential classes of
components: abrasives, lubricants, fillers, metals, and binders. The terminology and definition of fric-
tion materials’ classes are used with latitude and many times in a misleading way. However, in certain
cases the absence or the presence of a fundamental raw material can characterize the friction material
in a marked manner.

* In memoriam of Francesco Cinque.

58-1
58-2 Design for Lubrication and Tribology

The future trends for automotive friction materials are mostly led by NVH concerns, weight reduc-
tion, environmental impact with special emphasis of heavy metal content reduction, and material wear
(loss of material into the environment) reduction.

58.2  Brakes as Kinetic Energy Dissipaters


Vehicles, and machines in general, need to decrease their velocity or to stop their motion in a controlled
manner. Brakes allow these machines to decelerate in a controlled manner with a deceleration greater
than that provided spontaneously. There are many types and concepts. The function of brakes is to con-
vert kinetic energy into other forms of energy or to prevent the transformation of potential energy into
kinetic energy (stationary brakes).
The most ancient and primitive brakes (e.g., for carriages), made with friction materials based on
common natural products such as leather and wood, utilized friction to convert kinetic energy into
heat. In spite of the many types of existing brakes, they have a common conceptual framework: a force F,
which may have different origins such as mechanical, magnetic, etc., intervenes and results in a friction
force FT, which depends on F, and it is opposite to the direction of motion.
The following equation of energy conservation can be utilized for all dissipative systems including
brakes:

ek + ∑ u = 0,
j
j (58.1)

where
ėk is the kinetic energy dissipation rate (which is negative)
u̇j are the rates of the different types of energy produced by dissipation, for example, heat production

The kinetic energy Ek dissipated in a defined time interval t can be defined as

∑ ∫ u dt = − ( E
j
j
t
k − Ek0 ) . (58.2)
0

In general, the production rates of the converted energy can be expressed in the following way:

 
∑ u = −v ⋅ F (F ),
j
T (58.3)

where
v⃗ is the velocity
F T⃗ (F) is the aforementioned friction force
FT(F) takes the following form:

FT (F ) = a ( X )F , (58.4)

where α(X‾ ) is a factor dependent on a set of variables X‾ , and the force F that causes FT to arise.
In the case of dry friction between surfaces, α(X‾ ) becomes μ, the coefficient of friction, and F is W
the normal load:

FT (W ) = a ( X )F = mW . (58.5)
Brake and Clutch 58-3

It is noteworthy to point out, as reported in previous chapters, that the coefficient of friction is not
constant [4], as stated by the third law of friction, and it is affected by many variables such as speed,
temperature, and load.
In dry friction brakes (drums and disc brakes), the kinetic energy is converted mainly into heat but
also into other types of energy such as acoustic and mechanic waves, triboreactions, mechanical work to
produce wear, heat of phase transitions, and at high temperatures, infrared radiation, as well as visible
light, is emitted by metallic surfaces of rotors.
Equation 58.1 can be expressed more explicitly by indicating some types of energy conversions
involved in disc brakes, such as heat production q̇, chemical affinities of different triboreactions Ai with
positive reaction rate ξ̇ i, while other types of triboemissions are associated to other types of energy flows
u̇j, such as those due to triboelectrons, acoustic emissions, wear loss, etc.:

ek + q + ∑x A + ∑ u = 0.
i
i i
j
j (58.6)

In Equation 58.6, the dominating dissipative term is doubtless heat production, usually representing
more than 90% of the amount of kinetic energy dissipated, and for this reason brakes should also work
as efficient heat exchangers in order to avoid temperature increase to critical values that can affect brake
efficiency in a dramatic way. Indeed, the power dissipated by brake pads can be of the order 1–2 kW
per square centimeter for many vehicles. It is noteworthy to point out that the first four equations (58.1
through 58.4) are valid for any type of brake, even for magneto-rheological fluid brakes, which are based
on viscosity enhanced by orientation of ferromagnetic particles by means of a magnetic field [5].
Although friction materials are used also for clutches, in this case they have an opposite purpose; in
fact, clutches have the function to transmit kinetic energy and not to dissipate it. So engineers will try
to minimize each dissipative term u̇j and the entire rate of dissipation ėk, while preserving the kinetic
energy and maintaining a certain torque.
The basic principles governing clutches will be provided in Section 58.3 together with design
features.

58.3  Brake Types and Design


There are different types of brakes [6–8]. We can distinguish between two main classes: service brakes
and stationary brakes. The service brake is applied during driving or machine operation, while the
stationary brake only keeps a vehicle stopped during parking, so that a separate mechanical system is
needed for stationary brakes without fluids in order to avoid any failure due to leakage. Usually, the
typical system to be decelerated is a rotor with the exception of a few applications (e.g., lift): the rotor
can be a drum or a disc. The rotor counterpart is the friction material supported on a back plate, called
brake pads for disc and shoes for drums. The average coefficient of friction, usually accepted in Europe,
is around 0.4, while in the United States, a coefficient of friction between 0.35 and 0.37 is preferred.
A common and easy way to increase torque is to increase the number of friction surfaces; for example,
the friction torque is divided among several discs in a multiple disc brake for heavy-duty application, in
order to avoid excessive pressure.
In the case of a disc, the actuator can be a caliper or an axial piston (aircraft and industry machines).
In the case of a drum, there are many solutions: (1) external shoes (railways and industry machines),
(2) inner shoes (ground vehicles and industry machines), (3) brake bands that cover almost all the cir-
cumference of the drum (industry machines). In fact, there is another class of brakes called wet brakes,
where there are several alternated discs of steel and friction material merged in a fluid lubricant. During
braking application, the discs are packaged and the thin fluid film generates a drag force under bound-
ary lubrication conditions.
58-4 Design for Lubrication and Tribology

58.3.1  Friction Relationships


58.3.1.1  Disc Brake
The disc brake is widely used in ground vehicles, especially on passenger cars. There are two basic
types:

1. The sliding disc brake (Figure 58.1a), in which the pressure force is supplied by one or more pis-
tons on the side of the caliper. The caliper itself can slide over two pins so that the braking force is
transferred also to the brake pads opposite to the piston.
2. The fixed disc brake (Figure 58.1b), in which the pressure force is supplied by one or more pistons
on both sides of the caliper.

The braking torque Mf is calculated according to the following formula:

M f = mNRm , (58.7)

where
μ is the friction coefficient
N is the clamping force due to the piston
Rm is the average radius of the brake pad to which the friction force FT is applied

Compared with a drum brake, it requires a higher pressure but is simplest to build (even if more expen-
sive) and has a better and more uniform braking action due to good dissipation of thermal energy.

58.3.1.2  Ball Ramp Brake


A special type of disc brake is the ball ramp brake (Figure 58.2), particularly used for agricultural and
construction applications, and it consists of a ball ramp expander mechanism. This device applies a
pressure force against an array of discs (friction plates and steel counter plates), which are respectively

Pistons

Caliper

Brake pads —
friction material

Rotor (brake disc)

FIGURE 58.1  Sliding disc brake (left) and fixed disc brake (right).
Brake and Clutch 58-5

B
Ra

Force’s line of action φ


on the connecting rod

T
F
0
o o
B
J
F

Line of action of B–B section


Rb
the control force F
o

B Ra : Rod’s radius
Rb : Rolling ball’s radius
N α : Ramp’s angle
β : Angle between shaft and connecting rod
X : Complementary angle
α φ : Ball’s diameter

FIGURE 58.2  Sketch of a ball ramp brake with parameters.

engaged with the drive axle and the housing of the machinery. The braking torque is calculated accord-
ing to the following equation:

2F0 Ra
Mf = , (58.8)
⎡ tan a Rb ⎤
cos b ⎢ ⋅ − 1⎥
⎢⎣ tan f Ra ⎥⎦

where
F0 is the force applied to the rod
Ra is the rod radius
α is the inclination radius of the ramp
ϕ is the friction angle so that tan ϕ is the friction coefficient μ
β is the angle between shaft (brace) and the connecting rod
Rm is the average radius of the friction lining
Rb is the rolling balls’ radius (see Figure 58.2)

58.3.1.3  Drum Brake


The drum brake is normally fitted on the rear axle of small cars, trailers, and trucks. The self-­energizing
shoe is forced into the drum by the movement of the drum itself. A self-energizing shoe is mounted
so its friction surface leads or is ahead of its pivot point; the de-energizing shoe has the opposite
behavior. With reference to Figure 58.3, servo and self-energizing action is produced by hooking the
58-6 Design for Lubrication and Tribology

Anchor pin
Heel

on
ati
rot
Piston

el
he
W

Self-energizing
(primary)


Hub


De-energizing
(secondary)
Adjusting
device

Toe

Vehicle direction

FIGURE 58.3  Sketch of drum brake.

heel of the primary shoe to the toe of the secondary shoe. When the wheel cylinder forces the top ends
of the shoes against the revolving brake drum, it will try to carry the forward shoe around. As the
primary shoe attempts to revolve, it will jam the secondary shoe against the single anchor pin. This
stops both shoes and produces a binding effect that actually helps the shoes apply themselves. This
servo and self-energizing action reduces the amount of needed pedal pressure.
Note how the primary shoe in Figure 58.3 attempts to rotate in the direction of the drum. Since the
adjusting screw connects it to the toe of the secondary shoe, the heel of the secondary shoe is jammed
against the anchor pin. The arrows illustrate the braking force direction. When the vehicle is traveling
in reverse, the secondary shoe becomes the primary shoe.
The torque can be expressed in the following way for each shoe:

M f = 2m p0 R2bb, (58.9)

where
p0 is the pressure on the shoe surface
R is the drum’s inner radius
b is the width of the brake shoe
β is half of the angle of the brake shoe (see Figure 58.3)

58.3.1.4  Band Brake


For a band brake, a steel band is wrapped around a drum and it is tightened by means of a lever.
Normally, one of the ends—the one toward the incoming motion (see the arrow indicating rotation
direction of the drum)—is fixed to the frame of the machinery, because there is major tensile strain
(Figure 58.4) and the control force acts where the tension is lower.
Brake and Clutch 58-7

T1
Q

b
L

T0

FIGURE 58.4  Sketch of a band brake.

Therefore, this type of brake band can be used when the drums rotate in one direction. In cases where
the drum can rotate in reverse, another type of control lever must be used.
When the total lining lap angle is ϑ and T1 and T0 are the tensile forces on the ends of the brake
band, then

T0 = T1e mJ , (58.10)

where μ is the coefficient of friction (Figure 58.4). When R is the drum radius, the following equation
can be written for the braking torque Mf :

M f = (T0 − T1 ) R. (58.11)

58.3.2  Coefficient of Friction


The friction coefficient depends on many variables, and there is a great difference between dry or wet
applications. In dry applications, the coefficient of friction lies between 0.3 and 0.5, while in wet applica-
tions it lies between 0.05 and 0.2. In wet applications, a traditional organic friction material exhibits a
steady coefficient of friction at temperatures between 100°C and 200°C. Above 200°C, the friction coef-
ficient begins to decrease, and above 350°C/400°C, the friction material can be irreparably damaged.
Sintered metal materials, such as those based on copper alloys, work better at high temperatures, but,
at low temperatures, the friction coefficient is lower than that of an organic material.
For wet applications, the operation temperature can reach a maximum temperature of 180°C. The
limit is mainly due to the boiling point of oil. The oil for wet applications has to be chosen accurately
because of the different additives used by manufacturing companies. The working pressure range is
normally between 0.5 and 1 MPa; however, some materials can operate up to 3 MPa.

58.4  Clutch Types and Design


The function of a clutch is to transfer a torque between two systems at different velocities. The rotor in
this case is a metallic disc made of either steel or gray cast iron; the counterpart is the friction material
58-8 Design for Lubrication and Tribology

that also has a disc shape. Clutches can be dry or wet. Moreover, the multiple disc solution is utilized
both for dry and wet clutches, which is a common and easy way to increase torque.
The transmittable torque is

De − Di
M f = nmPh , (58.12)
4

where
n is the number of pairs of friction surfaces (1 in the case of a single disc with two facings)
P is the pressure provided by the spring
h is the efficiency of the system
De is the external diameter
Di is the inner diameter of the friction disc

An important parameter related to clutches is the response time, “the transient period,” that is, the time
needed to reach steady speed. The sliding between parts should be as short as possible to avoid excessive
wear and high temperatures. However, torque, which is too high, may cause uncomfortable vibration, so
the pressure of engagement must be modulated.

58.5  Brake Noise and Vibrations


58.5.1  Noise Onset
Noise, vibrations, and harshness (NVH) are a constant worry of brake system developers. The attempts
to correlate measures of frequencies and intensities to harshness have mostly failed. Actually, the
determinant NVH tests of a development process are always based on human feeling and statistical
approaches in automotive tests. Among the factors that affect this kind of assessment are cultural and
socioeconomic aspects of the driver evaluator.
In this section, a brief outlook is provided on the sources and initiation of brake noise, as well as ways
to mitigate it. Every structure that behaves as a harmonic oscillator with a given mass and stiffness may
be resonant or have uncontrolled movement if the exciting force has the same frequency as that given
by the following equation:

1 k
f = , (58.13)
2p m

where
f is the frequency
k is the elastic constant
m is the mass

Although the analytic equations describing the behavior of real oscillators are more complex than
Equation 58.13 and with more resonance frequencies, the most general conclusions that can be inferred
from Equation 58.13 are valid also for complex systems such as brakes.
In the brake system, the disc works as a resonance box (i.e., like a trumpet) so that noise starts and is
amplified by the disc itself. The rotor has many resonance modes in the audible range, and usually there
are seven or more frequencies up to 10 kHz. The pad has from one to five resonance frequencies in the
same range. The noise problem arises when one or more resonance frequency peaks of both the pad and
rotor overlap. Wear of both parts shifts the resonance modes to lower frequencies. This explains why
sometimes brakes are noisy and sometimes not, and this is due to the fact that for finite periods some
Brake and Clutch 58-9

k c

FIGURE 58.5  Sketch of a mass-spring-damper system (m: mass, k: spring elastic constant, c: damping coefficient
of the viscous damper).

resonance peaks of both parts overlap. Brake noise is a function of thickness and mass variation of both
parts. Since brake pad wear is much higher than that of the disc, it can be said that this phenomenon
depends mainly on pad wear. Thermal history is also important, because high temperatures can induce
several chemical transformations, which may change mass and mechanical and tribological properties
of the brake pad. Some of the ways to shift and alter resonance frequencies in order to reduce noise are
linked to the geometry and shapes of brake pads through slots and chamfers; however, there are other
effective approaches.
What can we do in order to avoid the resonance of the system? According to Equation 58.13,

1. We may decrease k
2. We can increase m
3. We can increase the damping factor of the system

The first solution often implies weak structures with low mechanical strength. The second solution sug-
gests adding weight to the caliper or other brake components, which is counter to the current trend of
vehicle weight reduction. Luckily, every physical system has damping properties.
Indeed, real oscillators can be schematically described as a mass linked simultaneously to an elastic
spring and to a viscous damper, as shown in Figure 58.5.
Therefore, it is better to have a close look at the damping properties of the brake components’ materi-
als, but it should be remembered that the exciting force is only due to the friction force variation during
the brake disc contact during braking. Moreover, the variation of the friction force is mainly due to
the variation of contact area, contact nature (adhesive, abrasive, deformation), and phases in contact
between the pad and disc during sliding. The main difference between noise and vibrations is essentially
the frequency range. The human body perceives low-frequency oscillations (long amplitude) as vibra-
tions and high-frequency oscillations (short amplitude) as noise. The vibrations’ frequencies are below
1 kHz while noise reaches up to 20 kHz, that is, the upper limit of the human audible range.

58.5.2  Noise and Vibration Damping


The damping ability of materials is due to their viscoelastic properties, and since friction force generates
a shear stress, the shear modulus is the most adequate to determine the damping ability of both fric-
tion materials and rotors. The shear modulus is the ratio between the shear stress σ and shear strain γ.
Under oscillating shear stress (as in noise conditions for friction materials), the following equations are
obtained for stress and strain:

s = s 0e iw t , (58.14)
58-10 Design for Lubrication and Tribology

g = g0e i(w t +d ) , (58.15)


where
δ is the phase lag or loss angle
i is −1
σ 0 and γ0 are the amplitude values

For full elastic materials δ is 0, and for a full viscous material, it is π/2.
For viscoelastic materials such as friction materials, a complex modulus G* can be defined, which is
given by the following equation:

s s 0 −id
G* = = e = Gʹ + iGʹʹ. (58.16)
g g0

The damping ability of materials can be expressed through tangent δ (tan δ), which is given by the
ratio between the shear loss modulus G″ (the out-of-phase or viscous component) and the shear storage
modulus G′ (the in-phase or elastic component) of the considered material:

Gʹʹ
tan d = . (58.17)

Since the loss modulus is related to dissipative processes, while the storage modulus is related to conser-
vative processes, high values of tan δ correspond to high damping factors. Both the disc and caliper are
usually made of gray cast iron, and the only chance of increasing damping properties of these compo-
nents is to increase the carbon content. However, an excessive amount of graphite flakes (or spherules in
the caliper) in the cast-iron matrix weakens the structure.
However, there is another much utilized solution. Either vibrations or noise begin due to a friction
force variation, but we can put a damping barrier between the brake pad back plate and the caliper,
called either an insulator or antinoise shim, fixed on the back plate. This damping barrier is usually
a multilayered shim, a wafer-like structure with a thickness of around 0.1 mm and alternate sheets of
metal and rubber-based materials with high damping properties, in order to maximize the tan δ value
of the antinoise shim.

58.6  Friction Materials’ Compositions, Processes, and Properties


Friction materials have a wide automotive utilization, both for clutches and brakes. In the last two
decades, certain changes have occurred in friction material composition and engineering; unfortu-
nately, no real breakthroughs were achieved in this field. The most significant innovations were due to
the banning of the use of asbestos about 20 years ago in several countries. Asbestos was one of the main
components of friction materials for its multifunctional properties, such as high thermal stability, low
thermal conductivity, etc. Moreover, asbestos provided a stable friction coefficient. However, this forced
change has resulted in more complex and expensive material formulas with many components.
There are many definitions of friction materials such as ceramic, metallic, semimetallic, non-asbestos
organic (NAO), low met, low steel, hybrid, and so on. Each definition is often arbitrary, and often fol-
lows commercial appeal rather than real conceptual differences among materials, for example, the uti-
lization of a different binding agent. These definitions are mostly linked to the quantity or absence of
a determined compound or component that provides specific desired or undesired characteristics, but
all these materials follow a common general composition scheme [9], in which the goal is to achieve
Brake and Clutch 58-11

a balance among the abrasive, lubricant, and metallic systems for any application. The weight ratio
between abrasives and solid lubricants (including mild abrasives) is usually in the range of 1.5–0.5.
The highest values of this ratio usually correspond to friction materials with mild abrasives such
as NAO. For lubricants, the weight ratio of graphite and sulfides is usually between 1.5 and 1. High
operation temperatures (T > 300°C) need a high content of sulfides. Finally, the ratio of abrasives and
nonferrous metals is usually between 1.5 and 1. A high content of nonferrous metals stabilizes the
coefficient of friction at high temperatures (T > 400°C) especially for materials with low content of
aggressive abrasives. The components of the common mixtures are divided among seven main catego-
ries: abrasives (alumina, zircon sand, silicon carbide, etc.), solid lubricants such as graphite and sulfides
(CuS, ZnS, SnS, SnS2, etc.), metals (iron powder, steel fiber, zinc, tin, copper, and alloys), fillers such as
barite (BaSO4) and coke, rubbers, inorganic (rock wool) and organic fibers (aramid, polyacrylonitrile
PAN, and cellulose fibers), and phenolic resins. Each category of components has different functions,
although not always clearly defined. For example, abrasives are used to abrade the oxide film formed
on the rotor surface; lubricants mitigate the effect of abrasives and lower wear of both the disc and
brake pad. Figure 58.6 shows some example of raw materials widely used. Figure 58.6a clearly shows
the lamellae of graphite by scanning electron microscope (SEM), the most common solid lubricant that
reduces friction and wear below 300°C.

499X 5 kV WD : 15 mm S : 00000 P : 00003


100 µM

5 µm
3243 Cot 2635 chinese

(a) (b)

100 µm 50 µm
50. 0× 20. 0 kV 10 mm CL : 5. 0

(c) (d)

FIGURE 58.6  SEM images of (a) graphite lamellae (Courtesy of Asbury Carbons, Asbury, NJ), (b) polyacryloni-
trile (PAN) fibers with fibrils, (c) Sicacell, a raw material based on silica and cellulose fibers, and (d) sponged iron
powder. (Courtesy of Itaprochim S.r.l., Milan, Italy.)
58-12 Design for Lubrication and Tribology

NAO material:
High organic content
Organics : resin and rubber High amount of copper and its
Copper alloys
alloys Low carbon-based materials
Fibers
Mild abrasives
Inorganic ceramic fibers
Graphite
Fillers

Sulfides Low-met materials:


Low organic content
Steel (fiber, powder)
Copper and its alloys
High carbonaceous materials
Mild abrasives High amount of strong abrasives

FIGURE 58.7  NAO materials or materials derived from NAO represent the latest improvement in conventional
friction materials. The pie chart shows a typical composition of NAO.

Figure 58.6b shows polyacrylonitrile (PAN) fibers with fibrils that retain ceramic powders during
processing and mixture handling, decreasing dry mixture segregation of low-density materials as well
as dust. Usually, a certain amount of aramid or PAN fibers, as well as steel fibers were introduced into
the early asbestos-free mixtures to substitute for the framework made by asbestos fibers. Moreover,
aramid fiber also exhibited high thermal stability. Figure 58.6c shows a synthetic silicate product with
a low content of cellulose fiber, which is used to decrease the thermal conductivity and lower densi-
ties of ceramic materials. In Figure 58.6d, a sponge particle of iron powder, which is a typical metallic
component of metallic friction materials, is shown. Improvements led to materials that have little or no
steel fibers (NAO materials). Figure 58.7 schematically displays the main ratios among the formula com-
ponents of the new-engineered materials. It is possible to notice the tendency to obtain more ceramic
formulas without or with a small amount of iron. Currently, NAO materials or materials derived from
NAOs are those most adequate to satisfy the NVH requirements of the major automobile manufacturers
worldwide.
The production process can be described through the following steps: (dry or wet) mixing, (hot or
cold) press molding, grinding, thermal treatments (infrared, radio frequencies, microwave, or convec-
tive ovens), painting, and a variable number of other operations such as insulator and spring assemblies.
The critical physical properties of brake pads are density, compressibility, shear strength, and Young’s
modulus. Compressibility is a property linked with brake pedal response and noise. Front axle pads
have typical compressibility values between 150 and 180 μm, while values of rear axle pads can be much
lower at 100 μm or less (around 90 μm). There are many different types of friction materials for clutches,
including metallic sintered materials. Nevertheless, the general conceptual composition scheme pre-
sented for brake pad materials is valid also for clutch materials with variations depending on the differ-
ent requirements.
Metallic sintered friction materials represent a different class of materials. Their matrix is made of
sintered metallic powders of iron or copper alloys containing abrasives and solid lubricants. The sin-
tered metallic matrix has a higher thermal stability than resins and so they are preferred for applica-
tions where a high density of energy is dissipated, as in railway applications or thin pads as required for
motorcycles.
Carbon–carbon materials are commonly used for aircraft brakes; however, they require an expensive
manufacturing process in which a pyrolyzed matrix of carbon fibers bound by a resin is subsequently
densified by decomposition of an organic gas at high temperatures [10]. Nevertheless, these materials
have high mechanical and thermal strength, and have excellent friction performance at high tempera-
tures (above 300°C) with a stable coefficient of friction near or above 0.4. Their top performance at high
temperatures (low level of friction fading) is especially appreciated in Formula 1 competitions. On the
other side, carbon ceramic materials (CCM), a carbon fiber framework embedded in silicon carbide,
have been used for rotors in high-performance cars, but they are too expensive for mass production.
Brake and Clutch 58-13

58.7  Friction Couple: Brake Pads and Rotors


The brake pad and rotor interface determines the friction characteristics and most of the other brake
performance criteria. Most common rotors are made of gray cast iron. Gray cast iron exhibits a good
damping factor due to its composite heterogeneous microstructure and contains graphite flakes, which
provide a solid lubricant that reduces wear. These characteristics together with low price make this
material a good choice for serial rotor production. Brake pads are made of composite materials, which
contain ceramics, graphite, metals, sulfides, elastomers, and phenolic resins, whose composition has
been described in the previous section. Brake friction tests show that there is a running-in period during
which the coefficient of friction increases till it reaches a steady state, during braking applications under
the same conditions. Then, if an iterative cycle is stopped and the temperature of the friction couple
decreases in a significant way, upon restarting of the cycle, the coefficient of friction does not begin
from the steady-state value or from the initial one, but from an intermediate value [1]. This simple fact
observed many times during friction tests shows that two conditions are fundamental: (1) surface state
and (2) temperature. Both sliding and abrasive wear (two- and three-body) generate tracks and defor-
mation on the surfaces, and temperatures favor the oxidation of surfaces and debris, modifying surfaces
in an irreversible manner. The transient coefficient of friction during running-in reflects the changes
occurring in the interface between the pad and rotor. This interface needs energy to initialize and main-
tain wear, and tribochemical reactions, such as the oxidation of the gray cast-iron surface and wear
debris. This energy is provided by the dissipation of kinetic energy. Adhesion and deformation also dis-
sipate the kinetic energy used to sustain the mentioned wear and tribochemical irreversible processes.
The transfer of material from the friction material surface to the rotor surface and back are other
aspects to be taken into account while attempting to describe this friction couple. The surface of the
brake pad is covered by a film of variable thickness (2–10 μm) as seen on the top of a section of friction
material observed in an SEM (Figure 58.8) [11].
This film is the true friction surface, which is continuously regenerated by the oxidized wear debris
coming from the disc. The oxides of this film exhibit all the oxidation states of iron, with the most
typical composition very close to Fe3O4 [1]. The disc is covered by a thin film of iron oxides with a com-
position close to FeO1.3 of 5–10 μm thickness [3], which is the main source of the wear debris. The most

10 µm Mag = 1.58 k × Signal A = OBSD Date: Feb 3, 2006


WD = 16 mm
EHT = 20.00 kV Photo no. = 1921 Time: 14:47:09

FIGURE 58.8  SEM image of the section of a friction material that underwent a friction test. On the friction sur-
face, it is possible to distinguish a layer of variable thickness (2–10 μm) darker than the composite matrix. (From
Dante, R.C. et al., Tribol. Int., 42, 958, 2009. With permission.)
58-14 Design for Lubrication and Tribology

external part of this oxide layer contains oxidized metals and ceramics coming from the pad and form-
ing a composite upper film.
Although the transfer of material from the pad to the disc surface is quantitatively much lower than
that transferred from the disc to the pad surface, the composition of the most external film of the disc
can affect the friction behavior of the couple, especially if the pad material does not contain aggressive
abrasives with large particle size (100–200 μm). This is especially true in the case of the so-called NAO
friction materials, which have low levels of aggressive abrasives.
It is possible to find on the disc surface many metals (depending on pad composition) coming from
the pad such as copper, tin, zinc [2,3], which are either in a metallic state or part of the composite most
external film.
The friction couple exhibits all types of wear, but the main wear mechanism for the disc surface
is oxidative, while also important are deformation wear and abrasive wear. The pad wear is mainly
three-body abrasive wear. It should be noted that the friction material binder, for example, phenolic
resins, undergoes pyrolysis, and at high surface temperatures the surface is weakened by the oxidation
of the binder. This is another source of wear that occurs at high temperatures (more than 400°C on the
interface) and becomes particularly relevant for thermal histories marked by high density of dissipated
power, such as for heavy trucks, and mountain descent operation of automobiles. The oxidative/ther-
mal decomposition process tends to destroy the external oxide film due to the stresses generated by the
volume changes caused by the loss of resin mass, and the increased porosity. In addition, many of the
particles that are embedded in the resin matrix are released and cause the three-body wear. At very high
density of dissipated power, pad material undergoes very severe wear that can be attributed to the failure
of the binder. Figure 58.9 shows that, at values of dissipated power around 3.5 kW cm−2, four different
friction materials are severely worn with a wear rate of 2–3 μm s−1 [11], that is, close to full removal of
the oxide film every second.
Taking into account these phenomena, it can be understood that many R&D activities are focused
to strengthening the binding properties of the friction material matrix. A specific problem that

3.5

N3
3 N1
N2
M1
2.5
Wear rate (µm s–1)

1.5

0.5

0
0 0.5 1 1.5 2 2.5 3 3.5 4
–8 –1 –2 –2
p [ 10 KJ s µm (kW cm )]

FIGURE 58.9  Wear rate n for materials N1, N2, N3, and M1. The unit of the dissipated power density p, expressed
as 10−8 kJ s−1 μm−2, is equivalent to kW cm−2. (From Dante, R.C., et al., Tribol. Int., 42, 958, 2009. With permission.)
Brake and Clutch 58-15

concerns interface properties is disc thickness variation (DTV) [3], which, in many cases, provokes
vibrations during braking applications at high speeds. DTV is due to an uneven wear of disc surface
that amplifies the initial waviness of the disc [12]. Cold vibration is noticeable during deceleration
and at low to medium pressures, situations that occur frequently when slowing down slightly on
the motorway. Cold vibration seems to be linked to the balance between disc wear and growth of
the oxide layer film at very low pressures, for example, city braking applications or in conditions of
residual torque without braking. Hot vibration occurs after use at high temperature and its mecha-
nism is due to complex factors.
Hot vibration seems to be linked to hot spots and consequent localized disc surface hardening due to
the formation of metastable martensite. This can cause uneven wear of the disc surface. More compress-
ible friction materials often solve this undesired phenomenon by better coupling with the counterface.
Another phenomenon, which has caused many discussions about friction material engineering, rotor
composition and manufacture, is metal pickup. The formation of embedded metallic iron particles of
variable size (from micrometers to millimeters in size) has been found on friction material surface after
sliding. These metallic particles may provoke serious disc scoring, noise, and weaken the friction mate-
rial surface. Materials with high porosity, large surface cavities, and weak structure favor the embed-
ding of these particles. Metallic iron debris coming from the worn disc surface is thought to accumulate
in material surface cavities, and friction shear stresses plastically deform and alloy the iron particles to
form large hard particles. This hypothesis implies that iron debris is not fully oxidized. In this case, gray
cast-iron microstructure, especially graphite lamellae’s density and distribution, as well as rotor surface
roughness, are important factors. Since iron debris is rapidly oxidized in air, tribochemical explana-
tions for generation of large particles support a sudden reduction of iron debris by potential sources of
electrons such as carbon and hydrogen present in organic compounds of friction materials, which are
activated by the high-energy density of friction dissipative processes. Apart from this tribochemical
initiation, generally the formation of metal pickup is explained through shear flow alloying [13].
In any case, the presence of large cavities on the friction material surface favors the hosting of metallic
iron pickup particles. The weakness of the friction material structure is doubtless a factor that contrib-
utes to the enhancement of this phenomenon. This is in accordance with the fact that wet friction con-
ditions, and especially salt solution, that weaken the friction material structure, enhance metal pickup
formation in steel materials. In addition, dry metal pickup formation is observed in NAO materials that
have an intrinsic weak structure.

58.8  Future Trends


As stated in the previous section, the banning of asbestos caused the most significant changes of the last
decades in friction materials. Usually, a certain amount of aramid fibers, as well as a great amount of
steel fibers were introduced into the early asbestos-free mixtures to substitute for the framework previ-
ously provided by asbestos fibers. High content of steel fibers causes oscillations of the friction force,
which may excite brake resonance frequencies. Later improvements led to materials that have low or no
steel fiber content (NAO).
This trend has led to material structural problems. Indeed, material wear issues became more com-
mon mainly due to the lack of steel fibers, representing a further challenge to material engineering. In
order to reduce wear, great quantities of nonferrous metals, and metal sulfides, such as copper, were
introduced into formulas. Since prices of nonferrous metals have considerable fluctuations, low price
policies will lead to automotive friction materials with a low content of these metals.
Environmental issues concerning heavy metal (e.g., copper) emissions due to brakes are now seri-
ously considered [14], and many attempts are underway to reduce copper in formulas. In the United
States, The Brake Pad Partnership [15] reached consensus that the most effective course of action
would be to pursue legislation that reduces the amount of copper used in brakes to an insignificant
amount in a phased manner. The resulting bill, SB 346, which became law in September 2010, places
58-16 Design for Lubrication and Tribology

a 5% weight limit on the copper amount used in brakes sold in California by 2021, and reduces
that percentage to 0.5% by 2032. The brakes developed to meet these requirements must also meet
all applicable safety and performance standards. Other states will follow the example of California
legislation.
The recovery of braking energy is pursued in the reduction of both fuel consumption and greenhouse
gas emissions. Hybrid cars already allow us to recover part of the braking energy, and probably there
will soon be other types of systems that will allow us to save even more energy.
The recovery of part of this energy implies that brake does not need to dissipate it. Therefore, the
demand of high performances by brake pads will be lowered leading to a new class of mild friction
materials with mild abrasives, less metals, no steel, no heavy metals, for example, copper (either
in metallic state or compounds) with the exception of iron, less sulfides, few lubricants and fibers.
Research on raw materials will lead to new classes of raw materials where morphology and fine struc-
ture will be taken into account. Performance will be assured through composite raw materials that
will couple together more friction characteristics in order to minimize the reduction or disappear-
ance of some specific compound. Therefore, coated particles with a specific core and a thin sur-
face layer with different characteristics could be attractive for these new classes of friction materials.
High-performance cars are expected to adopt ceramic/carbon composite rotors that will lead to dif-
ferent friction materials where carbon-based components from graphite and coke-based materials
will dominate.

References
1. R. Dante, E. Galetto, and D. Venezia, Chemical transformations induced by friction, Proceeding of
the Fourth European Conference on Advanced Materials and Processes, Venice, Italy, September 1995,
pp. 25–28.
2. R. Dante, B. Tron, and L. Longo Vaschetti, Wear mechanism of the brake disc studied with the analy-
sis of the evolution of the sliding surface, Proceedings of the 29th ISATA (International Symposium
on Automotive Technology and Automation), Florence, Italy, June 3–6, 1996.
3. E. Sales, D. Venezia, and R. Dante, Effects of surface oxidation of the brake disc on D.T.V (Disc
Thickness Variation) creation, Proceedings of the 29th ISATA (International Symposium on Automotive
Technology and Automation), Florence, Italy, June 3–6, 1996.
4. W.H. Li and H. Du, Design and experimental evaluation of a magnetorheological brake, The
International Journal of Advanced Manufacturing Technology, 21(7), 508–515, 2003.
5. I.M. Hutchings, Tribology—Friction and Wear of Engineering Materials, Butterworth-Heinemann,
Oxford, U.K., 1992.
6. R. Giovannozzi, Costruzioni di Macchine I, 3rd Edn., Patron Editore, Bologna, Italy, 1980.
7. G. Belforte, Meccanica Applicata alle Macchine, Libreria Tecnica Editrice Dott. Ing. V. Giorgio, Turín,
Italy, 1983.
8. G. Bongiovanni and G. Roccati, Freni, Librería Editrice Universitaria Levrotto & Bella, Turín, Italy,
1990.
9. R.C. Dante, D. Rivacoba Benavides, and M. Fernández, Development of a friction material formula-
tion without metals by Taguchi design of experiments, Proceedings of the 18th Brake Colloquium &
Engineering Display, San Diego, CA, October 2000.
10. J.D. Buckley and D.D. Edie, Carbon-Carbon Materials and Composites, William Andrew Publishing/
Noyes, Park Ridge, NJ, 1993.
11. R.C. Dante, F. Vannucci, P. Durando, E. Galetto, and C. Kajdas, Relationship between wear of friction
materials and dissipated power density, Tribology International, 42, 958–963, 2009.
12. R.C. Dante and C. Navire, Cold wear of brake disc related to brake induced vibrations, International
Journal of Vehicle Noise and Vibrations, 1(3/4), 287–305, 2005.
Brake and Clutch 58-17

13. K. Nukumizu, T. Kobayashi, T. Abe, and M. Unno, Study of the formation mechanism of metal
pick-up on the frictional surface of a disc brake pad, 26th Annual SAE Brake Colloquium and
Exhibition, San Antonio, TX, October 2008.
14. J.H.J. Hulskotte, M. Schaap, and A.J.H. Visschedijk, Brake wear from vehicles as an important source
of diffuse copper pollution, Proceedings of the Tenth Specialised Conference on Diffuse Pollution and
Sustainable Basin Management, Istanbul, Turkey, September 2006.
15. Brake Pad Partnership: http://www.suscon.org/bpp/index.php (accessed June 2010)
59
Automotive Tribology
59.1 Introduction..................................................................................... 59-1
59.2 Engine................................................................................................ 59-2
Edward P. Becker 59.3 Transmission.................................................................................... 59-7
General Motors Company References.....................................................................................................59-8

59.1  Introduction
There are many definitions of the word “automobile” and many claims to its invention. However, it is
clear that the first patent for a successful road vehicle powered by an internal combustion engine was
issued to Karl Benz in Germany in 1886 for his “Motorwagen” (Figure 59.1). Although Mr. Benz would
probably be confounded by the shapes, sizes, and capabilities of today’s automobiles, he would neverthe-
less have no difficulty understanding their basic construction. The majority of automobiles today still
use an internal combustion engine, which delivers power to the wheels by mechanical means (gears and/
or chains), a mechanism for an occupant to move one or more wheels to steer the vehicle, and a throt-
tling mechanism to control the speed of the motor.
A commercially successful automobile must balance many, sometimes conflicting, requirements.
Consumers will often choose an automobile based on such factors as power, durability, reliability, and
quality. In addition, many legislative demands are placed on automakers by various governments, such
as fuel economy, emissions, and freedom from hazardous substances. All of these must be considered
during the design phase, and all have great impact on the tribology of the automobile.
In terms of converting the energy of the fuel into mechanical energy to propel the vehicle, there
are various estimates of the overall efficiency of modern automobile internal combustion engines [1,2].
A typical breakdown is the following:

Exhaust 30%
Cooling 30%
Pumping loss 5%
Drivetrain friction 20% Piston assembly 9%
Bearings 5%
Valvetrain 2%
Transmission 4%
Power to wheels 15%

Thermodynamic considerations dictate that some energy will be lost (primarily through exhaust and
cooling). The challenge for the tribologist is to reduce the energy lost through friction and therefore
increase the amount of energy available to propel the vehicle and drive the accessories.

59-1
59-2 Design for Lubrication and Tribology

FIGURE 59.1  Benz Motorwagen.

59.2  Engine
Reciprocating internal combustion engines are the power plant of choice for the vast majority of auto-
mobiles and trucks on the road today. The basic hardware of such an engine is illustrated in Figures
59.2 and 59.3. It is interesting to note the similarities between these two engines, as Figure 59.2 depicts
a typical engine from c. 1924, while Figure 59.3 is a c. 2010 engine. The basic principle of operation is to
convert the chemical energy of a fuel first into heat, then into mechanical energy.
One of the most fundamental characteristics of an engine is the number of strokes, or reversals of
direction, the piston completes in order to generate power. For automotive transportation, the four-
stroke cycle is typically used. The four strokes are the following:

1. Intake: One of the valves (the intake) opens when the piston is near the top of the cylinder. As the
piston moves downward, either air (in a fuel-injected engine) or an air–fuel mixture (in a carbu-
reted engine) is drawn into the cylinder.
2. Compression: As the piston moves upward again, both valves are closed and the contents of the
cylinder are compressed.
3. Power: As the piston nears the top of the cylinder again, fuel is added in a fuel-injected engine (the
fuel is already mixed with air in a carbureted engine) and the mixture is ignited. The combustion
process releases heat and generates a variety of combustion products such as water vapor and
carbon dioxide. The pressure rises rapidly within the cylinder, and the pressure forces the piston
downward.
4. Exhaust: As the piston reaches the bottom of the cylinder, some of the energy generated by the
combustion of the fuel has been transferred to the crankshaft by the piston via the connect-
ing rod. Another valve (the exhaust) now opens, and the momentum of the crankshaft moves
the piston upward again, sweeping the combustion products out of the cylinder. As the piston
nears the top of the cylinder, the exhaust valve closes and the intake opens, and the cycle
repeats.

Engines are further classified according to the number and arrangement of cylinders (in-line 4, V6, and
V8 are common), total displacement, valve location (overheard valve is typical), camshaft placement (in
block, as in Figure 59.2, or overhead, as in Figure 59.3), means of initiating combustion (spark ignition
or compression ignition), and fuel (gasoline, alcohol, natural gas, and diesel are common). More details
of general engine design can be found in Reference 3.
Automotive Tribology 59-3

Oil plug
rocker arm cover Valve stem
Valve spring
Adjusting ball Valve cage
nut
Lock nut
Valve cage
Water jacket
Valve

Exhaust
Spark plug manifold
cover Intake
manifold
Combustion
space
Hot air
chamber
Push rod

Valve push rod


cover

Wrist pin
Valve lifter Cylinder
cap Piston
Valve lifter
guide clamp
Valve lifter
spring
Valve lifter
guide
Valve lifter
Cam roller Connecting rod
pin
Cam roller Crank case
Cam shaft Crank shaft
Cam

FIGURE 59.2  Internal combustion engine schematic, c. 1924.

Many diverse surfaces interact and have the potential to experience wear when converting the chemi-
cal energy of the fuel into the mechanical energy. The severity of these surface interactions is reduced, to
a greater or lesser extent, by the presence of engine oil. The oil provides different functions in different
regions of the engine. Lubrication conditions are often subdivided into boundary, mixed, and hydrody-
namic domains according to the Stribeck curve (Figure 59.4), which also shows the lubrication regimes
in which various engine components usually operate.
Base oils and additives are discussed in detail in Part II. In general, engine oil is composed of about
80% base oil and 20% additives. The base oil can be a mineral oil derived from refined crude oil or a
synthetic oil built up from chemical feedstock.
59-4 Design for Lubrication and Tribology

FIGURE 59.3  Cutaway of a modern internal combustion engine, c. 2010.

Boundary Hydrodynamic
1. Mixed
Coefficient of friction

Piston rings
0.1
Piston skirt
Valve train
0.01 Engine bearings

0.001

Viscosity × Speed
Unit load

FIGURE 59.4  Stribeck diagram, showing typical operating regions of internal combustion engine components.
Automotive Tribology 59-5

The additive package in engine oil usually contains the following:


• Antioxidants are sacrificial molecules which work to prevent oxidation of base oils and other
additives. Typical antioxidants are dithiophosphates, amines, and phenols.
• Dispersants prevent sludge and deposit formation by keeping foreign contaminants well dis-
persed (suspended) in the oil. Typical dispersants are “ashless” compounds, that is, they do not
contain metallic elements.
• Detergents perform functions similar to dispersants. However, they also neutralize acidic com-
bustion and oxidation products and are ash (metallic) containing. The most common detergent in
engine oil is calcium sulfonate.
• Antiwear compounds react with metal surfaces to form a protective chemical barrier, which helps
prevent surface wear and welding of asperities between moving parts. The most common anti-
wear additive in engine oil is zinc dialkyldithiophosphate, usually abbreviated ZDDP.
• Pour-point depressants prevent the formation of wax crystals at low temperatures, which prevents
the oil from turning into an unpumpable gel.
• Viscosity modifiers are polymer molecules which thicken the oil at high temperatures but have
less effect at low temperatures. This permits the use of lower viscosity base stocks (for better low-
temperature pumpability) without the oil becoming too thin at high temperature.
In general, the lubrication requirements of the internal combustion engine are determined by the vari-
ous surfaces in relative motion within the engine. Each of these interactions has unique characteristics
which must be accommodated by the lubricant. The goal in each case is to minimize the energy lost at
these interfaces and minimize or eliminate wear.
Crankshaft to crankshaft bearing: The crankshaft is the component that transfers rotational energy
from the engine. As it turns, sliding occurs between the crankshaft and the engine block, as well as
between the crankshaft and connecting rod. These two interfaces form a journal bearing, as described in
the section on journal bearings in this handbook. The space within these bearings is filled with engine oil,
and the energy dissipated in these bearing is primarily a function of oil viscosity. Proper alignment and
clearance of these bearings are required for good engine performance, so these components are produced
from strong and stiff materials to minimize deformation under load. Common choices for crankshaft
materials are ductile cast iron or low-alloy steel. Connecting rods are usually steel, although titanium
alloys have been used in a few cases. Engine blocks were traditionally produced from gray cast iron; how-
ever, cast aluminum has become the material of choice for modern engines due to weight savings.
Vehicle engines are generally shut down for long periods. During these times, the shaft will settle to
the bottom of the journal and squeeze out the oil film. Also, solid particles can become entrained in the
oil. These can be from external sources such as dust or internal sources such as wear particles. If these
particles are larger than the minimum clearance between the shaft and journal, they have the potential
to damage the bearing. In order to prevent sticking during engine start-up and provide the possibility
of trapping a limited amount of debris, a soft, compliant bearing is desirable. The ability to trap debris
is called embedability [4].
To meet these contradictory requirements, bearing inserts are used for both the connecting rod and
the engine block. These inserts are usually made of a steel backing (for strength and stiffness), with a soft
coating (for embedability and to prevent adhesion when the engine is off). For many years, the coatings
were lead-based Babbitt alloys. However, environmental concerns have led to aluminum-based coatings
becoming common.
In most engines, the oil is provided under pressure to these bearings, so in this case the most impor-
tant properties of the lubricant are viscosity and heat capacity. The viscosity must be sufficiently high,
even at high temperature, to maintain a hydrodynamic film between the surface and sufficiently low,
even at low temperatures, to allow the pump to maintain oil flow into the bearings. Heat capacity is
important as energy will be dissipated in the form of heat as the oil film is sheared, and the oil must
absorb some of this heat without an unacceptable rise in temperature.
59-6 Design for Lubrication and Tribology

Piston skirt to cylinder block: The piston skirt to cylinder block is one of the primary contributors to
total engine friction [5]. The piston must fit within the cylinder with sufficient clearance to avoid seizure.
However, too much clearance can result in unacceptable noise [6]. Most automotive engine pistons are
cast from aluminum-silicon alloys, which are considerably lighter than the cast iron pistons of the past,
and have resulted in a significant reduction in engine mass and vibration. Also, the higher thermal
conductivity of aluminum helps prevent overheating at the top of the piston. Most of the excess heat is
conducted away by the cylinder block, and some oil is delivered to the pistons by the rotating crankshaft.
However, sometimes, pistons require additional cooling. This is usually provided by adding devices to
direct a jet of oil onto the underside of the pistons.
Cylinder bores are usually made from gray cast iron, which has a lower coefficient of thermal expan-
sion than aluminum. This creates a design challenge, since a piston with adequate clearance at running
temperature may be too loose (and hence noisy) at low temperature. To reduce friction and prevent
scuffing of the piston, oil must be supplied to the cylinder bore walls. Oil flung from the crankshaft is
usually sufficient for this purpose. The clearance between the piston skirt and cylinder wall is so small
at high temperature that special coating is applied to most pistons, such as nickel ceramic composites or
molybdenum disulfide [7,8]. These coatings also reduce the friction in the interface between the piston
skirt and cylinder wall.
As the piston moves through the cylinder, the velocity changes from zero (at the top and bottom of
each stroke) to a maximum value near the center of the stroke. The viscosity of the oil should be suf-
ficient to generate a full hydrodynamic film over most of the stroke, and sufficient thermal conductivity,
to facilitate heat movement from the piston to the block.
Piston rings to cylinder block: The piston rings function as a set of sliding seals to separate the com-
bustion gasses above the piston from the crankcase environment below. The most common arrangement
in automotive engines is a set of three rings. The upper compression ring, lower compression ring, and
oil control ring can be seen in Figure 59.1. The ring-block sliding interface is the major contributor to
total engine mechanical friction [9].
As with the piston skirt, oil usually reaches the cylinder bore surface by being thrown from the
crankshaft after flowing through the bearings. Some oil is necessary for the compression rings to func-
tion properly, but oil that escapes past the top compression ring into the combustion chamber is lost. The
oil control ring ensures that only the necessary amount of oil reaches the compression rings.
The upper compression ring experiences the highest loads and oil temperatures, and it must provide
a good seal to the cylinder surface with very little engine oil. To provide acceptable durability, this ring
is usually made from either nitrided stainless steel or from steel coated with molybdenum.
Piston pin to piston: The piston pin transfers force from the piston to the connecting rod. The inter-
face between the pin and the piston is also a type of journal bearing, however, in the case the motion is
not full rotation. In engines with a fixed pin design, the pin is press-fit into the connecting rod, and the
pin motion against the piston is fully reversed partial rotation. In the floating pin design, the pin is free
to rotate within both the rod and the piston, and the resulting motion is indeterminate. The floating pin
has been shown to reduce the operating temperature of the pin boss and is therefore the preferred design
[8]. In either case, the velocity of the pin is not sufficient to generate a full fluid film between the surface,
and a condition of boundary lubrication results.
Automotive engine pistons are usually cast from aluminum-silicon alloys. Piston pins are made from
low or medium carbon steel, which is usually carburized to generate a very hard surface. Lubricant
is usually provided to this interface from two sources. Some oil is flung upward by the motion of the
crankshaft and arrives at the underside of the piston. In addition, oil that is scraped from the cylinder
walls by the oil control ring can flow down the piston into the pin bore. It has been shown that increasing
the oil supply to this interface decreases the tendency for scuffing [10].
Camshaft to follower: As the camshaft rotates, it presses against a flat or roller surface, which
reciprocates to open and close the valve. The interface between the camshaft and follower is unidirec-
tional sliding between nonconformal surfaces. Although engines are designed to provide oil to this
Automotive Tribology 59-7

interface, it is likely that oil will be scarce on occasion. For example, when starting a cold engine, the
cams will begin turning before oil pressure is sufficient to pump the oil to the top of the engine. Only
a few material combinations are used successfully in this application, and even those wear sufficiently
during the life of an engine to require periodic adjustment or the use of self-adjusting hydraulic
elements.
In this case, the lubricant functions primarily to reduce friction at the sliding interface, to provide a
chemical film to resist wear through the action of antiwear additives, and, in the case of engines with
hydraulic valve lifter of hydraulic valve timing adjustments, to provide energy in the form of hydraulic
pressure to actuate those systems.
Valve to cylinder head: The valves control the flow of gasses into and out of the cylinder. The valve
reciprocates within the valve guide. Engine valves are usually made of specialty steel alloys to resist
oxidation and corrosion. In addition, wear-resistant coatings such as chromium are often used. Modern
cylinder heads are usually made from cast aluminum alloys. The valve guides are typically made from
steel by the powder metal process.

59.3  Transmission
The main drawback of the internal combustion engine (compared to steam or electric motors) is that it
cannot provide high torque at low speed, and the engine must continue turning even when the vehicle
stops or the motor will stall. This requires a means for allowing the engine to turn at high speed even
when the vehicle is traveling slowly or not at all. The device that varies the ratio of rotation between the
engine and the wheels is called a transmission.
A modern automatic transmission is shown in Figure 59.5. The principal components of the trans-
mission are the torque converter (not shown), clutch assemblies, input and output shafts, and gear sets.
The torque converter is attached to the engine crankshaft and turns at engine speed. The torque con-
verter is filled with automatic transmission fluid (ATF), which acquires momentum from the converter
and exerts pressure on a turbine, which is connected to the transmission input shaft. The input shaft
is connected to the output shaft through a series of gear sets. Clutches are engaged or disengaged by
hydraulic pressure to determine which gear set is active at any given time, which in turn determines the
ratio of rotational speed between the input and output shafts. In this manner, the engine can provide
high torque at low vehicle speed to the wheels for acceleration and continuous power at lower engine
speed when cruising at high vehicle speeds.

FIGURE 59.5  Cutaway of a modern automatic transmission.


59-8 Design for Lubrication and Tribology

The ATF is an oil-based lubricant, similar to engine oil but requiring different performance
characteristics. Most ATFs contain about 85% base oil and 15% additives, and around 250 ppm red dye,
to make the ATF visually distinct from engine oil.
ATF has less additives overall than engine oil, particularly because ATF is not exposed to the high
temperatures and reactive chemicals in the engine combustion chamber. Therefore, ATF does not
require as much antioxidant additive. ATF also functions as a hydraulic fluid to move valves and
engage clutches. Like engine oil, ATF must protect against wear in heavily loaded components (in
this case, gears), but unlike engine oil, ATF must provide for high and controllable friction between
components (in this case, clutch plates). For this reason, ATF contains different additives specifically
to maintain friction behavior. A discussion of the ability of ATFs to transmit torque is contained in
Reference 11.

References
1. Taylor, C. M., Automobile engine tribology—Design considerations for efficiency and durability,
Wear, 221, 1–8 (1998).
2. Toten, G. E. (ed.), Handbook of Tribology and Lubrication, Vol. 1, Application and Maintenance,
Taylor & Francis, Boca Raton, FL (2006).
3. Taylor, C. M., The Internal Combustion Engine in Theory and Practice, The MIT Press, Cambridge,
MA (1985).
4. Ludema, K. C., Friction, Wear, Lubrication—A Textbook in Tribology, CRC Press, Boca Raton, FL
(1996).
5. Goenka, P. K., R. S. Paranjpe, and Y. R. Jeng, FLARE: An integrated software package for friction and
lubrication analysis of automotive engines, Part 1: Overview and applications, SAE International
Paper No. 920487, Warrendale, PA (1992).
6. Kageyama, H., T. Suzuki, and T. Ochia, Numerical study on the three dimensional contact pressure
and deformation of piston skirt, SAE International Paper No. 2001-08-0082, Warrendale, PA (2001).
7. Funatani, K., K. Kurowawa, P. A. Fabiyi, and F. M. Puz, Improved engine performance via use of
nickel ceramic composite coatings (NCC Coat), SAE International Paper No. 940852, Warrendale,
PA (1994).
8. Rao, V. D. N., D. M. Kabat, D. Yeager, and B. Lizzote, Engine studies of solid film lubricant coated
pistons, SAE International Paper No. 970009, Warrendale, PA (1997).
9. Ting, L. L., A review of present information on piston ring tribology, SAE International Paper No.
852355, Warrendale, PA (1985).
10. Takiguchi, M., M. Oguri, and T. Someya, A study of rotating motion of piston pin in gasoline engine,
SAE International Paper No. 938142, Warrendale, PA (1992).
11. Linden, J. L. et al., A comparison of methods for evaluating automatic transmission fluid effects
on friction torque capacity—A Study by the International Lubricant Standardization and Approval
Committee (ILSAC) ATF Subcommittee, SAE International Paper No. 982672, Warrendale, PA
(1998).
60
Turbomachinery Tribology
60.1 Mechanical Component Design and Analysis............................60-1
Bearings  •  Seals  •  Rotor Systems
60.2 Mechanical Testing, Diagnostics, and Prognostics....................60-8
Test Methods  •  Diagnostics: Methods of Detecting Degree of
William D. Degradation  •  Prognostics
Marscher 60.3 Closure.............................................................................................60-11
Mechanical Solutions, Inc. References...................................................................................................60-12

60.1  Mechanical Component Design and Analysis


The term “turbomachinery” is generally considered to include fluid machinery that turns at a constant
rotational speed for any given operating condition. As such, the term may be used to represent steam
and gas turbines (axial and radial), gas expanders, centrifugal and axial compressors and fans, and
centrifugal and high-flow pumps. All of these machines share a set of components that are conceptu-
ally similar from machine to machine. Key components include a casing, also known as a “frame” or
“housing”; shaft bearings mounted in the casing; a rotating shaft and typically a coupling (nearly “rigid”
or purposely “flexible”) that attaches it to a mating machine; and fluid elements consisting of some sort
of disk or hub with attached blades that either work on the fluid (in a compressor, fan, or pump) or are
worked on by the fluid (in a turbine or expander). There are also stationary passages within the machine
that convey the fluid to or from the blading in a manner that is intended to lead to maximum effective-
ness and efficiency, and sealing that prevents unwanted leakage.
As turbomachines age, their operational effectiveness (e.g., thermodynamic efficiency) and reliability
degrade, typically due to one of the following four factors: (1) wear of various forms (including surface
erosion), (2) fouling of the process flow passages or the bearing lubricant, (3) corrosion and/or oxidation,
and (4) metal fatigue. The focus of this chapter will be only tribological issues, involving wear, erosion,
and lubrication.

60.1.1  Bearings
The interface between the surface speed of a rotating shaft and the stationary housing that ultimately
supports the rotor system has always been problematic. Depending upon rotor speed, lateral and axial
thrust loads placed on the rotor by the process, and the machine’s environment, a variety of bearing
solutions have been developed and successfully applied in turbomachinery applications. In each case,
the key issue is to maintain shaft location with sufficient precision, over an extended period of time
hopefully at least equal to the required useful life of the turbomachine. In many applications, such as
power plant, petrochemical, and aerospace prime movers, this goal is typically not achieved, and instead
the fallback goal becomes achieving a reasonable mean time between overhaul (MTBO), for example,

60-1
60-2 Design for Lubrication and Tribology

minimum 5 years in industrial applications. Typically, bearing types are selected to maximize MTBO,
within purchase cost and lifetime maintenance cost constraints, and in the case of aerospace applica-
tions, weight constraints. More details on bearing selection, application, and lubrication are given by
Marscher [1].

60.1.1.1  Rolling Element Bearings


Rolling element bearings are represented by various styles of ball bearings and cylindrical or “spherical”
(barrel-shaped) roller bearings, the fundamental concept of which is reported to date back to Roman
times. The benefit of this style bearing is that it serves as a self-contained “cartridge” that can be fit onto
a shaft and pressed into a housing, providing instant ability of the shaft to rotate relative to the housing
with minimal friction and wear. Typically, some form of lubricant is required to achieve low wear at
useful running speeds and/or significant loads on the shaft. The amount of lubricant required is small
if its “lubricity” is high in terms of sufficient adhesion to the metal asperities and sufficient viscosity to
maintain a film thickness between the rotating and stationary components, which is greater than the
combined asperity heights of the separated surfaces. In aerospace research applications, it may be a sur-
prise that vapor phase lubricants have been able to achieve this goal. In industrial applications, grease
is typically used, and contrary to intuition, overfilling the bearing cavity with grease is more often a
problem than provision of insufficient lubricant. Likewise, in the case of oil-lubricated rolling element
bearings, the oil should pool only to the middle of the lower ball or roller. When overfilling with either
grease or oil, there is more lubricant present than needed, and this extra fluid is churned by the bearing,
shear from which leads to considerable heat generation, enough to damage the bearing through exces-
sive thermal growth or to quickly degrade the lubricant both through a lower viscosity than anticipated
by design (viscosity drops dramatically with temperature) and through rapid oil oxidation. In high
speed or highly loaded applications, this issue competes with the need to have sufficient lubricant flow
present to remove frictional heat generation even if churning is avoided. The approach in this case is
an oil jet directed at the loaded roller interface and “scavenged” shortly thereafter, or an “oil mist” that
prevents oil pooling at the bearing, once again requiring a good scavenging strategy.
Concerning system dynamics, there is a common misconception among industrial turbomachinery
engineers that rolling element bearings essentially result in near metal-to-metal contact between the
shaft and housing, and therefore no motion of significance can occur between the two. Therefore, shaft
displacement proximity probes are not mounted near rolling element bearings in situations where they
would be a probe of choice for fluid film bearings. In reality, the effective stiffness of a rolling element
bearing is typically somewhat higher but of the same order of magnitude as a comparably sized well-
designed fluid film bearing. Furthermore, depending upon the internal clearances and mounting clear-
ances of a rolling element bearing, it can allow a relatively large nonlinear lateral motion of the shaft in
the bearing under light load until the clearances are taken up, and this must be taken into account in
setting maintainable wear ring and labyrinth seal clearances, for example.
Another important aspect of rolling element bearings with regard to system dynamics is their very
low damping. Among all forms of shaft support, rolling element bearings have the least energy absorp-
tion by the nature of their construction. This can result in excessive vibration response when, for exam-
ple, running speed traverses critical speeds, or can even result in rotordynamic instability if net modal
damping is not provided by other elements of the rotor or rotor support system.

60.1.1.2  Journal/Fluid Film Bearings


Journal bearings involve a shaft being supported by a fluid film within a cavity surrounding the shaft.
These bearings occur in several categories. The simplest form is the plain journal bearing.
In plain journal bearings, the bearing stationary cavity is a round cylinder, typically with at least
one axial groove and perhaps a circumferential groove for instance at mid length, where the grooves
facilitate lubricant distribution. The shaft lateral support in the plain bearing is by a hydrodynamic film
(a “fluid wedge” of increased static pressure that at least partially opposes any lateral load or deflection
Turbomachinery Tribology 60-3

imposed on the shaft). In the process of developing the film pressure distribution, the hydrodynamic
action of the lubricant attempting to flow through the “pinch” point of the shaft’s eccentric deflected
location leads to a pressure bubble skewed to the side of the incoming fluid. In addition, the lubricant
flows at roughly half the shaft rotational speed since the fluid at the cavity wall is stationary while that
at the shaft is at shaft surface speed. The skewed pressure creates a force vector perpendicular to the
shaft motion, in addition to the vector that opposes the external applied load. If the rotational speed of
the shaft is well below its first lateral natural frequency (its first “critical speed”), this so-called cross-
coupled force simply encourages light whirl of the rotor.
However, if the rotor is above its first critical speed, particularly if it is above twice the first bending
critical speed, then the cross-coupled force is not responded to until the rotor whirls ¼ turn, due to the
90° “phase lag” between force and response, which is the nature of whirling at the critical speed. This
leads to the rotor responding in a manner that further closes the “pinch point” between the rotor and
the bearing cylindrical wall, in turn increasing the cross-coupling force for the next quarter revolution,
and so the cycle continues until damaging rubbing wear is likely to occur. This process is known as
“rotordynamic instability” and is the Achilles heel of plain journal bearings operated at relatively high
speeds. Another perspective on the cause of rotordynamic instability is that the cross-coupled force vec-
tor is in the same direction as, but with opposite sense to, the damping force, which is primarily a drag
force vector opposing the instantaneous path of the shaft whirling motion. Disruption of coherent whirl
of the bearing lubricant (or annular seal flow at wear rings in pumps; Massey [2]) discourages rotor
asynchronous whirl in concert with such fluid whirl and discourages rotordynamic instability. Most of
the complications introduced into more sophisticated journal bearing types are intended to accomplish
this. Other design enhancements are typically intended to directly decrease cross-coupling, or to over-
come cross-coupling by increasing damping, such that the damping vector amplitude is guaranteed to
be greater than the cross-coupling vector it opposes.
The more complex forms of journal bearings include lemon bore bearings, fixed pad bearings, tilting
pad bearings, specially grooved bearings (e.g., herringbone groove gas bearings), and foil bearings. The
details of these bearings are provided in other chapters of this handbook.
Another device similar in configuration to a plain journal bearing is the squeeze film damper (SFD).
The operational difference is that the shaft and bearing cylindrical shell are both stationary in the SFD.
This allows the ends of the SFD to be statically sealed, such as by an o-ring. By trapping the lubricant
between the shaft and hollow cylinders, the seals ensure that the lubricant is squeezed through the clear-
ance in order to allow it to move out of the way of the vibrating shaft. This situation results in the “Stokes
effect” inertia term of the Navier–Stokes equation to dominate, resulting in dissipation of energy, from
the viewpoint of the rotor system. Thus, the SFD may add damping to otherwise poorly damped rotors
and can be particularly useful in providing damping to rotors supported by rolling element bearings,
such as aircraft gas turbine engines.

60.1.1.3  Magnetic Bearings


Magnetic bearings have been used both as the primary means of rotor support as well as a method of
adding damping. In either case, the concept is to use a magnetic field, usually set up to attract the rotor
to the bearing stator, acting opposite in sense to lateral loads on the rotor. In the case of the bearing, the
loads being reacted are the external loads imposed on the rotor system, while in the case of the damper,
the magnetic field is set up to counter any cross-coupled force and to act as a drag force to discourage
whirling velocity of the rotor.
Magnetic bearings began decades ago as devices controlled by expensive, complex analog circuitry.
The devices employed as “backup”, “auxiliary”, or “touch-down” bearings were typically off-the-shelf
rolling element bearings that were kept disengaged from supporting load or rotating by a clearance at
either the inner or outer race. The problem with the analog circuitry was the difficulty of designing and
later tuning the bearings to fit a specific application, adding significant cost and risk to new installations.
The rolling element auxiliary bearings, while they have worked well in many applications, in others have
60-4 Design for Lubrication and Tribology

disintegrated in the first or second time they were called into use, because of roller fretting damage and/
or lubricant loss or degradation over the preceding months or years. These issues have made magnetic
bearing adoption slower than many expected.
Modern magnetic bearings possess control systems based on digital circuitry, practical thanks to
advances in modern microprocessors, and designing or tuning for a new application is basically a
software issue. Modern auxiliary bearings have taken on a broader range of forms and sophistication,
increasing the power-loss and/or coast-down reliability of magnetic-bearing-supported rotor systems.
The main remaining impediments to magnetic bearing adoption in many applications, particularly
aerospace, are their relatively high weight (e.g., due to copper coils), bulky geometry compared to alter-
natives, and relatively low specific load-carrying capability (order of 1 MPa [145 psi] on a practical basis).
On the positive side of the ledger, they eliminate the bearing lubrication system and permit optimized
rotor response to critical speed resonances, maximized rotor damping, and vectorial cancellation of
even unexpected loads occurring within an amplitude range and frequency response within the capa-
bility of the bearing. High-temperature (ceramic) insulation and various cooling schemes are available
to allow their use in even extreme temperature situations and other forms of hostile environments.

60.1.2  Seals
Seals in turbomachinery consist of contacting and noncontacting varieties. Representative examples are
labyrinth seals and mechanical seals, respectively.

60.1.2.1  Labyrinth Seals


The most common forms of noncontact seals at sliding or rotating interfaces are labyrinth seals or some
sort. In compressors and turbines, these consist of multiple “fingers” that have close clearance at their
tips and small plenums in between the fingers. This creates a multiple orifice effect, reducing leakage
considerably more than a smooth surface of the same clearance and length.
Such seals are sometimes run against surfaces that preferentially wear, in order to preserve the integ-
rity of the fingers during and after a significant rub interaction. These types of surfaces are known
as “abradable.” Certain porous ceramics (e.g., lightly sintered yttrium-stabilized zirconia), compressed
metal fiber surfaces, fluoropolymer coatings, and metal/polymeric composites (e.g., aluminum poly-
ester) are examples. In each design, the trade-off is to develop a surface material or coating that has
good thermal resistance and structural integrity, as well as excellent erosion resistance, and yet is easily
grooved or removed during a high-speed rub interaction.
Rather than abradable materials, some manufacturers (or retrofitters) of steam turbines have flex-
ibly mounted the labyrinth seals by circumferentially segmenting them and spring-loading the seg-
ments together. In some cases, such “retractable seals” exhibit their flexibility only during start-up
processes.
Another approach to preserving low leakage after a rotor/stator interaction (e.g., in an aircraft engine
during a high-G maneuver load) is to use brush seals. These seals have seen application not only in
aerospace but also in steam turbines and industrial gas turbines and compressors. These seals consist
of thousands of short small-diameter fibers that are inserted into the stator at an angle (usually about
45°) relative to the rotor surface rotation. The resulting seal can operate for long periods with a slight
nominal interference and works similar to the brushes that homeowners sometimes fix to the bottoms
of their front doors. The fibers typically are made of superalloy and may be laid up in multiple rows like
a labyrinth or combined with labyrinths by placing them to partially fill some of the inter-finger cavi-
ties. In actual operation, the fibers typically aerodynamically entrain a thin film of gas under the fiber
tips, so they do not truly rub, in spite of nominal interference in the stationary condition. The pressure
drops such seals can resist are limited to typically order of 700 kPa (about 100 psi) differential pressure.
A variation on this concept that shows promise to tolerate larger pressure drops is the film-riding seal.
In some cases, the retractable seal concept used in some labyrinth seals has been applied to brush seals
Turbomachinery Tribology 60-5

to help decrease the distress on them during start-up and shutdown, when they are most vulnerable
because of minimal aerodynamic film under the bristles.
In centrifugal pumps, labyrinth seals at the inlet of front shrouds of impellers, as well as in balance
device clearances, function to decrease leakage much as they do in turbines and compressors. However,
in pumps, the fingers are sometimes removed when tight clearances permit it, forming “plain seals.” In
the case of plain seals, and to a lesser extent the labyrinth or “grooved seals” as pump manufacturers
term them, besides the leakage reduction there is a hydrostatic bearing effect that beneficially takes place
at the seal, helping to support the rotor. This is the so-called Lomakin effect and is maximum when
clearance is minimum, with no fingers or grooves. This effect may be calculated or determined by test
on an operating pump (Allaire [3], Childs [4], Black [5], Marscher [6]).

60.1.2.2  Mechanical Seals


Mechanical seals do come in forms that truly contact, but more often, they operate in turbomachinery
by permitting a slight amount of leakage that is sufficient to hydrodynamically “lubricate” the seal,
maintaining separation of the rotating versus nonrotating seal faces (typically but not always forming
an axial plane) by a very small clearance. Because small abrasive particles can enter into this zone, typi-
cally at least one of the faces, and sometimes both, is a very hard ceramic or refractory material. If the
application is such that some rubbing may occur at time, then the other surface is typically a grade of
carbon or carbon composite that would be able to self-lubricate the interface. The various designs and
implementations of mechanical seals are covered in considerable depth in other parts of this handbook.

60.1.3  Rotor Systems


A rotor “system” of a turbomachine typically involves the shaft, the fluid impellers (or in the case of an
axial machine, the “bladed disks”), seals to prevent fluid leakage from one stage to another or from the
discharge of a stage back to its inlet (usually some form of labyrinth seal), seals to prevent leakage to or
from the environment (often these are mechanical seals), the radial and axial thrust bearings, possibly
an axial thrust-balancing hydrostatic device (especially in multistage machines in order to minimize
the size of the thrust bearing required), and the machine casing or “frame” or “housing,” including such
load bearing subassemblies such as nonrigid bearing housings. If the casing/housings are very massive
and rigid compared to the rotor, the rotor system is often truncated to include only the simplified rotat-
ing components in any dynamic analysis of the rotor behavior.

60.1.3.1  Rotordynamic Analysis


Rotordynamic analysis includes calculation of torsional and lateral (and sometimes axial) natural fre-
quencies and mode shapes, known as a “critical speed analysis.” If any natural frequencies are predicted
to be within some margin, often ±15%, of known strong excitation frequencies such as caused by imbal-
ance (1× running speed), or if such excitations pass through a poorly damped natural frequency on the
way up to full-speed operation, a more comprehensive analysis is typically performed. Such analysis
usually includes a forced response analysis, in which the quantitative vibration response at key locations
(e.g., bearings, seals, and blade tip clearances) is calculated based on worst-case excitation forces and
minimum anticipated damping. The results of this are compared to appropriate specifications, such as
ISO 7919 [7] for shaft vibration and ISO 10816 [8] for bearing housing motion. If damping is not high
for a mode with a frequency below running speed, a rotor stability analysis should be performed as well.
Typically, this is a damped eigenvalue analysis in which the net modal damping for subsynchronous
(below running speed) natural frequencies is calculated and interpreted in terms of logarithmic decre-
ment (“log dec”), approximately equal to two pi times the modal damping divided by the critical damp-
ing. Depending on the machine and the analyst, typically a log dec below somewhere in the range of
0.1–0.2 is considered at risk of rotordynamic instability. Theoretically, instability actually occurs when
true log dec becomes equal to or less than 0.
60-6 Design for Lubrication and Tribology

In an operating machine in an actual installation, or at least on the manufacturer’s test stand, rotor-
dynamic characteristics and overall system vibration levels can be determined by test, as discussed in
Section 60.2.

60.1.3.2  Bladed Disk Evaluation


In axial turbomachinery, in which the process fluid moves axially rather than radially (more or less)
through the machine, the part that acts on the fluid or is acted on by the fluid is known as a bladed
disk. If the blades and disk are integral, as is often true in smaller gas turbines, the name is shortened
to “blisk.” A blisk has the simplicity of fewer parts but has the disadvantage that no relative motion can
occur at the zone where each blade is attached to the pancake-like or drum-like disk. Such attachments
have various designs, with the most common being a “dovetail” or the similar Tee-root, with one “root
leg” on each side and another common attachment possessing multiple root legs that form a “fir tree,”
where either the blade or the disk can form the exterior shape of a Christmas tree, and the opposing part,
disk or blade, respectively, has an interior similar shape. In the case of both dovetails and fir trees, and in
simpler designs where a pin or dowel is driven crosswise through radial root legs, the blade surfaces that
interface with the disk attachment and face the outer diameter (OD) are loaded by centripetal force on
the blade’s mass, and thermal distortion can force these interface loads to skew or even reverse in local
areas. As a result, some zones are lightly enough loaded that they microslip under buffeting loads for
the pressure and velocity fields that are the unsteadiness that represents even “steady” operating condi-
tions in a turbomachine of any kind, particularly aerospace turbomachinery of any kind, and industrial
turbomachinery where the unsteadiness is exacerbated at off-design.
This suggests the negative result of rapid fretting wear at the root. However, designers have learned
through experience that the materials and/or coatings will tolerate this microslipping with little or no
fretting wear. In fact, they have converted a possible problem into an advantage, in that root interfaces
are recognized as an important source of bladed disk damping which assists in minimizing vibration
response of the bladed disks when vane or nozzle pass harmonic frequencies are close to a bladed disk
natural frequency. If this is insufficient for vibration control at all operating conditions, other rubbing
interfaces are often introduced or taken advantage of if they serve another purpose such as gas path
sealing. For example, the shroud bands common in steam turbines, shrouds at the blade OD or at “part-
span” and the “damping wires” present in steam turbines, large turbochargers and expanders, and some
gas turbines (even with blisks) can provide substantial damping, and fortunately in an increasing man-
ner if a vibration mode becomes active enough to cause slip.
Other damping modes being experimented with at the present time include electromagnetic and
enhanced aerodynamic forces opposing blade motion of relative vibration and piezoelectric or encased
powder components in the blade body or in the root that are “exercised” by any blade vibration.
Resonant vibration control is key to bladed disk reliability because literally thousands of natural
frequencies exist between zero and the second harmonic of vane passing frequency. It was once believed
that the only bladed disk modes that needed to be avoided were the cantilever blade modes, in which the
blade could be analyzed like a swimming pool diving board. Starting with Grinsted’s experiments in the
1940s [9], and now with the assistance of the finite element analysis (FEA) method of computer-based
analysis of complex mechanical components and systems, engineers have come to understand that each
blade cantilever mode acts as a single degree of freedom in the bladed disk system; the total number of
modes will be the same as the total number of degrees of freedom, including not only the number of
cantilever modes in the frequency range of interest times the number of blades on a disk, but also the
number of “plate” modes of the circular disk, biased at the rim by the blade mass as it participates in that
particular mode. Any shrouds likewise have their own degrees of freedoms and therefore modes and
interact with other bladed disk modes if those bladed disk mode frequencies are similar in value to the
shroud mode in question. These bladed disk mode frequencies involve the blades moving tangentially or
axially or twisting and the disk material either cooperating in that motion or resisting it. This results in
the blades and/or disk developing mode shapes that look a bit like multi-petaled flowers, with one petal
Turbomachinery Tribology 60-7

moving toward the viewer and the adjacent one moving away. Between the petals are so-called nodal
diameters of zero vibration. Conversely, there are circularly lines of zero vibration at various diameters,
known as “nodal circles.” With some exceptions beyond the scope of the present discussion, each pos-
sible combination of nodal diameters and circles has two orthogonal (independent motion) natural
frequencies associated with them.
The net result is as follows: many, many modes, typically a large number of which have frequencies
close to significant excitation frequencies such as vane (“nozzle”) passing frequency, first and second (at
least) harmonics. Fatigue damage is typically prevented by one of several issues: (1) the excitation force
is weak, for example, by special shaping or distribution (e.g., clocking at uneven angles) of the vanes;
(2) the damping is sufficiently high; or (3) the distribution of the excitation force (typically pulsating
static pressure field) in space has a pattern that is poorly matched to the natural frequency’s mode shape,
causing an effect a bit like two children on a seesaw trying to both move up or both move down at the
same time. The potential for the latter situation is best determined with the aid of Grinsted’s interfer-
ence diagram (natural frequency versus “nodal diameter” of a given bladed disk mode, with excitation
speed range represented by separate skew lines), in the author’s experience. Most analysts prefer a form
of the Campbell diagram, in which natural frequency is computed against rotor speed, with excita-
tions represented by ray lines from the origin, and intersections with natural frequencies only counted
if engine order (i.e., 1xRPM = 1EO, 2xRPM equals 2EO, etc.) is equal to nodal diameter number. The
Campbell diagram does not easily represent so-called sum-and-difference excitations, in which nozzle
pass frequencies can excite the bladed disk not only when number of nozzles equals nodal diameter, but
also, for example, when the absolute value of the number of nozzles minus the number of blades equals
nodal diameter number. The latter has been responsible for many unexpected bladed disk failures, in
the author’s experience.
In addition to potential fatigue problems, if insufficient damping is present when a strong excita-
tion lines up with mode shape, tribology also is important to blade health in terms of erosion of the
surfaces of the blades, blade platform (to which the root is attached), stator vanes/platforms, disk cavity
(between platforms/stage pieces), and flow passage end walls. Erosion in these zones can be caused by
particulate matter in the flow stream (e.g., sand particles or, in certain areas of the world, volcanic ash),
by trace amounts of corrosive chemicals in the hot-gas flow (e.g., “sulfidation” in aircraft gas turbines),
by two-phase fluids (e.g., precipitation of liquid in refrigeration compressor gas flows at off-design
conditions), and by cavitation (either as part of rain drop erosion or due to insufficient net positive
suction head [NPSH] in pumps and hydraulic turbines). Strategies to control erosion include using,
of course, erosion-resistant materials, such as Stellite and super-austenitic stainless steel, but certain
types of composite or ceramic coatings may also provide a degree of thermal protection as well in the
hot sections of gas turbines. However, material improvements have limits, and other approaches are
also effective.
The use of computational fluid dynamics (CFD) has become a useful approach to allow designers
to trade slight performance compromises in order to optimize flow path streamlines in a manner that
alters either local velocities or impingement angles of the flow on the affected surfaces. In some cases,
the erosion streamline optimization actually improves efficiency of the stage. Decreases in erosion rates
of over an order of magnitude can be achieved in worst-case locations.
In pumps and hydraulic turbines, cavitation erosion can be significantly reduced or eliminated by
increasing the local static pressure to ensure it remains above vapor pressure at the temperature of the
pumped fluid. Cavitation normally occurs at higher flows, perhaps above the design flow, since high flow
results in an increased pressure drop between the fluids’ original source (e.g., a sump) and the pump
inlet or “suction.” For parallel, consistent flow, this can be achieved with a booster pump, for example,
when the extra cost and complexity are practical. However, sometimes cavitation occurs at flow rates
below design flow, due to local flow velocity increases or vortices that can be eliminated, rather than
pressure being too low in the bulk flow. In such cases, good design practice or, in more-complicated
situations, CFD can be used to “smooth” the flow and eliminate the low-pressure pockets. The same is
60-8 Design for Lubrication and Tribology

true when pumps are run at flows far below what is designed in that the angle of attack is so poor at the
pump blades (also called vanes in a pump) that stalling takes place, like on an airplane wing cocked at
too high an angle into the airstream.
The stalling typically leads to vortex formation that resembles a small tornado or cyclone, which
twists off the suction side of the blade and spins the vortex tip across the surface of the pressure side of
the neighboring blade. This is an extreme form of what is known as “suction recirculation” and can lead
to cavitation damage unexpectedly on the high-pressure side of the blade rather than the low-pressure
or “suction” side, where “classical” cavitation would occur if NPSH (i.e., absolute pressure at the inlet
divided by fluid density) was simply too low at higher flow rates. At even lower flows, the process tran-
sitions to the discharge of a centrifugal pump’s impeller, causing “discharge recirculation,” which can
(even more surprisingly) cause cavitation erosion pitting on the pump impeller OD blades or even on
the discharge volute tongue or diffuser vanes. Although sufficiently high NPSH can cure this, the level
of pressure is so high this is seldom practical. Therefore, the best approach is to ensure the inlet flow
angle is more consistent with the suction blading of the impeller, avoiding stalling. Once again, CFD can
be valuable in accomplishing this. Alleviating stall recirculation is not only useful to avoid cavitation
erosion, but also avoids greatly increased pressure oscillations and resulting vibration of the pump rotor
system, reducing wear of bearings and seals.

60.1.3.3  Integral versus Built-Up Rotors


In performing dynamic analysis of overall rotor systems, the mathematics to predict rotor critical
speeds is fairly well understood if the bearing and seal rotordynamic coefficients can be accurately pre-
dicted. This latter aspect is not trivial and should be based on either component-specific experiments,
calibrated bulk flow models (e.g., from the rotordynamic consortia at University of Virginia, Virginia
Tech, or Texas A&M), or comprehensive 3-D CFD models run at various eccentricities. However, even
the mechanical shaft/sleeves/disks as an assembly can be a challenge if the rotor is “built up” from the
components, rather than being made from a single piece of steel (“integral”). The API-684 guideline
(produced by the American Petroleum Institute, Washington, DC [10]) for turbomachinery rotordy-
namics recommends in built-up rotors that loose fits be ignored in lateral rotordynamic analysis, and
suggests the same for press-fit components as well, since centripetal forces tend to relieve radial fit-ups,
although they warn that the true result will be machine specific. The author’s experience is that even
light press fits cause rotor systems to behave as if they are integral for torsional dynamics, but it takes a
moderately heavy press fit (class 3 or heavier) for the same to be true for lateral rotordynamics. Therefore,
the author agrees with API relative to effect of light press fits on lateral rotordynamics, but otherwise
recommends that press-fit rotor components be treated as integral. If the components are loose fit, the
typical practice, as also recommended by API, is to consider the full inertia of the component but to
ignore its stiffness. In the case of large effective diameter components (e.g., impellers) held firmly axially
at the front and rear by components rigidly attached to the shaft, for example, turbocharger compres-
sor wheels, the wisdom of this approach is questionable and should be confirmed by a 3-D FEA before
construction or an operational modal test after build.

60.2  Mechanical Testing, Diagnostics, and Prognostics


Instrumentation probes used for mechanical reliability and vibration testing typically include piezo-
electric accelerometers on stationary components and eddy current proximity probes mounted within
bearing housings observing the motion and eccentricity of the rotating shaft in two perpendicular
directions. Vibration versus time from the proximity probes is then evaluated in the form of either
vibration amplitude versus time linear plots or two-dimensional (2-D) “orbit” plots that exhibit the shaft
centerline’s motion perpendicular to the centerline. In either case, the levels are compared to available
clearances, factoring in mode shape at the primary frequency in the case of “flexible” rotors (rotors with
Turbomachinery Tribology 60-9

the first bending natural frequency close to or below the running speed). Depending upon the standard
invoked, generally somewhere between 1/3 and 2/3 of the available clearance is permitted before the
machine is considered in “alarm.” Most standards require the machine to be shut down or “tripped” if
¾ or more of the clearance is utilized. Vibration data on the stationary structure, typically the bearing
housing, may be used in addition or as an alternative to shaft vibration displacement. In the case of bear-
ing housing, vibration typically is considered in terms of its velocity level rather than the acceleration
form in which it is acquired, thereby more strongly emphasizing lower- versus high-frequency com-
ponents. A thorough discussion of this topic is provided by Marscher [11]. The allowable velocity limit
depends upon the standard used, such as ISO 10816 [8], as well as the machine type and application. A
typical acceptance limit range is 0.12–0.20 in./s (3–5 mm/s) rms.

60.2.1  Test Methods


There are many international standards for turbomachinery. In addition to various company propri-
etary standards and specifications, open standards include the ASME Power Test Codes (e.g., PTC-6 for
steam turbines [12]); the ASME Boiler & Pressure Vessel Code (e.g., Section XI (11) for machinery vibra-
tion testing in pressure systems [13]); the API machinery specific codes such as API-610 for centrifugal
pumps [14], API-612 for steam turbines [15], API-616 for gas turbines [16], and API-617 for compressors
[17]; and the ISO machinery vibration standards ISO 7919 for rotor vibration [7] and ISO 10816 for sta-
tionary component vibration [8].

60.2.2  Diagnostics: Methods of Detecting Degree of Degradation


Methods of detecting mechanical problems in turbomachinery vary depending upon the machine and
its application and value. Vibration analysis, acoustic analysis, electric current spectrum analysis, oil
monitoring analysis (OMA), and thermography are all useful in specific situations. Other chapters in
the handbook cover OMA, and thermography and electric current spectrum analysis details are outside
the scope of the handbook. Acoustic analysis will be considered a subset of vibration.
Vibration in high-value turbomachinery typically involves measuring shaft motion in two orthogo-
nal directions near each bearing housing, in addition to some representative measurements on station-
ary structures (typically bearing housings) using accelerometers. In lower-value machinery, typically
only bearing housing readings are taken, and perhaps only by portable “walk-around” instrumentation.
Ideally, for good diagnostic potential, such measurements should be in three orthogonal directions on
each bearing housing of the driver as well as the driven machine. If the decision is to measure only one
location, then the location most likely to exhibit symptoms for the greatest variety of problems is hori-
zontal vibration on the driven end bearing housing of the driven machine.
Typical mechanical problems exhibited by vibration frequency spectra and/or vibration versus time
2-D “orbits” are illustrated in Figures 60.1 through 60.6 [18].

60.2.3  Prognostics
Once a problem has been detected, a critically important problem becomes the prediction of remaining
useful life. Basically, such predictions typically fall into two regimes:
1. Prediction of fatigue life, given the steady and oscillating load and therefore stresses
2. Prediction of wear, due to various tribological environments and loading cycles
The first is outside the scope of this chapter but is discussed well in various ASM and SAE handbooks
on this subject (e.g., [19–21]). The second is the subject of ongoing research by many organizations (e.g.,
Hydraulic Institute [22]) and government agencies and tends to be very system and component specific.
60-10 Design for Lubrication and Tribology

Orbit:
1 mil
Normal High 1 × N

Common causes:
a. Mechanical unbalance
b. Misalignment
Spectrum: (usually high 2 × also)
c. Bent shaft
3 mils
Harmful effects:
Vibration

2 mils a. Internal rubbing


on bearings and seals
b. Overload of rolling
1 mil
Normal element bearings
range
1×N 2×N 3×N 4×N 5×N
Frequency

FIGURE 60.1  1× running speed issues. (Reproduced from Booser, E.R. (ed.), Rotating Machinery Vibration and
Condition Monitoring Chapter by W.D. Marscher, Tribology Data Handbook, CRC Press, New York, pp. 944–955, 1997.
With permission.)

Orbit:

Or

Spectrum:
3 mils
Vibration

2 mils

1 mil

1× 2× 3× 4× 5×
Frequency
Common causes: a. Mechanical misalignment
b. Looseness in bearing retention
c. Severe shaft or bearing
housing crack
Harmful effects: a. Internal rubbing
b. Coupling wear
c. Shaft fatigue

FIGURE 60.2  2× running speed issues. (Reproduced from Booser, E.R. (ed.), Rotating Machinery Vibration and
Condition Monitoring Chapter by W.D. Marscher, Tribology Data Handbook, CRC Press, New York, pp. 944–955, 1997.
With permission.)

A representative success has been in the area of rolling element bearings, as covered in detail elsewhere in
this handbook. In short, depending upon the specific style of such bearings, the statistical life of the bear-
ing (the so-called L10 or B10 life) has been found proportional to the steady load as well as oscillating load
on the bearing, to a power somewhere between 3 and 3.3. Therefore, if a given degree of vibration can be
successfully reduced to a bearing load for a given machine, the bearing life can be predicted statistically.
For example, according to the author’s research, detailed vibration measurements can be taken in several
key locations and then extrapolated to the bearing location with the aid of FEA of the machine type in
question. A similar approach can be taken also for clearance utilization at labyrinth seals, as well as within
journal bearings, and degree of dynamic misalignment from vibration permissible in mechanical seals.
Turbomachinery Tribology 60-11

Orbit:

Spectrum:
3 mils

Vibration
2 mils
1 mil

1× 2× 3× 4× 5×
Frequency
Common causes: a. "Gap B" too tight
b. Discharge recirculation
c. Flat or damaged volute tongues
d. Internal resonance of
diffuser walls or vanes

Harmful effects: a. Fatigue in instrumentation


wire connections or drain
pipe connections
b. If internal resonance is the
cause, fatigue cracking of the
resonating part

FIGURE 60.3  Vane or blade passing frequency issues. (Reproduced from Booser, E.R. (ed.), Rotating Machinery
Vibration and Condition Monitoring Chapter by W.D. Marscher, Tribology Data Handbook, CRC Press, New York,
pp. 944–955, 1997. With permission.)

Orbit: Type 1: Type 2:

"Half-
speed"

Spectrum:
× % of running speed

1× 2× 3× 4× 5×
Subsynchronous peak
Possible causes:
Type 1:
a. x = 40% – 49% : Bearing instability
b. x = 50% Exactly: Severe rub
(or exactly 1/3 and 2/3)
c. x = 5% – 30% : Diffuser stall
Type 2:
d. x = 65% – 95% : 1. Impeller stall
2. Suction recirculation

FIGURE 60.4  Subsynchronous issues. (Reproduced from Booser, E.R. (ed.), Rotating Machinery Vibration and
Condition Monitoring Chapter by W.D. Marscher, Tribology Data Handbook, CRC Press, New York, pp. 944–955, 1997.
With permission.)

60.3  Closure
Turbomachinery rotor systems depend upon a variety of tribological processes in order to function reliably.
It is important that the rotor maintain its position within tight clearances in order to achieve high efficiency,
since larger than required clearances would imply leakage of hard-fought-for pressure back to its low-pressure
state. In addition, vibration must be kept low, or fretting wear will occur at fit-up locations, especially in built-
up rotors and nonintegral bladed disk blade roots. The maintenance of proper position/eccentricity and low
vibration in a high-speed rotor implies the need for excellent bearings and bearing lubrication. Finally, the
need for maximum thermodynamic efficiency in this energy-conscious world emphasizes the criticality of the
60-12 Design for Lubrication and Tribology

Orbit:

Spectrum:
3 mils

Vibration
2 mils

1 mil

1× 2× 3× 4× 5× 6× 7×
Frequency
Common causes: a. Scratch in the shaft or
chrome plate thickness
variations at Bentley probe
b. Rubbing
c. Very loose "rattling" internal
component, such as impeller
Harmful effects: a. From cause (a), none
(except stomach problems!)
b. Premature wear from (b)
c. Progressively looser fits and
worsening situation from (c)

FIGURE 60.5  Time-unsteady high-harmonic issues. (Reproduced from Booser, E.R. (ed.), Rotating Machinery
Vibration and Condition Monitoring Chapter by W.D. Marscher, Tribology Data Handbook, CRC Press, New York,
pp. 944–955, 1997. With permission.)

∆t
Finput

τ t

Orbit:

Spectrum:
Xresponse

2π f
τ

∆t
Possible causes:
a. Rub consistently over part of orbit
b. Pinched or misaligned seal
c. Journal bearing worn to oval shape

FIGURE 60.6  Time-steady high-harmonic issues. (Reproduced from Booser, E.R. (ed.), Rotating Machinery
Vibration and Condition Monitoring Chapter by W.D. Marscher, Tribology Data Handbook, CRC Press, New York,
pp. 944–955,, 1997. With permission.)

maintenance of running clearances as well as blade and vane flow-path surface shapes, requiring good wear
and erosion control. This chapter has discussed the evaluation procedures and criteria for these issues, as an
aid to those involved in ensuring the mechanical reliability of high-performance turbomachinery of all types.

References
1. Marscher, W.D., Bearing Application and Lubrication Chapter, Modern Marine Engineer’s Manual,
3rd edn., Cornell Maritime Press, Centreville, MD, 1999.
Turbomachinery Tribology 60-13

2. Massey, I., Subsynchronous vibration problems in high speed multistage centrifugal pumps,
Proceedings of the 14th Turbomachinery Symposium, October, Houston, TX, pp. 11–16, 1985.
3. Allaire, P.E. and Flack, R., Dynamics of short eccentric plain seals with high axial Reynolds number,
J. Spacecraft AIAA, 15(6), 341–347, 1978.
4. Childs, D., Finite length solutions for rotordynamic coefficients of turbulent annular seals, ASME/
ASLE Lubrication Conference, ASME Paper 82-LUB-42, October 1982.
5. Black, H.F., Effects of fluid-filled clearance spaces on centrifugal pump vibrations, 8th Turbomachinery
Symposium, Texas A&M University, College Station, TX, pp. 29–37, c. 1979.
6. Marscher, W.D., Analysis and test of multistage pump wet critical speeds, Tribology Trans., STLE,
34(3), 445–457, 1991.
7. ISO 7919 Machinery vibration standards for rotor vibration, International Standards Organization,
Geneva, Switzerland, 1996.
8. ISO 10816 Machinery vibrtion standards for stationary component vibration, International
Standards Organization, Geneva, Switzerland, 1995.
9. Grinsted, B., Nodal pattern analysis, Proc. IMechE, 166, 309–326, 1952.
10. API 684, 1st edn., Tutorial on the API Standard Paragraphs Covering Rotor Dynamics and Balance
(An Introduction to Lateral Critical and Train Torsional Analysis and Rotor Balancing), American Pet.
Institute, Washington, DC, c. 1996.
11. Marscher, W.D., The relationship between pump rotor system tribology and appropriate vibration
specifications for centrifugal pumps, Proceedings of the IMechE 3rd European Congress on Fluid
Machinery for the Oil and Petrochemical Industries, May, The Hague, the Netherlands, 1987.
12. ASME Power Test Code PTC-6 for Steam Turbines, American Society of Mechanical Engineers,
New york, 2004.
13. ASME Boiler and Pressure Vessel Code, Section XI, American Society of Mechanical Engineers,
New york, 2010.
14. API-610 for Centrifugal Pumps, 11th edn., American Pet Institute, Washington, DC, c. 2011.
15. API-612 for Steam Turbines, 5th edn., American Pet Institute, Washington, DC, c. 2003.
16. API-616 for Gas Turbines, 5th edn., American Pet Institute, Washington, DC, c. 2011.
17. API-617 for Compressors, 7th edn., American Pet Institute, Washington, DC, c. October 2002.
18. Booser, E.R. (ed.), Rotating machinery vibration and condition monitoring chapter by W.D.
Marscher, Tribology Data Handbook, CRC Press, New york, pp. 944–955, 1997.
19. ASM Handbook Vol. 19, Properties and Selection: Fatigue and Fracture, American Society of Metals,
Cleveland, OH, 1996.
20. ASM Handbook Vol. 18, Friction, Lubrication and Wear of Materials, American Society of Metals,
Cleveland, OH, 1992.
21. SAE Fatigue Handbook, 3rd edn., Society of Automotive Engineers International, Dearborn, MI,
c. 1997.
22. HI/ANSI Standards, 2010 edn., The Hydraulic Institute, Parsippany, NJ, c. 2000.
61
Natural and Artificial
Human Joints
61.1 Natural Human Joints..................................................................... 61-1
Physiology of Natural Synovial Joints  •  Tribology of Natural
Francis E. Kennedy Synovial Joints  •  Diseases and Injuries Affecting Tribological
Dartmouth College Performance of Natural Human Joints
61.2 Artificial Human Joints.................................................................. 61-4
Douglas W. Total Hip Prosthesis  •  Total Knee Prosthesis  •  Shoulder, Ankle,
Van Citters and Other Artificial Joints
Dartmouth College References................................................................................................... 61-16

61.1  Natural Human Joints


Some of the most effective bearings, that is, those with lowest friction and wear, known to man are those
in our own bodies. The human body has remarkable bearings between the moving bones of our arms
and legs that enable easy motion of the joints and provide trouble-free performance for decades in most
people. These bearings are characterized by the presence of layers of articular cartilage on the contacting
surfaces of the bones, and between the contacting surfaces there is a natural lubricant, synovial fluid,
which promotes low friction coefficient (on the order of 0.01) and negligible wear of the cartilage over
the lifetime of the individual. Because of the presence of the synovial fluid lubricant, the movable joints
are called synovial joints or diarthrodial joints.
There are many different synovial joints in the human body, and they differ from each other in the motion
that they allow. For example, the elbow joints are hinge joints, which allow free rotation in one plane, while
the hip and shoulder joints are ball-in-socket joints, which allow free rotation about all three axes. The most
complex joint is the knee joint, which allows rotation about two axes as well as sliding translation.

61.1.1  Physiology of Natural Synovial Joints


A schematic diagram of a typical synovial joint is shown in Figure 61.1 (Dumbleton, 1981). The joints
are enclosed in a fibrous capsule, within which is found a synovial membrane, which secretes the syno-
vial fluid and also provides nutrients to the joint’s tissues. The contacting surfaces of the bone are
covered by a thin layer of hydrated soft tissue, articular cartilage, which acts as the bearing surface.
In addition, some joints, such as the knee, may contain a pad of fibrocartilage called the meniscus,
which helps separate the bones, thus providing load support and helping maintain proper joint motion.
Stability of the joint is maintained by ligaments and tendons that hold the bones together, and motion
is provided by muscles that are attached to the bones. Blood is supplied to the joint’s tissues by arteries,
although the only portion of the cartilage that receives blood is at the rear surface where it attaches to
the bone; the remainder of the articular cartilage is avascular. There are no nerves in either articular
cartilage or the synovial membrane, although nerves are present in the other tissues of the joint.

61-1
61-2 Design for Lubrication and Tribology

Periosteum

Nerve
Artery
Muscle
Ligament

Fibrous capsule

Synovial fringe

Synovial
membrane

Meniscus
Fat pad

Tendon Articular cavity


containing
Bursal wall synovial
fluid

Epiphyseal
line Articular
cartilage

FIGURE 61.1  Schematic diagram of a typical synovial joint. (Reprinted from Tribology of Natural and Artificial
Joints, J.H. Dumbleton, Copyright 1981, with permission from Elsevier.)

The most important components of a synovial joint from a tribological point of view are the syno-
vial fluid and the articular cartilage, since they are responsible for the low friction and low wear of
natural synovial joints. Synovial fluid is a viscous fluid that is produced in the synovial membrane.
Its viscosity ranges from about 0.05 to 1.5 Pa s and is a strong function of patient age and health,
as well as the velocity of joint motion. Synovial fluid contains hyaluronic acid in a concentration of
protein (mucin) (Dumbleton, 1981). Because of the long-chain protein molecules, the fluid’s viscous
behavior is non-Newtonian; it is a thixotropic fluid, that is, its viscosity decreases and the fluid
thins when subjected to increased shear rate (Schurz and Ribitsch, 1987). In addition to hyaluronic
acid, the fluid contains other components, including lubricin, albumin, and phospholipids, that have
important effects on its tribological behavior. Based on work by Schurz and Ribitsch (1987), Hills
(1989), Murakami et al. (2007), and others, Trunfio-Sfarghiu et al. (2007) summarized those effects
as follows:
• Hyaluronic acid tends to increase the viscosity of healthy synovial fluid by up to a factor of 2 at low
shear rates, but at high shear rates, the effect of hyaluronic acid on viscosity is not very significant.
• Albumin modifies the rheology of the synovial fluid and helps protect against cartilage wear.
• Phospholipid bilayers are major contributors to the effectiveness of synovial fluid as a boundary
lubricant.
• Lubricin helps fix the phospholipid bilayers to the surface of healthy cartilage and protects them
against oxidation.
Natural and Artificial Human Joints 61-3

In addition to its critical role in lubricating human joints, synovial fluid also provides the necessary
nutrients for articular cartilage, protects the cartilage against enzyme activity, and acts as a medium for
getting blood supply to the joint (Ateshian and Mow, 2005).
The articular cartilage that covers the ends of articulating bones in synovial joints is a unique
­material. In a healthy human joint, the cartilage thickness ranges between 0.5 and 5 mm (Ateshian and
Mow, 2005). The tissue is composed of a porous and permeable fibrous solid matrix, along with inter-
stitial water and dissolved electrolytes. In a typical healthy joint, the articular cartilage contains about
68%–85% water, 10%–20% collagen fibers, and 5%–10% proteoglycans (by wet weight) (Mow et al.,
2005). Water content increases from the deepest zone in the cartilage to the superficial zone (Ateshian
and Mow, 2005). The composition of menisci is similar, but with more collagen and fewer proteoglycans
(Mow et al., 2005). In general, the solid matrix of articular cartilage consists of a network of oriented
collagen fibers in a high concentration of proteoglycan gel (Mow et al., 2005). The primary function of
the collagen fibers is to provide stiffness and tensile strength to the cartilage, while the proteoglycans
enable the cartilage to resist compressive stress. The structure of the matrix varies through the depth of
the cartilage, with the collagen fibers being densely packed and oriented parallel to the surface in the
top layer, less densely packed and more randomly oriented in the middle zone, which comprises 50%
or more of the cartilage thickness, and organized in fiber bundles that are attached to the subchondral
bone in the deepest zone (Dumbleton, 1981; Mow et al., 2005). In general, because of the complex orga-
nization of the collagen network and proteoglycans, the solid matrix exhibits anisotropic, inhomoge-
neous, viscoelastic behavior that is best analyzed by a poroviscoelastic or biphasic mechanical model
(Mow et al., 2005). Like other connective tissue, articular cartilage is metabolically active and undergoes
a continuous process of degradation and synthesis during which the solid matrix is renovated in healthy
individuals (Creamer and Hochberg, 1997).

61.1.2  Tribology of Natural Synovial Joints


Synovial joints are subjected to a wide range of loading conditions and motion during daily activi-
ties and are required to provide low friction and negligible wear for years in healthy individuals. For
example, during normal walking motion, the hip and knee joints encounter contact forces of several
times body weight at heel-strike and toe-off points in the walking cycle, and those joint forces increase
when descending stairs or running downhill. These forces may be repeated several million times each
year of a person’s life. For most individuals, the joints are able to provide extremely low friction (friction
coefficients ranging from 0.005 to 0.04) have been measured in hips and knees (Unsworth et al., 1975),
and they do this with almost no wear.
For several decades, researchers have attempted to understand how human synovial joints achieve
their extraordinary tribological performance. Attempts to treat synovial joints as hydrodynamic bear-
ings, such as that of Higginson and Unsworth (1981), have shown that the predicted fluid film thickness
in a typical synovial joint, considering hydrodynamic action alone, will be in the range of 10−8–10−7 m,
much too thin to separate the cartilage surfaces. Subsequent elastohydrodynamic (EHL) and micro-
EHL models, such as that of Dowson and Jin (1986), which included elastic deformation and roughness
of the cartilage surfaces, showed that micro-EHL may be a viable lubrication mechanism for synovial
joints under low load conditions. However, even with the addition of squeeze-film effects, EHL film
thicknesses are too small to be sustained without cartilage contact in conditions of high loads and slow
velocities, so mixed or boundary lubrication conditions may prevail during much of the motion cycle
in human hip and knee joints. Studies of the boundary lubrication characteristics of the constituents
of synovial fluid have found that the lipidic components (surface-active phospholipids, SAPL) are the
major contributor to the effectiveness of synovial fluid as a boundary lubricant in natural knee joints
(Hills and Crawford, 2003). The SAPL adsorb on the cartilage, forming a thin hydrophobic layer that
reduces friction coefficient to less than 0.01 (Hills and Crawford, 2003). Other recent studies have shown
that the effectiveness of the lipidic bilayers is augmented by the synergistic presence of other components
61-4 Design for Lubrication and Tribology

of the synovial fluid, such as lubricin, which help maintain a boundary lubricant film between cartilage
surfaces (Trunfio-Sfarghiu et al., 2007). Although the boundary-lubricating capability of synovial fluid
can play an important role in maintaining low friction in human joints, a boundary lubricant alone
cannot account for the very low friction between cartilage surfaces because cartilage has been shown
to produce elevated friction coefficient (up to 0.04) after prolonged loading, only to return to very low
values (<0.01) upon allowing sufficient time for recovery, without the addition of fresh boundary lubri-
cant (Ateshian and Mow, 2005). To explain this discrepancy, another mechanism has been proposed,
that of “weeping” lubrication (McCutcheon, 1962) or lubrication by pressurization of interstitial fluid
(Ateshian and Mow, 2005). This mechanism relies on the fact that interstitial fluid in the pores of the
articular cartilage is pressurized when the joint is loaded. Studies have shown that the in vivo interstitial
fluid pressures during normal walking average about 5–6 MPa, rising to over 25 MPa when the joint is
heavily loaded (Morrell et al., 2005). Some of this pressurized fluid is exuded from the surface ahead of
the moving contact during cartilage deformation, and the exuded fluid can then help keep the cartilage
surfaces separated when the contact moves forward, thus keeping the friction low. This mechanism was
modeled in detail by Ateshian and coworkers, using biphasic models of cartilage deformation behavior
(Ateshian et al., 1998; Ateshian and Mow, 2005). In summary, it appears that several different lubrica-
tion mechanisms, primarily boundary lubrication and lubrication by pressurized interstitial fluid, con-
tribute to the effective lubrication of human synovial joints, with the result being low friction coefficient
(about 0.005–0.02), with very little damage to cartilage in most individuals throughout their lives.

61.1.3 Diseases and Injuries Affecting Tribological


Performance of Natural Human Joints
Failure of natural joint may occur due to injury, such as a torn ligament or a torn meniscus, or to disease,
such as arthritis. From a tribological point of view, the most significant failure mode is osteoarthritis, a
disease which is caused by the breakdown and wear of the articular cartilage of one or more of the joints.
Osteoarthritis is also known as degenerative arthritis.
Osteoarthritis is the most common source of joint pain, particularly in older persons. It occurs when
the normally balanced process of degradation and synthesis of the solid matrix of the cartilage is upset,
resulting in degradation (e.g., wear and fracture) of the cartilage. A variety of factors affect this deg-
radation, including age, weight, genetics, gender, disease, trauma, and mechanical loading (Creamer
and Hochberg, 1997). It is generally felt, though, that mechanical loads are a primary factor and there
are several proposed mechanisms for this detrimental mechanical action. One proposed mechanism is
repeated high pressure of the interstitial fluid, resulting in cyclic higher stress in the solid matrix and
gradual deterioration of the cartilage through fibrillation and crevassing (Morrell et al., 2005). An alter-
native mechanism is through high impact stress in the solid matrix, resulting in failure of the cartilage
matrix, as evidenced by fissures in the cartilage (Arteshian and Mow, 2005). In either case, the cartilage
is gradually removed, resulting in bone/bone contact, joint pain, and decreased joint function, often
accompanied by swelling of the joint.
Some joint problems, especially those caused by injury, can be remedied by relatively minor surgery
(e.g., ligament repair), but severe arthritis is usually best treated by a total joint replacement.

61.2  Artificial Human Joints


61.2.1  Total Hip Prosthesis
Hip joint replacement surgery generally is performed to relieve the pain of arthritis or to treat severe
fractures of the hip. Although there had been earlier attempts to replace parts of the human hip joint
with artificial components, most modern total hip prostheses are derived from the pioneering low
friction arthroplasty designed and implanted by Sir John Charnley in Lancashire, England, in 1962.
Natural and Artificial Human Joints 61-5

FIGURE 61.2  Typical artificial total hip prosthesis.

Since that time, hip replacement surgery has become a widely used and successful orthopedic surgi-
cal procedure. In the year 2005, over 250,000 total knee replacement surgeries were performed in the
United States, and that number is predicted to rise to over 380,000 annually by the year 2020 (Kurtz
et al., 2007).
A total hip replacement surgery, or total hip arthroplasty (THA), involves the removal of the ace-
tabulum (the articular cartilage-covered socket in the pelvic bone) and the femoral head (the articular
cartilage-covered ball at the top of the femur) and replacing them with man-made components (Mayor
and Collier, 1994). A typical total hip prosthesis is shown in Figure 61.2.
A modern total hip replacement consists of an acetabular component that fits in the pelvis where it
replaces the acetabulum, and a femoral component that extends into the medullary canal of the femur
and has a ball-like head that replaces the femoral head. The acetabular component may have a backing
shell that is embedded in the pelvis, coupled with a wear-resistant bearing surface, called the cup or
liner, that either is part of the backing shell or is snapped in place in the shell. The femoral component
may be a single piece or may have a separate femoral stem upon which is fit a ball with a smooth surface;
a Morse taper is used to enable the femoral ball to fit securely on the femoral stem.

61.2.1.1  Materials of Total Hip Prostheses


There are effectively three different material combinations used in current total hip prostheses. The
most common is a metal-on-polymer prosthesis featuring an acetabular bearing surface made of ultra-
high-molecular-weight polyethylene (UHMWPE) in a metallic socket generally made of a titanium
alloy (Ti-6Al-4V, ASTM 136). The ball of the femoral component is usually made of a cobalt-chrome-­
molybdenum alloy (CoCrMo, ASTM F75 or Vitallium™) that is quite hard and has a polished contact
surface (about 0.01 μm Ra). The composition of the CoCrMo alloy is generally approximately 60%–65%
Co, 25%–30% Cr, and 5%–7% Mo (Ratner et al., 2004). The femoral head may be integral with a CoCrMo
femoral stem or may be fit on a femoral stem made of either Ti-6Al-4V or CoCrMo. An alternative
61-6 Design for Lubrication and Tribology

TABLE 61.1  Mechanical Properties of Orthopedic Prosthesis Materials


ASTM Modulus of Ultimate Tensile Tensile Yield Hardness
Material Designation Elasticity (GPa) Strength (MPa) Strength (MPa) (GPa)
Polymers
Unirradiated UHMWPE ASTM F648 850 46–58 22
Irradiated UHMWPE 940 43–55 24
Highly cross-linked UHMWPE ASTM F2565 740 43–51 19–21
Ceramics
Al2O3 366 310 20–30
ZrO2 201 420 12
Metals
Stainless steel ASTM F138 190 930 241–820 1.3–1.8
CoCrMo ASTM F75 210–250 655–1275 207–950 3–4
Ti-6Al-4V ASTM 136 116 965–1100 897–1034 3
Source: Data from Collier, J.P. et al., Clin. Orthop. Relat. Res., 414, 289, 2003; Ratner, B.D. et al., eds., Biomaterials Science,
2nd edn., Elsevier Science, San Diego, CA, pp. 535–536, 2004; Currier, B.H. et al., J. Arthroplasty, 22, 721, 2007; Kurtz, S.M.,
The UHMWPE Handbook, 2nd edn., Elsevier, London, U.K., 2009.

metallic femoral component material is stainless steel, generally 316L stainless steel (ASTM F138), but
that material is not used as widely in the United States as is CoCrMo.
In recent years, several alternative material combinations for hip prostheses have been used instead
of the more common CoCrMo ball on UHMWPE acetabular cup, partly because of concerns about
wear of the UHMWPE material. One alternative is to replace the polyethylene acetabular component
with CoCrMo and to have a metal-on-metal (MOM), or CoCrMo-on-CoCrMo, bearing couple. In some
European countries, stainless steel-on-stainless steel hip joints are also available. The MOM hip joints
eliminate the problem of polyethylene wear, but raise some other issues, as will be discussed later. Two
other hip joint materials that have come into common use in recent years are ceramics, specifically
alumina (Al2O3) and zirconia (ZrO2). Ceramic femoral balls may be fit on metallic (e.g., titanium or
CoCrMo) femoral stems using a Morse taper fit, and these ceramic balls may be in contact with acetabu-
lar liners made of either UHMWPE, for a ceramic-on-UHMWPE joint, or ceramic, for a ceramic-on-
ceramic (COC) joint, or metal (CoCrMo), for a ceramic-on-metal (COM) joint.
The metallic acetabular and femoral components may be fixed in the pelvis and femur, respectively,
by either of two methods. Since the days of Sir John Charnley, most metallic components have been
cemented in place in the bones using poly(methylmethacrylate) (PMMA) bone cement. During the hip
replacement operation, PMMA bone cement is put in place in a doughy form, and it solidifies in place to
fill the space between metallic component and bone. In fact, bone cement is actually a grout and not an
adhesive, but it holds the metallic component securely in the bone (Mayor and Collier, 1994). In recent
years, there has been an increase in the use of noncemented or cementless prostheses, in which the
bone-contacting (nonarticulating) surfaces of the metallic component are manufactured with a porous
metallic surface into which bone can grow to fix the components in place (Mayor and Collier, 1994).
Frequently the porous surface is coated with hydroxyapatite in order to promote bone growth into the
components.
Typical properties of hip prosthesis materials are given in Table 61.1.

61.2.1.2  Kinematics, Loading, and Lubrication of Total Hip Prostheses


The hip joint is a ball-in-socket joint that is subjected to rotation about at least two axes during normal
activities. Studies have shown that motion within the contact area between femoral head and acetabular
liner is multidirectional in nature and wear of the polyethylene acetabular component has been found
to be significantly affected by the multidirectional motion (Bragdon et al., 1996; Wang et al., 1997). For
Natural and Artificial Human Joints 61-7

this reason, most testing of wear behavior of artificial hip materials is done either in multidirectional
hip simulators or in pin-on-disk test devices that simulate the multidirectional sliding motion (Saikko
and Ahlroos, 1999; Burroughs and Blanchet, 2001).
Loads in the hip joint vary considerably during a walking cycle, with the peak joint force ranging
from about four times body weight at the toe-off point during normal walking to a peak of about seven
times body weight during rapid downhill walking or running (Paul, 1976). The orientation of the resul-
tant force also changes during the walk cycle as the femur goes through flexion-extension, abduction-
adduction, and internal/external rotation motions. Current hip simulators attempt to duplicate the load
and orientation changes when testing hip prostheses.
During the implantation of the artificial hip joint, the synovial membrane is disrupted, and the syno-
vial fluid is drained, but the synovial membrane is reattached at the end of the operation. Within a few
months, the membrane is once again producing synovial fluid which lubricates the artificial joint. The
fluid film thickness in a hip joint replacement with metallic ball in UHMWPE socket was analyzed
by Jalali-Vahid et al. (2001) using an elastohydrodynamic lubrication (EHL) model for ball-in-socket
configuration. The predicted lubricating film thickness was generally too small in relation to the sur-
face roughness of the liner and head to maintain separation between the two articulating surfaces.
Therefore, one can expect that the contact would be in the mixed or even the boundary lubrication
regime during part of the cycle of daily activities. For this reason, the boundary lubricating capabilities
of the joint’s synovial fluid are important. Studies of the effect of joint fluid on the tribology of total hip
joint arthroplasty have shown that the lubricating capability of synovial fluid in metal-on-polyethylene
(MOP) hip arthroplasty is somewhat patient dependent, but generally proteins in the fluid are respon-
sible for reducing friction in this couple (Mazzucco and Spector, 2004). Recent work has found that the
boundary lubricants may be less effective with polyethylene implants than with natural cartilage or
even metal (Trunfio-Sfarghiu et al., 2007). Ceramic components may provide better attachment for the
boundary-lubricating molecules of synovial fluid than metallic or polyethylene components, leading to
a reduction in friction (Mazzucco and Spector, 2004).
Because of the importance of the joint fluid in determining the tribological behavior of artificial hip
joints “in vivo,” it is important to use an appropriate lubricant whenever carrying out tribotesting of
artificial hip joint designs or materials “in vitro.” While synovial fluid is the biological joint lubricant,
it is not available in sufficient quantities to be used in wear testing or simulation. Studies of a number of
possible lubricant alternatives have shown that bovine serum produces wear rates and wear debris that
are similar to those that occur in vivo (Saikko and Ahlroos, 1999; Brown and Clarke, 2006). As a result,
nearly all testing of hip joint materials is carried out in bovine serum.

61.2.1.3  Wear and Related Failure Modes for Metal-on-Polymer Hip Prostheses
Although total hip prostheses are designed to withstand decades of use, some prostheses do fail and
require replacement. In some cases, there are problems in the days or months following implanta-
tion, and most of these problems are directly associated with the patient’s response to the opera-
tion or to the implant. These short-term problems, most of which are of biological origin, generally
have little to do with tribology or mechanical failure. On the other hand, in the longer term, that
is, at least 6 months postimplantation, the leading causes of implant failure are implant loosening,
instability, or other problems related to wear (Jacobs et al., 2007; Ulrich et al., 2008; Bozic et al.,
2009). The focus here will be on wear-related failure. When failure of an implanted hip prosthesis
occurs, the patient may feel considerable pain or discomfort, and as a result, it is usually necessary
to remove the failed implant and replace it with a new one in a revision surgery. Retrieval surgeries
are an expensive and somewhat risky operation and occur in large numbers. In 2005, there were over
40,000 hip joint revision surgeries in the United States alone, and that number is growing at about
the same rate as primary THA surgeries; as a result, it is expected that there will be over 67,000 hip
revision surgeries annually by the year 2020 (Kurtz et al., 2007). Hip revision surgeries are some-
what risky and quite costly, with the average cost of a hip revision in the United States in 2009 being
61-8 Design for Lubrication and Tribology

about $55,000 (Bozic et al., 2009), so it is important to reduce the rate of failure of primary total hip
replacements.
Wear of the UHMWPE acetabular liners of hip prostheses has been examined both by study of
retrieved implants to see what wear had occurred in vivo and by tribotesting using hip simulators.
Analyses of retrieved hip prostheses have shown that generally little wear or damage had occurred to
the metallic femoral heads, but that a large percentage of the retrieved UHMWPE acetabular liners had
evidence of wear (Sutula et al., 1995; Jasty et al., 1997). The wear modes found in acetabular liners that
had been implanted prior to the late 1990s included delamination, a subsurface-originated wear mode,
as well as burnishing, light abrasion, and other surface-originated wear modes (Sutula et al., 1995; Jasty
et al., 1997). Wear by delamination produces relatively large wear volumes, while the surface-originated
wear modes produce smaller wear particles and smaller wear volumes. In the mid-1990s, it was found
that the primary contributor to delamination wear was the presence of oxidized UHMWPE in the near-
surface region of the UHMWPE cups (Sutula et al., 1995). That oxidation was a by-product of the dam-
age that had occurred to the polymer during sterilization using gamma irradiation; the gamma rays
broke chemical bonds in the polyethylene molecules, producing free radicals that eventually combined
with diffused oxygen to result in oxidized polyethylene. The oxidized polyethylene has lower strength
and much lower ductility than unirradiated UHMWPE (Kennedy et al., 2000), and it was within the
layer of the oxidized material that much of the delamination wear originated (Sutula et al., 1995). After
that discovery, most implant manufacturers changed their sterilization method to limit the number of
free radicals in the polyethylene liners. This resulted in a lower incidence of delamination wear (Kurtz
et al., 1999), but it did not eliminate surface-originated wear modes such as burnishing and other mild
sliding wear modes (Collier et al., 2003).
Wear of the polyethylene acetabular liner of hip prostheses results in the production of millions of
small wear particles (Ries et al., 2001). The body tries to deal with these foreign body particles by encap-
sulating them in macrophages, setting in motion a complex cascade of biological signaling that ulti-
mately results in resorption of bone near the particles (Ingham and Fisher, 2005). This process, known
as osteolysis, is a direct result of surface-originated adhesive/abrasive wear of UHMWPE and leads to
loss of bone tissue. It has been found that the most biologically active wear particles, from the point of
view of osteolysis, are those of submicron size (Ingham and Fisher, 2005). Osteolysis can ultimately
result in loosening of the prosthesis, a major cause of THA failure (Min et al., 2005; Ulrich et al., 2008).
In addition to causing loosening of the prosthesis, osteolysis can sometimes result in bone fracture.
In an attempt to limit the amount of wear of the UHMWPE acetabular liners, new methods of pro-
ducing UHMWPE have been developed that result in a highly cross-linked material (Kurtz et al., 1999).
Although a variety of proprietary processing methods are used to produce cross-linked UHMWPE,
they all involve a combination of intensive gamma irradiation in order to break bonds in the linear
polyethylene molecules and heat to cause adjacent molecules to bond to each other (Kurtz et al., 1999).
Many investigators have shown through tribotesting that highly cross-linked UHMWPE has lower wear
rates than non-cross-linked UHMWPE (e.g., Endo et al., 2002), but wear of the cross-linked material
still does occur, and the wear process results in the generation of millions of particles of submicron size
that might be biologically active (Ries, 2001; Collier et al., 2003; Jacobs et al., 2007).
Beyond changing the polymer bearing to reduce wear, the metallic articulating component can be
modified. For instance, a titanium nitride coating has been applied to Ti-6Al-4V to improve the surface
hardness and decrease susceptibility to scratching. Early laboratory tests of this material combination
suggested wear reduction over a CoCrMo on UHMWPE bearing (Pappas et al., 1995). Studies of devices
retrieved at revision surgery suggest that the clinical performance of these devices does not mirror the
laboratory predictions. Raimondi and Pietrabissa (2000) report on four retrieved devices with increased
surface roughness at retrieval and increased polymer wear. Likewise, Harman et al. (1997) documented
in a retrieved device that wear debris can originate from a TiN-coated femoral head.
More recently, implants have been developed using a zirconium-niobium alloy exposed to oxygen in
a thermal diffusion process. These devices, therefore, have an oxidized zirconium layer that possesses a
Natural and Artificial Human Joints 61-9

nanohardness of 12–16 GPa, greater than two times that of CoCrMo (Long et al., 1998). This hardness,
coupled with a tough substrate, is cited as providing resistance to roughening during in vivo service,
thereby providing improved wear performance when coupled with a polyethylene bearing (Spector
et al., 2001). Clinical wear results at short durations show no statistical difference in wear when com-
pared with CoCrMo on polyethylene prostheses (Kadar et al., 2011).

61.2.1.4  Metal-on-Metal, Ceramic-on-Ceramic, and Ceramic-on-Metal Hip Prostheses


In recent years, in an attempt to avoid the problems related to wear of the polyethylene acetabular com-
ponents of total hip replacements, three other material combinations have become popular for THA:
metal-on-metal (MOM), ceramic-on-ceramic (COC), and ceramic-on-metal (COM). In the MOM
designs, both acetabular socket and femoral head are generally made of CoCrMo. For COC designs,
both femoral head and acetabular cup are made of either alumina or zirconia; the femoral head fits on
a metallic (CoCrMo or Ti-6Al-4V) femoral stem. COM designs have a ceramic (Al2O3 or ZrO2) femoral
head in contact with an acetabular socket made of CoCrMo. In general, the hip prostheses with metallic
or ceramic acetabular cups exhibit less volumetric wear than designs with UHMWPE cups (Fisher et al.,
2006). Both COC and COM designs have shown less volumetric wear than MOM in laboratory tests,
with COC and COM showing similar wear rates in laboratory tests (Williams et al., 2007). There are
insufficient clinical data to indicate whether all the laboratory conclusions will hold up in vivo, but early
results were promising and made these alternative material combinations good candidates for implan-
tation in young and healthy patients (Fisher et al., 2006). However, some in vivo tribological problems
have shown up for all three material combinations, and these raise a caution flag.
It has been found that even small amounts of wear in metallic sliding components of MOM designs
cause the release of metal ions, particularly Co and Cr, at levels that can result in implant failure
(Dumbleton and Manley, 2005; Mikhael et al., 2009). In some patients, the ions can cause hypersensi-
tivity reactions that appear similar to an infection of the hip; this may result in component loosening
and may require revision of the prosthesis and replacement of the metallic acetabular socket, fre-
quently with one made of UHMWPE (Mikhael et al., 2009). Such metal hypersensitivity reactions are
relatively rare but may affect as many as 5% of the population, particularly for Co and Ni (Waterman
and Schrik, 1985).
The wear rate of MOM devices is influenced by a number of parameters such as surface roughness,
radial clearance, implant position, and material selection, among others (Brockett et al., 2008; Morlock
et al., 2008; Kretzer et al., 2009). This may result in increased serum ion levels of patients and can be
amplified by a deficient renal system. Recently, the condition of arthroprosthetic cobaltism has been
identified in the literature (Tower, 2010). Wear of MOM designs can, in some cases, lead to a release of
cobalt ions which may reach a high enough concentration to influence neurological and cardiac func-
tion. Even in the well-lubricated MOM joint, it should be noted that the tribochemical environment can-
not be discounted. Wimmer and colleagues have noted that mechanical mixing in the top 200 nm of a
metallic bearing surface can create a protective, mixed nanocrystalline/organic material zone (Wimmer
et al., 2010). The durability of that mechanically mixed layer remains to be determined. Tribocorrosion
has also been found to have a dramatic effect on wear rates. Yan et al. report that corrosion-related dam-
age can account for more than 40% of material loss from an MOM (CoCrMo ball on CoCrMo cup) wear
couple (Yan et al., 2008).
Studies of various hip joint material combinations have revealed lower metal ion levels in patients
with COM designs compared with MOM designs (Williams et al., 2007; Isaac et al., 2009). This may be
in part to the reduction in tribocorrosion products, which have been identified as roughly half those in
MOM contacts (Yan et al., 2008). The COM construct is not yet approved for use in the United States
(as of 2010), and clinical results from Europe are not widely published. Further research in this area is
warranted.
COC hip prosthesis bearings have shown low wear rates and low biological reactivity of wear particles
(Fisher et al., 2006). Although some ceramic hip prosthesis components have fractured or chipped due
61-10 Design for Lubrication and Tribology

TABLE 61.2  Material Combinations for THA


Material Combination Relative Cost Clinical
(Femoral Head on 1—Lowest, Experience
Acetabular Cup) 4—Highest (As of 2011) Benefits Potential Risks
MOP 1 Most Long clinical experience, Wear debris can cause osteolysis,
low friction and wear, which may require revision
good biocompatibility,
mechanically forgiving
MOM 2 Moderate Lower friction and wear Metal ions and metallic wear debris
than MOP can cause biological reaction
requiring revision; sensitive to
surgical orientation
Ceramic on 3 Some Lower wear than MOP, Some polyethylene wear debris;
polyethylene (COP) no metallic debris isolated fracture of ceramic head
COM (not FDA 4 Least Lower wear and less ion Some metallic ion debris; some
approved as of 2010) debris than MOM friction-induced squeaking;
possible chipping or fracture of
ceramic
COC 4 Some Lowest friction and Friction-induced squeaking,
wear isolated chipping or fracture of
ceramic

to femoral neck to acetabular liner impingement, most have performed very well in vivo, and failure
rates have been low (Capello et al., 2008). COC hip devices are now considered a good option for young,
active arthroplasty patients for whom the expected implant life cycle presents increased risk of revi-
sion due to wear or wear debris-induced osteolysis with more traditional metal-on-polymer bearings.
Long-term resistance of the ceramic hip components to fracture has yet to be determined. An additional
clinical complication that has been receiving increased attention and documentation over the past few
years is the phenomenon of squeaking in COC hips. The incidence of squeaking in published studies
varies from less than 1% (Capello et al., 2008) to more than 20% (Keurentjes et al., 2008). The squeaking
can be persistent and audible, representing an unacceptable outcome for some patients. Recent studies
have shown that the squeaking is a result of friction-induced vibration that is driven by the rotation
of the hip during walking and is related to the small diametral clearance between femoral head and
acetabular cup (Currier et al., 2010).
Characteristics, including tribological benefits and concerns, for the various material combinations
available for total hip prostheses are summarized in Table 61.2.

61.2.2  Total Knee Prosthesis


Total knee arthroplasty (TKA) is a surgery that helps hundreds of thousands of persons around the
world who are afflicted with arthritis or other diseases of the knee joint. In the year 2005, over 450,000
total knee replacement surgeries were performed in the United States (Kurtz et al., 2007), with a similar
number being performed in the rest of the world. The rate of total knee implantation is growing very
rapidly, at a rate of 8%–10%/year, so it has been estimated that the number of TKA procedures will
increase to about 1.5 million per year by 2020 (Kurtz et al., 2007). TKA is already the most common
orthopedic joint replacement procedure in the United States and will undoubtedly become the most
common such procedure throughout the world in the near future.
In TKA, the contacting bone surfaces, including the articular cartilage, are removed and are replaced
by man-made components. Nearly all total knee prosthesis designs incorporate a polished metal-
lic femoral component, usually made of a CoCrMo alloy, in contact with a tibial bearing made from
UHMWPE. A typical knee prosthesis is shown in Figure 61.3 (Kennedy et al., 2010). In some designs,
Natural and Artificial Human Joints 61-11

Polished metal
(Co-Cr-Mo)

Polyethylene
bearing

Metallic plate
(Ti alloy)

FIGURE 61.3  Typical total knee replacement.

such as that shown in the figure, the polyethylene bearing is fixed in the metallic tibial plate (fixed
­bearing), while in other designs, the bearing is allowed to rotate on the flat surface of the tibial plate
(mobile bearing). In many fixed bearing designs, the tibial plate is made of a titanium alloy, whereas in
other cases, particularly for mobile bearing designs, the tibial plate may be made of CoCrMo. In some
cases, the metallic components are cemented in place in the femur and tibia using PMMA bone cement.
In some other cases, the bone-contacting (nonarticulating) surfaces of the metallic components are
manufactured with a porous metallic surface into which bone can grow to fix the components in place;
the result is a cementless or uncemented prostheses.

61.2.2.1  Kinematics and Loading of Total Knee Prostheses


The natural human knee joint has a very complex design to accommodate large amounts of angular
flexion (over 90°), rotation (up to 10°), and translation (of ± several mm). Artificial knee joints have
been designed to replicate the natural knee joint geometry as closely as possible so as to not inhibit the
motion of the knee. As a result, total knee replacements are characterized by a relatively nonconforming
contact subjected to a combination of rolling and sliding motion, accompanied by some rotation, during
the walk cycle.
In addition to the relatively complex geometry and kinematics of the knee joint, total knee replace-
ments also have to withstand the severe loading conditions faced by the knee. The knee is the most
heavily stressed joint in the human body, with the high contact stresses being a consequence of both
high contact forces and the nonconformal geometry of the knee joint. The contact forces on the knee
are usually slightly larger than those on the hip, varying from 2× to 4× body weight during normal
walking, up to double those values when walking downhill or descending stairs, and >10× body weight
during strenuous activities such as running downhill or lifting weights (Kuster et al., 1997). These high
contact forces, when acting on the nonconforming contact geometry of the knee prosthesis, result in
contact pressures that exceed 20 MPa during the heel-strike and toe-off points of the walking cycle
and can exceed 30 MPa during strenuous activities. For comparison, it should be noted that the yield
strength of typical UHMWPE used in knee bearings is about 19–24 MPa (Table 61.1). To complicate
matters, the knee joint is repetitively loaded millions of times per year, so long-term performance of
the joint replacements requires that they withstand millions of repetitions of high contact pressure in a
nonconforming contact.

61.2.2.2  Lubrication and Tribotesting of Total Knee Prostheses


The lubrication condition in a total knee prosthesis can have an important influence on wear of the
prosthesis. As described earlier, prior to knee replacement surgery, the human knee joint is lubricated
by synovial fluid. Soon after a knee prosthesis is implanted, the synovial membrane resumes production
of synovial fluid, so the prosthesis is similarly lubricated. The in vivo tribological performance of the
prosthetic bearings is therefore determined by the material’s lubricated wear resistance and not by dry
sliding behavior. Owing to the relative scarcity of human synovial fluid, nearly all laboratory testing of
61-12 Design for Lubrication and Tribology

artificial knee joints and prosthetic joint materials is done using bovine serum, which has been found
to produce wear results that are comparable to that occurring in prosthetic knee joints in vivo (Saikko,
2003). The articular surface of the polyethylene tibial bearing is subjected to oscillatory rolling/sliding
motion during joint motion at a rate of 1 to several Hz. Much tribotesting of total knee prostheses is
done in knee simulators that attempt to duplicate the kinematics of the knee and the load cycle that is
applied to it during walking (e.g., Young et al., 2000). Such knee simulator tests are done with bovine
serum as the joint lubricant. It has been found that most of the sliding motion in locations of high tribo-
logical intensity in knee prostheses is very close to unidirectional reciprocation, with little cross-motion
(Hamilton et al., 2005). In addition, the back surface of a tibial bearing for mobile knee designs sees only
oscillatory rotation, with no crossing motion. As a result, preliminary wear testing of materials for knee
prosthesis applications is generally done in either an oscillatory rolling/sliding tester (Van Citters et al.,
2007) or in an oscillatory pin-on-disk tester (e.g., Atwood et al., 2008).
Elastohydrodynamic analysis of the oscillatory lubricated conjunctions between the polymer bearing
and a polished metallic femoral component of a total knee prosthesis has shown that the thickness of the
lubricant (synovial fluid) film is so small that the contact is in the mixed lubrication regime throughout
a simulated walking cycle and enters the boundary lubrication regime at the times when the velocity
approaches zero as the motion direction changes (Jin et al., 1998; Mongkolwongrojn et al., 2010). The
film thickness in the conjunction follows the changes in load and velocity during a walking cycle; the
lowest value of film thickness occurs at the point in the walking cycle (at “toe off”) when the highest load
and lowest velocity occur. It has also been found that there is a squeeze film effect at the zero-velocity
portion of the motion that can prevent the film thickness from ever decreasing to zero during a cycle
(Mongkolwongrojn et al., 2010). Despite this, the lubricant film, especially near the change of motion
direction, is so thin that the conjunction is in the boundary lubrication regime for typical TKA bearings
in vivo. For this reason, the boundary lubrication properties of the synovial fluid in the knee are of great
importance.
Studies of the boundary lubrication characteristics of the constituents of synovial fluid have found
that the lipidic components (SAPL) are a major contributor to the effectiveness of the boundary lubri-
cant in natural knee joints (Hills and Crawford, 2003). The lipid layer has also been found to adsorb
on metallic components of knee prostheses, contributing to their relatively low friction and good wear
resistance (Hills and Crawford, 2003). Recent work, however, has found that the lipidic components are
less effective as boundary lubricants with polyethylene implants than with natural cartilage or even
metal (Trunfio-Sfarghiu et al., 2007). Therefore, the lubrication of prosthetic knee joints, even in the
presence of synovial fluid, does not necessarily protect the polyethylene knee bearings from wear ­during
service.

61.2.2.3  Wear and Other Tribological Failure Modes for Knee Prostheses
Total knee prostheses have a finite life, which is often less than 15 or 20 years, and thus a prosthesis may
have to be replaced in the patient’s lifetime. These revision surgeries currently number over 38,000/year
in the United States alone and are expected to increase to about 120,000 annually by 2020 (Kurtz et al.,
2007). Revision surgeries can be more complicated than the original surgery, with a lower rate of suc-
cess, and are frequently more costly. In fact, the total cost of replacing failed knee prostheses in the
United States amounted to over one billion dollars per annum in 2005 (Bozic et al., 2005). The rate of
failure of TKA is unacceptable, particularly because more knee prostheses are being implanted in even
younger patients and because the overall population is living to an older age. The primary reasons for
the replacement (or revision) of the knee prosthesis are pain in the joint and/or looseness of the prosthe-
sis. The root cause of failure in most cases is related to wear of the polyethylene bearings or complica-
tions that result from wear debris (Sharkey et al., 2002; Kurtz, 2009). Tribological failures such as these
necessitate thousands of expensive retrieval surgeries each year.
Analysis of worn knee components after retrieval has shown that wear on the articulating polyethyl-
ene surface can occur by subsurface-originated delamination or contact fatigue or by surface-originated
Natural and Artificial Human Joints 61-13

removal of small wear particles (Blunn et al., 1997; Williams et al., 1998; Sharkey et al., 2002; Kennedy
et al., 2003; Kurtz, 2009).
Analysis of subsurface-originated failures has concluded that much of the subsurface-originated dam-
age is caused by the interaction of contact stress with oxidized and weakened polyethylene (Kennedy
et al., 2003). The peak maximum shear stress occurs at about 1 mm beneath the contact surface, and
this is approximately the same location where oxidation of radiation-sterilized polymer reaches its peak
(Kennedy et al., 2000). The oxidation results from the reaction of atmospheric or biologically delivered
oxygen with unstabilized free radicals developed during ionizing radiation of the polymer. Oxidation
results in a drastic reduction in the ductility and a reduction in the tensile strength of the polyethylene,
and both of those factors result in easier fatigue crack initiation and propagation in the subsurface
region where oxidation is prevalent. To counteract the oxidation problem, new processing and steril-
ization methods were developed that have resulted in the production of cross-linked UHMWPE that
is more resistant to oxidation and wear (Kurtz et al., 1999). Despite this improvement in tibial bearing
materials, many UHMWPE knee bearings that were implanted prior to the year 2000 remain in patients
around the world and are continuing to oxidize in vivo; some of those bearings are still failing due to
subsurface-initiated wear or contact fatigue. In addition, surface-originated wear occurs, even for the
cross-linked UHMWPE bearings.
A large number of the small particles originate on the articulating surface through such surface-
originated wear processes as burnishing or mild scratching (Hamelynck and Stiehl, 2002; Sharkey et al.,
2002). As is discussed in Section 61.2.1.3, these wear particles can cause osteolysis, which often results
in loosening of the knee prosthesis. Contact in a typical knee joint is relatively concentrated, and the
contact location moves on the surface of the tibial bearing during a typical walking cycle. As a result,
the wear tends to occur in locations on the bearing where the contact conditions were most severe
(Currier et al., 2005). To counteract this, mobile bearing TKA designs have been created that allow the
UHMWPE bearing to rotate against the metallic tibial backplate in order to lower the contact stress
(Hamelynck and Stiehl, 2002). One outcome of the mobile bearing design is that wear can then occur
on the back (nonarticulating) surface as well as on the articulating surface. Recent studies of retrieved
mobile knee prostheses have shown that considerable wear (over 100 mm3/year) of the back surface of
mobile bearings can occur in vivo at short durations (Atwood et al., 2008); the amount of backside wear
and the size of the wear particles can be sufficient to cause osteolysis (Huang et al., 2002). At longer in
vivo durations (e.g., >2 years), wear rates typically decrease to below 54 mm3/year, less than fixed bear-
ing designs implanted for comparable duration (Atwood et al., 2008). Wear, wear particle formation,
and damaged appearance at the back surface have been found to be primarily due to abrasion by third
bodies (Atwood et al., 2006, 2008). Tribological testing has shown that PMMA bone cement particles
produce worn surfaces most like the damaged surfaces seen on many retrieved tibial bearings (Atwood
et al., 2006). Bone cement particles can be generated during and soon after implantation of the prosthe-
sis, and they can work their way into the sliding contact between polymer bearing and metal backplate.
Fixed bearing designs are supposed to have no relative sliding motion between the back surface of
the modular polyethylene tibial bearings and the metallic tibial plates into which they are snapped
into “fixed” position; therefore, it was expected that there would be no wear of the back surface of the
polyethylene. Analysis of retrieved bearings, however, showed that there was measurable wear of the
backsides of many of the bearings that had served in vivo (Conditt et al., 2004). It was also found that
there was unexpected relative motion of up to ±0.5 mm at the backside interface, leading to fretting-type
wear (Engh et al., 2001). Laboratory simulation of the backside wear process showed that the total wear
volume at the backside of a non-cross-linked UHMWPE bearing during low-amplitude fretting-type
motion against a rough Ti-6Al-4V tray could approach 100 mm3/year (Van Citters et al., 2009). If a
polished CoCrMo tray was used instead of a rough titanium tray, the wear volume would drop by about
70%. Additional wear reduction can be realized by the selection of cross-linked UHMWPE bearing
(Van Citters et al., 2009). The particles produced by backside wear could be sufficient in number to be
responsible for some of the osteolysis and implant loosening that occurs in vivo.
61-14 Design for Lubrication and Tribology

61.2.3  Shoulder, Ankle, and Other Artificial Joints


Total knee, total hip, and partial hip arthroplasty procedures account for the vast majority of arthro-
plasty surgeries in the United States, approaching 86% of all joint replacements in 2004 (Defrances and
Podgornik, 2006). Revisions of these devices account for an additional 9% of procedures. The remaining
5% of procedures are performed on the remainder of the joints in the body. Specifically, joint replace-
ment can be performed on the shoulder, ankle, spine, and elbow, as well as in the joints of the hand and
foot. While the majority of these procedures are performed as inpatient procedures and can therefore be
tracked using national surveys such as the National Hospital Discharge Survey, outpatient procedures
performed on smaller joints are poorly documented (Jacobs, 2008).

61.2.3.1  Artificial Shoulder Joints


Total shoulder arthroplasty, like other joint arthroplasty procedures, is performed primarily on patients
aged 60 years and older, often to ameliorate the pain and restricted range of motion associated with
arthritis. The biomechanics of the human shoulder are different from the knee and hip in that the joint
typically experiences up to 150° of abduction and 180° of flexion (Rockwood, 2009). Loads can reach
2000 N during daily activities (Anglin et al., 2000), but are typically modeled at the lower load of 600 N
(Terrier et al., 2007).
Accommodating large loads and a large range of motion requires a joint that can be likened to a
ball and socket joint with decreased conformity between the humeral head and the glenoid compo-
nent. The difference in diameters between the humeral head and glenoid bearing differs between and
within designs, but can approach 10 mm. The increased range of motion associated with a difference
in diameters has been associated with improved outcomes (Walch et al., 2002). This is likely due to the
reduction in shear loads leading to a reduction of bone stock at the nonarticular surface of the glenoid
component. The low conformity of the joint construct will also necessarily lead to high contact stresses.
A typical shoulder arthroplasty uses materials similar to conventional artificial knees and hips, wherein
a CoCrMo humeral head is articulated against a UHMWPE glenoid component. It has therefore been
hypothesized that decreased conformity between the articulating components of the shoulder will lead
to increased wear of the polymer bearing (Walch et al., 2002). However, this has not been documented
in the literature.
Artificial shoulders experience a higher complication rate than knees or hips, estimated at over 14%
in mid- to long-term studies (>2 year follow-up) (Bohsali et al., 2006). The majority of failures are associ-
ated with loosening of the components, with the artificial glenoid loosening at approximately five times
the rate as the humeral component (Bohsali et al., 2006). As with total knees and total hips implanted
before 1997, total shoulder devices were typically sterilized with gamma irradiation and stored in an
air-permeable package. High contact stresses and large magnitude motions led to gross mechanical
failure through cracking, delamination, and fatigue, as well as wear-related biological failure through
osteolysis (Scarlat and Matsen, 2001).
The glenoid can be metal backed, but is often cemented in place as a monoblock polymer device with
stabilization pegs on the nonarticular side. Metal-backed devices have been associated with higher rates
of mechanical and biological failure, including polyethylene wear, fracture of the metal component, and
loosening of the glenoid ingrowth surface (Boileau et al., 2002).
Devices have been introduced that transfer the polymer articulating surface to the humerus in an
effort to better accommodate the geometrical constraints associated with a small joint space and limited
bone stock for fixation in the glenoid. These “reverse shoulders” have been associated with increased
wear (Terrier et al., 2009) and have potentially high complication rates (Bohsali et al., 2006).

61.2.3.2  Artificial Spine Constructs


Interest in artificial disks for the spine has been steady for several decades. Artificial disks have been
available in Europe since the mid-1980s, but there is a shortage of literature on long-term clinical
Natural and Artificial Human Joints 61-15

outcomes. It is difficult to design for spinal applications due in part to the proximity to the neural tissue,
the limited bone stock available, and the motions required of the construct. Designs must accommodate
9° of flexion/extension, 4° of bending, and 4° of axial rotation, all while sustaining loads up to 2 kN [ISO
18192-1]. These prescribed testing parameters indicate a multidirectional motion on the intervertebral
construct, now well known to result in an increased wear rate over unidirectional motion. Grupp et al.
(2009) showed in a spine simulator test that wear rates were 20-fold higher in multidirectional simulator
motion than in unidirectional motion.
The majority of devices identified in the literature include a UHMWPE on metal articulation, similar
to materials used in artificial knees and hips. In vitro wear rates for these spine constructs range in the
literature from 2 mg/million cycles (Grupp et al., 2009) to almost 20 mg/million cycles (Nechtow et al.,
2006). The wear particles have been identified as submicron in size, and the potential for osteolysis has
been cited (Grupp et al., 2009). Adhesive-abrasive wear mechanisms have also recently been reported in
a retrieval study by Kurtz et al. (2007). Oxidation-induced mechanical failure of the retrieved devices
was also present (Kurtz et al., 2007).
Because of the sensitive area in which the devices are being implanted, researchers are further cau-
tioned to consider the effect of wear debris on surrounding tissue. Cunningham (2004) tested poly-
meric and metallic wear debris in an epidural application in rabbits. The results showed local reaction
to the debris and evidence of a more systemic reaction in the cerebral spinal fluid (Cunningham, 2004).
Further analyses of patient outcomes and retrieved tissue are thus warranted.
The in vitro and in vivo wear performance of UHMWPE bearings has spurred new research in alter-
native polymeric bearings for disk arthroplasty. For instance, polyaryletheretherketone (PEEK) has
been developed with and without carbon fiber reinforcement for cervical disk replacement (Grupp et al.,
2010; Langohr et al., 2011). Devices employing a PEEK on PEEK bearing couple show equivalent wear to
UHMWPE on metal articulations, and when the PEEK bearing surfaces are reinforced, the wear rates
are substantially lower (Grupp et al., 2010; Scholes and Unsworth, 2010; Langohr et al., 2011). Langohr
et al. (2011) report adhesive and abrasive wear after laboratory simulation, suggesting that PEEK wear
particles might be involved in third-body abrasion. Under appropriate micromotion conditions, this
could lead to a fretting-type wear, but similar results have not been reported elsewhere in the literature.
In vivo tests of carbon-fiber-reinforced PEEK bearings are underway given the promise that these
materials show for artificial joints.

61.2.3.3  Artificial Ankle Joints


Degenerative ankle disease is typically treated through arthrodesis, or fusion, of the affected joints.
While total ankle arthroplasty has existed since the 1970s, the procedure has not enjoyed the same
popularity as total hip arthroplasty and TKA. Early designs neglected to account for the high axial
and shear forces encountered during gait (>4× body weight and 80% body weight, respectively), result-
ing in high failure rates (Jackson and Singh, 2003). Modern designs, however, typically consist of a
three-component system, including two metallic components separated by a polymer bearing. In some
designs, the bearing is free to move between both of the metallic components, allowing translation to
occur on one articular surface, and flexion to occur at the other surface (e.g., Small Bone Innovations
STAR, Endotec Buechel-Pappas). In other designs, the motions are all accommodated by a single articu-
lar surface, with the polymer component locked into one of the two metallic components (e.g., DePuy
Agility, Tornier Salto Talaris, Wright INBONE).
Patient interest in total ankle arthroplasty is increasing, due in part to reports of improved patient
range of motion relative to fusion (Saltzman et al., 2009), but there is continued concern over outcomes
for these joints. Joint registry studies in Finland, Norway, and Sweden show 5 year survival rates con-
sistently below 90%, with loosening or instability cited as the primary reasons for revision (Skytta et al.,
2010). Few controlled studies exist that identify wear-related failure of total ankle surgeries. Vaupel et al.
(2009) examined eight devices and noted pitting, scratching, or abrasion in all eight, with osteolysis
identified in the majority.
61-16 Design for Lubrication and Tribology

The limited understanding of wear in the ankle has inspired some researchers to create joint simula-
tors to predict tribological outcomes of the joint. Affatato et al. (2007) modified a knee wear simulator
to test total ankle replacements. Their results show that simulator data for the ankle can be compared to
retrieved devices (Affatato et al., 2009), but it is clear that there exist numerous opportunities to perform
further research and modeling of this particular tribological system.

References
Affatato, S., A. Leardini et al. 2007. Meniscal wear at a three-component total ankle prosthesis by a knee
joint simulator. J Biomech 40(8): 1871–1876.
Affatato, S., P. Taddei et al. 2009. Wear behaviour in total ankle replacement: A comparison between an in
vitro simulation and retrieved prostheses. Clin Biomech 24(8): 661–669.
Anglin, C., U.P. Wyss et al. 2000. Mechanical testing of shoulder prostheses and recommendations for
glenoid design. J Shoulder Elbow Surg 9(4): 323–331.
Ateshian, G.A., V.C. Mow. 2005. Friction, lubrication, and wear of articular cartilage and diarthrodial
joints, in Basic Orthopaedic Biomechanics and Mechano-Biology, 3rd edn., V.C. Mow and R. Huiskes,
eds., pp. 447–494, Lippincott Williams & Wilkins, Philadelphia, PA.
Ateshian, G.A., H. Wang, W.M. Lai. 1998. The role of interstitial fluid pressurization and surface porosities
on the boundary friction of articular cartilage. ASME J Tribol 120: 241–251.
Atwood, S.A., J.H. Currier et al. 2008. Clinical wear measurement on low contact stress rotating platform
knee bearings. J Arthroplasty 23: 431–440.
Atwood, S.A., F.E. Kennedy et al. 2006. In-vitro study of backside wear mechanism on mobile knee bearing
components. ASME J Tribol 128: 275–281.
Blunn, G.W., A.B. Joshi et al. 1997. Wear in retrieved condylar knee arthroplasties. J Arthroplasty 12:
281–290.
Bohsali, K.I., M.A. Wirth et al. 2006. Complications of total shoulder arthroplasty. J Bone Jt Surg Am
88(10): 2279–2292.
Boileau, P., C. Avidor et al. 2002. Cemented polyethylene versus uncemented metal-backed glenoid com-
ponents in total shoulder arthroplasty: A prospective, double-blind, randomized study. J Shoulder
Elbow Surg 11(4): 351–359.
Bozic, K.J., P. Katz et al. 2005. Hospital resource utilization for primary and revision total hip arthroplasty.
J Bone Jt Surg Am 87: 570–576.
Bozic, K.J., S.M. Kurtz et al. 2009. The epidemiology of revision total hip arthroplasty in the United States.
J Bone Jt Surg Am 91: 128–133.
Bragdon, C.R., D.O. O’Connor et al. 1996. The importance of multidirectional motion on the wear of
polyethylene. Proc IMechE J Eng Med 210: 157–166.
Brockett, C.L., P. Harper et al. 2008. The influence of clearance on friction, lubrication and squeaking in
large diameter metal-on-metal hip replacements. J Mater Sci Mater Med 19(4): 1575–1579.
Brown, S.S., I.C. Clarke. 2006. A review of lubrication conditions for wear simulation in artificial hip
replacements. Tribol Trans 49: 72–78.
Burroughs, B.R., T.A. Blanchet. 2001. Factors affecting the wear of irradiated UHMWPE. Tribol Trans 45:
215–223.
Capello, W.N., J.A. D’Antonio et al. 2008. Ceramic- on-ceramic total hip arthroplasty: Update. J Arthroplasty
23(2008): 39–43.
Collier, J.P., B.H. Currier et al. 2003. Comparison of cross-linked polyethylene materials for orthopedic
applications. Clin Orthop Relat Res 414: 289–304.
Conditt, M.A., S.K. Ismaily et al. 2004. Backside wear of modular ultra-high molecular weight polyethyl-
ene tibial inserts. J Bone Jt Surg Am 86-A(5): 1031–1037.
Creamer, P., M.C. Hochberg. 1997. Osteoarthritis. Lancet 350: 503–509.
Natural and Artificial Human Joints 61-17

Cunningham, B.W. 2004. Basic scientific considerations in total disc arthroplasty. Spine J 4(Suppl. 6):
219S–230S.
Currier, B.H., J.H. Currier et al., 2007. In vivo oxidation of gamma-barrier sterilized ultra-high molecular
weight polyethylene bearings. J. Arthroplasty 22: 721–731.
Currier, J.H., D.E. Anderson, D.W. Van Citters. 2010. A proposed mechanism for squeaking of ceramic-on-
ceramic hips. Wear 269: 782–789.
Currier, J.H., M.A. Bill, M.B. Mayor. 2005. Analysis of wear asymmetry in a series of 94 retrieved polyeth-
ylene tibial bearings. J Biomech 38: 367–375.
Defrances, C.J., M.N. Podgornik. 2006. National Hospital Discharge Survey. National Center for Health
Statistics, Hyattsville, MD.
Dowson, D. and Z.M. Jin. 1986. Micro-elastohydrodynamic lubrication of synovial joints. Proc IMechE
J Eng Med 15: 63–65.
Dumbleton, J.H. 1981. Tribology of Natural and Artificial Joints, Elsevier, London, U.K.
Dumbleton, J.H., M.T. Manley. 2005. Metal-on-metal total hip replacement: What does the literature say?
J Arthroplasty 20: 174–178.
Endo, M., J.L. Tipper et al. 2002. Comparison of wear, wear debris and functional biological activity of
moderately cross-linked and non-crosslinked polyethylenes in hip prostheses. IMechE J Eng Med
216H: 111–122.
Engh, G.A., S. Lounici et al. 2001. In vivo deterioration of tibial baseplate locking mechanisms in contem-
porary modular total knee components. J Bone Jt Surg Am 83: 1660–1665.
Fisher, J., Z.M. Jin et al. 2006. Tribology of alternative bearings. Clin Orthop Relat Res 453: 25–34.
Grupp, T.M., H.J. Meisel et al. 2010. Alternative bearing materials for intervertebral disc arthroplasty.
Biomaterials 31(3): 523–531.
Grupp, T.M., J.J. Yue et al. 2009. Biotribological evaluation of artificial disc arthroplasty devices: Influence
of loading and kinematic patterns during in vitro wear simulation. Eur Spine J 18(1): 98–108.
Hamelynck, K.J., J.B. Stiehl. 2002. LCS Mobile Bearing Knee Arthroplasty, Springer, Berlin, Germany.
Hamilton, M.A., M.C. Sucec et al. 2005. Quantifying multidirectional sliding motions in total knee replace-
ments. ASME J Tribol 127: 280–286.
Harman, M.K., S.A. Banks, W.A. Hodge. 1997. Wear analysis of a retrieved hip implant with titanium
nitride coating. J Arthroplasty 12: 938–945.
Higginson, G.R., T. Unsworth. 1981. The lubrication of natural joints, in Tribology of Natural and Artificial
Joints, J.H. Dumbleton, ed., pp. 47–73, Elsevier, Amsterdam, the Netherlands.
Hills, B.A. 1989. Oligolamellar lubrication of joints by surface-active phospholipid. J Rheumatol 16:
82–91.
Hills, B.A., R.W. Crawford. 2003. Normal and prosthetic synovial joints are lubricated by surface-active
phospholipid, a hypothesis. J Arthroplasty 18: 499–505.
Huang, C-H., H-M. Ma et al. 2002. Osteolysis in failed total knee arthroplasty: A comparison of mobile-
bearing and fixed-bearing knees. J Bone Jt Surg Am 84: 2224–2229.
Ingham, E., J. Fisher. 2005. The role of macrophages in osteolysis of total joint replacement. Biomaterials
26: 1271–1286.
Isaac, G.H., C. Brockett et al. 2009. Ceramic-on-metal bearings in total hip replacement: Whole blood
metal ion levels and analysis of retrieved components. J Bone Jt Surg Br 91(9): 1134–1141.
Jackson, M.P., D. Singh. 2003. Total ankle replacement. Current Orthopaedics 17(4): 292–298.
Jacobs, J.J., ed. 2008. The Burden of Musculoskeletal Diseases in the United States, American Academy of
Orthopaedic Surgeons, Rosemont, IL.
Jacobs, C.A., C.P. Christensen et al. 2007. Clinical performance of highly cross-linked polyethylenes in
total hip arthroplasty. J Bone Jt Surg Am 89: 2779–2786.
Jalali-Vahid, D., M. Jagatia, Z.M. Jin, D. Dowson. 2001. Prediction of lubricating film thickness in UHMWPE
hip joint replacement. J Biomech 34: 261–266.
61-18 Design for Lubrication and Tribology

Jasty, M., D.D. Goetz et al. 1997. Wear of polyethylene acetabular components in total hip arthroplasty. An
analysis of one hundred and twenty-eight components retrieved at autopsy or revision operations.
J Bone Jt Surg Am 79: 349–358.
Jin, Z.M., D. Dowson et al. 1998. Prediction of transient lubricating film thickness in knee prostheses with
compliant layers. Proc Inst Mech Eng 212: 157–164.
Kadar, T., G. Hallan et al. 2011. Wear and migration of highly cross-linked and conventional cemented
polyethylene cups with cobalt chrome or Oxinium femoral heads: A randomized radiostereometric
study of 150 patients. J Orthop Res 29: 1222–1229.
Kennedy, F.E., B.H. Currier et al. 2003. Oxidation of ultra-high molecular weight polyethylene and its
influence on contact fatigue and pitting of knee bearings. Tribol Trans 46: 111–118.
Kennedy, F.E., J.H. Currier et al. 2000. Contact fatigue failure of ultra-high molecular weight polyethylene
bearing components of knee prostheses. ASME J Tribol 122: 332–339.
Kennedy, F.E., D.W. Van Citters, J.P. Collier. 2010. Tribological characteristics of polyethylene bearings of
knee prostheses. Int J Surf. Sci Eng 4: 166–174.
Keurentjes, J.C., R.M. Kuipers et al. 2008. High incidence of squeaking in THAs with alumina ceramic-on-
ceramic bearings. Clin Orthop Relat Res 466: 1438–1443.
Kretzer, J.P., J.A. Kleinhans et al. 2009. A meta-analysis of design- and manufacturing-related parame-
ters influencing the wear behavior of metal-on-metal hip joint replacements. J Orthop Res 27(11):
1473–1480.
Kurtz, S.M. 2009. The UHMWPE Handbook, 2nd edn., Elsevier, London, U.K.
Kurtz, S.M., O.K. Muratoglu et al. 1999. Advances in the processing, sterilization, and crosslink-
ing of ultra-high molecular weight polyethylene for total joint arthroplasty. Biomaterials 20:
1659–1688.
Kurtz, S.M., K. Ong et al. 2007. Projections of primary and revision hip and knee arthroplasty in the
United States from 2005 to 2030. J Bone Jt Surg Am 89: 780–785.
Kurtz, S.M., A. van Ooij et al. 2007. Polyethylene wear and rim fracture in total disc arthroplasty. Spine J
7(1): 12–21.
Kuster, M.S., G.A. Wood et al. 1997. Joint load considerations in total knee replacement. J Bone Jt Surg
79-B: 109–113.
Langohr, G.D.G., H.A. Gawel, J.B. Medley. 2011. Wear performance of all-polymer PEEK articulations for
a cervical total level arthroplasty system. Proc IMechE Part J: J Eng Tribol 225: 449–513.
Long, M., L. Riester, G. Hunter. 1998. Nano-hardness measurements of oxidized Zr-2.45Nb and various
orthopaedic materials. Trans Soc Biomater 21: 528.
Mayor, M.B., J.P. Collier. 1994. The technology of hip replacement. Sci Med 1: 58–67.
Mazzucco, D., M. Spector. 2004. The role of joint fluid in the tribology of total joint arthroplasty. Clin
Orthop Relat Res 429: 17–32.
McCutcheon, C.W. 1962. The frictional properties of animal joints. Wear 5: 1–17.
Mikhael, M.M., S.D. Hanssen, R.J. Sierra. 2009. Failure of metal-on-metal total hip arthroplasty mimicking
hip infection. A report of two cases. J Bone Jt Surg Am 91: 443–446.
Min, B-W., K-S. Song et al. 2005. Polyethylene liner failure in second-generation Harris-Galante acetabular
components. J Arthroplasty 20: 717–722.
Mongkolwongrojn, M., K. Wongseedakaew, F.E. Kennedy. 2010. Transient elastohydrodynamic lubrication
in artificial knee joint with non-Newtonian fluids. Tribol Int 43: 1017–1026.
Morlock, M.M., N. Bishop et al. 2008. Reasons for failure of hip resurfacing implants. A failure analysis
based on 250 revision specimens. Orthopade 37(7): 695–703.
Morrell, K.C., W.A. Hodge et al. 2005. Corroboration of in vivo cartilage pressures with implications for
synovial joint tribology and osteoarthritis causation. Proc Natl Acad Sci 102: 14819–14824.
Mow, V.C., W.Y. Gu, F.H. Chen. 2005. Structure and function of articular cartilage and meniscus, in Basic
Orthopaedic Biomechanics and Mechano-Biology, 3rd edn., V.C. Mow and R. Huiskes, eds., pp. 181–
258, Lippincott Williams & Wilkins, Philadelphia, PA.
Natural and Artificial Human Joints 61-19

Mow, V.C., R. Huiskes. 2005. Basic Orthopaedic Biomechanics and Mechano-Biology, 3rd edn., Lippincott
Williams & Wilkins, Philadelphia, PA.
Murakami, T., Y. Sawae et al. 2007. Micro- and nanoscopic biotribological behaviours in natural synovial
joints and artificial joints. J Eng Tribol 221: 237–245.
Nechtow, W., M. Hintner et al. 2006. Intervertebral disc replacement mechanical performance depends
strongly on input parameters. Paper 118: 52nd Annual Meeting of the Orthopaedic Research Society,
March, Chicago, IL.
Pappas, M.J., G. Makris, F.F. Buechel. 1995. Titanium nitride ceramic film against polyethylene: A 48 mil-
lion cycle wear test. Clin Orthop (317): 64–70.
Paul, J.P. 1976. Approaches to design—Force actions transmitted by joints in human body. Proc R Soc Lond
B 192: 163–172.
Raimondi, M.T., R. Pietrabissa. 2000. The in-vivo wear performance of prosthetic femoral heads with tita-
nium nitride coating. Biomaterials 21: 907–913.
Ratner, B.D., A.S. Hoffman et al. eds., 2004. Biomaterials Science, 2nd edn., Elsevier Science, San Diego, CA,
pp. 535–536.
Ries, M.D., N.K. Scott, S. Jani. 2001. Relationship between gravimetric wear and particle generation in hip
simulators: Conventional compared with cross-linked polyethylene. J Bone Jt Surg 83A: 116–122.
Rockwood, C.A. 2009. The Shoulder. Saunders/Elsevier, Philadelphia, PA.
Saikko, V. 2003. Effect of lubricant protein concentration on the wear of ultrahigh molecular weight poly-
ethylene sliding against a CoCr counterface. ASME J Tribol 125: 638–642.
Saikko, V., T. Ahlroos. 1999. Type of motion and lubricant in wear simulation of polyethylene acetabular
cup. Proc Inst Mech Eng H 123: 301–310.
Saltzman, C.L., R.A. Mann et al. 2009. Prospective controlled trial of STAR total ankle replacement versus
ankle fusion: Initial results. Foot Ankle Int 30(7): 579–596.
Scarlat, M.M., F.A. Matsen. 2001. Observations on retrieved polyethylene glenoid components.
J Arthroplasty 16(6): 795–801.
Scholes, S.C., A. Unsworth. 2010. The wear performance of PEEK-OPTIMA based self-mating couples.
Wear 268(3–4): 380–387.
Schurz, J., V. Ribitsch. 1987. Rheology of synovial fluid. Biorheology 24(4): 385–399.
Sharkey, P.F., W.J. Hozack et al. 2002. Why are total knee arthroplasties failing today? Clin Orthop Relat Res
404: 7–13.
Skytta, E.T., H. Koivu et al. 2010. Total ankle replacement: A population-based study of 515 cases from the
Finnish Arthroplasty Register. Acta Orthopaedica 81: 114–118.
Spector, M., M.D. Ries, R.B. Bourne, W.S. Sauer, M. Long, G. Hunter. 2001. Wear performance of ultra-high
molecular weight polyethylene on oxidized zirconium total knee femoral components. J Bone Jt Surg
Am 83(Suppl. 2): 80–86.
Sutula, L.C., J.P. Collier et al. 1995. Impact of gamma sterilization on clinical performance of polyethylene
in the hip. Clin Orthop Relat Res 319: 28–40.
Terrier, A., F. Merlini et al. 2009. Comparison of polyethylene wear in anatomical and reversed shoulder
prostheses. J Bone Jt Surg Br 91(7): 977–982.
Terrier, A., A. Reist et al. 2007. Effect of supraspinatus deficiency on humerus translation and glenohu-
meral contact force during abduction. Clin Biomech (Bristol, Avon) 22(6): 645–651.
Tower, S.S. 2010. Arthroprosthetic cobaltism: Neurological and cardiac manifestations in two patients with
metal-on-metal arthroplasty: A case report. J Bone Jt Surg Am 92(17): 2847–2851.
Trunfio-Sfarghiu, A.-M., Y. Berthier et al. 2007. Multiscale analysis of the tribological role of the molecular
assemblies of synovial fluid. Case of a healthy joint and implants. Tribol Int 40: 1500–1515.
Ulrich, S.D., T.M. Seyler et al. 2008. Total hip arthroplasties: What are the reasons for revision? Int Orthop
(SICOT) 32: 597–604.
Unsworth, A., D. Dowson et al. 1975. The frictional behavior of human synovial joints, Part 1. ASME J Lub
Technol 97: 360–376.
61-20 Design for Lubrication and Tribology

Van Citters, D.W., F.E. Kennedy et al. 2009. Backside wear of modular knee bearings; Clinical evidence
and in vitro simulation. Proceedings of the World Tribology Congress, September 6–11, Kyoto, Japan,
Paper D113.
Van Citters, D.W., F.E. Kennedy, J.P. Collier. 2007. Rolling sliding wear of UHMWPE for knee bearing
applications. Wear 263: 1087–1094.
Vaupel, Z., E.A. Baker et al. 2009. Analysis of retrieved agility total ankle arthroplasty systems. Foot Ankle
Int 30(9): 815–823.
Walch, G., T.B. Edwards et al. 2002. The influence of glenohumeral prosthetic mismatch on glenoid
radiolucent lines: Results of a multicenter study. J Bone Jt Surg Am 84-A(12): 2186–21891.
Wang, A., D.C. Sun et al. 1997. Orientation softening in the deformation and wear of ultra-high polyeth-
ylene. Wear 203/204: 230–241.
Waterman, A.H., J.J. Schrik. 1985. Allergy in hip arthroplasty. Contact Dermatitis 13: 294–301.
Williams, I.R., M.B. Mayor, J.P. Collier. 1998. The impact of sterilization method on wear in knee
arthroplasty. Clin Orthop Relat Res 356: 170–180.
Williams, S., A. Schepers et al. 2007. The 2007 Otto Aufranc Award. Ceramic-on-metal hip arthroplasties:
A comparative in vitro and in vivo study. Clin Orthop Relat Res 465: 23–32.
Wimmer, M.A., A. Fischer et al. 2010. Wear mechanisms in metal-on-metal bearings: The importance of
tribochemical reaction layers. J Orthop Res 28(4): 436–443.
Yan, Y., A. Neville et al. 2008. Tribo-corrosion analysis of wear and metal ion release interactions from
metal-on-metal and ceramic-on-metal contacts for the application in artificial hip prostheses.
Proc IMechE J Eng Tribol 222(J3): 483–492.
Young, S.K., T.S. Keller et al. 2000. Wear testing of UHMWPE tibial components: Influence of oxidation.
ASME J Tribol 122: 323–331.
62
Nuclear Reactor Power
Station Lubrication
62.1 Introduction..................................................................................... 62-1
Fission  •  Types of Nuclear Reactors
62.2 Effect of Radiation on Lubricants.................................................62-4
History and Requirements  •  Radiation Effects  •  Lubricants in Use
62.3 Condition Monitoring....................................................................62-9
62.4 Maintenance Scheduling.............................................................. 62-10
62.5 Maintenance Rule.......................................................................... 62-11
62.6 Other Considerations.................................................................... 62-11
62.7 Effect of Radiation on Elastomers............................................... 62-12
Ken J. Brown 62.8 Design Considerations for Equipment....................................... 62-14
Eco Fluid Center Ltd.
Probable Bounds  •  SCC of Fasteners and Components  • ​
Steven Lemberger Eliminating Cobalt  •  Commercial Items for Nukes
Lemberger Consulting 62.9 Summary......................................................................................... 62-15
Services, LLC References................................................................................................... 62-16

62.1  Introduction
Producing electrical power using nuclear reactors is similar to using coal, oil, or natural gas. But rather
than burning fossil fuel, nuclear fission heats the water. Depending on the design, water is circulated
through the top of the reactor or through separate steam generators. The heated water is used to turn
steam turbines coupled to electrical generators that produce electricity. The temperatures and pressures
are much lower than in fossil-fired boilers. In a boiling water reactor (BWR), they are about 285°C
(550°F) and 7.6 MPa (1100 psi), and in a pressurized water reactor (PWR), about 315°C (600°F) and
16 MPa (2250 psi). When using pulverized coal as a fuel, the steam conditions in a supercritical unit
can reach around 600°C (1100°F) and 30 MPa (4360 psi). The nuclear steam is at lower temperatures
and pressures, so the nuclear steam turbines require larger blades to extract the same energy for a given
megawatt rating. For efficiency reasons, the blade tips are close to Mach 1, causing the nuclear turbines
to run at lower revolutions per minute (rpm). These are sometimes called half speed machines which
operate at 1500 or 1800 rpm for 50 or 60 Hz electrical frequency, respectively. This affects the physical
size of the turbines and components like bearings and the configuration of the turbine rotors. With
lower temperatures and wet steam, care has to be taken to prevent erosion and impingement damage in
the steam path. Erosion/corrosion is particularly important for steam piping safety.
While the focus might be on generating the heat energy in the reactors, many supporting systems at
different locations are also required to remove unwanted heat. The main one is the large condenser to
cool the steam back into water after it has gone through the steam turbines, but cooling is also required
for such locations as the area around the reactors, cooling the working spaces and removing heat from
the spent fuel pool. To remove this heat, most stations are located adjacent to rivers or lakes and draw

62-1
62-2 Design for Lubrication and Tribology

in water to be used as service water, condenser cooling water, or treated to be used as boiler water or as
water for other coolers. This requires very large pumps. In locations with shortages of useful water, there
can be very large cooling towers with fans for air cooling. These have little to do with the reactor itself
and can be used on fossil fired stations.
Safety-related and backup equipment can be numerous, and the proper operation of the many tribo-
logical components in pumps, motors, fans, valves, sensors, electrical switches, circuit breakers, etc., is
essential if the equipment is to perform its design function as and when required. If components move,
then friction and wear are involved. Even if components do not move such as brakes, snubbers, sup-
ports, and fasteners, subcomponents can create friction and wear. Vibration from equipment operation
can cause fretting in equipment, structures, and piping.
The effect of radiation on materials must be considered, especially for lubricants and elastomeric seal
materials. Certain materials, for instance those containing heavy metals, halides, or sulfides, must be
avoided due to their possible impact on other materials and/or their radioactive isotopes.

62.1.1  Fission
Nuclear power plants generate the heat to produce steam through fission and the splitting of atoms of
uranium. The uranium fuel is manufactured as small, hard ceramic pellets packaged into long metal
tubes. The tubes are assembled together into fuel assemblies and inserted into the reactor. For enriched
fuel, the pellets consist of two types of uranium, U-238 and U-235. Most of the uranium in nuclear fuel
is U-238. In U-235 atoms, the nucleus, which is composed of protons and neutrons, is unstable and
releases neutrons as the nuclei break up. When the neutrons hit other uranium atoms, those atoms
also split, releasing neutrons of their own, along with heat. These neutrons strike other atoms splitting
them in a fission chain reaction. When that happens, fission becomes self-sustaining. The movement of
control rods inserted among the fuel assemblies regulates the nuclear reaction. The control rods con-
tain elements with a high neutron capture cross section and can include silver, indium, cadmium, and
boron, as well as their alloys and compounds. Liquid “poisons” such as boric acid for PRWs can also be
used as neutron absorbers to control the reaction. A reactor goes “critical” when the fission process is
self-supporting and not “prompt critical” as in a bomb. A nuclear explosion as such is not believed to
be possible but heat, radiation, and the possible formation of hydrogen gas must be taken into account.

62.1.2  Types of Nuclear Reactors


There are two main commercial reactor types: the PWR and the BWR. Both rely on a moderator to slow
down fast neutrons with the differences being the reactor pressure and hence water temperature. The
PWR is at a higher pressure and requires steam generators located away from the reactor but still in
the containment dome. The coolant from the reactor is pumped to the steam generators, often in four
separate loops or circulation systems, and then back to the reactor using reactor feed pumps. The steam
in a separate system on the other side of the steam generators goes to the steam turbines, after which it
is cooled and recirculated to the steam generators.
In a BWR, see Figure 62.1 from Reference 1, the water in the reactor is maintained at about 7.6 MPa
(1100 psi) so that it boils in the core at about 285°C (550°F). Because there is boiling and steam in the
reactor, there is no separate steam generator. In addition, the moisture separators can be in the reactor
vessel so steam from the reactor containment vessel goes directly to the steam turbines. The moisture
separators being on the top is a distinguishing characteristic of BWR’s so the control rods are at the bot-
tom and insert by traveling upward. While the temperature is lower in a BWR, the radiation fields in the
turbine hall are higher. This has some implications for the tribocomponents and maintenance.
With the higher pressure in a PWR primary loop of about 16 MPa (2250 psi), no significant boiling is
allowed; see Figure 62.2 [1]. The hot water is carried out of the reactor to a steam generator, which makes
the steam that goes to the steam turbine. The coolant outlet temperature is 300°C (572°F)–320°C (608°F).
Nuclear Reactor Power Station Lubrication 62-3

Typical boiling-water reactor

Containment
cooling system

4
Steamline

Reactor vessel
Turbine
generator
Separators
and dryers Heater
Feedwater Condenser
3 Condensate
pumps
Core
1, 2 Feed
pumps

Control
rods Demineralizer
Recirculation pumps

Emergency water
supply systems

FIGURE 62.1  BWR (U.S.Nuclear Regulatory Commission, Rockville, MD.)

Typical pressurized-water reactor

Steamline
Containment
cooling system

Steam
4 3 generator

Reactor Control
vessel rods
Turbine
generator

Heater
Condenser

Condensate
pumps
Coolant loop 2
Core

Feed 1
pumps

Demineralizer Reactor Pressurizer


coolant
pumps Emergency water
supply systems

FIGURE 62.2  PWR (U.S.Nuclear Regulatory Commission, Rockville, MD)

A Canadian deuterium uranium (CANDU) reactor is a form of a PWR with the major differences being
the use of horizontal pressure tubes that contain the fuel and the use of a heavy water moderator in which
a deuterium isotope replaces hydrogen in the water molecules (D2O instead of H2O). The pressure tube
arrangement has the advantage that with a fuelling machine the fuel bundles can be moved around and
changed while online.
62-4 Design for Lubrication and Tribology

In addition to coolants of light water and heavy water, other designs include gas-cooled reactors.
At the design stage are fourth-generation sodium-cooled fast reactors known as ASTRID (Advanced
Sodium Technological Reactor for Industrial Demonstration). In current reactors, the uranium fuel may
be enriched with plutonium, and studies are underway to use thorium instead of uranium. Thorium is
more plentiful than uranium and can be used to breed U-233 as nuclear fuel. The U-233 fuel is believed
to be better than U-235 because of better neutron economy and less long life waste. Most commercial
reactors are based on thermal or slower-moving neutrons, but there are other types. For example, the
liquid metal fast breeder reactor (LMFBR) is said to “breed” fuel because it produces fissionable fuel
during operation. Reactors based on the fusion process are also under development with one being
constructed by the International Thermonuclear Experimental Reactor (ITER) consortium in the south
of France.

62.2  Effect of Radiation on Lubricants


62.2.1  History and Requirements
There was significant effort in the 1950s and 1960s to develop radiation resistant lubricants for nuclear
reactors and space exploration. See References 2–16. As the designs progressed, it turned out that for
most of the equipment the radiation levels were not that high. While there are areas that are subject to
very high radiation levels, the components could often be designed to use liquid hydrocarbon–based
lubricants. Many oils could withstand up to 200 Mrads without significant degradation, and in general,
while greases tend to show more degradation, they can withstand similar exposures. For equipment
that has to operate during or after a situation, there can be higher levels of radiation. The exposure var-
ies depending on the reactor type, but it is generally about 200 Mrads. This can be called a design basis
event (DBE) and can include such things as a loss of coolant accident (LOCA). The conditions to which
the equipment will be subjected and for which it must remain fully functional are part of the environ-
mental qualification (EQ) requirements.
In addition, the lubricant has to be in an acceptable condition from the start till the end of any such
event. In the case of many types of grease, radiation first softens them after which they can harden.
Consequently, considerations should not only be given to any “final” hardened state but that the design
of the bearing system and enclosures are appropriate so at its softest the grease cannot run out of the
bearing or the area in which it is required. Then as the grease stiffens, it has to be ensured that this will
not cause excessive torque, lubricant starvation, or excessive operating temperatures. Fortunately, much
of the safety-related equipment only has to run for a relatively short time. Radiation can also have a
significant effect on elastomers used in seals. Some elastomers become too hard and are no longer func-
tional, which can lead to lubricant leaks.
During the EQ, the equipment might be tested with full-scale exposure to steam, heat, and possibly
irradiation. If required, the equipment might also be tested on shaker tables to simulate the ground
movement during an earthquake. See Figure 62.3 from Reference 17, for the test conditions used to
simulate a steam leak for a motor-operated valve (MOV) actuator. In this case, the conditions include
heat, steam, and pressure for set periods of time and the radiation dose or exposure to which the lubri-
cant might be subjected.

62.2.2  Radiation Effects


Radiation can change hydrocarbon molecules by cross-linking (or polymerization) and cleavage. In
many ways the effects of radiation are similar to that resulting from heating and can include discol-
oration and evolution of gases (typically hydrogen), followed by changes in viscosity of an oil and in
consistency of a grease. The consequences are directly related to the amount and type of radiation. The
amount is generally given in rads where one rad is 100 ergs/g.
Nuclear Reactor Power Station Lubrication 62-5

260 256°C /469 kPag


(492°F / 68 PSIG)

232
185°C / 469 kPag 185°C / 469 kPag
(385°F / 68 PSIG) (385°F / 68 PSIG) Cycle test
204 174°C / 779 kPag
174°C / 779 kPag (346°F / 113 PSIG)
(346°F / 113 PSIG)
164°C / 586 kPag
Temperature (°C)

(328°F / 85 PSIG)
177 156°C / 448 kPag
196°C / 469 kPag (312°F / 65 PSIG)
196°C / 469 kPag (385°F / 68 PSIG)
138°C / 248 kPag
149 (385°F / 68 PSIG) (280°F / 36 PSIG)
126°C / 138 kPag
(259°F / 20 PSIG)
121

Start spray
93 AT 12 min

49°C/0 kPag
(120°F/0 PSIG) End spray
66
a b c d e f g

38
0 10 40 2 4 10 12 3 5 6 10 40 2 4 10 12 9 12 24 2 3 4 5 10 20 30
Sec Min Hours Sec Min Hours Days

FIGURE 62.3  MOV EQ conditions.

0.5 MeV gamma

1 MeV gamma ray +

Spur
Delta ray

Photoelectron

Compton electron

FIGURE 62.4  Effect of gamma radiation.

These degradation mechanisms were discussed by Bolt [18] and showed that gamma rays and fast
neutrons deposit energy in organics almost entirely by interaction with electrons. Figure 62.4 illustrates
this interaction with a high-energy (1 MeV) gamma ray. In the initial encounter, an electron is ejected.
The gamma ray gives up about half its energy and is scattered where it can have encounters with other
atoms. The ejected electron can cause ionization in subsequent reactions. The electron is often derived
from an inner orbit of the original atom, leaving a vacancy to be filled from an outer orbit. The energy
lost by the new electron in dropping to the lower orbit is produced in an x-ray and eventually in another
electron (photoelectron).
The original ejected electron (Compton electron) can also cause ejections of secondary electrons from
atoms along its path. These secondaries are called delta rays if they are energetic enough to ionize other
atoms. Ionizations and excitations from all electrons occur in small clusters, called spurs. The spurs are
more dense at the end of the tracks, i.e., at low-energy levels. The low-energy electrons from this process
are eventually absorbed.
62-6 Design for Lubrication and Tribology

R1—CH2—CH2—R2— Hydrocarbon

R1—CH1—CH2—R2 R1—CH2—CH2—R2+ + e-
(positive ion
or excited state)

(R1—CH2—CH2—R2)+ (R1—CH2—CH2—R2)*
(ion molecular reaction
products)

R1—CH2. + R2—CH2.
R1—CH2—CH2—R2 (free radicals)
R1—CH2.

R1—CH2—R1—CH—CH2—R2 R1—CH2—CH2—R2
(free radical propagation) (disproportionation) (radical recombination)

R1—CH=CH—R2 + H2
Olefin formation
Polymer product
by
radical recombination

FIGURE 62.5  By-products of hydrocarbon irradiation.

Fast neutrons affect organics by a similar mechanism; however, they have no charge and do not cause
ionization directly. Instead, they transfer energy by elastic (or billiard ball–like) collisions with nuclei
and form recoil protons. These protons are the charged particles that cause ionization. In hydrocarbons,
the recoil protons are more likely to come from hydrogen atoms, but some can come from carbon nuclei
in the molecule.
The effect of radiation on nonionic compounds may cause formation of ions, radicals, and elec-
tronically excited species which make the compounds more reactive with themselves or with the
environment. The free radicals can further combine or condense. See Figure 62.5 from Pirro and
Wessol [19].
Table 62.1 from the Electric Power Research Institute (EPRI) [20] shows the maximum radiation
exposure that resulted in failure of the base fluid. Each dose measure covers an order of magnitude
plus there can be variations in fluid and the testing. The actual consequences of radiation on a for-
mulated product will depend on the additives used so the table is only a rough guide. For example,
viscosity index (VI) improver additives tend to be significantly degraded so that the viscosity at higher
temperatures will be reduced. This might be problematic if an automatic transmission fluid (ATF) or a
high-VI mineral oil–based hydraulic fluid is used over wide temperature ranges. Pirro and Wessol [19]
also reported that silicone antifoams are destroyed at low radiation doses. In general, a higher aromatic

TABLE 62.1  Base Fluid Maximum Radiation Limits


Base Fluid
Phosphates
Methyl silicones
Siloxanes, silicates diesters
Mineral oil, polyalkenes (PAO)
Polyethers
Poly alkylaromatics (phenyl ethers)
Polyphenlys
Rads 106 107 108 109 1010 1011
Nuclear Reactor Power Station Lubrication 62-7

STO A STO A - base oil STO A used

40

Viscosity (cSt @ 40°)


35

30

25

20
0 0.2 0.4 0.6 0.8 1
Radiation dosage (Mrads)

FIGURE 62.6  Effect of radiation on oil viscosity.

content of the fluid is associated with better radiation resistance. For the power stations, this has been
of concern because more Group II oils are replacing the older Group I steam turbine oils (STOs) com-
monly also used to lubricate pump, motor, and fan bearings.
There is considerable information available on the effect of radiation on power stations lubricants.
The EPRI guide [20] is Revision 4, which is only available to EPRI members. Revision 3 is available for
downloading from their website at www.epri.com. A large volume of work was done on the effect of
gamma radiation for nuclear powered aircraft. Fainman et al. [21] is a very good collection of such work.
EPRI also issues Lube Notes on an annual basis on lubrication issues relating to power stations, and
a number of these pertain to radiation issues [22–27]. The older issues can be downloaded from their
website. For some of the other material specifically applicable to power stations for various applications,
see [28–36]. These can be excellent user-based starting points for material and for other references.
The effect of radiation on a lubricant used in power stations was reported by Kottcamp, Nejak, and
Kem [10]. Figure 62.6 shows an identical effect of radiation exposure on viscosity for both the base
oil and used oil of commercial STOs. In both cases, the viscosity increased by about 50%. The tests
went up to 270 Mrads which would cover many worst-case scenarios. Figure 62.7 from [10] shows that
the Turbine Oil Oxidation Stability Test (TOST) results decreased considerably from 2250 to 220 h.
Similarly, the Rotary Pressure Vessel Oxidation Test (RPVOT) times decreased from >400 to 108 min.
This was for a new fully formulated oil. The foaming tendencies for ASTM D892 Sequences I and III
increased significantly but not as much in Sequence II. The foaming tendency seemed to increase after
90 Mrads. This can be important regarding false indications from oil level detectors, oil being carried
over stand tubes in vertical motors, and load-carrying capabilities during events.
The earlier data were for solvent refined oils and older additive technology. The effect of radiation on
the newer Group II oils and additive technology was addressed in [22–26]. In [22], four new oils were

TOST RPVOT

40 1
0.8
PRVOT (min)

35
Tost (h)

0.6
30
0.4
25
0.2
20 0
0 0.2 0.4 0.6 0.8 1
Radiation dosage (Mrads)

FIGURE 62.7  Effect of radiation on oxidative stability.


62-8 Design for Lubrication and Tribology

TABLE 62.2  Effect of Radiation on Group II STO

Viscosity, cSt % Viscosity


Viscosity Increase at RPVOT Acid
Oil Sample 40°C 100°C Index (VI) 100°C (min) Number
Reference 29.84 5.24 106 394 0.06
100 Mrads 33.08 5.62 108 7 35 0.1
220 Mrads 37.84 6.17 110 18 30 0.05
Turbine oil A 32.79 5.73 116 1015 0.08
100 Mrads 36.04 6.13 117 7 27 0.06
220 Mrads 42.12 6.85 119 20 26 0.04
Turbine oil B 30.41 5.3 106 1428 0.07
100 Mrads 33.03 5.62 108 6 138 0.10
220 Mrads 37.92 5.6 111 17 50 0.04
Turbine oil C 32.12 5.52 108 1128 0.05
220 Mrads 44.19 6.88 112 25 26 0.04

TABLE 62.3  Effect of Radiation on Grease Worked Penetration (60 Strokes)


Fresh Grease Aged 300 h at 150°C (300°F) Aged and Irradiated to 220 Mrads
Calcium complex 334 204 318
Calcium sulfonate 329 341 399

exposed to 100 and 220 Mrads, and some results are shown in Table 62.2. There was a 27% viscosity
increase at 40°C which is less than the data reported by Kottcamp et al. [10]. Plus, it showed a similar
significant decrease in the RPVOT at about the same radiation level. In the earlier data, the viscosity
increase occurred before 90 Mrads, and in the EPRI data, before 100 Mrads. It is essential that the oil
can provide the required performance characteristics during an event; possibly having to change the oil
afterward is less important.
With radiation exposure, greases showed softening followed by hardening. In [27,28], EPRI reported
on irradiation up to 220 Mrad on a previously used calcium complex–thickened grease and on the
replacement calcium sulfonate–thickened grease. The calcium sulfonate thickener technology has some
advantages including good compatibility and inherent extreme pressure (EP) characteristics, as well as
high dropping points, good mechanical stability, and excellent corrosion protection. See Table 62.3 for
the radiation effects on grease penetration.
In summary, up to about 220 Mrads irradiation, oil can be expected to show an increase in
viscosity, a darkening of color, evolution of gases, increase in foaming, and a loss in oxidation
resistance. Oils with VI improvers can also show considerable decreases in viscosity. Greases tend
to show softening but not necessarily affecting the dropping point, indicating that the thickener
system was not compromised.

62.2.3  Lubricants in Use


It is not unusual to have 30–40 different lubricants in a nuclear plant. These are necessary to properly
lubricate the variety of equipment which is classified as safety related or non-safety-related, depending
on their design function. Safety-related equipment is necessary for the safe shutdown of the plant while
non-safety-related equipment performs support functions. Some of the safety-related equipment is envi-
ronmentally qualified, which means if there are any changes to the equipment, including the lubricant,
the equipment needs to be requalified, which can take a long time and lots of paperwork. The variety of
equipment in a nuclear plant includes pumps, fans, blowers, motors, gearboxes, couplings, cranes, hoists,
valve actuators, electrical switches and breakers, control mechanisms, door hinges, etc. In the case of
Nuclear Reactor Power Station Lubrication 62-9

cranes, lubrication is required for the motors, gearboxes, rails, sheaves, brakes, and the wire rope. Also,
for the wire rope in the spent fuel pool cranes, there are sometimes limits on the mineral oil content of
any lubricant because of concerns about contaminating the water in the pool. Restrictions can apply to
the ingredients of other lubricants that might drip or leak into or onto other equipment or systems.
The lubricants can include a number of mineral oil–based fluids, synthetics and silicone fluids.
Synthetic fluids are generally used because of their longer service lives, higher VI, and/or greater resis-
tance to high-temperature degradation. Fire-resistant phosphate ester fluids can be used in steam valve
isolation valves, and silicone fluids are often used in hydraulic snubbers that control piping and com-
ponent movement during an earthquake. The greases can include thickeners based on lithium com-
plex, polyurea, clay, and calcium sulfonates. In addition, there can be greases with more resistance to
centrifugal separation. These can be used in gear and grid flexible shaft couplings, although disc and
diaphragm types that do not require lubrication are also used. Vertical pumps tend to have solid cou-
plings. The specific use for an application is usually mandated by approvals for the EQ. A few sprays are
used for wire rope and chain lubrication, and an antiseize compound is used for fasteners. For the latter,
a nuclear grade is required because of the restrictions on components that might cause stress corrosion
cracking (SCC).
It is important to know where the equipment is located when the lubricant is selected. For example,
some equipment hoists are located outside and exposed to extreme temperatures. For these, the design
bounds could be as high as 48°C (120°F) degrees in the summer and −40°C/°F degrees in the winter. The
lubricant should be suitable for both conditions and everything in between.
PWRs and BWRs usually have several diesel-driven generators for emergency electrical power that
require a suitable engine oil. In the United States, many of the diesels are older types that were designed
for use with high-sulfur fuels. Single viscosity grade oils with higher total base number (TBN) were
used. The fuel is changing to lower sulfur contents, possibly plant sourced and possibly with alcohols, so
the diesel engine oils and systems have to cope. Concerns can include the following: vapor locks, flash
points, the corrosion of fuel tanks, and effects on elastomers used in seals and hoses. However, because
they are part of safety-related systems, the oil cannot just be changed to a new one. In the CANDU
stations, the emergency power generators are driven by gas turbines. The gas producer turbines could
be either aero derivative requiring ester-based lube oils or industrial based using mineral oils or PAOs.
Generally the power turbine and generators used an STO. The requirements for these are not demand-
ing with only test starts on a regular basis. But being idle most of the time has other consequences
including the need to control the water content in the oils and fuels. While the equipment is indoors,
the oil in wall-mounted oil coolers can still be chilled so that low-temperature characteristics, and in
particular the cloud point of the lube oils, can be an issue. Otherwise, cold conditions can cause inter-
mittent problems with oil flow and high differential pressures on filter elements.

62.3  Condition Monitoring


Condition monitoring of lubricated equipment can include a variety of methods, including observation
and inspections, functional testing, stroking with diagnostics for motor operated value (MOV) actua-
tors, vibration testing, thermography, and oil analysis. Grease condition monitoring is not as common,
but assessments are often made on the condition of the grease in MOVs using a combination of appear-
ance, smell, and tactile observations.
The extent and type of condition monitoring to be performed depends on the importance of the
equipment and whether or not it can be accessed during operation. If located in the reactor building,
oil samples or other readings might only be available at 18 month intervals. Sampling during operation
is feasible, but more difficult and exposes a worker to higher radiation dose levels. Since each employee
is carefully monitored for radiation exposure, the accumulated exposure in any one period may limit
future work availability. As a result, one consideration is whether the components, including the oil,
can easily make it to the next reactor outage (RO). It is not desirable to derate or do work in midcycle.
62-10 Design for Lubrication and Tribology

TABLE 62.4  ASTM D-4378 STO Testing


Frequency Test
Daily Appearance
Weekly Color
1–3 months Acid number, cleanliness, and water content
3–6 months Viscosity
6–12 months RPVOT or voltammetry or FTIR
Yearly Rust test

In addition, for safety-related equipment, it is not appropriate to exceed the bounds for the condition of
the oil or component to which it was aged prior to the EQ. In the case of oils and greases, this has been
taken as a 5 year service life.
Another consideration for testing is the residual radiation content of the lubricant samples. While the
levels are generally low, some test procedures require more handling and/or higher temperatures, which
generate more fumes. As a result, some test equipment might have to be located in hot cells limiting
what tests can be readily performed. In general, the testing performed and the limits are in accordance
with those used for other equipment in the power plant. For example, often, a regular STO is used for
lubrication of motor radial and thrust bearings. These have good oxidation resistance, good corrosion
resistance, and few additives that might be adversely affected by radiation. Other than the main boiler
feed pumps, many other pump sets are vertical with the motor on top of the pump. Generally, the pump
axial forces are taken by a thrust bearing in the motor and these can be tilting pad or rolling element
types. The pump radial forces can be taken up by the impeller wear rings and other internal compo-
nents so that these are essentially water lubricated. These vertical pump set arrangements include the
large reactor feed pumps, also called recirculation pumps in BWRs, reactor coolant pumps in PWRs,
and primary heat transport pumps in CANDUs. The motor bearing of reactor coolant pumps typically
uses ISO VG46 STO. The permanent instrumentation for equipment condition monitoring can include
the following: up-thrust bearing temperature, down-thrust bearing temperature, upper radial bearing
temperature, and lower radial bearing temperature, as well as the upper oil cooler water outlet tempera-
ture and lower oil cooler water outlet temperature. While they do not necessarily have remote readings
for oil temperature, oil pressure, or oil level, they could have alarms for oil level and the thrust bearing
hydrostatic oil lift pressure. If the oil level drops below the minimum, the oil level will be restored to the
middle of the level gauge by the equipment operators even in midcycle. The upper oil reservoir in the
case given earlier is 760 L (200 gal). For oil condition monitoring, a sample can be taken at every reactor
refueling outage (1½–2 years). At one BWR station, the pump motor oil is monitored using viscosity,
water content, oxidation stability by RPVOT Method “A” ASTM D-2272, particle counts, trace metals
by ICP, a ½ μm patch test for weight, and if the particle counts are high, then analytical ferrography.
Guidelines for the oil sampling of the main steam and feed pump turbines are given in ASTM D-4378
[37] and the balance of plant equipment in ASTM D-6224 [38]. See Table 62.4 for the D-4378 recommen-
dations. These tests and monitoring frequencies are to be fine tuned for specific equipment and to take
into account historical issues or experiences at other plants. A few stations are experiencing issues that
could be related the greater use of some Group II oils. Foaming and possibly a greater tendency to form
varnish deposits are of concern, but these are not tests currently recommended by ASTM as routine.
However, it is given that the flash point and foam and water separability are tested if contamination is
suspected.

62.4  Maintenance Scheduling


There are many requirements and guidelines for what maintenance is done and when, at nuclear power
stations. Most major work is done when the reactor requires changing some or all of the fuel. Other work
Nuclear Reactor Power Station Lubrication 62-11

TABLE 62.5  Capacity Factors at U.S. Nuclear


Power Plants
Year Capacity Factor (%)
1980 56
1985 58
1990 66
1995 77
2000 88
2005 89
2009 92

is done at different frequencies. An RO is generally about every 18 months and lasts at least 15 days. Before
every RO, Maintenance inspects the condition of the new fuel rods. The target is zero defects. Other work is
done based on multiples of the RO, for example, a feedwater pump steam turbine disassembly every 4 cycles
or 10 years with steam stop and control valve work at the same time. The frequency can be based on input
from the original equipment manufacturers (OEM) but must take in account similar equipment experi-
ence elsewhere. Many larger components have user groups where one can meet to discuss common issues.
For nuclear power to be commercially viable, operating and maintenance costs had to be reduced,
and one aspect was to improve uptime. Also called plant capacity factor with 100% being never shut-
down. A series of institutional changes have facilitated the process: During the 1980s, the Institute of
Nuclear Plant Operators (INPO) was created to share technical information. The new fuel designs lasted
longer which permitted the time between refueling outages to be increased from 12 months to some-
times 2 years. Refueling outages have also been cut from an average of 104 days in 1990 to 41 days in
2009. The capacity factors at U.S. nuclear power plants went from 56% in 1980 to >90%. See Table 62.5
from Reference 39. This required improvements in maintenance and condition monitoring. Higher-
capacity factors means the in-service time is longer and the time to do maintenance is less. For some
equipment, it can also mean that these are used more, while for other equipment like shutdown cooling
pumps and steam valves, it can mean that these are used less. Both have implications that must be con-
sidered to ensure functionality. For example, if servo-valves are used less frequently, then the limits on
fluid condition might have to be tightened because silt cannot be purged.

62.5  Maintenance Rule


In the United States, Maintenance Rule (10CFR50.65) for nuclear stations became effective in 1996.
This rule was established to monitor performance of systems, structures, and components (SSCs) and
to fulfill the requirements of Nuclear Regulatory Commission (NRC) 10CFR50.65. The objectives were
to ensure that important systems are capable of performing their intended functions and that failures
resulting in reactor trips and unplanned safety system actuations are minimized. For power station
operators, compliance with the Maintenance Rule means monitoring to demonstrate the effectiveness
of the preventive and corrective maintenance, predictive maintenance, and routine condition monitor-
ing and trending programs to prevent failures. When there are failures, the Maintenance Rule requires
cause determination and implementation of corrective actions. This justifies more in-depth oil analyses
of failed tribological components. It also encourages the review of experience at other stations. If the
component is safety related, the laboratory used must be properly qualified.

62.6  Other Considerations


The maintenance and inspections must be consistent with Nuclear Electric Insurance Limited (NEIL)
requirements. NEIL insures electric utilities for damage to insured sites, nuclear decontamination
62-12 Design for Lubrication and Tribology

expenses incurred at insured sites, physical loss at such sites, certain premature decommissioning costs,
and the costs associated with certain long-term interruptions of electricity supply.
NEIL provides “should” and “shall” activities pertaining to condition monitoring, inspections, and
maintenance. They provide a credit of 125 points for quarterly analysis of oil samples used to determine
the presence of wear metals or contaminants. For PWRs, NEIL allows a credit of 50 points for vibration
and lubricating oil analysis every fuel cycle for the reactor coolant pumps and motors. For other elec-
tric motors, motor generator sets, and pumps, there is a credit of 30 points for quarterly testing if that
equipment is operating continuously and once a fuel cycle for equipment operating only for surveillance
testing. The NEIL points are important because they are used to determine the cost of the coverage. This
supports the need for oil analysis as part of routine condition monitoring and requires selecting the
most appropriate tests and control limits.

62.7  Effect of Radiation on Elastomers


Most lubrication systems will have elastomeric seals, gaskets, and/or hoses. Bearings may have nylon
retainers or seals in sealed-for-life bearings. Elastomers are reportedly about 10 times more sensitive
to radiation than lubricants. This is shown in Figure 62.8 [20], in that some elastomers show degrada-
tion at only 106 rads. One of the more common elastomers for use with water and steam is ethylene
propylene rubber (EPR) but it is not suitable for use with many mineral oils because it can swell.
Extra diligence is required to use the proper elastomer. When irradiated, many elastomers increase in
hardness and show a decrease in elongation. See Figure 62.9 from Broadway et al. [40], for the change
in elongation of irradiated Buna N. In Broadway’s test, the degradation was more severe in a closed
(sealed ampoule) than exposed to atmosphere. The degradation was even more severe in a vacuum with
a 50% loss in tensile strength after exposure to 65 Mrads. This shows the importance of environmental
considerations. Figure 62.10 from the same reference 40 shows the change in tensile strength in which
it first increases and then decreases. The hardening and loss of elongation can affect the ability to seal
and make seals more prone to cracking and failure. These changes can cause leaks and more lubricant
seepage. Lubricant loss can be a concern regarding the minimum operating levels, as can smoke or fires
as a result of lubricants dripping on hot surfaces. Elastomeric tests in air are not necessarily indicative
of what might happen in service with specific lubricants. Broadway et al. [40] reported that some elas-
tomers showed less degradation when immersed in fluids. Exceptions to the general hardening with
irradiation are reported by King et al. [41] to be butyl and Thiokol (a polymer of ethylene polysulfide)
rubbers that softened and became liquid. Much of the referenced work in the literature on the effects
of radiation on elastomers was done in the 1960s; thus, more recent products are not mentioned. An
example of a newer product is polyether ether ketone that reportedly can have very good radiation
resistance as well as good tribological characteristics [42].
In addition, the polymer degradation by-products must be considered. In particular those elastomers
that contain halogens like the fluorine in fluoroelastomers and the chlorine in neoprene and polychlo-
rinated ethylene elastomers can possibly form halides. King et al. [40] reported that hydrogen fluoride
evolved with elastomers made with copolymers of hexafluoropropylene and vinylidene fluoride and that
hydrogen chloride evolved with PVC when irradiated. This may cause SCC of stainless steels or other
alloys used for components, piping, and/or fasteners. Other restricted ingredients include heavy metals
and sulfur. A wide range of restrictions can apply and do not always agree, so it is prudent to determine
what is required for specific applications. Any material that is suspect is not allowed in the reactor area
or where it might come into contact with the water used in the reactor. Concerns of cross contamination
and/or the wrong product being used at some time in the life of the unit can cause further restrictions
on what can be used and where.
Because there can be considerable variation in the compounds of the same elastomer type and the
testing, data in the literature can vary widely. So it is important to have measures in place to ensure
Nuclear Reactor Power Station Lubrication 62-13

Degree of change

Elastomers Nil Moderate High


Acrylics
Butyls
Fluoroelastomers
Hypalons
Natural rubbers
Neoprenes
Nitriles
Polysulfides
Silicones
Styrenes
Urethanes
Vinylpyridines

Others

Polyethylenes
Polyfluorocarbons
Poly (vinyl chlorides)
Silicone resin-glass
Ceramics
Metals

105 106 107 108 109


Gamma dose (rads)

FIGURE 62.8  Effect of radiation on elastomers.

that the compound being used is appropriate and has been suitably tested. The most prudent approach
is to assume the most conservative result or do verification testing on the actual compound used. For
safety-related equipment, this can be particularly important. It is also necessary to have quality assur-
ance at the stations with the appropriate purchasing and verification procedures to ensure that the
compound ordered was the one received. One concern is substitution for whatever reason regardless

Closed system Open system Vacuum


90.0
80.0
Percent elongation

70.0
60.0
50.0
40.0
30.0
20.0
10.0
0.0
0.0 20.0 40.0 60.0 80.0
Radiation dosage (Mrads)

FIGURE 62.9  Effect of radiation on the elongation of Buna N.


62-14 Design for Lubrication and Tribology

Closed system Open system Vacuum


6.0

Tensile strength (kg)


5.0
4.0
3.0
2.0
1.0
0.0
0.0 10.0 20.0 30.0 40.0 50.0 60.0 70.0
Radiation dosage (Mrads)

FIGURE 62.10  Effect of radiation on the tensile strength of Buna N.

of any reported “equivalence.” Supporting data are warranted. Simple tests can include those for elas-
ticity using rebound type hand held devices, measuring the specific gravity and/or doing a constitu-
ent analysis on a sample from a batch. Thermal gravimetric analysis (TGA) and differential scanning
calorimetry (DSC) can be helpful for polymer characterization. Some o-rings can also be supplied with
color coded dots or color in the whole o-ring. This can be helpful as long as it was approved with those
color agents.
Elastomer selection includes suitability for a specific function but also to what chemicals it might
be exposed to during service and/or maintenance. For example, EPR elastomers have good resistance
to radiation, water, and phosphate ester fire-resistant fluids but not to mineral oils. Consequently, it is
important to consider what is being used from manufacturing to end use including corrosion protecting
compounds that might be used for storage and during maintenance. In the case of mechanical seals, a
lubricant is sometimes used to help slide the o-rings over a shaft or sleeve. If these are made of EPR, then
mineral oil–based products are not suitable. In some cases, silicone-based oil or paste can be used. With
phosphate ester fluids, just use a drop of that fluid.

62.8  Design Considerations for Equipment


62.8.1  Probable Bounds
No significant radioactive releases outside the reactor and auxiliary containment structures are permit-
ted, so a great many factors must be considered. Expected events include storms of the century, aircraft
strikes, and explosions on any nearby railway lines or roads, plus the consequences of “natural” events
such as flooding, tornadoes, and/or earthquakes. For the most part, the design considerations are for the
structures and the building envelope.
These events can have an impact on tribological considerations for pump, motors, fans, turbines,
valves, and the like as can seismic events and steam pipe breaks. In the case of structures and piping, an
analysis can be done, but in the case of smaller and more separate components such as valves, testing
is done on a shaker table. There are also different considerations depending on whether the component
has to operate during or after an event. Both conditions require consideration of the bearings, seals, and
any associated lubrication equipment such as lines, filters, and monitoring equipment. Anything that
penetrates the oil sump and can break off or affect functionality must be considered. As a consequence,
the review process can be very time consuming even for the addition of a simple component such as
a pin-type sample valve or even a new and hopefully better filter. It must be shown that there will not
be any adverse impact on the mechanical integrity of the component nor on the ability to perform its
intended function.
Nuclear Reactor Power Station Lubrication 62-15

62.8.2  SCC of Fasteners and Components


SCC can be of concern especially with the extensive use of austenitic stainless steels and high-strength
alloy steels in nuclear power plants. Insulation materials, dye penetrants, and lubricants can all promote
SCC. Reference 43 shows that antiseize pastes must be chosen with care. The compounds of concern
can depend on whether the application is wet or dry, with MoS2 being restricted in wet environments.
For example, the failure of a number of turbine discs was attributed to the use of MoS2 during assembly.
Similarly, there have been issues with mild steel and low alloy bolts because of boric acid corrosion [44].

62.8.3  Eliminating Cobalt


In other industries, cobalt is widely used in wear-resistant weld overlays and in wear-resistant inserts
in pumps, turbines, valves, and mechanical seal faces. In nuclear reactors, cobalt-containing mate-
rial removed by wear and/or corrosion can circulate through the reactor and become radioactive. This
increases the overall radiation levels making operation, maintenance, and disposal issues more compli-
cated. As a result, cobalt-free materials are used [35].

62.8.4  Commercial Items for Nukes


Nuclear plant operation requires the replacement of parts, as well as complete components. As a result
of the reductions in new nuclear power plant construction, some suppliers had discontinued 10CFR50,
Appendix B quality assurance programs, which had provided guidance on the design, manufacture, and
application of safety-related components. In such cases, operators must procure qualified replacement
items from alternative suppliers or purchase replacement parts as commercial grade and dedicate them
for safety-related use. With the increased use of commercial grade items, a guideline was developed to
ensure their conformance to regulatory requirements. The lubricants can be purchased from the origi-
nal equipment supplier who has sufficient documentation to show that the product is suitable. In such
cases, the supplier is likely to be audited by members of the Nuclear Procurement Issues Committee
(NUPIC) and they also would be expected to audit the lubricant manufacturer and possibly even dis-
tributors. NUPIC will audit companies even if they are International Standards Organization (ISO)
registered because NUPIC has a different standard. The utilities can also purchase commercial grade
lubricants if approved for that application, but they must then do their evaluation and testing required to
show that the lubricant is what was specified and ordered. As an example, in the case of grease for MOVs,
this might include some or all of the following: appearance, color, penetration, FTIR, and dropping
point—the purpose being to verify that the thickener is what was ordered and that there is no mixing.
Some utilities also do elemental analyses. See Reference 36 for guidance from EPRI.

62.9  Summary
For most equipment in nuclear power plants, the operating conditions on a daily basis are not necessar-
ily more severe than for other forms of steam-powered generating plants, and in fact, the steam pres-
sures and temperatures can be much lower. The differences are the dire consequences of radiation leaks,
the cost associated with repairs, and the need to maintain the utmost reliability. This must be for the
design life of 40 years and the life extension of 20 years.
Few products remain the same over a few years let alone 60 years, and in any case no 60 year service
data are available for nuclear reactors, so maintaining reliability requires a continuous effort. In addi-
tion, both the capacity and capacity factors of nuclear stations have increased already. While further
improvements might be more difficult, it is expected that they are possible and will require new and/
or improved tribocomponents and/or procedures. Future reliability will require ensuring that these are
correct.
62-16 Design for Lubrication and Tribology

References
1. http://www.eia.doe.gov/nuclear/, downloaded October 26, 2010.
2. Carroll, J.G. and Calish, S.R., Some effects of gamma radiation on commercial lubricants, Lubr. Eng.,
13(7), 388–392 (1957).
3. Carroll, J.G., Bolt, R.O., and Hotten, B.W., Radiation resistant greases, Lubr. Eng., 13(3), 136–140
(1957).
4. King, J.A. and Rice, W.L.R., The effects of nuclear radiation on lubricants, Lubr. Eng., 13(5), 279–283
(1957).
5. Dynamic tests spark nuclear lube program, Lubr. Eng., 13(7), 380–383 (1957).
6. Fischer, D.J., Zack, J.F. Jr., and Warrick, E.L., Radiation stability of silicone greases, Lubr. Eng., 15(10),
407–409 (1959).
7. Carroll, J.G., Bolt, R.O., Cunniff, C.E., and Heyl, P.T., Field tests on a radiation-resistant grease, Lubr.
Eng., 18(2), 64–70 (1962).
8. Hausman, R.F. and Booser, E.R., Application problems with petroleum lubricants in nuclear power
plants, Lubr. Eng., 1(4), 199–202 (1957).
9. Carroll, J.G. and Bolt, R.O., Lubrication in the presence of nuclear radiation, Lubr. Eng., 12(5),
305–309 (1956).
10. Kottcamp, C.F., Nejak, W.W., and Kem, R.T., The effects of high energy ionizing radiation of turbine
oil performance, Tribol. Trans., 2(1), 7–12 (1959).
11. Calkins, V.P. and Collins, C.G., General radiation damage problems for lubricant and bearing-type
materials, Tribol. Trans., 1(1), 87–90 (1958).
12. Carroll, J.G. and Bolt, R.O., Development of radiation resistant oils, Tribol. Trans., 2(1), 1–6 (1959).
13. Cox, D.B., Oberright, E.A., and Green, R.J., Dynamic and static irradiation of nuclear power plant
lubricants, ASLE Trans., 5(1), 126–133 (1962).
14. McHugh, K.L. and Stark, L.S., Properties of a new class of polyaromatics for use as high-temperature
lubricants and functional fluids, ASLE Trans., 9(1), 13–23 (1966).
15. Hourani, M.J., Hessell, E.T., Abramshe, R.A., and Laing, J., Alkylated naphthalenes as
high-performance synthetic lubricating fluids, Tribol. Trans., 50(1), 82–87 (2007).
16. Zaslavsky, Y.S., Gorbach, V.A., Zhukov, N.A., Frunza, V.I., and Astratova, N.M., On the methods of
dynamic irradiation tests for evaluation of radiation stability of lubricating oils, Tribol. Trans., 16(3),
216–222 (1973).
17. Rotork Control Ltd., Electric actuators for nuclear powerplants, Figure 3.2, Publication E250E, Issue
05/05.
18. Bolt, R.O., Effects of radiation on lubricants, in: Handbook of Lubrication, Vol. 1, Chapter 14, ed.
Booser, E.R., CRC Press, Boca Raton, FL (1985).
19. Pirro, D.M. and Wessol, A.A., Lubrication fundamentals, in: Nuclear Reactors and Power Generation,
Chapter 14, Marcel Dekker, New York (2001).
20. EPRI Lubrication Guide, Revision 4, EPRI Report 1019518, December 2009.
21. Fainman, M.Z., Krasnow, M.E., Kaufman, E.D., Reynold, O.P., Thistlethwaite, R.L., and Wolford, O.C.,
The behavior of fuels and lubricants in dynamic test equipment in the presence of gamma radiation,
WADC Technical Report 58-264, March 1958, 177 pp.
22. EPRI Lube Note Number 2, Radiation Exposure Study on Modern Turbine Oils, October 2009.
23. EPRI Lube Note Number 1, The Effect of Group II Base Oils in Turbine Oils, October 2007.
24. EPRI Lube Note Number 1, Effect of Turbine Oil on Radiation Resistance, December 2010.
25. EPRI Lube Note Number 6, RULER on Irradiated Turbine Oils Shows Depleted Inhibitors, December
2010.
26. EPRI Lube Note Number 2, Lube survival after a LOCA, November 2002.
27. EPRI Lube Note Number 8, Replacing ExxonMobil Nebula EP Grease with Cor-Tek MOV Long
Life—First EQ results, November 2002.
Nuclear Reactor Power Station Lubrication 62-17

28. Comparative Analysis of Nebula and MOV Long Life for Limitorque Main Gearbox Applications,
EPRI Report 1003483, December 2002.
29. Flexible Shaft Coupling Maintenance Guide, EPRI Report 1007910, December 2003.
30. Reactor Coolant Pump/Reactor Recirculation Pump Motor Lubrication Oil System Maintenance
Guide, EPRI Report 1013456, December 2006.
31. Oil Lubrication Guide for Rotating Equipment, EPRI Report 1019517, December 2009.
32. Effective Practical Greases Practices, EPRI Report 1020247, October 2010.
33. Crane Maintenance and Application Guide—Maintenance and Application of Overhead Cranes
(Containment, Fuel Handling, Turbine Building, and Intake Structure), EPRI Report 1000986,
December 2000.
34. Plant Support Engineering: Elastomer Handbook for Nuclear Power Plants, EPRI Report 1014800,
August 2007.
35. Compilation and Evaluation of NOREM™ Test Results: Implications for Valve Applications, EPRI
Report TR-109343, June 1999.
36. Supplemental Guidance for the Application of EPRI Report NP-5652 on the Utilization of
Commercial Grade Items, EPRI Report TR-102260, March 1994.
37. ASTM D-4378, Standard Practice for the In-Service Monitoring of Mineral Turbine Oils for Steam
and Gas Turbines, 1997.
38. ASTM D-6224, Standard Practice for the In-Service Monitoring of Lubricating Oil for Auxiliary
Power Plant Equipment, 1998.
39. Ref: http://www.eia.gov/totalenergy/data/annual/txt/ptb0902.html, March 31, 2011.
40. Broadway, N.J. and Palinchak, S., The Effect of Nuclear Radiation on Elastomers and Plastic
Components and Materials, REIC Report No. 21 (Addendum), Battelle Memorial Institute, August
31, 1964, 443 pp. AD454056.
41. King, R.W., Broadway, N.J., and Palinchak, S., The Effect of Nuclear Radiation on Elastomers and
Plastic Components and Materials, REIC Report No. 21, Battelle Memorial Institute, September 1,
1961, 375 pp., AD267890.
42. Victrex Properties Guide, PROPENT/301, Figure 34.
43. Czajkowski, C.J., Testing of Nuclear Grade Lubricants and Their Effects on A540 B4 and A193
Bolting Materials, BNL-NUREG-35629, National Association of Corrosion Engineers Annual
Meeting, March, Boston, MA, 1985.
44. Chang, T.Y., Regulatory Analysis for the Resolution of Generic Safety Issue 29: Bolting Degradation
or Failure in Nuclear Power Plants, U.S. Nuclear Regulatory Commission, NUREG-1445, September
1991.
63
Space Mechanism
Lubrication
63.1 Mechanism Requirements.............................................................. 63-1
Operational Considerations  •  Lubrication Issues
63.2 Solid versus Liquid Lubrication.....................................................63-3
Liquid Lubricants  •  Solid Lubricants
63.3 Mechanism Components............................................................. 63-11
Bearings  •  Lubricant Film Thickness  •  Cage
Impregnation  •  Lubricant Loss Mechanisms  •  Gears and
Harmonic Drives  •  Slip Rings
Stuart Loewenthal 63.4 Concluding Remarks..................................................................... 63-15
S. Loewenthal & Associates References................................................................................................... 63-15

63.1  Mechanism Requirements


Lubrication requirements for space mechanisms differ in several respects from terrestrial and aviation
applications. The most obvious is the extreme environment of space. Environmental temperatures on
outboard components can range from −100°C to 100°C, and some instruments operate at cryogenic
temperatures. Vacuum levels are lower than 10−6 torr and radiation levels of up to 30°K rad or higher
are not uncommon for exposed hardware. Furthermore, the inability to replace anomalous mechanisms
(except for nonstandard case of on-orbit serviceable units) and the considerable cost of satellites have
put a large premium on the development and application of robust tribological solutions.
Today’s commercial and military satellites are typically designed for mission lives of 12–15 years.
Antenna gimbals, solar array drives, optical scanner instruments, and space reaction wheels have lubri-
cation systems that cannot be replenished, which put a difficult burden on the lubricant. Complicating
matters further is the need to demonstrate life margins of 1.5–2X in a reasonable time period. Validation
life testing is often completed very late in the production cycle so heavy reliance on heritage lubrication
methods is customary.
Weight and size constraints require motors for space actuators to be small and light. Consequently,
torque margins are meager. Torque increases from lubricant degradation are thus a significant con-
cern so a bearing may be considered failed even before significant wear occurs. Also, solar array
panels, thermal control materials, and optical instruments have stringent particulate (debris) and
molecular film (lubricant outgassing) requirements. These not only drive the selection of the lubricant
but usually require the design of mitigation features such as labyrinth seals, lubricant anticreep bar-
rier films, etc.

63-1
63-2 Design for Lubrication and Tribology

63.1.1  Operational Considerations


The lubrication approach needs to be tailored to the specific mechanism application (Figure 63.1).
Applications range from one-time restraint/hold-down mechanisms to solar array drives that track
the sun at 1 rev per 90 min and further to high-speed attitude control devices such as reaction wheels,
momentum wheels, and attitude control gyroscopes. Bearings in these high-speed components typi-
cally rotate at many thousands of rpm often accumulating many billions of revolutions over the mission.
Such components are among the most difficult to ensure near flawless reliability so spacecraft normally
carry at least one spare unit.
Launch restraint and deployment mechanisms are typically designed for a minimum of 50 cycles of
ground test operations even though they may only need to be actuated once on orbit (Figure 63.1). These
types of mechanisms are not particularly sensitive to bearing or pivot smoothness and mainly require
that there is adequate torque margin to ensure mission success. Grease-lubricated preloaded ball bearings,
bonded MoS2-lubricated spherical bearings, and, occasionally, self-lubricated sleeve bearings are typical.
Liquid lubricants are normally used with medium-speed solar array drives used to orient solar panels
to the sun and despin mechanisms used to continuously point antennas to the earth. They generally
contain speed reducers on the order of 100–160 to 1 ratio for torque amplification, which reduces the
size of the motor at the expense of higher speeds and more stress cycles. European solar array drives
generally use ion-plated lead lubrication.
The mechanisms shown in 63.1 have no special requirement for precision pointing so smoothness
of operation is not a driving factor. However, scanner instruments such as filter wheels and pointing
mirrors generally require a high degree of positional accuracy/repeatability. The servo control system
is generally sensitive to bearing torque irregularities sometimes referred to as torque ripple. This means
that lubricant degradation may cause a torque problem. High-quality ball bearings lubricated with
premium quality liquid lubricants are the norm.
Perhaps the most demanding applications are high-speed momentum and reaction wheel bearings
and those for navigational gyroscopes run at speeds on the order of 15,000 rpm and have cycle life
requirements in the billions of revolutions category. For these applications, premium ball bearings

Attitude

Rate
gyro΄s
Attitude
control
10 – 20K rpm
Momentum/ 500 HZ
reaction
Spin- wheels
stabilize
satellites 3 – 10K rpm

De-spin 100 HZ
drives
Leo
50 – 100 rpm
Space station 1 HZ
rotary
joints
Synchronous
orbit 1 Rev/90 mins
Satellite
solar array/ 10–4 HZ
antenna drives
Ascent
1 Rev/day Speed affects life and lubrication approach
Restraint 10–5 HZ Low speed tens to thousands of cycles dry/wet lubes
mechanisms
easier to life test
No revs/day High speed billion of cycles wet lubes hard to life test
0 HZ

FIGURE 63.1  Space mechanism lubrication life requirements.


Space Mechanism Lubrication 63-3

(class ABEC 7 or ABEC 9) and high-quality synthetic hydrocarbon space oils/greases coupled with
stringent processing control with substantial acceptance testing are critical to meeting mission lives of
up to 15 years.

63.1.2  Lubrication Issues


Surveys have indicated that improper lubrication is the primary cause of spacecraft mechanism
anomalies. For example, [1] identify tribology problems ranging from wear with potentiometers due
to ineffective lubrication to high spin bearing torque in heavily loaded momentum wheels due to insuf-
ficient lubricant (inadequate initial amount, loss via transport process, chemical degradation, etc.).
A vast collection of space mechanism tribology issues can be found in the Proceedings of the Aerospace
Mechanisms Symposium and its European counterpart: European Space Mechanisms and Tribology
Symposium. Many of these lessons learned from these symposia are captured in Reference 2. Good
design practices per applicable design standards such as AIAA mechanism and NASA design stan-
dards [3,4] in combination with strict controls on materials and processes plus ample testing help
avoid lubrication problems such as those shown in Table 63.1.

63.2  Solid versus Liquid Lubrication


Space lubricants can be either solid or liquid, and their selection is application dependent. Table 63.2
compares their strengths and weaknesses. Solid lubricant coatings and films are generally insensitive
to temperature and are useful down to cryogenic temperatures, whereas liquid lubricants are generally
limited to working temperatures no colder than approximately −35°C due to viscous drag, although
some perfluoropolyether (PFPE) oils can operate at somewhat colder temperatures. These limits are
sometimes remedied with strip heaters at the cost of extra electrical circuit complexities and power.
Solid lubricants have negligible volatility, which is an advantage near optical surfaces compared to
liquid lubricants. Liquid lubricants require labyrinth seals in combination with low-vapor-pressure oils
to mitigate potential optical contamination issues. However, solid lubricant films generally have shorter
operational lives than liquid lubricants because they can’t be replenished. Still, bearing life still may
be sufficient particularly when self-lubricating composite ball retainers are used to transfer additional
lubricant like PTFE to the balls and races.

63.2.1  Liquid Lubricants


Properties of some commonly used space liquid lubricants are shown in Table 63.3 [5–10]. Lubricant
viscosity affect not only bearing drag particularly at cold operating temperatures but the ability to form

TABLE 63.1  Problem Areas for Space Lubricants


General Liquid Solid (Dry) Films
Depletion/wear out Polymerization/degradation Improper surface preparation (adhesion)
Long-term storage Outgassing Oxidation in air
Friction/drag in vacuum Insufficient lubricant film thickness (EHD) Insufficient/excessive thickness
Corrosion protection Operating temperature limits Particulate debris generation
Contact stress limits Creep/migration properties in 1 G Debris buildup in bearings
Galling and fretting wear Compatibility with system materials Poor thermal conductivity
Inadequate lubricant quantity Dewetting Vacuum storage
Improper run-in Lubricant/additive volatility Coating spalling/delamination/debonding
Boundary lubrication wear
Poor electrical conductivity
Accelerated life testing validity
63-4 Design for Lubrication and Tribology

TABLE 63.2  Relative Merits of Solid and Liquid Space Lubricants


Oils and Greases
Advantages Disadvantages
Relatively long-lived High viscous drag at low temperatures
Easily applied Can outgas on optics
Low friction and noise Tendency to migrate
Relatively good thermal conductivity Difficult to accelerate life test
Will replenish until supply exhausted Poor electrical conductivity
Solid Lubricants
Advantages Disadvantages
Negligible volatility Limited life, no replenishment
Excellent low speed performance Sensitive application process
Wide temperature range (Cryo) Wear debris causes torque noise
Will not migrate from contact Particulate contamination
Can be electrically conductive Shorter life when tested in air
Can more easily accelerate life test Poor thermal conductivity

a thin elastohydrodynamic (EHD) lubricant film between the ball and the races at high temperatures.
Vapor pressure is a measure of the lubricant’s outgassing characteristics. This property not only governs
the loss of lubricant from vacuum evaporation but is also a significant factor in determining likelihood
of optical contamination.
Space oils and greases almost always need a synthesized hydrocarbon to enhance chemical and physi-
cal stability. Although less common, super-refined mineral oils such as SRG 60 and KG-80 have been
used to lubricate navigational gyroscope bearings, which have demanding life requirements (speeds up
to 15,000 rpm and mission lives in excess of 10 years). They are refined (purified) from conventional
mineral oils to enhance stability and improve low temperature viscosity characteristics approaching
that of synthetic hydrocarbons.

63.2.1.1  PFPE Lubricants


PFPE oils have a long-standing space mechanism heritage. They have excellent thermal and oxida-
tive stability. Straight chained (linear) PFPEs have exceptionally low volatility and cold pour char-
acteristics (see Figure 63.2). Branched PFPEs have higher film strength and wear protection at the
expense of higher volatility and higher cold-temperature viscosity. PFPEs have exceptional shelf life.
The study in Reference 11 concluded that there were no detrimental performance effects from PFPE
grease after 17 year storage in a spare rudder brake actuator. Despite their long-standing heritage,
PFPEs are considered to be poor boundary lubricants. This is due to their chemical reactivity with
Fe in steel [12,13] and formation of higher Lewis acids even under moderate sliding/stress conditions
and aggravated by their inability to accept conventional antiwear additives [14]. However, progress
has been made developing PFPE-compatible antiwear additives [15]. Also, longevity can be improved
using more chemically inert, ceramic-coated (TiN) or solid ceramic (Si 3N4) bearing balls as shown
by Reference 16.

63.2.1.2  Hydrocarbon Lubricants


Hydrocarbon oils such as the super-refined mineral oils, esters, and polyalphaolefins have inherently
good boundary performance, can accept additives, but have poor vapor pressure characteristics and
higher pour points than PFPEs (Table 63.3). They have traditionally been used in high cycle life gyro-
scope and attitude control wheel applications where long lubricant life is essential and housing enclo-
sures plus heaters can be used to mitigate their vapor pressure and pour point limitations.
TABLE 63.3  Space Liquid Lubricant Propertiesa
Vapor Pressure, Torr Pressure-
Specific Pour Viscosity
Viscosity, cSt Temperature, °C
Molecular Gravity Point, Coefficient,
Type Trade Name Weight Cold @ 40°C @ 100°C g/cc RT 20 100 125 °C GPa−1 @ 38°C
Mineral Oil
Super refined SRG 60 [5] 77.6 9.44 0.88 1 × 10−8 [5] −12
Super refined KG-80 [5] 520 @ 20°C −6
1 × 10 [5] −9
Coray 100
Space Mechanism Lubrication

Synthetic
MAC Nye synthetic 1000 80,500 cSt 108 14.6 0.841 1.7 × 10−13 [6] 1.4 × 10−10 [6] 4 × 10−7 [8] −55 11 [7]
oil 2001A 75,000 cP @ −40°C 3 × 10−11 @ 25°C 5.3 × 10−8 [10]
(Pennzane®)b [10]
MAC Nye synthetic 26,000 @ −40°C 52.9 9 0.841 1 × 10−7 [6] 1.2 × 10−7 [10] 8 × 10−4 [8] −63
oil 1001A 3.5 × 10−11 @ 25°C
(Pennzane) [10]
PFPE Brayco 815Z 10,000 6500 cSt 12,000 cP 148 45 1.85 3 × 10−12 [8] 8 × 10−8 [8] −72
@ −40°C
PFPE Fomblin Z-25 9500 263 @ 20°C 157 49 1.85 2.9 × 10−12 [6] 9.8 × 10−9 [6] −66 18 [6]
PFPE Kryton 143 3700 5000 @ −32°C 77 10.3 1.89 1.5 × 10−6 [6] 3 × 10−4 [6] −40 30 [6]
AB
Neopentyl ester Nye UC-7 75 @ 20°C −56 5 [7]
PAO Nye 186A 103 14.6 0.85 6.5 × 10−8 [9] −48 12.5 PAO-188
[7]
PAO Nye 179 7720 @ −40°C 28.5 5.5 0.825 9 × 10−7 [9] −62
Source: Derived from Sathyan, K. et al., Tribol. Int., 43, 259, 2010; Jones, W.R. and Jansen, M.J., Space tribology, NASA TM 2000-203324, 2000; Spikes, H.A., Estimation of the pressure-
viscosity coefficient of two silahydrocarbon fluids at two temperatures, Imperial College of Science, Tribology Section, London, U.K., Report TS024b/96, 1996; Venier, C.G. et al., Comparative
physical and tribological properties of three Pennzane fluids, SHF X-1000, SHF X-2000, and SHF X-3000, Proceedings of the 10th European Space Mechanism and Tribology Symposium,
September 24–26, San Sebastian, Spain, ESA SP-524, pp. 337–340, 2003; Fusaro, R.L. and Khonsari, M.M., Liquid lubricants for space application. NASA TM105198, 1992; Nye Lubricants
Inc., Nye Synthetic Oil 2001 and 1001A data sheets.
a Properties derived from lubricant vendor data sheets unless otherwise noted.

b Rheolube 2000 grease (NSO 2001 oil + sodium thickener) has a vapor pressure 5.5 E−8 at RT due to thickener outgassing. Also “A” designates the unformulated oil.
63-5
63-6 Design for Lubrication and Tribology

1 E+00

KG-80 Mineral oil


Krytox 143 AC

Nye 1001A (Neat)


Nye 179 PAO
1 E–02

Vapor pressure, torr @ 20°C


1 E–04

Brayco 815Z

Fomblin Z25

Nye 2001A (Neat)


1 E–06

1 E–08

1 E–10

1 E–12
(a)

100,000
90,000
80,000
Kinematic viscosity, cS

70,000
Pennzane 2001
60,000 Pennzane 1001
50,000 "Bray 815 Z"
40,000
30,000
20,000
10,000
0
–50 –30 –10 10 30
(b) Operating temperature, °C

FIGURE 63.2  Vapor pressure (a) and cold-temperature viscosity (b) of some common space oils. (From
Sathyan, K. et al., Tribol. Int., 43, 259, 2010; Jones, W.R. and Jansen, M.J., Space tribology, NASA TM 2000-203324,
2000; Venier, C.G. et al., Comparative physical and tribological properties of three Pennzane® fluids, SHF X-1000,
SHF X-2000, and SHF X-3000, Proceeding of the 10th European Space Mechanisms and Tribology Symposium,
September 24–26, San Sebastian, Spain, ESA SP-524, pp. 337–340, 2003; Fusaro, R.L. and Khonsari, M.M., Liquid
lubricants for space application, NASA TM105198, 1992. With permission.)

Multiply alkylated cyclopentane (MAC) fluids, introduced in the early 1990s, have friction and
boundary lubrication performance on a par with the best traditional hydrocarbon oils but also possess
outstanding volatility characteristic (low vapor pressure) comparable to linear PFPEs. However, their
cold-temperature viscosity characteristics are not as attractive (Figure 63.2). For example, flight bearing
lives with MAC-based space oils exceeded those with PFPE oils by more than an order of magnitude,
as shown in Figure 63.3, from flight-like scanner bearing vacuum life tests reported in Reference 17.
Differences between linear and branched PFPE lubricant life were statistically insignificant.
MAC oils that showed similar life results were found from ball-on-plate, spiral orbit tribometer (SOT)*
data compiled from two sources [18,19], as shown in Figure 63.4. Higher contact stress significantly

* The SOT is essentially a thrust bearing with a single ball loaded between two flats. Its rolling ball geometry better
simulates the contact in a ball bearing than does pure sliding testers such as the pin-on-disk. The ball is wetted with
an extremely small amount of lubricant (typically 50 μg) and tested under high contact stress to reduce test time.
Consequently, SOT results are more useful as a screening tool for determining relative lubricant/material life but are not
an appropriate substitute for full-scale bearing tests.
Space Mechanism Lubrication 63-7

95
90

70

50
Bray 815Z
Percent of bearings failed

7.3

10

5
8.2
131
Krytox 143AC Pennzane 2001

0.1
1 5 10 50 100 500 1000
Life, million cycles

FIGURE 63.3  Weibull plot of lubricant life advantage of MAC versus PFPE from angular ball bearing life tests
(From Bazinet, D.J. et al., Life of Scanner bearings with four space liquid lubricants. Proceedings of the 37th Aerospace
Mechanisms Symposium, May 19–21, Galveston, TX, NASA CP-2004-212073, pp. 333–341. With permission.)

shortened lubricant life. There was no clear life advantage for the MAC oil formulated with an antioxi-
dant and antiwear additive versus the same oil with no additives.
MAC greases also had more than an order of magnitude higher durability than PFPE greases from
SOT test results in Reference 20. This was true for MAC-based grease thickened with either a sodium
soap or PTFE particles (MAPLUB®) [20]. Adding MoS2 particles to the grease to enhance boundary
mode performance had a negligible effect on durability.
The relatively poor viscosity index of Pennzane 2001 oil has limited its usefulness to applications
above approximately −25°C (Figure 63.2) unless heaters are used. Reference 8 reports on a less viscous
version of this oil referred to as SHF X-1001 (an earlier designation of Pennzane 1001). It extends the
lower operating temperatures to approximately −35°C without sacrificing lubricant wear life. However,
its lower viscosity is accompanied by higher volatility (Figure 63.2).

63.2.2  Solid Lubricants


Solid (dry) lubricant films/coatings can typically be found in slip rings, one-time deployment
devices, cryogenic instruments, some despin mechanisms, and solar array drives [21,22]. Ion-plated
lead has been used for decades on European solar array drives [23,24]. Solid lubricants are useful
when cycle life requirements are modest particularly when liquid lubricants may cause a problem,
either when operating temperatures are extreme or when molecular contamination on adjacent
optics is a threat.
63-8 Design for Lubrication and Tribology

100,000
Oil MAC oil
(formulated and
2001A
unformulated)
2001
10,000 Z25
143AC

Normalized life, orbits/µg


1,000

PFPE oil
(linear and branched)
100

10
0 0.5 1 1.5 2 2.5
Mean hertz stress, GPa

FIGURE 63.4  Effect of contact stress on lubricant life of MAC and PFPE oils from spiral orbit tribometry. (Derived
from Jansen, M.J. et al., Relative lifetimes of several space liquid lubricants using a vacuum spiral orbit tribometer
(SOT), NASA TM 2001-210966, 2000 [open symbols]; From Buttery, M., An evaluation of liquid. Solid, and grease
lubricants for space mechanisms using a spiral orbit tribometer, Proceedings of the 40th Aerospace Mechanism
Symposium, May 7–9, NASA CP-2010-216272, pp. 59–71, 2010 [closed symbols]. With permission.)

Solid lubricant types and application processes should be tailored to the given application. Examples
range from gold-plated slip rings in combination with MoS2 alloyed brushes to ball bearings with bear-
ing races ion-sputtered with MoS2 films. Bonded molybdenum disulfide (MoS2) coatings are popular for
use with single-use deployment joints and release devices.
Diamond-like carbon (DLC) coatings have been investigated [25] for space use but are not very com-
mon. The carbon in the DLC coatings is generally graphite, which requires humidity to function prop-
erly. Consequently, their friction and wear rates are poor in dry environments such as a space vacuum
(Table 63.4). However, friction coefficients as low as 0.02 can be achieved in a vacuum provided hydro-
gen is introduced into the DLC microstructure (hydrogenated) [25]. However, air performance suffers
not unlike MoS2 films when tested in ambient environments [26].

TABLE 63.4  Comparison of Friction and Wear of Selected Lubricant Films in Contact with 440C Steel Balls
from Pin-on-Disk Tests
Ion-Sputtered Bonded Ion-Plated Ion-Plated Ion-Sputtered PA CVD
Environment MoS2 MoS2 Lead Silver DLC DLCa
Friction Vacuum 0.07 0.045 0.15 0.2 0.7 0.54
coefficient μ Air 0.1 0.14 0.39 0.43 0.12 0.07
Film wear life, Vacuum 274100 >1 M 30300 364800 <10 <10
passes to μ = 0.3 Air 277400 113600 82 8 >1 M >1 M
Wear rate, Vac-film 0.09 0.06 1.5 0.088 57 11
10−6 mm3/N-m Vac-ball 0.0025 0.0013 0.76 0.024 320 180
Air-film 0.24 2.4 3.7 55 0.17 0.1
Air-ball 0.15 0.081 0.36 12 0.041 0.023
Source: Derived from Miyoshi, K., Friction and wear properties of selected solid. Lubricating films: Case studies, NASA
TM-2000-107249/chapter 6, 2000.
a Plasma-assisted, chemical-vapor-deposited DLC.
Space Mechanism Lubrication 63-9

63.2.2.1  Molybdenum Disulfide


Coatings containing lamellar solid MoS2 are arguably the most commonly used space solid lubricant.
They typically produce lower coefficients of friction and longer life in a space vacuum than other types
of solid lubricants (Table 63.4 [26]). Also, processes to apply them are well developed. The wear rate and
friction of MoS2 films are adversely affected by moisture in air environments, which needs to be consid-
ered during ground testing and storing mechanisms. This degradation is caused by chemical reaction
with water vapor and oxygen in the air to produce molybdic oxide (MoO3) [27]. Soft metal silver and lead
films will also oxidize in air penalizing performance (Table 63.4).
Burnishing, incorporation into a bonded coating, and ion sputtering are the main ways of applying
MoS2 lubricant to the contact surface with many material and process variations within each category.
Burnishing involves rubbing MoS2 powder into the surface texture producing modestly adherent plate-
lets of lubricant. It is the simplest technique but is the least durable [27,28]. It is mainly suited for in situ
field repairs of light duty components (one shot latches and hinge pins). Bonded MoS2 coatings involve
mixing MoS2 particles into a binder resin. They can produce relatively long wear lives (Table 63.4), but
are generally too thick (≈10 to 15 μm) for ball bearings and other critical clearance applications, for
example, see Table 15.4 in Reference 27.
Vacuum deposition ion-beam sputtering techniques also produce long-lived solid films. Their film
thinness (<1 μm) and high adherence make them excellent candidates for ball bearing applications.
However, the sputtering process requires careful control of the relevant parameters, and availability of
sputtering vendors is limited.
High-velocity spray impingement of tungsten disulfide (WS2) is only 0.5 μm thick. It has the
potential for lubricating precision components, but there is large variation in life from different
vendors and coating life is 2–3 orders of magnitude shorter than sputtered MoS2 wear life according
to Reference 29.

63.2.2.2  Bonded Coatings


MoS2 lubricants are generally applied in the form of a heat-cured bonded coating where the MoS2 par-
ticles are relatively uniformly dispersed in a binder, not unlike pigments mixed into a paint resin. The
binder, typically a phenolic (organic) or silicate (inorganic) resin, improves adhesion to the part and
provides a relatively large concentration of MoS2 powder in the binder (typically MoS2-to-binder ranges
from 1:1 to 2:1), which enhances durability [27,30].
Organic binder systems generally have better wear life than inorganic systems, which are brittle [27],
but the latter are better suited to high radiation and higher temperature environments in addition to
their compatibility with liquid oxygen [30]. A small amount (10%) of graphite is sometimes added to
increase wear life in humid air [27]. Property data and supplier information on more than 250 bonded
solid and liquid lubricants for space applications can be found in Reference 31.
Bonded coatings are generally not selected for critical clearance applications such as ball bearings.
They are more useful for gears, deployment sliding joints, and other more open contacts where debris-
induced torque noise and wear debris contamination are not critical.
Premature bonded-coating failures are often the result of improper surface preparation that may
have jeopardized good adhesion leading to debonding. Pretreatments are dependent on the type of
metal surface but normally involve mechanical surface roughening (vapor or sandblasting) and/or
chemical (phosphating, chemical etch, etc.). Pretreatments can increase coating wear life by orders of
magnitude [32].

63.2.2.2.1 Ion-Beam-Sputtered Films


Vacuum deposited, ion-sputtered MoS2 films are generally the preferred solid lubricant for ball bearings
and other precision clearance applications. Long wear life can be achieved even with thin (<1 μm) films
[33–35]. Also, thickness variation is very controlled. The sputtering deposition process effectively builds
63-10 Design for Lubrication and Tribology

Sputtered MoS2 film


ML = Multilayer, RFM = Radio frequency
magnetron, DC=Direct current, None = no coating
ML
ML
ML
Br +PTFE
RFM
cages
DC
None
ML
Unfailed at
RFM
test cutoff
PTFE DC
cages DC
RFM
ML
DC
Polyimide DC
cages ML
RFM
Failed at
RFM 50 N–mm
PTFE cages DC
+ SiN balls
ML
DC
ML
DC
MoS2 coated ML
balls
RFM
RFM

0 5 10 15 20 25 30 35 40 45 50
Millions of gimbal cycles

FIGURE 63.5  Life of sputtered MoS2 angular contact ball bearings in vacuum. (From Loewenthal, S.H. et al.,
Tribol. Trans., 37(3), 505, 2007. With permission.)

up the coating of few molecular layers at a time. Equally important is the high coating adhesion. This is
achieved through ion cleaning molecular contaminants from the part’s surface prior to coating along
with the highly energetic bombardment of the part surface with the coating atoms [30,33,34].
MoS2 films can be layered or co-sputtered with metallic (Au, Ag, or Ni) dopants to improve ball
bearing durability. Ball bearing life and torque signature were investigated in Reference 35 with three
types of ion-sputtered MoS2 films. These included multilayer (ML) MoS2 films separated by thin layers
of Ni for improved ductility. Also, DC triode (DC)-sputtered MoS2 films containing about 35% Sb2O3
were evaluated along with radio frequency magnetron (RFM) films sputtered at high deposition rates
for increased film density. Life tests were conducted in a vacuum over a simulated duty cycle for a
space payload gimbal. Self-lubricating PTFE containing ball retainers were required for long life (see
Figure 63.5). Bearings with polyimide retainers, silicon nitride ceramic balls, or steel balls sputtered
with MoS2 film suffered early torque failure from debris buildup, irrespective of the type of sputtered
race film. Comparable grease-lubricated bearings produced lives that were generally longer with lower
torque signatures [35].

63.2.2.3  Soft Metallic Coatings


Low shear strength metals such as gold, silver, and lead have also found use on satellites because their
friction is insensitive to temperature or vacuum and they possess high electrical and thermal conduc-
tivity. Solid or electroplated gold and silver alloys are routinely used for slip rings and slip ring brushes.
Some use gold wire brushes on gold rings [36], while others use gold alloyed with silver brushes. Gold
Space Mechanism Lubrication 63-11

0.6
0.5

Torque, N mm
0.4 Peak
0.3
0.2
0.1 Mean
(a) 0
0.2

Torque, N mm
0.15

0.1 Peak

0.05
Mean
(b) 0
0.05
Torque, N mm

0.04
0.03
0.02 Peak
0.01
Mean
0
(c) 0 0.2 0.4 0.6 0.8 1 1.2 1.4
Revolution, millions

FIGURE 63.6  Torque comparison of ball bearings in vacuum lubricated with (a) gold, (b) lead, and (c) MoS2 thin
films. (Derived from Roberts, E.W., Tribol. Int., 23(2), 95, 1990.)

plate is standard for electrical connector pin contacts. Ion-plated lead films have found widespread use
on bearings and gears for European solar array drives [24], although liquid lubricants continue to find
favor in the United States. Ion-plated lead has excellent adhesion; low, temperature-independent torque
down to cryogenic temperatures; and very long life in a vacuum [33]. However, operation in air must be
limited to fairly low speed and short duration due to oxidation. Furthermore, torque noise is not on a
par with liquid lubricants.
Ion-plated lead 20 × 42 mm duplex angular-contact ball bearings outfitted with a leaded bronze cage
(Figure 63.6) produced steady low torque for >1.3 million revolutions when tested in a vacuum [33].
Torques were even better for bearings with ion-sputtered MoS2 bearing races and a Duroid® PTFE com-
posite cage. Bearings with ion-plated gold race produced average torques similar to lead but had substan-
tial higher torque peaks (Figure 63.6) attributed to the presence of gold wear particles as a by-product of
its tendency to work harden [33].

63.3  Mechanism Components


63.3.1  Bearings
Oil- or grease-lubricated ball bearings are generally the first choice for most space actuators, drives, and
gimbals. However, self-lubricating plain bearings are useful for some mechanism applications.
Sleeve or pivot bearings made from self-lubricating polymers such as PTFE reinforced with fiber
glass and polyimides filled with a small percent of MoS2 have proven useful for limited life applica-
tions such as deployment hinges and latches. They are not only of simple construction but can provide
redundancy by having inner and outer surfaces free to rotate. However, polymers generally have poor
thermal conductivity and limited load and speed capabilities relative to ball bearings nor can they be
easily preloaded to avoid gaps or to provide high stiffness.
63-12 Design for Lubrication and Tribology

63.3.2  Lubricant Film Thickness


Liquid lubricants form a protective film thickness to separate the surface asperities of ball bearings and
gear contacts when the parts are in motion. The lubricant film is subjected to Hertzian contact stresses
on the order of hundreds of MPa when entrained in the contact. This greatly increases the viscosity of
the fluid to where it behaves like a glassy solid tending to separate the surfaces. The resulting separation
is generally referred to as the EHD film thickness and is reasonably predictable:

hc
Hc = = 2.69U 0D.67G0.53W −0.067 (1 − 0.61e −0.73k ) (63.1)
Rx

Expressions for each of these subterms can be found in Chapter 18 or from other references (e.g., [37])
and therefore will not be repeated here. However, the nondimensional EHD central film thickness term
Hc is mainly sensitive to the product of lubricant’s absolute viscosity η and entrainment velocity through
the dimensional speed parameter UD ∝ η. It is moderately sensitive to the lubricant’s pressure-viscosity
coefficient α (dimensionless material term G ∝ α) and relatively insensitive to normal load (dimension-
less load term W). Kinematic viscosity η and pressure-viscosity coefficient α. for typical space lubricants
are given in Table 63.3.
Maximizing the lubricant film thickness is a key contributor to component life (see Figure 63.7). The
lubricant film parameter or lambda ratio (λ) is a measure of the lubricant’s effectiveness where λ is the
ratio of the EHD central film thickness to the composite surface roughness of the contacting surfaces
(root-mean-square surface finishes of ball and race). It is also an indicator of the contact’s operating
lubrication regime. This in turn is an indicator of its likely failure mechanism (i.e., wear or fatigue pit-
ting). Although not precisely defined, the EHD regime typically ranges from λ approximately 1.5 to 3
[37]. This is generally considered to be a desirable lubrication regime for bearing and gear contacts since
predicted lubricant life is not significantly reduced as it is for lower λ values. Note that the value of the
lubrication-life correction factor (Figure 63.7) to be applied is not precisely known since this factor is
intended to be applied to fatigue life (pitting) predictions [37] and space bearings/gears almost never fail
from fatigue.
Transition to the boundary lubrication regime begins below a λ less than approximately 1. Here,
surface asperity interaction grows more intense as λ diminishes until there is considerable rubbing and

3.5
Hydrodynamic
Lubrication-life correction factor

3.0
EHD

2.5 Mixed

2.0 Boundary

1.5
Hard
1.0 boundary
starved
0.5

0.0
0.1 0.2 0.5 1 2 5 10
Lubricant film parameter (λ ratio = EHD film thickness/composite surface roughness)

FIGURE 63.7  Lubrication regimes and relative bearing fatigue life as a function of EHD central film thickness
and contact surface roughness. (Derived from Zaretsky, E.V. (Ed.), STLE Life Factors for Rolling Bearings, STLE
SP-34, Park Ridge, IL, 1992.)
Space Mechanism Lubrication 63-13

potential for wear. This region is considered by some to be hard boundary lubrication. In this region,
there is a complex chemical and physical interaction of the lubricant with the surface that cannot be
easily quantified. Reliance on lubricant life test data is therefore a necessity.
Many space mechanisms rotate very slowly, such as those that operate at low earth orbit rates
(≈ 1 rev/90 min). The contacts spend majority of operating life in the boundary lubrication region. This
not only has life implications but complicates the ability to do accelerated life test since speeding up the
test rate can unintentionally improve the mode of lubrication and make the test nonconservative.
Other regions of lubrication shown in Figure 63.7 include the mixed region, which is the mixture of
boundary and EHD. On the other end of the λ curve, components rotating at high speeds and have light
loads may approach the hydrodynamic regime (λ > 3) where the surface film is considerably thicker than
the composite surface roughness and life is maximized. Unfortunately, few space bearings operate in the
hydrodynamic region. This is due in part to starvation effects from insufficient time between ball passes
for the lubricant to flow back into the ball track causing the lubricant film thickness to fall short of its
theoretical prediction. This starvation effect is more prevalent with grease lubrication and can occur at
much lower speeds where space mechanism components generally operate. (Lugt’s survey [38] is a good
source of information on the complexities of grease lubrication in general and contains detailed discus-
sion on grease flow and film thickness starvation effects.)

63.3.3  Cage Impregnation


As previously discussed in Section 63.2.2, solid lubricated bearings generally use self-lubricating cages
to separate the balls. Phenolic cages or ball retainers are generally specified when used with oil and
grease lubrication, although brass, polyimide, and PTFE cages have also been used. Phenolic cages made
from tubes of wound cotton fabric bonded with phenolic resin matrix typically have a small amount of
porosity (approximately 3%–6%). It was thought that an oil-impregnated porous cage in combination
with the absorbent cotton fibers would act as an oil reservoir and deliver oil back to the bearing as the
initial oil/grease charge was being depleted. However, the testing in Reference 39 showed that no net oil
is delivered to a poorly lubricated bearing even with a fully impregnated ball retainer. In fact, a partially
impregnated retainer continues to absorb oil even when operating. Furthermore, it takes more than a
week of soaking the retainer in oil after vacuum drying it out to reach impregnation levels of approxi-
mately 40%–60% of the retainer’s total oil capacity [40].

63.3.4  Lubricant Loss Mechanisms


Space mechanisms are generally once lubricated for the entire mission life. Prevention of lubricant loss
mechanisms such as oil migration and outgassing are therefore critical for long life. Also, chemical con-
tamination of the surface leading to dewetting is yet another concern.
Outgassing rate is primarily determined by the volatility of the lubricant in a vacuum at operating
temperature. Normal practice for space mechanism acceptability is to show that materials used in its
construction, including lubricants, have a total mass loss (TML) of less than 1% and that the collected
volatile condensable material (CVCM) is less than 0.1% when heated to 125°C for 24 h in a vacuum per
the ASTM E-595 standard [41]. The low-vapor-pressure lubricants shown in Table 63.3 are preferred.
An outgassing calculation is normally necessary to determine if there will be sufficient lubricant
left in the bearing at the end of the mission. A study in Reference 42 was conducted to evaluate the
effectiveness of labyrinth seals of different geometries in reducing the mass loss of three common space
oils versus the predicted amount. The theoretical outgassing rate is proportional to the area of the vent
opening and the vapor pressure of the lubricant and inversely proportional to the square root of the
absolute temperature in accordance with Equation 63.2. The measured mass loss showed approximate
agreement with prediction for the two more volatile oils, but the loss of the MAC oil was significantly
underpredicted [42]:
63-14 Design for Lubrication and Tribology

0. 5 −1
⎛ M⎞ ⎛ 0.375L ⎞
Q = 0.0436Ppdb ⎜ ⎟ ⎜⎝ 1 + ⎟ (63.2)
⎝ T⎠ b ⎠

where
Q is the mass loss in g/s
P is the fluid vapor pressure in torr
d is the diameter of annular seal (mm)
b is the gap width (mm)
L is the path length (mm)
M is the molecular weight
T is the absolute temperature (K)

Oil migration is another loss mechanism to be concerned about. Oils have a tendency to migrate
away from the contact working surfaces over time. This can lead to oil loss during long storage in the
presence of gravity. Orienting bearing openings not in the direction of gravity is prudent for long stor-
age, and periodic exercising may also be necessary for reliability. The low surface tension of most space
oils is a desirable attribute. It enhances the ability to wet the working surface. However, low surface
energy materials will readily creep along the surface unless they encounter a material with lower surface
energy. At that point they will have difficulty crossing. Exceptionally low energy anticreep barrier films
have been developed that can be carefully applied near a bearing to inhibit oil migration.
Another prudent design practice is to require an in-process wettability test to be performed on space
bearing surfaces during acceptance testing. This is to guard against the unlikely possibility of surface
chemical contamination. The wettability test usually consists of inspecting surfaces for oil beading after
being wetted with a highly solvent diluted oil film.

63.3.5  Gears and Harmonic Drives


Liquid lubricants are commonly used to lubricate speed reducer gears and harmonic drives (high ratio
compact drives with a flexible tooth member) commonly used on solar array drives. Although solid-
film-lubricated gears are not as common, life data for instrument and worm gears can be found in
References 31 and 43.
Gear drives are generally more demanding than rolling-element bearings. There is considerable tooth
sliding at the entry and exit of the tooth mesh in addition to rougher surface finishes. Gear tests con-
ducted in a vacuum [44] on a variety of dry coating and material combinations reported that plasma-
nitrided gears from En40B nitriding steel had the lowest total wear rate. Hard anodized layer on the
aluminum gears showed early cracking. Bray 601 grease-lubricated AISI 440C versus plasma-nitrided
gears showed no measureable wear.
A variety of different lubricants have been successfully used on harmonic drives in different space
applications according to Reference 45. The most common are MAC- and PFPE-based greases. Gold
plating on the drive teeth has been used to reduce chemical degradation of the grease and extend
longevity.

63.3.6  Slip Rings


Electrical power and signal across continuously rotating joints are typically transferred using slip rings,
although roll rings have also been used occasionally [36]. Applications include transferring power from
solar arrays to spacecraft batteries through slip rings in solar array drives, despin mechanisms, control
moment gyroscope CMG gimbals, etc. Traditionally, two more brush blocks comprised mainly of 85%
silver with 12% MoS2 and 3% graphite [36,46] are in sliding contact with rotating rings of either coined
Space Mechanism Lubrication 63-15

silver or alloyed rings with a silver plate. Self-lubricating brush slip rings perform well in vacuum but are
vulnerable in air to oxidation, which can increase signal noise [46].
Another configuration is single filament wire (normally two per channel) made from hard alloyed
gold sliding in V-shaped rings plated with a hard gold flash. The main advantage according to Reference
46 is tighter ring spacing resulting in higher current density. Also, the space at the bottom of the V
allows room for wear debris collection. A variation is the multifiber wire brushes with flattened ends
that sit in V-shaped grooves, again gold on gold. The multiple fibers mean that the individual load on
the fiber is significantly smaller than that of a single fiber contact. The scrubbing action against the sides
of the V helps maintain low electrical resistance and leave some room at the bottom of the V for debris.
However, the axial spacing is about twice that of the monowire brush design [46].
For long life, low-vapor-pressure liquid lubricants have been used. However, PFPE oils because of
their tendency to degrade in boundary regime sliding would not be able to compete with MAC or PAO
synthetic hydrocarbon oils [46]. Also, attention has to be paid to oil outgassing and migration.

63.4  Concluding Remarks


Lubricated for life space mechanisms that need to perform for mission lives of up to 15 years or longer
in the hostile environment of space presents a significant challenge in the selection of the appropriate
lubrication approach and its test verification. Tribal knowledge of what does or does not work well is
therefore of paramount importance. The review presented here had space to only touch upon the high
points of a very broad and demanding topic. Fortunately, there is a wealth of design and user experi-
ence information available to the designer in the open literature to supplement what was reported here.
References 47–53 should be of value to the reader.

References
1. Fleischauer, P.D. and Hilton, M.R. 1991. Assessment of the tribological requirements of advanced
spacecraft mechanisms. Aerospace Report No. TOR-0090(5064)-1.
2. Fusaro, R. (ed.). NASA Space Mechanisms Handbook. 1999. NASA TR1999-206988.
3. Moving Mechanical Assemblies for Space and Launch Vehicles. 2005. AIAA S-114-2005.
4. Lubrication, friction and wear. NASA Space Vehicle Design Criteria. 1971. NASA SP-8063.
5. Sathyan, K., Hsu, H.Y., Lee, S.H., and Gopinath, K. 2010. Long-term lubrication of momentum
wheels used in spacecraft—An overview. Tribol. Int., 43, 259–267.
6. Jones, W.R. and Jansen, M.J. 2000. Space tribology. NASA TM 2000-209924, pp. 17–18.
7. Spikes, H.A. 1996. Estimation of the pressure-viscosity coefficient of two silahydrocarbon fluids at
two temperatures. Imperial College of Science, Tribology Section, Report TS024b/96, London, U.K.
8. Venier, C.G., Jones, W.R. Jr., Jansen, M.J., and Marchetti, M. 2003. Comparative physical and tribo-
logical properties of three Pennzane® fluids, SHF X-1000, SHF X-2000, and SHF X-3000. Proceedings
of the 10th European Space Mechanisms and Tribology Symposium, September 24–26, San Sebastian,
Spain, ESA SP-524, pp. 337–340.
9. Fusaro, R.L. and Khonsari, M.M. 1992. Liquid lubricants for space application. NASA TM105198.
10. Nye Lubricants Inc., Nye Synthetic Oil 2001A and 1001A data sheets.
11. Munafo, P. et al. 2004. Orbiter Rudder Speed Brake Actuator Braycote Grease Independent Technical
Assessment/Inspection (ITA/I) NASA Engineering and Safety Center Report. NASA RP-04-03/
03-003-E.
12. Carré, D.J. 1988. The performance of perfluoropolyalkylether oils under boundary lubrication con-
ditions. Tribol. Trans., 31, 4, 437–441.
13. Mori, S. and Morales, W. 2008. Tribological reactions of perfluoroalkyl polyether oils with stain-
less steel under ultrahigh vacuum conditions at room temperature. Wear, 132, 111–121. NASA/
TM-2008-215184.
63-16 Design for Lubrication and Tribology

14. Street, K.W. Jr. 2009. Liquid space lubricants examined by vibrational microspectroscopy. NASA
TM-2008-215184.
15. Gschwender, L.J. and Snyder, C.E. Jr. 2009. U.S. air force perfluoropolyalkylether experiences. Tribol.
Trans., 52, 165–170.
16. Carré, D.J. December 1990. The use of solid ceramic and ceramic hard-coated components to pro-
long the performance of perfluoropolyalkylether lubricants. Surf. Coating Technol., 43–44, Part 2,
609–617.
17. Bazinet, D.J., Espinosa, M.A., Loewenthal, S., Gschwender, L., Jones, W.R. Jr., and Predmore, R. 2004.
Life of scanner bearings with four space liquid lubricants. Proceedings of the 37th Aerospace
Mechanisms Symposium, May 19–21, Galveston, TX, NASA CP-2004-212073, pp. 333–341.
18. Jansen, M.J., Jones, W.R. Jr., Predmore, R.E., and Loewenthal, S.H. 2000. Relative lifetimes of several
space liquid lubricants using a vacuum spiral orbit tribometer (SOT). NASA TM 2001-210966.
19. Buttery, M. 2010. An evaluation of liquid, solid, and grease lubricants for space mechanisms using a
spiral orbit tribometer. Proceedings of the 40th Aerospace Mechanisms Symposium, May 7–9, NASA
CP-2010-216272, pp. 59–71.
20. Marchetti, M., Jones, W.R., Jr., and Sicre, J. 2002. Relative lifetimes of MAPLUB® greases for space
applications. NASA TM 2002-211875.
21. Hilton, M.R. and Fleischauer, P.D. 1992. Applications of solid lubricant films in spacecraft. Surf. Coat.
Technol., 54/55, 435.
22. Jones, W.R. Jr. and Jansen, M.J. 2005. Lubrication for space applications. NASA CR-2005-213424.
23. Briscoe, H.M. and Todd, M.J. 1983. Considerations on the lubrication of spacecraft mechanisms.
Proceedings of the 17th Aerospace Mechanisms Symposium, May 5–6, Pasadena, CA, NASA CP-2273,
pp. 19–37.
24. Rowntree, R.A. and Todd, M.J. 1989. A review of European trends in space tribology and its applica-
tion to spacecraft mechanism design. Mater. Res. Soc. Symp. Proc., 140, 21–34.
25. Erdemir, A. and Donnet, C. 2006. Tribology of diamond-like carbon films: Recent progress and
future prospects. J. Phys. D: Appl. Phys., 39, R311–R327.
26. Miyoshi, K. 2000. Friction and wear properties of selected solid. Lubricating films: Case Studies.
NASA TM-2000-107249/chapter 6.
27. Didziulis, S.V., Lince, J.R., Carré, D.J., and Hilton, H.R. 1999. Lubrication, NASA Space Mechanisms
Handbook, Chapter 15. NASA TP-1999-206988, Cleveland, OH.
28. Rittenhouse, J.B. and Singletary, J.B. 1969. Space Materials Handbook. NASA SP-305, Cleveland, OH,
pp. 216–227.
29. Anderson, M.J., Cropper, M., and Roberts, E.W. 2007. The tribological characteristics of dicronite.
Proceedings of the 12th ESMATS, September 19–21, Liverpool, U.K., ESA SP-653.
30. Bhushan, B. and Gupta, B.K. 1991. Handbook of Tribology: Materials, Coatings, and Surface
Treatments. McGraw-Hill Book Company, New York.
31. McMurtrey, E.L. 1985. Lubrication Handbook for the Space Industry (Part A: Solid Lubricants; Part B:
Liquid Lubricants). NASA TM-86556, Cleveland, OH.
32. Gresham, R.M. October 2003. Solid film lubricants: Unique products for unique lubrication. Tribol.
Lubr. Technol., 59(10) 28–31.
33. Roberts, E.W. April 1990. Thin solid lubricant films in space, Tribol. Int., 23(2), 95–104.
34. Roberts, E.W. 1986. Towards an optimized sputtered MoS2 lubricant film. 20th Aerospace Mechanism
Symposium, May 7–9, Cleveland, OH, NASA CP-2423, pp. 103–119.
35. Loewenthal, S.H., Chou, R.G., Hopple, G.B., and Wenger, W.L. July 2007. Evaluation of ion-sputtered
molybdenum disulfide bearings for spacecraft gimbals. Tribol. Trans., 37(3), 505–515.
36. Lince, J.R. 1999. Electrical contact ring assemblies. NASA Space Mechanisms Handbook, Chapter 16.
NASA TP-1999-206988, Cleveland, OH.
37. Zaretsky, E.V. (Ed.). 1992. STLE Life Factors for Rolling Bearings, STLE SP-34, Park Ridge, IL.
38. Lugt, P.M. 2009. A review on grease lubrication in rolling bearings. Tribol. Trans., 52, 470–480.
Space Mechanism Lubrication 63-17

39. Bertrand, P.A., Carré, D.J., and Bauer, R. April 1995. Oil exchange between ball bearings and cotton-
phenolic bail-bearing retainers. Tribol. Trans., 38(2), 342–352.
40. Bertrand, P.A. 1993. Oil absorption into cotton-phenolic material. J. Mater. Res., 8(7), 1749–1757.
41. ASTM E595-07 Standard test method for total mass loss and collected volatile condensable mate-
rials from outgassing in a vacuum environment, ASTM International, West Conshohocken, PA,
p. 19428.
42. Anderson, M., Freeman, S., and Roberts, E. 1999. Evaporative losses of vacuum-compatible oils
through labyrinth seals. Proceedings of the 8th European Space Mechanism and Tribology Symposium,
September 29–October 1, Toulouse, France, ESA SP-438, pp. 225–231.
43. Rowntree, R.A. and Vine, M.K. 1987. Thin solid films for satellite gears. Proceedings of the 3rd
European Space Mechanisms and Tribology Symposium, September, Madrid, Spain, ESA-SP-279,
pp. 149–154.
44. Stevens, K.T. 1983. The tribology of gears for satellite applications. Proceedings of the First European
Space Mechanisms and Tribology Symposium, October 12–14, Neuchâtel, Switzerland, ESA-SP-196.
45. Schafer, I., Bourlier, P., Hantschack, F., Roberts, E.W., Lewis, S.D., and Forster, D.J. 2005. Space lubrica-
tion and performance of harmonic drive gears. Proceedings of the 11th European Space Mechanism
and Tribology Symposium, September 21–23, Lucerne, Switzerland, ESA SP-591, pp. 21–23.
46. Phinney, D.1986. Slip ring experience in long duration space applications. Proceedings of the 20th
Aerospace Mechanisms Symposium, May 7–9, Cleveland, OH, NASA CP-2423, pp. 45–54.
47. Zaretsky, E.V. 1990. Liquid lubrication in space. Tribol. Int., 23(2), 75, and NASA RP-1240.
48. Fusaro, R.L. (Ed.) 1999. NASA Space Mechanisms Handbook, NASA TP-1999-206988, Cleveland, OH.
49. Miyoshi, K. 2007. Solid lubricants and coatings for extreme environments: State-of-the-Art Survey.
NASA TM- 2007-214668, 2007.
50. Shapiro, W., Murray, F., and Howarth, R. 1995. Space mechanisms lessons learned Study
Volume I—Summary. NASA TM-107046.
51. Anon. 2008. NASA standard materials and processes requirements for spacecraft. NASA-STD-6016.
52. Staugaitis, L. (ed.). 1975. NASA Materials Guidelines. NASA SP-3094.
53. Kawashima, N., Hiraoka, N., and Yoshii, Y. 1996. Space tribology in Japan. Proc. I Mech. E. 210(J3),
173–178.
64
Magnetic Storage
64.1 Introduction.....................................................................................64-1
Structure of a Modern Hard Disk Drive  •  Reduction in Slider’s
Flying Height to Nanoscale Spacing
64.2 Historical Background: Head-Disk Interface and
Air-Bearing Sliders..........................................................................64-5
64.3 Generalized Reynolds Equation....................................................64-7
Numerical Approach
64.4 Slider’s Static Performance.............................................................64-8
Intermolecular Force  •  Raised Temperature and Altitude
64.5 Slider’s Dynamic Performance......................................................64-9
Contact Model  •  Intermolecular Force  •  Load/Unload  •  Shock
64.6 Thermal Flying-Height Control Sliders..................................... 64-11
Nan Liu 64.7 Outlook............................................................................................64-12
University of Helium-Filled Hard Disk Drives  •  Bit-Patterned Media  •  Heat-
California, Berkeley Assisted Magnetic Recording  •  Lube-Surfing Recording
64.8 Conclusion...................................................................................... 64-14
David B. Bogy
University of
Acknowledgments.....................................................................................64-15
California, Berkeley References...................................................................................................64-15

64.1  Introduction
Hard disk drives (HDDs) were invented by R. Johnson in 1954 and were introduced by IBM 2 years
later as an alternative to the then existing magnetic storage devices such as magnetic drums and floppy
disks (Harker et al., 1981; Stevens, 1981). The first HDD, weighing over one ton and shown in Figure 64.1
(Wikipedia, 2010), does not bear much similarity with its modern counterparts. It used 50 disks to store
information data, which were read and written by two transducers. Each transducer was embedded in
an externally pressurized head, as shown in Figure 64.2 (Harker et al., 1981). During the HDDs’ opera-
tion, compressed air was supplied to the head such that the head was positioned hundreds of microm-
eters from the disk. Since this HDD had only one pair of heads, the heads needed to move from one disk
to another in order to read/write information data on different disks. The first generation HDDs needed
regular maintenance, and IBM leased them to their customers for $3500 per month, instead of selling
them (Wikipedia, 2010).
The next two decades saw much improvement and technological innovation, such as lightly loaded
self-acting air-bearing sliders, lubricated disks, and voice coil servomotors, applied to HDDs, which
finally led to the appearance of IBM’s Winchester HDD. Despite their large size and small capacity
when compared to modern HDDs, the Winchester HDDs had an almost identical inner structure and
working principle as the modern HDDs, which are sometimes also called Winchester-type HDDs. The
Winchester HDD was so reliable that regular maintenance was no longer needed, and HDDs could be
sold to customers at a reasonable price, which paved the way to the popularization of IBM’s personal

64-1
64-2 Design for Lubrication and Tribology

FIGURE 64.1  First HDD. (From Wikipedia, http://en.wikipedia.org/wiki/Early IBM disk storage#IBM 350, 2010.)

Air in

Head
Air
Air
Disk
Air Air
Head

Air in

FIGURE 64.2  Hydrostatic slider used in the first-generation HDDs.

computer (PC). Further improvement in the HDDs’ performance appeared in the following several
decades, which enabled HDDs to dominate the magnetic recording data storage market.
HDD performance mainly refers to its capacity, speed, and reliability. Capacity is characterized
by the HDD’s areal density defined as the number of bits that can be stored in 1 in.2 on the magnetic
disk. This density is now approaching the superparamagnetic limit, which limits further miniatur-
ization to a magnetic grain. Accordingly, the HDDs’ annual compound growth rate has decreased
from 100% during the 1990s to about 40%, which is still larger than the growth rate for the number
of transistors in a chip predicted by Moore’s law (Grochowski and Halem, 2003). The HDDs’ areal
density is expected to be larger than 10 Tbit/in.2 in the next decade as the emerging techniques such as
bit-patterned media and heat-assisted magnetic recording (HAMR) become mature (Shiroishi et al.,
2009; Wood, 2009). An HDD’s speed is characterized by its disk rotation speed and its track seeking
time, which has decreased from 10 ms in 1990s to the current 3 ms, and an HDD’s mean time between
failure (MTBF), characterizing the HDD’s reliability, has increased from less than 50,000 h in 1990s
to the current 1,500,000 h.
HDDs can have different form factors that refer to the size of the HDD’s disk diameter. In the his-
tory of HDDs, several form factors have appeared and then disappeared. Most recently, the 0.8 and
1.0 in. HDDs with a capacity of tens of gigabytes disappeared due to the competition from solid-state
“flash” drives with a comparable capacity. Nowadays, HDDs mainly have three form factors: 3.5 in.
Magnetic Storage 64-3

for desktop application, 2.5 in. for laptop application, and 1.8 in. for mobile application such as MP3
players and smart phones. Despite their different form factors, all the HDDs have essentially the same
structure.

64.1.1  Structure of a Modern Hard Disk Drive


A modern HDD with its cover plate removed is shown in Figure 64.3. The HDD usually contains up to
four magnetic disks to store information. Figure 64.4 shows a close-up diagram of a typical head-disk
interface (HDI). The information data are stored on a 20–50 nm thick magnetic layer deposited onto the
substrate. On top of the magnetic layer, an ultrathin diamond-like carbon (DLC) overcoat of 2–5 nm
thickness is deposited to protect the magnetic layer from impact, corrosion, and wear. The DLC overcoat
is covered by a layer of lubricant of thickness less than 2 nm to reduce the DLC film’s wear and further
protect the magnetic layer.
The information data stored on the magnetic disk are read and written by a read-write transducer,
made by integrated circuit (IC) technology and embedded in a self-acting air-bearing slider. The slider
is attached to the end of a suspension, which itself is connected to a voice coil motor (VCM). When data
at different locations need to be manipulated, the VCM drives the suspension so that the slider can be
positioned over the designated location. For the HDDs’ reliability, the slider needs to fly stably over the
disk, that is, the slider’s flying attitude (height, pitch, and roll) should change no more than 10% when it
is positioned over different locations on the disk. This stable flying attitude is achieved through a preci-
sion design of the pattern on the slider’s surface facing the disk, known as the air-bearing surface (ABS).
A typical ABS design is shown in Figure 64.5 where different colors represent different etch depths.
During the HDDs’ operation, the magnetic disk rotates at a speed up to 15,000 revolutions per minute
(RPM), producing a linear speed at the slider around 10–30 m/s (Liu et al., in press). The rotating disk
drags air (due to its viscosity) into the region between the slider and the disk, known as the HDI, and the

Cover mounting holes


(cover not shown)

Base casting

Spindle

Slider (and head)

Actuator arm

Actuator axis Case


mounting
Actuator holes

Platters

Ribbon cable
(attaches heads
SCSI interface to logic board)
connector
Jumper pins
Jumper Power Tape seal
connector

FIGURE 64.3  Modern HDD.


64-4 Design for Lubrication and Tribology

VCM Disk

Actuator

Sus
Spr pen
ing sion

Slid
er
Lubricant
Transducer
layer

Disk

Lubricant layer ~ 1 nm
Magnetic layer DLC overcoat ~ 2 nm
~50 nm

Substrate

FIGURE 64.4  HDI and structure of a modern HDD.

0.7

0.5
y (mm)

0.3

0.1

0
0 0.2 0.4 0.6 0.8
x (mm)

FIGURE 64.5  Typical ABS design.

air gets compressed and rarefied by the slider’s ABS, so that the total air pressure on the ABS balances
the load applied by the suspension and the slider is kept away from the disk at a designated distance.

64.1.2  Reduction in Slider’s Flying Height to Nanoscale Spacing


Information data are read out and written onto the disk through the read-write transducer embedded
in the slider. The density of information that can be stored on the disk is determined by the ability
Magnetic Storage 64-5

y (mm)
1

0
0 1 2 3 4
x (mm)

FIGURE 64.6  ABS design for an early hydrodynamic slider. The vertical line on each rail represents the ramp
commonly used in early sliders.

of the transducer to distinguish two adjacent magnetic bits, which is, to some extent, determined by
the strength of the magnetic signals at the read-write transducer. According to the Wallace magnetic
spacing law (Wallace, 1951), the strength of a reading signal decreases exponentially with the distance
between the magnetic layer and the transducer, known as the magnetic spacing (HMS), consisting
of the slider’s flying height or the air gap thickness, which is the distance between the slider and the
lubricant surface, the lubricant thickness, and the thickness of DLC overcoats. During the last several
decades, the lubricant thickness has been reduced from tens of nanometers to around 1 nm, while
the thickness of the DLC overcoats has been reduced to several nanometers. The most significant
reduction comes from the slider’s flying height: a reduction from several micrometers to less than
10 nm, which is achieved through increasing the rotational speed of the disk and carefully designing
the slider’s ABS. The slider’s ABS is becoming more and more complex, which brings in robustness
and reliability: it has evolved from the simple two-rail configuration in the early sliders, as shown in
Figure 64.6, to a typical modern sliders shown in Figure 64.5, and much more complicated designs
are not uncommon in state-of-the-art commercial HDDs. Designing a complex ABS requires a rapid
knowledge of the flying performance of a specific design, which is only achievable through numerical
simulations.

64.2 Historical Background: Head-Disk


Interface and Air-Bearing Sliders
The theoretical and numerical studies of a slider’s flying performance started with the design of
Winchester HDDs, which introduced the hydrodynamic self-acting sliders similar to that shown in
Figure 64.6. The flying height of earlier sliders was quite large—more than 1 μm—and the air flow in the
HDI could be described by continuum theory (Harker et al., 1981; Stevens, 1981). Since the slider’s flying
height h is significantly smaller than the slider’s length L and its width b, which are on the order of mil-
limeters, the length scale for air flow parallel to the disk is significantly larger than that perpendicular
to the disk, and continuum theory could be reduced to lubrication theory (Reynolds, 1886). In view of
the compressibility of air, Gross (1959) reduces the momentum conservation equation for the air flow,
known as the compressible Navier–Stokes equation, to

∂p ∂ 2u
=m 2 (64.1)
∂x ∂z
64-6 Design for Lubrication and Tribology

∂p ∂ 2v
=m 2 (64.2)
∂y ∂z

where
p is the pressure
μ is the air viscosity
x is the slider’s length coordinate
y is the width coordinate
z is the direction perpendicular to the disk
u and v are the components of air velocity along the x and y directions

These two equations are supplemented with the no-slip boundary conditions:

u = U D at z = 0 and u = 0 at z = h (64.3)

v = VD at z = 0 and v = 0 at z = h (64.4)

where
h is the slider’s local flying height, as shown in Figure 64.4
UD and VD are the components of the disk’s velocity along the slider’s length and width directions

Given h ≪ L and h ≪ b, the pressure difference across the HDI can be neglected, and the air velocity u and
v can be obtained from Equations 64.1 through 64.4 as functions of pressure gradients. The obtained air
velocity, when combined with the ideal gas law ρ ∝ p and an integral of the mass conservation equation
h h h h
∂r ∂ru ∂rv ∂rw
∫ ∂t
dz +
∫ ∂x
dz +
∫ ∂y
dz +
∫ ∂z
dz = 0 (64.5)
0 0 0 0

as well as the no-slip boundary conditions for w, leads to the compressible Reynolds equation:

∂ ⎡ 3 ∂P ⎤ ∂ ⎡ 3 ∂P ⎤ ∂
(64.6)

∂X ⎣
PH
∂X
− Λ X PH ⎥ +
⎦ ∂Y ⎢⎣ PH ∂Y − ΛY PH ⎥⎦ = s ∂T [ PH ]

where
X = x/L, Y = y/L, P = p/pa is the dimensionless pressure, pa is the ambient pressure
H = h/hm is the dimensionless air gap thickness, hm is a reference air gap thickness, often taken as the
minimum value
Λ X = 6mU D L/( pahm3 ) and ΛY = 6mVD L/( pahm3 ) are bearing numbers in the x and y directions
s = 12mw L2 /( pahm2 ) is the squeeze number, ω is angular frequency of the rotating disk
T = ωt is the dimensionless time, t is the time

Analytical solutions of Equation 64.6 only exist for some special cases such as the one dimensional case
with Λ X → ∞ and σ = 0 (Gross et al., 1980). Real configurations encountered in HDDs require numerical
solution of Equation 64.6, which was pioneered by Michael (1959) using the finite difference method to
solve a steady (time-independent) version of Equation 64.6. The numerically predicted load agreed with
that obtained in experiments (Brunner et al., 1959).
Magnetic Storage 64-7

The first time-dependent study was carried out by Tang (1971) who numerically solved Equation 64.6
by using the finite different discretization for both the time and space. He also used a capacitor head to
measure the dynamic response of the head flying over a wavy disk surface. The maximum spacing varia-
tion obtained from experiments agreed with numerical results.
Tang’s approach to measuring a slider’s dynamic performance required a specifically made capacitor
head and was not applicable for real heads used commercially. It is also limited by the accuracy achieved
by the capacitor. Both of these two limitations were removed by Miu et al.’s introduction of laser Doppler
interferometry (Miu et al., 1984), which is now the standard approach to experimental study of a slider’s
dynamic response.
As the slider’s flying height decreased over the years, the discrete nature of the air appears to be more
and more important. An indication of this importance is the Knudsen number Kn = λ/h, where λ is the
mean free path of air molecules defined as the average distance the air molecules can travel between two
collisions (Vincenti and Kruger, 1965). It is generally agreed that the discrete nature needs to be considered
for Kn > 0.01 (Schaaf and Chambre, 1961; Bird, 1994). The first attempt to include this effect was to intro-
duce the first-order slip boundary condition for the air velocity, which states that the difference between
the boundary velocity and the velocity of the gas adjacent to the boundary is proportional to the gradient
of the gas velocity at the boundary, for example, u − UD = β∂u/∂z at z = 0, where β is the proportional-
ity constant (Burgdorfer, 1959). Further improvement introduced more refined slip boundary conditions
such as the second-order (Hsia and Domoto, 1983) and 1.5th-order slip boundary conditions (Mitsuya,
1993). The modified Reynolds equations coming from these refinements can be cast into the same form:

∂ ⎡ ∂P ⎤ ∂ ⎡ 3 ∂P ⎤ ∂
∂X ⎢⎣
QPH 3
∂X
− Λ X PH ⎥ +
⎦ ∂Y ⎢⎣QPH ∂Y − ΛY PH ⎥⎦ = s ∂T [ PH ] (64.7)

where

Kn
Q =1+ 6 for first-order slip
PH

2
Kn Kn ⎞
Q =1+ 6 + 6 ⎛⎜ ⎟ for second-order slip
PH ⎝ PH ⎠

Numerical studies of a slider’s flying performance based on Equation 64.7 agreed with experiments
when the slider’s flying height was about several tens of nanometers (Hsia and Domoto, 1983).

64.3  Generalized Reynolds Equation


Since continuum theory, even when supplemented with slip boundary conditions, only applies when
Kn < 0.1, further decrease in the slider’s flying height required a more refined theory for describing
the air flow in the HDI. Given that the air speed in the HDI is much less than the speed of sound in
air, Fukui and Kaneko (1987, 1988) reduced the full Boltzmann equation to a linearized form of the
Boltzmann equation and, under the same observation as the continuum Reynolds equation, that is,
h ≪ L and h is less than 1 μm, derived the generalized Reynolds equation. Since the Reynolds equa-
tion is a mass conservation equation, it implies that the mass flow rate passing a cross section in the
HDI is a linearized combination of that part due to Couette flow and that part due to the Poiseuille
flow. In Equation 64.6, these two parts are calculated through continuum theory. In the generalized
Reynolds equation, the Couette flow part is still calculated by the formula obtained from the con-
tinuum theory, which turns out to be a good approximation for the rarefied gas, whereas a correction
64-8 Design for Lubrication and Tribology

factor is applied to the Poiseuille flow part. Fukui and Kaneko’s equation can also be cast into the
same form as Equation 64.7 except here Q is a function of Kn obtained from a look-up table (Fukui
and Kaneko, 1990). Kang et al. (1999) later improved Fukui and Kaneko’s model by introducing a
correction factor for the Couette flow part. However, the effect of this improvement on a slider’s fly-
ing performance is often negligible (Chen and Bogy, 2010), and, thus, Fukui and Kaneko’s original
equation is still widely used.

64.3.1  Numerical Approach


The early numerical approaches for studying a slider’s flying performance were all based on the finite dif-
ference method. Since this method solves the differential equation, Equation 64.6 or Equation 64.7, and
does not guarantee mass conservation, some unphysical phenomena could appear. Modern approaches
are all either based on the finite volume method (FVM) or finite element method (FEM) (Wahl and
Talke, 1994). The FVM solves the integral form of the differential equation, while the FEM solves the
weak form. Both of them automatically guarantee the conservation of mass and thus provide more accu-
rate numerical predictions. Lu and Bogy (Lu, 1997) and Hu and Bogy (Hu, 1996) first implemented the
FVM to study a slider’s steady and dynamic performance. Their codes evolved into the CML air-bearing
design program, which is widely used in the HDD industry and has been proven to be robust. This same
implementation was later adopted by Data Storage Institute at Singapore to develop the ABSOLUTION
code (Hua and Liu, 2006).

64.4  Slider’s Static Performance


A slider’s static performance refers to its performance at a steady state including its flying attitude
(height, pitch, and roll) and air pressure distribution on the ABS. Numerically, this corresponds to a
state where the total pressure and torque on the ABS balances those applied by the suspension. Thus,
numerical approaches usually involve numerical solution of the steady version of Equation 64.7 at a
given attitude of the slider and another scheme to determine how to change the attitude such that force
and torque balance can be achieved.
To improve an HDD’s reliability, it is desired that the strength of the magnetic signal at the read-
write transducer is the same for all locations on the disk, which requires that the slider’s flying height
at the transducer remains constant over the disk. By tuning up the ABS profile, current solvers can
reduce the variation in flying height to less than 1 nm. Figure 64.7 shows a typical pressure on a mod-
ern ABS. The pressure peak is near the transducer and serves as a high-stiffness spring helping to
stabilize the slider when it flies over bumps or rough surfaces.

20
25
15
15
(p– p0)/p0

5 10

–5 5
0.8
1
0.4 0.8
y/L 0.6 0
0.4
0 0 0.2 x/L

FIGURE 64.7  Air pressure on a modern ABS where L = 0.85 mm and p 0 = 1 atm.
Magnetic Storage 64-9

64.4.1  Intermolecular Force


As the slider’s flying height decreases, microscopic forces start to manifest themselves. Intermolecular
force, most importantly the van der Waals force, appears when the distance between two objects
becomes less than about 10 nm (Israelachvili, 1992). Wu and Bogy (2002) investigated the effect of the
van der Waals force on a slider’s static performance and found that this force is no longer negligible
when the slider’s minimum flying height is less than 10 nm.

64.4.2  Raised Temperature and Altitude


Due to the heat generated during an HDD’s operation, the temperature inside the HDD can be as high
as 100°C, which affects the physical properties of air, namely, the mean free path and the viscosity, both
of which are involved in Equation 64.7. Generally speaking, increasing the mean free path reduces the
number density of air molecules and the corresponding load capacity, leading to the decrease of the
slider’s flying height (Cha et al., 1996; Liu and Bogy, 2009), whereas increasing the viscosity increases
the slider’s flying height, as seen in numerical studies. As the temperature increases, both the mean free
path and viscosity increase leading to opposite effects on the slider’s flying height. In most cases, the
effect of the mean free path is larger, resulting in a decrease of the slider’s flying height with the tempera-
ture (Cha et al., 1996; Liu and Bogy, 2009).
Since most HDDs are not sealed, the pressure inside the HDD is the same as that outside. So when
HDDs are used in raised altitudes, as, for example, in high mountains or on airplanes, the ambient pres-
sure inside the HDD decreases, which increases the mean free path. But this induces only a small change
in the viscosity since viscosity is not a strong function of pressure, so the net effect is a decrease in the
slider’s flying height (Cha et al., 1996).

64.5  Slider’s Dynamic Performance


A slider’s dynamic performance refers to its transient response to external disturbances such as surface
roughness of the disk. Modern numerical approaches use either FVM or FEM to discretize the time-
independent part, that is, all the terms except the first one in Equation 64.7, and a finite difference
method to discretize the time.
In early numerical studies, the effect of the suspension was reduced to a constant force and torque
applied on the slider, and the dynamics of the suspension itself was neglected (Tang, 1971). More accu-
rate prediction requires the inclusion of the suspension dynamics. Due to the complex structure of the
suspension, a full analysis of the suspension dynamics based on the FEM is usually not cost-effective.
Several reduction approaches exist. The simplest one is to model the suspension as a spring-damper
system with required parameters extracted from experiments. A more refined approach is to use model
analysis to retain the first few vibration modes of the suspension so that the most important dynamics
response of the suspension can be recovered (Gupta, 1997).
Figure 64.8 shows a slider’s transient response to an external disturbance. Due to the damping of the
suspension and that induced by the air viscosity, the vibration amplitude decreases with time, and the
slider’s response finally approaches a steady state. From this transient response, the slider’s vibration
mode and other dynamics properties of the system, such as the total damping, can be extracted.

64.5.1  Contact Model


As the slider’s flying height decreases to the nanometer scale, occasional contacts between the slider
and the disk are inevitable. The first model to consider contact between two rough surfaces was that
derived by Greenwood and Williamson (1966) based on the Hertz contact theory for two elastic bodies
64-10 Design for Lubrication and Tribology

17

15

Minimum FH (mm)
13

11

7
0 0.04 0.08 0.12 0.16
Time (ms)

FIGURE 64.8  Slider’s transient response to an external disturbance.

(Johnson, 1985) and statistical theory for the surface roughness. Fuller and Tabor (1975) later intro-
duced adhesion into the Greenwood–Williamson model by using the Derjaguin–Muller–Toporov model
(Derjaguin et al., 1975), which assumes that the adhesion is only important in the contact region. More
sophisticated models introduce the disk’s and the slider’s plastic deformation as well as a more realistic
model of the surface roughness (Chang et al., 1987; Kogut and Etsion, 2002). These more refined models
require more computational resources and longer computation time, and they often do not produce
results much different from those based on the G-W model with adhesion (Chen and Bogy, 2007). Thus,
the G-W model with adhesion is still widely used to predict a slider’s dynamic performance.

64.5.2  Intermolecular Force


Using the same approach to modeling the intermolecular force in the static case, Thornton and Bogy
(2003) studied how intermolecular forces would affect a slider’s dynamic performance. They found that
the intermolecular force can cause the slider to lose its flying stability and make contact with the disk.
More accurate models of the intermolecular force, considering the layered structure of the disk, do not
change this qualitative finding (Gupta and Bogy, 2006).

64.5.3  Load/Unload
Load/unload refers to a technique used in state-of-the-art HDDs to park the slider when the HDD is
not in operation. Not-so-recent HDDs used a landing zone at the inner diameter of the disk to park the
slider in contact-start-stop (CSS) mode. In this mode when the HDD is powered on, the slider slides in
contact over the landing zone, which is rough enough to reduce the adhesion between the slider and the
disk, until air pressure is high enough to lift the slider off the disk. This approach has two limitations:
the start-up process induces wear on the slider, and the existence of a loading zone reduces the usable
region on the disk for storing information data. These limitations were overcome by the load/unload
technique, which instead of using a landing zone introduced a ramp at the outer diameter of the disk to
engage the extended tip of the suspension and park the slider when it is not in use. The ramp height is
higher than the disk. When the information data need to be transferred, the slider moves off the ramp
and is supported by the air pressure on the ABS (Suk and Albrecht, 2002).
By design, the slider is not rigidly fixed to the suspension, and its possible distance from the suspen-
sion load dimple is constrained by a spring (often referred to as the “flexure”) existing between the
suspension and the slider, as shown in Figure 64.4. When the slider is loaded onto the disk and flies
in steady state, the air pressure on the ABS pushes the slider against the flexures load dimple through
Magnetic Storage 64-11

which the suspension load is applied. However, during the load/unload process, the air pressure is ini-
tially not large enough, and the contact between the slider and the suspension transmits little or no
force. To numerically model the load/unload process, the interaction between the suspension and the
slider needs to be refined in the dynamic simulation. The simplest model is to regard the force between
the slider and the suspension as proportional to the distance between them when they are not in contact
then becoming the same as the static load case when they are in contact (Zeng et al., 1999).

64.5.4  Shock
Shock refers to a sudden high impact applied to the HDD, and it occurs, for example, when an HDD
is dropped and crashes to the ground as can happen in mobile applications. The robustness of HDDs
against shock is critical in their use in portable devices. Numerically, this process can be simulated using
the dynamic simulator with the shock modeled as an initial impulse to the slider or the suspension. The
interaction between the suspension and the slider can be modeled the same way as in the load/unload
case, but the suspension needs to be modeled more accurately since high impact usually involves large
deformation of the suspension due to possible strong contact between the slider and the disk (Bhargava
and Bogy, 2007). In order to get more accurate and reliable results for shocks, the response of other
structural parts of the HDD must also be modeled, such as the disk, spindle motor, and its mounting
structure.

64.6  Thermal Flying-Height Control Sliders


An increase in an HDD’s capacity always requires the reduction in the slider’s flying height at the trans-
ducer. This is achieved traditionally by designing a different ABS, which is cumbersome and time-
consuming. To partly overcome this difficulty, the HDD industry recently introduced the thermal
flying-height control technique (Meyer et al., 1999; Kurita and Suzuki, 2004; Liu et al., in press). In
this technique, a heater element is embedded in the slider near the transducer, as shown in Figure 64.9.
During the HDD’s operation, power is applied to the heater element, and, due to thermomechanical
coupling, the slider protrudes near the heater element and the transducer so that the physical spacing at
the transducer gets reduced.
The complex structure of the heater element and the read-write transducer prohibits analytical stud-
ies, and instead numerical approaches are widely used and have been proven robust (Juang et al., 2006;
Juang and Bogy, 2007). This approach iterates between two steps: in one step, the air pressure in the HDI
is solved using FVM or FEM method for a given geometry of the slider, and, in the other step, the slider’s
deformation is obtained from the FEM method given the air pressure on the ABS and the slider’s flying
attitude (Zheng et al., in press). The process continues until some convergence criterion is met such as

Susp
ensi
on Heater
Shield

Slide
Lubricant r
layer Heater
h

Disk
U
Reader Writer
transducer transducer

FIGURE 64.9  Sketch of a thermal flying-height control slider.


64-12 Design for Lubrication and Tribology

20

Sea level

Protrusion amount (nm)


2 km
5 km
15

10 Read transducer

5
–50 –40 –30 –20 –10 0 10 20
Distance from the AlTiC trailing edge (µm)

FIGURE 64.10  Deformed profile of a thermal flying-height control slider in operation. (From Zheng, J. et al.,
Tribol. Lett., 40, 295, 2010. With permission.)

the average difference between the slider’s flying heights obtained in two adjacent steps is less than a
given value.
Due to the power applied to the heater element, the temperature on the ABS is not the same as that of
the slider, and, strictly speaking, the air flow in the HDI cannot be described by the isothermal Reynolds
equation. Instead, the full Boltzmann equation needs to be solved. However, due to the smallness of the
temperature difference when compared to the temperature of the ABS or the disk, the thermal problem
can be decoupled from the air flow problem, and the interaction between them is one-way: the air flow
part affects the thermal part but not vice versa (Sone, 2006).
In the numerical approach, the air flow affects the heat transfer through the model for calculating the
heat flux from the ABS. Given the smallness of the temperature gradient, this flux can be divided into
three parts: heat conduction, viscous dissipation, and radiation. The viscous dissipation is induced by
the internal friction of the air, and the radiation is induced by the different temperatures of the slider
and the disk. These two contributions have been shown to be negligible when compared to the heat con-
duction (Ju, 2000; Chen et al., 2009; Liu et al., 2010; Zheng et al., 2010).
The current model for calculating the heat conduction was originally derived from the first-order slip
theory (Zhang and Bogy, 1999; Chen et al., 2000), and it has been shown to be accurate enough for cur-
rent systems (Liu et al., 2010).
Figure 64.10 shows a zoomed-in profile of a TFC slider’s transducer area in operation. The slider’s
deformation is confined in a region of tens of micrometers (Zheng et al., in press). It is seen that the
profile is not very sensitive to elevated altitude.

64.7  Outlook
To compete with solid-state drives (SSDs) with ever increasing performance, the HDDs need to have a
better reliability and larger capacity. Several approaches that can contribute to these improvements are
discussed in this section.

64.7.1  Helium-Filled Hard Disk Drives


Modern HDDs are not sealed and are open to ambient air. It has been often proposed to seal the HDD
with a helium atmosphere. Most of the advantages of filling the HDDs with helium come from helium’s
low density and high thermal conductivity when compared to air. Helium’s low density reduces power
consumption of HDDs, part of which is positively related to the gas density. Since the helium’s viscosity
Magnetic Storage 64-13

is comparable to that of air, helium’s low density also suppresses the turbulence in HDDs, which reduces
nonrepeatable vibration of the slider and increases the position accuracy of the read-write transducer.
Helium’s high thermal conductivity helps to dissipate heat generated during the HDD’s operation and
reduces the maximum temperature rise, which, when combined with the inert property of helium, helps
to protect the disk from corrosion (Sato et al., 1988; Aruga et al., 2007). The difficulty with this approach
lies in the sealing of the HDDs. However, with the progress in manufacturing, this approach seems to
be more achievable than before (Liu et al., 2009).
Numerical investigations of the flying performance of a slider in helium follow the same approach as
that in air with the only difference being that the physical properties of helium are used. The aforemen-
tioned advantages have all been confirmed by experiments or numerical studies (Sato et al., 1988; Aruga
et al., 2007). Liu et al. (2009) recently extended this investigation to a TFC slider flying in air–helium
gas mixtures by using established approaches to calculate the physical properties of gas mixtures. They
found that, for a specific commercial ABS design, slightly less power is required for a given flying-height
reduction when the mixture is composed of about 60% helium and 40% air.

64.7.2  Bit-Patterned Media


Increasing the HDD’s capacity was traditionally achieved through squeezing the size of the magnetic bit,
which usually consists of several magnetic grains. Accordingly, the grain size also needs to be reduced.
Given such a small size of the current magnetic grain, further reduction makes the grain thermally
unstable: the thermal fluctuation will reverse the magnetic field of the bit and induce loss of information
stored on the disk (Bandic and Victora, 2008). Two approaches exist to overcome this difficulty: using
more magnetically stable materials and using the current materials but keeping the size of the magnetic
grain. Patterned media is the way to realize the second approach and the HAMR is to realize the first
one (Bandic and Victora, 2008).
In patterned media, the magnetic grains are deposited on the pattern on the disk, and each grain cor-
responds to one magnetic bit, in contrast to several grains forming one bit in traditional recording. The
difficulty in realizing this approach lies in the mass production of patterns on the disk, which requires
more sophisticated facility than that currently used in the semiconductor industry. Nanoimprint seems
to be the solution, but it still needs to be improved to meet the requirement of the HDD industry (Dobisz
et al., 2008).
Duwensee et al. (2009) investigated a slider flying over a discreet track media using the direct simu-
lation Monte Carlo (DSMC) method, and they concluded that the generalized Reynolds equation still
applies to the air flow between a slider and a patterned disk. Li et al. (2009) investigated the static per-
formance of a slider flying over a patterned disk, as shown in Figure 64.11, by numerically solving the
generalized Reynolds equation with the FEM. They showed that increasing the pattern height leads to
larger change in the slider’s flying height when compared to that for a slider flying on a smooth disk.
Given the small size of the pattern, which is about 20 nm, Li et al.’s approach requires substantial com-
putation resources. Since the pattern height is small compared to the air-bearing thickness for most
regions in the HDI, the pattern can be viewed as a small perturbation to the air bearing, and the air flow
can be approximately described by the generalized Reynolds equation for a smooth disk supplemented
with averaged contribution from the perturbation due to the existence of patterns on the disk. This

Slid
er

Disk

FIGURE 64.11  Slider flies over a patterned disk. Note the figure is not to scale.
64-14 Design for Lubrication and Tribology

Slid
er

Lubricant film
Disk

FIGURE 64.12  Lube-surfing recording.

approach seems to be able to capture the main features of the slider’s static and dynamic performances
(Gupta and Bogy, 2006).

64.7.3  Heat-Assisted Magnetic Recording


Information data are more difficult to be written on a disk with magnetic materials with a high magnetic
stability (high coercivity), which can be overcome by raising the temperature of the material to a certain
value, known as the Curie temperature. HAMR achieves this through locally heating the disk with a
laser (Kryder et al., 2008). The main difficulty with this approach lies in the integration of the laser into
the system. Given that much progress has been accomplished (Challener et al., 2009; Stipe et al., 2010),
this approach may be more achievable than the pattern media.
Since the laser spot has a diameter on the order of tens of nanometers, preliminary results showed
that the generalized Reynolds equation is still capable of describing air flow in the HDI, whose accu-
racy, however, still needs to be investigated through experiments. The most critical tribology issue in
this approach lies in the lubricant. The currently used lubricant becomes thermally decomposed under
laser heating and needs to be replaced by some alternatives (Kryder et al., 2008). Wu (2007) numeri-
cally investigated the deformation and instability of the lubricant film in HARM by including lubricant
evaporation, thermocapillary effect, and thermoviscosity. He argued that new lubricants with low vola-
tility and small surface tension are required for the HARM system. More numerical and experimental
studies are needed for clarifying or confirming these findings.

64.7.4  Lube-Surfing Recording


Increasing HDDs’ capacity requires the decrease in the magnetic spacing. To achieve an areal density
of 10 Tbit/in.2, the magnetic spacing needs to be reduced to less than 3 nm. One approach to achieve
this is to make the protrusion of a TFC slider come immersed in the lubricant film but not in full
contact with the disk, as shown in Figure 64.12. Since only a small region of the ABS is immersed in
the lubricant film, the ABS is still supported by the air bearing on most regions of the ABS, and the
air flow can still be approximately described by the generalized Reynolds equation (Liu et al., 2009).
But capillary force and adhesion force between the protrusion and the lubricant film as well as the
disk need to be included in the analysis. Zheng and Bogy (2010) used an adhesion model considering
the adhesive contact between the slider and the disk in the presence of a lubricant film (Stanley et al.,
1990) to investigate the instability of the slider’s flying attitude. They found that the slider’s static flying
becomes unstable when it becomes immersed in the lubricant film. Experiments are needed to confirm
this finding.

64.8  Conclusion
HDDs have been serving as the dominant magnetic recording device for over a half century and are
predicted to be so in the near future (Burr et al., 2008). This is made possible partly by the continuing
Magnetic Storage 64-15

increase in the HDDs’ capacity, which requires a reduction in the slider’s flying height at the transducer,
which now approaches 5 nm. Numerical approaches for studying the slider’s performance provide indis-
pensable tools for designing more robust sliders with an ever deceasing flying height at the transducer.
By incorporating the necessary physics, such as a realistic contact model and a heat transfer model, the
numerical approach has been extended to handle a slider with a flying height at the transducer less than
10 nm, which is much beyond the expectation a few years ago. The authors have confidence that the
numerical approaches will continue to serve the HDD industry provided improvements in the physical
models are appropriately incorporated when needed.

Acknowledgments
The authors thank the Computer Mechanics Laboratory (CML) at the Department of Mechanical
Engineering in University of California, Berkeley, for supporting this work. The authors also thank
Ms. Jinglin Zheng at CML for providing Figures 64.8 and 64.10.

References
Aruga, K., M. Suwa, K. Shimizu, and T. Watanabe. 2007. A study on positioning error caused by flow
induced vibration using helium-filled hard disk drives. IEEE Trans. Magn. 43: 3750–3755.
Bandic, Z. Z. and R. H. Victora. 2008. Advances in magnetic data storage technologies. Proc. IEEE 96:
1749–1753.
Bhargava, P. and D. B. Bogy. 2007. Numerical simulation of operational-shock in small form factor hard
disk drives. J. Tribol. 129: 153–160.
Bird, G. A. 1994. Molecular Gas Dynamics and the Direct Simulation of Gas Flows. New York: Oxford
University Press.
Brunner, B. K., J. M. Harker, K. E. Houghton, and A. G. Osterlund. 1959. A gas film lubrication study part
III experimental investigation of pivoted slider bearings. IBM J. Res. Dev. 3: 260–274.
Burgdorfer, A. 1959. The influence of the molecular mean free path on the performance of hydrodynamic
gas lubricated bearing. J. Basic Eng. 81: 94–100.
Burr, G. W., B. N. Kurdi, J. C. Scott, C. H. Lam, K. Gopalakrishnan, and R. S. Shenoy. 2008. Overview of
candidate device technologies for storage-class memory. IBM J. Res. Dev. 52: 449–464.
Cha, E., C. Chiang, J. Enguero, and J. J. K. Lee. 1996. Effect of temperature and attitude on flying height.
IEEE Trans. Magn. 32: 3729–3731.
Challener, W. A. et al. 2009. Heat-assisted magnetic recording by a near-field transducer with efficient opti-
cal energy transfer. Nat. Photonics 3: 220–224.
Chang, W. R., I. Etsion, and D. B. Bogy. 1987. An elastic-plastic model for the contact of rough surfaces.
J. Tribol. 109: 257–263.
Chen, D. and D. B. Bogy. 2007. Intermolecular force and surface roughness models for air bearing simula-
tions for sub-5 nm flying height sliders. Microsyst. Technol. 13: 1211–1217.
Chen, D. and D. B. Bogy. 2010. Comparisons of slip-corrected Reynolds lubrication equations for the air
bearing film in the head-disk interface of hard disk drives. Tribol. Lett. 37: 191–201.
Chen, D., N. Liu, and D. B. Bogy. 2009. A phenomenological heat transfer model for the molecular gas
lubrication system in hard disk drives. J. Appl. Phys. 105: 084303.
Chen, L., D. B. Bogy, and B. Strom. 2000. Thermal dependence of MR signal on slider flying state. IEEE
Trans. Magn. 36: 2486–2489.
Derjaguin, B. V., V. M. Muller, and Y. P. Toporov. 1975. Effect of contact deformations on the adhesion of
particles. J. Colloid Interface Sci. 53: 314–326.
Dobisz, E. A., Z. Z. Bandic, T.-W. Wu, and T. Albrecht. 2008. Patterned media: Nanofabrication challenges
of future disk drives. Proc. IEEE 96: 1836–1846.
64-16 Design for Lubrication and Tribology

Duwensee, M., F. E. Talke, S. Suzuki, J. Lin, and D. Wachenschwanz. 2009. Direct simulation Monte Carlo
method for the simulation of rarefied gas flow in discrete track recording head/disk interfaces.
J. Tribol. 131: 012001.
Fukui, S. and R. Kaneko. 1987. Analysis of ultra-thin gas film lubrication based on linearized Boltzmann
equation (influence of accommodation coefficient). JSME Int. J. 30: 1660–1666.
Fukui, S. and R. Kaneko. 1988. Analysis of ultra-thin gas film lubrication based on linearized Boltzmann
equation: First report–derivation of a generalized lubrication equation including thermal creep flow.
J. Tribol. 110: 253–261.
Fukui, S. and R. Kaneko. 1990. A database for interpolation of Poiseuille flow-rates for high Knudsen num-
ber lubrication problems. J. Tribol. 112: 78–83.
Fuller, K. N. G. and D. Tabor. 1975. The effect of surface roughness on the adhesion of elastic solids. Proc.
R. Soc. Lond., Ser. A 345: 327–342.
Greenwood, J. A. and J. B. P. Williamson. 1966. Contact of nominally flat surfaces. Proc. R. Soc. Lond., Ser.
A 295: 300–319.
Grochowski, E. and R. D. Halem. 2003. Technological impact of magnetic hard disk drives on storage sys-
tems. IBM Systems J. 42: 338–346.
Gross, W. A. 1959. A gas film lubrication study part I some theoretical analyses of slider bearings. IBM
J. Res. Dev. 3: 237–255.
Gross, W. A. et al. 1980. Fluid Film Lubrication. New York: Wiley.
Gupta, V. 1997. Air bearing slider dynamics and stability in hard disk drives. PhD thesis, Department of
Mechanical Engineering, University of California, Berkeley, CA.
Gupta, V. and D. B. Bogy. 2006. Effect of intermolecular forces on the static and dynamic performance of
air bearing sliders. J. Tribol. 128: 197–208.
Harker, J. M., D. W. Brede, R. E. Pattison, G. R. Santana, and L. G. Taft. 1981. A quarter century of disk file
innovation. IBM J. Res. Dev. 25: 677–689.
Hsia, Y. T. and G. A. Domoto. 1983. An experimental investigation of molecular rarefaction effects in gas
lubricated bearings at ultra-low clearances. J. Lubr. Technol. 105: 120–130.
Hu, Y. 1996. Head-disk-suspension dynamics. PhD thesis, Department of Mechanical Engineering,
University of California, Berkeley, CA.
Hua, W. and B. Liu. 2006. Mechanism studies of the multiple flying states of the air bearing slider. Tribol.
Int. 39: 649–656.
Israelachvili, J. N. 1992. Intermolecular and Surface Forces, 2nd edn. New York: Academic Press.
Johnson, K. L. 1985. Contact Mechanics. New York: Cambridge University Press.
Ju, Y. S. 2000. Thermal conduction and viscous heating in microscale Couette flows. J. Heat Transfer 122:
817–818.
Juang, J. Y. and D. B. Bogy. 2007. Air-bearing effects on actuated thermal pole-tip protrusion for hard disk
drives. J. Tribol. 129: 570–578.
Juang, J. Y., D. Chen, and D. B. Bogy. 2006. Alternate air bearing slider designs for areal density of 1 Tb/in2.
IEEE Trans. Magn. 42: 241–246.
Kang, S. C., R. M. Crone, and M. S. Jhon. 1999. A new molecular gas lubrication theory suitable for head-
disk interface modeling. J. Appl. Phys. 85: 5594–5596.
Kogut, L. and I. Etsion. 2002. Elastic-plastic contact analysis of a sphere and a rigid flat. J. Appl. Mech. 69:
657–662.
Kryder, M. H. et al. 2008. Heat assisted magnetic recording. Proc. IEEE 96: 1810–1835.
Kurita, M. and K. Suzuki. 2004. Flying-height adjustment technologies of magnetic head sliders. IEEE
Trans. Magn. 40: 332–336.
Li, H., H. Zheng, Y. Yoon, and F. E. Talke. 2009. Air bearing simulation for bit patterned media. Tribol. Lett.
33: 199–204.
Liu, B. et al. 2009. Lube-surfing recording and its feasibility exploration. IEEE Trans. Magn. 45: 899–904.
Magnetic Storage 64-17

Liu, N. and D. B. Bogy. 2009. Temperature effect on a HDD slider’s flying performance. Tribol. Lett. 35:
105–112.
Liu, N., J. Zheng, and D. B. Bogy. 2009. Thermal flying-height control sliders in hard disk drives filled with
air-helium gas mixtures. Appl. Phys. Lett. 21: 213505.
Liu, N., J. Zheng, and D. B. Bogy. 2010. Predicting the flying performance of thermal flying-height control
sliders in hard disk drives. J. Appl. Phys. 108: 016102.
Liu, N., J. Zheng, and D. B. Bogy. 2011. Thermo-mechanical aspect of thermal flying-height control sliders
for hard disk drives. Math. Mech. solids. 16: 694–705.
Lu, S. 1997. Numerical simulation of slider air bearings. PhD thesis, Department of Mechanical Engineering,
University of California, Berkeley, CA.
Meyer, D. W., P. E. Kupinski, and J. C. Liu. 1999. Slider with temperature responsive transducer positioning.
U.S. Patent No. 5,991,113.
Michael, W. A. 1959. A gas film lubrication study part II numerical solution of the Reynolds equation for
finite slide bearings. IBM J. Res. Dev. 3: 256–259.
Mitsuya, Y. 1993. Modified Reynolds equation for ultra-thin film gas lubrication using 1.5-order slip-flow
model and considering surface accommodation coefficient. J. Tribol. 115: 289–294.
Miu, D. K., G. Bouchard, D. B. Bogy, and F. E. Talke. 1984. Dynamic response of a Winchester-type slider
measured by laser Doppler interferometry. IEEE Trans. Magn. 20: 927–929.
Reynolds, O. 1886. On the theory of lubrication and its application to Mr. Beauchamp Tower’s experi-
ments, including an experimental determination of the viscosity of olive oil. Philos. Trans. 177:
157–234.
Sato, I., K. Otani, S. Oguchi, and K. Hoshiya. 1988. Characteristics of heat transfer in a helium-filled disk
enclosure. IEEE Trans. Compon., Hybrids, Manuf. Technol. 11: 571–575.
Schaaf, S. A. and P. L. Chambre. 1961. Flow of Rarefied Gases. Princeton, NJ: Princeton University Press.
Shiroishi, Y., K. Fukuda, I. Tagawa, H. Iwasaki, S. Takenoiri, H. Tanaka, H. Mutoh, and N. Yoshikawa. 2009.
Future options for HDD storage. IEEE Trans. Magn. 45: 3816–3822.
Sone, Y. 2006. Molecular Gas Dynamics: Theory, Techniques, and Applications. Boston, MA: Birkhauser.
Stanley, H. M., I. Etsion, and D. B. Bogy. 1990. Adhesion of contacting rough surfaces in the presence of
sub-boundary lubrication. J. Tribol. 112: 98–104.
Stevens, L. D. 1981. The evolution of magnetic storage. IBM J. Res. Dev. 25: 663–675.
Stipe, B. C. et al. 2010. Magnetic recording at 1.5Pb m−2 using an integrated plasmonic antenna. Nat.
Photonics 4: 484–488.
Suk, M. and T. R. Albrecht. 2002. The evolution of load/unload technology. Microsyst. Technol. 8: 10–16.
Tang, T. 1971. Dynamics of air-lubricated slider bearings for noncontact magnetic recording. ASME
J. Lubr. Technol. 93: 272–278.
Thornton, B. H. and D. B. Bogy. 2003. Head-disk interface dynamic instability due to intermolecular forces.
IEEE Trans. Magn. 39: 2420–2422.
Vincenti, W. G. and C. H. Kruger. 1965. Introduction to Physical Gas Dynamics. New York: Wiley.
Wahl, M. H. and F. E. Talke. 1994. Numerical simulation of the steady-state flying characteristics of a
50-percent slider with surface texture. IEEE Trans. Magn. 30: 4122–4124.
Wallace, R. L. 1951. The reproduction of magnetically recorded signals. Bell Syst. Tech. J. 30: 1145–1173.
Wikipedia. 2010. http://en.wikipedia.org/wiki/Early IBM disk storage#IBM 350
Wood, R. 2009. Future hard disk drive systems. J. Magn. Magn. Mater. 321: 555–561.
Wu, L. 2007. Modelling and simulation of the lubricant depletion process induced by laser heating in heat-
assisted magnetic recording system. Nanotechnology 18: 215702.
Wu, L. and D. B. Bogy. 2002. Effect of the intermolecular forces on the flying attitude of sub-5 nm flying
height air bearing sliders in hard disk drives. J. Tribol. 124: 562–567.
Zeng, Q. H., M. Chapin, and D. B. Bogy. 1999. Dynamics of the unload process for negative pressure sliders.
IEEE Trans. Magn. 35: 916–920.
64-18 Design for Lubrication and Tribology

Zhang, S. and D. B. Bogy. 1999. A heat transfer model for thermal fluctuation in a thin slider/disk air bear-
ing. Int. J. Heat Mass Transfer 42: 1791–1800.
Zheng, H. et al. 2010. The effect of thermal radiation on thermal flying-height control sliders. IEEE Trans.
Magn. 46: 2376–2378.
Zheng, J. and D. B. Bogy. 2010. Investigation of flying-height stability of thermal flying-height control
sliders in lubricant or solid contact with roughness. Tribol. Lett. 38: 283–289.
Zheng, J., D. B. Bogy, S. Zhang, and W. Yan. 2010. Effect of altitude on thermal flying-height control
actuation. Tribol. Lett. 40: 295–299.
65
Diagnostics
65.1 Introduction..................................................................................... 65-1
65.2 Asset Management.......................................................................... 65-1
65.3 Nondestructive Evaluation.............................................................65-2
Visual Inspection  •  Liquid Penetrant Inspection  •  Magnetic
Particle Inspection  •  Eddy Current Inspection  •  Acoustic Emission
Detection  •  Ultrasonic Inspection  •  Radiographic Inspection
65.4 Condition Monitoring....................................................................65-5
Vibration Monitoring  •  Oil Monitoring  •  Thermal
Monitoring  •  Performance Monitoring
Richard S. Cowan 65.5 Tribosystem Applications.............................................................65-12
Georgia Institute Hydraulic Systems  •  Fluid Film Bearings
of Technology References................................................................................................... 65-15

65.1  Introduction
Just as medical patients would be less than satisfied if their doctors needed to open their bodies for
routine health checks, the opening of a machine to evaluate if it is performing satisfactorily can often
be much more harmful than good. Accordingly, it makes sense to use diagnostic techniques to assess
the condition of a tribosystem in just the same way as a medical doctor uses symptoms to assess the
condition of humans. The purpose of this chapter is to present technologies for machinery diagnosis and
suggest how they may be readily applied.

65.2  Asset Management


When something fails, the limitations of a design can be quickly understood. While success enables one
to see the possibilities for reducing conservatism by subjecting machinery to more demands, it is only
in failure that boundaries are defined and much can be learned to increase safety and reliability and
decrease manufacturing and operating costs. By probing failures to their root cause, the results should
ultimately impact some aspect of manufacture and or use.
In planning for the possibility of failure, one may wish to compare a set of measurements for a given
product to a set of specifications. The designer or the customer determines these specifications, usu-
ally called tolerance limits. Mathematically, tolerance limits are a set of bounds between which one
can expect to find any given proportion of a population. If one knows or assumes that a parameter is
normally distributed with mean and variance, then tolerance limits can be constructed at any level of
confidence for a proportion of the population using the tables of the cumulative normal distribution.
Measurements outside the bounds would likely be symptomatic of a problem.
Theoretically, all processes can be characterized by a certain amount of variation if measured with
an instrument of sufficient resolution. The key task in the application of such statistical control [1] is in

65-1
65-2 Design for Lubrication and Tribology

determining the appropriate variables on which to apply the control, obtaining sufficient data on which
to base a decision, and establishing a system of maintenance to ensure that an action can be taken when
the data suggest that a problem is imminent.
Approaches taken to maintain systems have evolved over time. Given product demand, historical
behavior, and management attitude, the maintenance plans of today will likely be unique to the sys-
tem of interest, incorporating one or more philosophies based on failure mode, timing, reliability, or
condition.
Failure-based maintenance requires little, if any, advanced planning. It typically takes the form of
waiting until a failure occurs, at which time the reaction is to repair the damage. Time-based main-
tenance evolved due to the economic impact of failure-based maintenance. It is a system of address-
ing upkeep at fixed periods of time, independent of the condition of the equipment. It is very effective
when the performance and condition of equipment are related to the passage of time and where the
maintenance tasks can be carried out efficiently, such as an oil change. Condition-based maintenance is
safer and economically more attractive than either the failure-based or time-based methods. Evaluation
based on condition requires management support, as resources must be allocated for the devices used
to obtain information about the behavior of the system or structure, as well as establish documentation
and historical information for comparative use. Trained personnel are required to implement and sus-
tain the effort. In an effectively run program, the savings incurred from reduced maintenance should
offset these costs. Success or failure is dependent on good knowledge of the condition of the item or
system of interest, generally obtained via nondestructive evaluations (NDEs) and/or condition monitor-
ing approaches.

65.3  Nondestructive Evaluation


NDE is an interdisciplinary field of study that is concerned with the development of analysis and mea-
surement technologies [2,3] for the characterization of materials by noninvasive means. A variety of
inspection and detection techniques are available to provide users with the speed, accuracy, and cost
efficiency needed to probe, identify, and diagnose features of importance in evaluating tribosystem
performance.

65.3.1  Visual Inspection


When performed with human sight, visual inspection is relatively inexpensive and has no need for
special equipment. Good eyesight, good lighting, and the knowledge of what to look for are required.
Visual inspection can be enhanced by various methods ranging from the use of magnifying glasses to
optical probes or boroscopes.
Preparation can range from wiping the surface of interest with a cloth to blast cleaning and treating
with chemicals to show detail. To aid in identifying defects that are not easily tracked by the eye, replica-
tion techniques and other surface treatment methods often enhance the inspection process. In response
to situations where items need to be monitored quickly or with tolerances too small to be analyzed by
the human eye, machine vision can be used. A machine vision system can inspect objects accurately
even within complex or confusing scenes.
In a typical machine vision application, items undergoing inspection are positioned under proper
lighting in front of one or more video cameras. A lens forms an image of the item on the camera sen-
sor, which generates an analog signal that is subsequently digitized into pixels that represent light
intensity at points on the item and collectively form an image of the item. The image is processed by
specialized digital computers to select or amplify key features, or convert the image into measure-
ments, such as size and location. From these measurements, the machine vision system can qualify,
accept, or reject an item based on decision thresholds or classifications established by the operator
through experience.
Diagnostics 65-3

65.3.2  Liquid Penetrant Inspection


Liquid penetrant inspection is a method that is used to reveal a surface-breaking flaw by the release
of a colored or fluorescent dye. The surface of the part under evaluation is coated with a penetrant
in which a visible or fluorescent dye is dissolved or suspended. The penetrant is pulled into surface
defects by capillary action. After a waiting period, the excess penetrant is cleaned from the surface of
the sample. A white powder or developer is then sprayed or dusted over the part. The developer lifts
the penetrant out of the defect, and the dye stains the developer. By visual inspection under white
or ultraviolet light, the visible or fluorescent dye indications, respectively, are identified and located,
thereby defining the defect.
Penetrant inspection can be used on any material with a relatively smooth, nonporous surface and
usually takes less than an hour to accomplish for most parts. It is essential that the material is carefully
cleaned; otherwise, the penetrant will not be able to get into the defect. If the surface penetrant is not
fully removed, misleading indications will result.

65.3.3  Magnetic Particle Inspection


Magnetic particle inspection is a method that can be used to find surface and near-surface flaws in ferro-
magnetic materials. The technique uses the principle that magnetic lines of force or flux will be distorted
by the presence of a flaw in a manner that will reveal its presence.
Upon magnetizing the item of interest, iron particles are dusted over it, or flowed over it if they are
suspended in a fluid such as kerosene. A surface defect will form a magnetic anomaly, which attracts and
holds the particles, giving a visual indication of a defect. An inspector views the object being evaluated
and makes a decision of acceptance or rejection. As surface irregularities and scratches can give mis-
leading indications, it is necessary to ensure careful preparation of the surface before magnetic particle
inspection is undertaken. Under optimum conditions, the probability of detecting a flaw as small as
2 mm is extremely likely.
The evaluation time is typically a few minutes for relatively small parts, 10 kg or less. Large parts
must be done in sections, extending evaluation time. As the part must be ferromagnetic, this technique
cannot be used on most stainless steels. While magnetic particle inspection is primarily used to find
surface-breaking flaws, it can also be used to locate subsurface flaws. Its effectiveness, however, dimin-
ishes depending on the flaw depth and type.

65.3.4  Eddy Current Inspection


Eddy current inspection is an electromagnetic approach to flaw detection that can only be used on con-
ductive materials. A coil, bearing an alternating current (typically 10 Hz–10 MHz), creates an alternat-
ing magnetic field that is used to induce circulating electric currents (eddy currents) in the component
to be inspected. These currents flow in a thin skin beneath the surface adjacent to the coil. The thickness
of this skin, between 5 μm and 10 mm, is dependent on the shape, size, and operating frequency of the
coil, the proximity of the coil to the component, and the conductivity of the targeted material. As the
coil scans a region of interest, the impedance in the coil is altered when the eddy currents are distorted
by the presence of defects or material variations. This change is measured and displayed in a manner
that indicates the type of flaw or condition of the material.
Eddy current evaluations can be made at different material depths using coils of differing sizes at a
range of frequencies. Given eddy currents are established through induction, there is no need for the
coil to contact the component. Therefore, the surface can be painted or coated. Depending on surface
condition, it is usually possible to find cracks as small as 0.1 mm deep. In automated applications, where
scanning speeds can be as high as a meter per second, best results are obtained when the scan direction
is normal to flaw orientation.
65-4 Design for Lubrication and Tribology

65.3.5  Acoustic Emission Detection


Acoustic emissions (AEs) are stress waves produced by sudden changes in the internal structure of a
material. Possible causes are crack initiation and growth, crack opening and closure, dislocation move-
ment, and material phase transformation. As most of the sources of AEs are damage related, the detec-
tion and monitoring of these emissions are commonly used to predict material failure.
Wideband transducers (typically 50 kHz–1 MHz) are normally used to detect the random bursts of
deformation, characteristic of an AE event. Made of a thin film of piezoelectric material, such as lead zir-
conate titanate (PZT), these sensitive devices emit an electric charge proportional to the applied stress.
As AE signals are generally very weak, a charge amplifier is connected to the AE transducer to minimize
noise interference and prevent signal loss. The amplified signal, which is in analog form, can then be
sent via coax cable to a filter for noise removal and routed to a signal conditioner and/or computer for
further analysis.
AE signals may be recorded continuously by one or more relatively small sensors mounted on the sur-
face of the structure being examined for progressive damage. When the AE transducer senses a signal
over a certain threshold, an AE event is captured. The number of times the signal rises and crosses the
threshold is the count of the AE event. The time period between the rising edge of the first count and the
falling edge of the last count is the duration of the AE event. The time period between the rising edge of
the first count and the peak of the AE event is called the rise time. The area under the envelope of the AE
event is the energy. These features and others are correlated with defect formation and failure.
The advantages of using AEs for assessing the health of a tribosystem include the following: (1) the
entire system can be monitored from several locations, (2) the system can be tested in use (without tak-
ing it out of service), (3) microscopic changes can be detected if sufficient energy is released, (4) source
location can be estimated using multiple sensors, and (5) continuous monitoring with alarms is possible.
The main disadvantages are that AE systems can only estimate qualitatively how much damage is in the
material and approximately how long the system will last. Hence, other NDE methods are still needed
to do more thorough examinations.

65.3.6  Ultrasonic Inspection


Ultrasonic inspection is an approach that uses high-frequency waves to evaluate the quality of a tribo-
logical material. Pulses of ultrasonic energy, commonly emitted via a piezoelectric transducer placed on
the object of interest, penetrate the material and are subsequently altered as they travel through it due to
attenuation, reflection, and scattering. The resulting output pulse is detected, processed, and interpreted
based on its relation to the input pulse. This can be accomplished through a pulse-echo arrangement,
where one transducer is used to emit and receive ultrasound, or through a pitch-catch mode, where one
or more transducers are strategically placed to catch the output pulse.
For potentially hazardous and hostile environments, laser-based technology can be used for gen-
erating the ultrasonic signal. A laser beam, targeted at the material of interest some meters away, can
produce broadband ultrasonic vibrations up to frequencies of about 100 MHz without damaging the
surface. These vibrations act as sources of compression, shear, and Rayleigh surface waves that pass
through the object of interest. Noncontact detection of these waves can be realized using interferometry
to monitor the induced surface for resulting effects. Hence, no contact medium is required for ultrasonic
inspection.
Ultrasonic inspection can be applied to metals as well as carbon composites and ceramics. It is most
often used to search for flaws within materials, for example, cracks, voids, porosity, and delamination.
As the ultrasonic wave penetrates the object, defects will reflect the signal in a characteristic manner,
which can be interpreted by observing the amplitude of the received pulse and the time taken to arrive.
Whenever the configuration of the object under examination permits, a two- or three-dimensional
image of the interior of the object can be made, showing reflections of the ultrasound for use in locating
Diagnostics 65-5

regions of concern. Flaw-size resolution (typically 1 mm in length) is dependent on both the speed of
sound in the material and the frequency of the ultrasonic wave.
Because this approach to NDE is complex, considerable training and skill are required. Best results
are obtained if the object being evaluated is flat or has spherical or cylindrical symmetry. As ultrasonic
waves can attenuate rapidly in certain materials, pulse-echo measurements on thick (more than a few
inches) parts can be difficult to perform. A high-resolution scan of large parts and scans of parts with
odd shapes or convoluted surfaces can take many hours. Setup time typically takes less than an hour.

65.3.7  Radiographic Inspection


Radiographic inspection uses radiation as the source for identifying abnormal characteristics within a
material. A source of radiation (x-ray or gamma ray) is directed toward the part under evaluation with a
sheet of radiographic film placed behind it. This film is then processed, and a semitransparent image of
the object is obtained as a series of shades between black and white. The density of the image that results
is a function of the quantity of radiation transmitted through the object, which in turn is inversely pro-
portional to its atomic number, density, and thickness.
The image can reveal internal defects such as voids, cracks, or inclusions in the material. It can also
expose internal clearances between parts in an assembly and any displacement of internal components.
Laminations, tight cracks, or cracks parallel to the film plane are not readily discernible. The method is lim-
ited by material thickness to approximately the equivalent of 0.1 m of lead, 0.3 m of steel, or 0.6 m of concrete.
Special precautions must be taken when performing radiography. The operator must use a protective
enclosure or appropriate barrier and provide warning signals to ensure there are no hazards. Setup typi-
cally takes a few minutes, the exposure typically 1–10 min, and film processing up to 10 min.

65.4  Condition Monitoring


Any engineering system will fail to operate satisfactorily at some time in its useful life, induced by many,
potentially unrelated, factors. As a result, there are a variety of sensors and approaches [4,5] commonly
used to assess the behavior of a tribosystem, as noted in this section.

65.4.1  Vibration Monitoring


Vibration is defined as a periodic motion about an equilibrium position. Its duration and magnitude
depend upon the degree of damping the effected materials possess and the phase relationships between
the mechanism that perturbs the system and the response that is obtained. Vibration may be forced
through unbalance, rub, looseness, and misalignment or freely self-excited through internal friction,
cracking, and resonance.
Once generated, vibration can be transmitted from its source to other components or systems. When
it reaches unacceptable levels, material wear and tear processes are accelerated, which in turn initiate
various failure mechanisms. Hence, by monitoring for the presence and change of vibration [6], the
consequences of avoidable breakdowns can be minimized.
As vibration exhibits a unique pattern or signature of motion, analysis aims to provide information
concerning its amplitude and predominant frequencies using data transmitted from sensor pickup. The
means to accomplish this may vary from direct measurement in the time domain to the sophisticated
application of several mathematical methods in both the time and frequency domains.
The first step in the monitoring of vibration is to capture an accurate recording of it, normally over a
period of time. The devices typically used for this task are transducers, which convert numerous types of
mechanical behavior into proportional electronic signals. Transducer outputs are usually converted into
voltage-sensitive signals that may be processed with various electronic instruments. Industrial trans-
ducers used for measuring dynamic characteristics can fall into the three distinct, yet mathematically
65-6 Design for Lubrication and Tribology

related, categories of acceleration, velocity, and displacement. Selection of a sensor proportional to these
parameters depends on the frequencies of interest and the signal levels involved.
Accelerometers are preferred for use in most vibration monitoring applications. These acceleration
measuring devices are fully contacting probes that are mounted directly onto a mechanical element (e.g.,
a bearing housing). They are useful for detecting low to very high frequencies (15–20,000 Hz) and are
available in a wide variety of general-purpose and application-specific designs. The piezoelectric accel-
erometer is versatile and reliable. It contains a piezoelectric crystal that emits an electrical charge when
a mechanical load is applied. It also contains a mass, which, when accelerated by vibration, imposes a
definable force against the crystal. The resulting output charge from the crystal, being proportional to
the force, which is proportional to the acceleration, is converted by an integrated circuit to a voltage
signal that can be interpreted by a recording instrument.
The rugged, solid-state construction of industrial piezoelectric sensors enables them to oper-
ate under harsh environmental conditions, unaffected by dirt, oil, and most chemical atmospheres.
Standard transducers may be purchased for operation up to 300°C, and custom accelerometers have
been built to withstand temperatures in excess of 650°C. Most piezoelectric sensors contain internal
electronics that can amplify a small acceleration signal, and if desired, convert it into a velocity or
displacement signal.
Velocity sensors, generally used for low- to medium-frequency measurements (10–1500 Hz), obtain
absolute velocity measurements of machine elements. Traditional velocity coil sensors or vibrometers
use an electromagnetic (coil and magnet) system to generate the velocity signal without the need of an
external power source. They are fully contacting probes that are mounted directly onto the structure to
be monitored.
Displacement sensors are typically used to measure changes in position and clearance. An LVDT
(linear-variable differential transformer) converts the rectilinear motion of an object to which it is
attached into a corresponding electrical signal by means of a series of inductors positioned in a hol-
low cylindrical shaft and a solid cylindrical core. An RVDT (rotary-variable displacement transformer)
similarly senses angular position. Piezoelectric transducers (doubly integrated accelerometers) have
also been used to measure displacement, yielding an output proportional to the absolute motion of a
structure. Noncontact proximity probes, such as eddy current sensors, can monitor the relative motion
between the proximity sensor mounting point and a target surface. These devices are generally used to
sense shaft vibration relative to bearings or some other support structure.
For a piezoelectric sensor, two key parameters to consider include sensitivity and frequency range.
In order to select the optimum frequency range, it is necessary to determine the frequency requirement
of the application. This may already be known from vibration data collected from similar systems or
applications; otherwise, the best approach is to place a test sensor at various locations on the system of
interest and evaluate the data collected. Given most high-frequency sensors have low sensitivities and,
conversely, most high-sensitivity sensors have low-frequency ranges, it is generally necessary to com-
promise between the two.
The collection, transmission, and recording of vibration information should be done efficiently to
avoid signal degradation. Signals must be protected from the many forms of interference generated by
electrical and mechanical components during normal circuit operation. Transmission lines should be
shielded from magnetic and electrostatic fields with appropriate materials. Consideration should be
given to cable length, temperature, and power need.
The output from vibration sensors generally requires some degree of conditioning and statistical
interpretation to provide meaningful information that can be used to detect a potentially detrimental
change in system performance. In addition to associating a given wave form to its arithmetic mean,
geometric mean, root mean square (RMS), standard deviation, kurtosis, and skewness, a variety of other
parameters have found the favor of diagnosticians including frequency, phase, peak value, crest factor,
and power spectral density.
Diagnostics 65-7

When the vibration signature is composed of a mixture of sinusoidal waveforms of different ampli-
tudes, frequencies, and phase differences, it is usually necessary to transform the data from the time
domain into the frequency domain. This may be accomplished through a fast Fourier transform (FFT),
which is a mathematical operation that decomposes the signal into a finite number of discrete frequency
components. To remove some of the unwanted noise from a signal, enveloping may be used. This tech-
nique identifies the bursts of high frequencies that occur at a regular rate and passes them through high-
frequency filters to eliminate low-frequency components.
Wavelet analysis provides an alternative approach to traditional signal processing methods, for
example, Fourier analysis, in breaking a signal up into its constituent parts. Wavelets are mathematical
functions that divide data into different frequency components and then study each component with
a resolution matched to its scale. The driving impetus behind wavelet analysis is its ability to localize
portions of signals in time (space) as well as scale (frequency), which provides an advantage over Fourier
methods in that potential problems can be more accurately represented when the signal contains dis-
continuities and sharp spikes. The technique is ideal for analyzing signals that show time delays between
pulses or pulse shape changes.
A successful vibration monitoring program using signature analysis requires a commitment to col-
lecting a sufficient amount of vibration data, well in advance of having to take corrective action. Whether
using visual or numerical techniques, data representing satisfactory machine performance must first be
established and trended to establish satisfactory levels of alarm and response.

65.4.2  Oil Monitoring


Oil is required to perform a number of functions including reducing friction, cooling components, and
cleaning load-bearing surfaces. Over time, it is likely to degrade, lose its lubrication properties due to
chemical breakdown, and become contaminated by a buildup of particles caused by component wear. A
number of techniques are available [7,8] for analyzing the condition of oil and any wear particles that are
present. Results can provide information that is very useful in determining the actual state of machine
performance and lead to improved efficiencies and cost savings in equipment operation.

65.4.2.1  Chemical and Physical Testing


The chemical and physical testing of oil necessitates that a representative sample of the oil supply be
taken for analysis. In circulating oil systems, the best location to obtain this sample is at a live zone,
upstream of filters where wear debris particles are likely to concentrate. Usually, this means sampling
from fluid return or drain lines. The sample should ideally be taken while the machine is running at
its normal load, speed, and work cycle with the lubricant at normal operating temperature. Since an
important objective in oil analysis is to obtain results of integrity, considerable care must be taken to
avoid contaminating the contaminant. Sampling intervals are commonly scheduled at a frequency that
may be keyed to recommended drain intervals or operating hours.
A number of parameters are available for describing the condition of oil. Those identified in Table
65.1 can be tested for by the user, but special instrumentation and skill are often required. For this rea-
son, analysis is commonly performed by lubricant companies or outside laboratories, offering relatively
inexpensive service.

65.4.2.2  Spectrography
Spectrography is an analysis method that identifies the presence of particular elements within an
oil. Results can be compared to those of an as-new oil sample in identifying deleterious particles
and the rate in which they generate. A typical report from this test would summarize the results for
the customer and recommend remedial action. Table 65.2 provides a summary of spectrographic
techniques.
65-8 Design for Lubrication and Tribology

TABLE 65.1  Oil Condition Monitoring


Parameter Measurement Impact
Viscosity Flow resistance Increase is symptomatic of degradation upon
oxidation or the addition of an incorrect grade of
lubricant. Reduction is indicative of contamination
Total acid number Level of acid content Increase is symptomatic of degradation through
oxidation. Reduction is indicative of additive
depletion
Total base number Concentration of the additives that At 50% reduction, an oil change is advised
counteract acid buildup
Specific gravity Mass of a fluid per unit volume referenced Change indicative of contamination
to a standard temperature
Flashpoint Temperature at which sufficient vapor is Change indicative of contamination
given off to momentarily form a flammable
mixture with air
Pour point Lowest temperature at which flow is Change indicative of contamination
observed when cooled
Water content Water content 25 ppm or greater can reduce load-carrying capacity
and negatively react with additive package
Gravimetric Dry weight of solid material present in a Verifies cleanliness levels given from an alternative
analysis specific volume (usually 100 mL) method

TABLE 65.2  Spectrographic Techniques


Method Measurement Process
Atomic absorption Concentration of particular elements Sample vaporized in flame, producing atoms that absorb
via absorption of light particular wavelengths of light. Concentration
determined from strength of absorption
Atomic emission Optical emission from excited atoms Sample is vaporized in flame when atomization occurs.
(AES) Light is emitted, producing a unique wavelength of the
activated metal, upon reverting to former stable state
X-ray fluorescence Concentration of emitted light from Similar to AES, XRF uses x-ray, gamma ray, or charged
(XRF) excitation of atoms particle beam to stimulate characteristic x-ray
emission from elements in the sample
IR Observed difference in how IR radiation Molecules in a sample absorb IR at specific wave lengths
interacts with different elements in the IR spectrum
Raman Wavelength and intensity of inelastically Raman effect is generated when a laser is directed onto
scattered light from molecules sample, exciting molecules that scatter a fraction of
this excitation light at shifted wavelengths
corresponding to different energy levels in the material

65.4.2.3  Ferrography
Ferrography is an analysis method that identifies the presence of ferrous wear particles suspended in oil
using a magnetic field to separate them according to their size. Two techniques are used: direct reading
(DR) and analytical.
DR ferrography provides a direct measure of the amount of ferrous wear metals present in a sample
of oil. The particles are separated and measured by drawing a sample of oil through a collector tube
that lies over a magnetic plate. Larger particles in the oil (greater than 15 μm) are strongly attracted
to the magnet and accumulate at the entrance of the collector tube. Smaller particles, which are only
weakly attracted by the magnet, deposit equally along the length of the collector tube. By measuring the
Diagnostics 65-9

blockage of light using fiber optics, one at the entrance of the collector tube and the other just further
up the collector tube, the quantities of large particles, denoted DL , and small particles, denoted DS, are
determined.
Trending of the DL and DS readings reveals changes in the wear mode of the system. For example, an
increase in the DL value indicates that the system has entered into an abnormal wear mode. In compari-
son, an increase in the DS value can indicate an increase in system corrosion (corrosion wear particles
are typically less than 3 μm in size). When an abnormal wear mode is detected, analytical ferrography
is used for more detailed analysis.
Analytical ferrography allows an analyst to visually examine the wear particles present in an oil
sample by use of a microscope. Oil is passed over a glass slide that rests on a magnetic plate. The ferrous
particles line up from the largest wear particles to the smallest in rows along the length of the slide,
called a ferrogram. Nonferrous wear particles can be distinguished from ferrous particles as they are
deposited randomly across the length of the slide.
Under high magnification, the particles are readily identified and classified according to their mor-
phology (size, shape, texture, etc.). A trained analyst can differentiate between varieties of wear particle
types and assess the cause of such wear. Used in conjunction with spectroscopic oil analysis, ferrography
verifies a wear condition.

65.4.2.4  Magnetic Chip Detectors


For this monitoring technique, a magnetic plug is inserted into the lubrication system before the filter
to pick up ferrous chips created due to material wear and tear. The rate at which particles appear and
the size of the particles can give an indication as to the likelihood of a machine malfunction. Larger
particles are often generated near failure. In systems where the consequences of failure can be severe,
continuous monitoring may be appropriate.

65.4.2.5  Particle Counting


A system of oil cleanliness classification was initially ratified in 1974 by the International Standards
Organization and later updated in 1987 and 1999. Identified as ISO 4406, this standard is used to
describe a theoretically infinite range of contamination levels in oil via a three-class code representing
the total number of particles per mL of oil greater than 4, 6, and 14 μm, respectively.
To manually count, size, and evaluate the types of particles present in oil, oil is drawn through a
membrane of known pore size and the trapped particles are counted by viewing the membrane under
a microscope at various magnifications. As the technique is tedious, statistical counting techniques are
usually employed.
Automatic particle counters have been developed, which have certain limitations and give erroneous
counts if not used correctly. One approach uses the principle of fluid flow decay in which oil is passed
through a screen of known mesh. The presence of contaminant gradually blocks the pores of the screen.
By knowing the number of pores in the screen and the volume of oil that passes, the number of particles
greater than the pore size per unit volume can be inferred by the degree of blockage. The disadvantage
of using this technique is that it assumes a predetermined size distribution without actually measuring
the number of particles.
The most common approach in automatic particle counting uses a light blockage principle. Typically,
a known volume of oil (usually 5 mL) is injected through a very small sampling cell. On one side of the
cell is a beam of laser light and on the other side, a detector. As particles pass through the cell, they are
counted as they block the beam of light by casting a shadow on the detector. The drop in light intensity
received by the detector is proportional to the size of the particle blocking the light beam. In this way,
both the number and size of the particles can be measured. An instrument is usually set up to measure
particles in five different micron size ranges: 5–15, 15–25, 25–50, 50–100, and greater than 100. Results
are expressed as the total number of counts per mL of oil.
65-10 Design for Lubrication and Tribology

65.4.2.6  Image Analysis


Image analysis provides information through visual means. One of the simpler methods of image analy-
sis in assessing used oil is the patch test, a method by which a specified volume of fluid is filtered through
a membrane filter of known pore structure. All particulate matters in excess of an average size, deter-
mined by the membrane characteristics, are retained on the surface. Thus, the membrane is discolored
by an amount proportional to the particulate level of the fluid sample. Visually comparing the test filter
with standard patches of known contamination levels determines acceptability for a given fluid. The
membrane could also be scanned using a video camera, with the resulting image being converted into a
signal for processing by an image-analyzing computer (IAC). Particles are identified in relation to their
gray scale or color contrast with the surface of the membrane. The computer then applies the logic to
provide particle count and a host of shape parameters, including dimension, area, and circumference.
Potentially, this method overcomes the tiresome task of visual inspection; however, the technique has
some shortcomings. If particles cluster together, editing is necessary to avoid interpreting the mass as a
single particle. If there is a similarity between the color of the particle and background, incorrect sizing
can occur.

65.4.3  Thermal Monitoring


Thermal energy is often generated and transferred when a system begins to experience trouble, possi-
bly attributed to friction, overloading, chemical reactivity, and/or insulation damage. Manifested as an
unexpected temperature rise, action is warranted to determine the cause. Thermal monitoring methods
available to help diagnose a situation fall into two categories: contact and noncontact [9,10].
Contact methods for measuring temperature infer that something is placed on or within the surface
of the component being assessed. Thermal paints and crayons fall in this category. They are easy to apply
on the surface of interest and change color when a particular temperature range is reached, giving an
observer a quick indication of a thermal event. Thermometers, such as the sealed liquid-in-glass type,
consist of a glass tube containing liquid. If the temperature increases, the liquid expands and rises in
the tube. The temperature is then read on an adjacent scale. Mercury is commonly used as the liquid for
measuring temperatures ranging from −15°C to 540°C. Other instruments, often used in permanent
temperature control and measurement systems, include the thermocouple, RTD (resistance-tempera-
ture detector), and thermistor.
A thermocouple is a temperature transducer consisting of two wires of different metals (e.g., iron/
constantan) joined at both ends. If the two junctions between the metals are at different temperatures,
an electric current will flow around the circuit at a voltage directly proportional to the difference in
temperature between the two junctions. When used for measuring temperature, one of the two junc-
tions is placed in contact with the material whose temperature is to be monitored. The temperature of
the second junction must be either known or controlled to at least the accuracy expected for measure-
ment. A meter used to measure the voltage of the circuit may be calibrated directly to read in degrees of
temperature or provide the user a direct voltage reading to be converted into temperature via a standard
thermocouple chart. Thermocouples are self-powered, rugged, and are capable of measuring within a
large temperature range (−200°C to 1800°C).
An RTD is constructed with a wire coil or a thin layer of metal to form a precision resistor. The
resistance value changes very accurately and repeatedly in a positive direction when heated. By incor-
porating the resistor in an electric circuit and attaching it to the material whose temperature is being
measured, temperature can be determined from the change in resistance with the use of an equation
relating the two variables. RTD assemblies can be used in a wide variety of configurations to give the
highest accuracy of temperature measurement. An RTD is considered to be very stable, measuring
temperatures in the range of −260°C to 850°C. It tends to be expensive and requires a current source.
Diagnostics 65-11

Platinum is usually used as the resistive material because it is chemically inert and exhibits repeatable
resistance-temperature characteristics.
A thermistor is generally described as a thermally sensitive resistor constructed with metal oxides
formed into a bead and encapsulated in epoxy or glass. Its resistance exhibits a nonlinear large negative
change as it is heated. Used in the same way as an RTD, the change in resistance recorded during a small
temperature change of a thermistor is several times greater than an RTD, making measurement easier.
The device tends to be low in cost, yet fragile. Its temperature range is limited (−80°C to 300°C).
The noncontact approach to measuring temperature uses the principle that all objects emit electro-
magnetic waves from their surface in proportion to their warmth. By focusing this form of radiation
from its source onto a sensor, its intensity can be interpreted and displayed as a temperature. The main
advantage of this method, applied in pyrometers and infrared (IR) imaging, is that large areas can be
quickly surveyed at a safe distance.
A pyrometer is a device with sensors that accept a range of radiant energy wavelengths from the vis-
ible light portion to the IR portion of the electromagnetic spectrum. A detector, usually focused on a
select point of the surface of interest through a suitable lens, converts the radiant energy emitted by the
surface into an electrical signal that is interpreted to provide a temperature reading. One type of pyrom-
eter, the optical or brightness pyrometer, requires manual adjustment based on what is viewed through
a sighting window. A second type, the two-color ratio pyrometer, compares the spectral radiance at two
wavelengths to identify the temperature.
The total radiation pyrometer is commonly used for industrial applications, where temperature mea-
surement from −150°C to 4000°C is needed. It responds reliably to all wavelengths in both the visible
and IR portions of the spectrum, with exception to those altered by components of the system, for
example, lenses, filters, and atmosphere. Calibration of the instrument is required, typically using a
blackbody, which is an object having a surface that does not reflect or pass radiation. Temperature limits
and speed of response are dependent on the type of radiation detector used.
Thermal images are pictures of heat using the radiated energy emitted from an object. They are cre-
ated by an IR thermal imager, a device that captures a portion of the radiated energy through a variety of
scanning camera techniques, yielding a spatial map of temperatures. The instrument typically incorpo-
rates a cooling system to avoid the effects of system noise created by the sensing of its own temperature.
IR imaging systems are available with a wide range of capabilities, features, and prices. Scan speeds
range from real time to seconds per image. Detectors range from being simple, single-element, and
thermoelectrically cooled to state-of-the-art, multielement focal plane detector arrays, incorporating
closed cycle Sterling coolers. Spot temperature measurement accuracy as high as ±1% can be achieved
in a range from −20°C to 2000°C when adequate compensation is given for the emissivity of and dis-
tance to the target region, and a reliable method is used to locate the point of interest in the image.
Measurements are taken in real time, and processing takes only a few minutes.

65.4.4  Performance Monitoring


Performance monitoring uses process information to quantify or characterize normal machine behav-
ior and highlight abnormal characteristics that often indicate deterioration. Information, representative
of parameters identified through observation and measurement, is collected over time for interpretation
and benchmarking using manual and computer methods [11,12].

65.4.4.1  Flow Rate


The flow rate of a fluid entering or exiting a component is one of the basic parameters of a hydraulic
system and is often a direct measure of its performance. A variation in flow can be indicative of leakage
or internal wear. Numerous types of flowmeters are available for closed-piping systems, of which dif-
ferential pressure and velocity meters are commonly used. The differential pressure flowmeter is based
65-12 Design for Lubrication and Tribology

on the premise that a pressure drop across the meter is proportional to the square of the flow rate. It
has a primary element that causes a differential pressure in the closed-piping system and a secondary
element to measure the differential pressure and provide the readout that is converted to an actual
flow value.
Orifices and Venturi tubes are examples of differential pressure meters. The former is a flat piece of
metal with a specific-size hole bored in it. It is installed in the pipe between two flanges and constricts
the flow of fluid to produce a differential pressure across the plate. The latter is a section of pipe with a
tapered entrance and straight throat. As the fluid passes through the throat, its velocity increases, caus-
ing a pressure differential between the inlet and outlet regions. In each device, pressure taps at entrance
and exit are used to measure the pressure difference used to calculate flow rate.
Velocity meters operate linearly with respect to the flow rate. The most adaptable velocity meter is the
turbine meter, which consists of a multiple-bladed rotor mounted within a pipe, perpendicular to the
fluid flow. The rotor, or turbine, rotates as the fluid passes through the blades. The rotational speed is
directly related to flow rate and can be sensed by either a capacitive, inductive, or magnetic pickup. Flow
rates from 0.5 to 1000 L/min have been accurately measured using this device.
A second type of velocity meter uses ultrasonics to determine the flow rate in a piping system. One
method uses a piezoelectric crystal mounted on the outside of the pipe to emit an ultrasonic signal of
known frequency into the flow stream. Upon contacting contaminant particles, bubbles, or discontinui-
ties, the signal is reflected back to a receiver element. Because the fluid causing the reflection is moving,
the frequency of the returned pulse is shifted. The frequency shift is proportional to the fluid velocity.
This portable technique is very accurate in measuring flow velocities from 0 to 10 m/s.
An alternative ultrasonic approach uses two transducers mounted on opposite sides of the pipe at a
45° angle to the direction of fluid flow. The speed of a signal traveling between the transducers increases
or decreases with the direction of transmission and the velocity of the fluid being measured. A time-
differential relationship proportional to the flow velocity can be obtained by alternating the signal trans-
mission in either direction. A limitation of this time-of-travel meter is that the fluid being measured
must be relatively free of particles to minimize signal scatter and absorption.

65.4.4.2  Current Signature Analysis


In motor-driven rotating equipment, current signature analysis provides a nonintrusive method for
detecting mechanical and electrical problems. The current signal can be measured using an ammeter,
either clipped on to the motor circuit or installed in the machine control panel. A spectrum analyzer
and or commercial computer with signal conditioning capability is needed to process the signal and
provide diagnoses.
When an electric motor drives a mechanical system, it experiences variations in load caused by gears,
bearings, and other conditions that may change over the life of the motor. The variations in load caused
by each of these factors in turn causes a variation in the current supplied to the motor. These variations,
though very small, modulate the 60 Hz (or 50 Hz) carrier frequency.
Upon demodulating the signal from the carrier using Fourier techniques, abnormal signal charac-
teristics can be identified, representing a variety of failure precursors including rotor bar deterioration,
stator phase imbalance, and increased friction forces. A single measurement of current is not likely to
be informative, but comparison with prior readings can give immediate results. Hence, recording the
current at regular intervals is advised in order to identify changes or trends toward failure.

65.5  Tribosystem Applications


As summarized in Table 65.3, there are several different types of measurements that can be used to
facilitate the diagnosis and prognosis of machines, containing such triboelements as bearings, gears,
and seals. The technical requirements of each of these are very different, and as a result, there is a ten-
dency to view the technologies in competing roles. Yet, as can be seen by using hydraulic systems and
Diagnostics 65-13

TABLE 65.3  Summary of Diagnostic Approaches


Monitoring Technique Elements Monitored Machine Diagnosis Measurement
NDE Rotating equipment Cracks Flaw size
Visual inspection Structures Decay Frequency
Penetrant/particle Leakage
inspection
Eddy current inspection Wear
AE monitoring
Ultrasonic monitoring
Radiography

Vibration Bearings Imbalance Displacement


Signature analysis Gears Looseness Velocity
Rotors Misalignment Acceleration
Shafts Wear Spike energy

Oil analysis Bearings Contamination Composition


Chemical/physical testing Gears Degradation Contaminants
Spectrography Rotors Fracture
Ferrography Lubricant Wear
Magnetic chip detection
Particle counting
Image analysis

Thermal monitoring Bearings Chemical reaction Temperature


Contact methods Coolant Fracture
Noncontact methods Gears Friction
Lubricant Overload
Motors Wear

Performance monitoring Filters Blockages Current


Flow Seals Failed rotors Flow rate
Current monitoring Motors Worn brushes Pressure

fluid film bearings as examples [13], best results are realized if the technologies can be used to comple-
ment each other.

65.5.1  Hydraulic Systems


Many of the failures in a hydraulic system show similar symptoms: a gradual or sudden loss of high
pressure, resulting in loss of power or speed in the cylinders. In fact, the cylinders may stall under light
loads or may not move at all. Often, the loss of power is accompanied by an increase in pump noise,
especially as the pump tries to build up pressure. Any major component (pump, relief valve, directional
valve, or cylinder) could be at fault, with the cause attributed to one or more reasons, including high
contamination levels, wrong oil viscosity, high-temperature operation, and cavitation.
Contaminants of hydraulic fluid include solid particles, air, water, or any other matter that impairs
the function of the fluid. Particle contamination accelerates wear of hydraulic components. The rate
at which damage occurs is dependent on the internal clearance of the components within the system,
the size and quantity of particles present in the fluid, and system pressure. Particles smaller than 5 μm
are highly abrasive. If present in sufficient quantities, these invisible “silt” particles cause rapid wear,
destroying hydraulic pumps and other components, hence highlighting the importance of monitoring
hydraulic fluid cleanliness levels at regular intervals. If the high levels of silt particles present in the
65-14 Design for Lubrication and Tribology

hydraulic fluid are identified and the problem rectified early enough, the damage to a hydraulic pump
and the significant expense of its repair can be avoided.
Most hydraulic systems will operate satisfactorily using a variety of fluids, including multigrade
engine oil and automatic transmission fluid (ATF), in addition to the more conventional antiwear (AW)
hydraulic fluid, provided the viscosity is correct. Viscosity is the single most important factor when
selecting a hydraulic fluid. Note that as the temperature of a petroleum-based hydraulic fluid increases,
its viscosity decreases. If fluid temperature increases to the point where viscosity falls below the level
required to maintain a lubricating film between the internal parts of the component, damage will result,
thus highlighting the importance of monitoring fluid temperature. When a hydraulic system starts to
overheat, the system must be shut down so as to identify the cause for subsequent repair.
Cavitation occurs when the volume of hydraulic fluid demanded by any part of a hydraulic circuit
exceeds the volume of fluid being supplied. This causes the absolute pressure in that part of the flow
circuit to fall below the vapor pressure of the hydraulic fluid, resulting in the formation of vapor bubbles
within the fluid, which implode when compressed. Cavitation causes metal erosion, which damages
hydraulic components and contaminates the hydraulic fluid. In extreme cases, cavitation can result in
major mechanical failure of pumps and motors. While cavitation commonly occurs in the hydraulic
pump, it can occur just about anywhere within a hydraulic circuit. In a hydraulic valve, metal erosion
in the body of the valve can be so severe that the valve is no longer serviceable, thus highlighting the
importance of checking the operation and adjustment of circuit protection devices, including anticavi-
tation and load control valves, at regular intervals.
The vast majority of hydraulic systems in operation today have leaks, most of which are planned. They
are designed with a specific function in mind and, in many cases, are documented by the original equip-
ment manufacturer as the amount of acceptable leakage under normal operating conditions. Internal
planned leakage is typically small orifices or pathways that allow a fluid from a higher pressurized zone
of a system to travel into a lower pressurized zone to lubricate, clean, and cool a specific component
or area. These planned internal leaks do not allow the fluid to exit the hydraulic circuit, so there is no
visual indication of its presence. The most common cause of excessive internal leakage is wear of com-
ponent surfaces during normal operation. Leakage can also result from poor system design, incorrect
component selection, poor quality control tolerances during the manufacturing of a component, and
incorrect overhaul of rebuilt components. System performance, reliability, and increased operating tem-
peratures are the first visual signs of excessive internal leakage. Identifying and controlling hydraulic
system leakage require an in-depth approach to record keeping and surveillance. In addition, dedication
to performing repairs and/or modifications aimed at the root causes of the leaks, along with a method
of monitoring, will ensure that the repairs are effective. Low fluid viscosity or excessive heat (reducing
the effective viscosity of a fluid) will increase leakage rates. This form of internal leakage reduces system
performance and decreases fluid film strength, which will also result in premature wear of the equip-
ment surfaces and the fluid’s properties.
Eventually, all of the aforementioned conditions will affect hydraulic system performance and, ulti-
mately, company profits. Detection of unplanned internal leakage, in most cases, would rely on specific
tools to examine the location and quantity of the leak. Performance issues or the inability of a circuit
to perform its designed function typically triggers the installation of flowmeters in various locations.
Noncontact IR thermometers are useful for nonobtrusive measurement of operating temperatures
of equipment. An abnormal temperature increase at a relief valve could indicate that the valve is in a
bypassing condition. This bypass condition will generate heat locally in the component; in many cases,
the anomaly would have gone undetected by monitoring the system reservoir temperature because of
system coolers or dissipation of heat throughout the system.
Ultrasonic detection has proven to be another effective method of determining high-pressure or
high-velocity leaks in various locations of valve and cylinder leakage. This method enables the localiza-
tion of the internal leakage, but similar to temperature reading, the results are not quantifiable into the
amount of leakage.
Diagnostics 65-15

The only quantifiable method is to measure the flow or quantity of fluid loss in a given time frame
using a flowmeter or other related test equipment. Detection and quantification of the fluid consump-
tion is the first step in external leak control. Up-to-date reservoir management records must be main-
tained to determine when, by whom, and how much fluid was required to top up a reservoir. These
records should be used along with visual inspections to determine the location and the leak rate of any
detected anomalies. Quantification of the leakage rate and location will allow for the opportunity to
prioritize the repairs.
In many cases, the source and quantity of the leaks cannot be determined, as they are difficult to
see. To alleviate this problem, dyes sensitive to black light have been formulated to assist in the location
and identification of external of leaks. This liquid dye is formulated to be compatible with the existing
hydraulic fluid and machine surfaces. The dye is mixed into the reservoir after which the mixture will
emit a bright green/yellow glow when struck by the rays of a black light. This method of visual detec-
tion helps determine whether the fluid being viewed is from an active leak from the system in question.

65.5.2  Fluid Film Bearings


A fluid film bearing is used in a wide variety of machines to support and guide a rotating shaft. In its
most basic form, the shaft is contained within a stationary close-fitting partial or full cylinder separated
by a film of fluid. The fluid, most commonly oil, is supplied to the clearance space between the surfaces
to create a wedge support that prevents metal-to-metal contact as rotation occurs.
Many application-dependent bearing designs have been created, each developed to use the fluid
film to position the shaft or rotor at an optimum location relative to the bearing housing. Application
of a disturbing force will move the shaft from this position, altering the pressure field within the fluid
and hence the forces within the bearing. The forces generated normally tend to restore the shaft to its
optimum position, but, sometimes, they can lead to an unstable motion or self-excited vibration. For
an oil film bearing, the vibration, often referred to as oil whirl, usually coincides with the average oil
velocity, which is typically less than half of the shaft rotational frequency. When this whirl frequency
coincides with a rotor resonance, a severe vibratory state can occur. Known as an oil whip condition,
the fluid film support becomes unstable, leading to severe bearing and shaft deterioration from metal-
to-metal contact.
To be assured that a machine maintains balance, stability, and proper alignment with the bearing
supports, vibration monitoring is needed to assess the relative motion between shaft and bearing. A
noncontacting proximity probe, such as an eddy current transducer, is typically used. On small, less
critical machines, one probe per bearing may be adequate, which will measure radial vibration for the
plane in which the shaft moves away from and toward the mounted probe. On larger machines, two
probes per bearing mounted 90° apart from each other are advised to assess the total radial vibration by
measuring the shaft displacements within their respective planes.
To complement vibration analysis, the fluid used to support the shaft should be periodically moni-
tored. Its physical and chemical properties must be evaluated and maintained to assure that the film
performs as designed and provides sufficient damping to limit vibration transmission. An increase in
temperature and or the presence of such elements as aluminum, tin, lead, copper, zinc, and iron, com-
monly used as bearing materials, may indicate the onset of wear from metal-to-metal contact.

References
1. Montgomery, D.C., Introduction to Statistical Quality Control, Wiley, New York, 1996.
2. Bray, D. and Stanley, R.K., Nondestructive Evaluation: A Tool in Design, Manufacturing, and Service,
CRC Press, Boca Raton, FL, 1997.
3. Sollier, T., Premel, D., and Lesselier, D. (eds.), Electromagnetic Nondestructive Evaluation (VIII), IOS
Press, Amsterdam, the Netherlands, 2004.
65-16 Design for Lubrication and Tribology

4. Rao, B.K.N., Handbook of Condition Monitoring, 1st edn., Elsevier Science, Ltd., Oxford, U.K., 1996.
5. Solomon, S., Sensors Handbook, McGraw-Hill, New York, 2010.
6. DeSilva, C.W., Vibration Monitoring, Testing and Instrumentation, CRC Press, Boca Raton, FL, 2007.
7. Denis, J., Briant, J., and Hipeaux, J-C., Lubricant Properties, Analysis and Testing, Editions Technip,
Paris, France, 2000.
8. Slaymaker, R.R., Bearing Lubrication Analysis, Wiley, New York, 1955.
9. Leigh, J.R., Temperature Measurement and Control, P. Peregrinus (IEE), London, U.K., 1988.
10. Childs, P.R.N., Practical Temperature Measurement, Butterworth-Heinemann, Oxford, U.K., 2001.
11. Hunt, T., Condition Monitoring of Mechanical and Hydraulic Plant: A Concise Introduction and Guide,
Chapman and Hall, New York, 1996.
12. Tavner, P.J., Condition Monitoring of Rotating Electrical Machines, Institution of Engineering and
Technology, London, U.K., 2008.
13. Totten, G.G., Handbook of Hydraulic Fluid Power Technology, Marcel Dekker, New York, 2000.
66
Tribology Testing
66.1 Types of Testing...............................................................................66-1
66.2 Testing Delineated by Assembly Complexity..............................66-2
66.3 Preparing for Tribology Testing....................................................66-2
66.4 Parameters........................................................................................66-3
Load  •  Friction Force (Output Force Measurement)
•  Motion  •  Temperature  •  Pressure (or Vacuum)
•  Surface Finish, Geometry, and Cleanliness of Test
Specimen  •  Replicas  •  Cleanliness  •  Material
Properties  •  Lubrication Regime or Wear Mode
66.5 Selection of Type of Tester..............................................................66-9
Standard Tests  •  Custom Apparatus  •  Modification of Existing
Apparatus  •  Conversion/Incorporation of Actual Component
into Tester
66.6 Metallurgical Techniques Supporting Tribology Testing.......66-11
66.7 Tribology Test (Other than Friction and Wear Testing).........66-12
Lubricant Properties  •  Lubricant Analysis  •  Traction
Testing  •  Grease Testing  •  Bearing Fatigue Life
Testing  •  Corrosion Testing  •  EHD Film Thickness
66.8 Design of Experiment...................................................................66-14
66.9 Break-In or Wear-In Phenomena................................................66-15
66.10 Accelerated Testing.......................................................................66-15
66.11 Statistical Analysis of Friction and Wear Data.........................66-15
Terry L. Merriman 66.12 Role of Modeling in Tribology Testing......................................66-15
Battelle References...................................................................................................66-16

66.1  Types of Testing


Relatively few texts consider tribology testing as a whole [1], yet once the subject is broken down into
the specific type of testing desired, there are more numerous sources of information. Examples of text
on specific aspects include the Wear Control Handbook [2], the Tribology Data Handbook [3], Friction,
Lubrication and Wear Technology [4], and the Tribology Handbook [5].
For the most part, the challenge will be to find a source of information that is pertinent to the spe-
cific test desired. The following discussion considers how the field of tribology testing may be logically
categorized.
Tribological testing includes, but is not limited to, friction, wear, and lubrication testing. Friction
testing is a specialized area that can be reviewed in brief with reasonable completeness. The subjects
of wear and lubrication cover such a broad range that perhaps the best approach is to point to other
detailed resources. Wear may include but is not limited to

66-1
66-2 Design for Lubrication and Tribology

• Materials in dry sliding contact, which may include metals, ceramics, and polymers
• Materials in lubricated contact, which may include surface coatings, paints, boundary lubricants,
and solid lubricants
• Erosive wear, a specific topic involving liquid impingement
• Wear with corrosion effects, a subtopic wherein the environment augments the wear processes
A next level of detail is definition of the type of wear mechanism. Wear mechanisms are typically
described as adhesive wear, abrasive wear, rolling wear, fretting, galling, thermal, or oxidative. The
reader will find that these terms are sometimes used slightly differently or preferably, yet there is no
reason to be deterred by these distinctions. The necessary point is to find a test that models sufficiently
the engineering system under consideration so that design decisions can be adequately addressed. One
consideration in selecting tests is whether testing should be done as an entire mechanism or assembly,
at a component level, or at an elemental level.

66.2  Testing Delineated by Assembly Complexity


Tribology testing is usually done to support an emerging design or assembly. Ultimately, production of
useful products will be the eventual goal of testing. An assembly may be a complex interaction of many
tribological components—a convenient example is an automobile. An automobile will contain hun-
dreds of tribological interfaces such as journal bearings, rolling element bearings, gears, piston rings,
and valves. The following example considers increasing specificity. An automobile may be tested at
• Fleet level, such as a group of taxis or trucks operating in a city or state
• Assembly level, for example, a transmission from a vehicle
• Component level, such as one shift solenoid within a transmission
• Interface level, such as a seal inside a solenoid
Fleet data from taxis or commercial trucks are often part of their industrial management. Tribological
failures account for much of the eventual attrition of the fleet. Data gathered at the fleet level are useful
yet generalized. The data will be pertinent to the application, but the parameters of the experiment will
be generally known only to a higher level of definition, that is, the car was driven one hundred miles.
At the next level of specificity, a particular system on the car might be subjected to tribological test-
ing. A transmission, for example, is tested for some period of time to determine if there are any com-
ponents with inclinations toward tribological distress. This testing is more likely to be performed at a
test bench specifically designed to accommodate the transmission as a whole. Testing at this level would
again show good similitude since the actual article is under test and the test parameters will be less
global, more specific than that of the entire car.
Perhaps in this example, the transmission will show generally satisfactory life; however, one or more
components may require additional tribological attention. In choosing a testing strategy, is it more effi-
cient to test the component itself to evaluate tribological performance or is it desirable to break the
component down to a specific interface of interest? This decision may be guided by whether or not the
conditions at the tribological interface can be ascertained with reasonable certainty. To employ a stan-
dard test, the parameters in the actual component will have to be known in detail. An American Society
for Testing and Materials (ASTM) [6] or equivalent standard test would be considered to be the most
specific on the spectrum described here. Parameters will be well defined, and similitude to the actual
component will be dependent on selection and values of the parameters.

66.3  Preparing for Tribology Testing


An important step in preparing for tribological testing is gaining an understanding of the tribological
mode that is occurring in an actual part and considering the related critical parameters. If the incorrect
Tribology Testing 66-3

type of test is conducted, then the results are not applicable or useful for design. The value of data gath-
ered about dry wear will be questionable if the actual part is lubricated. Data gathered about mild wear
processes will be useless if the actual part experiences severe galling and seizure. Therefore, step one in
selecting or designing a useful tribological test is to do some logical sleuthing.

66.4  Parameters
The parameters that influence wear in machine components include the following:
• Load
• Motion
• Temperature
• Pressure (or vacuum)
• Surface geometries and textures
• Material properties
• Lubrication regime
The experimenter should systematically write down a description of the problem with the following
guidelines in mind.

66.4.1  Load
Load should be measured normal to the contact interface. Contact pressures may be calculated for vari-
ous purposes, but load is typically the parameter of control in friction and wear studies. This concept
is tied to concepts of real area of contact and basic (Amonton’s) friction models. For efficiency, these
concepts will not be explained here; they are documented elsewhere.
Normal load in tribology testing is the load component that is at 90° to the sliding interface. It is
important to measure or predict the magnitude, direction, and variations of this load in service. The
wear tester selected will need to be capable of producing the same load. Light loads may sometimes be
applied through dead weights, that is, simple calibrated weights pressing by gravity on the specimen
holder.
If required loads rise above several hundred pounds, dead weights become unwieldy. The next step
up is to use a lever arm with mechanical advantage to multiply the dead weight. By this means, mul-
tiplications of several times are practically possible, and thus, the load to be applied may approach a
thousand pounds. Simple pulleys or gearing can be combined with a lever arm to further gain mechani-
cal advantage. As one example, researchers produced a number of simple three button testers in the
1980s by modifying sturdy drill press frames. The quill of the drill press applied the load, and the lever
arm (normally manually operated) was loaded with a weight pan. Through mechanical means, the total
applied load approached one thousand pounds.
Load ranges for friction testing can range much higher than a 1000 lb (0.45 MT). Airframe bearing
tests, for example, must mimic the loads experienced when a commercial airliner hits the landing strip.
These loads can be 15,000–20,000 lb (6.8–9.1 MT). The closing action of a gate valve in a nuclear power
plant application is another example of a highly loaded interface. To produce these loads in a test appa-
ratus, one technique employs pneumatic air bags of the type used to support semitractor trailer beds.
These air bags are readily available at relatively low cost and can be controlled with simple regulators to
add or release pressure to maintain a fixed pressure in the air bag. When combined with a mechanical
advantage (lever arm), this is a reliable means of maintaining a substantial load over a long test. The
pressures used in the air bags are normal shop pressures and thus with suitable precautions can be eas-
ily arranged.
For high loading, the experimentalist may need to resort to hydraulic cylinders. Hydraulics range
from small power packs that produce several thousand pounds up to a full-scale MTS servohydraulic
66-4 Design for Lubrication and Tribology

test system that may supply as much as 500,000 lb (227 metric ton). The hydraulic power supply to con-
trol and feed such a system is sufficiently costly that this type of work is normally contracted with an
established laboratory. A hydraulic or MTS-type system has the advantage that the load cycle can be
programmed to follow a complex input profile if desired.
No matter how the load is applied, there are a couple of principles to keep in mind. The load should
be measured directly in line with the test specimen if at all possible as otherwise the friction force
will introduce a torque that will affect the normal force. Any mechanical arrangements to multiply
the load can and will introduce some parasitic friction if not taken into account. The more closely the
load measurement is moved directly in line with and close to the test specimen, the least likely error
will creep in. Direct load measurement is often possible through the introduction of a compression
load cell. With the introduction of piezoelectric technology, load cells have large ranges yet good
precision. The experimenter need assure that the load cell is rated for sufficient capacity that it may
not be accidentally overloaded and destroyed. Some extra margin of capacity should be designed;
however, keep in mind that the resolution of the load cell is normally expressed as some percentage
of its full-scale range.
It is important that the specimen is properly aligned so that the load is applied evenly over the test
area. An improperly aligned specimen will not change the total applied load; however, for a misaligned
specimen, edge loading will increase the unit loading or contact pressure. Whereas the friction coeffi-
cient is less sensitive to load variations, the wear regime can be quite sensitive to contact pressure. With
a properly aligned specimen, better consistency will be achieved in experiment results. Other than geo-
metric checks while constructing the apparatus and specimens, probably the surest manner in checking
alignment will be to examine a wear scar after a short test period. Symmetry in the wear scar will be a
sensitive indicator of good alignment of the specimen and even load application.

66.4.2  Friction Force (Output Force Measurement)


The normal load on the specimen is thought of as the input parameter in friction and wear testing. The
friction force produced is thought of as a resultant or output parameter. Therefore, in addition to the
applied load or input load, the experimenter must measure the friction force produced and required to
move the specimen. Whereas the input load is often steady or at least applied in one direction, a slid-
ing, reciprocating, or rotational specimen creates specific challenges for measuring the friction force
output. If the test is a sliding motion (reciprocating and rotational will be considered presently), then it
is often convenient to place a load cell immediately in line with the specimen. Specifically, the specimen
being tested will generally be contained in some larger holder or sled arrangement. This often facilitates
attachment of a load cell. Several cautions are in order for this load cell. The load cell will have to be
selected to be good for both tension and compression; these types are slightly less common and possibly
less sturdy than simple compression load cells. Any twisting or out-of-plane force applied to this load
cell will introduce false readings as well as likely damage the load cell. To this end, the specimen holder
for a reciprocating test is often mounted on a linear slide table with sturdy crossed cylinder bearings.
These tables when properly adjusted can produce little parasitic friction keeping in mind any friction in
your apparatus be nulled out.
Next, consider how the friction force can be measured on a rotating specimen whether the test
specimen is either in continuous rotation or reciprocating in a clockwise and counterclockwise pat-
tern through a fixed angle. As an example, in a geometry such as a pin-on-disk, three-button test, or
four-ball tester, the problem is simplified by mounting one-half of the rotating setup on a low friction
(typically rolling element) bearing and restraining this part from rotating by resting a lever arm against
a stationary load cell. The input motion and force are through the vertical quill that holds a specimen
holder rigidly attached to the quill. The bottom specimen is a round plate mounted on a rolling ele-
ment bearing, which is capable of withstanding the test load and still rotates quite freely. This bottom
round plate has an arm that extends, and the length of the arm serves as a mechanical multiplier on the
Tribology Testing 66-5

measured torque. The arm is retrained against a simple compression load cell; in this case, a test rotat-
ing continuously in one direction is assumed. The arm need not be rigidly attached to the load cell, and,
indeed, if the force is transmitted through a pivot point such as a bearing ball, the experimenter elimi-
nates unwanted twisting and only measures the rotational torque induced in the plate when through the
frictional interface of the apparatus.
A ring-on-block test uses a variation of this principle to measure the output friction torque. In this
test, a stationary block is brought to bear against a rotating ring. The block will tend to move in a tan-
gential direction to the wheel but is restrained by an arrangement that eventually bears on a load cell.
A ring-on-block tester is customarily designed with flexures that allow the stationary block to transmit
force in the tangential direction but are rigid in restraining movement in other directions. The study of
these types of flexures can be instructional if a custom friction tester or modification of such a tester is
to be made.
Finally, an alternative is considered where in place of measuring the force at the stationary part of the
apparatus, the equipment is instrumented to measure the torque input working against friction at the
specimen interface. The simplest version of this is a journal bearing where the torque input is measured
as a means to quantify the friction at the shaft to bearing interface. The torque divided by the radius is
a measure of the friction force in this case. The most straightforward means of measuring torque input
on a rotating shaft is a torque cell produced for this purpose. The torque cell has strain gages built onto
its shaft and slip rings to transmit the torque measurement. The experimenter can build these elements
into the hardware by instrumenting the input shaft; however, the shaft itself must be typically of reduced
cross section to create suitably measurable strains. For the relatively small expense, it is often simpler to
use a commercial torque cell.

66.4.3  Motion
The relative movement of the parts in contact should be known, again including the correct magnitude,
direction, and variations or specific motion profiles. Motion in a friction or wear test normally consists
of a back and forth linear sliding motion, a reciprocating motion, or a simple rotation. Magnitude of
motion helps delineate the type of wear. For example, small motions, 0.003 in. (0.076 mm) or less, may
indicate fretting is possible. Reciprocal sliding motions of 0.040 in. (1 mm) or less may entrap wear
debris in both the component and wear tester. Reciprocal sliding of greater than 0.040 in. (1 mm) in the
component may be similar to unidirectional sliding in the wear tester.
Mechanical crank arm mechanisms can readily produce a reciprocating sliding action. The throw
adjustment on the arm is used to set the total stroke. These type mechanisms are typically driven by
electric motors. With the advent of low-cost AC frequency controllers and closed loop feedback control,
the ability to accurately control test speeds has improved in recent years.
Under extreme loads or speeds, however, mechanical or hydraulic means are sometimes required to
maintain controlled speeds throughout a test. The controls of distance and time of travel are incorpo-
rated into the power pack for the hydraulic cylinder.
Controlled motion can be particularly challenging if the friction coefficient changes severely during
a test. A key element in planning this stage of the experiment is to estimate what the friction forces may
be at all points in the experiment. If mild wear modes can be assured through controlling test param-
eters such as material choice and lubrication, then the test may well proceed in an orderly fashion with
friction remaining well within expected boundaries. If there is a possibility of severe wear and adhesive
welding, the experimenter should keep in mind the friction coefficient can go above unity and the forces
transmitted to the prime mover can be destructive.
This brings to mind the admonition that a safe and reasonable stop to an experiment must be consid-
ered beforehand. If a specimen fails and forces increase catastrophically, much expensive instrumenta-
tion can be destroyed if an orderly stop has not been properly engineered. Safety measures can include
physical or electronic braking and or weak links (mechanical fuses) built into the apparatus.
66-6 Design for Lubrication and Tribology

66.4.4  Temperature
Temperature is often either monitored or controlled throughout tribology testing. The temperature
will modify basic properties of the materials and lubricants under test. The temperature can also
affect the wear/lubrication mechanism or mode. A complete temperature profile for the test should
therefore be planned. Both heat input and output must be considered. Temperature can be controlled
during a test by manipulating the environment. The specimen can be flooded or immersed in a fluid
bath that is heated or cooled. In one series of experiments related to a high-pressure oxygen tur-
bopump, liquid nitrogen was used flowing through the test apparatus to remove large amounts of
excess heat buildup. To a lesser extent, a similar control can be achieved through convective cooling
with airflow.
Measurement of temperature in a wear or friction apparatus is often achieved through thermo-
couples (or thermistors). Thermocouples are easy to implement in most cases. They come in a wide
variety of profiles that affect their time response. They have few drawbacks. Their time constant can
be a bit slow, yet in most friction or wear experiments, the temperature is changing gradually enough
that it is not a problem. There are many types of thermocouples to choose from; those with the small-
est mass generally have the fastest response. A chief note with regard to thermocouples is to remem-
ber the thermocouple only measures the temperature at the juncture where it is applied. By drilling
or electrodischarge machining and carefully placing a thermocouple close to the test surface, fairly
accurate measurements are made. If the thermocouple is placed some distance from the actual con-
tact interface, then some thermal modeling may be required to extrapolate the interface temperature.
If the specimen is rotating, then it may again be necessary to carry out the temperature measurement
through slip rings or telemetry instrumentation. Both are a not easy to install but are nonetheless reli-
able once worked out properly. Do not neglect the possibility that a conventional thermocouple with a
loose loop of extension wire can often be applied to a moving test specimen.
Another method for measuring temperature on moving or awkward locations is infrared tempera-
ture sensors. Simple transducers are available that produce analog voltages proportional to temperature.
Some considerations for infrared sensors are as follows. These transducers have a fixed spot size (it varies
sensor vendors) that they average temperature over. Also, the time constant, that is, frequency response
of infrared sensors, can be slower than that of thermocouples. Within these limitations, they can be
useful. There is a broad range of cost for these sensors varying by a factor of 10 or more. More expensive
units can image entire component during an experiment although again pay attention to the frequency
response of the instrument chosen.

66.4.5  Pressure (or Vacuum)


A limited subset of experiments may be required to be tested in either a pressurized or vacuum environ-
ment. Examples include the following:

• Wear testing of Freon replacements in the 1970s; these tests necessitated a pressurized environ-
ment to be successfully accomplished. A high-pressure chamber commonly referred to as a reac-
tor was adapted to contain the test.
• Fundamental wear tests conducted in an SEM environment [7].
• Development and testing of lubricants for space mechanisms conducted in deep vacuum to mimic
conditions on orbit.

In the limited number of cases where vacuum or pressure is required, then the experimenter must tackle
the substantial problem of enveloping the relevant parts of the test in a safely designed pressure vessel.
This will, in many cases, necessitate a specialized tester.
Tribology Testing 66-7

66.4.6  Surface Finish, Geometry, and Cleanliness of Test Specimen


The effect of surface finish and geometry on friction and wear coefficients is documented in technical
papers and journals. Suffice it to say, this is an important parameter that must be properly controlled.
A ground surface will often be the preparation of choice because this will economically bring a metallic
specimen in the range of 5–20 μin. as a general rule. The grinding process will also generally provide a
reliable round or flat geometry. Rolling element bearing surfaces require additional polishing to bring
the typical roughness down to 5 μin. or less. Flatness or roundness of specimens is important because
variations in geometry will invariably affect lubrication modes and wear scar geometries. The small
radius or lubrication groove in a specimen will have to be taken into account in any wear volume calcu-
lations. In general, the experimenter should follow any standard test guidelines closely in the prepara-
tion of specimens.
If the test is one of the actual component, then a surface roughness measurement before testing is
at least called for. There are several options for documenting surface roughness including contacting
(stylus) profilometers or more recently a variety of noncontacting optical instruments for measur-
ing roughness. In most cases, a record of the average roughness (Ra) and perhaps a printout or digi-
tal record of the actual surface profile will be sufficient. In wear testing, it is often possible to take a
digitized profile of the original surface that may be compared to another profile taken after testing to
calculate wear volumes. This technique is most helpful if an unworn area of the specimen remains for
reference since the worn areas removed are quite small. Another, trick-of-the-trade in quantifying
small amounts of wear is to make a microindentation in the surface of a wear test specimen by using a
Knoop hardness indenter. Changes in the small diamond-shaped crater dimensions can be tracked to
quantify mild wear [8].

66.4.7  Replicas
Often, it may be desirable to take a replica or at least a magnified photograph by light microscopy to doc-
ument the surface condition before testing. In mild wear test scenarios, changes in the surface charac-
teristic may not surpass even the original machining marks in the part. If the reader is not familiar with
replicas, these are simply applications of materials such as dental facsimile polymers that will produce a
highly accurate reproduction of the surface. Reproduction capability can be down to the microinch level
if replicas are applied skillfully. Replicas of fine surface features can be made by application of Duco®
Cement that is lifted from the test specimen with Scotch Tape®. These replicas can then be examined by
light microscope or scanning electron microscope.

66.4.8  Cleanliness
It is important to establish and maintain the cleanliness of the specimens under test. Metal speci-
mens in specific have wide variations in their test performance depending on the cleanliness. The
concept involved is that any oxide or material of any sort left on the specimen during machining can
serve as a boundary lubricant and will affect friction. A specimen in its purest form is referred to as
having a nascent or bare metal surface that can normally only be achieved by careful preparation.
Contamination is often introduced by a machining fluid or oil that will leave behind a film. There
are industrial and military standards for cleaning specimens before testing. The general gist of these
protocols is that any rough oils or contaminants can be removed through scrubbing with a detergent
solution. This step is generally followed by a series of washes with various solvents in a specific order.
The final test specimen will then be sufficiently clean as to be immediately susceptible to rust or oxi-
dation, so the specimen must either be cleaned immediately prior to testing or kept in a desiccator to
preserve its condition.
66-8 Design for Lubrication and Tribology

66.4.9  Material Properties


The wear test should use the correct materials of suitable manufacturing method and similar hardness
and roughness when compared to the component being modeled. Details of the material alloy are criti-
cal to tribology tests. For example, specifying a “stainless steel” is not sufficient for a tribology test. The
exact type of steel used in the component as well as its metallographic state must be determined. This
could be done by reference to production drawings for a component, yet even these can be incorrect.
Knowing the material specification, it is still prudent to confirm the material makeup by EDAX or simi-
lar test and at least the surface hardness. Cross-sectioning may be necessary to confirm the interior state
or heat-treat history of a material confidently. The salient point is that tribology test data from material
with an unknown pedigree will be suspect. The materials used in the specimen should closely match
the component modeled. Often, small test samples can be cut from the actual component themselves.

66.4.10  Lubrication Regime or Wear Mode


The experimenter should anticipate the wear or lubrication mode present in the mechanism under test.
Selection of a suitable test requires this be known; few tests allow for evaluation of wear or friction under
widely divergent modes or regimes. Definitions of wear modes and lubrication regimes are documented
elsewhere, so they will not be repeated here. Perhaps less well known, however, are guidelines as to how
to quantitatively evaluate which mode applies. Figure 66.1 is an attempt to diagram some of the con-
siderations that provide clues as to which type of test is applicable. In cases where the components are
dry or unlubricated, this will entail whether the likely modes include adhesive wear or abrasive wear. In
adhesive wear, one should determine if mild wear or more severe forms such as galling are likely. A less
frequent but nonetheless important mode might include fretting.
In lubricated contacts, one should consider whether the interface is likely to be boundary, mixed,
hydrodynamic, or elastohydrodynamic (EHD). Less frequent but important modes for lubricated con-
tacts might include erosion and surface fatigue (for rolling elements). A traditional means to map lubri-
cation modes is a generalized Stribeck curve, as shown in Figure 66.2. The curve is primarily intended
to describe the type of contact occurring in a rolling or sliding lubricated contact. High loads, slows
speeds, nonconformal contacts, high temperature and low viscosities, and marginal or starved lubrica-
tion will produce results in the area labeled as boundary film. In this region lubricant films are on the
molecular level, surface asperities are in full contact and friction coefficients may be 0.05–0.15. Heat
generation and wear can be significant and increase quickly with load in this region.
In Figure 66.2, the following conditions may promote movement into a mixed-film regime:
• If the speed in the lubricated application is higher
• If the load in the application is less, or the contact geometry is more conformal
• If the lubricant viscosity is higher
Under mixed film conditions, the lubricant film begins to separate surfaces by hydrodynamic pressure.
Intermittent contact depends on surface roughness. There is partial hydrodynamic support with a cer-
tain amount of metal-to-metal (or surface-to-surface) contact. Wear and friction decrease as compared
to boundary lubrication, as does temperature generation.
If the conditions that promote mixed film develop sufficiently, then the interface may move into full
hydrodynamic film that provides low friction and low wear. With full hydrodynamic operation, contact
typically only occurs during start-up and stopping.
When evaluating a mechanism that contains rolling elements, EHD may apply. EHD is a subset
of hydrodynamics wherein the contact pressures are sufficiently high as to cause deformations of the
surfaces. Examples of mechanisms that may exhibit EHD include rolling element bearings (as distin-
guished from journal bearings) and gearing of various types. Gearing, however, can tend to move back
into a mixed lubrication mode under less than ideal conditions.
Tribology Testing 66-9

Determine loads, velocities,


temperatures, materials,
contact geometries, and surface
roughnesses in the system

Interface lubricated
with coatings, liquid,
or grease lubricants?

Yes No

Evaluate lubrication
regime: Evaluate wear
boundary, mixed film, mode
hydrodynamic, or
elastohydrodynamic

Hydrodynamic
lubrication
Third body or Erosive or
Lowloads, good Two dry bodies in
abrasive material corrosive liquid in
lubrication (plentiful fluid contact
Journal bearing present? fluid stream
with higher viscosity),
higher speeds, low tester
surface roughnesses

Dry or wet sand


Intermediate loads, Erosion/corrosion
Block-on- rubber wheel
partial or starved impingement test Favorable or
Mixed film ring test, abrader
lubrication, lubrication pin-on- difficult conditions?
intermediate speeds disk test

Boundary High loads,


Light loads,
lubrication poor material match-up,
High loads, good material match-up,
molecular level lubrication, low speeds high speeds
non-conformal contacts,
high surface roughnesses Does severe
No wear or galling Yes
occur?

Extreme pressure tests: Mild adhesive Aggressive wear


Four-ball (constant load), and galling/weld
four-ball ep (increasing load), wear possible
pin-on-disk test, point possible
pin-and-vee,
three-button test
crossed-cylinder

High loads, Pin-on-disk,


Three-button,
concentrated contacts, three-button,
Elastohydrodynamic block-on-ring, or
rolling or converging Bearing testers block-on-ring, or
lubrication application specific
geometry, low application specific
geometries
roughnesses geometries

FIGURE 66.1  Considerations for selecting friction and wear tests.

66.5  Selection of Type of Tester


Once the parameters of interest are known and a reasonable guess (engineering estimate) has been made
as the lubrication or wear mode, then one proceeds to select a tester. Two major directions here will be
either selection of a standard tester or construction of a custom tester.

66.5.1  Standard Tests


A standard test will likely be one covered by ASTM standards. The ASTM, parts 23, 24, and 25, con-
tains descriptions of (hundreds of) standardized tests related to petroleum products and lubricants.
66-10 Design for Lubrication and Tribology

1
Absence of
boundary
lubricants

Coefficient of friction
0.1

0.01

Boundary Mixed Hydrodynamic


0.001
zn/P

FIGURE 66.2  Conceptual map of coefficient of friction in various lubrication modes.

An ASTM test will typically include descriptions of scope of the test, summary of method, apparatus,
procedures, calculations, precision, and a line drawing or photograph of the apparatus.
The advantages of selecting one of these tests will include the following:
1. A test plan will have already been carefully documented in the ASTM description of test. The test
will have essentially been debugged by others that have gone before, and thus, the expected results
will be reasonably predictable.
2. Data may exist in the literature that can be readily compared to your test results.
3. When results are published, others will be familiar with the test protocol and what the results
imply. The results may carry an air of respectability since they were conducted “in accordance
with” a standard.
One of the most telling drawbacks to the ASTM tests is a statement that is included in the introduction
of many of the standards. This caveat says “Significance … No correlation has been established between
this test method and service.” Unless one is engaged in scholarly pursuits, the very point of conducting
tests is to obtain information that is directly related to service of the component being tested. This may
motivate one to undertake or employ a custom apparatus.

66.5.2  Custom Apparatus


If careful research into the subject, both literature and standards, reveals that no existing test will ade-
quately model the component under evaluation, then one may be persuaded that a custom test apparatus
is required. The custom apparatus has the following elements in its “learning curve”:
1. A test plan/apparatus design will have to be written. If not done by a seasoned campaigner, this
process can be fraught with delays and budget overruns. In short, tribology tester development
can be harder than anticipated. Friction is described in introductory engineering as a coefficient
whose value is essentially a constant for a given material combination that does not vary with
load, temperature, velocity, or surface condition. In practice, one finds that friction is a complex
quantity inextricably connected to all of the aforementioned.
2. Data produced by a custom apparatus are specific to that apparatus. It can be compared to other
data only if results are normalized by some recognized parameter such as a wear coefficient, PV
limit, or fatigue life.
A great advantage of a custom apparatus is that data will likely be significant to your component and
have a strong correlation with service. There are institutions and researchers that have dedicated their
Tribology Testing 66-11

careers to certain types of tribology tests. To avoid an appearance of commercialism, names will be
left out; however, to be practical, a diligent inquiry will reveal specialized practitioners of certain tri-
bology tests. Examples of specialized areas include friction of plastics or friction of compact disk and
tape storage media. Sources for these custom testers will be research institutions, universities, national
laboratories, laboratories of armed forces branches, NASA, NIST (National Institute of Standards and
Technology; formerly National Bureau of Standards), and research laboratories of commercial compa-
nies. Once an avid experimentalist in one of these resources is identified, they are typically willing to
arrange testing since it furthers their particular field of interest.

66.5.3  Modification of Existing Apparatus


To increase your likelihood of success, one may decide to either modify one of the standard tests or
adapt an actual component and incorporate it into a tester. Modification of a standard tester can be
relatively straightforward. One might replace a small DC motor and belt drive with an AC motor that
achieves much higher horsepower and speed. Another common modification is to take a standard tester
and introduce a load mechanism to greatly increase its range. One might switch a small dead weight out
for a hydraulic cylinder, for example. Modification of an existing tester can reduce your learning curve
and increase likelihood of a successful experiment if one primarily needs to extend the range of one or
two test parameters.

66.5.4  Conversion/Incorporation of Actual Component into Tester


A manufacturer may often have a ready supply of an existing component that is under development
or is in service and has tribological failures. Such a component can be converted into a pertinent tri-
bological tester. Indeed, many of the ASTM standards are pieces of commercially available equipment
that has been used sufficiently to be embodied into a standard. If one decides to use an existing com-
ponent as the test bed, then the component will need to be mounted sturdily and external utilities
such as loads, motions, etc., applied to mimic actual service. The notes in general from the parameter
discussions earlier may give clues as to how this may be done. The apparatus itself can become a test
bed. Good practice should be applied to predict deflections of members under load. Fatigue in support
bearings or shafts can often occur and disrupt a test. Premature wear of connection points in the drive-
train can experience wear failures. A substantial margin is recommended when selecting and design-
ing the apparatus, so the experimenter can concentrate on the test of interest rather than the apparatus
itself becoming a project.

66.6  Metallurgical Techniques Supporting Tribology Testing


Several metallurgical techniques are used in conjunction with tribology testing. Examples include
hardness measurements, scanning electron microscopy (SEM), and particle analysis by energy disper-
sion and x-ray analysis (EDAX). A tribologist typically requires access to a metallurgical laboratory.
Photographs by light microscopy at low magnifications are a good method to document a wear speci-
men before and after a test and compare wear features of bench test specimens with that of the compo-
nent from field service. Additionally, it is sometimes helpful to take a worn part, particularly if testing
an actual component, into an SEM. Under the high magnification of the SEM, the type of wear that is
occurring can be more readily evident. Also, if the SEM is equipped with EDAX, then small individual
wear particles can be identified by their elemental constituents. Hardness measurements are typically
done to provide information needed to calculate wear coefficients. Most wear models involve the hard-
ness of the mating materials.
66-12 Design for Lubrication and Tribology

66.7  Tribology Test (Other than Friction and Wear Testing)


It may be appropriate to mention some types of tribological test that are important yet will be consid-
ered out of scope for this chapter.
Tribology testing may include the following:
• Lubricant properties, viscosity, pressure viscosity, and temperature viscosity
• Lubricant analysis, specifically in-service analysis
• Traction testing
• Grease testing
• Bearing testing, including fatigue life
• Materials testing related to mechanical properties
• Corrosion testing
• EHD film thickness measurement

66.7.1  Lubricant Properties


Literature is generally available for commercially available lubricating oil in the form of data sheets. The
data sheet includes data for viscosity at several temperatures as other commonly used parameters. These
general data are often most economically gathered by standard testing laboratories where the efficiency
of the test is improved by volume of tests conducted. Providers of lubricants will often sponsor and
provide these basic data.
Some design efforts may require data that are more exotic and not so broadly available. Information
of this type may be generated by custom test apparatus operated by the manufacturer if the need is suf-
ficiently recurrent. An example would the pressure-viscosity or temperature-viscosity coefficient for a
lubricant. The Air Force sponsored research for development of much lubricant data in the 1960s and
1970s. The data are specifically useful in the design of turbine engines (bearings) and other EHD inter-
faces. An inherent difficulty is that such industry specific data will be difficult to locate. Universities,
technical organizations such as NASA, and laboratories for Army, Air Force, and Navy can all be sources
for public domain information on some data.

66.7.2  Lubricant Analysis


There is an entire branch of testing related to testing of lubricants in service to monitor their perfor-
mance and degradation. These tests are categorized under condition monitoring or lubricant analysis.
These tests may include particle analysis (suspended contaminants and wear debris), particle counts that
are typically reported as distributions ranked by particle size, entrained water, oxidation analysis, and
analysis for decomposition products including additives. These tests are typically conducted by labo-
ratories dedicated to this purpose. Samples of lubricants are taken at routine intervals and shipped in
vials to companies who provide standard analyses at competitive prices. This type of data is often evalu-
ated to determine if either the lubricant itself or the mechanism of interest is experiencing tribological
breakdown. A simple example would be to follow changes in metallic particle count and size distribu-
tion to detect excessive bearing wear. By tracking oil analysis sample results over the life of a particular
machine, trends can be established that can help eliminate costly repairs or maintenance. A common
use of the analysis is to determine the proper interval at which the oil supply must be replaced. Although
lubricant analysis is an important type of tribology test, it is not a friction or wear test in its own right.

66.7.3  Traction Testing


Generation of slip-traction curves is of particular interest for transmission fluids. Slip-traction behavior
is often significant in a rolling interface also; however, the sliding may be masked within the contact
Tribology Testing 66-13

zone. Many people assume that a rolling element bearing experiences pure rolling, but in fact realities
of geometry make pure rolling across an interface almost impossible. The amount of sliding in a “rolling
contact” can generate heat that will reduce the lubricant viscosity and affect (lower) the film thickness at
the interface. In mechanisms such as transmissions and clutches, there is intentional gross slip between
the components in some regimes of operation. Normally in these mechanisms, as the normal force is
increased, the gross slip is reduced, that is, the transmitted force across the interface also increases and
thus generates the traction coefficient of these fluids. The slip-traction behavior is important both to
standard lubricants (oils) in EHD elements such as rolling element bearings or gears as well as traction
fluids designed for clutches. The generation of a slip-traction curve for a fluid normally involves an appa-
ratus wherein two cylinders are initially rolling with aligned axis (zero slip). The axis of one cylinder is
then systematically skewed, which introduces sideslip to the interface; two cylinders at right angles have
a slip ratio that approaches infinity (i.e., pure sliding). The same test has also been conducted by simply
varying the speed of one cylinder relative to the other. This approach normally requires a healthy sized
motor to force the slip condition. Slip-traction tests, in summarizing, are important tribological data
but tend to be generated with custom apparatus, by a limited number of facilities.

66.7.4  Grease Testing


Grease-lubricated bearings are common because of the low cost and simplicity if one can avoid the addi-
tional complexity of a liquid lubricant system. Since greases are in common use, there is a battery of tests
available that can well be considered tribology test. The ASTM standards [6] cover these tests in detail,
and so there is little value in repeating them here. A few examples are as follows:
• Cone Penetration of Lubricating Grease (working penetration) ASTM D217
• Dropping Point of Lubricating Grease ASTM D566
• Extreme Pressure ASTM D 2596 Four-Ball
• Anti-wear ASTM D 2509 326 Timken Method
• Four-Ball Wear ASTM D 2266 239
• Oxidation Resistance ASTM D 942 142 Bomb Oxidation
• Wheel Bearing Life ASTM D 3527
• Apparent Viscosity ASTM D 1092 At 16 shear rates
• Leakage ASTM D 1263 Wheel Bearing Leakage
The data sheet for any commercial grease often contains results from such standard tests.

66.7.5  Bearing Fatigue Life Testing


Another subset of tribology testing is the calculation of fatigue life of rolling element bearings. This is
the basis for all load ratings presented for bearings offered in manufactures catalogs. Although this is
not friction or wear testing, it is certainly fundamental tribology testing. For reference on this subject,
STLE Life Factors for Rolling Bearings [9] covers the subject in good detail and has references to the
many original texts and researchers for fatigue life calculations.

66.7.6  Corrosion Testing


Wear and corrosion can involve synergistically combine and result in significant mutual interactions
beyond the individual contributions of either. Corrosive wear can be induced by either liquid solutions
or gaseous atmospheres.
Although there are many tests for wear, and many tests for corrosion, there are fewer tests for cor-
rosive wear. ASTM G119-09, Standard Guide for Determining Synergism Between Wear and Corrosion
[10] is one reference.
66-14 Design for Lubrication and Tribology

Experience has shown that a custom apparatus that provides the atmosphere expected in ser-
vice is a good approach to corrosion testing. This can often be accomplished more expeditiously by
starting with a standard wear tester and providing a chamber for containment of a corrosive atmo-
sphere. Examples include the extensive work on the development of tribology tests for the adiabatic
diesel circa 1980. Also, testing in seawater or in a pure deionized water environment is an applica-
tion that can yield interesting challenges. These are just representative examples of corrosive test
environments.

66.7.7  EHD Film Thickness


In the 1960s, there was a considerable amount of experimentation done to support the development
of modern EHD theory. Much of this work was done with a twin disk apparatus or ball on disk. These
types of apparatus have some mechanism for measurement of film thickness between the rotating ele-
ments, whether by electrical resistance measurements, x-ray techniques, or optical fringe techniques.
This field of study produced vitally important data for development of EHD modeling. Since this field is
now mature, less of this type of research and testing is currently pursued.

66.8  Design of Experiment


It is good practice for the design of experiment to be formally written up. This process forces the experi-
menter to document aspects that may otherwise be missed. A formal test description can be reviewed
by all shareholders to determine if the test will produce the desired result. The design of experiment can
include elements such as the following:
1. Test apparatus—a hardware description, conceptual sketch of test apparatus, and drawings and
photographs of the apparatus if available.
2. Test specimen description—released drawings if available with callouts for dimensions, toler-
ances, materials, surface roughness, and process notes such as hardening or coating steps. If for-
mal drawings are not available, then even pencil sketches are desirable.
3. Material and process certifications for the test specimen materials.
4. Test instrumentation/data requirements—description of all instrumentation to be used including
instrumentation precision and accuracy. Calibration certifications for each instrument, typically
traceable to NIST, can be appended to the experiment report.
5. Test parameters and tolerances—ranges for parameter values. There is a form published by
Czichos [11] that can be useful for organizing parameter information. A test matrix can be an
alternate form for organizing parameter values for each test. This include duration and stopping
or failure criteria for tests.
6. Test procedure—detailed steps of assembly, test conduction, and disassembly.
7. Test records—a logbook should be maintained during each test run and can include a record
of start and stop times, cycles accumulated, maintenance performed on the test apparatus,
and any other pertinent information related to the test. Data sheets can also be used to track
experiments and results; however, it seems there is always some unanticipated detail that can
be recorded in a logbook that may eventually be useful in reporting. Data sheets can be used in
conjunction with a logbook. A logbook can be a comforting backup to computer data record-
ing system.
8. Test report—at the conclusion of a test series, the recorded data can be summarized in report
form and can contain all the aforementioned plus
a. Photograph of specimens before and after
b. Summary and plots of recorded test parameters generally versus time or number of cycles
c. Any test malfunctions resulting in test interruption or complete test abort
Tribology Testing 66-15

The experimenter should always log unusual events during a test. Events that seem anomalous at the
time can be important when summarizing and considering the data at a later time and may eventually
lead to inventions and breakthroughs.

66.9  Break-In or Wear-In Phenomena


Friction and wear tests may behave differently if the test is preceded by, or introduced with, a break-in
period where the load, speed, etc., are reduced for an initial break-in period. The material couple in the test
can sometimes exhibit superior performance (lower friction wear, or higher PV values) if first allowed to
break in. This phenomenon is probably associated with mild polishing at the interface, which can lower the
surface roughness (polish) and eliminate any unintentional high spots or sharp edges in the test specimens.
Even though gross alignment is, and should be, done carefully for a test, a break-in procedure can allow
the specimens to polish until a high degree of conformity to the intended geometry is achieved. To some
degree, this is rather cheating the test. A good guideline may be that if a break-in procedure will be allowed
for the final intended design, then perhaps a break-in period for the test is allowable. If, in the intended
design the parts are likely to be immediately run under full load and exclude development of a favorable
surface finish, then it is a better simulation to run the wear test under similar challenging conditions.

66.10  Accelerated Testing


The conduction of an accelerated test is an area of engineering expertise in itself. Acceleration of general
types of testing is generally done by increasing the level of one of the test parameters such as load, speed,
or temperature. For tribology testing, this may not be practical. Increasing any one of these parameters
may shift the wear or lubrication mode. If speed is increased, the interface may end up in a hydrody-
namic region when it was intended to study boundary lubrication wear. Also, with increased speed,
excessive heat can build up and degrade the lubricant or the material specimen through softening or
tempering. Many examples could be given, and the end result is that although accelerated testing models
can be valid for some types of testing, it is not always necessary or expedient for friction and wear tests.

66.11  Statistical Analysis of Friction and Wear Data


Friction and wear data can sometimes carry a reputation for high standard deviations. The problem can
be discussed as both deviation of values for a given test and deviation when comparing data from different
sources. As discussed earlier, the values measured are quite sensitive to control of the test parameters. Good
consistency is probable with similar specimens on the same apparatus if test parameters are under control. In
this case, a repetition of several trials and conduction of standard statistical evaluations generally improves
the accuracy of reported results. Since tribological tests can be expensive, particularly if the test runs for an
extended period (days or weeks), a few repetitions may suffice if the phenomena are deterministic.
In using handbooks, a single coefficient-of-friction or wear coefficient number may be sought. This is
one condition where wide ranges in reported values may occur. Complete friction and wear data from a
tribology test will include initial (break-in), generally reach some plateau value in the middle of the test,
and perhaps attain a higher value as the test enters a more severe mode or stopping criteria. The com-
plete description of the conditions under which a single value was obtained is lost if the result is distilled
down to single value coefficients. Thus, the reputation for wide variations in tribology data, although
undeserved, may persist.

66.12  Role of Modeling in Tribology Testing


Analytical modeling can be important in tribology testing in two different ways. The models can be
used to design the experiment or to interpret the data. A few examples will show what models can be
66-16 Design for Lubrication and Tribology

applied. As a general comment, experiment data are much more generally useful if some model of fun-
damental processes is used.
Models as simple as the calculation of a Sommerfeld number can be predictive for experiments. There
are general rules for what may be expected from an interface based on such calculations. A reasonable
expectation of the transition from mixed to full film lubrication can be achieved as an example.
Generalized models for lubricant film thickness are readily available from many sources [2,12].
Heat transfer models are useful for both designing and interpreting test results. Heat transfer models
are often more easily obtainable now that finite element analysis or models are commonly connected
with drawing software (CAD) packages.
Another type of modeling often applied to tribology testing is the calculation of wear coefficients.
Reduction of test results to a wear coefficient allows comparison of data between tests.
After any model has been verified by testing its validity against real world data, the model can be used
for parametric studies. If an adequate model of the experiment is achieved, then the experimenter can
explore by exercising the model through a range of possible parameters. The experimenter must keep
in mind that the model prediction or extrapolation generally becomes more reliable when applied to
parameter ranges covered in experiments and less conclusive as extended to ranges not tested.

References
1. Bayer, R.G., Mechanical Wear Fundamentals and Testing, 2nd edn., Marcel Dekker, New York, 2004.
2. Peterson, M. and Winer, W. (eds.), Wear Control Handbook, ASME, New York, 1981.
3. Booser, E.R., Tribology Data Handbook, CRC Press, Boca Raton, FL, 1997.
4. Blau, P.J. (ed.), ASM Handbook: Volume 18: Friction, Lubrication, and Wear Technology, ASM
International, Materials Park, OH, 1992.
5. Neale, M.J. (ed), Tribology Handbook, Elsevier, Amsterdam, the Netherlands, 2003.
6. ASTM Standards, Petroleum Products and Lubricants, Vol. 5.01 (I): D 56-D 3348, Vol. 5.02 (II): D
3427-D, Vol. 5.03 (III): D 5769-D, and Vol. 5.04 (IV): D 6730–latest, ASTM International, West
Conshohocken, PA.
7. Glaeser, W.G., Wear experiments in the scanning electron microscope, Wear, 73, 1981, 371–378.
8. Glaeser, W.G., Wear measurement technique using surface replication, Wear, 40, 1976, 135–137.
9. Zaretsky, E.V. (ed.), Life Factors for Rolling Bearings, STLE, Park Ridge, IL, 1992.
10. ASTM Standard G119-09, Standard Guide for Determining Synergism between Wear and Corrosion,
ASTM International, West Conshohocken, PA.
11. Czichos, H., A systems analysis data sheet for friction and wear tests and an outline for simulative
testing, Wear, 41, 1977, 45.
12. Hamrock, B.J., Lubrication of Machine Elements, NASA Reference Publication 1126, August 1984.
Tribology/Lubrication Engineering

materials and treatments, which are currently the


fastest growing areas of tribology, with announce-
ments of new coatings, better performance, and
new vendors being made every month. The final
section presents components, equipment, and
designs commonly found in tribological systems.
VOLUME II It also examines specific industrial areas and
Theory an d D e s i g n their processes.

Features
• Contains an abundance of material
Since the publication of the best-selling first relevant to today’s technology needs
edition, the growing price and environmental • Includes peer-reviewed contributions
cost of energy have increased the significance from leading authorities in lubrication
of tribology. Handbook of Lubrication and and tribology
Tribology, Volume II: Theory and Design, • Presents critical data for the design
Second Edition demonstrates how the principles of lubrication and tribological-related
of tribology can address cost savings, energy equipment
conservation, and environmental protection. This • Explores wear materials, treatments, and
second edition provides a thorough treatment of coatings—some of the most burgeoning
established knowledge and practices, along with areas of tribology
detailed references for further study.
Sponsored by the Society of Tribologists and
Written by the foremost experts in the field, the Lubrication Engineers, this handbook incorporates
book is divided into four sections. The first reviews up-to-date, peer-reviewed information for tackling
the basic principles of tribology, wear mechanisms, tribological problems and improving lubricants
and modes of lubrication. The second section and tribological systems. The book shows how
covers the full range of lubricants/coolants, the proper use of generally accepted tribological
including mineral oil, synthetic fluids, and water- practices can save money, conserve energy, and
based fluids. In the third section, the contributors protect the environment.
describe many wear- and friction-reducing

6908X
ISBN: 978-1-4200-6908-2
90000

9 781420 069082

You might also like