0% found this document useful (0 votes)
713 views

Estática Kassimali

Engineering Mechanics. Statics -KassimallI TOMO 1 Inglés

Uploaded by

Jhoel Cc
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
0% found this document useful (0 votes)
713 views

Estática Kassimali

Engineering Mechanics. Statics -KassimallI TOMO 1 Inglés

Uploaded by

Jhoel Cc
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/ 600

EN

MECHANICS

Das Kassimali Sami


U.S. Customary Units and Their SI Equivalents
Quantity U.S. Customary Unit SI Equivalent

Acceleration ft/s? 0.3048 m/s?

Angular acceleration rad/s? 1.0 rad/s?

Angular velocity rad/s 1.0 rad/s

Area ne 0.0929 m?

Energy 1.356 J

Force 4.448 kN
4.448 N

Impulse 4.448N-s

Length 0.3048 m
25.4 mm

Mass lb mass 0.4536 kg


oz mass 28.35 g
slug (lb-s?/ft) 14.594 kg
ton (2000lbs) 907.18 kg
Moment Ib + ft © 1.356 N-m

Moment of inertia of in* 0.4162 x 10°mm*


an area = 0.4162 x 10m’?

Moment of inertia of lb - ft - s? 1.356 kg - m?


amass

Power lb « ft/s 1.356 W

Pressure lb/ft? 47.88 Pa


Ib/in? 6.895 kPa

Velocity ft/s 0.3048 m/s

Volume (liquid) gal 3.785. L

Volume (solid) ft® 0.02832 m*°


in? 16.39 cm?

Work ft - Ib 1.3558 J
kw-h 3.60 x 10°J
SI Units and Their U.S. Customary Equivalents
Quantity SI Unit U.S. Customary Unit

Acceleration m/s? 3.281 ft/s?

Angular acceleration rad/s? rad/s?

Angular velocity rad/s rad/s

Area 10.764 ft?

Energy 0.7375 lb - ft

Force 0.22482 Ib
0.22482 kip
Impulse 0.22482 Ib-s

Length 3.281 ft
0.6215 mi

Mass 0.06854 slug


Moment 0.7375 Ib - ft
Moment of inertia of 8.63 x 10° ft*
an area

Moment of inertia of 0.7375 Ib - ft - s*


a mass

Power 0.7375 lb «ft/s


Pressure 1.45 x 10“Ib/in?
Velocity 3.281 ft/s
Volume (liquid) 0.2642 gal
Volume (solid) 35.311 ft?
cm? 0.061 in®

Work 0.7375 lb - ft
Digitized by the Internet Archive
In 2022 with funding from
Kahle/Austin Foundation

httos://archive.org/details/engineeringmecha0000dasb_g8q3
ENGINEERING
MECHANICS
iy
ENGINEERING
MECHANICS

Braja M. Das
Asiam Kassimali
Sedat Sami

Department of Civil Engineering and Mechanics


Southern Illinois University at Carbondale

IRWIN
Burr Ridge, Illinois
Boston, Massachusetts
Sydney, Australia
© RICHARD D. IRWIN, INC., 1994
All rights reserved. No part of this publication may be
reproduced, stored in a retrieval system, or transmitted, in
any form or by any means, electronic, mechanical,
photocopying, recording, or otherwise, without the prior
written permission of the publisher.

© Copyright 1987, Wayne Coble, Coble Studios, Inc.

Associate editor: Kelley Butcher


Marketing manager: Robb Linsky
Production manager: Ann Cassady
Designer: Laurie Entringer
Art manager: Kim Meriwether
Compositor: CRWaldman Graphic Communications
Typeface: 10/12 Times Roman
Printer: R. R. Donnelley & Sons Company

Library of Congress Cataloging-in-Publication Data

Das, Braja M., 1941-


Engineering mechanics : statics / Braja M. Das, Aslam Kassimali,
Sedat Sami.
p. cm.
Includes index.
ISBN 0-256-11452-8 (5.25” Version) ISBN 0-256-11453-6 (3.5” Version)
1. Mechanics, Applied. 2. Statics. I. Kassimali, Aslam.
II. Sami, Sedat. III. Title.
TA350.D36 1993
620.1'03—dc20 91—42975
Printed in the United States of America

1 2:3 4 36788190 DOES 0) 9° 8 716558453


PREFACE

Engineering mechanics, concerned with the study of equi- deal with two-dimensional analysis using the scalar formu-
librium and motion of rigid and elastic bodies, is regarded lation. In Chapter 3, the basic concepts are explained in de-
as essential to the basic education of an engineer. Since the tail using both the scalar (intuitive) as well as the vector
problems confronted by today’s engineers are seldom re- approaches. The chapter contains a large number of solved
stricted to one’s own specialization, it is imperative that the examples as well as unsolved problems, with the goal of
engineering student become thoroughly grounded in the fun- preparing the students for upper-level courses in the areas
damental principles of mechanics so necessary for the so- of mechanics of materials, and analysis and design of struc-
lution of many problems. A major objective of the authors tures and machines. In Chapter 4, however, the vector ap-
has been to present, in a coherent and systematic fashion and proach is emphasized for solving three-dimensional prob-
by emphasizing the useful application, a fundamental treat- lems. Distributed forces (Chapter 5) are covered before the
ment of the principles of mechanics. It was particularly im- analysis of trusses, frames, and machines (Chapter 6), thus
portant to illustrate the application of these principles to enabling Chapter 6 to include examples and problems deal-
problems encountered in various fields of engineering. ing with the analysis of structures subjected to distributed
We have paid special attention to the degree of clarity loads. The International System of Units (SI) and the U.S.
that should be within the grasp of an average sophomore Customary System of Units (USCS) have both been used
with prior knowledge of algebra, geometry, trigonometry, throughout the problems at the end of each chapter.
and calculus. The examples offered in the book are repre- This book can be covered in a three-semester-hour
sentative applications of the fundamental principles devel- course. However, at the discretion of the instructor, certain
oped previously. The problems have been designed with the sections of the first nine chapters and all of Chapter 10—
goals of familiarizing the student with real-life problems and indicated with an asterisk (*)—could be omitted without
developing in them an appreciation for their own powers of loss of continuity.
analysis and the effective use of mathematical modeling. Each chapter has a brief chapter outline and introduction,
This volume, the subject of which is statics—the equi- many illustrations, and a summary section. To illustrate the
librium of bodies—is divided into 10 chapters and an ap- application of methods and equations developed in the text,
pendix. The organization of subject matter may be consid- there is an abundance of worked examples in each chapter
ered somewhat unconventional. The concept of coplanar (over 140 total). Example solutions are detailed and clearly
forces and the equilibrium of two-dimensional rigid bodies explained for maximum understanding. Homework prob-
(Chapter 3) is covered in a separate chapter from that dealing lems appear at the end of each chapter. These problem sets
with the three-dimensional force systems and the equilib- represent a wide range of applications and progress from
rium of rigid bodies subjected to such forces (Chapter 4). simple to more complex. There are approximately 900
Traditionally, the two- and three-dimensional cases have homework problems in all.
been covered together using vector notations. However, the The solutions to all of the examples and homework prob-
authors believe that too much reliance on vector notations lems have been double checked by two independent accu-
for two-dimensional problems encourages a ‘‘plug-and- racy checkers. We have taken every effort to provide you
chug’’ attitude among students and deprives them of the with an error-free book. Any remaining errors can be
intuitive sense necessary to understand some important basic brought to the attention of the authors for correction in future
concepts such as the moment of a force, couple, and equi- editions. Special thanks to James Matthews, Jeff Filler, and
librium. Furthermore, the majority of upper-level engineer- Charles Atz for their diligent efforts at finding errors.
ing courses, of which the mechanics courses are prerequisite, The Instructor’s Manual, available from the publisher to

Vv
vi Preface

adopters of the text, contains complete typeset solutions to editor, Bill Stenquist, and Kelley Butcher, associate editor,
the homework problems. Transparency masters of important of Richard D. Irwin, Inc., for their constant help and support
figures and examples are also available from the publisher. during the period of the development of the manuscript. In
It is almost impossible to list and give proper credit to all addition, the manuscript was class tested at Valparaiso Uni-
those who have aided in the preparation of this book. How- versity by Professor Kassim Tarhini. We would like to thank
ever, the authors are especially indebted to Janice Das and him and his students for trying the manuscript in the class-
Maureen Kassimali for their tireless efforts in preparing the room and for providing valuable feedback to help us shape
manuscript for publication. Thanks are due to our former the final draft of the text. The authors are indebted to:

Philip J. Guichelaar Donald A. Grant Gholam H. Nazari J. Edward Anderson


Western Michigan University University of Maine North Dakota State Boston University
Kenneth Oster at Orono University Han-Chin Wu
University of Missouri William W. Predebon Ali E. Engin — University of Iowa
at Rolla Michigan Technological The Ohio State University W. F. Swinson
Paul C. Chan University Roberto A. Osegueda Oakridge National Lab
New Jersey Institute Robert J. Schultz University of Texas Thalia Anagnos
of Technology Oregon State University at El Paso San Jose State University
James R. Matthews Richard J. Leuba M. A. M. Torkamani Carl Vilmann
University of New Mexico North Carolina State University of Pittsburgh Michigan Technological
Richard B. Reimer University D. Joanne Wilson University
University of California William L. Bingham University of Wisconsin David Oglesby
at Berkeley North Carolina State at Platteville University of Missouri
James F. Devine University Tury L. Maytin at Rolla
University of South Florida Kenneth R. Johnson Clarkson University
George G. Adams Indiana University, Purdue Ghassan Tarakji
Northeastern University University at Ft. Wayne San Francisco State
John C. McWhorter John W. Bird University
Mississippi State University University of Nevada at Reno

for their helpful comments, suggestions, and critical reviews


of the manuscript.
Braja M. Das
Aslam Kassimali
Sedat Sami
CONTENTS

List of Symbols xi

INTRODUCTION TO RIGID-BODY MECHANICS 1 2.5 Rectangular Components of a Force 27


1.1 Introduction 2 2.6 Unit Vectors 29
1.2 Historical Background 2 2.7 Resultant of Several Forces from Their Rectangular
1.3. Basic Concepts 3 Components 3/
1.4 Fundamental Principles 4 2.8 Equilibrium of Particles 33
1.5 Scalars and Vectors 6 2.9 Free-Body Diagrams 35
1.6 Units of Measurement 8 2.10 Common Types of Connections 38
1.7. Accuracy of Numerical Calculations J/5 (Concurrent Forces in Three Dimensions)

pees: Cl roblen Solving a> 2.11 Rectangular Components of a Force in Space 40


a 2.12 Direction Cosines of a Force Vector 42
Gy Tens” ue 2.13 Representation of a Force Using Unit Vectors in
Problems 16 Three Dimensions 45
2.14 Representation of a Force Vector with Known
FORCE AND EQUILIBRIUM OF PARTICLES 78 Magnitude and Line of Action 47
2.1 Introduction 19 2.15 Resultant of Concurrent Forces on a Particle
(Concurrent Forces in Two Dimensions) in Three Dimensions 49
2.2 Resultant of Two Forces 1/9 2.16 Equilibrium of a Particle in Three Dimensions 5/
2.3. Resultant of Several Concurrent Forces on a 2.17 Summary 52
Particle 23 Key Terms 53
2.4 Components of a Force 26 Problems 54

vii
viii Contents

3 EQUILIBRIUM OF RIGID BODIES (Equilibrium in Three Dimensions)


IN TWO DIMENSIONS 7g 4.7 Equilibrium of Rigid Bodies 1/94
3.1 Introduction 79 4.8 Types of Supports and Connections 194
3.2 Rigid Bodies 79 4.9 Statically Determinate Structures 197
3.3. Forces on Rigid Bodies and the Principle of 4.10 Free-Body Diagrams 201
Transmissibility 5&0
4.11 Procedure for Analysis of Reactions 20/
(Coplanar Force Systems) 4.12 Summary 209
3.4 Rigid Bodies in Two Dimensions 6&2 Key Terms 2//
3.5 Moment of a Force—Scalar Approach 6&3 Problems 2//
3.6 Cross Product of Vectors 87
3.7 Moments as Vectors 90 CENTER OF GRAVITY, CENTROID,
AND DISTRIBUTED FORCE 227
3.8 Varignon’s Theorem 93
Sal Introduction 228
3.9 Procedure for Determining Moments 94
5.2. Centroid—Definition 230
3.10 Couples 100
3.11 Resolution of a Force into a Force-Couple (Two-Dimensional Problems—Areas and Lines)
System 106 5.3. Centroid of an Area by Integration 233
3.12 Resultants of Nonconcurrent Coplanar Force 5.4 Centroid of a Line by Integration 238
Systems //0
5.5 Centroid of Composite Areas 239
(Equilibrium in Two Dimensions) 5.6 Centroid of Composite Lines 243
3.13 Equilibrium of Rigid Bodies 1/9 *5.7 Theorems of Pappus and Guldinus 244
3.14 Equilibrium Equations in Scalar Form 1/9 *5.8 Distributed Load on Beams 248
3.15 Equilibrium of Rigid Bodies Subjected to Forces at *5.9 Hydrostatic Force on Submerged Surfaces 250
Two and Three Points /23
(Three-Dimensional Problems—Volumes)
3.16 Types of Supports and Connections 1/25
5.10 Centroid by Integration 255
3.17 Statically Determinate Structures /27
5.11 Centroid of Composite Volumes 259
3.18 Free-Body Diagrams 1/34
5.12 Summary 263
3.19 Procedure for Analysis of Reactions 1/38
Key Terms 264
3.20 Summary /46
Problems 265
Key Terms 1/48
Problems 1/49
ANALYSIS OF STATICALLY DETERMINATE
STRUCTURES 286
6.1 Introduction 287
4 EQUILIBRIUM OF RIGID BODIES
INTHREE DIMENSIONS 763 6.2 Internal Forces at Connections 287
6.3. Trusses 294
4.1 Introduction 1/64
6.4 Assumptions for the Analysis of Trusses 296
(Three-Dimensional Force Systems)
6.5 Arrangement of Members of Simple Plane
4.2 Moment of a Force about a Point /64 Trusses 298
4.3 Dot Product of Vectors 170 6.6 Statically Determinate Plane Trusses 300
4.4 Moment of a Force about an Axis 173 6.7 Analysis of Plane Trusses by the Method of
4.5 Couples 177 Joints 303
4.6 — Resultants of Nonconcurrent Three-Dimensional 6.8 Analysis of Plane Trusses by the Method of
Force Systems /8/ Sections 3/5
Contents

O19) Space Trusses 324 9.3 Radius of Gyration of an Area 470


6.10 Frames and Machines 330 9.4 Polar Moment of Inertia and Radius of Gyration 472
6.11 Summary 35/ 9.5 Parallel Axis Theorem for Moment of Inertia
Key Terms 352 of an Area 475
Problems 352 9.6 Moment of Inertia of Composite Areas 479
*9.7 Product of Inertia of an Area 485
DISTRIBUTED LOAD—ANALYSIS OF BEAMS *9.8 Product of Inertia of Composite Areas 489
AND CABLES 369 *9.9 Principal Moment of Inertia of an Area 492
7.1 Introduction 370 *9.10 Mohr’s Circle for Moment of Inertia of an Area 496
(Beams) 9.11 Moment of Inertia and Radius of Gyration of
7.2. Types of Beams 370 Masses 500

7.3. Internal Forces 372 9.12 Mass Moment of Inertia by Integration 503
7.4 Types of Loads ona Beam 373 9.13 Mass Moment of Inertia of Composite Bodies 508
7.5 Shear and Moment Diagrams 373 9.14 Summary 5//

7.6 Relations between Distributed Load, Shear Force,


Key Terms 5/3
and Bending Moment 378 Problems 5/4

(Cables)
*7.7 Cable Carrying Concentrated Loads 384 10 WORK AND ENERGY 927

*7.8 Cable Carrying Distributed Loads 389 *10.1 Introduction 528

*7.9 Cable Subjected to Its Own Weight 394 *10.2 Basic Concept of Work 528
7.10 Summary 397 *10.3 The Principle of Virtual Work 532
Key Terms 398 *10.4 Procedure for Analysis 534
Problems 398 *10.5 Potential Energy 538
*10.6 Equilibrium—Principle of Stationary Potential
Energy 542
FRICTION 477
8.1 Introduction 4/2 *10.7 Stability of Equilibrium—Principle of Minimum
Potential Energy 545
8.2 Mechanics of Dry Friction 4/2
*10.8 Summary 548
8.3 Analysis of Some Dry Friction Problems 4/5
Key Terms 549
8.4 Wedges 424
Problems 549
8.5 Square-Threaded Screws 428
8.6 Belt Friction 43/
*8.7 Frictional Resistance on Thrust Bearings—Disc Appendix 555
Friction 436 A SI Prefixes 555
*8.8 Journal Bearings 438 Conversion Factors 556
*8.9 Rolling Resistance 440 Specific Weight of Common Materials Spy
8.10 Summary 44/ Mathematical Expressions 558
Key Terms 443 Properties of Areas and Homogeneous Bodies
moaw S61
Problems 443

Answers to Selected Problems 568


MOMENT OF INERTIA 462
of Introduction 463
O22 Moment of Inertia of an Area 465 Index 576
a

¢ _—
2 i,

‘ i> = : - _ 7

i ir d= heh ge }
5 F a : :
. api sae a vied Te ae
ee — a

LA eed

i }
a >,

\ = — : har ay Ht fi
rey

ao rm one ostiad
x we . ent
3 re cxang
; vd eae’td sh
sa wee bao si beached Biovailtot aan
. rer pena -
= raid sd eine “, aii ae
ay’ ike Cites area
oer © set I Fotis &
Ai‘“s rid id’ba
Le: By of

ant 5 meymesgtetnd
pivesoese hit Gees
Tks ted Ne
Se hae ofSail
SYMBOLS

Area G Location of the centroid; universal constant of


gravitation
Reactions at A in the x, y, and z directions,
respectively g Acceleration due to gravity
Scalar components of the reactions at A in the x, y, H Height, distance
and z directions, respectively. i
Height
Acceleration vector I
Moment of inertia
. < ib +
Acceleration, distance, ie ;
I I
CG Moment of inertia about an axis passing through
the center of gravity
Width Hee by Moment of inertia with respect to the m and n axes,
Distance, coefficient of rolling resistance respectively

Flange width Inn Product of inertia with respect to the m and n axes

Constant, compressive axial force !, Principal moment of inertia

Reactions at D in the x, y, and z directions, 11,0 Moment of inertia with respect to the x, y, and z
respectively. axes, respectively

Scalar components of the reactions at D in the x, y, Ve I,,, I, Centroidal moment of inertia with respect to the
and z directions, respectively centroidal axes, x’, y’, and z’, respectively

Distance, moment arm Es Product of inertia with respect to the x and y axes

Force vector Tey Product of inertia with respect to the centroidal


axes x’ and y’
Force
i Unit vector in the x direction
Components of F in the x, y, and z directions,
respectively Jo Polar moment of inertia

Scalar components of F in the x, y, and z J Unit vector in the y direction


directions, respectively k Unit vector in the z direction
Force vector k Spring constant, radius of gyration
Force ko Polar radius of gyration

xi
xii Symbols

Centroidal polar radius of gyration Web thickness


Radius of gyration with respect to the x, y, and z Shear force (vector)
axes, respectively
Potential energy, volume, shear force
Length, distance, lead
Force vector
Length, distance
<
=s=x
Force, weight, work
Component of length L in the x, y, and z
Distributed load
directions, respectively
Coordinate of a point
Moment vector
Coordinate of a point
moment
Centroidal coordinate
Reaction moment at support A; moment about
point A Centroidal coordinate of an element

Moment vector about axis ab Coordinate of a point, moment arm


Resultant couple (vector) Centroidal coordinate

Mass Centroidal coordinate of an element

Normal reaction vector Coordinate of a point


Normal reaction Centroidal coordinate

unit vector Centroidal coordinate of an element

Origin Angle, direction angle of a vector with respect


to the x axis
Location of the centroid
Angle, direction angle of a vector with respect
Force vector
to the y axis
Force, load
Angle, direction angle of a vector with respect
ds Pitch, constant (specific distance), hydrostatic to the z axis, specific weight
pressure
Elemental length
Force vector
Virtual displacement
Force, weight
Virtual work
Force vector, resultant force vector
5x, dy, 5z Components of 5S in the x, y, and z directions,
FO
> Force, resultant force, reaction, radius of respectively
Mohr’s circle
Bex Coefficient of kinetic friction
Distance, radius
Bs Coefficient of static friction
Radius of the earth
b Angle, angle of friction
Position vector (for centroid of a body)
bx Angle of kinetic friction
Force vector, displacement vector
ob, Angle of static friction
Force, cable length, surface area
Density
Force vector, torsional moment vector
Normal stress
Force, cable tension, torsional moment, tensile
Shear stress
axial force
Angle, angle of revolution
Thickness

Flange thickness
CHAPTER

INTRODUCTION TO
RIGID—BODY MECHANICS

1.1 Introduction OUTLINE


1.2 Historical Background
1.3. Basic Concepts
1.4 Fundamental Principles
1.5 Scalars and Vectors
1.6 Units of Measurement
1.7. Accuracy of Numerical Computations
1.8 Process of Problem Solving
1.9 Summary
Key Terms
Problems
Chapter 1 Introduction to Rigid-Body Mechanics

1.1 INTRODUCTION
Mechanics is the study of the effects of forces acting on bodies. It provides
the fundamental principles by which the mechanical interaction and the motion
of bodies are related, described, and understood. The application of engineer-
ing mechanics principles to physical situations has led to the solution of count-
less problems, such as the design of high-rise buildings, bridges, dams, bicy-
cles, airplanes, and various types of machinery for manufacturing processes.
The design and development of artificial satellites would not have been pos-
sible without the basic understanding of the fundamental principles of engi-
neering mechanics.
Engineering mechanics can be divided into three broad categories:

1. Mechanics of rigid bodies.


2. Mechanics of deformable bodies.
3. Mechanics of fluids.

The study of the mechanics of rigid bodies assumes that the body under
consideration is rigid and does not deform even when subjected to large forces.
Mechanics of deformable bodies relates to real solids, which are not rigid and
will deform to a certain degree, even under the application of small forces.
Mechanics of fluids relates to the behavior of liquids, which are virtually in-
compressible, and gases, which are compressible. This text enumerates the
principles of rigid bodies.
Rigid-body mechanics can in turn be divided into two parts: statics and
dynamics. Volume 1 of this text deals with statics, in which we study the
mechanical interaction of rigid bodies at rest or in motion with constant ve-
locity. Dynamics, the study of rigid bodies in motion, is the subject of Vol-
ume 2.
This introductory chapter contains the definitions of such basic terms as
space, time, force, particle, rigid body, and mass, as well as vector and scalar
quantities, which are essential to the study of rigid-body mechanics. The laws
of motion and gravitation are listed. A detailed discussion of the two funda-
mental systems of units of measurement is also presented.

1.2 HISTORICAL BACKGROUND


Mechanics is probably the oldest branch of physical science. Principles of
mechanics were used to build the pyramids in ancient Egypt. Archimedes
(287-212 B.c.) derived relationships for the equilibrium of levers. Stevinus
(1548-1620) studied the principles of inclined planes. He also employed the
principles of parallelograms of forces for addition of vectors, which we will
Archimedes (287-212 B.c.) describe in Section 1.5. Other ancient works address the principles of the pulley
Greek mathematician and
and wrench. Galileo Galilei (1564-1642) contributed immensely to the prin-
inventor
(By permission of Oxford University ciples of dynamics by conducting experiments with the pendulum and with
Press.) falling bodies. The most important contributions to the development of modern
1.3. Basic Concepts

engineering mechanics, however, came from the scientific works of Sir Isaac
Newton (1642-1727). He formulated the laws of motion and the law of grav-
itational attraction between bodies, which we will describe in Section 1.4. After
publication of the laws of motion during the latter part of the 18th century,
there came rapid development of the principles of engineering mechanics due
to the work of Varignon, Euler, d’Alembert, Laplace, and others.

1.3 BASIC CONCEPTS


Certain fundamental definitions are essential to the study of statics and dynam-
ics of rigid bodies. These will be specified in this section.

Galileo Galilei (1564-1642)


Space Italian physicist and
mathematician who wrote
Space is a geometric region occupied by a body or bodies. This region may Dialogues Concerning Two New
be one, two, or three dimensional and will be defined in terms of linear and Sciences; with this book, he
angular measurements relative to a specified coordinate system. began the formal study of
dynamics and mechanics of
materials
Time (Courtesy of the Library of
Congress.)
Time is an absolute quantity that measures the succession of events. It is a
basic quantity in dynamics. In statics, however, time is not directly of concern.

Force

Force is the action of one body on another body. The interaction can occur
by direct contact, such as a person pushing or pulling a box. It can also occur
through a distance by which the bodies are physically separated. Examples of
this type of force are gravitational, magnetic, and electrical. Force is com-
pletely characterized by its magnitude, direction, and point of application. It
is a vector quantity, which we will describe in Section 1.5.

Concentrated Force
A concentrated force is one assumed to act at a point on a body. Jean le Rond d'Alembert
(1717-1783)
French physicist who contributed
Particle d’Alembert’s principle to the
study of rigid bodies
A particle is an object whose deformations and rotations are either negligible (Courtesy of the French Embassy,
or not of interest. A particle is an object whose motion can, for our analysis, Press and Information Division.)

be represented quite adequately by specifying the position, velocity, or accel-


eration of a single representative point in that body. For the purposes of me-
chanics, a particle is not a body of negligible dimension, but rather a body
whose dimensions are of no current interest.
Chapter | Introduction to Rigid-Body Mechanics

Rigid Body
A rigid body is an object whose size and shape are assumed to remain un-
changed under the influence of external forces. The relationship between those
forces and the changes in position and orientation that they induce in rigid
bodies is the subject of rigid-body mechanics. Since all real objects are de-
formable to some extent under the influence of forces, we may treat an object
as a rigid body only when such deformations are not of interest to us in achiev-
ing our desired purposes. Thus, among the set of real objects (whose motions
can be described by deformation, translation, and rotation), rigid bodies are
those objects whose deformations are either negligible or not of interest.

Mass
Mass is a quantity used to measure the translational inertia of a body and
represents the resistance of matter to a change in velocity. The mass of a body
characterizes the mutual gravitational attraction with another body. Two bodies
having the same mass will be attracted by the earth in a similar manner.

1.4 FUNDAMENTAL PRINCIPLES


Newton’s Laws of Motion
In 1687, Sir Isaac Newton published his treatise, The Principia, in which the
Sir Isaac Newton (1642-1727)
fundamental principles describing the motion of a particle were developed.
English mathematician who
formulated the laws of motion
These principles, known as Newton’s laws of motion, are essential to the study
and gravity of rigid-body mechanics. These laws of motion can be expressed in the fol-
(Deutsches Museum, Miinchen.) lowing manner.

FIRST LAW A particle remains at rest or continues to move along a straight line with a
constant speed if it is not subjected to an unbalanced force.

SECOND LAW A particle acted upon by an unbalanced force experiences an acceleration


that is directly proportional to the force and has the same direction as the
force.

The second law is mathematically expressed in a consistent set of units as

F=ma (1.1)
1.4 Fundamental Principles

where F is the force acting on the particle, m is the mass of the particle, and
a is the acceleration. Boldface for force and acceleration indicates that these
are vector quantities (see Section 1.5).

The forces of action and reaction between interacting bodies have the same THIRD LAW
line of action, are equal in magnitude, and are opposite in direction.

This law is basic to our understanding of force. It states that forces always Force
rom the
occur in pairs. As an example, consider a chandelier hanging by a cord from
‘, chandelier
a ceiling (Figure 1.1a). The chandelier is exerting its own weight on the cord. to the cord
This force has a magnitude of F and is directed downward (Figure 1.15). To Force
countereffect this downward force is the equal upward force of the cord acting from the
cord to the
on the chandelier (i.e., F), thus preventing the chandelier from falling. chandelier

Newton’s Law of Gravitation

Newton also proposed a law that mathematically describes the gravitational


force of attraction between two particles. Expressed in a mathematical form,
(a) (db)

jee
mm= 2 (1.2) Figure 1.1
:

where F = gravitational force of attraction ee


m,,m, = masses of the two particles under consideration (Figure 1.2) my
r = distance between the centers of the two particles (Figure 1.2)
G = universal constant of gravitation, which has a magnitude of ae
6.673..< 10 wma)ke 57 vs
tA
For a small body (which can be considered a particle) of mass m near the Times t
surface of the earth, being attracted by the earth of mass M, the force of
attraction is
My,
i Mm GM he
= = m
nena Cia
Figure 1.2
where r, = radius of the earth. This equation can be rewritten as

F=mg (1.3)
where g = GM/r2 = acceleration due to gravity. The magnitude of the ac-
celeration due to gravity, g, varies slightly from place to place on the earth’s
surface (ranging from 9.78 m/s? to 9.82 m/s”); however, the average value is
Chapter 1 Introduction to Rigid-Body Mechanics

9.81 m/s? (or 32.2 ft/sec”). The gravitational attraction of the earth on a body
of mass m is called the weight, W, of the body. Thus

Weight of a __ Mass of the |,Acceleration due (1.4)


body (W) body (m) to gravity (g) ;

Although Equations (1.1) and (1.3) are in a similar form, it is important not
to confuse Newton’s second law of motion with Newton’s law of gravitation.
In Equation (1.3) the acceleration due to gravity, g, is a constant. Objects do
not have an acceleration of g unless they are in free-fall.

1.5 SCALARS AND VECTORS


In mechanics problems one will encounter both scalar and vector quantities.
Scalar quantities have magnitude only. Examples of scalar quantities are time,
temperature, mass, and volume. Mathematical operations involving scalar
quantities follow the usual laws of algebra.
Vector quantities have magnitude and direction and obey the rules of vector
algebra. Examples of vector quantities are force, velocity, acceleration, dis-
placement, moment, and momentum.
An example of a force vector is shown in Figure 1.3. The force has a
magnitude of F, and it is directed at an angle 6 with respect to the horizontal.
The magnitude is represented by the length of the line. The arrow defines the
sense, or direction. Point A is called the tail of the vector, and point B is the
tip of the vector. Vectors in the text will be shown by boldface letters; thus,
the vector shown in Figure 1.3 can be written as either F or AB. When writing
longhand, it is usually shown as F (or F) or AB (or AB). The magnitude of
Figure 1.3 the vector in the text will be shown in lightface italic (i.e., F or AB). In long-
hand, it is shown as |F |(or |F}) or |AB| (or |AB)).

Terminology for Vectors


In statics, various types of vectors are encountered, such as fixed vector, free
vector, sliding vector, equal vectors, negative vector, concurrent vectors, col-
linear vectors, and coplanar vectors. Each is briefly described here.
O
A fixed vector has a unique point of application in space. Figure 1.4a shows

Vi
(a) Fixed vector
a fixed force vector, F, whose point of application is O. The action of a force
on a deformable body is specified by a fixed vector at the point of application
of the force.
A free vector can act anywhere in space provided its direction and mag-
nitude are retained; it is not uniquely associated with any given point or line
in space (Figure 1.4b). An example of a free vector is the velocity of a car
moving along a straight path, as shown in Figure 1.5. Assuming the car is
(b) Free vector rigid, the velocity, v, of the car can be described accurately whether it acts at
Figure 1.4 pomntA, Bb: Co Dee.
1.5 Scalars and Vectors

Sa a
So 7

we

ae a .
3)
(c) Sliding vector WA
(f) Concurrent vectors

F F’

(d) Equal vectors “


(Note: F =F’) _-*,

F on
Za, eae

7 “<r oe te
g
(e) Negative vector (g) Collinear vectors

(h) Coplanar vectors


Figure 1.4 (continued)

A sliding vector can be applied at any point along the line of action as
shown in Figure 1.4c. An example of a sliding vector is shown in Figure 1.6.
If we are interested in the resulting motion of the car, then the force F applied
at point O or at point O’ will produce the same consequence. Hence, F is a
sliding vector.
Equal vectors have the same magnitude and direction, as shown in Figure
1.4d. Note that F = F’.
Figure 1.4e shows a negative vector, —F. This vector has the same mag-
nitude as the vector F, but it is in a direction opposite to that of vector F.
Concurrent vectors are those whose lines of action pass through the same
point. Figure 1.4f shows the vectors F,, F,, F;, and F, whose lines of action
pass through the common point O. These are concurrent vectors.
Collinear vectors have the same line of action. In Figure 1.4g, the vectors
F,,, F,, and F, have the same line of action and, hence, are collinear. Figure 1.6
Chapter 1 Introduction to Rigid-Body Mechanics

Coplanar vectors are those whose lines of action lie in the same plane, as
shown in Figure 1.4h.

Parallelogram Law of Addition of Vectors


At the outset of this section, we mentioned that vector quantities satisfy the
parallelogram law of addition: To explain this law, let us consider two free
vectors P and Q (Figure 1.7a), which are coplanar. According to the parallel-
ogram law of addition, the vectors P and Q can be replaced by a resultant
vector R, which is the diagonal of a parallelogram formed with the vectors P
and Q as its sides. This is shown in Figure 1.7b. The magnitude and direction
of the resultant R can be determined using the laws of sines and cosines from
trigonometry. Thus, from Figure 1.7, using the law of cosines, the magnitude
of R is

R = VP2 + Q? — 2P0 cosB . Wee


and, using the law of sines

le Q R
Epes Soy eee (1.6)
sina siny sinB
(Db)
Figure 1.7
Thus, we can say that the resultant vector R is found by summing P and Q

R=P+Q (1.7)

This is called vector addition.


Note that the algebraic addition of the magnitudes of P and Q will not give
the magnitude of R, unless P and Q happen to be collinear. This means that,
in general, P + OQFR.

1.6 UNITS OF MEASUREMENT


Fundamental quantities commonly used in mechanics are length, time, mass,
and force. The magnitudes of these quantities are measured by arbitrarily
chosen divisions called units. In any consistent system, three of these units
may be defined independently; the fourth must be derived, according to New-
ton’s second law. Two common systems of units are now used in the United
States:

1. SI units (International System of Units).


2. U.S. Customary units.
1.6 Units of Measurement

SI Units

The International System of Units is a modern version of the MKS metric


system (meter, kilogram, second) and is intended for worldwide standardiza-
tion. This is an absolute system of units: the fundamental quantities measured
by the base units include mass and not force, and are thus independent of the
location of measurement.
The SI units can be divided into three basic classes: base units, supple-
mentary units, and derived units." The three base units are meter (m) for
length, kilogram (kg) for mass, and second (s) for time. Radian (rad) for plane
angle is a supplementary unit.
Derived units are formed by combining base units, supplementary units,
and other derived units according to the algebraic relations linking the corre-
sponding quantities. Quantities such as force, energy, power, pressure, fre-
quency, acceleration, velocity, area, and volume have derived units. The most
important derived unit, for force, is called the newton (N) and is defined as
the force that gives an acceleration of 1 m/s? to a mass of 1 kg. The kilogram
is the mass of a platinum standard kept in France.
From Equation (1.1) we have seen that (Figure 1.8) ———— Acceleration=1 m/s?

Force = Mass X Acceleration

Therefore

Force=1N
1 N =(1 kg)(1 m/s”) = 1 kg - m/s? (1.8)

From Equation (1.4) the weight of a body, which is the gravitational force
acting on the body, can be given as
W = mg

Thus, the weight of a body having a mass of | kg is (Figure 1.9) Figure 1.8

(1 kg)(9.81 m/s?) = 9.81 kg - m/s* = 9.81 N


or, in general,

W (N) = m (kg) X g (m/s”)


Table 1.1 presents a summary of SI units frequently used in solid mechanics.
When a quantity is very large or very small, the units may be modified by
using the appropriate prefixes shown in Table 1.2.
According to Table 1.2, for length we can write that

1000 m = 1 X 10° m = 1 km

* ASTM Standard E-380.


10 Chapter | Introduction to Rigid-Body Mechanics

Acceleration due to
gravity, g= 9.81 m/s?

Weight, W=9.81N
Figure 1.9

Table 1.1 SI units


Class Quantity Unit Formula

Base Length Meter


Mass Kilogram
Time Second

Supplementary Plane angle Radian Rad


Derived Acceleration Meter per second
squared
Angular acceleration Radian per second
squared
Angular velocity Radian per second
Area Square meter
Energy Joule
Force Newton
Impulse Newton - second
Moment Newton - meter
Moment of inertia Meter to the fourth
of an area power
Moment of inertia Kilogram : meter
of a mass squared
Power Watt
Pressure Pascal
Velocity Meter per second
Volume (liquids) Liter
Volume (solids) Meter cubed
1.6 Units of Measurement 11

Table 1.2 SI prefixes


Multiplication factor Prefix Symbol
1 000 000 000 = 10° giga G
1.000 000 = 10° mega M
1000 = 10° kilo k
100 = 10? hecto* h
LOO: deca* da
Onan deci* d
OO silOme centi* c
0001 —s10ine milli m
0.000 001 = 10~° micro WL
* To be avoided except for measurement of areas and volumes.

Similarly,

12500) mi="12.5103m —=212.5km


2.85 km = 2850 m
286 mm = 0.286 m
Examples using prefixes for mass and force are as follows:
IDO sis. 20104 — 2 ke
1.2 Mg = 1200 kg
1 g = 0.001 kg
2500 N = 2.5 kN
3200 KN = 3.2 MN
As noted in Table 1.2, the use of such prefixes as hecto, deca, deci, and
centi should be avoided in the measurement of length, mass, or force. They
are, however, allowed for the measurement of area and volume. So we can
write for area:

ixdmee (th. dm)7—=.G.0> 4am) 7=310 4.7


Ven = (lem) =(10 2m) —"10=" m-
1 mm? = (1 mm)* = (107? m)? = 107° m?
Similarly, for volume measurements:
ldo = (l dm) =10—'myi— 10° 2m
ficm = (cm) = (104m) = 10;° m*
fmm = mm)? =U0* my = 10 ° mm
Non-SI units that may be used with the SI system are shown in Table 1.3.
Chapter | Introduction to Rigid-Body Mechanics

Table 1.3 Non-SI units used with the SI system


Quantity Unit Symbol Definition
Time Minute min 1 min = 60s
Hour h 1h = 60 min
Day d 1d = 24h
Week, month, etc.
Plane angle Degree A (T
1° = |—— Jrad
180
Mass Metric ton t ti slOske

U.S. Customary Units


The U.S. Customary units, in which the three base units are foot (ft) for length,
pound (1b) for force, and second (s) for time, are still widely used by American
engineers. The U.S. system is a gravitational system, as the base unit for force
is dependent on the earth’s gravitational attraction and is thus notabsolute.
The second is the same unit as in the SI system. One foot is equal to 0.3048 m.
The pound is the weight, at sea level and 45° latitude, of a platinum standard
kept at the National Bureau of Standards in Washington, D.C. The measure-
ment of mass is given by a derived unit, and the unit is the slug. One slug is
the mass that is accelerated at 1 ft/s’ when a force of 1 lb is applied. From
Equation (1.1),

F=ma
1 Ibi (i! slugs) <i it/s2)
or

1 lb
igsino— = Lilbe s*/ft (1.9)
1 ft/s?

The pound as the unit of force is widely accepted in the study of statics,
and poses no problems. In dynamics, however, where force and mass are both
of concern, it is important to be careful when using the units of force and mass.
If the weight of a body (W) is given, its mass can be obtained from Equation
(1.3) as

m (slug) = a AID)
gy)

The acceleration due to gravity, g, is 32.2 ft/s”.


Also, in many cases, kip (kilopound) and ton are used as units of force.
Note that 1 kip = 1000 lb, and 1 ton (U.S.) = 2000 lb.
1.6 Units of Measurement 13

Conversion of Units
Sometimes we need to convert from one system of units to another. Common
conversions are given here.

Length By definition,

1 ft = 0.3048 m (1.10)
To find the equivalence of a mile in meters,

1 mile = 5280 ft = 5280(0.3048 m) = 1609 m


Hence,

1 mi = 1609m (ULI

In a similar manner, to find the equivalent value for inches,

: 1 1
lin. = D ft = gp 00 3048 m) = 0.0254m

Hence,

1 in. = 0.0254 m = 25.4 mm (1.12)

Force In U.S. Customary units, the unit of force is the pound (lb). From
Equation (1.4) we know that W = mg. A mass that at sea level weighs one
pound (force) has a value of 0.4536 kg. At sea level g = 9.807 m/s.’ There-
fore,

1 Ib = (0.4536 kg)(9.807 m/s”) = 4.448 kg - m/s? = 4.448 N


Hence,
1 lb = 4.448N (1.13)

Mass_ As mentioned, the unit of mass in U.S. Customary units is the slug.
The numerical relationship between the two systems is found to be
lib 4.448.N
l slug = 1 Ib-s*/ft = = 1459N-s*/m
1 ft/s? 0.3048 m/s?
From Equation (1.8), we know that 1 kg = 1 N-s?/m. Therefore,

1 slug = 14.59 kg (1.14)


Conversion factors for quantities in U.S. Customary units to SI units and
SI units to U.S. Customary units are given inside the front cover.

* Although g at a latitude of 45° at sea level is 9.807 m/s’, it is generally taken as 9.81 m/s?
for problem solving.
14 Chapter | Introduction to Rigid-Body Mechanics

EXAMPLE 1.1

Convert the velocity of a car moving at 22 ft/s to km/h.

Solution We know that 1 ft = 0.3048 m = 0.3048 * 1073 km. We also know


that 1 s = 1/3600 h. Thus we write the conversion as

22 ft (0.3048 x 10-3 km) (3600


22.00 ft/s = 2p (25) =24.14km/h <

i} i} t

Dimension check> (f) (=) (£)


s ft h
L

SpE
EEE
EXAMPLE 1.2

Convert the moment of 148 N - m applied to a wrench to Ib - in.

Solution We know that 1 N = 0.22482 lb. Also,

12 in. .
1 m = 3.281 ft = (3.281 n(22) = 39.372 in.

t t t

é : in. ;
Dimension check—> (ft) (2) ie (iT)

So, we calculate

0.22482 *)(2 n ll 1310.04 lb- in.


14
8N-m. =
(148 N .
m( iN <
1m

Dimension check—> (XM - 1h) (?) i (Ib - in)


LETS$.|5
eS
1.9 Summary 15

1.7 ACCURACY OF NUMERICAL COMPUTATIONS


Most statics and dynamics problems can be solved by using pocket electronic
calculators, which provide accurate numerical values to a number of decimal
points. However, the number of significant figures in the answer should not
be greater than the number of figures that can be justified by the accuracy of
the given data. For example, consider a case where the load can be measured
with a possible error of 10 Ib. If the load measures 900 Ib, it means that,
accounting for the measurement error, the load could be 900 + 10 lb. So, the
percentage error could be
10 Ib
———
900 4 10000 =e11 percent
(

Hence the required accuracy in the answer must not be greater than 1.11 per-
cent. (Nevertheless, more figures are usually retained in the intermediate cal-
culations to enhance the computational accuracy of the answer.)

1.8 PROCESS OF PROBLEM SOLVING


The best way to learn the principles of statics and dynamics is by solving
problems. Each problem may differ slightly, due to the shape of the body or
bodies involved along with the given force system, but all of the problems in
mechanics can be solved using only the basic principles stated in Section 1.4
and the theorems and relationships derived from those principles. Problem
solving is a very logical and orderly process. The following is a general step-
by-step outline for this process.
1. Read the problem carefully.
2. Develop a feeling for the physical situation involved.
3. Draw the necessary diagram(s) neatly, showing the forces acting on the
bodies, with any mathematical or physical idealizations necessary.
Neat diagrams will help you think logically. Give special attention to
free-body diagrams.
Tabulate the data given in the problem.
5. Use the proper relationships to obtain the desired quantities. Pay
attention to the correctness of units for the quantities involved.
6. Use judgment to see if the solution appears reasonable.

1.9 SUMMARY
In this chapter, we have covered the following:

1. The definition of space, time, force, concentrated force, particle, rigid


body, and mass (Section 1.3).
16 Chapter 1 Introduction to Rigid-Body Mechanics

2. Newton’s laws of motion and gravitation (Section 1.4). Based on


Newton’s second law of motion, a force can be given by the relation
F=ma (1.1)

Also, based on Newton’s laws of gravitation, the weight of a body is


given as
W =meg (1.4)

3. The fundamental definition of a vector and the terminology for vectors


(Section 1.5).
4. The two systems of measurement, SI units and U.S. Customary units,
for quantities generally used in mechanics (Section 1.6).
5. The accuracy of numerical computation and the process of solving
mechanics problems (Sections 1.7 and 1.8).

nn

base unit 9 scalar 6


derived unit 9 SI unit 9
force 3 space 3
law of cosines & supplementary unit 9
law of sines § time 3
mass 4 U.S. Customary unit /2
parallelogram law of addition 8 universal constant of gravitation 5
particle 3 vector 6
rigid body 4

SECTION 1.6
1.1. A steel box weighs 420 lb. Express the weight in KN.

1.2 A golf ball weighs 1.62 oz. Determine its mass in kg.

1.3 A car moves with a speed of 55 miles/h. Determine the speed in m/s.

1.4 A body has a mass moment of inertia of 7.80 < 10~? kg - m?. Determine its
moment of inertia in Ib - ft - s?.
1.5 Convert an impulse of 40.21 lb -s to SI units.

1.6 Convert a gas pressure of 200 MPa in a cylinder to Ib/in.’.


Problems 17

1.7. The specific weight of concrete is 150 Ib/ft?. Convert the specific weight to
kN/m?.

1.8 The density of water is 1.94 slug/ft*. Express the density in kg/m?.

1.9 Convert an energy of 24 ft - lb to Joules.

1.10 The specific weight of wrought iron is 480 Ib/ft*. Express the weight in kN of
a piece of wrought iron having a volume of 0.25 m?.

1.11 Given the following relationship:

where m mass
S| ll velocity
x = distance
a = aconstant

Determine the unit of a in the SI system.

1.12 Refer to Problem 1.11. Determine the unit of a in the U.S. Customary system.
CHAPTER

FORCE AND EQUILIBRIUM


OF PARTICLES

OUTLINE 2.1 Introduction

Concurrent Forces in Two Dimensions

2.2 Resultant of Two Forces


2.3 Resultant of Several Concurrent Forces on a Particle
2.4 Components of a Force
2.5 Rectangular Components of a Force
2.6 Unit Vectors
2.7. Resultant of Several Forces from Their Rectangular Components
2.8 Equilibrium of Particles
2.9 Free-Body Diagrams
2.10 Common Types of Connections

Concurrent Forces in Three Dimensions

2.11 Rectangular Components of a Force in Space


2.12 Direction Cosines of a Force Vector
2.13 Representation of a Force Using Unit Vectors in Three Dimensions
2.14 Representation of a Force Vector with Known Magnitude
and Line of Action
2.15 Resultant of Concurrent Forces on a Particle in Three Dimensions
2.16 Equilibrium of a Particle in Three Dimensions
2.17 Summary
Key Terms
Problems

18
2.2 Resultant of Two Forces 19

2.1 INTRODUCTION
In Chapter 1, we defined force as the action of a body on another body. When
represented mathematically, force is a vector quantity which has both mag-
nitude and direction. In this chapter, we will study the effects of forces and
the equilibrium of particles. The principles for the equilibrium of particles
discussed here are also applicable to many rigid bodies of different sizes and
shapes, provided the forces acting on the given body are concurrent. The chap-
ter is divided into two parts: the first part discusses concurrent forces in two
dimensions, and the second part considers concurrent forces in three dimen-
sions. Forces in two dimensions are called coplanar forces.
In each part, we will first study how to replace two or more forces by a
single force that has the same effect as the original forces. This single force
replacing the original forces is called the resultant. We will follow this by
enumerating the relationships among the forces for equilibrium of particles.

CONCURRENT FORCES IN TWO DIMENSIONS

2.2 RESULTANT OF TWO FORCES


Consider a problem where a body can be assumed to be a particle, such as
point A shown in Figure 2.la, with two forces P and Q acting on it. These
two forces are concurrent; that is, their lines of action pass through the same
point. If it is desired that the two given forces be replaced by a single force

(d)
Figure 2.1
20 Chapter 2 Force and Equilibrium of Particles

R, which will produce the same effect on the particle, then R is the resultant
of P and Q, or

R=P+Q (2a)

Equation (2.1) is a vector addition equation, which we discussed briefly in


Section 1.5. As pointed out in that section, vector additions follow the par-
allelogram law. This means that the addition of vectors P and Q can be ac-
complished by constructing a parallelogram with the vectors as its sides. The
diagonal of the parallelogram passing through A is the resultant R shown in
Figure 2.16. The magnitude of the resultant can be obtained from the length
of the diagonal, and its direction can be given by the angle 6 with respect to
the horizontal.
Vector addition can also be accomplished by applying the triangle rule.
This is illustrated in Figure 2.1c, in which vector P is first drawn from A, and
then vector Q is drawn from the tip of vector P. The resultant R can now be
found by joining the tail of vector P with the tip of vector Q. The resulting
triangle will be referred to as the force triangle.
It is important to point out that the construction of the parallelogram in
Figure 2.1b does not depend on the order in which the vectors P and Q are
selected. This means that vector addition is commutative, or

P+Q=Q+P_ (2.2)
The commutative law of vector addition can be checked while applying the
triangle rule. In Figure 2.1d, vector Q is first drawn from A. Vector P is then
drawn from the tip of vector Q. The resultant R is found by joining the tail of
Q with the tip of P. This resultant R has the same magnitude and direction as
that shown in Figure 2.1c. Compare Figures 2.1c and 2.1d with Figure 2.15.
Figure 2.1c represents the bottom half of the parallelogram shown in Figure
2.1b; Figure 2.1d represents the top half of the same parallelogram. So the
triangle is simply one half of the parallelogram, and it makes no difference
which half is used.
Figure 2.2a shows the resultant R obtained by the addition of two vectors
P and Q acting on a particle A. The subtraction of a vector Q from a vector
P means that we are adding a negative vector, in this case —Q, to P (Figure
2.2b). We write this as

P-—-Q=S (23)

If two coplanar forces P and Q are applied at points such as A and B (Figure
2.3), we can move them along their lines of action to their point of concurrency,
Figure 2.3 O. The resultant R can then be determined using the parallelogram law.
2.2 Resultant of Two Forces 21

Procedure for Vector Addition

Addition of two vectors to obtain the direction and magnitude of the resultant
can be done using trigonometry. The following is a step-by-step procedure for
the trigonometric solution. A second method for obtaining the direction and
magnitude of the resultant, using the rectangular components of the vectors,
will be discussed in Section 2.7.

Trigonometric Solution

1. Considering a reference axis, determine the magnitude and direction of


the vectors to be added. In Figure 2.4a, let the magnitudes of P and Q be
40 lb and 80 lb, respectively. P makes an angle of 8, = 30° with the horizontal
(reference axis), and Q makes an angle of 6, = 80° with the horizontal.
2. Draw the force triangle ABC as shown in Figure 2.45. Note that the angle
B = 0, + (180° — 6,) = 180° — (8, — 9).
3. The line AC represents the resultant R. Use the law of cosines to find the
Reference axis
magnitude of R; that is,
(a)

R=N\ P2102 — 2P0 cos B (2.4)

In this example, P= 40 1b, Q = 80 lb, and B = 180° — (80° — 30°) = 130°.


Thus,

R = V (40 lb)? + (80 Ib)? — (2)(40 Ib)(80 1b)(cos 130°) = 110.06 lb

4. Determine the direction of R by using the law of sines:

sin y ~ sina sin B 2

For this problem, rewriting the equation and substituting the known values
gives
Figure 2.4
re Q : a 80 :
Qa — sin |(
= )siB |
= sin |G ; )si130 e |
— 33.8 ale

So,

0; = 0, + a = 30° + 33.84° = 63.84°


5. Write the answer as

R = 110.06lb A 63.84°
22 Chapter 2 Force and Equilibrium of Particles

EXAMPLE 2.1

The hook shown is subjected to two forces P and Q. If P = 300


Ib and Q = 600 lb, determine the magnitude and direction of the
resultant R.

Solution The force triangle is shown. Using the law of cosines

R= VPE+O—2PO
cons
We know that P = 300 Ib, Q = 600 Ib, and B = 180 —
[(90 + 30) — 45] = 105°. Thus,

R = V(300 Ib)? +(600 Ib)? —(2)(300 1b)(600 Ib)(cos 105°)


= 737.0 lb
P=3001b
Using the law of sines

Re caaee.
sinB sina

a cere
= sin WN
|(E)sin 8|oe,
sin 000
| ae
(cin 1059

= 23.15°
6 =a + 45° = 23.15° + 45° = 68.15°
Therefore,

R = 737.0 lb A 68.15° <


2.3 Resultant of Several Concurrent Forces on a Particle 23

EXAMPLE 2.2

The two forces P and Q are concurrent at A. Given P = 6 kN,


what should be the magnitude of Q so that the resultant R is
vertical? Also determine the magnitude of the resultant.

Solution The force triangle is shown. Using the law of sines

ss anil
sing sin y

Note that P = 6 kN, a = 50°, and y = 30°. So, we can


compute

Q=9.19kN <
Again using the law of sines to determine the magnitude of R,

R= (2
sin
eer = (= au JeekN)
sin 30
R=11.82kN <

EXAMPLE 2.3
Two forces P and Q are applied to a steel flagpole at A. The
magnitude of P is 200 Ib. Determine the magnitude of the smallest
force Q necessary to make the resultant horizontal. Also deter-
mine the magnitude of the resultant R.

Solution The force triangle is shown. Note that, for the


resultant to be horizontal, we have several choices; that is, Q
can be Q,, Q,, Q;, ... . However, the smallest magnitude of Q
would be when Q = Q,. Thus,

OQ = P sin 30° = (200 Ib) sin 30° Q=100lb <


R = P cos 30° = (200 Ib) cos 30° R=173.21lb <
24 Chapter 2 Force and Equilibrium of Particles

2.3 RESULTANT OF SEVERAL CONCURRENT


FORCES ON A PARTICLE
The resultant of more than two concurrent forces acting on a particle can be
determined by using the triangle rule. Consider the vectors F,, F,, and F,
acting at A (Figure 2.5a). To find the resultant, we first determine the resultant
S which is the sum of vectors F, and F, (Figure 2.5b). We now add vectors
S and F; to find the resultant R (Figure 2.55). Thus,

F,+F,+F;,=(F,+F,) +F,;=S+F,=R (2.6)

From Figure 2.5c, we see that the resultant R can be determined by adding
the vectors in a tip-to-tail fashion and then joining point A to the tip of the last
vector plotted. This is known as the polygon rule for addition of vectors. It is
important to note that the resultant could also have been found as

F,+F,+F,=F,+(F,+F,)=R Cas

This proves that vector addition is associative.


We have seen in Section 2.2 that vector addition is commutative. Thus
(Figure 2.5d)

’F,+F,+F, =(F,+F,)+F; =F, + (F, + F3)


(2.8)
=F,+ (F, + F,)

| Note: R has the same magnitude and direction |

(b) (d)
Figure 2.5
2.3 Resultant of Several Concurrent Forces on a Particle 25

EXAMPLE 2.4

Three forces are acting at A in the direction shown. Given that F,


= 80 lb, F, = 100 lb, and F, = 200 Ib, find the resultant R.

Solution

S=F.F,
S = VF? + F3 — 2F,F, cos B
We have that F, = 80 lb, F, = 100 Ib, and B = 150°. Thus
we write

S = V (80 Ib)? + (100 1b)? — (2)(80 1b)(100 Ib)(cos 150°)


= 173.9 lb
Also, using the law of sines,
F, S 100 Ib 173.9 Ib
: 5 or ; = Sa ee
sina sing SINQMESINEI Og
100
a =
= sin
sin! |
(Ee)
——— cin
i 150 - |= 16.71 So

Now we can find @, as

0, = 30° + a = 30° + 16.71° = 46.71°


The law of cosines gives
R= VS?
+ F3 — 2SF, cos8
We now have that S = 173.9 lb, F,; = 200 lb, and6 =
106.71°. So,
R=
V (173.9 Ib)? + (200 Ib)? — (2)(173.9 Ib)(200 1b)(cos 106.71°)
= 300.4 Ib
Sie he
sin sin 106.71°

"= sin} (B)osin106.2119)

200
= sin] (2sin 106.119 = 39.62°

0, = 0, + 39.62° = 46.71° + 39.62° = 86.33°

R = 300.4 lb A 86.33° <


26 Chapter 2 Force and Equilibrium of Particles

2.4 COMPONENTS OF A FORCE


In Sections 2.2 and 2.3, we saw that two or more force vectors can be replaced
by a single force vector, called the resultant. The reverse is also true: a given
force vector can be replaced by two or more force vectors, which are called
components. Components are obtained by constructing lines parallel to the
given axes. The principle for obtaining the components of a vector R is shown
in Figure 2.6. Two cases may arise.

m
Case |
Given The magnitude and direction of R, and the line of action of the com-
ponents such as the m and n axes as shown in Figure 2.6a.

Determine The magnitudes


of the components.

Solution Referring to Figure 2.6a, OB represents the component P and oc


represents the component Q. Referring to the triangle OAB and using the law
of sines,
Cae tOme R
sinB sina _ sin(180° — a — B)
Thus
= R sin B oe nae Acs a
Figure 2.6
psi( 180 oa) sin(180° — a — B)

Case Il
Given The magnitude and direction of R and the magnitude and direction of
one component such as P (Figure 2.65).

Determine The line of action and the magnitude of the other component such
as Q.

Solution Referring to Figure 2.6b, OA represents the force vector R and


OB represents the component P. Draw line AB and then draw line On parallel
to AB. Line AC is then drawn parallel to Om. OC is the required component
Q. Referring to the triangle OAB and using the law of cosines,
Q? = P? + R* — 2PRcosa
or
Q = VP? + R? — 2PR cosa
From the law of sines,

Q R P
sina sin@~ sinB
2.5 Rectangular Components of a Force 27

Thus,

EXAMPLE 2.5

A force R has a magnitude of 1000 Ib. The magnitude of its


R= 1000 lb
component P along Om is 1200 lb. Determine the magnitude and
direction of its other component Q.

Solution The required force triangle is drawn. From this we


can write
Q? = R* + P? — 2RP cos 60°

Substituting the known values and solving for Q gives

Q = V(1000 Ib)? +(1200 Ib) =(2)(1000 1b)(1200 Ib)cos 60°


= 1114 lb
From the law of sines,

1000 _ 1200 _ 1114


sina sinB sin60°
O ano
P=12001b
eres Cane: ©) Bee ae

8 = 180° — a = 180° — 51.02° = 128.98° Q = 1114 bA 128.98° <

2.5 RECTANGULAR COMPONENTS OF A FORCE


The rectangular components of a force can be obtained by resolving the force
into two components with their lines of action at right angles to each other. In
Figure 2.7a, the rectangular components of F are F, and F,. The component
F, is in the direction of the x axis, and F, is in the direction of the y axis. In
Figure 2.7b, the rectangular components are F, and F}, respectively, in the
direction of the x’ and y’ axes. This is a special case of the discussion presented
in Section 2.4. Usually the x and y axes are chosen in the horizontal and vertical
directions, respectively (Figure 2.7a); however, they can be chosen to have
any orientation. Therefore, for rectangular components

F,. = F cos 8 and F; =F sin (2.9) ;


Figure 2.7
28 Chapter 2. Force and Equilibrium of Particles

Also, using the Pythagorean theorem,

F = VF2 + F2 (2.10)
We can now use sign conventions for the rectangular components of a force.
When the component is directed in the positive direction of an axis it will have
a positive sign, and when it is directed in the negative direction of an axis it
will have a negative sign. This is shown in Figure 2.8.
In Figure 2.8a (0° < 8 < 90°)
F, is positive (+)
F, is positive (+)

In Figure 2.85 (90° < 6 < 180°)


F,, is negative (—)
F, is positive (+)

Figure 2.8
2.6 Unit Vectors 29

In Figure 2.8c (180° < 6 < 270°)


F,. is negative (—)
F, is negative (— )

In Figure 2.8d (270° < @ < 360°)


F,. is positive (+)
F, is negative (— )
Note that the angle 6 is measured counterclockwise from the positive x axis.

EXAMPLE 2.6

A force F having a magnitude of 80 N is applied to a cedar box


located on a sloped loading platform. Determine the components
of the force along the x and y axes.

Solution A simplified diagram for the force at point O is


shown, from which we can see that

@ = 270° + 75° = 345° (or — 15°)


F.. = F cos 8 = (80 N) cos 345°
F.=7727NC) <
and y
F, = F sin @ = (80 N) sin 345° = —20.71N
F,=20.71NQ) <

——
F

2.6 UNIT VECTORS (a)


The product of a vector F (Figure 2.9a) with a scalar of magnitude m yields mF (mis positive)
a vector that has a magnitude of mF. This vector mF will have the same (b)
direction as the vector F provided m is positive (Figure 2.9b). If m is negative,
the direction of mF will be opposite to that of vector F (Figure 2.9c).
This concept of the multiplication of a vector by a scalar can be used to <————_—_—_—_____-
represent the rectangular components of a force vector described in Section mE (mis negative)
2.5. If we take i to be a vector along the x axis, with a magnitude of unity, (c)
then i is a unit vector in the x direction. Similarly, let jbe a unit vector in the — Figure 2.9
30 Chapter 2. Force and Equilibrium of Particles

y direction. The unit vectors i and j are referred to as Cartesian unit vectors
along the positive directions of the x and y axes, respectively. Referring to
Figure 2.10, we can then write that

Ld Des A
However, using unit vector notation,

Eee i
t t t
Vector Scalar Vector
0 i Fo= igi

Figure 2.10
Eee oe j
t i t
-Vector Scalar Vector

So F expressed in Cartesian vector form will be

F=F,i+F,j (2.11)
where F,=F cos 0
F, =F sin®

EXAMPLE 2.7

Express forces F, and F, in the Cartesian vector form, given


8, = 40°, 0, = 240°, F,; = 250 lb, and F, = 300 lb.

Solution

F, = (F, cos 0,)i + (F, sin®,)j


= (250 lb) (cos 40°)i + (250 Ib) (sin 40°)j
F, = 191.5i+160.7jlb <
F, = (F,cos8,)i + (Fp sin83)j
= (300 lb cos 240°)i + (300 lb sin 240°)j
F, = —150.0i = 2598 jlb <

Fy = 300 Ib
2.7 Resultant of Several Forces from Their Rectangular Components 31

2.7 RESULTANT OF SEVERAL FORCES


FROM THEIR RECTANGULAR COMPONENTS
In Section 2.3, we discussed the method of finding the resultant of several
concurrent coplanar forces using the triangle rule. The resultant can also be
determined by considering the rectangular components of all forces involved.
Figure 2.1la shows forces F,, F,, F3, Fy, ... acting on a particle at O. The
components of these forces in the x direction can be given as

F,,. = F,,i = F, cos 0,i


F,, = F,i = F, cos 0,1
F,, = F3,i = F3 cos 031
(2.12)
Fy. = F4,i = F4 cos 0,41

Similarly, the components in the y direction can be given as

Fi, = 1,3 = F; sin9;j


F,,2y = F,,j
2yJ = F, D sin 05j
23 (2.13)
(a)
F,, = FJ = F, sin 03j
Fy, = Fy j = Fy sin 6,j

The components in the x direction can then be added to determine a single


force R,:

R,=F,,+ F,, + Fs,4+ Fa,+-::

=(F 4, + Fs, +14, oo

(2.14)
= (F, cos 6, + F, cos 0, + F; cos 0, + F, cos 0, + --:)i
=R,i

Similarly, adding the components in the y direction to determine a single force


R,:yi

(d)
Figure 2.11
= (Fy, + Foy + Fs) + Fy °° 95
(215)

= Rij

The resultant (Figure 2.11) of all forces at O can now be written as

R=R,+R,=Rit+R,j (2.16)
32 Chapter 2 Force and Equilibrium of Particles

The magnitude of the resultant R is

R= VR aR Qi

Equation (2.17) is in a form similar to Equation (2.10), which we wrote using


the Pythagorean theorem.
The direction of R can be obtained as

ac
0 =tan” |— (2.18)

EXAMPLE 2.8

Three forces act on a hook attached to the wall. (a) Express the
resultant R as a Cartesian vector. (b) Determine the magnitude
and direction of the resultant R.

Solution

Part a. The following table can be prepared from the diagram.

F; 0; F,,. = F; cos 0; F,, = F, sin 0;


(N) _(degree) (N) (N)
100 30 86.6 50.0
80 60 40.0 69.3
200 290* 68.4 — 187.9

LF, = 195.0N=R, %Fy = —68.6N=R,


* Also can be —70°

Hence,

R = 195.0i — 68.6j(N) <


Partb. R= V(R,)? + (R,)? = V (195.0)? + (— 68.6)
R=206.7N <
R =
j= wn-*(2) = wn-1(=22) = —19,38°

9 = 19.38° <
2.8 Equilibrium of Particles 33

2.8 EQUILIBRIUM OF PARTICLES


A particle is in a state of equilibrium if, and only if, its acceleration is zero.
This means that the resultant of all forces acting on it is equal to zero. This
is actually a restatement of Newton’s first law. In Figure 2.12, a particle is
subjected to two forces, F, and F,. The Cartesian vector representations of F
these vectors are

Ley Sea and Me Se haa O

The components associated with the unit vector i are not present in the above
EB}
equations since the lines of action of the two forces are in the y direction. The
resultant R of these forces is Figure 2.12

As stated above, for equilibrium, the resultant of all forces must be zero, or

Fie

Thus, for equilibrium of the particle, the two forces F, and F, must have the
same magnitude and must be opposite in direction.
In Section 2.3 we discussed the procedure for determining the resultant of
several forces acting on a particle by drawing a force polygon. Figure 2.13a
shows a similar force polygon to obtain the resultant R of three forces (F,,
F,, and F) acting on a particle A. As we can see from this figure, the resultant
R is not equal to zero and hence, by Newton’s first law, the particle is not in
a state of equilibrium. However, equilibrium of the particle is possible if a
fourth force, F,, having the same magnitude as R and opposite in direction to

F, Fy

Equal in magnitude and opposite in direction

(a) (b)
Figure 2.13
34 Chapter 2. Force and Equilibrium of Particles

R is applied to the particle as shown in Figure 2.135, in which the force


polygon is closed. Note that
R ae F, — 0

However,

Re Kia ke) and Fe iG ea a

So that

(Rot Fai (R, + Fy)j =a)

The preceding equation will be satisfied if

Ra S= 0 k= —-K

Rotel 0) Lr = ee

A force such as F,, which holds the particle in equilibrium when any number
of forces are applied to it, is called a reaction.
Based on the above discussion we can conclude that, for two-dimensional
problems, if a particle is acted upon by a number of forces (F,, F5, F3, .. .)
and the particle is in equilibrium, then

>F,=0 and 3F=0 (2.19)

where 2F,=Fi,+F,,+hs,+-~
DH Rd he alsbye at ge a ue

Equations (2.19) are the necessary and sufficient conditions for equilibrium of
a particle.
2.9 Free-Body Diagrams 35

ee 2.9

A ball bearing is subjected to three forces, F,, F,, and F;. Given Jy
F; = 300 N, for equilibrium, determine the magnitudes of F,
and F,.

Solution Writing the rectangular components of the forces in


Cartesian vector form gives

F, = F, cos 30°i + F, sin30°j


F, = F, cos 300° + F, sin 300°j
F; = 300 cos 140°i + 300 sin 140°j
The resultant of these forces is
R=F,+F,+F,
= (F, cos 30° + F, cos 300° + 300 cos 140°)i
+ (F; sin 30° + F, sin 300° + 300 sin 140°)j
For equilibrium in the x direction we write

Se 0
F, cos 30° + F, cos 300° + 300 cos 140° = 0
or 0.866F, + 0.5F, — 229.81 = 0 (a)
or 0.866F, + 0.5F, = 229.81
Also, for equilibrium in the y direction,

2F, = 0
F, sin 30° + F, sin300° + 300 sin 140° = 0 (b)
or 0.5F, — 0.866F, = — 192.84
Solving Equations (a) and (b) gives the magnitudes of the
unknown forces:
F,=1026N <
F, =282.0N <
Note: The problem could have been solved using 8, = —60°
| eles than 300°.

2.9 FREE-BODY DIAGRAMS


The discussion provided in Section 2.8 relates to the equilibrium of particles.
The principle governing the equilibrium of particles can be extended to deter-
mine the unknown forces in a body or a group of bodies, by isolating a particle
and drawing a sketch showing all the forces acting on it. This type of sketch
36 Chapter 2 Force and Equilibrium of Particles

is called a free-body diagram. In this sketch, the direction and the magnitude
of all the known forces should be properly labeled; however, for a force with
a known line of action but unknown magnitude, the directional sense must be
assumed. After determining the solution to the problem, if the magnitude of
the force is found to be negative, then the actual directional sense is opposite
to that assumed originally. This is because the magnitude of a force is always
positive.
The forces in a free-body diagram will include (a) all active forces, which
include weight, magnetic force, and electrostatic force, and (b) all reactive
forces, which are sometimes referred to as constraint forces or reactions. The
reactive forces result from the constraints or supports that prevent the motion.
The relation between action and reaction is described by Newton’s third law
of motion.
For an example of a free-body diagram, consider a box A resting on a
frictionless inclined surface (Figure 2.14a). A force F is applied to the box to
keep it from sliding down. For drawing the free-body diagram of the box, the
active forces are:
1. The weight of the box, W.
2. The force F.

The box cannot penetrate the surface of the inclined plane; thus, there is a
(a)
constraint force or reaction exerted by the plane on the box. Let this constraint
force be equal to N. We can now draw the free-body diagram of the box by
removing the plane and accounting for its presence by the reaction N, as shown
in Figure 2.14b. Realize that, in drawing the free-body diagram, the box is
modeled as a particle. This means that we assume a concurrent force system
exists.
Once the free-body diagram is drawn, we can use Equations (2.19) to solve
for the unknown forces shown in the free-body diagram, or

(b)
SF.=0 3F,=0
Figure 2.14 Note that these two equations can be used to solve for only two unknown
forces. Hence, under certain circumstances, more than one free-body diagram
each corresponding to different particles in a system may be required to solve
for all unknown forces. The following example problems show how free-body
diagrams can be used in problem solving.
EXAMPLE 2.10

Two cables AB and BC are tied at B and support a weight of 350


Ib. Determine the tensions in the cables. Neglect the weight of
the cables.

Solution

Step 1. Choice of particle to draw the free-body diagram.


Since we need to find the tensions in the cables (which
are reactive forces), it is essential that we isolate a particle upon
which at least one of these forces (or both forces) act. For this
problem, point B is ideal since both forces act at that point (that
is, the forces are concurrent at B).

Step 2. Drawing of free-body diagrams. The free-body T,4 (force on } T,- (force on particle
é : ‘ : BC
diagram of the particle at B is shown. In the free-body diagram, particle at B | at B by cable BC)
T,,, is the tension in the cable AB, T,- is the tension in the yl A
cable BC, and T,, is the tension in the cable BD. The |
|
arrowheads show the directional sense of these forces on the | i
particle at B. 40° \60
The magnitude of the tension in the cable BD can be found B
by drawing a free-body diagram of the box as shown.

Step 3. Solution of Tp, and Tc. For the free-body diagram


of the box, for equilibrium in the y direction, 2F, = ()),
Therefore,

Ln 350 = 0 or Tpg = 350 |b

We know (from Newton’s third law) that |

Tppg = Tgp = 350 |b (a) Tpp

Now, for the free-body diagram of the particle at B,


>F, = 0. Thus along the x axis,

—Tz, cos 40° +Txc cos 60° = 0

Solving for Tz,,


cos 60°
My Pac =) = 0.653Tgc (b)

In a similar manner, in the y direction, 2F, =). ence

Combining Equations (a) and (b) into Equation (c) we get


ine = 2 Moy

From Equation (b), 7,4 = 0.6537gc = (0.653)(272)


Tz, = 178 lb <
Note: The magnitudes of T,, and Tg¢ are positive as calculated
above; hence, the assumed directional senses of these forces
are correct.

37
38 Chapter 2 Force and Equilibrium of Particles

EXAMPLE 2.11

Three cables (AB, BC, and CD) support two forces. Determine
the tensions in the cables and the angle 6 for equilibrium.

Solution The free-body diagram of the particle at B is shown.


T,,4 and Tg are the tensions in the cables BA and BC,
respectively. Now we consider equilibrium along the x axis by

iCOS OD matael 7COSi 0s 10 (a)


LF, = 0 =p

Similarly,

XFy=0 +T7p, sin 65° — Tgcsin50° — 200 N = 0 (b)


Simultaneous solution of Equations (a) and (b) gives

Tac = 326.80 N and Tp, = 496.74 N <«

The free-body diagram of the particle at C is shown, in


which Tc, is the tension in cable CD. For equilibrium,

LF, = 0 —Tc, cos 50° + Tepsin 0 = 0 (c)


We know that
Tex = Tac = 326.80 N (d)
Again,

XF,=0 Teg sin 50° + Tep cos 8 — 400 N = 0 (e)


Simultaneously solving Equations (c), (d), and (e), we obtain

Tep = 258 N and O=5455—<

2.10 COMMON TYPES OF CONNECTIONS


The most common types of connections encountered in equilibrium problems
are cables, cables passing over frictionless pulleys, and linear elastic springs.
Following are brief descriptions of the physical principles for each of these,
to help in drawing free-body diagrams.

Cables

A cable, generally assumed to be perfectly flexible and inextensible, can only


be subjected to tension (that is, a pulling force). It cannot be subjected to
compression, or the cable will collapse. The tension always acts in the direction
2.10 Common Types of Connections

of the cable. Unless otherwise specified, the weight of the cable is assumed to
be negligible.

Cables Passing over Frictionless Pulleys


Figure 2.15 shows a cable passing over a frictionless pulley. In real life, pulleys Frictionless pulley
are not frictionless; however, frictionless pulleys are necessary practical ac-
commodations we make for expediency of computation. The magnitude of the
tension in the cable is constant while passing over the pulley; that is, T, =
T,. This can easily be proved by taking the moment of the forces about the
center of the pulley. The concept of moment is described in Chapter 3.
S eS: oO

sag
Linear Elastic Springs
F
Generally speaking, any material that develops elastic restoring force may be
considered as a spring. Figure 2.16a shows a linear elastic spring whose
=) 1
unstretched length is L. A linear spring, when acted upon by a force F, goes
Figure 2.15
through a deformation AL, which is proportional to the force F. We may write
this as
AL«F
or

F=k(AL)
where k = stiffness, or spring constant

Note that if F is positive, AL is positive (Figure 2.16b); if F is negative, AL


is negative (Figure 2.16c). The weight of the spring is usually neglected.

+1
(a) Initial condition

LV) ——F
cei elvis
(b) Spring in tension

F F

: AL
LA
(c) Spring in compression
Figure 2.16
40 Chapter 2. Force and Equilibrium of Particles

EXAMPLE 2.12

K 25 an + A box is attached to a spring. The spring is unstretched in position


1. If the box is pulled by a force P and remains in equilibrium in
position 2, determine (a) the magnitude of the weight of the box,
ce W, and (b) the magnitude of P. The spring constant is given as
= : /b= k = 15: 1b/in.
C= 20in he

as Solution Let the length of the spring in the stretched position


Frictionless AB’ be called L,. To compute this length and find the angle
surface 8, we can apply the Pythagorean theorem,
Position |
L, = VL? + (25in.)? = V(20)? + (25)? = 32.02 in.

DS
@= in '(25)= Sil See

The free-body diagram of point B’ (in the stretched position) is


shown. We now consider equilibrium in the x and y directions.

LF, =0 —F sin 51.34°+ P=0 (a)


2F, =0 +Fcos51.34° -W=0 (b)
Now, using the relationship between the magnitude of the force
F and the elongation of the spring, we can write

Ww F = (AL)(k) = (L; — Lyk)


= (32.02 in. — 20 in.)(15 Ib/in.) = 180.3 Ib
Using Equation (a),

P = Fsin51.34° = (180.3)(sin 51.34°)


P=140.8lb <

Again, using Equation (b)

W = F cos 51.34° = (180.3)(cos 51.34°)


W=112.6lb <

CONCURRENT FORCES IN THREE DIMENSIONS

2.11 RECTANGULAR COMPONENTS


OF A FORCE IN SPACE
Figure 2.17 shows a force vector F acting at the origin of a rectangular coor-
dinate system having x, y, and z axes. The origin of the coordinate system is
O. Let F,, F,, and F, be the components of the force vector F in the x, y, and
z directions, respectively. Referring to the box shown in the figure, it can be
seen that Oabc is a rectangle whose sides are formed by the vectors F, and F,.
2.11 Rectangular Components of a Force in Space 41

The line Ob is the diagonal of the rectangle and represents the vector F’. Using
vector addition,

F.+F, =F’
Again referring to Figure 2.17, Obeg is a rectangle whose sides are formed by
vectors F’ and F,. The diagonal of this rectangle is Oe and represents vector
F. Thus

F’ + | ei

Hence, it follows that

F=F,+ F, +F, (2.20)

Note that Z.

Figure 2.17
F' = F cos 0,
The magnitudes of the rectangular components can then be given as

F,.
= F' cos 9, = F cos 6, cos 0,
hf sin 0; (2.21)
F, = F' sin 0, = F cos @, sin 0,

From Equation (2.21) we can derive the following:

F2 + F2 + F? = (F cos 6, cos 6,)” + (F sin 6)


+ (F cos 6, sin 05)?
= F?[(cos* 8, + sin? @,)cos” 6, + sin? 0,] = F?
or

F = VF2 + F2 + F (2.22)

Equation (2.22) states that the magnitude of a force is equal to the positive
square root of the sum of the squares of the components.
42 Chapter 2 Force and Equilibrium of Particles

EXAMPLE 2.13

A force F has a magnitude of 120 N. The rectangular components


F and F, of the force are + 60 N and — 80N, respectively. (Note:
The negative sign for F, means that the directional sense of the
force component is in the negative direction of the y axis.) We
know that the directional sense of F, is in the positive direction
of the z axis. Determine 6,, 85, and F,. (Note: Notations for 6,,
9,, and F, are the same as in Figure 2.17.)

Solution Using Equation (2.21),

Fi Fesingg;
N

6,= sin-1(2) = sin" (=) = —41.8° or — 138.2°

F.. = F cos 0, cos 85

If6, = —41.8°,
F,. = 60 N = (120)[cos(—41.8°)]cos 6, = 89.46cos 0,

60
= cos~!{ —— } = 47.88°
pa (.)
If 8, = —138.2°,
F, = 60 N = (120)[cos(— 138.2°)]cos 0,

6, = cos (2) = 113 2)L2y

F. = F cos 8, sin 8,

If 6, = —41.8° and 6, = 47.88°,


F, = (120 N)[cos(— 41.8°) sin(47.88°)] = 66.35N
If 8, = —138.2° and 0, = 132.12°,
F, = (120 N)[cos(— 138.2°) sin(132.12°)] = —66.35N
Since F, is positive (given),
0, = —41.8° <
0, = 47.88° <
F,=66.35N <

2.12 DIRECTION COSINES OF A FORCE VECTOR


Figures 2.18a, 2.185, and 2.18c show a force vector F enclosed within a ‘‘box’’
from which we can write that
2.12 Direction Cosines of a Force Vector 43

F.. = F cos a (see Figure 2.18a)


F, = F cos B (see Figure 2.18) (23)
F, = F cos y (see Figure 2.18c)

where a = Z eOa
B = Z eOg
y = Z eOc

4 y

(a) (b)
44 Chapter 2. Force and Equilibrium of Particles

In obtaining Equation (2.23), it is important to realize that Oae (Figure


2.18a), Oge (Figure 2.18b), and Oce (Figure 2.18c) are right triangles with
90° at a, g, and c, respectively. The cosines of the angles a, B, and y, which
define the direction of the force vector F, are called direction cosines. Com-
paring Equations (2.21) and (2.23), we see that

COS & = cos 8, cos 85


cos B = sin 0, (2.24)
cos y = cos 8, sin@,

Thus

cos~ a.a + cos* 2 B + cos* Deen


y = 1 (2.25)

EXAMPLE 2.14

y For the force F shown, determine:

a. The components F,, F,, and F,.


b. The angles a, B, and y that the force F makes with the x, y,
and z axes.
F=1001b
Solution

Part a. In this case, 8, = 50° and @, = —30°.


F,, = F cos 8, cos 8, = (100 Ib)(cos 50°)[cos(— 30°)]
F, = 55.67 lb <
F, = F sin 8, = (100 Ib)(sin 50°)
F, = 76.60 lb <
F, = F cos, sin 8, = (100 Ib)(cos 50°)[sin(— 30°)]
F, = —32.14lb <

Part b. Utilizing these values to find the angles, we write

F
a= eos-(%) = cos (357) a= 56.179 <

B= 40.0° <

y¥=cos! FE = COS a peas y = 108.75° <


F 100 :
2.13 Representation of a Force Using Unit Vectors in Three Dimensions 45

2.13 REPRESENTATION OF A FORCE USING


UNIT VECTORS IN THREE DIMENSIONS
In Section 2.6, for two-dimensional problems, the concept of unit vectors i and
j directed along the x and y axes was introduced. For three-dimensional prob-
lems in a rectangular coordinate system we can use the unit vectors i, j, and
k directed, respectively, along the x, y, and z axes as shown in Figure 2.19.
The components of a vector F in a rectangular coordinate system (Figure 2.20)
can be written as

F, = F,i 10s ad F, = F,k (2.26)

If we combine Equations (2.20) and (2.26), we find that

F=F,i+F,j+Fk (2.27)
a
Zz

Also, combining Equations (2.23) and (2.27), we can write that Figure 2.19

F = F(cos ai + cos Bj + cos yk) (2.28)

Let us now define a unit vector n in the direction of the force F. The unit
vector n has a magnitude of unity as shown in Figure 2.21 and can be ex-
pressed as
n = cos ai + cos Bj + cos yk (2.29)

where cos a, cos B, and cos ¥ are the direction cosines of force F. Combining
Equations (2.23) and (2.29), we have

Zz

Figure 2.20 Figure 2.21


46 Chapter 2 Force and Equilibrium of Particles

ieee Ys bs
n =i
pe FIja F 2
(2.30)

Also, from Equations (2.28) and (2.29) it is apparent that


F = F n
L i 7 (2.31)
vector scalar vector
Thus, unit vector

EXAMPLE 2.15

A force F having a magnitude of 50 N acts on a hook attached


to the wall.

a. Determine the angles a, B, and y that the force makes with


the x, y, and z axes.
b. Write an expression for the unit vector n in the direction of
the force F.
c. Express F in terms of unit vectors i, j, and k.

Solution

Part a. For the given force, 8, = —25° and 0, = +35°.


From Equation (2.24) we write

a =cos '(cos 0, cos 85) = cos” ![cos(— 25°) cos(35°)]


a = 42.06 <
B =cos~ ‘(sin 0,) = cos” ![sin(—25°)] b=11s =<
y = cos '(cos 6, sin 6) = cos~ '[cos(—
25°) sin(35°)]
y = 58.68° <
Part b. From Equation (2.29) the unit vector is

n=cosai+cos
Bj +cos yk
= cos 42.06° i + cos 115° j+ cos 58.68°k
n = 0.742 i— 0.423 j+ 0.520k <

Part c. Using Equation (2.31)

F = F n= ‘50N)(0.742 i — 0.423 j + 0.520 k)


F = 37.10 i — 21.15 j+ 26.0 k(N)
2.14 Representation of a Force Vector with Known Magnitude and Line of Action 47

2.14 REPRESENTATION OF A FORCE VECTOR WITH


KNOWN MAGNITUDE AND LINE OF ACTION
In many cases, the magnitude of a force will be known and two points along
its line of action will be specified. It may be necessary to express the force
vector in terms of its rectangular components. The method of accomplishing
this may be explained by using Figure 2.22, in which the line of action of the
A
force vector F is AB. The coordinates of point A are x,, y,, and z,, and those
pA Oli 2a)
of point B are x,, y>, and z,. The length L of AB is

[eas A Bo od
Ea oiBe 32)

where

L, = Xp - xX, Le =o = 4 Lo = 25— 2 (2233)


Figure 2.22
As in Equation (2.27), we can write that

De ar, i york

The unit vector n along the line AB can then be expressed similarly to
Equation (2.30), as
fh OWines ge
n=—i+—j+—k (2.34)
L LE L

where L = Vi2+
12 +L?

from Equation (2.32). Thus the force vector is

F =Fn=F(—i+—j+
n i J —k (2335)
e es ‘
+8 Chapter 2 Force and Equilibrium of Particles

| EXAMPLE 2.16

y(ft) A force F has a magnitude of 95 lb and is directed along line AB.


Express F in terms of F,, F,, and F,.

Solution From the figure, we see that the coordinates of point


A are

x,= +4 ft y,= +2ft z,=Oft


and the coordinates of point B are
x(ft)

The magnitude of the components of vector AB in the directions
of the x, y, and z axes are computed from

L,=% =X, = (=1) — (4) =o tt

Ly = yo — yy = (—2.5) — (2) = —4.5ft


= 4 3) (0) ait
The length of AB is

= VL? AL? te Lee V (3) 4 5) 3) sit


Lee Lee
F=Fn=F Pig Jeeak

= 050 Ga) (an) (re) = — 64.45 i — 58.01 j + 38.67 k (Ib) <

EXAMPLE 2.17

Refer to Example 2.16. Calculate the angles a, B, and y, which Therefore,


define the direction of the force.
a = 132.7° <
Solution The unit vector in the direction of the force F is

Ike ae = 127.60 <


n= it dt k= cos.ol + cos Bj = cos yk

y= 66° <
2.15 Resultant of Concurrent Forces on a Particle in Three Dimensions 49

2.15 RESULTANT OF CONCURRENT FORCES


ON A PARTICLE IN THREE DIMENSIONS
If a number of concurrent forces F,, F,, F3, ..., are acting on a particle in
space, the resultant R of these forces can be obtained by adding them vecto-
rially, thus

Fee een] Mato


aK
F, =F), 1 see. | + Fok

R= 3F,i+ 2F,j+2F,k
Using the rectangular components, the resultant R can be expressed as
Re Reais kayak kK SP i 2h) ok
From the preceding equation, it follows that

R, = =F, R, = 2F, R, = 2F, (2.36)

and the magnitude of R is then

Re (ry ry OF) G27)

Following Equation (2.23), the angles that the resultant makes with the x,
y, and z axes are given as

R
as cos-'(Fs) 6B = cos 7: (7) y= cos) (2.38)

We can use these methods to determine the magnitude and direction of the
resultant of several concurrent forces in space.
EXAMPLE 2.18

The tensions (F4,, F4c, F4p) in cables AB, AC, and AD are 100 Solution
lb, 130 lb, and 95 lb, respectively. Determine the resultant R of
Thane free eA Vectorial representation of F,4z. The coordinates of point A
are (0, 4 ft, 0), and the coordinates of point B are (0, 0,
6 ft). Let the length of AB be L,. Then
L,,=(0—0)=0
Ly = (0 —4) = —4ft
L,,= (6— 0) = 6ft
Ly = Vy ty? i) = 0) (= 4G)
= 7.21 ft
Using Equation (2.35),

x eee Ly pe Ey Ae Ly, k
AB — !'aB 7m 1 rm J ie

0
= 1001b)(a)
(Jie =A| ee 6
(=): (5 )a
|
F,, = —55.48 j + 83.22 k (Ib)
Vectorial representation of F,4-. Coordinates of points A and
C are (0, 4 ft, 0) and (4 ft, 0, 6 ft), respectively. Let the
length of AC be L,. So L3, = 4 ft, L,, = —4 ft, Lo, = 6 ft,

rer (2he (Ee


and Le 805 fe

-oo (a8)
F,- = 63.03 i — 63.03 j+ 94.55 k (Ib)
Vectorial representation of F,p. Proceeding in a manner
similar to that for F,, and Fy4.,

wo =Ful (4) +2) +(2 )a|


1p ale

Thus we find that


Xp
— Xa \.
SE)
Mis
a
OWA Ve Zp
aE (|
— 2a

Fp = 95 | (5) oP (=); +e (.)«|


4.47 4.47 4.47
Fp = 42.51 i — 85.01] (1b)
Calculation of the resultant. Since we know that
R = Fy, + Fuc + Fap
then
R = (63.03 + 42.51)i + (—55.48 — 63.03 — 85.01)j
+ (83.22 + 94.55)k
R= 105.541) — 20352) + L/7al7 & Gb) ee

50
2.16 Equilibrium of a Particle in Three Dimensions 51

2.16 EQUILIBRIUM OF A PARTICLE


IN THREE DIMENSIONS
A particle subjected to several forces will be in a state of equilibrium if the
resultant R of these forces equals zero. However, the resultant can only equal
zero if its rectangular components are equal to zero, or

R, = >F,=0 R, = 2F, =0 R, = >F,=0 (2.39)

It is important to note that only problems with three unknowns can be solved
using Equation (2.39).

EXAMPLE 2.19
y(ft)
A weight of 120 lb is being supported by three cables. Calculate
the tensions Ty,, Tog, and Ty in the cables.

Solution The free-body diagram of the particle at O is drawn.

Vectorial representation of the forces. As before, we write the


force vector equation for cable tension in OA as

ae On\s y= 0, gy =
eT A - 4 k
is on(Loa ) (Loa )i (Loa
From the figure we have that x, = 3 ft, y, = 0, andz, = 0.
Thus

Lon VX, — 0) (yg = 0) 4", = 0)


=VG20)2-40=07 POH 02=3 fh 2(ft) W=1201b
io
and

a0 O=0\) /0=0
loa = Tos|(2°) (2); cE (=) |= Toai (a)
Similarly,
xp-0\. [y,—0\. a
Top =T. sol PA Fae k
iy ol (lt ) (Hise ) ((i W = —120j (Ib) (d)
T..=T bee, ie aa. ‘om pao. k For equilibrium,
OB 80?1.\ 539 Js N59 )> 9\5189 R = To, + Top + Toc + W =0
= 037115, 0.7427.) — 0.557T>,k (b) So, as ER, = 0, ER, = 0, and ER, = 0,
ean. Sheet 0Sit 053757 = 0 (c)
Toc = Toc|(S—")i fs(2
>= *)
ye (<= “| SR, = 0.742Top+ 0.53Toc— 120 = 0 (f)
os ; as ? a ER, = —0.557Top + 0.662T9¢= 0 (g)
oc Toc|(=4)ia ()3 i (3.)«| Simultaneously solving Equations (e), (f), and (g), we obtain
= —0.53T9c i + 0.53T
oc j + 0.662T9. k (c) T= 82.73 1b Top = 101.67 Ib Toc = 84.93lb <
Sys Chapter 2. Force and Equilibrium of Particles

2.17 SUMMARY
In this chapter, we have covered the following:
Ee Two or more concurrent coplanar forces acting on a particle can be
added by using trigonometric relationships (that is, the laws of sines
and cosines), to obtain the magnitude and direction of the resultant.
Law of cosines:
R=NV P*-+ 07 — 2PC cos B (2.4)
Law of sines:
P 7) R
sin y ~ sina sin B cz)
Vector addition is associative.
F,+F,+ F, =(F, + F,)+ F, =F, + (F,+ F;)=R
Vector addition is commutative.
F,+F,+F,=(F, + F,)+F;=F,+
(F, + F) . (2.8)
=F,+(F,+F,=R
For two-dimensional problems, the magnitude of the rectangular
components of a force F can be given as

F,. = F cos 0 and Fo —F sing (2.9)

Similarly, for three-dimensional problems


F..= F cosa k= F.cos B F,=F cos y (2.23)
where cos a, cos B, and cos y are direction cosines of the force vector
F, and
cos? a + cos* B + cos” y = 1 (2.25)
A force F in the Cartesian vector form can be written as follows. For
the two-dimensional case,
Te et (2.11)
where
| IA et Sr (2.10)
For the three-dimensional case,
BrP nel yar Kk Q29
where
| INET esa
anew Ck (2.22)

Note that i, j, and k are the unit vectors in the x, y, and z directions,
respectively.
For three-dimensional problems, if n is a unit vector along the line of
action of a force vector F, then
Key Terms 53

F=Fn (2.31)
where
th Ta
eelFi oyF J tk
F
ie ;
(2.30)

7. The resultant R of a number of concurrent forces F,, F,, F3,..., can


be expressed for two-dimensional problems as

Ree | (2.16)
where
he = >F, 1 = aE

Ria R2 RG (2517)

O==tane
an 1( 3
R, (2.18
el )

and, for three-dimensional problems, as

R=R,i+R,j+Rk
= 2F,i+ 2F,j+ Fk
The directions of the resultant with respect to the x, y, and z axes are

a=cos! ie B= "cose Ry
R R es= TCOSme Ky
R (2.38)

8. For equilibrium of a particle subjected to a number of forces, in two-


dimensional problems
LF, =0 and XF, =0 (2.19)
and in three-dimensional problems

LF, =0 LF, =0 XF,=0 (2:39)


9. We described the procedure for drawing the free-body diagram by
isolating a particle in a body or a group of bodies in equilibrium
(Section 2.9).
10. We presented physical principles needed to consider common types of
connections such as cables passing over pulleys and springs; this will
aid in drawing free-body diagrams (Section 2.10).

associative (vector addition) 22 coplanar force /9


cable 38 direction cosine 44
Cartesian unit vector 30 equilibrium 33
commutative (vector addition) 20 free-body diagram 36
component (of a force) 26 frictionless pulley 39
54 Chapter 2. Force and Equilibrium of Particles

linear elastic spring 39 resultant /9


polygon rule 22 spring constant 39
reaction 34 triangle rule (vector addition) 20
rectangular component (of a unit vector 29
force) 27

PROBLEMS

SECTION 2.2
2.1 Two concurrent forces F, and F, are shown. Given F, = 180 lb, F, = 300 lb,
6, = 30°, and 8, = 115°, determine the resultant of F,; + F,.
2.2 If F, = 30N, F, = 80N, 0, = 40°, and 6, = 110°, determine the resultant of

2.3 IfF, = 100N, F, = 220N, 0, = 40°, and 6, = 120°, determine S where


S =F, — F..
Figure P2.1, P2.2, P2.3, P2.4, and P2.5 2.4 Show that the resultant of the two concurrent forces F, and F, can be expressed
as
R= VF? + F2 + 2 F,F, cos (0, — 9,)
2.5 Show that the resultant R of the two concurrent forces makes an angle 0, with
the horizontal where
weed ig Me te
0, = 0, + sin 7 sin(8, — 0,)

2.6 The top of a rigid post is being pulled by two ropes. The tensions in the ropes,
T, and T,, are 500 N and 800 N, respectively. Determine the magnitude and
direction of the resultant force at A.

A Ro oa

Figure P2.6 Figure P2.7 and P2.8

2.7 Two forces F, and F, are applied to a bracket. Determine the magnitude and
direction of the resultant, given F, = 250 lb, F, = 120 lb, a, = 45°, anda, = 30°.
2.8 If F, = 300 N, what should be the magnitude of F, so that the resultant is
horizontal? Also determine the magnitude of the resultant. Let a, = 60° and
a, = 40°.
2.9 Two forces are applied to the bracket. If F, = 80 lb, determine the magnitude
of F, so that the resultant is directed along AB. Also determine the magnitude of the
Figure P2.9 resultant.
Problems 55

2.10 A nail is pulled vertically upwards by applying force to two strings. Let
F, = 30 lb, 8, = 40°. The resultant is vertical and has a magnitude of 50 lb.
Determine F, and 8).
2.11 F, and F, are two forces applied to a welded bracket. Given F, = 1000 N,
F, = 3000 N, 0, = 25°, and 8, = 20°, determine the magnitude and direction
of the resultant force on the bracket.

Figure P2.10

Figure P2.11

2.12 A box is subjected to two forces F, and F,. Given F, = 100 lb and F, = 300
Ib, determine the magnitude and direction of the resultant force at A.

Figure P2.12

2.13 Two forces F, and F, are applied to a hook. Given F 1 = 200 N, what should
be the smallest possible force F, so that the resultant is directed along OA? Also
determine the angle a. Figure P2.13
Chapter 2. Force and Equilibrium of Particles

2.14 A boat is being towed along a narrow channel by two ropes attached at O.
Given the force F, = F,/m, show that the magnitude of the resultant force on the
boat, directed along the channel, is R = F,(m sina + sin B).

2.15 In Problem 2.14, if R,; = 800 lb, R = 1248.5 lb, a = 30°, and B = 45°,
determine the magnitude of F,. The resultant is directed along the channel.

2.16 The forces P and Q have magnitudes of 80 KN and 94 kN, respectively. The
resultant of these two forces is directed vertically upwards (along Ox) and has a
magnitude of 100 KN. Determine the angles a and £.

Figure P2.14 and P2.15

Figure P2.16

SECTION 2.3
2.17 Three forces F,, F,, and F, act at O. Given F, = 100 lb, F, = 200 Ib, and
F; = 250 lb, determine the magnitude and direction of the resultant R.

2.18 Determine the magnitude and direction of the resultant R for the three forces
F,, F,, and F;. Given F,; = 80 N, F, = 150 N, and F, = 270N.

F,=2001b

F3=2501b

Figure P2.17 Figure P2.18


Problems 57

2.19 Three forces act at the top of a pole. Given F, = 150 N, F; = 200 N, and
6@ = 50°, determine the magnitude of F, such that the resultant R has a magnitude of
420 N. Also determine the direction of the resultant R.

2.20 Given F, = 100 lb, F, = 190 lb, and @ = 50°, determine the magnitude of
F, such that the resultant of the three forces is 300 lb.

SECTION 2.4
2.21 Determine the magnitudes of the components of F along Am and An. Let
F = 120 lb anda = 20°.

2.22 The magnitude of the components of F along Am and An are 170 Ib and 210
Ib, respectively. Determine the magnitude and direction of F. Figure P2.19 and P2.20

30°

Figure P2.21 and P2.22

F=3601b

Figure P2.23 and P2.24

2.23 Determine the magnitude of the components of F along Am and An, given F
= 250 N anda = 100°.

2.24 The magnitudes of the components of F along Am and An are 70 N and 40 N,


respectively. Determine the magnitude and direction of F.

2.25 A force F is applied to a box. If the magnitude of the component along Am is DS


300 lb, determine the magnitude of the component along An and the angle a. Figure P2.25
58 Chapter 2. Force and Equilibrium of Particles

2.26 A force F acts at B of a frame. Given a, = 45° = a, a, = 30°, and


F = 220N, determine the magnitudes of the components of F along lines CB and
AB.

2.27. The force F has components along CB and AB. Let a, = 40°, a, = 50°, and
F = 250 lb. If the component of Fc, acting along CB is 180 Ib, determine a3.

2.28 Resolve the force F into components along BA and BC (i.e., Fz, and F,,). Let
F = 350 N anda = 60°.

2.29 Determine the magnitude of F such that the magnitudes of the components
along BA and BC are 260 N each. Let a = 60°.

Figure P2.26 and P2.27

Figure P2.28 and P2.29

2.30 The force F has components F',, and Fz- along BA and BC. Determine the
angle a so that Fp¢/Fpz, = 1.2.

Figure P2.30

SECTIONS 2.5 AND 2.6


2.31 If the magnitude of each of the forces F,, F5, F3, Fy, F;, F,, F>, and Fs is
150 N, determine the rectangular components (i.e., F, and F,) for each of the forces,
Figure P2.31 and express each of them in Cartesian vector form.
Problems 59

2.32 Express F, and F, in terms of unit vectors i and j. Let F,; = 50 lb and
F, = 80 lb.

F,=50Ib
F,=801b
Figure P2.32

2.33 Express F,, F,, and F; in terms of unit vectors i and j, given F; = 150N,
F, = 200N, andF; = 100N.

F,=200N

Figure P2.33
60 Chapter 2. Force and Equilibrium of Particles

2.34 The tension T in the cable has a magnitude of 380 Ib. Express T in Cartesian
vector form.

Sift

ee ee ie

Figure P2.34

2.35 Express force F, in Cartesian vector form, given F, 300 N.

2.36 Express force F, in Cartesian vector form, given F, 250 N.

2.37 Express force F, in Cartesian vector form, given F; = 480 N.

Figure P2.35, P2.36, and P2.37

2.38 The box on the inclined plane has a mass of 32 kg. Express the weight of the
Figure P2.38 box in Cartesian vector form.
Problems 61

2.39 The box on the inclined plane is acted upon by a force F having a magnitude
of 150 Ib. Express F in Cartesian vector form.

Figure P2.39

2.40 If the magnitude of the component of the force F in the y direction is 120 Ib,
(a) determine the magnitude of F, and (b) express F in Cartesian vector form.

2.41 A force F is applied to the bracket fixed to the wall. It produces a component
of 85 N parallel to the wall in the positive x direction and a component of 98 N
perpendicular to the wall in the positive y direction. Determine the magnitude and
direction (0) of the force.

x
Figure P2.41
Figure P2.40

SECTION 2.7
2.42 Determine the magnitude and direction of the resultant of the forces described
in Problem 2.6.

2.43 Determine the magnitude and direction of the resultant of the forces described
in Problem 2.17.
62 Chapter 2. Force and Equilibrium of Particles

2.44 Determine the magnitude and direction of the resultant of the forces described
in Problem 2.18.

2.45 Determine the magnitude and direction of the resultant of the forces described
in Problem 2.32.

2.46 Determine the magnitude and direction of the resultant of the forces described
in Problem 2.33.

2.47 Three forces act at the top of a pole. The magnitudes of F, and F, are 250 Ib
and 300 Ib, respectively. Determine the magnitude of F; if the resultant R is directed
vertically downward. Let a = 45°.

2.48 The magnitude of the forces F,, F,, and F; are 100 N, 150 N, and 250 N,
respectively. Determine the angle a if the resultant force is vertical.

Figure P2.47 and P2.48

SECTIONS 2.8, 2.9, AND 2.10


2.49 Two cables, AB and BC, are tied at B supporting a weight W. Given 8, = 30°,
6, = 50°, and W = 480 Ib, determine the magnitudes of the tensions T;, and Tz.
in cables BA and BC, respectively.

2.50 Solve Problem 2.49 with the following: 6, = 45°, 8, = 60°, and
W = 1200N.

Figure P2.49, P2.50, and P2.51


Problems 63

2.51 Determine the magnitudes of the weight W and tension T,, in cable BA, if
6, = 40°, 8, = 60°, and the tension in cable BC = 280 N.

2.52 ABC is a cable passing over a frictionless pulley at B where a force F is


applied. Let h, = 12 in. and h, = 24 in. The length of the cable ABC is 50 in.
Determine the magnitude and the direction of the force F such that the tension in the
cable is 65 |b.

Figure P2.52

2.53 Two cables, AB and BC, are tied at B supporting a weight of 75 lb. Determine
the magnitudes of the cable tensions T,, and Tg.

2.54 A weight of 300 lb is being supported by a cable and a frictionless pulley


arrangement as shown. Determine the force P required for equilibrium.

Figure P2.53

Figure P2.54
64 Chapter 2 Force and Equilibrium of Particles

2.55 A weight of 180 lb is being supported by a cable and pulley arrangement.


What should be the magnitude and direction of the force P for equilibrium?
Let B = 65°.

2.56 Determine the magnitude of the force P so that the crate weighing 205 N will
be in equilibrium. The cables are tied at B.

Figure P2.56
Figure P2.55

2.57 Given W, = 88 lb, W, = 128 lb, a, = 30°, and a, = 20°, determine the
tension in the cables AB, BC, and CD and the angle a; for equilibrium. Note: The
cables are tied at B and C.

2.58 Repeat Problem 2.57 with the following: W, = 150 N, W, = 75N,


a, = 40°, anda, = 30°.

2.59 IfW, = 100 lb, a, = 40°, a, = 30°, and a, = 65°, determine the weight
W>.

2.60 Ifa, = 55°, a, = 30°, a; = 65°, and W, = 160 Ib, determine the weight
W,.

Figure P2.57, P2.58, P2.59, and P2.60


Problems 65

2.61 AB and BC are two cables tied at B. Given F = 50 N anda, = 40°,


determine the magnitude of the angle a, for which the tension in cable AB is as
small as possible.

T=) ON)

Figure P2.61

2.62 The equilibrium position of two springs AB and BC supporting a weight


W = 500 N is shown. Determine the unstretched length of the springs. Let
kag = 500 N/m and kg = 800 N/m.

W=SO00N

Figure P2.62

2.63 The unstretched length of the spring AB is 1.5 m. The equilibrium position of
the cable BC and the spring under an applied force P is shown. Determine P.

1.5 m=Unstretched
length of spring k =600 N/m

Figure P2.63
66 Chapter 2 Force and Equilibrium of Particles

SECTION 2.11
2.64 Determine the magnitude of the rectangular components of force F,, given
F, = 85 lb.

2.65 Determine the magnitudes of the rectangular components F,, F,, and F, for the
force F,, given F, = 350N.

2.66 Determine the magnitude of the rectangular components of the force F3, given
F, = 200 lb.

2.67 F is a force and the magnitude of its component F’ in the x-z plane is 65 lb.
Xx Calculate the rectangular components F,, F,, and F,.

2.68 The components of the force F are as follows: F, = 100 N, (ie 180 N, and
F, = 220 N. Calculate (a) 6,, (b) 85, and (c) F.

2.69 Repeat Problem 2.68 with the following: F, = 68 lb, F, = 36 Ib, and
F, = 98 lb.

2.70 Repeat Problem 2.68 with the following: F, = —60 Ib, F, = 75 Ib, and
Figure P2.64, P2.65, P2.66, P2.71, P2.72, and P2.73 li —teypy IIo) :

SECTION 2.12
2.71 Determine the angles a, B, and y that the force F, described in Problem 2.64
makes with the x, y, and z axes.

2.72 Determine the angles a, B, and y that the force F, described in Problem 2.65
makes with the xX, y, and-z axes.

2.73 Determine the angles a, 8, and y that the force F, described in Problem 2.66
makes with the x, y, and z axes.

Figure P2.67 and P2.74

Figure P2.68, P2.69, P2.70, P2.75, P2.76, and P2.77


i
Problems 67

2.74 Determine the angles a, B, and y that the force F described in Problem 2.67
makes with the x, y, and z axes.

2.75 Determine the angles a, 8, and + that the force F described in Problem 2.68
makes with the x, y, and z axes.

2.76 Determine the angles a, 8, and + that the force F described in Problem 2.69
makes with the x, y, and z axes.

2.77 Determine the angles a, 8, and y that the force F described in Problem 2.70
makes with the x, y, and z axes.

2.78 A force F acts at the top of a pole. Let F = 120 lb. The direction angles a
and £ are 60° and 35°, respectively. Determine the magnitude of the rectangular
components F,, F,, and F,.

Figure P2.78 and P2.81

2.79 The rectangular components of a force F are as follows: F, = 30 N,


F, = —48N, and F, = 110 N. Determine the magnitude of F and the direction
angles a, B, and y.

2.80 A force F having a magnitude of 800 lb acts at the top of a pole. Determine
the rectangular components F,, F,, and F,, and the direction angles a, B, and y.

2.81 For the force F applied at the top of a pole, if the direction angles a = 60°
and y = 60°, show that
_ Nn 1
—_—+-—
1
qoaN lis<N Ag np

SECTION 2.13
2.82 For a force expressed as
F = — 100i + 220j + 82k(N)
determine its magnitude and direction angles. Figure P2.80
68 Chapter 2. Force and Equilibrium of Particles

2.83 Repeat Problem 2.82 for F = 65i — 120j — 90k (Ib).

2.84 Repeat Problem 2.82 for F = —85i + 68j — 100k (Ib).

2.85 Write the equation for unit vector n in the direction of application of a force
F = 28i — 60j + 55k (Ib).

2.86 Write the equation for unit vector n in the direction of application of a force
F = —72i — 80j + 100k (N).

2.87 The direction angles a and £ of a force F are, respectively, 40° and 60°. Also,
F, = +48.43 lb.
a. Determine the magnitude of the force F and the direction angle y.
b. Express the unit vector n in the direction of application of force F in terms of i,
j, and k.

2.88 The direction angles a and ¥y of a force F are, respectively, 65° and 45°. Also
given Fy = -— 100,N.

a. Determine the direction angle B.


b. Express the unit vector n in the direction of application of force F in terms of i,
j, and k.

2.89 Repeat Problem 2.88 with the following changes: a = 70°, y = 30°, and
Fi = 50.25N.

SECTION 2.14
2.90 The force F, has a magnitude of 95 lb. Express F, in terms of unit vectors i, j,
and k. Also determine the direction angles a, B, and y.

3)

: Bons eae
Figure P2.90, P2.91, and P2.92

2.91 The force F, has a magnitude of 160 lb. Express F, in terms of unit vectors i,
j, and k. Also determine the direction angles a, B, and y.

2.92 The force F; has a magnitude of 220 lb. Express F; in terms of unit vectors i,
j, and k. Also determine the direction angles a, 8, and y.
Problems 69

2.93 A plate is supported by cables AB and CD. If the magnitude of tension T,, in
cable AB is 140 N, write the vector expression for T,, (i.e., in terms of i, j, and k).
Also determine the direction angles for T4,.

2.94 If the magnitude of the tension T;,p in cable CD is 75 N, write the expression
for Tcp in terms of i, j, and k.

200 mm

200 mm

[: ee mm
Zz be—275 mm—
Figure P2.93 and P2.94

2.95 A plate OABC is supported by a vertical rod DE. The magnitude of the axial
force F on the rod at D directed along DE is 45 Ib. Write the vector expression
for F. Note: Point D is midway between A and B.

Figure P2.95
70 Chapter 2. Force and Equilibrium of Particles

2.96 AB is a tower guy wire. The tension T, in the guy wire AB at B is 8 KN.
Determine the components T,,, 7),, and T;,. Also determine the direction angles a,
B, and y.
2.97 AC is a tower guy wire. The tension T, in the guy wire AC at A is 10.5 KN.
Determine the components T},, T>,, and T>,. Also determine the direction angles a,
B, and y.

(Note: OA = 40 m)

Figure P2.96 and P2.97

2.98 OA is a boom supported by cables AB and AC. If the tension T in the cable
AB at A is 50 lb, express T in terms of the unit vectors i, j, and k. Also determine the
direction angles.

Figure P2.98
Problems 71

2.99 The line of action of a force F is AB. The magnitude of F is 100 N. If x, =


3 m and yg = 2 m, determine zz such that F, = 45 N.

2.100 If F = 200N, x, = 3m, and yz = 1.5 m, determine Z, such that F, =


60 N.

Figure P2.99 and 2.100

2.101 A circular plate is in the xz plane. It is supported by a hinge at O and a cable


AB. The radius of the plate is 2.5 ft. If the magnitude of the tension T in the cable

B| (0,3 ft, 0)

Zz
Figure P2.101
72 Chapter 2. Force and Equilibrium of Particles

AB at A is 75 lb, express T acting on the plate in terms of i, j, and k. Also determine


the direction angles a, B, and y for T.

SECTION 2.15
2.102 Two forces acting at a point are as follows:
, = 200i — 260j + 302k (N)
F, = — 120i + 180j — 208k (N)

Determine (a) the resultant R, (b) the magnitude of R, and (c) the direction angles a,
8, and y of the resultant R.

2.103 Repeat Problem 2.102 with the following:


F, = —80i + 60j — 35k(N)
F, = —55i — 60j + 45k (N)

2.104 Three forces act at O. Given

F, = 75i + 80j (N)


F, = 110N, 0, = 55°, 6, = 30°
F; = 200 N, yg = 2m, zg = 1.5m

determine (a) the resultant R, (b) the magnitude of R, and (c) the direction angles a,
B, and y of the resultant R.

2.105 Given

F, = 48 1b, x, = 3.2 ft, y, = 2.5ft


F, = 65 Ib, 8, = 55°, 8, = 45°
Figure P2.104, P2.105
F; = 90 Ib, yz = 4 ft, zg = 3 ft
determine (a) the resultant R, (b) the magnitude of R, and (c) the direction angles a,
8, and y of the resultant R.

2.106 Two forces T, and T, act at B. Find the resultant T, and also the direction
angles a, 8, and y of the the resultant Tp. Let

T, = 150 N, T, = 300 N
X,=0,y, =2m,z,=3m
Xp = 2.5m, yg = —2.5 m,z, = 1.8m
X¢ = 1.2 m, yo = 1.75 m, and z- = —0.8 m

2.107 The tensions in cables AC and BC are 300 lb and 150 Ib, respectively.
Determine the resultant force R at the support C.

2.108 The tension in cable BC is 180 lb. Determine the tension in cable AC if the
resultant R at the support C has a vertical component R, = —250 Ib.
(Xp,VB>Zp)
2.109 Two forces F, and F, act on the top of a pole. Given

Figure P2.106 F, = 350i + 400j + 450k (N)


Problems 73

Z
Figure P2.109

determine the magnitude and direction angles a, 8, and y of F, when the resultant
R = 550 j.

2.110 A weight W = 20 kN is hanging vertically at the end of a rod OA. AB and


AC are two cables. Given the tensions T,, = 15 kN and T,, = 25 kN, determine
the resultant R of the cable tensions at A and also its direction angles a, B, and y. Figure P2.110

SECTION 2.16
2.111 Four forces act at O. Given

F, = 400
N, x, = 1.5m, y,=2m,z,= —2m
F, = 300 N, xg = 2 m, yg = 3 m, zg = 1.8m
F, = 200 N
74 Chapter 2. Force and Equilibrium of Particles

(Xp, Vp5Z8)
be

(Xe, VesZc) ;
«
Oe X4:V4>Z,)
A>+A2“A
se
>< BA

z
/
Figure P2.111

determine the magnitude and direction angles of F, necessary for equilibrium to exist
at O.

2.112 Three cables AD, BD, and CD support a weight W. If W = 500 lb, determine
the tension in the cables.

2.113 If the tension in the cable AD is 250 lb, determine the weight W.

2.114 If the tension in the cable BD is 250 lb, determine the weight W.

v(ft)

Figure P2.112, P2.113, and P2.114

2.115 A circular plate which weighs 550 N is being supported by three cables AD,
BD, and CD. Given h = 2 m and radius of plate = 0.8 m, determine the tension
in the cables.
Problems 75

Figure P2.115

2.116 The rod OA supports a weight W. AB and AC are two cables. If the force in
the rod acts along its axis and W = 180 lb, determine the force in the rod and the
tension in the cables for equilibrium.

2.117 Ifthe tension in the cable AB is 125 Ib, determine the axial force in the rod,
the tension in the cable AC, and the weight W for equilibrium.

(0, 3 ft,—4 ft)

A
(0,3 fty4ft) |.

Figure P2.116 amd P2.117

2.118 Three cables AB, AC, and AD resist a force F. Given F = 400i — 300j + Zi

250k (N). Determine the tension in the cables. Figure P2.118


76 Chapter 2. Force and Equilibrium of Particles

2.119 A thin plate is attached to three cables DA, DB, and DC. Determine the
tension in the cables. Given L, = L, = 12 in.

Figure P2.119 and P2.120

2.120 Solve Problem 2.119 with L; = 12 in. and L, = 18 in.

2.121 A tripod supports a weight of 280 N. Determine the magnitude of the force
on each leg of the tripod. Given OB = OC = OD = 1 m.

| Weight = 280 N

| Weight = 400 N

v4,

Figure P2.121

2.122 The legs of the tripod in Problem 2.121 have been rearranged. If the tripod
Zz
has to support a weight of 400 N, determine the magnitude of the force on each leg
Figure P2.122 of the tripod. Given OB = OC = OD = 1 m.
Problems 77

2.123 A weight of 150 lb is held by two cables AB and AC and a horizontal force
F. Determine the magnitude of the tension in each cable and the force F.

iz

Figure P2.123

2.124 A weight of 800 N is held by three cables AB, AC, and AD attached to the
top of three posts. Determine the magnitude of the tension in each cable.

(Note: OF =OF = OG =3 m;
BE = CF=DG =4m;
OA =2m)
Figure P2.124
CHAPTER

EQUILIBRIUM OF RIGID
BODIES IN TWO
DIMENSIONS

OUTLINE 3.1. ‘Introduction


3.2 Rigid Bodies
3.3. Forces on Rigid Bodies and the Principle of Transmissibility

Coplanar Force Systems

3.4 Rigid Bodies in Two Dimensions


3.5 Moment of a Force—Scalar Approach
3.6 Cross Product of Vectors
3.7 Moments as Vectors
3.8 | Varignon’s Theorem
3.9 Procedure for Determining Moments
3.10 Couples
3.11 Resolution of a Force into a Force-Couple System
3.12 Resultants of Nonconcurrent Coplanar Force Systems

Equilibrium in Two Dimensions

3.13 Equilibrium of Rigid Bodies


3.14 Equilibrium Equations in Scalar Form
3.15 Equilibrium of Rigid Bodies Subjected to Forces at Two
and Three Points
3.16 Types of Supports and Connections
3.17 Statically Determinate Structures
3.18 Free-Body Diagrams
3.19 Procedure for Analysis of Reactions
3.20 Summary
Key Terms
Problems
78
3.2 Rigid Bodies 79

3.1 INTRODUCTION
In Chapter 2, we discussed how to determine the resultants of concurrent force
systems. We also saw that for a particle to be in equilibrium, the resultant of
the concurrent forces acting on it must be zero. This analysis is not only
applicable to particles of negligible dimensions, but also to bodies of finite or
definite size, provided the forces acting on the body are concurrent. Such a
body, subjected to concurrent forces, can be treated as a particle for the purpose
of analysis, without involving the dimensions of the body.
Many engineering structures and machines are subjected to forces that act
on different parts of the body and are not concurrent. Such bodies cannot be
treated as particles, and their dimensions must be considered in the analysis.
In this and the following chapter, we will consider the equilibrium of rigid
bodies subjected to general force systems. The procedures for determining the
resultants of nonconcurrent force systems will also be discussed.
The main difference between the analysis of particles and the analysis of
rigid bodies is that a particle, when subjected to forces, only tends to translate,
whereas a rigid body has the tendency to translate as well as rotate under the
action of a general force system. The condition of zero resultant force used
for particles in the previous chapter is therefore no longer sufficient by itself
to ensure equilibrium of a body under the action of a general force system.
A new concept, the moment of a force as a measure of the rotational effect
of the force, will be introduced in this chapter and will be used to develop an
additional condition necessary for the equilibrium of rigid bodies.
While all engineering structures and machines are three-dimensional bod-
ies, many of them are subjected to forces lying in a single plane. For analysis,
such bodies can be treated as two-dimensional bodies subjected to coplanar
force systems. The objective of this chapter is to study the analysis of coplanar
force systems and the equilibrium of rigid bodies under the action of such
forces. The equilibrium of rigid bodies subjected to three-dimensional forces
will be examined in the next chapter.

3.2 RIGID BODIES


In Section 1.3, we defined a rigid body as one that is perfectly rigid and does
not deform under the action of forces. Even though all physical bodies deform
and change shape or dimensions when subjected to forces, in most engineering
structures and machines, the deformations under service or operating condi-
tions are so small that their effect on the conditions of equilibrium or motion
of the bodies can be neglected. The assumption of rigidity considerably sim-
plifies the analysis of forces acting on the body because we need not calculate
the deformations and changes in dimensions of the body. The consideration
of deformations, however small they may be, and the resulting strains and
stresses that develop within the material of the body, is essential for the safe
design of structures. Such analysis of deformations and internal stresses and
strains, however, belongs in the area of mechanics of deformable bodies, and
is beyond the scope of this textbook on the statics of rigid bodies.
80 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

A few types of structures deform excessively under loads and, therefore,


cannot be treated as rigid bodies. Examples of these include springs, described
in Section 2.10, and structures made of materials such as soft plastics and
rubber. The equations of equilibrium or motion of such structures should
be based on their deformed shapes to take into account the effect of large
deformations.

3.3 FORCES ON RIGID BODIES AND THE PRINCIPLE


OF TRANSMISSIBILITY
Forces acting on rigid bodies may be classified as external forces and internal
forces.

External Forces

An external force may be defined as the action of another body on the body
under consideration. External forces may be further classified as applied forces
and reaction forces. Applied forces, also called loads, have a tendency to move
the body. These forces are usually known in the analysis. Reaction forces,
usually called reactions, are the forces exerted by the supports on the body
under consideration. These forces tend to prevent the motion of the body and
keep it in equilibrium. The reactions are usually the unknowns in the analysis
and are determined by using the equations of equilibrium of the body. The
state of motion or equilibrium of the rigid body as a whole depends solely on
the external forces acting on it.

Internal Forces
Internal forces are the forces exerted on a particle, portion, or component of
the body by the rest of the body. According to Newton’s third law, internal
forces always exist in equal-but-opposite pairs. Thus, each particle, portion,
or component exerts back on the rest of the body the same internal forces
acting upon it in opposite directions. When considering the state of equilibrium
or motion of the body as a whole, these internal forces cancel each other and,
therefore, do not appear in the equations of equilibrium or motion of the entire
body. Nonetheless, the consideration of internal forces is essential for the de-
sign of structures; we will study internal forces in detail in Chapters 6 and 7.

Principle of Transmissibility
The concept of a rigid body is very useful in that the external forces acting on
it can be treated as sliding vectors. In other words, a force acting at a certain
point on a rigid body may be moved to any other point along its line of action
without changing the external conditions of equilibrium or motion of the rigid
body. This is called the principle of transmissibility. For example, the force
Figure 3.1 F acting on the car shown in Figure 3.1 may be applied at point A (pulling the
3.3 Forces on Rigid Bodies and the Principle of Transmissibility 81

car) or at point B (pushing the car), or at any other point along its line of
action, and the resulting motion of the car will remain the same. As another
example, the effect of the force F on the rigid body shown in Figure 3.2 will
be the same whether the force acts at point A or B or C or D, or at any other
point on its line of action. Thus, the forces acting on rigid bodies may be
adequately described by their magnitudes, directions, and lines of action. The
points of application of the forces need not be specified.
The principle of transmissibility is very useful for determining the resultants
of external force systems acting on rigid bodies, and for establishing the equa-
tions of equilibrium or motion of such bodies. However, this principle may
not be used to determine the internal forces and stresses that develop within
the material of the body, or to determine the deformations of the body. The
internal forces, stresses, and deformations depend not only on the magnitudes,
directions, and lines of action of the forces acting on the body, but also on the
points of application of these forces. In such analyses, the forces acting on the
bodies are treated as fixed vectors rather than as sliding vectors.
Consider, for example, the bar ABC shown in Figure 3.3a. The bar is sup-

G G
Figure 3.2
|
B i
| b
¢

F
A A

(b) (c)

) y }

Car Cer Cor

(Cs

F
oe 3,5
B

A
F

(d) (e) (f)


Figure 3.3
82 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

ported at the top by a frictionless hinged support and is subjected to a down-


ward force of magnitude F acting at point A on the bottom edge of the bar.
Using the principle of transmissibility, we can move force F to any point along
its line of action. In Figure 3.3b, the force has been moved to point B, and in
Figure 3.3c, it is acting at point C on the top edge of the bar. The free-body
diagrams of the bar for the three locations of F are shown in Figures 3.3d
through 3.3f, in which C), represents the reaction force exerted on the bar by
the support. Since the bar is in equilibrium under the action of two collinear
forces, we can use the equation of equilibrium Fe = 0 to determine the
magnitude of the support reaction C,. Thus,
SF, =0
C,-F =0
CSE
Note that the above relationship between the applied force F and the support
reaction C,, remains the same whether F acts at point A, point B, or point C,
as we can see in Figures 3.3d, 3.3e, and 3.3f. In fact, the above equation of
equilibrium will not change regardless of the point at which the force F is
applied as long as that point is on the line of action of F. This is in accordance
with the principle of transmissibility.
Now let us focus our attention on what may be happening inside the bar.
When F acts at point A of the bar (Figures 3.3a and 3.3d), the two forces F
at points A and C tend to pull the bar apart. The entire bar from A to C is in
tension, and since, in reality, the bar is not perfectly rigid, it will elongate or
stretch slightly. When the force F acts at point B, we can see from Figures
3.3b and 3.3e that only the portion BC of the bar will be in tension and will,
therefore, stretch. The portion AB will not be subjected to any internal forces
or stresses and will not deform. Finally, when the force F acts at point C as
shown in Figures 3.3c and 3.3f, the entire length ABC of the bar will be free
of tension and will remain undeformed. Clearly, the deformation and the in-
ternal forces and stresses (in this case, tension) in the bar depend on the point
of application of the force F. The principle of transmissibility, therefore, may
not be used to determine the internal forces, stresses, and deformations of
bodies.
In this chapter and the next, we will be concerned with the determination
of external forces acting on bodies that may be treated as rigid, and with the
conditions of equilibrium of such bodies; thus, the principle of transmissibility
may be used in the analysis without reservation.

COPLANAR FORCE SYSTEMS

3.4 RIGID BODIES IN TWO DIMENSIONS


If all the forces acting on a body lie in a single plane, the body can be treated
as a two-dimensional or planar body for analysis. Such a body can be ade-
quately defined by its dimensions in the plane of the forces. The third dimen-
3.5 Moment of a Force—Scalar Approach 83

sion of the body, or its depth, in the direction perpendicular to the plane of
the forces, is not needed in the analysis and is usually not specified.
The analysis of two-dimensional systems is simpler than that of three-
dimensional systems mainly because they are easier to visualize and the in-
tuitive scalar approach can often be used. We will present the analysis of two-
dimensional problems in this chapter using both scalar and vector notations.
The vector formulation will then be extended in the following chapter to ana-
lyze three-dimensional systems. The vector approach is usually necessary for
dealing with three-dimensional problems.

3.5 MOMENT OF A FORCE—SCALAR APPROACH


The moment of a force about a point is defined as the force times the perpen-
dicular distance from the point to the line of action of the force. The moment
is a measure of the rotational effect of the force. The term torque is also
sometimes used for the moment.
The magnitude of the moment about a point O, of a force F, is given by

Mo =Fd (3.1)
where M, = magnitude of the moment about point O, F = magnitude of the
force, and d = perpendicular distance from the point O to the line of action
of the force F. The perpendicular distance d is called the moment arm or lever
arm. From Equation (3.1) we can see that the unit of the moment is force times
length. The commonly used units of moment are pound-foot (Ib - ft) or pound-
inch (Ib - in.) in the U.S. Customary system, and newton-meter (N - m) in the
SI system. For large magnitudes of moments, the prefixed units of kilopound-
foot (kip - ft) or kilopound-inch (kip - in.) are commonly used in the U.S. Cus-
tomary system, and kilonewton-meter (KN - m) in the SI system.
The sense of the moment, which may be either clockwise or counterclock-
wise, is the same as the sense in which the force tends to rotate the body about
the point. These senses are usually depicted by curved arrows: the clockwise
sense by Q) and the counterclockwise sense by V).
As an example, consider the wrench being used to tighten a bolt shown in
Figure 3.4a. The force F is applied vertically downward on the handle of the
wrench. The magnitude of the moment of this force about the bolt center O is
simply the product of the magnitude of the force, F, and the moment arm, or
Point O
the perpendicular distance from point O to the line of action of F, which in
this case is the length of the wrench, /. (Note that the line of action of F is
perpendicular to the axis of the wrench.) Therefore, M, = FI. Since the force
F is tending to rotate the wrench and the bolt in the clockwise direction, the
sense of moment Mj is clockwise. It should be realized that the force need
not actually rotate the wrench and the bolt—the moment is produced as long (d)
as the force has the tendency to rotate even when it cannot actually rotate the Figure 3.4
84 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

body. The moment M, is completely specified as


M, = FIQ)
From this relationship, we can see that the magnitude of the moment depends
on two quantities: magnitude of the force, and moment arm. In other words,
we can change the moment magnitude by changing either the force magnitude
or the moment arm, or both.
Now consider the wrench as it is being used to loosen the bolt as shown in
Figure 3.4b. The force F is being applied at an angle @ with the vertical di-
rection, while the wrench axis remains horizontal. Note that since the force is
not perpendicular to the wrench axis, the moment arm d is not equal to /. The
moment arm can be calculated by considering the right triangle ABO shown
in Figure 3.4b. Since the angle A of the triangle ABO is equal to (90 — 8),
the moment arm d = / sin (90 — @) = /cos 8. The moment being applied to
the bolt about point O is then
Moy = Fl cos 6C)
Note that the sense of My is counterclockwise because F has the tendency to
rotate the wrench and the bolt in the counterclockwise direction.
Next, consider the wrench in the inclined position as shown in Figure 3.4c.
The force F is acting at an angle a with the axis of the wrench, and tends to
rotate it in the clockwise direction. The moment arm d (again, the perpendic-
ular distance from O to the line of action of F) is determined from the right
triangle ABO to be d = / sin a. Therefore, the moment of F about point O is

Mo, = Fl sin aQ)


Finally, consider the case of the wrench subjected to the force F which is
collinear with the wrench axis as shown in Figure 3.4d. Since the point O lies
on the line of action of F, the moment arm from O to the line of action of F
is zero, or d = 0. Therefore, the moment of F about point O, Mo, is zero. The
force F here has no rotational effect on the wrench and the bolt; it has only
the tendency to pull on the body under consideration.
3.5 Moment of a Force—Scalar Approach 85

EXAMPLE 3.1

For the frame shown, determine (a) the moment of the 60-lb
vertical force about point B; and (b) the horizontal force applied mae
at C that produces the same moment about point B, and the mo- <a
ment of this force about point A.

Solution

a. From the figure, we can see that the 60-lb force tends to
rotate the structure clockwise about the point B. Therefore, the 8 ft
sense of the moment about B is clockwise. The moment arm
from B to the line of action of the 60-Ib force is

a—i(4riticoss304——5'40: tb

The magnitude of the moment about B is then

Mz = (60 1b)(3.46 ft) My, =207.6lb-ftQ) <


b. Since the moment about B must be clockwise, the sense of
the horizontal force F must be to the right as shown. The
moment arm from B to the line of action of F is given by

d = (4 ft) sin 30° = 2 ft


The magnitude of F that will produce the moment of 207.6
Ib - ft about point B can be determined from the relationship
M, = Fd. Thus we have
207.6 Ib - ft = F(2 ft)
F = 103.8 Ib F =103.81bC) <
From the figure, we can see that the sense of the moment of F
about point A will also be clockwise because F tends to
rotate the structure in that direction about A. The moment arm
from A to the line of action of F is

d = (8 ft) + (4 ft) sin 30° = 10 ft


Therefore, the magnitude of the moment of F about A is

M, = (103.8 Ib)(10 ft) M4 = 10381b-ftQ) <


EXAMPLE 3.2

If a force F is applied to the structure as shown, (a) derive the


expression for the moment at point A in terms of the magnitude
of the force F and the angle 8; (b) determine the moment at A if
1 = 3m,F = 500N, and 6 = 30°; and (c) for /= 3 m,determine
the force of the smallest magnitude that produces a counterclock-
wise moment of 900 N - m about A.

Solution

a. We see from the figure that the moment arm can be


expressed as
d=I/sin0

The magnitude of the moment about A is therefore given by

M,=Fisin0 <
Note: From the above equation of M,, we can see that for 0° <
6 < 180°, the value of sin 0, and hence the magnitude of M,,
will be positive. This indicates that the sense of M, will be
counterclockwise as initially assumed in the figure. However,
for 180° < 6 < 360°, the value of sin 0, and therefore the
magnitude of M,, will be negative, indicating that the sense of
M, is opposite to that initially assumed; that is, it is clockwise.

b. Substituting the numerical values of /, F, and @ in the


expression for M,, we obtain
M, = (500 N)(3 m) sin 30° = 750 N-m
M,=750N-mQ) <
c. Since it is given thatM, = 900 N-m C), we can write
the moment equation M, = Fd as

900 N- m= Fd
From this relationship, we can see that in order for the force
magnitude F to be the smallest, the moment arm d from point A
must be the largest. As shown, the largest moment arm d = |
develops when F is perpendicular to the axis AB of the
structure. Substituting d = / = 3 m in the above equation,

900N-m=F3m) F=300N® <


This result can also be obtained by examining the expression
M, = Fl sin 9, derived in part a. We can see that since / =
3 m is constant, for a given positive (counterclockwise) value
of M,, F is the smallest when sin 0 is the largest. The
largest value of sin 9 = 1 occurs when 6 = 90°, or when F is
perpendicular to the axis AB of the structure. Substituting the
values of M, = 900 N- m,/ = 3 m, and sin 8 = 1 into
the equation M, = F/ sin 0, we obtain F = 300N.
3.6 Cross Product of Vectors 87

3.6 CROSS PRODUCT OF VECTORS


Before we discuss the concept of moment as a vector quantity, it is useful to
understand a type of vector multiplication called the cross product of vectors,
also referred to as vector product.
The cross product of two vectors P and Q, shown in Figure 3.5, is defined (C=IPO@
as a vector C, which has a magnitude of

C=PO sin 6 (3.2)

and a direction perpendicular to the plane containing the vectors P and Q, with
the sense specified by the right-hand rule. The symbol X is used to denote
the cross product. For example, if vector C is the cross product of vectors P
and Q, it is written as
Figure 3.5

C=PxXQ (3.3)

It is important to note that the angle @ in the scalar Equation (3.2) is the
angle between the lines of action of the vectors P and Q, as shown in Figure
3.5. This angle should be measured so that it will always be between 0° and
180°, or 0° = 6 € 180°.
The application of the right-hand rule to determine the sense of vector C
is illustrated in Figure 3.6. If we curl the fingers of our right hand from the
direction of the first vector P, toward the direction of the second vector Q,
then the extended thumb points in the direction of vector C. In applying the
right-hand rule, it is important to maintain the order in which the vectors are
multiplied. This is because cross products are not commutative; that is,

PxQ#QxXP
This can be verified by comparing Figures 3.6 and 3.7, which show the cross
products C = P X Q and D = Q x P, respectively. The vectors C and D Figure 3.6
have the same magnitude, both equal to PQ sin 9, and the same line of action.
However, as we can see, they have opposite senses. Therefore, D = —C, or

PXQ=-(QxXP) (3.4)

The reader is urged to apply the right-hand rule to the two cross products
shown in Figures 3.6 and 3.7, by assuming that the vectors P and Q are acting
in the plane of the page of this book. In the case of the cross product C =
P X Q (Figure 3.6), the thumb of the right hand should point outward or away
from the page, which is the sense of vector C. In the second case of the cross
product D = Q X P (Figure 3.7), the right-hand thumb should point inward DOG ==
or into the page, which is the sense of vector D. Figure 3.7
88 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

A useful property of cross products is that they are distributive; that is,

PX(Q+S+T)=PXQ+PXS+PXT (3.5)

Cross Products of Cartesian Unit Vectors


In Chapter 2, we defined the Cartesian vectors i, j, and k as vectors of unit
magnitude in the mutually perpendicular x, y, and z directions, respectively.
Our objective here is to determine the cross products of these unit vectors with
themselves and with each other.
Consider the mutually perpendicular coordinate system shown in Figure
3.8, with the unit vectors i, j, and k directed along the positive x, y, and z axes,
respectively. The cross product of the unit vector i with itself or i X i produces
a vector with magnitude equal.to (7)(i)(sin 0°) = (1)(1)(0) = 0; thusi X i =
0). Similarly, we can show that the cross products j X j and k X Kk are also
equal to zero.
Next, consider the cross product i X j. This cross product produces a vector
with magnitude equal to (/)(j)(sin 90°) = (1)(1)(1) = 1, or a unit vector. This
unit vector must be perpendicular to the plane containing the i and j vectors,
with its sense specified by the right-hand rule. From Figure 3.9a, we can see
Figure 3.8 that this unit vector is directed along the positive z axis; that is, it is the unit
vector k. Therefore, i X j = k.
The cross product j X ican be determined directly by using Equation (3.4).
jxi=-@xj=-k

Figure 3.9
3.6 Cross Product of Vectors 89

We can verify this from Figure 3.9b by curling the fingers of the right hand
from j to i (as shown by the curved arrow): the thumb points in the negative
z direction.
The cross products of the remaining pairs of unit vectors are similarly de-
termined (Figures 3.9c through 3.9f). The cross products of all pairs of unit
vectors are listed here:

ixi=0 ixj=k ixk=-j


jxi=-k jxj=0 jxk=i (3.6)
kxXi=j kxXj=-i kxk=0
A convenient way to establish the cross products of unit vectors is to draw Counterclock
wise
a circle and write the letters i,j, and k on it in counterclockwise order as shown =F

Me
in Figure 3.10. The cross product of any two unit vectors that are in counter-

=
j
clockwise order is the positive third vector. Conversely, the cross product of
any two vectors that are in clockwise order is the negative third vector. For
example, j X k = i, andk X j = —i. i
It is important to realize that the expressions for the cross products of unit ee ote

vectors given in Equation (3.6) are valid only for the right-handed coordinate —

systems, which are the only type of Cartesian coordinate systems used in this
Clockwise
textbook. A Cartesian coordinate system is considered to be right-handed if,
Figure 3.10
as we curl the fingers of our right hand from the direction of the positive x
axis toward the direction of the positive y axis, the thumb points in the direction
of the positive z axis (Figure 3.11).

Cross Product in Cartesian Vector Form

By using the cross products of unit vectors determined in the preceding para-
graphs, we can develop an expression for the cross product of two Cartesian
vectors in terms of their rectangular components. Consider two arbitrary vec-
tors P and Q, given in terms of their rectangular components as

Pee Pee | dk
0407 30 Janlzk
If the vector C is the cross product of vectors P and Q, then we can write
Figure 3.11 Right-handed coordinate system
Cee ex Oe tate Ne kK XO O52 Qk)

Using the distributive property of cross products, we write the expression as

CaF OMX) tirO0 Xj) + 2.0.0 Xk)


+ P,O.G X i) + P,O,G X j) + P,O.G x k)
+ P.O,(k X i) + P,O,(k X j) + P.0.(k X k)
90 Chapter 3 Equilibrium of Rigid Bodies in Two Dimensions

By performing the cross products of unit vectors from Equation (3.6) we obtain

C= PLO, k= POO Me Oi tO ai
After rearranging terms, we can write

C=PXQ=(€,Q, — P,Q,)i + (P,Q, — P.Q,)i (3.7)


+ (P,Q, — PyQ,)k
From Equation (3.7), we can see that the rectangular components of the vector
C are
C= Orta,
CG $ EOF - ieley (3.8)

C, oe enor, 7 ON

Equation (3.7) is usually expressed in the more convenient form of a deter-


minant as

i ieenk
Ce Px OP ar mer (3.9)
OC; OF OQie

To show that Equations (3.7) and (3.9) are equivalent, we expand the deter-
minant on the right side of Equation (3.9) by minors of the first row. Thus,
ier idfin edd2
P,
POP
QO. Q $
OPE
QO. Q,
| P SIS
Q Zz Q
PAUP
ee
Qz
Pi,
Q BF
oP
ee
Qy

= 10,0, P03) iP, Ope OD) KUO rae


= (P,Q, — P,Q,)i + P,0, — P,O)j + P,Q, > PyO)k
which is the same as the right side of Equation (3.7).

3.7 MOMENTS AS VECTORS


The moment of a force F about a point O may also be defined as the cross
product of a position vector r from point O to any point on the line of action
of F, and the force vector F. Thus

Mo =rXF (3.10)

The moment as defined by Equation (3.10) is a vector, the moment vector,


directed along the axis passing through point O, and perpendicular to the plane
containing the position vector r, and the force vector F. The sense of the
moment vector Mz is specified by the right-hand rule.
3.7 Moments as Vectors 91

To show that this vector definition of moment is consistent with the scalar Moment
definition presented in Section 3.5, let us consider the moment of a force F axis

about point O of the wrench shown in Figure 3.12. The position vector r can
be chosen from point O to any point on the line of action of F. In other words,
r does not have to be perpendicular to F. In Figure 3.12, we have chosen the
position vector r from point O to a point P on the line of action of force vector
F. From Equation (3.2), the magnitude of the moment vector Mz is given by

Mo ='F sin0

As we can see from Figure 3.12, the product r sin 8 = d, where d is the
moment arm or the perpendicular distance from point O to the line of action
of force F. Therefore,
M, = Fd
Figure 3.12
which is in agreement with the scalar definition of the moment magnitude
given by Equation (3.1).
As stated previously, the line of action of the moment vector Mj, passes
through point O and is perpendicular to the plane containing the position vector
r and the force F. This line of action of Mo is called the moment axis. Although
we usually refer to the moment in two dimensions in terms of a point, it is
actually the moment about an axis perpendicular to the plane and passing
through the point.
The sense of Mz is given by the right-hand rule: If we curl the fingers of
the right hand from the direction of the position vector r toward the direction
of the force vector F, then the extended thumb points in the direction of the
moment vector M,. The example in Figure 3.12 shows that the sense of My
is upward because the curl of the fingers has the counterclockwise sense. Note
that the sense of this curl is the same as that in which the force tends to rotate
the body. In other words, the thumb of the right hand gives us the sense of
the moment vector, while the curl of the fingers indicates the sense of rotation
caused by the force. If one of these two senses is known, the other can be
determined by the right-hand rule. In three dimensions, the moment vectors
are denoted by straight arrows pointing in the direction of the moment vector
(direction of the thumb) on top of which a curved arrow is added to indicate
the rotational tendency of the force (curl of the fingers).
Figure 3.13 shows two examples of such three-dimensional moment vec-
tors. Note that the curved arrow is not necessary to specify the sense of these
moment vectors; it is added merely to distinguish moment vectors from force
vectors. Moment vectors obey the law of vector addition and, like force vec-
tors, can be treated as sliding vectors. We will discuss moment vectors further
in Chapter 4.
In two-dimensional systems, since moment vectors are always perpendic-
ular to the plane of the structure and the forces, the senses of moments are Figure 3.13 Moment vectors
92 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

simply denoted by clockwise or counterclockwise curved arrows (Q) or QO)


as discussed in Section 3.5.

Moment in Cartesian Vector Form

Consider a rigid body in two dimensions, subjected to a force F as shown in


Figure 3.14. The moment of F about a point A is given by the cross product

M,=rxF
where r is a position vector from point A to a point P on the line of action of
F. If we resolve the vectors F and r into their rectangular components as shown
in the figure, then we can write

y 2X A(0, 0, 0)
-
ay gia Wao a and Y= int ied
Substituting these expressions of r and F into Equation (3.9), we get
Le jee
Figure 3.14
M, = |" Try 0
Fe ie 0.

Upon expanding the determinant, we obtain the expression for M, in Cartesian


vector form as

Mi = (Fa kok (3.11)

Note that M,, is perpendicular to the plane of the body (xy plane), as expected.
The magnitude of M, is given by

M, =7,F, — ry F, (3.12)
A positive value for the magnitude M, indicates that the moment vector M,
points in the positive z direction (Figure 3.14)—the sense of the moment is
counterclockwise. Conversely, a negative value for M, indicates a clockwise
sense of the moment.
When the origin of the Cartesian coordinate system is located at point A,
about which the moment is desired, then the position vector can be directly
written in terms of the x and y coordinates of point P on the line of action of
the force. In Figure 3.14 we can see thatr, = x andr, = y; that is, the position
vector r = xi + yj. Equation (3.12) then becomes
M, =xF, — yF, (3.13)
However, when the origin of the coordinate system is not located at the
point A where the moment is desired, but at some other point O as shown in
Figure 3.15, then the position vector from A to a point P on the line of action
of F can be written as
3.8 Varignon’s Theorem 93

Figure 3.15

where r, and rp are the position vectors from the origin O to points A and P,
respectively. We see from Figure 3.15 that these two position vectors can be
written in terms of the coordinates of points A and P as

ry salty)
rp =Xpi + ypj
Subtracting r, from rp gives

r= Gp — X,)it (yp — Ya
Substituting r, = xp — x4 andr, = y, — y, into Equation (3.12), we obtain
the expression for the magnitude of the moment as

Mirae aay (yer vary (3.14)


In Equations (3.12) through (3.14), the moment of the force is given as the
sum of the moments of the rectangular components of the force. For example,
in Equation (3.13), the first term, xF,, is the moment of the component F,j
about point A (Figure 3.14), and its sense is counterclockwise or positive. The
second term, — yF,, is the moment of component F,i about A, and it is negative
because its sense is clockwise.

3.8 VARIGNON’S THEOREM


The method of determining the moment of a force about a point as a sum of
the moments of the components of the force about the point provides the most
convenient means of determining moments in many engineering applications.
94 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

The idea was initially developed by the French mathematician Varignon


(1654—1722), and is commonly known as Varignon’s theorem. It states that
the moment of the resultant of concurrent forces about any point is equal to
the sum of the moments of the components about the point.
Consider a concurrent system of forces acting on a body as shown in Figure
3.16. The resultant of the force system is
R=F,+F,+F,+F,
The moment of the resultant R about point O is given by
Mo =rxXR=rxX
(F, + F, + F; + Fy)
where r is the position vector from point O to the point of application of the
forces, P. Using the distributive property of vector cross products as given by
Equation (3.5), we can write

Mo =rxXR=rxXF,+rxXF,+rxXF,+rxk, G13)

Figure 3.16 Equation (3.15) shows that the moment of the resultant R about O is equal to
the sum of the moments of the components F,, F,, F3, and F, about O. As
shown here, Varignon’s theorem may be applied to any number of concurrent
forces acting in any arbitrary direction.

3.9 PROCEDURE FOR DETERMINING MOMENTS


The following step-by-step procedure can be used to determine the moment
of a force acting in the plane of a two-dimensional body about a point on the
body.

1. Select the orientation of the x-y coordinate system. Usually the x and y
axes are chosen in the horizontal (positive to the right) and vertical (positive
upward) directions, respectively. However, they may be oriented in any con-
venient direction provided that they are perpendicular to each other. If the
dimensions of the structure are given in an inclined direction, selecting the x
(or y) axis in that direction may considerably expedite the solution (see Ex-
ample 3.5).
2. Select the point of application of the force anywhere along the line of
action of the force, or even outside the physical boundaries of the body. For
some problems, the analysis is simplified by selecting a point of application
of the force so that the line of action of one of its rectangular components
passes through the point about which the moment is to be determined, thereby
eliminating the moment of that component (see Example 3.4).
3. Resolve the force into components in the x and y directions.
4. Adopt a sign convention for the senses of the moments. The counterclock-
wise sense is assumed to be positive in the examples given in this chapter, to
be consistent with the vector definition of the moment, and the clockwise sense
is considered to be negative.
5. Determine the moment about the specified point by algebraically summing
3.9 Procedure for Determining Moments 95

the moments of the components of the force about the point. The magnitude
of the moment of the x force component is given by the product of the mag-
nitude of the force component and the distance in the y direction from the
point in question to the line of action of the component. Similarly, the mag-
nitude of the moment due to the y force component is equal to its magnitude
times the moment arm, which is the distance in the x direction from the point
to the component’s line of action.

EXAMPLE 3.3

Determine the moment of the 3-kKN force about points A and B


of the pole.

Solution—Scalar Approach

Coordinate system. A coordinate system with origin at point A


and with the x and y axes in the horizontal and vertical
directions, respectively, is chosen.

Point of application of the force. Point C at the top of the pole


is the point of application of the 3-KN force.

Rectangular components of the force. The magnitudes of the


components of the 3-kKN force in the x and y directions are

F..= F cos 8 = (3 kN) (cos 30°) = 2.60kN


F, = F sin 0 = (3 KN) (sin 30°) = 1.50kN

Sign convention. A counterclockwise sense of the moment is


assumed to be positive, and a clockwise sense, negative.

Moment about A. We can see that the vertical or y distance


from A to the line of action of the horizontal component F, is
5 m. Also, F, has the tendency to rotate the body in the
clockwise sense about point A; thus, its moment about A is
considered negative.
As for the vertical component Moe since its line of action
passes through point A, its moment about A is zero. Therefore

M, = —(2.6 KN)(5 m) = —13.0kN-m


M, = 13.0kN-mQ) <
Moment about B. Since the line of action of the vertical
component F,, passes through the point B, its moment about B

(continued)
96 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

EXAMPLE 3.3 (concluded)

is zero. Therefore, the moment of the 3-kN force about B is


equal to the moment of its horizontal component F,. Thus

M, = —(2.6 kN)(2 m) = —5.2 kN: m


M,=5.2kN-mQ) <

Solution—Vector Approach

The 3-kN force can be expressed in the Cartesian vector form


as
F = 2.61 + 1.5j (KN)

Moment about A. The moment of F about point A is given by


the cross product
M, =lac XF

where r4c is the position vector from A to the point of ;


application C of the force F. From the figure, we can see that

Myc = (5 m)j
Therefore,

My, = lac X F = (5 m)j X [(2.6 KN)i + (1.5 KN)j]


= (5 m)(2.6kN)(j X i) + (Sm)(1.5 kN) X j)
= (13)§ X i) + (7.5) X j) (KN- m)

Recalling from Equation (3.6) that }X i = —kandj X j =


0, we obtain

M, = —13k (kKN-m) <


Note that the negative magnitude of the moment vector implies
a clockwise sense of rotation about A. (If we point the thumb
of our right hand in the negative z direction or into the plane of
the page, the curl of the fingers is clockwise.)

Moment about B. The position vector from B to the point of


application of the force C is given by

Mec = (2 m)j
Therefore,
Mg, = rac X F = (2 m)j X [(2.6KN)i + (1.5 KN)j]
M, = —5.2k(kKN-m) <
3.9 Procedure for Determining Moments 97

EXAMPLE 3.4

A bracket is attached to the flange of a column. Determine the


moment of the 250-lb force about the connection point A.

Solution—Scalar Approach

Coordinate system. A coordinate system with its origin at


point A and with x and y axes in the horizontal and vertical
directions, respectively, is chosen.

Point of application of the force. In this solution, we will first


use the actual point of application of the 250-lb force, shown
as point B. Following this, we will present an alternate solution
using a different point of application of the force for
comparison.

Rectangular components of the force. In this problem, the


direction of the 250-lb force is not given as an angle with
respect to the x or y axis, but in terms of the slope of its line of
action. The numbers 3 and 4 on the vertical and horizontal
sides, respectively, of the small right-angled triangle drawn on
the line of action of the force imply that the ratio of the
magnitudes of its components in the vertical and horizontal
directions is equal to *%4 or

From the right triangle, labeled abc, on the line of action of the
force, we obtain the hypotenuse ab as equal to

ab = V(4? + GY =5
If 6 denotes the angle between the line of action of the force
and the x axis, then from the triangle abc, we obtain

3 + 3
in§@=—=0.
sin 8 5 0.6 cos 0 =-=0.8
5 tanan 8 =—=0.75
4

The magnitudes of the components of the 250-lb force in


the x and y directions can now be computed as

F,. = F cos 8 = (250 1b)(0.8) = 200 Ib


F, = F sin 6 = (250 1b)(0.6) = 150 Ib

Moment about A. Considering the counterclockwise sense as


positive, and observing that the moment of F, about A has a

(continued )
EXAMPLE 3.4 (concluded )

clockwise sense, whereas the moment of F,, about A has a


counterclockwise sense, we can write

M, = —(200 Ib)(3 ft) + (150 Ib)(2 ft) = —3001b =ft


M, = 300lb-ftQ) <
Alternate Solution—Scalar Approach By moving the force
F = 250 lb along its line of action to point C, we can eliminate
the moment of the F,, component because its line of action
then passes through A. The location of point C is determined by
considering the right triangle BCD, with the angle B equal to
8, which is the angle of inclination of the force F with the x
direction. For the triangle BCD,

CD_ an 9
BD =<" {i =
iw

As BD = 2 ft,
3 3
CD = (3)a0 - (2)eft) = 1.5 ft

Therefore, the moment arm in the y direction from point A to


the line of action of the horizontal component F, is equal to
AD — CD = 3 — 1.5 = 1.5 ft. The moment aboutA can now
be computed as
M, = —(200 lb)(1.5 ft) = —300 lb - ft
M, = 300lb-ft@Q) <

EXAMPLE 3.5

A uniform 5-m boom weighing 2 kN is supported by a cable. The


tension in the cable produces the same moment about the support
point A as that caused by the weight of the boom, but with an
opposite sense. Determine (a) the tension in the cable if a = 30°,
and (b) the direction and the magnitude of the minimum tension
in the cable.

Solution—Scalar Approach

Coordinate system. Since the dimensions of the boom are


given in its longitudinal direction, and the direction a of the
cable tension T is specified with respect to the longitudinal axis
of the boom, it is convenient to orient the x axis in the
longitudinal direction of the boom as shown. The y axis is
perpendicular to the x axis, with the origin of the coordinate
system located at point A.
(continued)

98
EXAMPLE 3.5 (concluded)

Points of application of the forces. Since the actual points of


application of the weight and the cable tension, points C and B
respectively, are located on the x axis, the moments of the x
components of these forces about A will be eliminated. We will
use these points of application in the analysis.

Moment of the weight about A. The magnitudes of the


rectangular components of the 2-KN weight in the x and y
directions are

W,. = W cos 27° = (2)(cos 27°) = 1.78 kN


W, = W sin 27° = (2)(sin 27°) = 0.91 kN

Considering the counterclockwise sense as positive, we can


write the equation for the moment of the weight about A as
M, = —(0.91)(2.5) = —2.28 KN- m
The clockwise (negative) sense of M, can be visually checked.

Moment of the cable tension about A. The rectangular


components of the cable tension T in the x and y directions are
T, =T cosa pe Sine

The moment of the cable tension T about point A is given by

M, = +(Tsina)(5) = +5ST sina

The moment of T about A must be equal but opposite to the


moment of W about A, which is — 2.28 kN - m. Therefore, the
moment of cable tension T must be + 2.28 kN - m. Substituting
M, = +2.28 kN - m into the above expression, we obtain

2.28 = 51 sini (a)

a. If a = 30°, then the cable tension T can be determined as

DDS — Sines Om) T=0.91 kN <

b. From Equation (a), we can see that since the moment


(2.28 kN - m) and the moment arm (5 m) are constant and
cannot be changed, the cable tension T can only be reduced by
increasing sin a. Also, since sin a attains the maximum value
of 1 at a = 90°, the tension in the cable T will be minimum
when a = 90°, or when the cable is perpendicular to the boom
as shown. Therefore,
=90° <
By substituting a = 90° into Equation (a), we obtain
2.28 = 5T,,,,(sin 90°) = S5T in
From which we determine the magnitude of the minimum
tension in the cable as
Tmin = 0.46kN <

99
100 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

3.10 COUPLES
A couple is a pair of equal, opposite, and parallel forces (Figure 3.17a). The
distinguishing feature of couples is their tendency to rotate bodies without
causing any tendency for translation.
Consider a pair of equal and opposite forces, F and — F, shown in Figure
3.17a. The perpendicular distance between their parallel lines of action is d.
Because the two forces have equal magnitudes but opposite senses, their re-
sultant force is zero—thus the two forces, or the couple, cannot translate the
body. However, couples do tend to rotate bodies, because the sum of moments
of the two forces forming the couple about any point on the body is not zero.
Obviously, the couple shown in Figure 3.17a tends to rotate the body in the
counterclockwise direction.
The moment of a couple is a measure of the rotational effect of the couple,
Figure 3.17
which is the combined rotational effects of the two forces forming the couple.
The magnitude of this moment is given by

M = Fd e(GiL6}

where M is the magnitude of the moment of the couple, F is the magnitude of


the forces forming the couple, and d is the moment arm or the perpendicular
distance between the forces’ parallel lines of action. The units used for the
moments of couples are the same as for the moments of forces. Couples are
usually expressed in terms of their moments, with curved arrows indicating
the sense of rotation (Figure 3.17), either counterclockwise (C)) or clockwise
(©).
To illustrate the difference between the effects of a couple and the moment
of a single force, consider the lug wrench being used to tighten a bolt of a
wheel as shown in Figure 3.18. In Figure 3.18a, two equal and opposite forces
of 40 Ib are applied on the two handles of the wrench, thereby forming a
couple, with a clockwise moment of magnitude equal to (40 lb)(24 in.) = 960
Ib - in., about the bolt. Since the two 40-lb forces are in opposite directions,
they cancel each other, and no force is exerted on the bolt. In Figure 3.185,
only one force of 80 lb is applied. This force produces a moment of the same
magnitude, (80 lb)(12 in.) = 960 Ib: in., and the same sense about the bolt.
However, in this case, since there is no other force to cancel the 80-lb force,
this 80-lb force is also exerted on the bolt.
It is important to understand that unlike the moment of a force (Figure
3.19), which may change from one point to another as the moment arm from
the point to the line of action of the force changes, the moment of a couple
remains the same for all points on the body. This can be shown by summing
the moments of the two forces of the couple about an arbitrary point A on the
body as in Figure 3.20. Considering counterclockwise as positive, we can
(b) express the moment about A as
Figure 3.18 M, = F(d +a) — F(a) = Fd©)
3.10 Couples 101

which is the same as the moment of the couple given by Equation (3.16).
Considering the moment about another point B on the body,
M, = —F(b) + F(b + d) = Fd)
we obtain the same result. This shows that the moment of a couple is the same
for all points on the body. So, the moment of a couple depends only on the
magnitude of the forces forming the couple, the perpendicular distance be-
tween their parallel lines of action, and the rotational sense of the couple.

Couples as Vectors
The moment of a couple is defined as the cross product of a position vector r, Figure 3.19 Moments of a single force
from any point on the line of action of —F to any point on the line of action
of F, and the force vector F (Figure 3.21). Thus,

M=rxF Cr)

The moment as defined by this equation is a vector directed along an axis


perpendicular to the plane containing the two forces that form the couple, and
with the sense established by the right-hand rule.
To verify that this vector definition of the moment of a couple is consistent
with the scalar definition presented earlier in this section, let us consider the
moment of the couple shown in Figure 3.21. Note that the position vector r
can be chosen from any arbitrary point C on the line of action of —F to any
arbitrary point D on the line of action of F. Using Equation (3.2), we obtain
Figure 3.20 Moment of a couple
the magnitude of the moment vector M as

M =rF sin 8

From Figure 3.21, we can see that the product r sin 8 = d, where d is the
perpendicular distance between the lines of action of the forces F and —F.
Therefore

M=Fd
which is identical to the scalar definition of the magnitude of the moment
given by Equation (3.16).
As mentioned before, the line of action of M (moment axis) is perpendicular
to the plane containing the two forces forming the couple. The sense of M is
in the direction of the thumb of the right hand when its fingers are curled in
the direction in which the couple tends to rotate the body. In Figure 3.21, we
can see that the couple formed by F and —F tends to rotate the body coun-
terclockwise; thus, we curl the fingers of the right hand counterclockwise.
Since the thumb points in the upward direction, the sense of M is upward.
Unlike the moment vector of a single force whose line of action must pass
through the point about which the moment is evaluated, the moment vector of
a couple may be applied at any point on the body. To show this, consider the Figure 3.21
102 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

moment of the couple shown in Figure 3.22 about an arbitrary point A. The
sum of the moments of the two forces is given by
M, = Tap X F + Pac X (—F) = ap — Fac) X F=Prep XF
— an

which is the same as Equation (3.17). Obviously then, the moment vector of
a couple does not depend on the point about which the moment is evaluated.
Therefore, the moment of a couple may be treated as a free vector and may
Figure 3.22
be applied at any point on the body with its line of action perpendicular to the
plane of the forces forming the couple, and its sense representing the rotational
sense of the couple in accordance with the right-hand rule.

Equivalent Couples
Couples formed by different pairs of forces are considered to be equivalent if
they have the same moment vector M. Equivalent couples have the same effect
on rigid bodies. Consider the four couples acting on a block as shown in Figure
3.23. Since all of these couples produce the same moment vector M, in the
upward direction and with magnitude of 30 Ib - ft, they are considered equiv-
alent and have the same effect on the block. Note that the pairs of forces
forming the four couples are different. In Figure 3.23a, the pair of 15-lb forces
act in the plane ABCD and have a moment arm of 2 ft. They form a counter-
clockwise couple with a moment of 30-Ib - ft magnitude. The moment vector
is perpendicular to the plane ABCD and is directed upward as shown. In Figure
3.23b, the two 30-lb forces lie in the plane ABCD and are separated by 1 ft.
They form a couple whose moment vector is identical to that of Figure
3.23a in terms of the magnitude, direction, and sense. The pair of 10-lb forces
in Figure 3.23c lie in the lower plane EFGH, which is parallel to the upper
plane ABCD. These 10-lb forces have a moment arm of 3 ft; hence they also
produce a counterclockwise couple with a moment of 30 Ib - ft. The moment
vector in this case is perpendicular to the plane EFGH and is directed upward.
Since the planes EFGH and ABCD are parallel, the moment vector M is also
perpendicular to the plane ABCD, and will have the same rotational effect on
the body as the moment vectors in Figures 3.23a and 3.23b. Therefore, the
couple of Figure 3.23c is equivalent to the couples of Figures 3.23a and 3.23b.
Similarly, we can see that the couple depicted in Figure 3.23d is equivalent to
the others.
Next, consider the couple shown in Figure 3.24a. Even though the two
15-lb forces with a moment arm of 2 ft form a counterclockwise couple with
a moment magnitude of 30 Ib: ft, this couple is not equivalent to those in
Figure 3.23. This is because the forces acting in the inclined plane BDJ/ pro-
duce a moment vector M perpendicular to this inclined plane, whereas the
direction of M of couples in Figure 3.23 is vertical. The couple in Figure 3.245
is also not equivalent to those previously considered because its sense is clock-
wise. Finally, the couple of Figure 3.24c has a moment of magnitude 15 lb - ft,
3.10 Couples 103

= 30 lb-ft

15 1b

M=30|b:ft
co
cu

101b

(c) (d)
Figure 3.23 Equivalent couples

~ fM=30\b-ft
CA =30 lb-ft Va ee

101b

(b) (c)
104 Chapter 3 Equilibrium of Rigid Bodies in Two Dimensions

which is not the same as the 30-Ib - ft couples shown in the other examples.
Note that the four couples of Figure 3.23 are equivalent to each other, whereas
the three couples of Figure 3.24 are neither equivalent to each other nor to any
in Figure 3.23.
From the above discussion, we realize that in order for couples to be equiv-
alent, they must have moment vectors with the same magnitude, direction, and
sense. For the direction of the moment vectors to be the same, the pairs of
forces forming equivalent couples must lie either in the same plane, or in
parallel planes.
=
EXAMPLE 3.6

1501b 1501b A couple consisting of two 150-lb forces is acting on the beam.
(a) Determine the moment of the couple, and (b) replace this
couple by an equivalent couple consisting of two vertical forces
on-\e acting at the supports.

|en 10 ft —+}-5 ei Solution—Scalar Approach


2 ft
a. The moment arm is 17 ft. Hence, the magnitude of the
moment of the couple is given by

M = Fd = (150 1b)(17 ft) = 2550 Ib - ft

We see that the couple consisting of two 150-Ib forces tends to


M=2550 lb-ft rotate the beam in the counterclockwise sense. Therefore, the
sense of the moment of the couple is also counterclockwise.
M = 2550 Ib: ftQ) <
The moment of the couple is shown in the figure.

255 Ib 255 lb b. Since the two vertical forces acting at the support points A
and B must form a counterclockwise couple, the force at A
A B must act downward, the force at B, upward. The moment arm
or the horizontal distance between points A and B, which
are the points of application of the forces, is 10 ft. Therefore,
10ft the magnitude of the forces can be computed from the equation
M = Fdas
2550 Ib - ft = F(10 ft) F=255lb <

y Solution—Vector Approach

a. The moment of the couple is given by the cross product


M = r X F. The position vector r, joining the lines of action
of the forces, can be written as r = 17i, and the force vector
given by F = 150j. Therefore, the moment of the couple is

M =r X F = (17 ft) i X (150 Ib)j= 2550k (Ib - ft)


M = 2550k (Ib: ft) <«
(continued)
3.10 Couples 105

EXAMPLE 3.6 (concluded)

The sense of M in the positive z direction implies a counter-


clockwise sense of rotation for the couple as shown. It should
be noted that M could also have been computed by using
the position vector r = —17i from the line of action of F to
that of —F, and the force F = —F = — 150j, as

M =r X F = (—17i) X (— 150j) = 2550k (Ib - ft) oe M = 2550k


(Ib: ft)
b. As shown in the figure, the position vector from the support
A to B isr = 10i, and the vertical force can be written in the
vector form as F = Fj. The magnitude of the force F can
be computed from the cross product M = r X Fas

2550k = (101) X Fj = 10Fk

from which we obtain

Fa) 010 F = 255j (lb) <

EXAMPLE 3.7
75 Ib
Determine the moment of the couple acting on the rectangular ie 323
7° : oy) i
plate shown. ge 0.5 ft
- 3 ft +|
Solution—Scalar Approach When forces are inclined with
respect to the dimensions of the body as given, it is usually
convenient to resolve them into rectangular components
and determine the moment of the couple as the sum of the
moments of two couples formed by the components of
the forces. The x and y components of the 75-Ib forces are
F,, = (75 Ib)(cos 32°) = 63.6 Ib
F, = (75 Ib)(sin 32°) = 39.7 Ib
As shown, now we have replaced the single couple consisting Ay

of two 75-lb forces with two couples, one formed by the two 63.615 | 39.71b 75 Ib
63.6-Ib horizontal components with a moment arm of 0.5 ft, yee 32°
and another by the two 39.7-Ib vertical components with a «-___Sa9
75ibe | a7, 63.6ieIb
moment arm of 3 ft. Considering counterclockwise moments as
|. 3 ft -|
positive, we can determine
M = + (63.6 Ib)(0.5 ft) + (39.7 Ib)(3 ft) = 150.9 Ib - ft Co lb-ft
M = 150.91b-ftQ) <
Note: The sense of the moments of both couples formed by F,
and F, components is counterclockwise or positive.
106 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

3.11 RESOLUTION OF A FORCE INTO


A FORCE-—COUPLE SYSTEM
As stated in Section 3.3, the principle of transmissibility enables us to move
a force to any point on its line of action without altering its effect on a rigid
body. In this section, we will discuss how a force can be moved to a point
which is not on its line of action, without changing the conditions of equilib-
rium or motion of the rigid body. Such transfers are necessary to determine
the resultants of nonconcurrent forces acting on rigid bodies, which we will
consider in the next section.
Consider a force F acting at a point A on a rigid body as shown in Figure
3.25a. In order to move this force to another point, B, which is not on the line
of action of F, we apply two equal but opposite forces, F and —F, at point B
as shown in Figure 3.255. Since F and — F at B cancel each other, they have
no effect on the rigid body. A careful look at Figure 3.255 reveals that the
original force F at point A and the opposite force — F at point B form a couple
with a moment of magnitude M, = Fd, where d is the moment arm or the
perpendicular distance of B from the line of action of F at A. Note that the
moment of this couple is the same as the moment of F acting at the initial
point A about the new point B.

Figure 3.25

Since the moment of a couple is the same at all points on a rigid body, this
couple can be applied anywhere on the body. In Figure 3.25c, the two forces
forming the couple (F at A and —F at B) are replaced by the moment of the
couple with a curved arrow indicating its sense. Note that we are now left with
only one force F, which is acting at point B as desired. The three force systems
shown in Figure 3.25 are equivalent in the sense that they will have the same
external effect on the rigid body.
Therefore, a force acting at any point on a rigid body may be moved to any
other point if a couple, with moment equal to the moment of the force acting
at the initial point about the new point, is added on the body.
This transformation is shown using vector notation in Figure 3.26. The
moment of the couple is given by
M,=rxF
where r is the position vector from the new point B to the initial point A as
shown in Figure 3.26b. Figure 3.26c shows that the line of action of the mo-
3.11 Resolution of a Force into a Force-Couple System 107

(a) (b) (c)


Figure 3.26

(a) (b)
Figure 3.27

ment vector of the couple M, is perpendicular to the plane containing the


vectors F and r, and its sense is given by the right-hand rule. Since in this
particular case the sense of the moment of F about B is clockwise, vector M,
is directed downward in Figure 3.26c. It should be noted that the subscript B
of the vector Mz simply implies that the moment of this couple is equal to the
moment of F about B. Since M, is the moment of the couple, it is a free vector
and may be applied anywhere on the body.

Reduction of a Force-Couple System to a Single Force


In many engineering applications, we encounter problems that require the re-
duction of a given force-couple system to a single force that has the same
effect on the rigid body. Since this is the reverse of the transformation de-
scribed in the preceding paragraphs, we achieve it by reversing the procedure
just described.
Consider a force-couple system, consisting of a force F at point A and a
couple of moment M acting on a rigid body as in Figure 3.27a. Figure 3.27b
shows another system consisting of the single force F acting at a different
point B on the body. The two systems will be equivalent if the moment of F
at B about A is equal to the moment of the couple M given in Figure 3.27a.
In other words, the couple that must be added to the body due to the transfer
108 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

of F from point B to point A in Figure 3.27b, should be equal to the couple of


moment M given in Figure 3.27a. From Figure 3.27b, the magnitude of the
moment of F about A is given by
M, = Fd
Substituting M, = M in the above equation, we obtain
M
d=
F
Therefore, the new point B (or the new line of action of F) should be located
with respect to the initial point A so that the moment arm (or the perpendicular
distance from A to the new line of action of F through B) is equal to M/F,
and the sense of the moment of F about A is the same as the sense of the given
couple.
Thus, a force-couple system acting on a rigid body may be reduced to a
single force by moving the force to a point selected so that the moment of the
force acting at the new point about the initial point is equal to the moment of
the given couple. .
Since the force F is a sliding vector, it is not restricted to the single point
of application B, but may be considered to be acting at any point on its new
line of action passing through B. The equation of the new line of action of F
(which must be parallel to its initial line of action) can be obtained by using
Equation (3.13) derived previously in Section 3.7. In Figure 3.27c, F is shown
to be acting at an arbitrary point C on the new line of action. The moment of
F about A can be written in terms of its components F, and F, as [Equation
(3.13)]
M, = edly = Vie

Note that in this equation, counterclockwise moments are assumed to be posi-


tive, and x and y are the coordinates of point C with respect to the x-y coor-
dinate system with origin at A as shown in Figure 3.27c. Setting M, = M,
which is the moment of the couple of the initial force-couple system (Figure
3.27a), we obtain the equation of the new line of action of F as

MSO eels
The x and y intercepts of this line can be obtained by setting y = O and
x = O, respectively.
3.11 Resolution of a Force into a Force-Couple System 109

EXAMPLE 3.8
Point O
A 50-N force is applied to the handle of a wrench. Replace this
force by an equivalent force-couple system with the force acting
at the bolt center O.

+300 |

Solution—Scalar Approach By applying two equal but


opposite 50-N forces at O, we can see that the couple, formed
by the 50-N upward force at O and the 50-N downward force at
A, has a clockwise sense and a moment of magnitude

Mo = Fd = (50 N)(300 mm) = 15,000 N - mm


Mo = 15,000N-mmQ) <
Therefore, the 50-N force on the handle of the wrench is
equivalent to a 50-N downward force at the bolt center O and a
clockwise couple of moment 15,000 N - mm.

15,000 N':mm

Solution—Vector Approach The moment of the couple of the wy,


equivalent force-couple system is equal to the moment of F at
A about O, or
Mj=rxF
As shown, r is the position vector from O to A and is given by
r = 300i. The force vector F can be written as F = —50j.
y
Therefore, the moment of the couple is
Mo = (300i) X (—50j) = — 15,000k F=—50j(N)
M, = —15,000k (N- mm) <

The equivalent force-couple system at O, consisting of F =


—50j (N) and M, = —15,000k (N - mm), is shown. Since the
moment of the couple Mz is a free vector, it could have been
applied anywhere on the wrench.
110 Chapter 3 Equilibrium of Rigid Bodies in Two Dimensions

EXAMPLE 3.9

A 200-lb force and a couple consisting of a pair of 50-Ib forces


are acting on a lever as shown. Reduce this force-couple system
to an equivalent single force. Determine the point of application
of the equivalent single force on the lever.

Solution—Scalar Approach The moment of the couple


formed by two 50-lb forces is
M = 50(2) = 100 Ib - ft)
Since this couple may be applied anywhere on the lever, it is
applied at A, the point of application of the 200-lb force.
In order to eliminate the couple, the 200-Ib force must be
moved from A to another point B so that its moment about A is
equal to that of the couple. Since the couple moment is
clockwise, the moment of the 200-lb force at B about point A
must be clockwise; therefore, point B must be located to the
right of point A as shown. The moment arm or the perpendicular
(horizontal) distance from A to the line of action of the 200-lb
force at B can be determined from the relationship M = Fd as
100 = (200)(d) d=0.5 ft
The distance AB in the longitudinal direction of the lever is
then given by

0 1D
AB = le Ont
cos 70°

Thus, the equivalent system consists of the 200-lb force


acting at B with

OB = 5.46 ft <

3.12 RESULTANTS OF NONCONCURRENT


COPLANAR FORCE SYSTEMS
When analyzing the conditions of equilibrium or motion of rigid bodies, it is
generally necessary to reduce the given system of forces acting on the bodies
into simpler systems, called resultants, that have the same external effect on
the bodies. The resultants of concurrent coplanar force systems were consid-
ered in Chapter 2. In this section, we will present the procedures for deter-
mining the resultants of nonconcurrent coplanar force systems.

Reduction of a System of Forces to a Force-Couple System


The procedure for reducing a system of forces and couples acting on a rigid
body to an equivalent one force—one couple system is essentially an extension
3.12 Resultants of Nonconcurrent Coplanar Force Systems 111

Figure 3.28

of the concept of resolution of a single force into a force-couple system as


discussed in the preceding section. Consider a general system of coplanar
forces and couples acting on a rigid body as shown in Figure 3.28a. This
system may be reduced to an equivalent system consisting of one resultant
force and one resultant couple. First, each force must be transferred to a
common reference point O by adding to the body, for each force, a couple
with moment equal to the moment of the force about O. As shown in Figure
3.28), this step produces an equivalent system consisting of a system of forces
that is concurrent at point O and a system of couples that may be applied
anywhere on the body. The concurrent forces in Figure 3.285 are then added
112 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

vectorially to obtain a resultant force R,

R =F
Similarly, all the couples are added vectorially to obtain a resultant couple
Mao, with
Mro = =M (3.18)
Since we are considering only coplanar force systems in this section, the mo-
ment vectors of all the couples, including the resultant couple, are perpendic-
ular to the plane of the forces. Therefore, the moment of the resultant couple
is simply the algebraic sum of the moments of all the forces about point O
plus the moments of any couples acting directly on the body, or

Mro = 2M (3.19)

The resultant force-couple system at point O thus obtained is shown in Figure


3.28c. This force-couple system is equivalent to the initial system of forces
and couples given in Figure 3.28a and, therefore, will have the same external
effect on the rigid body.
It is usually convenient to determine resultants by resolving the forces into
their x and y components, and then obtaining the x and y components of the
resultant force by algebraically summing the x components and the y com-
ponents, respectively, of all the forces acting on the body. Thus,

R,=3F, R= 3A,
The resultant force R is then given by
Re hie ke
The resultant couple is determined by algebraically summing the moments of
the x and y components of all the forces about the moment center O, as well
as the moments of all the couples acting on the body.

Reduction of a Resultant Force-Couple System


to a Single Resultant Force
The procedure described in the preceding paragraphs reduces a general system
of forces and couples to a one force—one couple system acting at any arbitrarily
chosen point on the rigid body. This resultant force-couple system may be
further reduced to a single resultant force by using the procedure discussed in
Section 3.11. The force-couple system at point O (Figure 3.28c) consisting of
the resultant force R and the resultant moment Mgo, is reduced to the single
resultant force R by moving R from its point of application O to a new point
A so that its moment about O, while acting at the new point A, is equal to the
moment Mav of the resultant couple (Figure 3.28d). In order for the moment
3.12 Resultants of Nonconcurrent Coplanar Force Systems 113

about O of R at A to be equal to Mav, the new point A must be located so


that the perpendicular distance from the initial point O to the new line of action
of R through point A is d = Mpo/R, and the sense of the moment of R about
O is the same as the sense of My (Figure 3.28d).
The single resultant force R acting at point A as shown in Figure 3.28d is
equivalent to the initial system of forces and couples given in Figure 3.28a,
and hence will have the same external effect on the rigid body. The equation
of the new line of action of R is given by
Mro = xR, — yRy (3.20)
In Equation (3.20), x and y are the coordinates of an arbitrary point B on the
new line of action of R, with respect to the x-y coordinate system with origin
at O as shown in Figure 3.28e.

Simple Forms of Coplanar Force Systems


Three simple types of coplanar force systems are described here. Being able
to identify these systems may expedite the analysis of their resultants.

Concurrent Force Systems When all the forces are concurrent at the same
point and no couple is acting on the body, the sum of moments about the point
of concurrency is zero. The resultant force R acting at the point of concurrency
may be calculated using the procedures described in Chapter 2.

Parallel Force Systems When the lines of action of all the forces are parallel
to each other, analysis can be expedited by orienting one of the coordinate
axes in the direction of the forces. The resultant force R is parallel to the forces
with its magnitude equal to the algebraic sum of the forces.

Systems with Zero Resultant Force When the resultant force of a system
of forces is zero (i.c., R = SF = 0), its resultant couple Mpo may not
necessarily be zero. Such a system of forces cannot be reduced to a single
resultant force, but it may be reduced to a single resultant couple.

EXAMPLE 3.10

A cantilever beam is subjected to the system of forces shown. 1S kN SKN


Determine the resultant force-couple system at the support. |
sb
eee Alen is i>}

(continued)
EXAMPLE 3.10 (concluded)

Solution—Scalar Approach

Coordinate system. Since the lines of action of the three given


forces are in the vertical direction, a coordinate system with
origin at the support point O and with x and y axes in the
horizontal and vertical directions, respectively, is chosen.
Mrg= 65 kN-m
Resultant force. The component of the resultant force in the x
direction is R, = 0. The y component of the resultant force is

R, = —20
+ 15-5 = —10kN
So, the magnitude of the resultant force is R = 10 KN and it
acts in the downward direction as shown.

Resultant couple. The moment of the resultant couple is equal


to the algebraic sum of the moments of the three forces about
the moment center O. Considering counterclockwise moments
as positive, we obtain

Mro = 20(5) — 15(3) + 5(2) = 65 KN: m


Thus, the resultant force-couple system at O is
R=10kNQ) and Mgy=65kN-m(Q) <
This resultant system is equivalent to the original system of
three forces, and has the same external effect on the beam.

Solution—Vector Approach The three given forces are


written in vector form as
F, = —20j F, = 15j F,= —Sj

Resultant force. The resultant force R is given by

R= XF=F,+F,+F;,
R = —20j + 15j — 5j = —10j (kN)

Resultant couple. The moment of the resultant couple is


obtained by summing moments of the three forces about O:
Meo = >M=r, XF,+r,XF,4+1r,XF;
As shown, the position vectors from O to the lines of action of
F,, F,, and F,, respectively, are
r, = —di r,= —3i r3= —2i
Using the vector expressions of the forces and the position
vectors, the moment of the couple is determined as

Mro = (— Si) X (— 20) + (—3i) X (5j) + (— 21) X (— Sj)


= 100k — 45k + 10k = 65k (kN - m)
Thus, the resultant force-couple system at O is

R = — 10j (KN) and Mero = 05K (KN -m) <

114
3.12 Resultants of Nonconcurrent Coplanar Force Systems 115

EXAMPLE 3.11

Determine the resultant of the forces and couples acting on the 7 kip 5 kip
beam shown. 30 kip-ft ee kip-ft
Ooaas eS
Solution—Scalar Approach

Coordinate system. The origin of the x-y coordinate system is ime aot =, ft
eSett 2 ft 2 ft
located at the left support point A, which will be used as the
reference point to sum moments. Note that since the 3-kip force
passes through A, its moment will be eliminated.

Resultant force. The x and y components of the resultant force Mp4 =103.5 kip-ft
are
R,=0 R,=3-7-5=—9kip
Thus, the magnitude of the resultant force is R = 9 kip and it
is acting in the downward direction as shown.

Resultant force-couple system at point A. We determine the


moment of the resultant couple by summing the moments
of the forces about A and the moments of the couples.
Considering the counterclockwise moments as positive, we
write
Mea = —7(8.5) — 5(12) + 30 — 14 = —103.5kip- ft
Therefore, the resultant force-couple system at point A consists
of the force R = 9 kip and the couple of moment Mp, =
103.5 kip : ft Q) as shown.

Location of the single resultant force. Since the sense of the


couple Mp, is clockwise, the resultant force R must be
moved to the right from point A a horizontal distance of

The single resultant force is located at point B, which is 11.5 ft


to the right of A. We can verify from the figure that the sense
of the moment of R acting at B about A is indeed clockwise.
Thus the resultant force is

R=9 kip ® at a distance of 11.5 ft from the left supportA <


This single resultant force at point B is equivalent to the
resultant force-couple system at A, and to the given system of
forces and couples.
116 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

-—

EXAMPLE 3.12

Determine the resultant of the forces and the couple acting on the
plate shown.

Solution—Scalar Approach

Coordinate system. The origin of the x-y coordinate system is


located at point A, which we will use as the reference point
for the summation of moments. Point A is selected because the
lines of action of the 2.5-kN and 2-kN forces pass through it;
thus, the moments of these two forces about A will be zero.

Rectangular components of forces. The components of the


four forces in the x and y directions are

Fa =0 F,, =3 kN
F,,°=(2:5)(cos 184°) =237 KN
F, = (2.5)(sin 18.4°) = 0.79 kN
F3,=2KN F;3,=0
F 4, = (1.5)(cos 60°) = 0.75 kN
Fy, = (1.5)(sin 60°) = 1.30kN
Resultant force. The x and y components are

R, = SF, II= 2.37 — 2 + 0.75 = 1.12 kN


R, = 3F, II= —3 + 0.79 — 1.3 = —3.51 kN

(continued)
3.12 Resultants of Nonconcurrent Coplanar Force Systems

EXAMPLE 3.12 (continued)


SS

y The magnitude of the resultant force is


R= V(R,)? + (R,)° = V(1.12)? + (= 3.51)? = 3.68 KN
and its direction with respect to the x axis is given by

MM. - 3.85 Nn

R,=3.51kN
R=3.68kN

Resultant force-couple system at point A. Since the lines of


action of forces F, = 2.5 kN and F; = 2 KN pass through
point A, their moments about A are zero. The moment arms of
F, and F,,, are the horizontal distances of 0.75 m and 0.5 m,
respectively. The moment arm of F,, is the vertical distance DE
shown, which is determined from the similar triangles ABC
and DEC by writing
DE _ AB
DEMEAG
Noting that AB = 0.5 m, AC = 1.5 m, and DC = 1 m, we
obtain
8) Siloear IEP Sifos) 1)(0.5
; d =e ie
(@eS)

Considering the counterclockwise moments as positive, we


determine the moment of the resultant couple as

Mp, = —3(0.75) + 0.75(0.33) — (1.3)(0.5) — 1.2


= —3.85 kN-m

Location of the single resultant force. The equation defining


the new line of action of the resultant force is

Mra = xR, = yR,.

(continued)
EXAMPLE 3.12 (concluded)

Substituting numerical values, we obtain example, by setting y = 0, we obtain the x intercept of the line,
or the point of application of the resultant force on the x axis;
= Sifels) = 28 —=Skt) = CL)
that is,
or
fy=0 iS SS = Ilhan
3.51x + 1.12y = 3.85
ee 351
The line of action of R and its point of application on the x axis
which defines the line of action of the single resultant force.
are shown. Thus, the resultant force is
Any point on the line defined by this equation can be considered
as the point of application of the single resultant force. For R = 3.68 kN % 72.3° at a distance of 1.1m fromA <

EXAMPLE 3.13

Determine the resultant of the three forces acting on the plate


ry shown.
2 ft
fo 4ft Solution—Vector Approach
60° Rectangular components of forces. The three forces can be
866 Ib 1500 1b
3 ft 30° rc e written in vector form as
t t
aa |= —750i — 433] (Ib)
866 lb F, = —750i + 433] (1b)
F, = 1500i (1b)
Peper) Ohi:
Resultant force. We may then write
By
R = SF = (— 750i — 433j)lb + (— 750i + 433j)lb + (1500i)Ib
R=0 <

Resultant couple. Even though the resultant force of the given


F,,= 7501 system of forces is zero, its resultant couple may not necessarily
Di
be zero. The position vectors from the origin A of the x-y
F, = 866lb #F— coordinate system to the lines of action of the three forces are
. F,=15001b
30 Base r, =0 r, = 2i — 3j (ft) r, = 4i — 2j (ft)
The moment of the resultant couple is given by

Mz, = 2M = X(r X F)
= (2i — 3j) X (— 750i + 433j) + (41 — 2j) x (15001)

5 = 866k — 2250k + 3000k


Mr, = 1616k(Ib-ft) <
Thus, the resultant of the given system of three forces is a
couple of moment 1616 lb - ft GC)as shown.

Mp =1616 lb-ft

118
3.14 Equilibrium Equations in Scalar Form 119

EQUILIBRIUM IN TWO DIMENSIONS

3.13 EQUILIBRIUM OF RIGID BODIES


A rigid body is in equilibrium if, at rest initially, it remains at rest when
subjected to a system of forces and couples. In order for a body to be in
equilibrium, the resultant force-couple system of the forces and couples acting
on the body must be zero:

R= XF=0
G21)
M, = >M=0

These equations, called the equations of equilibrium for rigid bodies, repre-
sent the necessary and sufficient conditions for equilibrium. They indicate that
the forces and couples acting on a body in equilibrium neutralize or cancel the
effects of each other and, therefore, have no net external effect on the body.
Such a system of forces, therefore, does not disturb (or accelerate) the initial
state of rest of the body, which remains at rest under the action of the forces.
If a rigid body initially in motion is subjected to a force system that satisfies
the equations of equilibrium, it will remain in motion with a constant velocity
because the force system cannot accelerate it. Such a body is also considered
to be in equilibrium. However, the term equilibrium is commonly used to refer
to the state of rest of bodies, and will be used in this context in the present
volume.

3.14 EQUILIBRIUM EQUATIONS IN SCALAR FORM


Since in this chapter we are focusing our attention on the equilibrium of two-
dimensional rigid bodies subjected to coplanar forces, the equations of equi-
librium can be expressed in more convenient scalar form in terms of the rec-
tangular components of forces. Rewriting the first of the two conditions given
in Equations (3.21) in Cartesian vector form as

R=2F =2(,i+ Bj = 2Fi+ 2Fj=0


we can see that in order for R to be equal to zero, the two summations >F,
and &F, must both be equal to zero, or }F, = 0 and {F, = 0.
As for the second condition given in Equations (3.21), we recognize that
the moment vectors of all the couples and of all the forces about any point in
the plane of the body, as well as the moment vector of the resultant couple,
will be directed perpendicular to the plane of the body. Therefore, the moment
equilibrium condition can be written as
M,k = =Mk = 0
from which we obtain the scalar equation {M = 0.
Thus, the necessary and sufficient conditions for the equilibrium of two-
dimensional rigid bodies subjected to coplanar forces are given by the three
120 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

scalar equations of equilibrium:

LF, =0 2F, =0 2M =0 G2z

The first two conditions indicate that the algebraic sums of the x and y com-
ponents of all the forces acting on the body in equilibrium must be zero. The
third condition states that the algebraic sum of the moments of all the forces
about any point in the plane of the body plus the moments of any couples
acting on the body must be zero. All three equilibrium conditions must be
satisfied in order for the body to be in equilibrium.

Alternate Forms of Equilibrium Equations


The equations of equilibrium as expressed in Equations (3.22) provide the
most convenient means of analyzing many problems involving equilibrium of
two-dimensional bodies. However, in some cases, the analysis may be sim-
plified by using one of two alternate forms of these equations.
One alternate system of equilibrium equations is

2F,=0 >~M, =0 Mz = 0" (3.23)

provided that the line connecting points A and B is not perpendicular to the
p axis, which may be chosen in any arbitrary direction. Equations (3.23) are

(c) (d) (e)


Figure 3.29
3.14 Equilibrium Equations in Scalar Form 121

equations of equilibrium because they ensure that the resultant force-couple


system of the forces and couples acting on the body is zero. This can be shown
by considering a rigid body subjected to an arbitrary system of forces and
couples as shown in Figure 3.29a. This system can be reduced to a resultant
force-couple system at a point A as shown in Figure 3.29b. The first equilib-
rium equation, 2F,, = 0, requires that the component of R in the direction of
the p axis must be zero, where the p axis may be in any arbitrary direction.
This condition implies that if the resultant R is not zero, then it must be
perpendicular to the p axis as shown in Figure 3.29c. Next, if the second
equilibrium equation 2M, = 0 is also satisfied, then the resultant couple
Mra, = 9, and we are left with the resultant force R at A with its line of action
perpendicular to the p axis as shown in Figure 3.29d. As for the third equilib-
rium equation 2M, = 0, we can see from Figure 3.29d that as long as point
B is not on the line of action of R (i.e., line AB is not perpendicular to the
p axis), the equation 2M, = 0 can be satisfied only if R = 0. Thus, Equations
(3.23), when satisfied simultaneously, ensure that the body is in equilibrium
(Figure 3.29e).
The other alternate set of equilibrium equations is

2M,=0 3M,=0 %M.=0 (3.24)

provided that the points A, B, and C are not on the same straight line. Like the
preceding two sets of equilibrium equations, Equations (3.24), if satisfied,
ensure that the resultant force-couple system of the forces and couples acting
on the body is zero. For example, consider a rigid body subjected to an arbi-
trary system of forces and couples as shown in Figure 3.30a. This system is
reduced to the resultant force-couple system at point A shown in Figure 3.305.
The first equilibrium equation }M, = 0, if satisfied, means that the resultant
couple Mr, = 0 (Figure 3.30c). The second equation, }M, = 0, if satisfied,
implies that either R = 0 and hence the body is in equilibrium, or point B is
on the line of action of R as in Figure 3.30c. Since point C may not be collinear
with A and B, it cannot be on the line of action of R if B is on it. Therefore,
if {Mc = 0 (ie., all three equations are satisfied) we are assured that R = 0
and hence the body is in equilibrium.
It is important to understand that each of the three sets of equilibrium equa- (d)
tions given by Equations (3.22), (3.23), and (3.24) contains three independent
Figure 3.30
equations. All three equations in a given set must be satisfied to ensure that
the body is in equilibrium. While any one of the three sets may be used in
analyzing equilibrium of rigid bodies, equations from different sets should
not be mixed nor should they be combined to obtain additional equations of
equilibrium.
122 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

EXAMPLE 3.14

A plate is subjected to four forces and a couple as shown. Is the


plate in equilibrium?

io
4 Im
3

2m
Solution To determine whether the plate is in equilibrium, we
15 kN:m
evaluate the three equations of equilibrium given in Equations
(3.22) as follows:

@zF, =-9+9=0
8 kN @sF, =8-20+12=0
@=M, = —8(@) + 9(1) + 20(1.5) — 15 = 0
20 kN Note that the moment center was chosen at A to eliminate the
moment of the 15-kN force.
Since the four forces and the couple acting on the plate
satisfy the three equations of equilibrium, the plate is in
equilibrium.

First Alternate Solution Applying the alternate set of


equations given in Equations (3.23), we write

@sr, =-9+9=0
@=M, = —8(3) + 9(1) + 20(1.5) — 15 =0
@>M, = —20(1.5) + 9(1) + 12) — 15 =0
As seen from the figure, the points A and B are chosen so that
the line AB is not perpendicular to the x axis.
Since the three equilibrium equations are satisfied, the plate
is in equilibrium under the action of the given forces and the
couple.

Second Alternate Solution Using the alternate system given


in Equations (3.24), we obtain

@2=M, =
lI —8) + 9(1) + 20(1.5) — 15 =0

@=M, = —20(1.5) + 9(1) + 123) — 15 =0


@2M- = 9) — 20(1.5) — 9(2) + 12) — 15 =0
As shown in the figure, the three moment centers A, B, and C
are chosen so that they do not lie on the same straight line.
Since the three equilibrium equations are satisfied, the plate is
in equilibrium.
3.15 Equilibrium of Rigid Bodies Subjected to Forces at Two and Three Points 123

3.15 EQUILIBRIUM OF RIGID BODIES SUBJECTED


TO FORCES AT TWO AND THREE POINTS
In many engineering applications, we encounter structures that are in equilib-
rium under the action of forces acting at only two or three points. Analysis
can be considerably expedited by recognizing the following characteristics of
such systems.

When only two forces act on a rigid body in equilibrium, they must be STATEMENT
equal, opposite, and collinear.

Such a body is commonly referred to as a two-force member. Figure 3.31


shows a rigid body subjected to two forces acting at points A and B. In order
for the resultant force of the two forces to be zero, the forces must be equal
in magnitude but with opposite senses. For their resultant couple to also be
equal to zero, the two forces must be collinear; that is, they must have the
same line of action.
When several forces act at two points on a body, the concurrent forces
acting at the two points may be combined into their respective resultant forces.
The two resultant forces thus obtained must be equal, opposite, and collinear
if the body is in equilibrium.
Figure 3.31

When three forces act on a rigid body in equilibrium, they must be either STATEMENT
concurrent or parallel.

Such a body is commonly referred to as a three-force member. A rigid body


subjected to three forces acting at points A, B, and C is shown in Figure 3.32.
We can see that if any two of three forces are concurrent at point D, and the
third force is not, then the nonconcurrent force will cause a resultant moment
about the point of concurrency D of the other two, thereby violating the state
of equilibrium of the body. If any two of the three forces are parallel, they
may be considered to be concurrent at infinity. The line of action of the third
force must intersect this point of concurrency of the other two forces at infinity:
it must be parallel to the lines of action of the other two forces. Therefore, if
Figure 3.32
a rigid body is in equilibrium under the action of three forces, the three forces
must be either concurrent or parallel.
When more than one force acts at three points on a body, the concurrent
forces acting at the three points may be combined into their respective resultant
forces. The three resultant forces thus obtained must be either concurrent or
parallel if the body is in equilibrium.
124 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

EXAMPLE 3.15

A 500-Ib signboard is supported in the horizontal position by two


cables. Determine the tension in each cable and the angle of in-
clination of the cable on the right-hand side.

Solution Since the board is in equilibrium under the action of


three forces that are not parallel, the forces must be concurrent.
The location of the point of concurrency D can be determined
by using lines of action of the 500-Ib force and cable tension T,
which are specified in the problem. From the right triangle
ADE shown, we obtain

DE = AE (tan 42°) = 4(0.9) = 3.6 ft


The direction of the cable tension 7, can be determined from
the right triangle BDE as

an-1(28)
tan”
3.6
*| —) = 60.9°

a= 60.9° <

The magnitudes of the cable tensions 7, and T, are


determined by using the equations of equilibrium:

@ SF, = -T,(cos 42°) + T>(cos 60.9°) = 0


from which we obtain the relationship between 7, and T, as
cos 42°
1b = 1) SS ]] = WS
‘ (2 | ’ (a)
Using the second equilibrium equation
OS, = T,(sin 42°) + T,(sin 60.9°) — 500 = 0
from which we obtain
0.677, + 0.87T, = 500
Substituting Equation (a) into Equation (b), we have

0.67T, + 0.87(1.53T,) = 500


2.01T, =500 T, = 249 lb <
Substituting the magnitude of T, into Equation (a), we obtain
T, = 381 lb <
Note that since the three forces are concurrent, the third
equilibrium equation {M, = 0 is automatically satisfied.
3.16 Types of Supports and Connections 125

3.16 TYPES OF SUPPORTS AND CONNECTIONS


Supports and connections are the devices used to restrict movement of one
rigid body with respect to another. In order to maintain bodies in equilibrium,
we prevent their movements due to the action of applied forces, by attaching
them to, or by supporting them on, other bodies or the ground. While applied
forces tend to move (translate and/or rotate) the bodies, supports prevent the
movements by exerting opposing or reactive forces on the bodies to neutralize
the effects of the applied forces, thereby keeping the bodies in equilibrium.
These reactive forces and couples exerted by the supports or connections are
called reactions. The type of reaction that may develop at a support point or
a joint depends on the type of supporting device or connection used at that
point. If a support prevents translation in a certain direction, then a reaction
force is exerted on the body in that direction. Similarly, if a support prevents
rotation of a body about a certain axis, then it must exert on the body a reaction
couple with a line of action in the direction of that axis. The magnitudes and
senses of the reactions depend on the system of forces and couples applied to
the body and are determined by solving the equations of equilibrium of the
rigid body. Such determination of reactions constitutes an essential first step
in the analysis and design of all engineering structures and machines.
Some common types of supporting devices and connections used for two-
dimensional bodies subjected to coplanar forces are shown in Table 3.1. These
supports and connections are classified into three categories: I, I, and III,
depending on whether they exert one, two, or three reactions on the body.

Category! Each of the supports shown in this category can prevent translation
in only one direction—each exerts one reaction force on the body in the di-
rection in which the translation is prevented. Since the direction of the reaction
force is known, the only unknown is its magnitude, which is determined by
applying the equations of equilibrium of the rigid body. The first three types
of supports—rockers, rollers, and frictionless or smooth guides—exert a re-
action force that is directed normal to the supporting surface, whereas the link
support exerts a reaction force in the direction of its axis. These supports are
considered to be reversible in the sense that the reaction force may act either
into or away from the body.
The two types of supports listed in Category Ia are not reversible. The cable
support can prevent translation of the body in its direction provided the sense
of translation is such that it causes the cable to be taut or in tension. It cannot
prevent translation and, therefore, cannot exert a reaction force on the body
when the sense of translation causes the cable to become slack. Hence, the
reaction force exerted by the cable is in the direction of the cable and must
act away from (not into) the body. Similarly, a smooth or frictionless surface
can prevent translation of the body only when the sense of translation is di-
rected into the surface, but not when the sense of translation is directed away
Table 3.1 Supports and connections for bodies subjected to coplanar forces

Category Types of supports or connections Types of reactions Number of unknowns



One

The reaction force is directed normal to the

= {
supporting surface and may act either into
or away from the body. The magnitude of
the reaction force is the unknown.
Rollers F

4
9 One
0 The reaction force is in the link’s direction
and may act either into or away from the
F body. The magnitude of the reaction force
Link is the unknown.

P F One
rau The reaction force acts away from the body in
la k the direction of the cable. The magnitude of
the reaction force is the unknown.
Cable
Sa

One
The reaction force acts away from the surface
Smooth surface (into the body) in the direction normal
; to the surface. The magnitude of the reaction
lr force is the unknown.

Two
The reaction is a force of unknown magnitude
and direction. The reaction force is usually
Smooth represented by its rectangular components.
hinged or The magnitudes of the two components are
Smooth hinged or pinned the two unknowns.
pinned support connection Fy
Two
The reaction is a force of unknown magnitude
and direction, which acts away from the surface
Ila (into the body). The reaction force is usually
represented by its rectangular components.
Rough surface The magnitudes of the two components are
the two unknowns.
Three
The reactions consist of a force of unknown
magnitude and direction, and a couple of
unknown magnitude. The reaction force is
usually represented by its rectangular
components. The magnitudes of the two force
Rigid components and the magnitude of the moment
connection
of the reaction couple are the three unknowns.

126
3.17 Statically Determinate Structures

from it. Therefore, it can only exert a reaction force normal to the surface,
with the sense away from the surface and into the body.

Category || The hinged support shown in this category can prevent translation
in any direction but cannot prevent rotation; thus, it can exert a reaction force
on the body whose magnitude and direction are both unknown. However, it is
usually convenient to represent the reaction force by its rectangular compo-
nents, with the magnitudes of the two components as unknowns, in the equi-
librium equations of the rigid body.
Hinged connections, formed by connecting two or more members or com-
ponents by a frictionless hinge or pin, are used in many engineering structures.
The translation of each member at such a joint is the same (i.e., the relative
translation between the members is prevented). The reaction force exerted by
a hinged connection on a member is of the same type as a hinged support: a
force of unknown magnitude and direction, or two rectangular force compo-
nents of unknown magnitudes. Analysis of hinge or pin-connected structures
will be presented in Chapter 6.
The rough surface shown in Category Ila can exert a reaction force of
unknown magnitude and direction on the body only when the sense of the
translation is directed into the surface. Therefore, the sense of the reaction
force must always be directed away from the surface or into the body. When
the reaction force is represented by its rectangular components, the sense of
the component that is perpendicular to the surface must be directed away from
the surface.

Category lll The fixed support shown in the table can prevent translation in
any direction, as well as rotation—it can exert a reaction force of unknown
magnitude and direction, and a reaction couple of unknown magnitude. The
reaction force is usually represented by its rectangular components; the three
unknowns (the magnitudes of the two force components and the moment of
the couple) are determined from the equations of equilibrium of the rigid body.
The rigid connection prevents any relative translation and rotation between the
members or components at the joint, thus the translation and rotation of each
member at the joint is the same. The reactions exerted by a rigid connection
on a member are of the same type as in the case of a fixed support.

3.17 STATICALLY DETERMINATE STRUCTURES


In the preceding section, we discussed the reactions that can be exerted by the
various types of supports and connections on a body. In this section, we will
examine the number and arrangement of reactions necessary for the equilib-
rium of a rigid body subjected to coplanar forces.
A body is considered to be statically determinate externally if all of its
reactions can be determined by solving the equations of equilibrium. Since
128 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

there are three equations of equilibrium for two-dimensional bodies subjected


to coplanar forces, they cannot be used to determine more than three reactions.
Also, in order to satisfy the three equilibrium equations to ensure that the body
remains in equilibrium under the action of a general coplanar force system,
the body must be supported by at least three reactions.
Consider, for example, the rectangular plate subjected to four forces as
shown in Figure 3.33a. The plate is supported by a hinged support at point A
and a roller support at point B. The free-body diagram of the plate in Figure
3.33b shows the plate removed or freed from the supports. Note that the free-
body diagram shows, in addition to the four known applied forces, the un-
known reactions being exerted on the plate by the two supports at points A
and B. Recall from the preceding section that a hinged support is capable of

‘. m—

800 N

800N _

2000 N

166.2N 3266.2N

(c)
Figure 3.33
3.17 Statically Determinate Structures 129

exerting two reaction force components on the body, and that a roller support
can exert only one reaction force which must be in the direction perpendicular
to the supporting surface. On the free-body diagram the two reaction force
components exerted by the hinged support at A are denoted by A, and A, in
the x and y directions, respectively. Similarly, the one reaction force, exerted
in the vertical direction by the roller support, is denoted by B,. The correct
senses of the three reaction forces are not known at this stage in the analysis
and are arbitrarily chosen. After we have determined the magnitudes of these
forces by solving the equilibrium equations, a positive magnitude for a reaction
will indicate that the assumed sense was correct, whereas a negative value will
imply a sense opposite to the one assumed on the free-body diagram. From
Figure 3.33b, we can see that without the three reactions, the plate cannot
remain in equilibrium under just the action of the applied forces. It is essential
that the free-body diagram show all the external (applied and reaction) forces
acting on the rigid body. A step-by-step procedure for constructing free-body
diagrams is presented in the following section. The magnitudes of the three
unknown reaction components (A,, Ay and By) can now be determined by
solving the three equations of equilibrium developed in Section 3.14.
We will determine the reactions of the plate by applying the equilibrium
equations given in Equations (3.22). To expedite the analysis, the equations
will be applied in such a manner that each contains only one unknown reaction,
thereby avoiding the need for solving simultaneous equations. Figure 3.33b
shows that two of the three unknown reactions, A,, and B,, are parallel to the
y axis. Therefore, the equilibrium equation {F, = 0 will contain only the third
unknown, A,, which can be determined from

@ Fr, = 0 800 + 2000 (cos 30°) + A, = 0


A, = —2532 N
x

The negative answer indicates that the sense of A, is opposite to the one as-
sumed on the free-body diagram. Therefore,

Ay= 2532 NO)


Since the two remaining unknowns, As and BY are acting at points A and B,
respectively, choosing either of the two points as the moment center will elimi-
nate one of the unknowns from the moment equilibrium equation. Using A as
the moment center, we write the moment equilibrium equation as

@©3IM,=0 —800(1) — 500(0.4) — 1600(0.9)


— 2000(cos 30°)(0.6) — 2000(sin 30°)(2.4) + B,(1.8) = 0
B, = 3266.2 N @)
Applying the remaining equilibrium equation, ie = 0, we determine the
third unknown, A,. Thus,

®ZF,=0 ~—500 — 1600 — 2000(sin 30°) + 3266.2 + A, = 0


A,=-166.2N
y
or A,=166.2NQ)
130 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

The magnitudes and the senses of the three reaction forces are shown in Fig-
ure; 3.336:
As all three of the reactions of the plate could be determined using the three
equilibrium equations, the plate is statically determinate externally. Some ad-
ditional examples of externally statically determinate bodies are shown in Fig-
ure 3.34. Note that each of the bodies is supported by three reactions that can
be determined by solving the three equations of equilibrium.

Static Indeterminacy and Instability


If a rigid body has more than three external reactions, all of the reactions cannot
be determined using the three equations of equilibrium. Such a body is con-
sidered statically indeterminate externally. Statically indeterminate bodies
have more reactions than necessary to maintain a body in equilibrium under a
general system of coplanar forces. Even though one or possibly two of the
reactions of such bodies can be determined by using the equilibrium equations,
these equations are not sufficient for complete determination of the reactions.
The excess reactions are called redundants, and the number of redundants is
referred to as the degree of indeterminacy. Thus, if a rigid body is supported
by r number of reactions (r > 3), then the degree of indeterminacy is given by
[=> (3.23)

Some examples of statically indeterminate bodies are shown in Figure 3.35.


Figure 3.34 Examples of externally statically
The analysis of statically indeterminate bodies belongs in the study of me-
determinate bodies
chanics of deformable bodies and will not be considered herein.
If a body has less than three external support reactions, the reactions are
not sufficient to prevent all possible movements of the body in its plane. Such
a body cannot remain in equilibrium under a general system of coplanar forces
and, therefore, is considered statically unstable externally. Statically unstable
bodies are also referred to as partially constrained bodies.
Consider, as an example, a beam supported on two rollers as shown in
Figure 3.36a. It is obvious from this figure that the two roller supports cannot

1=6—3=3
(b) (c)
Figure 3.35 Examples of externally statically indeterminate bodies
3.17 Statically Determinate Structures 131

po
Vv P

Lise ae
=
Figure 3.36 Partially constrained body not in equilibrium

prevent translation of the beam in the horizontal direction, but can prevent
translation in the vertical direction and its rotation. The beam is therefore only
partially constrained or supported and is considered statically unstable.
The free-body diagram of the beam is shown in Figure 3.36b. Applying the
three equations of equilibrium given by Equations (3.23), we have

@EmM,=0 -(Peosayay+BiL)=0 B= “SA*@


@©3M,=0 —A,(
+(Pcos
L) a)b =0 ee Te
Using the third equilibrium equation we obtain

@ SF, =P sina #0
Since neither P nor a is zero, the product P sin a # 0, and therefore, the
equilibrium equation }F,, = 0 is not satisfied. This indicates that the beam is
not in equilibrium in the x direction.
Even though statically unstable or partially constrained bodies cannot re-
main in equilibrium under any arbitrary system of coplanar forces and couples,
such bodies can be in equilibrium under certain systems of applied forces and
couples. These applied forces and couples must satisfy by themselves the equi-
librium equations that cannot be satisfied due to lack of support reactions. For
example, the beam considered in the preceding paragraphs is in equilibrium if
the applied load P acts in the vertical direction (a = 0°) as shown in Figure
3.37a. The magnitudes of the two reactions in this case can be determined by
132 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

(b)
Figure 3.37 Partially constrained body in equilibrium

substituting @ = 0° in the expressions for A, and B, developed previously, or


by applying the equilibrium equations }M, = 0 and {F, = 0. Thus,
Pa Pb
Beir d A,=—
ary - ey,
The third equilibrium equation }F, = 0 is now satisfied because sin 0° = 0;
that is,
@) SF, = P(sin «) = P(sin 0°) = 0
Thus, the partially constrained beam is in equilibrium under the vertical load
P. Note that this beam will be in equilibrium under any applied system of
coplanar forces and couples as long as all the forces contained in the applied
system satisfy the equation }F, = 0, which would otherwise not be satisfied
due to lack of reaction in the x direction.
The conditions of static instability, determinacy, and indeterminacy of two-
dimensional rigid bodies, supported by r number of reactions, are commonly
written in the following form:

r<3_ bodyis statically unstable externally


r=3_ bodyis statically determinate externally (3.26)
r> 3 bodyis statically indeterminate externally

Geometric Instability The first condition given above (r < 3) for static insta-
bility of bodies is both necessary and sufficient in the sense that if r < 3, the
3.17 Statically Determinate Structures 133

body is definitely unstable or partially constrained. However, the remaining


two conditions for static determinacy (r = 3) and indeterminacy (r > 3) are
necessary, but not sufficient conditions. In other words, these two conditions
simply tell us that the number of reactions is sufficient for stability; they do
not provide any information regarding their arrangement. A body may be
supported by a sufficient number of reactions, but may still be unstable (i.e.,
not completely restrained or supported to prevent any possible movement) due
to improper arrangement of supports. Such bodies are considered to be geo-
metrically unstable externally, or improperly constrained.
The two types of improper arrangements of support reactions that cause
geometric instability are shown in Figure 3.38. The beam in Figure 3.38a is
supported by three parallel cables. It can be seen from Figure 3.38a that even
though there are a sufficient number of reactions, 7,= 3, since all of them are
in the vertical direction (parallel to each other), they cannot prevent translation
of the beam in the horizontal (perpendicular) direction. The beam is therefore
geometrically unstable or improperly constrained.
The other type of improper arrangement of reactions that causes geometric
instability is shown in Figure 3.385. In this case, the beam is supported by a
hinged support at A and a roller at B; thus, there are a sufficient number of
reactions, r = 3. However, it can be seen from this figure that since the lines
of action of the three reaction forces are concurrent at support point A, they
cannot prevent rotation of the beam about that point. In other words, the mo-
ment equilibrium equation {M, = 0 will not be satisfied for a general coplanar
system of forces and couples applied to the beam; thus, the beam is geomet-
rically unstable or improperly constrained. (a)
From the foregoing discussion, we can conclude that in order for a two-
dimensional rigid body to remain in equilibrium under the action of any ar-
bitrary applied coplanar system of forces and couples, it must be restrained or
supported by at least three reactions, all of which must be neither parallel nor
concurrent.
Like partially constrained (statically unstable) bodies, improperly con-
strained (geometrically unstable) bodies can be in equilibrium under certain
systems of applied forces. However, unlike partially constrained bodies, the
reactions of improperly constrained bodies in equilibrium cannot be deter-
mined by solving the equations of equilibrium. For example, the beam of
Figure 3.38a is in equilibrium if the load P is applied in the vertical direction
(a = O°). But there are three unknown vertical reactions A,, B,, and C,, that (b)
cannot be determined by solving two equations of equilibrium, SF, = 0 and Figure 3.38 Improperly constrained bodies
SM = 0. (The third equilibrium equation, SF, = 0, is satisfied automatically
because, for a = 0°, there is no force acting in the x direction.)
Engineering structures such as buildings and bridges, which are expected
to be stationary (in equilibrium) under any arbitrary loading condition, must
always be properly constrained. Partial and improper constraints, however, are
commonly used for mechanical systems that are expected to move in certain
directions under particular loading conditions. A truck supported on wheels is
an example of such an improperly constrained body.
134 Chapter 3 Equilibrium of Rigid Bodies in Two Dimensions

In the remainder of this chapter, we will consider the analysis of reactions


of properly constrained statically determinate bodies, and partially constrained
bodies in equilibrium under given loading conditions. For these two types of
structures, all the support reactions can be determined by solving the equations
of equilibrium.

3.18 FREE—BODY DIAGRAMS


The free-body diagram of a rigid body is a diagram of the body isolated or
freed from its supports, and showing all the external (applied as well as re-
action) forces acting on it. The construction of the free-body diagram is per-
haps the most crucial step in the analysis of the equilibrium of rigid bodies.
The following step-by-step procedure may be used to draw free-body diagrams
of rigid bodies in equilibrium.

1. Identify the body whose free-body diagram is to be drawn.


2. Visually detach the body from its support and disconnect it from any other
bodies to which it may be connected. Draw the outline of this isolated body.
3. Show each known (applied) external force or couple on the diagram using
an arrow to indicate its direction and sense. Write the magnitude of each force
or couple by its arrow.
4. Show the weight of the body (if known) at its center of gravity by a
downward arrow. The center of gravity of a symmetric homogenous body is
located at its center of symmetry or the midpoint. (The procedure for locating
the centers of gravity of nonsymmetric bodies will be discussed in Chapter 5.)
5. Select the orientation of the x-y coordinate system. The x and y axes, which
must be perpendicular to each other, are usually oriented in the horizontal
(positive to the right) and vertical (positive upward) directions, respectively.
However, if the dimensions of the body and/or the lines of action of all
or most of the applied forces are in an inclined direction, selection of the x
(or y) axis in that direction may simplify the analysis.
6. At each point where the body has been disconnected from a support or a
connection, show the unknown reaction forces being exerted on the body. The
type and number of reaction forces that can be exerted by the various types of
supports and connections were given in Table 3.1. Each reaction force is rep-
resented on the free-body diagram by an arrow in the direction of its line of
action, which is known. Each reaction couple is shown by a curved arrow.
The senses of the reactions exerted by cable supports and smooth and rough
surfaces are known and should be correctly indicated. The senses of the re-
actions exerted by other types of supports are not known and may be arbitrarily
assumed. The senses of reaction force components can be conveniently as-
sumed to be in the positive x and y directions, and the senses of reaction
couples are usually assumed to be counterclockwise. Show the assumed senses
of reactions on the free-body diagram. The correct senses of the reactions will
be known only after we have determined their magnitudes by solving the
3.18 Free-Body Diagrams 135

equilibrium equations. A positive magnitude for a reaction will indicate that


the assumed sense was correct, whereas a negative magnitude will imply a
sense Opposite to the one assumed on the free-body diagram.
Since the magnitudes of the reactions are not known at this stage of the
analysis, the magnitude of each reaction should be denoted by a letter symbol
on the free-body diagram.
7. Finally, draw the dimensions of the body indicating the locations of all
the known and unknown external forces.

EXAMPLE 3.16

Draw the free-body diagram of the uniform cantilever beam 10 kip 5 kip
shown. The beam weighs 150 lb/ft. Can all the reactions be de-
termined from the equations of equilibrium? Explain. |

Solution |

Free-body diagram. The free-body diagram of the beam 18 ft AL 6 ft :


shows the beam removed from the fixed support at end A, and
shows all the external loads and support reactions acting on the
beam. Note that in addition to the two known forces of 10 kip
and 5 kip magnitudes, the diagram also shows the total
weight of the beam, which is calculated by multiplying the 10 kip 5 kip
given weight per unit length by the total length of the beam, or

W = (150 lb/ft)(24 ft) = 3600 Ib = 3.6 kip


The magnitude of the weight is expressed on the free-body
diagram in the same unit (kip) as used for the other two known
forces. Because the beam is uniform (symmetric), its center of
gravity G is located at the midpoint (at 12 ft from either end of
the beam).
Since a fixed support can exert two reaction force
components and one reaction couple on a body, the free-body
diagram shows the two unknown force components, A,. and A,,
and an unknown couple M, at end A. These three reactions
are being exerted on the beam by the wall or the fixed support
at A. The correct senses of the three reactions are not known.
The two reaction forces A, and A, are assumed to be in the
positive x and y directions, respectively, and the reaction couple
M,, is assumed to be counterclockwise as shown on the free-
body diagram.

Static determinacy. The beam is properly constrained by the


three reactions A,, A,, and M,. All three unknown reactions
can be determined by solving the three equations of equilibrium;
thus, the beam is statically determinate.
136 Chapter 3 Equilibrium of Rigid Bodies in Two Dimensions

EXAMPLE 3.17

Draw the free-body diagram of the bracket shown. Can all the
reactions be determined from the equations of equilibrium?
Explain.

support
]
Smooth
guide

Solution

Free-body diagram. ‘The free-body diagram of the bracket


shows all the external known and unknown forces acting on the
bracket. The pinned support at A exerts two reaction force
components on the bracket, which are denoted by A, and A, on
the free-body diagram, with their senses assumed to be in the
200 mm . _ positive x and y directions, respectively. The guide B prevents
translation in the direction perpendicular to itself, thereby
exerting an unknown reaction force F in that direction,
as shown on the free-body diagram. The sense of this reaction
is arbitrarily assumed.

Static determinacy. The bracket is supported by three


reactions: A,, A,, and F. The number of reactions is sufficient
to prevent any possible translation or rotation of the bracket.
However, the line of action of F passes through point A,
thereby making the three reactions concurrent at A. Therefore,
the reactions cannot prevent the rotation of the bracket about
point A. This can be verified by writing the moment equilibrium
equation,

© XM, = (—4 KN)(346.4 mm) = — 1385.6 kN: mm #0


Since this equilibrium equation cannot be satisfied, the bracket
is not in equilibrium. Therefore, the reactions cannot be
determined from the equations of equilibrium.
3.18 Free-Body Diagrams 137

EXAMPLE 3.18

Draw the free-body diagram of the uniform ladder shown. The


ladder weighs 150 Ib, and the weight of the painter is 200 Ib. Can Smooth
all the reactions be determined from the equations of equilibrium? wall
Explain.

\
\
\\
Solution ns Rough surface

Free-body diagram. The free-body diagram of the ladder


shows all the external forces acting on the ladder. Note that the
weight of the ladder is acting at its center of gravity G, which
is located at its midpoint, 7.5 ft from either end of the ladder.
The smooth vertical wall exerts a horizontal reaction A,
at end A of the ladder. The sense of this reaction is known, and
is directed away from the wall or to the right, as shown on the
diagram. The rough horizontal surface exerts two reaction force
components B, and B,, at end B of the ladder. The sense of
the vertical reaction B, is known and is directed upward (away
from the surface) as shown. The sense of the horizontal
reaction B, is not known and is assumed to be in the positive x eG
direction.

Static determinacy. Since the three unknown reaction forces


are neither concurrent nor parallel, they can be determined
by solving the three equations of equilibrium.
138 Chapter 3 Equilibrium of Rigid Bodies in Two Dimensions

3.19 PROCEDURE FOR ANALYSIS OF REACTIONS


The following step-by-step procedure can be used to determine reactions of
two-dimensional rigid bodies subjected to coplanar systems of forces and
couples.

1. Draw a free-body diagram of the body using the procedure described in


the preceding section.
2. Check the body for static determinacy. If the number of unknown reactions
is greater than three (7 > 3), the body is statically indeterminate; end the
analysis at this stage. If the number of unknown reactions is equal to three
(r = 3), but all of the three reaction forces are either concurrent at a single
point or are parallel, the body is improperly constrained; end the analysis at
this stage. Otherwise, the body is properly constrained and statically deter-
minate if r = 3, or partially constrained if r < 3; go to the next step.
3. Determine unknown reactions by solving the equilibrium equations
(LF, = 0, 2F, = 0, and XM = 0). Apply the equilibrium equations so that
each may involve only one unknown reaction, thereby avoiding the need for
solving simultaneous equations. The positive answer for the magnitude of a
reaction indicates that its sense as assumed on the free-body diagram is correct,
whereas a negative answer means that the correct sense of the reaction is
opposite to that assumed on the diagram.
4. If the body is properly constrained (i.e., 7 = 3), go to Step 5. Otherwise,
the body is*partially constrained by two or less reactions (r < 3). Use the
remaining equilibrium equations, which were not utilized in Step 3 in deter-
mining the reactions, to check the body for equilibrium. If all of the remaining
equations of equilibrium are satisfied, then the partially constrained body is in
equilibrium under the given loading condition; go to the next step. Otherwise,
the body is not in equilibrium; end the analysis at this stage.
5. Apply an equilibrium equation that was not used before to check the com-
putations. This alternate equilibrium equation should preferably involve all the
reactions that were calculated in the analysis. Usually, a moment equilibrium
equation is used for this purpose with the summation of moments taken about
a point that does not lie on the lines of action of the reaction forces. If the
analysis has been carried out correctly, this alternate equation will be satisfied.
EXAMPLE 3.19

Determine the reactions at the supports for the simply supported a4 ao 1000 N
beam shown. Neglect the weight and depth of the beam in the an
analysis.

Solution

Free-body diagram. The free-body diagram of the beam


To [2 866 N
AlN a N
shows the two known vertical forces of magnitudes 800 N and a
400 N, and the x and y components of the 1000-N inclined cl 'SOON

Teh
force. The magnitudes of the x and y components of the
inclined force are given by

F.. = 1000 (cos 60°) = 500 N ©)


F, = 1000(sin 60°) = 866 N @) Los

The free-body diagram also shows two unknown reaction the equilibrium equation {M, = 0 contains only one unknown,
components A, and A,, exerted on the beam by the hinged Av
support at A, and one unknown vertical reaction B,, exerted by @©smM,=0 —A,(10) + 800(7) + 400(5) + 866(2) =
the roller support at B. The senses of the three unknown
reactions are assumed to be in the positive x and y directions as Ay = 933.20N
shown on the free-body diagram. | The positive answer for A, indicates that our initial assumption
about its sense being vertically upward, or in the positive y
Static determinacy. The beam is supported by three unknown
direction, was correct; thus,
reactions, A,, A,, and B,, that are neither concurrent nor
parallel. Therefore, the penn is properly constrained (geometri- A,=933.2N(t) <
cally stable) and statically determinate; the three unknowns The only remaining unknown, B,, can now be determined by
can be determined by applying the three equations of applying the remaining equation of equilibrium,
equilibrium.
O> a0 933.2 — 800 — 400 — 866 + B, = 0
Support reactions. The three unknown reactions are determined
Bo 1132. 35N
by applying the equations of equilibrium (2F, = 0, 2F, = 0
and {M = 0) as follows: The positive answer for B, indicates that its sense is in the
Since two of the three reactions—A,, and B,—act in the vertical upward (positive y) direction as initially assumed; thus,
vertical (or y) direction, the equilibrium equation }F, = 0,
B,=11328N() <
which involves the summation of all the horizontal forces (in
the x direction), contains only one unknown, A,, which can be Note that in the above computations, in order to avoid the
determined directly from it: need for solving simultaneous equations, the equilibrium
equations were applied in such a manner that each equation
@ SF, =0 A, —500=0 contained only one unknown.
A, = 500 N
aIN a N mae
Ox
The positive answer for A, indicates that our initial assumption
about the sense of this reaction being to the right (in the SOON
positive x direction) was correct. Therefore,
933.2N
oa 7 ms
A,=500NC) < Onn 3m

Next, in order to determine A, we apply the moment


Checking of computations. Finally, to check our computations,
equilibrium equation {M, = 0, which involves the summation
we apply an alternate equation of equilibrium
of moments of all the forces about B. Since the line of action
of the other unknown B, passes through B, its moment about B © 3M. =0
is zero and it will not appear in the resulting equation. Hence, — 933.2(8) + 800(5) + 400(3) + 1132.8(2) = 0 checks

139
140 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

EXAMPLE 3.20

}~——9 ft Determine the reactions at the support for the frame shown.
}20 kip

Solution

Free-body diagram. The free-body diagram of the frame


shows the two known forces, and also the three unknown
reactions (two force components and one couple) being exerted
by the fixed support on the frame at end A. The senses of two
reaction forces A, and A, are assumed to be in the positive x
and y directions, respectively, and the sense of the reaction
couple M, is assumed to be counterclockwise as shown.

Static determinacy. The frame is properly constrained by the


three reactions A,, A,, and M,: it is statically determinate,
and the three unknown reactions can be determined by using
the three equilibrium equations.

Support reactions. Applying the three equations of equilibrium,

@ sr, =0 A 6
A,=7kip 4, =7kipG) <
Ose =0 -20+A,=0 A,=20kip()<
© =m, =0 A, = 20 kip
20(9) + 7(10) + M, = 0
M, = —250 kip - ft

The negative answer for M, indicates that its sense is opposite


to that initially assumed; that is, its correct sense is clockwise.
Therefore,

M, = 250 kip - ftQ) <


Checking of computations. We apply an alternate equilibrium
equation containing all three reactions determined above. By
summing moments about point B, we write

@ 3M, = —7(5) + 7(15) + 209) —250=0 checks


3.19 Procedure for Analysis of Reactions 141

EXAMPLE 3.21

A weight W is supported by a cable passing over a frictionless


pulley as shown. (a) Show that the magnitude of the tension in
the cable is constant, and (b) determine the reactions at the hinged
support at the center of the pulley.

Solution

Free-body diagram. The free-body diagram of the pulley and


the cable shows the known weight W, the unknown tension T in
the cable, and the two unknown reaction components A, and
A, exerted on the pulley by the hinged support at A. The sense
of the cable tension T, which is pulling away from the pulley,
is known, whereas the senses of A, and A, are not known
and are arbitrarily assumed to be in the positive x and y
directions, respectively.

Static determinacy. The three unknown reactions T, A,, and


A,, are neither concurrent nor parallel: the pulley is properly
constrained and statically determinate; hence, the three
unknowns can be determined from the three equations of
equilibrium.

Cable tension. The cable tension T can be determined directly


by applying the moment equilibrium equation about point A,
thereby eliminating the other two unknowns A, and A, from the
equation. So we write

@ 3M, =0 TRE Wik) =20 Tw <


Note that the cable tension is independent of the radius of the
pulley and the angle of inclination a. This relationship
shows that the magnitude of the tension in the cable remains
the same as the cable passes over the pulley. This property,
initially stated in Chapter 2 without proof, is usually applied by
inspection (i.e., without writing the moment equilibrium
equation) in the analysis of the cable-and-pulley arrangements.

Reactions at hinged support. Applying the two remaining


equilibrium equations, we obtain Checking of computations. By summing moments about point
B we obtain
©
i a) =0
A, — T(sin
® LF, =0
2 M0
A, =T sina =Wsina
A, = W sina C) < W cos a(R + R cosa) + W sina(R sin a) — W(1 + cos a)R
and = WRcos a + WR cos* a + WR sin? a — WR(1 + cosa)
(DSF =0 —T(cos a) + A, -—W =0 = WR cos a + WR(cos?
a + sin? a) — WR(1 + cos a)
Aga Ka + cos a) = WR cos a + WR(1)
— WR(1 + cosa)
A, = W(1 + cos a) @) < = WR(1 + cos a) — WR(1 +cosa)=0 checks
142 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

EXAMPLE 3.22

A triangular plate of 20-kg mass is subjected to two torques ap-


plied by the shafts. The center of gravity of the plate is located
at point G. Determine the reactions at the supports for the plate.

Solution

Free-body diagram. The free-body diagram of the plate is


Shafts
shown. The weight of the plate is determined by

W = mg = 20(9.81) = 196.2 N

Static determinacy. The plate is properly constrained by three


5310N-mm reactions, A,, B,, and B,, and is statically determinate.

Support reactions. The three unknown reactions are determined


by applying the three equilibrium equations as follows: since
two unknowns A, and B, are acting in the horizontal direction,
the third unknown B, can be determined directly by summing
forces in the vertical direction. Thus,

© 35,=0 — 196.2 + B, =0
B, = 196.2 N B,=196.2N@) <
By summing moments about B, we eliminate B, from the
equilibrium equation, which contains only one unknown, A,.
We write

© =m, =0
A,(270) + 196.2(170) — 5310 + 10,000 = 0
or

A, = —140.9N A,=1409N@ <

The third unknown B, is computed from the remaining


equilibrium equation,

@ 2F, =0
— 140.9 + B,=0 B. = 140.9 N
B,=1409NC) <
Checking of computations. We check the computations by
100 mm
summing moments about G:

© ~M, = 0
— 140.9(170) + 196.2(170) — 140.9(100)
170mm — 5310 + 10,000 = 0.9 N- mm =~0 checks
A The slight error of 0.9 N - mm is due to the rounding off of the
values of A, and B,. The results of the analysis are considered
30 aay 140.9N 5310N*mm
correct.
3.19 Procedure for Analysis of Reactions 143

EXAMPLE 3.23

A crate weighing 30 KN is being lifted by a crane attached to the


bed of a truck. Determine the reactions being exerted on the front
and rear axles of the truck. The truck weighs 75 kN and the crane
boom weighs 2.5 KN. The centers of gravity of the truck and the
boom are located at G and G,, respectively, as shown in the
figure.

Solution The free-body diagram of the truck is shown. Note


that A, and B, denote the reactions being exerted by the ground
on both front wheels (on the front axle) and both rear wheels
(on the rear axle), respectively. Since the truck is supported by
only two support reactions, A, and B,, it is partially constrained.
The two unknown reactions can be obtained by applying
the following two equations of equilibrium:

@ =m, = 0
— A,(5) + 75(3) — 2.5(1.91) — 30(4.33) = 0
A, = 18.07 KN
© 8, = 0
iW = 18 9D, = 95 =
B, = 89.43 kN
Equilibrium check. In order for the partially constrained truck
to be in equilibrium under the given loading condition, the third
equilibrium equation {F, = 0 must also be satisfied. From the
figure, we can see that since all the loads and reactions acting
on the truck are vertical, the equilibrium equation {F, = 0
is automatically satisfied. Therefore, the truck is in equilibrium
under the given loading condition.

Checking of computations. Summing moments about G, we


write
@ =u.
=0
= 18.07@)) + 89:43(3) — 2.54.91)
ae (255) a—m0) checks

Thus,

A, =18.07kN(@) <
B,=89.43kN(t) <
144 Chapter 3 Equilibrium of Rigid Bodies in Two Dimensions

EXAMPLE 3.24

Atte ott este Determine the reactions at the supports for the frame.

Solution The free-body diagram of the entire assembly or


system, that is, the frame and the two pulleys together, is
shown. The frame is properly constrained by three reactions,
A,, Ay, and By: it is statically determinate.
Applying the three equilibrium equations for the support
reactions,

@ SF. =0
A, + 16=0
A,=-16kip A,=16kip© <
@ =m, =0
~A,(10) + 15(4) — 16(6) = 0
A, = —3.6 kip A, =3.6kip@) <
© XF,=0
—3.6- 15+ 12+B,=0
B, = 6.6 kip B,=6.6kip (1) <
3.19 Procedure for Analysis of Reactions 145

EXAMPLE 3.25

A truck of weight P is traveling over a bridge at a constant speed.


(a) Derive the expressions for the reactions at supports A and B
in terms of the weight of the truck P and its position as measured
by the distance x; (b) draw plots showing the variations of the
vertical reactions at supports as functions of x; and (c) determine
the weight of the truck and its position if the vertical reactions at
supports A and B are 15 kip and 5 kip, respectively, in the upward
direction. Given: L = 60 ft.

Solution The structure is properly constrained by three


reactions A,, A,, and B,: it is statically determinate. P

a. Expressions for support reactions. By applying the three


equations of equilibrium, we obtain a as
l A,
@ SF, =0 A,=0 A,=0 < ai le
A, |
© 2M, xm, = 0
+— 1/3—r4-— 1 /3->}-1/3—>

b. The plots showing the above variations of A, and B, as


functions of x are shown on the right.

c. By substituting A, = 15 kip, B, = 5 kip, and L = 60 ft in


the expressions for A, and B,, we obtain
xe
is=P(2-%) (a)

x
s=r(=-1) (b)

(continued)
146 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

EXAMPLE 3.25 (concluded)

Multiplying Equation (b) by 3 and subtracting it from Equation Substituting x = 25 ft into Equation (a), we obtain
(a), we get 25

p(2- 2)-3°(4-1) = is= (2-75) P=20kip <


° ma Therefore, a truck weighing 20 kip, when located at a distance
2 eo 3 x re of 25 ft from the left end of the bridge, will cause the vertical
20 O20) reactions A, = 15 kip and B, = 5 kip.
Cee Finally, we check our computations by using the equilibrium
DSS0 = Sap
20 ll :
equation
weet Orr 6 A, +B,—-P=0
IS SE 5 = PSO
P = 20 kip checks

3.20 SUMMARY
In this chapter, we have learned the following:
1. Principle of transmissibility. A force acting at a point of a rigid body
may be moved to any other point along its line of action without
changing the conditions of equilibrium or motion of the body
(Section 3.3).
2. If all the forces acting on a body lie in a single plane, the body can be
treated as a two-dimensional body for analysis (Section 3.4).
3. The moment of a force about the axis normal to the plane through a
point O is defined as the force times the perpendicular distance
(Section 3.5):
Mo = Fd (358)
4. The moment of a force F about a point O is a vector defined as the
cross product of a position vector r from point O to any point on the
line of action of the force, and the force F (Section 3.7):

Lae eek
Mo =r, XB Fee aes (3.10)
PPS Fe
The direction of M;, is perpendicular to the plane containing r and F,
with its sense specified by the right-hand rule.
5. Varignon’s theorem. The moment of the resultant of concurrent forces
about a point is equal to the sum of the moments of the components
about the point (Section 3.8).
6. A couple is a pair of equal, opposite, and parallel forces. The magnitude
of the moment of a couple is equal to the product of the magnitude of
3.20 Summary 147

the forces forming the couple and the perpendicular distance between
their lines of action.
M=Fd (3.16)
The moment of a couple is a vector defined as the cross product
M=rxF (3.17)
The moment of a couple is a free vector with its line of action
perpendicular to the plane containing the forces, and its sense specified
by the right-hand rule (Section 3.10).
A force acting at a point on a rigid body may be moved to another
point if a couple, with the moment equal to the moment of the force
acting at the initial point about the new point, is added on the body.
A force-couple system acting on a rigid body may be reduced to a
single force by moving the force to a point so that the moment of
the force acting at the new point about the initial point is equal to the
moment of the given couple (Section 3.11).
A system of forces acting on a rigid body may be reduced to an
equivalent one force—one couple system at a point by the process of
moving each force to the point by adding a couple (as described in
item 7), and then adding all the forces and the couples, respectively, to
obtain the resultant force and the resultant couple.
The resultant force-couple system thus obtained may be further
reduced to a single resultant force by using the procedure described in
item 7 (Section 3.12).
A rigid body is in equilibrium if, at rest initially, it remains at rest
when subjected to a system of forces and couples. For a body to be in
equilibrium, the resultant force-couple system of the forces and couples
acting on it must be zero:
R=>F=0 M,=>M=0 (3.21)
These equations are called the equations of equilibrium for rigid bodies
(Section 3.13).
10. For two-dimensional rigid bodies subjected to coplanar forces, the
equations of equilibrium can be expressed in scalar form as

LF, =0 LF, =0 +M =0 G22)


Two alternate forms of the equilibrium equation are

LF, =0 >M,=0 ~M, =0 (3.23)


provided line AB is not perpendicular to the p axis, and

~M, =0 +M, = 0 >~M- =0 (3.24)


provided points A, B, and C are not collinear (Section 3.14).
ii. When only two forces are acting on a rigid body in equilibrium, they
must be equal, opposite, and collinear.
148 Chapter 3 Equilibrium of Rigid Bodies in Two Dimensions

When three forces are acting on a rigid body in equilibrium, they


must be either concurrent or parallel (Section 3.15).
12. Supports and connections are used to restrict movements of rigid
bodies to maintain equilibrium. The characteristics of common types of
supports and reactions used for two-dimensional bodies are summarized
in Table 3.1 (Section 3.16).
13. A body is considered to be statically determinate externally if all of its
reactions can be determined by solving the equations of equilibrium.
If a rigid body is supported by r number of reactions, then if
r<3 _ bodyisstatically unstable externally
(partially constrained)
(3.26)
r=3 bodyis statically determinate externally
r>3 _ bodyis statically indeterminate externally

Even when a body is supported by a sufficient number of reactions


(r => 3), if the reactions are concurrent or parallel, they cannot maintain
the body in equilibrium under a general coplanar system of forces and
couples. Such a body is called geometrically unstable or improperly
constrained (Section 3.17).

KEY TERMS
S990
ne ee ena

applied force 80 reaction 80


coplanar force system //3 reaction force 80
couple /00 resultant couple (coplanar force
cross product (of vectors) 87 systems) J//

degree of indeterminacy /30 resultant force (coplanar force


systems) ///
equations of equilibrium for rigid
bodies 119 resultant force-couple system
(coplanar force systems) /12
external force 80
right-hand rule 87
free-body diagrams for rigid
bodies /34 statically determinate bodies /27

geometric instability /33 statically indeterminate bodies /30

improperly constrained body /33 statically unstable structures /30


internal force 80 supports (for two-dimensional
bodies) /25
moment 83
three-force member /23
moment vector 90
two-force member /23
partially constrained body /30
principle of transmissibility 80
Problems 149

PROBLEMS
OEE SS Lt a a a

SECTION 3.5
3.1 Determine the moment of the force about point A. @~— 300N

3.2 Determine the moment of the force about points A and B.

3.3 Determine the moment of the force F = 300 Ib about point A on the lever, if ate
@ = 30°.

3.4 Determine the magnitude and direction of the smallest force F which produces yA
an 800 lb - ft counterclockwise moment about point A.
Figure P3.1
3.5 Determine the moment of the force F = 130 N about point A on the lever, if
@ = 30°.

3.6 Determine the magnitude and direction of the smallest force F which produces
a 120 N - m counterclockwise moment about point A.

a\ \ Figure P3.2

Oy ve
a
Figure P3.3, P3.4, and P3.7 Figure P3.5, P3.6, and P3.8

SECTION 3.9
3.7 and 3.8 Solve Problems 3.3 and 3.5 by resolving the force into components in ee
the longitudinal direction of the lever, and the direction perpendicular to it.

3.9 Determine the moment of the force F = 350 Ib about points A and B on the
lever if 8 = 40°. Use the scalar method.

3.10 Solve Problem 3.9 by using the vector method.

3.11 Determine (a) the magnitude and direction of the smallest force F that
produces a 1000-lb - ft counterclockwise moment about point A, and (b) the moment
of this force about point B. Use the scalar method.

3.12 Determine (a) the magnitude and direction of the smallest force F' that Zses
produces an 800-Ib - ft clockwise moment about point A, and (b) the moment of this a a
force about point B. Use the vector method. Figure P3.9, P3.10, P3.11, and P3.12
150 Chapter 3 Equilibrium of Rigid Bodies in Two Dimensions

3.13 A 12-N force is applied to a steering wheel as shown. (a) Derive the
expression for the moment at the center of wheel O in terms of the direction of the
force 0; (b) determine the direction of the force which produces the maximum
clockwise moment at the center O; and determine the magnitude of this moment. (c)
In what direction can the force be applied to produce the maximum counterclockwise
moment at O? Determine the magnitude of this moment.

Figure P3.13

3.14 Determine the total moment of all the forces about point A.

3.15 Determine the total moment of all the forces about point B.

3.16 Determine the magnitude and sense of the force F that produces the same
moment about the support point A as the total moment caused by three known forces
350 mm
about A.

4kip 8kip 6kip F

=LS
300mm 200 mm
Figure P3.14 and P3.15

Figure P3.16

ESrigh shesying) 3.17 A pole is supported by two cables as shown. Determine the unknown cable
Figure P3.17 tension T so that the total moment at the base O is zero.
Problems 151

Pa 2ft
spree
500 Ib

Figure P3.18

3kN

A B

Oe | se accanea
Figure P3.19

3.18 Determine the magnitude and direction of the smallest force F at point B that
produces the same moment about point A, as the total moment caused by the three
known forces, but with opposite sense.

SECTION 3.10
3.19 and 3.20 Determine the moment of the couple shown. Replace the given
couple by an equivalent couple consisting of the smallest forces acting at points A
and B. Use the scalar method.

3.21 Solve Problem 3.20 by using the vector method.

60°
450 Ib
PING ple
Figure P3.20 and P3.21
Chapter 3 Equilibrium of Rigid Bodies in Two Dimensions

3.22 and 3.23 Determine the magnitudes and senses of the two vertical forces F
exerted by the supports at points A and B to counteract the rotational tendency of the
applied couples. Use the scalar method.

3.24 Solve Problem 3.23 using the vector method.

3.25 The triangular plate is subjected to two torques applied by the shafts as shown.
Determine the magnitudes and senses of the two horizontal forces exerted by the
bolted supports at A and B to balance the rotational tendency of the applied torques.

6kN 2kN 6kN 2kN

ait oe ee
Figure P3.22

8 kip

300 Ibsin.

Figure P3.23 and P3.24

lm

180N

3m
300 mm

Cae 3 ft
A Aol

Figure P3.26 Figure P3.27 and P3.29

SECTION 3.11

3.26 through 3.28 Replace the given force by an equivalent force-couple system at
point A.

3.29 and 3.30 Replace the given force by an equivalent force-couple system at
Figure P3.28 and P3.30 point B.
Problems 153

3.31. A concrete column is subjected to a 13-kip axial load which is eccentric with
respect to its axis of symmetry through the center of gravity G. Replace the force by
an equivalent force-couple system at G.

3.32 through 3.34 Reduce the given force-couple system to an equivalent single
force. Specify its location from point A.

3.35 Determine the single equivalent force causing the given force-couple system at
the center of the bolt A if 9 = a = O°. Specify the location of the single force on
the handle of the wrench from point A.

3.36 Solve Problem 3.35 if 8 = 30° anda = 60°.

wee —_
be 2m eelTeens. m
Figure P3.32

ea a...
Figure P3.31 Figure P3.33

250 Ib-in.
0.6 kN

330kN-mm
Figure P3.34 Figure P3.35 and P3.36
154 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

7 kip 9 kip 12 kip 7 kip OkN

A LT 2.5m
a :
EN ee te 4KkN |"
2ft 3ft —_——
Figure P3.38 1.5m
<—6kN 5 kN a

2m

Figure P3.37 Figure P3.39

1Skip 9kip 6kip 20 kip 4kip ISkip 25kip


_ke
a_| cb
20°. Oly
175 kip -ft
30 kip «ft
Okinooh Sit 4h 3f 3ft
5 ft 6 flee ft i
Figure P3.40
pitt 3 ft
Figure P3.42

16kN
Figure P3.41

SECTION 3.12
3.37 through 3.42 Determine the resultant force-couple system at point A.
3.43 Determine the resultant of the four forces and one couple acting on the plate.
Specify the points where the resultant’s line of action intersects the edges of the
plate. Use the scalar method.
3.44 Solve Problem 3.43 by using the vector method.
3.45 Determine the resultant force-couple system at the center of gravity G of the
plate.

3.46 Determine the resultant of the four forces and one couple acting on the piping
system. Specify the location of the resultant on the pipe from point A. Use the
scalar method.
3.47 Solve Problem 3.46 by using the vector method.
Problems 155

600mm 400mm 300 Ib

4kN
G D | ae a 200lb ¢ A
SE 12 in. 450 lb-in.
S| it
300mm 50 1b :
wy125 Ib | 25 kN
26 in. ae 24 in.
5 kN 450mm 6 in. 6 in. 12kKN
°G Figure P3.46, P3.47, and P3.48 SOkN
6kN 2m
840 kN-mm 55 iz
mm een a
500 mm
4
2m

A 0772 FXBe
oe AB
10 kN EE
= >

500mm 500mm 2.5 m 2.5 m—>


Figure P3.43, P3.44, and P3.45 Figure P3.49 and P3.50

3.48 Determine the resultant force-couple system at point A.

3.49 Determine the resultant of the four forces and one couple acting on the frame.
Specify the points where the resultant’s line of action intersects the columns AC
and BD of the frame. Use the scalar method.

3.50 Solve Problem 3.49 by using the vector method.

SECTION 3.18
3.51 through 3.57 Draw the free-body diagram of the body shown. For the given
loading, can all the reactions be determined by solving the equations of equilibrium?
Explain.

tei e ne
emoccegt ieeeA Sone
seca
surfa
5 pee SS, ft
2 ft
Figure P3.52
156 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

500 Ib

Figure P3.53 a B
4 ft 4 ft iil eae
Figure P3.54

Smooth
Smooth Smooth Silotl splivecs ‘surface
sphere surface
Smooth
surface

Smooth
surface
Figure P3.55 Figure P3.56
surface
Figure P3.57

SECTION 3.19
3.58 through 3.67 Can all the reactions at the supports for the frame be determined
by solving the equations of equilibrium? If yes, determine the reactions. If no,
explain.

3.68 through 3.70 Determine the reactions at the support for the cantilever beam.

3.71 through 3.75 Determine the reactions at the supports for the beam.

3.76 through 3.78 Determine the reactions at the supports for the frame.

3.79 Determine the reactions at support A for the piping system. Assume support A
as a fixed support.

3.80 Determine the reactions at the supports for the boom.

3.81 A crane weighing 4 KN is supported by a hinge at A and a cable BC as shown.


Determine the reactions if F = 12 kN andd = 3.5 m.

3.82 For the crane in Problem 3.81, if the maximum allowable tension in the cable
is 150 kN, determine the maximum force F at a distance d = 5m that can be
carried by the crane.

3.83 For the crane in Problem 3.81, determine the relationship between the
maximum force F and its location, d, so that the tension in the cable is equal to the
maximum allowable value of 150 kN. Plot the maximum force, F, as a function
of its location d in the range 0.5 m =d = 6m.
Problems 157

— >
ve C 5 kip
2 ift

A
3 kip Rocker 3 ki Dp
a 4m—+}—4m—4 ee eee, sies|
Figure P3.58 Figure P3.59 Figure P3.60

20kN

G Cable

20kN
4m

ib

Figure P3.63

Cable

3 kip 5 kip
ee lie!
Figure P3.66

3kN te

eee eee era


Figure P3.67 Figure P3.68 Figure P3.69
158 Chapter 3 Equilibrium of Rigid Bodies in Two Dimensions

12kN
: fees a

._0:9m ,| 0.9m| tee: 13 ah Sifter be 6m + es


Figure P3.70 Figure P3.71 Figure P3.72

7 kip 9 kip 12kip 7kip ai 9kip 6kip 20kip


| esl st emt | ; fe Foes) SO kip-ft
nen 20 cen|al
wee DS pti =. 4°
A j
— ay Ae ae sical wk 60 Fa oa
Df Sat De Bite Sie aa Site Bai
Figure P3.73 Figure P3.74 Figure P3.75

100kN-m

1.2.5 m 2.5 m—>|


Figure P3.76 Figure P3.77

a
12 in_,.450 Ib-in.
50 tt ore |
125 Ib
sees 24 in.
+10 ft 10 ft——+ 26 in “Weinioint
Figure P3.78 Figure P3.79
Problems 159

Figure P3.80
Figure P3.81, P3.82, and P3.83

3.84 The moveable dump is attached to the bed of a truck by a hinged connection
at A and is supported in the position shown by the hydraulic cylinder BC. The dump
weighs 1200 Ib and its center of gravity is located at G. Determine the reactions at
hinge A and the force exerted by the hydraulic cylinder BC at B.

FAL,
Figure P3.84

3.85 A truck weighing 75 KN is in equilibrium on a 10° incline as shown.


Determine the reactions on the front and the rear axles of the truck and the tension in
the towline T for 8 = 0°.

3.86 Solve Problem 3.85 if 8 = 45°.

Figure P3.85 and P3.86


160 Chapter 3 Equilibrium of Rigid Bodies in Two Dimensions

3.87 and 3.88 Determine the reactions at the supports for the structure.

20 kip:ft

nelle Petals aan

Figure P3.87 Figure P3.88

3.89 A sign weighing 1200 lb is supported on a pole by a hinge at B and a pulley C


as shown. Determine the reactions at hinge B and tension in the cable.

i r=6in.

) G

Let ft
2ft

Figure P3.89 and P3.90

3.90 For the pole in Problem 3.89, determine the reactions at the fixed support A.

P 3.91 The reaction at support B for the beam is 18 KN in the vertically upward
A B direction. Determine the load P and the reactions at support A.

3.92 The reactions on the front and the rear axles of the truck are 12 kip and 8 kip,
respectively, in the vertical direction. Determine the weight W of the truck, and the
Die 5 sda geet horizontal distance a of its center of gravity from the front axle.
Figure P3.91
3.93 The vertical reactions at the supports A and B for the frame are 14 kN
downward, and 63 kN upward, respectively. Determine the load P and its location as
measured by the distance a. Given: 8 = 90°. Also determine the horizontal reaction
at A.
Problems

4m>+—7 m—+#-4 m
Figure P3.92 Figure P3.93 and P3.94

3.94 Solve Problem 3.93 for 6 = 75°.

3.95 The vertical reactions at supports A and B for the frame are 13 kip and 37 kip,
respectively. Determine the magnitude of the forces F, and F,. Also, determine the
horizontal reaction at A.

on ota is

cnery

Figure P3.95

3.96 Determine the maximum weight that can be lifted by the crane without tipping
the truck over. The truck weighs 18 kip and the weight of the boom is 800 lb. The
centers of gravity of the truck and the boom are located at G and G,, respectively, as
shown.

Figure P3.96
162 Chapter 3. Equilibrium of Rigid Bodies in Two Dimensions

3.97 Two smooth uniform spheres are supported by two smooth surfaces as shown.
Derive the expressions for the reactions being exerted on the spheres at contact
points in terms of the angle of inclination 0.

Figure P3.97

3.98 A beam is supported by two rollers on inclined surfaces as shown. Determine


the position a of the load for which the beam will be in equilibrium. Given: L =
9m, P = 20 kN, 6 = 45° anda = 30°.

3.99 Derive the expression for the position a of the load, in terms of the angles of
inclination 6 and a, corresponding to the equilibrium of the beam.

P
A B
ie CRU ee

Figure P3.98 and P3.99


CHAPTER

EQUILIBRIUM OF RIGID
BODIES IN THREE
DIMENSIONS

4.1 Introduction OUTLINE

Three-Dimensional Force Systems

4.2 Moment of a Force about a Point


4.3, Dot Product of Vectors
4.4 Moment of a Force about an Axis
4.5 Couples
4.6 _Resultants of Nonconcurrent Three-Dimensional Force Systems

Equilibrium in Three Dimensions

4.7 Equilibrium of Rigid Bodies


4.8 Types of Supports and Connections
4.9 Statically Determinate Structures
4.10 Free-Body Diagrams
4.11 Procedure for Analysis of Reactions
4.12 Summary
Key Terms
Problems

163
164 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

4.1 INTRODUCTION
In Chapter 3, we discussed the analysis of coplanar systems of forces and the
equilibrium of rigid bodies subjected to such forces. While many actual three-
dimensional structures can be considered as two-dimensional bodies subjected
to coplanar forces for the purposes of analysis and design, some structures,
due to applied loading or to the arrangement of their members and supports,
cannot be treated as two-dimensional. Such structures must be analyzed as
three-dimensional bodies subjected to three-dimensional force systems.
In this chapter, we will extend the formulations developed in Chapter 3 to
analyze general three-dimensional systems of forces and couples acting on
rigid bodies, and the equilibrium of rigid bodies subjected to such forces. The
basic concepts of a rigid body, the transmissibility of a force on a rigid body,
and the external and internal forces acting on rigid bodies remain the same in
both two and three dimensions. The main difference between the analysis of
two-dimensional and three-dimensional systems arises from the difficulty in
visualizing the orientation of bodies and forces in three-dimensional space.
Except for some simple cases, the intuitive scalar approach cannot be used for
the analysis of three-dimensional problems; the vector approach is more ap-
propriate for dealing with such problems.

THREE-DIMENSIONAL FORCE SYSTEMS

4.2 MOMENT OF A FORCE ABOUT A POINT


Consider a three-dimensional rigid body subjected to a force F as shown in
Figure 4.1. As discussed in Section 3.7, the moment of F about a point A on
the body is given by the cross product M, = r X F, where r represents the
position vector from point A to any arbitrary point P on the line of action of
F. The magnitude of the moment vector is given by M, = rF sin 8, where 0
represents the angle between the lines of action of r and F. As shown in Figure
4.1, the moment vector M, passes through point A and is directed perpendic-
ular to the plane containing the position vector r and the force F. The sense
of the moment vector M, is specified by the right-hand rule.
Moment vectors obey the law of vector addition and, like force vectors,
they can be treated as sliding vectors. To distinguish them from the force
vectors, the straight arrows representing the moment vectors are usually su-
perimposed by curved arrows indicating the rotational sense of the force, as
we see in the figure.
Figure 4.1
Moment in Cartesian Vector Form

By resolving the force F and the position vector r into their rectangular com-
ponents (Figure 4.2)

Be ia eeeKk
i el ak
4.2 Moment of a Force about a Point 165

Wilye Wee Tage

lie peel le

Maz Tf, —r,F,

Figure 4.2

we can express the moment of F about point A as


M,=rxF
Sa Wein | aur WS) (Fi+ Pye k)

Using the distributive property of cross products, and the cross products of
unit vectors determined in Section 3.6 [Equation (3.6)], we obtain

M,= (r= rF,)i he ror) I


(4.1)
SP FE = ry Fk

As discussed in Section 3.6, the above relationship can be expressed in the


determinant form
ie] @ ok
Ma ee (4.2)
IE es

By writing the moment vector M, in Cartesian vector form


My, = My,i + Maj + Mak (4.3)
and by comparing Equations (4.1) and (4.3), we can see that the magnitudes
of the rectangular components of the moment vector M, are given by

Be zy

May oD he a: Ae (4.4)
166 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

The three components M,,, M,,,, and M,, measure the rotational effect of the
force F about the x, y, and z axes respectively. A positive value for the mag-
nitude of a moment component indicates that the component points in the
positive direction of the axis. Conversely, a negative value indicates that the
moment component points in the negative direction of the axis.
When the origin of the Cartesian coordinate system is located at a point A
about which the moment is desired, as in Figure 4.2, then the position vector
r can be directly written in terms of the x, y, and z coordinates of point P
on the line of action of the force F. From this figure we can see that r =
xi + yj + zk. By substituting r, = x, r, = y, andr, = z into Equation (4.2),
we can write the moment vector M, as
iy) j Sek
M,=rxF=|x y z (4.5)
Ey ES ie

However, when the origin of the coordinate system is not located at point
A where the moment is desired, but at some other point O as shown in Figure
4.3, then the position vector r from A to a point P on the line of action of F
can be written as
r=rp-—T,

Figure 4.3
4.2 Moment of a Force about a Point 167

where r, and rp are the position vectors from the origin O to points A and P,
respectively. From this figure we see that these two position vectors can be
written in terms of the coordinates of points A and P as
r,=%X,i+ y,j + z,k Ip =Xpi + ypj + zpk
The position vector r, from A to P, can then be given as
r— O7— x, ia (ye — y4)j + Gp — z,)k

Substituting r, = xp — X4,7y = yp — Ya, and r, = Zp — z, into Equation


(4.2), we obtain the following expression for the moment M,, of the force F
about point A
i J k
M, =rxF=|@p-—x%) Or- Va) pam 2a) (4.6)
F.x Fy F,

EXAMPLE 4.1

A transmission tower is supported by three cables. The tensions


in cables CE and DG are 500 |b and 1200 Ib, respectively. De-
termine (a) the moments of the two cable forces exerted at points
C and D, respectively, about the base point A; (5) the total mo-
ment due to both cable forces about point A; and (c) the perpen-
dicular distance from point A to cable CE.

Solution A right-handed x-y-z coordinate system with origin


at point A is chosen as shown. Points C and D on the tower
are chosen as the points of application of the 500-lb and
1200-lb forces, respectively.
The magnitudes of the rectangular components of the cable
force T., which has a magnitude of 500 Ib, are

Tc, = —500(sin 30°)(sin 60°) = —216.5 1b


Tey = —500(cos 30°) = — 433.0 Ib
Tc, = 500(sin 30°)(cos 60°) = 125.0 Ib Thus, the force T, can be expressed in terms of its rectangular
Thus, components as
Te = — 216.51 — 433] + 125k (Ib) T,) = Tpn = 1200(0.188i — 0.941j — 0.282k)
The second cable force T, has a magnitude of 1200 lb and = 225.6i — 1129.2j — 338.4k (Ib)
is directed along line DG. The unit vector along line DG
(from D to G) is determined as a. Moments about point A. The moment of the cable force
, ; T, about point A is given by the cross product Myc = Tac X
n= (6 — Oi + @ = 30)j + (—9 — Ok T,, where r4- denotes the position vector from A to the
V(O)ea ao 30)e C= 9)"
= 0.188i — 0.941j — 0.282k (continued)
168 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

EXAMPLE 4.1 (concluded)

point of application C of the force T;. We can see that


T4c = 25j (ft). Therefore,

Mac = Fac X Te = 25j X (—216.5i — 433j + 125k)


= —(5412.5)(j X i) — (10,825)(j X j) + (3125) X k) Tp =12001b
Recalling the cross products of unit vectors from Equation
(3.6), we obtain

Myc = 5412.5k + 3125i (Ib - ft)


Myc = 3125i + 5412.5k (Ib: ft) <

From this, we find the magnitude of M,c- to be


Mac = V3125)? + (5412.5)? = 6250Ib-ft
The moment of a force about a point can also be computed
using the determinant given in Equation (4.6). Using this
alternate method to determine the moment of the cable force M47 = —7027i -1355.5k(Ib-ft)
Z
T, about point A, we have
M 4p = —10,152i-6768k (Ib-ft)
i j k
Map = |p —%4) (Yo — Ya) (pd — Za)
Tp. Tpy Tp,

i j k
= (00) G0 0). sO 0)
256 1129.2 3384
el 930 (ie pA! 0
7% Vd 1909) 2338.4 @21995.6) 3384
0 30
os ek |
= i(— 10,152 + 0) — j(O — 0) + k(O — 6768)
Map = —10,152i — 6768k (lb-ft) < The magnitude of the total moment is

Thus, the magnitude of M,, is Mar = V (—7027)? + (— 1355.5)? = 7157|b- ft

Map = V(— 10,152)? + (— 6768) = 12,200Ib-ft c. The perpendicular distance from point A to cable CE can be
determined by using the basic relationship M = Fd; that is,
the magnitude of the moment of a force about a point is equal
b. Total moment about point A. The total moment due to
to the magnitude of the force times the perpendicular distance
both forces, T. and Tp, about point A is given by the vector from the point to the line of action of the force. Thus, the
sum
perpendicular distance d from A to the cable CE, which is the
line of action of T;, is determined from the relationship
Mar = Mac + Map
= (3125i + 5412.5k) + (— 10,152i — 6768k) Myc = Ted
M,7 = — 70271 — 1355.5k (Ib: ft) < 6250 = 500d d=12.5ft <
4.2 Moment of a Force about a Point 169

EXAMPLE 4.2

A frame is subjected to a force F. If the force has a magnitude of


80 KN, with the direction angles a = 60°, B = 60°, and y =
135°, determine the moment of the force about point E of the
frame.

Solution A coordinate system with origin at support C of the


frame is chosen. The coordinates of point P, which is the
point of application of the force F, are

Xp=4m yp = 5m Zp=0

The magnitudes of the x, y, and z components of the 80-kN


force are

F.. = F(cos a) = 80(cos 60°) = 40 kN


F, = F(cos B) = 80(cos 60°) = 40 KN
F, = F(cos y) = 80(cos 135°) = —56.6 kN

Therefore, the force F can be written as

F = 40i + 40j —56.6k (KN)


The coordinates of point E are

Xp =0 yp =2.5m Zp =8m

The moment of F about point E is given by


i j k
Mz = |p — Xe) (Yp— ye) (Zp — Zz)
Ee FY 1

i j k
== (Ae) a (eee) (en)
40 40 — 56.6

el CS te) us 4 = Lia ay 5)
ay 40 —56.6 J 40 —56.6 40 40

= i(— 141.5 + 320) — j(—226.4 + 320) + k(160 — 100)


My, = Fep X F = (41 + 2.5j — 8k) X (401 + 40j — 56.6k)
M, = 178.5i — 93.6j + 60k (KN-m) <
= (160)(i x i) + (160)(i x fj) — (226.4)(i x k)
The moment M, can also be computed by performing the cross + (100)(j x i) + (100) x j) — (141.5) x k)
product rpp X F directly. The position vector rzp from E to P
is given as
— (320)(k x i) — (320)(k x j) + (452.8)(k X k)
= 160k + 226.4j — 100k — 141.5i — 320j + 320i
lep =Vp — Vg = (41 + Sj) — (2.5j + 8k)
M,, = 178.5i — 93.6j + 60k(kKN-m) <
= 41 + 2.5j — 8k(m)
The magnitude of the moment M,, is
M, = V (178.5) + (—93.6)? + (60)? = 210.3kKN-m
170 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

4.3 DOT PRODUCT OF VECTORS


We have thus far considered the moment of a force about a point. In many
engineering applications, it is necessary to determine the moment of a force
about a line or an axis, rather than a point. Before we discuss the concept of
moment of a force about an axis, it is useful to understand another type of
vector multiplication called the dot product of vectors, also referred to as the
scalar product.
The dot product of two vectors P and Q, shown in Figure 4.4, is defined
as a scalar of magnitude equal to the product of the magnitudes of vectors P
and Q times the cosine of the angle between them. The symbol : is used to
denote the dot product. For example, the dot product of vectors P and Q is
written as

Figure 4.4
D=P-Q=P0 cos0 (4.7)

where 6 is the angle between the lines of action of the vectors P and Q, when
placed tail to tail as shown in Figure 4.4. From Equation (4.7), we can see that
the dot product may also be considered as the product of the magnitude P of
P and the magnitude Q cos 0 of the component of Q in the direction of P;
or conversely, as the product of the magnitude Q of Q and the magnitude
P cos 0 of the component of P in the direction of Q.
Since the dot product of vectors does not depend upon the order in which
the vectors are multiplied (PQ cos 8 = QP cos 9), the dot products are com-
mutative, or

P-Q=Q°P (4.8)

Another useful property of dot products is that they are distributive:

P(Q> S+ T)=P-O-42PeS + Pel. (4.9)

Dot Products of Cartesian Unit Vectors

The dot products of the Cartesian unit vectors i, j, and k with themselves
or with each other can be determined using Equation (4.7), realizing that
cos 0° = 1 and cos 90° = 0. The dot product of the unit vector i with
itself is i- i = (1)(1)(cos 0°) = 1. Similarly, the dot productsj - j and k - k
are also equal to one.
Now consider the dot product of unit vectors i and j, which is i+ j =
(1)(1)(cos 90°) = 0. Since the dot products are commutative, j-i = i-j =
0. From the definition of the dot product as given in Equation (4.7), we can
see that the dot product of any two vectors that are perpendicular to each other
is always zero because cos 90° = 0.
4.3 Dot Product of Vectors 171

The dot products of all pairs of unit vectors are listed here:

i= i:j=0 itik=0
jii=o0 jj= jk=0 (4.10)
k:i=0 k:j = k-k=

Vectors in Cartesian Form

By using the dot products of unit vectors, we can develop an expression for
dot products of two vectors in terms of their rectangular components. Consider
two arbitrary vectors P and Q, given as
Vesa oe Lk O— Or O77 Ok
The dot product D of these vectors is given by

D=P-Q= (Pi + P,j + £,k)-(0,1 + Oj + @,k)


Using the distributive property of dot products, we can write

EEO SO er 200) et Ok)


tg O74 it OK j) eh Ok)
+ P,O.(K*i) + P,O,(kK-j) + P.Q.(k-k)
By performing the dot products of unit vectors [Equation (4.10)], we obtain

D =P-Q=P,O, + P,Q, + P.O, (4.11)


Thus, the dot product of two vectors is equal to the algebraic sum of the
products of the magnitudes of their corresponding rectangular components.

Angle between Two Vectors


The concept of dot product is useful for determining the angle between two
vectors, P and Q, given in terms of their rectangular components. Equating
Equations (4.7) and (4.11) gives

P-O = PO cos 0 = FOF sat O, + 20,


from which

Cos 0l—
ay Oe Weve ae
iO PQ
Thus, the angle 6 between the vectors P and Q is given by
2 IP ae ie ar 1ELO!
= 00s 1(E2) = cos # (SER TRO) (4.12)
172 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

Component of a Vector along a Line


Another important application of dot products is for determining the compo-
nent or projection of a vector along a given line or axis. Consider a vector P
as shown in Figure 4.5. We wish to determine the component of P along line
Oa, whose direction is known and is specified by the unit vector n. If the angle
6 between the vector P and line Oa can be determined, then the magnitude of
the component of P along Oa is given by
Poqg = P cos 9
The scalar P,,, also called the projection of P on line Oa, can be written in
vector form as
Poa = (Pog)n = (P cos 6)n

In two-dimensional problems, the angle @ is usually obtained directly from


trigonometric considerations, and the component Po, is determined using the
Figure 4.5 above equations. However, in three-dimensional problems, the angle @ usually
cannot be determined directly; in such cases, the magnitude of P,, is obtained
from the dot product

Pog = P+n = (P)(1)cos 8 = P cos 8 (4.13)

Note that the magnitude of the unit vector n is equal to | (i.e.,n = 1). Equation
(4.13) can also be written in terms of the rectangular components of P and n
as
Po, —P-n =F, cos 04k, Cos Bia Ff, cosy (4.14)
in which a, B, and y are the angles of line Oa with respect to the x, y, and z
axes, respectively, as shown in Figure 4.5. The component may be expressed
in the vector form as
Pog = (P*n)n (4.15)

Mixed Triple Product


The mixed triple product of three vectors R, P, and Q is defined as the dot
product of R with the cross product of P and Q, or

M=R-°(P X Q) (4.16)

where M is a scalar. Since the cross product P X Q yields a vector, the mixed
triple product may be considered as the dot product of two vectors.
Using the commutative property of the dot product of two vectors, we can
rewrite Equation (4.16) as

M=R:°:(P XQ) =(P XQ):R (4.17)


4.4 Moment of a Force about an Axis 173

Next, we develop an expression for the mixed triple product of R, P, and


Q in terms of their rectangular components. If vector C is the cross product
of vectors P and Q (i.e., C = P X Q), then we can write the mixed triple
product as

M=R-(PXQ)=R-C=R,C,+
RC, +RC,
Substituting the expressions of the rectangular components of the cross product
C from Equation (3.8) into the above expression, we obtain

M=R°:(P XQ)
= R,P,Q, — P,Qy) + Ry(PQ, — P,Q.) + RAP,Q, — PyQ,) (4.18)
Equation (4.18) is usually expressed in the more convenient form of a deter-
minant as

Rx R y R Zz

M=R-(PXQ)= P
. Poe)ke (4.19)
Q. OF O%

4.4 MOMENT OF A FORCE ABOUT AN AXIS


The moment of a force about an axis is defined as the component along the
axis of the moment of the force about any point on the axis. The moment about
an axis is a measure of the effect of the force to rotate the body about the axis.
Consider a rigid body subjected to a force F as shown in Figure 4.6. The
moment of F about an arbitrary point A on the axis ab is given by the cross
product

M,=rxF

where r is the position vector from point A on the axis to any point P on the
line of action of F. If n is the unit vector along the axis ab, then as we discussed
in Section 4.3, the magnitude of the component of M, along the axis ab is
given by the dot product
M,, =n-M, (4.20)
where M_,, is the magnitude of the moment of F about axis ab. Substituting Figure 4.6
M, = r X F into Equation (4.20), we can express M_, as the mixed triple
product

M,=n°'(r X F) (4.21)
174 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

When the three vectors n, r, and F are given in terms of their rectangular
components as
D7) are K
P= la tak
Re fic ek
then, using Equation (4.19), we can express the magnitude of the moment of
the force F about the axis ab in the determinant form

ny ny Z
Map Sails ry r, (4.22)
PS Fal ae

Note that M,, in this equation is a scalar quantity. Since the vector M.,, is
directed along the axis ab, it can be written as

M., = M,,n = (n'M,)n = [n-(r X F)Jn (4.23)

EXAMPLE 4.3

The space frame shown is subjected to a force F = —Si + 3j —


7k (kip) at joint H. Determine the moment of the force (a) about
line CG and (b) about line DE.

Solution

a. The magnitude of the moment of force F about line CG is


a given by the mixed triple product

Mee = Ncg' cu X F)
The unit vector ncg along line CG can be written as ncg = K.
The position vector r¢,,, from point C on line CG to the
point of application H of the force F, can be expressed in terms
of its rectangular components as ro, = 8i + 9k (ft), and the
force F is given as F = —Si + 3j — 7k (kip). Substituting
these expressions into the expression for Mag, we obtain
Meg = k-[(8i + 9k) X (—5i + 3j — 7K)
= k+(—27i + 11j + 24k) = 24 (kip -ft)
We can also determine M-, from the mixed triple product
Meg = cg * gy X F) with rgy = 8i (ft). Using the

(continued)
4.4 Moment of a Force about an Axis 175

EXAMPLE 4.3 (concluded)

determinant given in Equation (4.22),

nen an, 0 0 1 n G
Moe= |e I 07, = 8 0 0 Mog =24k
(kip-ft fs rh
Fe ee i a wR ah
G F H
San)
= 1
aif, =
24 kip - ft
j .

The moment of F about line CG, expressed as a vector, is

Mce = MceNce
Moc = 24k (kip: ft) <
b. From the figure we see that the unit vector np, along line
DE is given by

=
No
—S
2nppi(kip-ft)
DE — 8i — 6j + 9k
Ea Din
DED Paataa
A (= 8)" eG Oe
6) + (0)
= —0.59i — 0.45j + 0.67k
The position vector rp,, from point D to the point of application
H of F is rp, = 9k. The magnitude of the moment of F about
line DE is determined from
Lite Os i =O.) =0¢5 0.67
Mp Nl Ty ly) — 0 0 2)
eh IRS Se =5 3 ==]
tN
=U) =02'5
=—) 3
= —9(—1.77 — 2.25) = 36.2 kip - ft
Expressed in the vector form,

Mor = Mpepr = 36.2(—0.59i — 0.45j + 0.67k)


Mp- = —21.4i — 16.3j + 24.2k (kip: ft) <
EXAMPLE 4.4

A uniform platform weighing 8 KN is supported by a cable as


shown. The tension in the cable produces the same moment about
the hinged axis AB as that caused by the weight of the platform,
but with an opposite sense. Determine the tension in the cable if
a = 90°.

Solution A coordinate system with origin at point A is chosen


as shown.

Moment of weight about axis AB. We can write the weight of


the platform as W = — 8j (KN). The position vector r4g
from point A on the axis AB to the center of gravity (midpoint)
G of the platform is r4g = 1.0i + 1.5k (m). The unit vector
n along the axis AB is given by n = k. The magnitude of
the moment of W about axis AB is determined from

nn, MN, 0 @ il 1 0
Mis =) ee ald 0 1.5) =1 0 4
Pee ckigs tl ) =o
= —8 (kN - m)

So we may write

Mag
= Magn = —8 k (kN
- m)

Moment of cable tension about axis AB. Since the cable


tension T is directed along line CD, it can be written as
T = 7MNg¢p, in which T is the unknown magnitude and Ngp 1s
the unit vector along line CD. Then

ie Sj ek
OS eo + (1.8)
(—2)? TES + (2)
= —0.6i + 0.54j + 0.6k

Therefore,
T = —0.67i + 0.547j + 0.67k
The position vector from point A to the point of application C
of T is r4- = 2i + 1K (m). The magnitude T of the cable
tension is determined by using the condition that the moment of
T about axis AB must be equal but opposite to that of the
weight W, or M4, due to T must be equal to + 8k (KN - m).
Thus,
0 1
0
2 0 |
+8=| 2 0 1 =1 = 1.08T
0.6T —0.6T 0.54T
—0.6T 0.54T
T=741kN <
The cable tension can be expressed in the vector form as

T = Tncp = 7.41 (—0.6i + 0.54j + 0.6k)


T = —4.441 + 4.0j + 4.44k(KN) <

176
4.5 Couples 177

4.5 COUPLES
As stated in Section 3.10, a couple is a pair of equal, opposite, and parallel
forces. Consider a three-dimensional rigid body subjected to a pair of equal
and opposite forces, F and —F, as shown in Figure 4.7. The moment of the
couple formed by F and — F is given by the cross product M = r X F, where
r is a position vector from any point on the line of action of — F to any point
on the line of action of F. As shown in the figure, the line of action of the
moment of the couple M is perpendicular to the plane containing the two forces
forming the couple. The sense of M is specified by the right-hand rule. In
Section 3.10, we also saw that the moment of a couple is independent of the
point about which the moment is evaluated. Therefore, the moment of a couple
may be treated as a free vector and may be applied at any point on the body,
provided that its magnitude, direction, and sense are maintained.

Components and Resultants of Couples


Moment vectors of couples obey the law of vector addition. When more than
one couple is acting on a body, their moment vectors may be moved to a
common point and added vectorially to obtain the moment vector of the single
resultant couple, which has the same external effect on the rigid body. Con-
versely, the moment vector of a couple may be resolved into component vec-
tors representing moments of the components of the given couple.
In the case of a two-dimensional body, since all the forces lie in a single

Figure 4.7
178 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

plane, the moment vectors of all the couples are oriented in the same direc-
tion—perpendicular to the plane of the forces and the body. Therefore, the
moment vector of the resultant couple is always directed perpendicular to the
plane of the two-dimensional body, and its magnitude and sense can be de-
termined simply by algebraically summing the magnitudes (with appropriate
plus or minus signs indicating their senses) of the moments of the individual
couples. However, in the case of a three-dimensional system, the forces form-
ing the couples are not restricted to the same plane or parallel planes; hence,
the moment vectors of the couples are not restricted to a single direction, and
must be added vectorially to obtain the moment of the resultant couple.
Consider the three couples acting on a body as shown in Figure 4.8a. The
three moment vectors representing these couples are shown in Figure 4.8b.
Note that each moment vector is perpendicular to the plane containing the
forces forming the corresponding couple. The sense of each moment vector is
given by the right-hand rule. Since the moment vectors are free vectors, they
may be moved to any point on the body. In Figure 4.8c, the three moment
vectors are applied at a common point O on the body; they are added vecto-
rially in Figure 4.8d to obtain the moment vector of the resultant couple,
M=M,+M,+M;,
The moment of the single resultant couple M may then be resolved into com-
ponents in the x, y, and z directions if so desired, as in Figure 4.8e. All five
systems of couples shown in Figures 4.8a to 4.8e have the same external effect
on the rigid body.

(c)

a M=M,+M,+M,
(d)

Figure 4.8
4.5 Couples 179

EXAMPLE 4.5

Determine the moment of the couple shown.

Solution With the origin of the coordinate system at point


O as shown, the coordinates of the points of application A and
B, of the 250-lb forces, are

x4 = 12 + 12 (cos 30°) = 22.4 in.


y4 = —12 (sin 30°) = —6 in.
Z4
=0
and

Xp = 6 in. yp = 0 Zz = 30in.

Therefore, the position vectors from O to points A and B are


given by
r, = 22.4i— 6j(in.) sry = 61 + 30K (in.) 12in 38
“S/F = —250 k (Ib)
The position vector r,, from point A to B is ‘A
Tag =Vg — 1, = (6 — 22.4)i + (0 + 6)j + (30 — O)kK
= —16.4i + 6j + 30k (in.)
The 250-lb force F can be written in the Cartesian vector form =
as F = 250k (lb). The moment of the couple, formed by the Ve
two 250-Ib forces, is determined by summing moments Zr = 25 0ki(b)
about A, thus

M = rap X F = (— 16.41 + 6j + 30k)X (250k) y


M= 1500i+ 4100j (b:in)
= 4100j + 1500i
M = 1500i + 4100j (Ib- in.) <
The magnitude of M is
M = V(1500)? + (4100)? = 4366lb- in.

Alternate Solution The moment of the couple may also be


determined by summing the moments of the two forces about
any convenient point. It is not necessary that this point lie
on the line of action of one of the two forces. For example, by
summing moments about the origin of the coordinate system O,
we determine the moment of the couple:

M=r,X -—F+r,XF
= (22.4i — 6j) X (— 250k) + (6i + 30k) X (250k)
= 5600j + 1500i — 1500j
M = 1500i + 4100j (Ib- in.) <
180 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

EXAMPLE 4.6

Determine the resultant of the three couples shown.

Solution A coordinate system with origin at point O is chosen


as shown.

Moment of the couple formed by 500-N forces. The position


vector from the point of application A of —F, to the point
of application D of F, is given by

P 4p = 0.5i — 0.5j + 1.0k (m)


The 500-N force can be written in the Cartesian vector form as
F, = 500i (N). Therefore, the moment of the couple formed by
the two 500-N forces is

M, =Pap
X F,

saac ice? k
= 110.50 ae0 5m!
500 0 O
JO 0st a |) lose 1 en Commo
Fd (pa li Ga AG 500 0
= 500j + 250k (N - m)
The computation of M, can be somewhat simplified by
choosing point G, which is on the line of action of F,, as its
point of application, thereby eliminating the x component of the
position vector from the computation. The position vector
Tyg is
Tyg = —0.5j + 1k (m)
and the moment of the couple is

M, = rg X F, = (—0.5j + k) X (500i)
= 500j + 250k (N - m)

Moment of the couple formed by 350-N forces. We can see


that rz, = 0.3i — 0.4j + 1.0k (m) and F, = 350k (N). Thus,
the moment of the couple formed by the two 350-N forces is

fe aati eeeKk
M, = rp X F, = 0.3 -—0.4 1

0 0 350
Oe) er SPE Cate .|°3 04
0 350| ‘10 350 0 0
= — 140i — 105j(N - m)

(continued)
4.6 Resultants of Nonconcurrent Three-Dimensional Force Systems 181

EXAMPLE 4.6 (concluded)

Moment of the couple formed by 250-N forces.

Icr = —0.5i + 0.5j + 1.0K (m)


F, = 250(cos 30°)i — 250(cos 60°)j
= 216.5i — 125j (N)
i j k
M;=frerXF,=|-05 05 1
2165 1250
Oa ralbe=taye = alk =Ws 0.5
= = J +k
=i) iGO Plehs) .== 125)
= 125i + 216.5j — 45.75k (N - m)

Resultant couple. The moment of the resultant couple is Ye


M=M,+M,+M,
= (500j + 250k)
+ (— 140i — 105j) + (125i + 216.5j — 45.75k)
M = —15i + 611.5j + 204.25k(N-m) <
The magnitude of the moment of the resultant couple is
M = V(—15)? + (611.5)? + (204.25)? = 644.88N-m
Since the moment vectors of the three couples M,, M,, and
Ms, as well as the resultant couple M, are free vectors,
they may be applied at any point on the body. They are shown
to be acting at the origin O.
The moment of the resultant of the three couples can also be
computed by summing the moments of the six given forces
about any point on the body.
Z|

4.6 RESULTANTS OF NONCONCURRENT


THREE-DIMENSIONAL FORCE SYSTEMS
The procedure for determining the resultants of nonconcurrent three-dimen-
sional force systems acting on rigid bodies is essentially similar to that for the
coplanar systems presented in Section 3.12. This procedure is based on the
concept that any force F acting at a point A on a rigid body may be moved to
another point B, provided a couple, whose moment is equal to the moment
of the force F acting at the initial point A about the new point B, M =
182 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

I'p4 X F, is added on the body. This concept is illustrated graphically in Figure


4.9. Note that the moment of the couple M is perpendicular to the force F.

Reduction of a System of Forces to a Force-Couple System


Consider a general three-dimensional system of forces and couples acting on
a rigid body as shown in Figure 4.10a. This system may be reduced to an
equivalent system consisting of one resultant force at point O, and one resultant
couple. First, each force is transferred to the reference point O by adding to
the body, for each force, a couple with moment equal to the moment of the
force about O. As shown in Figure 4.10, this step produces an equivalent
system consisting of a system of forces concurrent at point O, and a system
of couples. Since the couples are free vectors, they have all been applied at
the reference point O. The concurrent forces are then added vectorially to

(a) (b) (c)


Figure 4.9

Figure 4.10
4.6 Resultants of Nonconcurrent Three-Dimensional Force Systems 183

obtain the resultant force R, with

R= SF (4.24)

Similarly, all the couples are added vectorially to obtain the resultant couple

Mpo = >M (4.25)


The resultant force-couple system at point O thus obtained is shown in Figure
4.10c. This force-couple system is equivalent to the initial system of forces
and couples given in Figure 4.10a, and will have the same external effect on
the rigid body.
It is usually convenient to determine the resultants in terms of their com-
ponents in the x, y, and z directions; thus the resultant force R is expressed as

R=R,i+R,j + Rk (4.26)
in which the components R,, R,, and R, are obtained by summing, respectively,
the x, y, and z components of all the forces acting on the body; that is,
R, = SF,, R, = XF,, and R, = YF,. Similarly, the resultant couple Mp, can
be written in terms of its rectangular components as
Mro = Maxi + May j+ Mek (4.27)

in which the components M;,, Mp,, and Mp, are obtained by summing the
moments, respectively, about the x, y, and z axes with origin at the reference
point O, of all the forces and the couples acting on the body.
It should be realized that the resultant force-couple system depends on the
choice of the reference point O. While the magnitude and direction of the
resultant force R remains the same regardless of where point O is located, its
line of action passes through O. The magnitude and direction of the resultant
couple Mp, depend on the location of point O about which the moments are
summed. The only exception to this occurs when the resultant force is zero
(R = 0), but the resultant couple is not zero. The resultant couple Mgy of
such a system does not depend on the location of the reference point. AsMpg
is a free vector, it may act at any point on the body.

Reduction of a Resultant Force-Couple System


to a Single Resultant Force
The procedure described in the preceding paragraphs can be used to reduce
any general three-dimensional system of forces and couples to an equivalent
one force—one couple system acting at any arbitrary point on the rigid body.
However, unlike the case of the coplanar force systems considered in Chapter
3 in which the resultant force and resultant couple are always perpendicular,
the resultant force R and the resultant couple Mgv of the three-dimensional
184 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

force systems in general are not always perpendicular to each other. Such
resultant force-couple systems cannot be further reduced to single resultant
forces.
If, though, for a given system of forces and couples, the resultant force R
and the resultant couple Mr,v are perpendicular, the system can be reduced to
a single resultant force. Recalling from Section 4.3 that the dot product of two
perpendicular vectors is zero, we can alternately state that the systems of forces
and couples for which the condition
R-Mpyo = 0 (4.28)
is satisfied can be reduced to single resultant forces, provided that R is not
zero. (When R = 0, the resultant of the given system of forces is the single
resultant couple Mzv.)
When the resultant force R and the resultant couple Mv of a given system
of forces and couples (Figure 4.11a@) are perpendicular to each other as shown
in Figure 4.11b, they may be reduced to the single resultant force using the
procedure described in Section 3.11. The resultant force R is moved from its
point of application O to a new point A so that its moment about O while
acting at the new point A is equal to the moment of the resultant couple Mav.
As shown in Figure 4.11c, this can be achieved by moving the resultant force
R along a line, perpendicular to the lines of action of R and Mgp, a distance
d = Mpo/R. Note that the point of application A of the single resultant R
must be located so that its moment about O has the same sense as that of Mzo.
The single resultant force R acting at point A (Figure 4.11c) is equivalent
to the initial system of forces and couples (Figure 4.11a), and hence will have
the same external effect on the rigid body. The equation of the line of action
of the single resultant R, passing through point A, can be obtained from the
relationship
rX R= Mego (4.29)

Ms
(a) (b) (c) (d)
Figure 4.11
4.6 Resultants of Nonconcurrent Three-Dimensional Force Systems 185

in which r = xi + yj + zk is the position vector from O to an arbitrary point


B on the new line of action of R, and x, y, and z are the coordinates of B with
respect to the x-y-z coordinate system with origin at O as shown in Figure
4.11d.

Simple Forms of Three-Dimensional Force Systems


Two types of three-dimensional force systems can always be reduced to a
single resultant force:
1. When all the forces acting on a body are concurrent at the same point, the
sum of moments about the point of concurrency is zero. The resultant force
R, acting at the point of concurrency, may be directly calculated by using the
procedures described in Chapter 2.
2. When the lines of action of all the forces acting on a body are parallel to
each other, as shown in Figure 4.12a, the resultant force R is parallel to the
forces, with its magnitude equal to the algebraic sum of the magnitudes of the
forces. As all the forces are parallel, their moments about any point O lie in
the same plane, which is perpendicular to the lines of action of the forces. The
moment of the resultant couple Mago, therefore, also lies in that plane, perpen-
dicular to the forces as well as R. Thus, the resultant force R and the resultant
couple Mz, are perpendicular to each other (Figure 4.125), and can be reduced
to a single resultant force using the procedure described previously.

(b)
Figure 4.12

Reduction of a Resultant Force-Couple System


to a Resultant Wrench
A system of nonperpendicular resultant force R and resultant couple Mgo
cannot be reduced to a single resultant force; such a resultant force-couple
system may, however, be reduced to a somewhat simpler form called the re-
sultant wrench.
Consider a nonperpendicular resultant force-couple system acting at a point
Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

O on the body as shown in Figure 4.13a. To reduce this system to the resultant
wrench, we first resolve the resultant couple Mv into rectangular components
Mp, in the direction of the resultant force R, and Ma, in the direction per-
pendicular to R as shown in Figure 4.13b. The magnitudes of the parallel and
perpendicular components of Mpv are
Mp, = Mao cos 9 Mrz = Mpo Sin8 (4.30)
When the angle 6 between R and Ma, cannot be readily determined by in-
spection, the component Mp, of Mgg in the direction of R may be obtained
by forming the dot product of Mzv with the unit vector n in the direction of
R. From Equation (4.15),

Mri = (Mgo"t)n (4.31)


The component Mz>, of Mav in the direction perpendicular to R, can then be
determined as

Mao = Mro — Mai | (4.32)

Next, the perpendicular component M,, is eliminated by moving R a dis-


tance d = M,/R along a line perpendicular to the lines of action of Mz, and
R (Figure 4.13c). The new point of application A of R is located so that its
moment about O while acting at A is equal to Mz>. Since Mp, is a free vector,
it may be moved from O to A, which is the point of application of R, as shown
in Figure 4.13d. This collinear resultant force-couple system is the resultant
wrench. The effect of the resultant wrench on the body is similar to the com-
bined thrust-and-twist effect of a wrench, a screwdriver, or a drill.

(a) (b) (c) (d)


Figure 4.13
4.6 Resultants of Nonconcurrent Three-Dimensional Force Systems 187

EXAMPLE 4.7

A pipe is subjected to three vertical forces. Determine the result- af oft


ant force-couple system at the support. : _
100 1b 75 |b

200 Ib ;
3 ft

Solution A mutually perpendicular right-handed x-y-z


coordinate system with origin at the support point O is chosen ;
as shown. The rectangular components of the three given forces i me!
can be written as
F, = — 100j (1b) F, = 75j (1b) F, = —200j (1b)
The resultant force R is given by
R= XF=F,+F,+F;
R = — 100j + 75j — 200j = —225j (1b)
The moment of the resultant couple is given by the sum of vA
moments of the three forces about O. Thus, Z
r,=4i+3k
Mro = 2M => XF) =r, XF, +r, X F, +1; XF;
in which r,, r>, and r; are the position vectors from the
reference point O to the lines of action of F,, F,, and F3,
respectively. As shown, these position vectors are
r,=2i(ft) 4r,=4i(ft) 41; =4i+ 3k (ft)
By substituting the expressions of the forces and the position
vectors into the equation of Mzo, we obtain
Mao = (21) X (— 100) + (4i) X (75j) + (4i + 3k) X (— 200)
= —200k + 300k — 800k + 600i
= 600i — 700k (Ib - ft)
Hence, the resultant force-couple system at the support point O
is
R = —225j (Ib) and Mero = 600i — 700k (Ib - ft) <
188 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

EXAMPLE 4.8

30N Determine the resultant of the system of forces and couples acting
on the 500-mm cube shown.

Solution A rectangular coordinate system with origin at point


O is chosen as shown. The origin O will also be used as the
reference point about which the resultant force-couple system
will be computed.

Resultant force.

R= 2F R = 70k — 50j + 20j + 30j — 30k — 40k


= (70— 30
— 40)k + (—50+ 20+ 30)j = 0

Resultant couple.

Maro = 2M
Maro = (0.5j) X (70k) + (0.5j + 0.25k) X (—S0j)
+ (0.5i + 0.5j + 0.5k) X (20j) + (0.5i + 0.5j)
X (30j) + (0.51 + 0.5j) X (— 30k) + (0.51) X (— 40k)
+ 30(cos 45°i — cos 45°k) + 50(—cos 45°i + cos 45°j)
= 35i+ 12.5i + 10k — 10i + 15k + 15j
= ilaylsp AO) <P Pili = PAL As == SIS),shor a= 335) Koy
= 8.35i + 70.36j + 3.79k(N - m)
Thus, the resultant of the given system of forces and couples is
y a couple of moment
M, = 8.35i + 70.36j + 3.79k(N-m) <
The magnitude of Mz is

Mp = V (8.35)? + (70.36)? + (3.79)? = 70.96N -m


Since the resultant couple is a free vector, it may be applied at
any point and is shown to be acting at the origin O.
Note: The moments of the forces about O could alternately
have been determined directly by inspection, instead of by
using the cross products as above. The plus or minus signs
should be assigned to these moments by applying the right-hand
rule. The reader is encouraged to verify the results of this
mae = 8.35i+ 70.36j +3.79k(N*m) problem by using this intuitive approach.
4.6 Resultants of Nonconcurrent Three-Dimensional Force Systems 189

EXAMPLE 4.9

Determine the resultant of the four parallel forces acting on the 10 kip
plate. 15 kip | 4 Ae
‘ , ‘ ® 0 ki Skip] 3 ft
Solution We select a coordinate system with the origin at ap
point O, which will also be used as the reference point 3 fy 3 ft
for summing moments. Since the 15-kip force passes through Z MS HAE ge 7
O, its moment will be eliminated. ca oi
The resultant force is

R=2F R = —15j — 20j — 10j — 5j = —50j (kip)

The moment of the resultant couple is determined by


summing the moments of the forces about the reference point
O.
Mero = 2M = Xr X F)
= (4i + 3k) X (—20j) + (41 — 3k) X (— 10))
+ (8i) X (—5j)
= —80k + 60i — 40k — 30i — 40k
30i — 160k (kip - ft) y
Therefore, the resultant force-couple system at point O consists
of the force R = —50j (kip) and the couple of moment R=—50j(kip)]—160k Maro= 30i —160k(k-ft)
Mero = 301i — 160k (kip - ft). ~ ie
It can be seen that the resultant force R and the resultant ee
couple Mpg are perpendicular to each other. This can be
verified by forming the dot product
R- Mao = (—50j):(30i — 160k)
= —1500(j-i) + 8000(j-k)
As j-i = Oandj-k = 0, then R* Mg = 0, which confirms
that R and Mg, are mutually perpendicular and, therefore,
may be reduced to a single resultant force.

Location of the single resultant force. The single resultant


force R must be located at a point A so that its moment about
O is equal to Mav. Let us assume that the (yet unknown)
coordinates of the point of application A of the single resultant
R are x, y, and z. Then the position vector from O to A is given
by ro, = Xi + yj + zk. Since the moment of R acting at A,

(continued)
190 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

EXAMPLE 4.9 (concluded)

about O, must be equal to Mpgv, we can write

Toa X R= Meo
(xi + yj + zk) X (—SO0j) = 30i — 160k
(—50x)k + (50z)i = 30i — 160k
By equating the i component on the left side of the above
equation to the i component on the right side, we obtain the
z coordinate of the point of application A of the single
resultant R:
50z = 30 z= 0.6 ft
Similarly, by equating the k components, we obtain
— 50x = — 160 x = 3.2 ft
Thus, the resultant force is

R = —50j (kip) at x = 3.2 ft andz=0.6ft <

The line of action of R must pass through point A whose


coordinates are x = 3.2 ft and z = 0.6 ft. Since R is a sliding
vector, it can act at any point on its line of action, which is
parallel to the y axis. Therefore, any value of the y coordinate
may be chosen for the point of application of R; as shown, R is
acting at point A, where y = 0.
Note: The coordinates of the point of application A of
resultant R could alternately be determined by eliminating the
x-component Mp, = 30i of the resultant couple Mgy by
moving R from O a distance of z = Mp,/R = 30/50 ft =
0.6 ft in the positive z direction; similarly, the z-component
Mr: = —160k could be eliminated by moving R a distance of
x = Mp./R = 160/50 = 3.2 ft in the positive x direction.
4.6 Resultants of Nonconcurrent Three-Dimensional Force Systems 191

EXAMPLE 4.10

Determine the resultant of the system of forces acting on the


bracket.

Solution A coordinate system with origin at point O is


chosen.

Rectangular components of the forces. The force F, with a


magnitude of 50 N is directed along line BC. The unit vector
along BC is

—400j + 300k
ee es = 0.8} +,0.6k
V(— 400)? + (300)
So,
F, = F\ngc = 50(—0.8j + 0.6k) = —40j + 30k (N)
The unit vector along the line of action AE of F, is

noe = — 400j 2 300k


V(300)?
+(—400)? +(—300)?
= 0.514i — 0.686j — 0.514k
So,
F, = Fyn,4,; = 70(0.514i — 0.686j — 0.514k)
= 36i — 48j — 36k (N)
Similarly, the unit vector along the line of action DO of F; is

Dpo = See rece = —0.707i — 0.707k

and
F, = F3npg = 30(—0.707i — 0.707k) = —21.2i — 21.2k (N)

Resultant force. R = XF = F, + F, + F;; thus


R = (— 40j + 30k) + (36i — 48j — 36k) + (—21.2i — 21.2k)
= 14.8i — 88j — 27.2k (N)
The magnitude of the resultant force R is

R = V (14.8)? + (—88)? + (—27.2)? = 93.3N

Resultant couple. Since the line of action of F; passes through


the reference point O, its moment about O is zero.

(continued)
192 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

EXAMPLE 4.10 (continued)

The position vectors from O to the points of application B


and E, respectively, of F, and F, are

r, = 400j (mm) and r, = 300i (mm)

The moment of the resultant couple is given by

Meo = 2M=r, XF, +r, X F,


jue jteik ie; k
=|0 400 0/+/300 0 Oo
0 -40 30 36 -48 —36
12,000i + 10,800j — 14,400k (N - mm)
y and it has a magnitude of

Mro = V (12,000)? + (10,800)? + (— 14,400)?


= 21,633.3 (N - mm)
Thus, the resultant force-couple system at point O consists of
the force R = 14.8i — 88j — 27.2k (N) and the couple of
moment Mey = 12,000i + 10,800j — 14,400k (N - mm).
Checking R and Mp, for perpendicularity,

R- Mao = (14.81 — 88j — 27.2k)


-(12,000i + 10,800j — 14,400k)
= 177,600 — 950,400 + 391,680 = — 381,120
As the dot product R - Mg, is not equal to zero, the resultant
force R and the resultant couple Mv are not perpendicular
and, therefore, cannot be reduced to a single resultant force.

Resultant wrench. To reduce the resultant force-couple system


at O to the resultant wrench, we determine the rectangular
components of Mp, in the direction of R and in the direction
perpendicular to R. The unit vector n in the direction of R is

ts
Rie!
gaa
: ;
88j — 27.2k)

= 0.1591 — 0.943j — 0.292k


The magnitude of the component Ma, in the direction of R is
determined from the dot product Mpy +n
Mp, = Mgo'n
(12,000i + 10,800j — 14,400k)
-(0.159i — 0.943j — 0.292k)
— 4071.6(N - mm)
(continued)
4.6 Resultants of Nonconcurrent Three-Dimensional Force Systems 193

EXAMPLE 4.10 (concluded)

The negative value indicates that the sense of Mp, is opposite


to that of the resultant force R. Written in the vector form,

Me = Mpn = — 4071.6 (0.159i — 0.943j — 0.292k)


— 647.41 + 3839.5j + 1188.9k (N - mm)
The component of Mag in the direction perpendicular to R is

Mp2 = Mro — Mari


= (12,000i + 10,800j — 14,400k)
— (— 647.41 + 3839.5j + 1188.9k)
= 12,647.41 + 6960.5j — 15,588.9k (N - mm)

The resultant wrench consisting of the resultant force R and


the resultant couple Mp,, which is collinear with R, must be
located at a point so that the moment of R about O is equal to
Ma,>. Assuming that the line of action of the resultant wrench
intersects the xz plane at point G with the coordinates x and
z as shown, we can write the position vector from O to G yy
as fog = xi + zk. As the moment of R acting at G about O
must be equal to Mp5, we write
Tog X R= Mp>
i j k
DG 0 Z = 12,647.41 + 6960.5j — 15,588.9k
See SON De Mp,= —647.41 + 3839.5j + 1188.9k(N-mm)

Expanding the determinant,


(88z)i + (27.2x + 14.8z)j + (— 88x)k
= 12,647.41 + 6960.5j — 15,588.9k v4 :
Equating the i, j, and k components, we obtain ; R= 14.81 —88j —27.2K(N)
88z = 12,647.4
27.2x + 14.82 = 6960.5
— 88x = —15,588.9
From these equations, we determine the coordinates of point G
of the resultant wrench as
and a collinear resultant couple,
x = 177 (mm) and z = 144 (mm)
Mp, = — 647.4i + 3839.5j + 1188.9k (N-mm) <
Thus, the resultant wrench, which is equivalent to the given
system of forces, consists of a resultant force, The point of application of the resultant wrench is located at
R = 14.81 — 88j — 27.2k(N) < x = 177 (mm) and
194 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

EQUILIBRIUM IN THREE DIMENSIONS

4.7 EQUILIBRIUM OF RIGID BODIES


A rigid body is in equilibrium if the resultant force-couple system of the forces
and couples acting on it is zero (Section 3.13), or

R=SF=0 M,=>M=0 (4.33)

These vector equations represent the necessary and sufficient conditions for
equilibrium of rigid bodies. The first equilibrium equation, R = =F = 0,
states that the vector sum of all the forces acting on the body is zero; the
second, M, = =M = 0, states that the vector sum of the moments of all the
forces about any point, plus the moments of any couples acting on the body,
is zero.
If the forces and the couples acting on the body are expressed in terms of
their rectangular components, then we can rewrite Equations (4.33) as

R=2F=2(@it+Fj+hkK)=2Fi+ 2h j+2ek=0
Mz = 2M = 3(M,i + M,j + M,k) = 2M,i + 2M,j + 2k =0

From these we can see that in order for R and M, to equal zero, each of the
six summations, 2F,, 2F,, &F,, 2M,, 2M,, and {M., must individually be
equal to zero. Therefore, the necessary and sufficient conditions for the equi-
librium of rigid bodies in three dimensions can be expressed by the six scalar
equations of equilibrium:

2) fs =0 25, = 0 LF, =0 (4.34)

>I) ees ey a igeaiy Kies

The first three equations indicate, respectively, that the algebraic sums of the
x, y, and z components of all the forces acting on the body in equilibrium must
be zero, while the last three equations indicate, respectively, that the algebraic
sums of the moments of all the forces and couples about the x, y, and z axes
must be zero. All six must be satisfied for the body to be in equilibrium.

4.8 TYPES OF SUPPORTS AND CONNECTIONS


The common types of supporting devices used to restrict movements of bodies
subjected to three-dimensional forces are shown in Table 4.1. These supports
are Classified into six categories according to the number of reactions they can
exert on the body.
4.6 Resultants of Nonconcurrent Three-Dimensional Force Systems 195

Category | The supports in this category prevent translation in only one


direction. Therefore, they exert one reaction force on the body, in the direction
in which the translation is prevented. Since the direction of the reaction force
is known, the only unknown is its magnitude, which is determined by applying
the equations of equilibrium of the rigid body.

Category Il The roller support on a guide in this category allows rotation


about an axis in any direction, but permits translation only in the direction of
the guide. Translations are prevented in the other two directions. Therefore,
this support exerts two reaction force components of unknown magnitudes on
the body.

Category Ill The two types of supports in this category prevent translation in
any direction but cannot prevent rotation. Therefore, they exert a reaction force
of unknown magnitude and direction on the body. It is usually convenient to
represent the reaction force by its rectangular components, with the magnitudes
of the three components as unknowns, in the equilibrium equations of the rigid
body.

Category IV The bearing support in this category permits translation in the


axial direction, and rotation about the axis of the shaft. Translations in the
radial directions and rotations about the axes perpendicular to the axis of the
shaft are prevented. Therefore, this type of support can exert two force com-
ponents and two moment components on the shaft. All four reactions act in a
plane perpendicular to the axis of the shaft. In many applications, bearing
supports are designed to support only radial forces but no couples. Such bear-
ing supports usually exert only small moments on the body, the effect of which
may be neglected in the analysis.

Category V The supports in this category are similar to the bearing support
of Category IV, except that these do not permit translation in the axial direc-
tion. Therefore, they can exert one additional reaction force component on the
body, in the axial direction.

Category VI The fixed support prevents translation as well as rotation in any


direction. Therefore, it can exert a reaction force and a reaction couple; the
magnitudes and directions of both are unknown. The reactions are usually
represented by their rectangular components; the six unknowns—the magni-
tudes of three force components and three moment components—are deter-
mined from the equations of equilibrium of the rigid body.
Table 4.1 Supports and Connections for Bodies Subjected to Three-Dimensional Forces

Category Types of supports Types of reactions Number of unknowns

One
The reaction force is directed normal to the supporting
surface and may act either into or away from
the body. The magnitude of the reaction force is the
unknown.

One
The reaction force is in the direction of the link and
may act either into or away from the body.
The magnitude of the reaction force is the unknown.

One

The reaction force acts away from the body in the


direction of the cable. The magnitude of the reaction
force is the unknown.

One
The reaction force acts away from the surface (into
the body) in the direction normal to the surface
The magnitude of the reaction force is the unknown.

Two

Two reaction force components act in a plane perpen-


dicular to the guide. Both reactions may act
either into or away from the body. The magnitudes
of the two components are the two unknowns.

Roller on guide

wy, Three
|
| The reaction is a force of unknown magnitude and
——x direction, which is conveniently represented by its
rectangular components. The magnitudes of the
a three components are the three unknowns.
Ball and socket

Three

The reaction is a force of unknown magnitude and direction,


Illa which acts away from the surface (into the body). The
reaction force is usually represented by its rectangular
components, acting into the body. The magnitudes of the
three components are the three unknowns.

Four

Two reaction force components and two reaction couples


IV act in a plane perpendicular to the axis of the shaft. The
magnitudes of the two force components and the two
moment components are the four unknowns.

Journal bearing
(continued)
196
Statically Determinate Structures 197

Table 4.1 (concluded)

Five

The reactions consist of three force components and two


couples. The moments of the couples act in a plane
perpendicular to the axis of the shaft or the hinge. The
magnitudes of the three force components and the two
moment components are the five unknowns.

Six

The reactions consist of a force and a couple, both of


unknown magnitude and direction. The reaction force
VI eee) — — X and couple are usually represented by their rectangular
components. The magnitudes of the three force
components and the three moment components are
the six unknowns.

Fixed support

4.9 STATICALLY DETERMINATE STRUCTURES


In Section 3.17, we examined the number and arrangement of support reactions
necessary for the equilibrium of rigid bodies subjected to coplanar forces. In
this section, we will extend those concepts to rigid bodies subjected to three-
dimensional forces.

Static Determinacy, Indeterminacy, and Instability


As we know, a body is considered to be statically determinate externally if all
of its support reactions can be determined by solving the equations of equilib-
rium. Since there are six equations of equilibrium for rigid bodies subjected
to three-dimensional forces, they cannot be used to determine more than six
reactions. Also, in order to satisfy the six equilibrium equations to ensure that
the body remains in equilibrium under the action of a general three-
dimensional force system, the body must be supported by at least six reactions.
An example of an externally statically determinate structure is shown in Figure
4.14a. The structure is supported by a short bearing support at A, exerting two
force components but no couples; a ball support at B exerting one force com-
198 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

(6) Externally statically indeterminate structure


Figure 4.14

ponent; and a ball-and-socket support at C, which exerts three force compo-


nents. As the structure is supported by six reactions that can be determined by
solving the six equations of equilibrium, it is statically determinate externally.
If a body has more than six support reactions, all of the reactions cannot
be determined using the equilibrium equations (although some of them may
be). Such a body is considered statically indeterminate externally. An example
of an externally statically indeterminate structure is shown in Figure 4.145.
This structure is similar to the statically determinate structure of Figure 4.14a
except that the bearing support at A is replaced by a ball-and-socket support.
The total number of support reactions is seven, one more than the six necessary
for equilibrium. (The analysis of statically indeterminate structures is beyond
the scope of this textbook.)
If a body has less than six external support reactions, the reactions are not
sufficient to prevent all possible movements of the body in space. Such a body
cannot remain in equilibrium under a general system of three-dimensional
forces and, therefore, is considered statically unstable externally or partially
constrained. As an example, the structure shown in Figure 4.15q is similar to
the one in Figure 4.14 except that the end A is now attached to a ball support.
As the total number of support reactions is now five, one less than the six
necessary for equilibrium, the structure is only partially constrained and con-
4.9 Statically Determinate Structures

(a) Partially constrained body not in equilibrium

y
|
|
CG |
27

(b) Partially constrained body in equilibrium


Figure 4.15

sidered statically unstable. It should be clear that under the action of the hor-
izontal applied load P, the structure cannot remain in equilibrium and will
rotate about the y axis passing through point C.
Even though statically unstable or partially constrained bodies cannot re-
main in equilibrium under any arbitrary system of forces and couples, such
bodies can be in equilibrium under certain systems of applied forces and cou-
ples, which satisfy by themselves the equilibrium equations that cannot be
satisfied due to the lack of support reactions. For example, the structure just
considered would be in equilibrium if the applied load P acted in the vertical
direction as shown in Figure 4.15b.
The conditions of static instability, determinacy, and indeterminacy of
three-dimensional rigid bodies supported by r number of reactions can be
written in the following form:

r<6_ bodyis statically unstable externally


r=6_ bodyisstatically determinate externally (4.35)
r>6_ body is statically indeterminate externally

Geometric Instability
The first condition given above, r < 6, is both necessary and sufficient
for static instability: when r < 6, the body is definitely unstable or partially
200 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

Figure 4.16 Improperly constrained bodies

constrained. However, the remaining two conditions for static determinacy


(r = 6) and indeterminacy (r > 6) are necessary, but not sufficient, conditions.
These two conditions tell us that the number of reactions is sufficient for
stability; but they provide no information regarding the arrangement of the
reactions. As we discussed in Section 3.17, a body may be supported by a
sufficient number of reactions but may still be unstable (i.e., not completely
restrained or supported to prevent any possible movement) due to improper
arrangement of supports. Such bodies are considered geometrically unstable
externally, or improperly constrained.
Two examples of improper arrangements of support reactions are shown in
Figure 4.16. The platform in Figure 4.16a is supported by six parallel cables.
Even though there are a sufficient number of reactions, , = 6, they are all in
the vertical direction (parallel to each other) and cannot prevent translation of
the platform in the horizontal direction. The structure is therefore geometrically
unstable, or improperly constrained.
In the case of Figure 4.165, a beam is supported by two ball-and-socket
supports at ends A and B. Thus, there are a sufficient number of reactions
(r = 6). However, since the lines of action of all six reactions intersect a
common axis AB, they cannot prevent rotation of the beam about axis AB: the
beam is geometrically unstable or improperly constrained.
4.11 Procedure for Analysis of Reactions 201

Like partially constrained (statically unstable) bodies, improperly con-


strained (geometrically unstable) bodies can be in equilibrium under certain
systems of applied forces. For example, the beam of Figure 4.16b would be
in equilibrium if the line of action of load P intersected axis AB. In the re-
mainder of this chapter, we will consider the analysis of reactions of properly
constrained, statically determinate bodies, and of the partially constrained
bodies whose support reactions can be determined using the equations of
equilibrium.

4.10 FREE—BODY DIAGRAMS


As in the case of two-dimensional analysis, the construction of the free-body
diagram is the most important step in the analysis of equilibrium for three-
dimensional bodies. Recall from Section 3.18 that the free-body diagram is a
diagram of the rigid body isolated or freed from its supports, and showing all
the external (applied, as well as reaction) forces acting on it.
The procedure for drawing free-body diagrams of three-dimensional rigid
bodies is essentially the same as that outlined in Section 3.18, except that a
three-dimensional right-handed x-y-z coordinate system must be used (Step 5),
and support reactions as given in Table 4.1 should now be specified (Step 6).
The student should review the step-by-step procedure in Section 3.18 before
proceeding with the analysis of reactions in the following section.

4.11 PROCEDURE FOR ANALYSIS OF REACTIONS


The following step-by-step procedure can be used to determine reactions of
rigid bodies subjected to three-dimensional systems of forces and couples.
1. Draw a free-body diagram of the body.
2. Resolve each force and couple applied to the body into its components
in the x, y, and z directions.
3. Determine unknown reactions by applying the equations of equilibrium
(ZF, = 0, 2F, = 0, 2F, = 0, 2M, = 0, 2M, = 0, and 2M, = 0).
It is usually convenient to obtain these scalar equations by first writing
the two vector equations, R = {F = 0 and Mz = =M = 0, and then
equating to zero the components of each of the two vector equations. To
avoid solving simultaneous equations, the moment center should be
chosen so that the lines of action of some of the unknown reaction
forces pass through the point, thereby eliminating those reactions from
the three moment equilibrium equations. Also, in some problems, the
solution can be simplified by summing moments about an axis passing
through two or more supports, thereby eliminating all the reaction forces
at those supports from the moment equilibrium equation.
202 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

EXAMPLE 4.11

Determine the reactions at the fixed support A for the pipe shown.

Solution The free-body diagram of the pipe shows the four


known applied forces and one known applied couple, and also
the six unknown reactions (three force components and three
moment components) exerted on the pipe by the fixed support
at A. The senses of the unknown reactions are assumed to be in
the positive x, y, and z directions. The origin of the mutually
perpendicular right-handed coordinate system is located at the
fixed support point A, which will also be used as the moment
center, thereby eliminating three unknown reaction force
components, A,, A,, and A., from the moment equilibrium
equations.
The structure is statically determinate, and properly
constrained; hence, the six unknown reactions can be determined
by applying the equations of equilibrium.

3ft Support reactions. We first apply the force equilibrium


equation }F = 0, which indicates that the vector sum of all the
—400j(1b)~3ggiib)
Se forces acting on the body must be zero. Hence,
2 ft
SF =0
2ft
—100i(1b) A,i + Aj + A,k + 150i — 400j — 100i — 300i = 0
—SOOi(1b-
ft) (A, — 250)i + (A, — 400)j + A,k = 0
Equating the i, j, and k components of the above vector
equation to zero, we obtain the following three scalar equations
of equilibrium:
=F, =0 A.-250=0 A,=2501b
A, = 250i (Ib) <
pw hen) A,-400=0 A,=4001Ib
A, = 400j(Ib) <
SF. =0 = | i=)

Next, we apply the moment equilibrium equation


<M, = 0, which indicates that the vector sum of the moments
of all the forces and couples acting on the body must be zero.
Hence,
>M, = 0
Maxi + My, + My.k + 3j X 150i + Gi + 6j) x (— 400))
— 500i + (3i + 6j + 2k) X (— 100i)
+ (3i + 6j — 2k) X (— 300i) = 0
(continued)
4.11 Procedure for Analysis of Reactions 203

EXAMPLE 4.11 (concluded)

Evaluating the cross products, we obtain

Maxi + Mayj + Mg,k — 450k — 1200k— 500i + 600k


— 200j + 1800k + 600j = 0
(M4, — 500)i + (My, + 400)j + (My, + 750)k = 0
Again, by equating to zero the i, j, and k components of the wy
above vector equation, we obtain three scalar equations of
equilibrium,
400 Ib
2M, =0 M,,
— 500 = 0 300 Ib
M,,. = 500 (lb - ft) 500 lb-ft
100 Ib
M,, = 500i (lb- ft) <
2M, =0 M,,
+ 400 = 0
May = —400 (Ib - ft) 750 lb-ft

M,, = —400j (lb-ft) < 250 |b


500 lb-ft
2M, =0 M,,
+ 750 = 0 400 Ib
M,, = —750 (Ib -ft)
M,. = —750k (Ib: ft) <

EXAMPLE 4.12

The beam shown is supported by a short bearing at A that can 0.75 m 70.75 m
exert only radial forces but no couples, a ball support at B, and a 500N
ball-and-socket support at C. Determine the reactions at the sup-
ports.

Solution As shown on the free-body diagram, the bearing


support at A exerts two force components A, and A, that are
perpendicular to the axis of the beam. The ball support at B
exerts one force component B, normal to the supporting |
OVS mie, O8/5m
surface; and the ball-and-socket support at C exerts three force
components, C,, C, and C,, on the beam. The senses of the
six unknown reactions are assumed to be in the positive x, y,
and z directions.
The origin of the x-y-z coordinate system is located at
support point C, also used as the moment center.

(continued )
204 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

EXAMPLE 4.12 (concluded)

Support reactions. Applying the force equilibrium equation,

LF =0
A,i + A,j + 300i + B,j — 500j + C,i+ C,j+Ck=0
(A, +300 + Cli (A, 4B, 500 + y+ CK =
By equating the i, j, and k components to zero, we obtain the
three scalar equilibrium equations,

2H 0
A, + 300 + C, =0 A, + C, = —300 (a)
So Oe
A,+B,—500+C,=0 A,+B,+C,=500 (b)
20
Gr=s0
Next, we write the moment equilibrium equation,

2M. =0
(=1.51 + 1k) X a+ A,j) + (1.5 + 05k) X 300i
+ (—1.5i) X B,j + (—0.75i) X (—500j) = 0
—1.5A,k + A,j — Ayi + 150j — 1.5B,k + 375k =0
=A,i + (A, 47-150)) 4 (— 1 3A, Bt sik 0
From which we obtain the three scalar moment equilibrium
equations

2M, =0 —A,=0 AU
XM, =0 A, + 150 =0 A,= —150N
2M, =0 =A al Bt oo)
Asa, = 0, we have

Bo 315 = 0 B, = 250N

Substituting the values of A,, A,, and B, into equations (a) and
(b), we obtain
A, + C, = 300 —150 + C,= -300 C,= —150N
A,+B,+C,=500 0+250+C,=500 C,=250N
The results of the analysis are shown. The total reactions at
supports A, B, and C can be written as
A= —I150i(N) <
B = 250j(N) <
C = —150i + 250j (N) <
EXAMPLE 4.13

A sign weighing 2 kip is attached to a pole. Determine the ten-


sions in the cables BE and CD and the reactions at the ball-and-
socket support at A. Neglect the weight of the pole.

Solution The free-body diagram of the pole and the sign is


shown. The origin of the coordinate system is located at
support A, also the moment center. Since the pole is supported
by only five reactions, it is only partially constrained. The
lines of action of all five reactions intersect the y axis;
therefore, they cannot prevent the rotation of the pole about the
y axis. However, since the only applied force acting on the
structure is the weight W, which does not tend to rotate
the structure about the y axis, the structure is in equilibrium.
The cable tensions Tp and Tp, are directed along the lines
CD and BE, respectively. Using the dimensions given, we can
write the unit vectors along CD and BE as

=i = IS) se ovis
DoD Ay 2
Ce re 2
US) om (5)2
= —0.302i — 0.905j + 0.302k
mes = 5 = ih) = Sk
Ves iy a5)
= —0.382i — 0.841j — 0.382k
Thus, the cable tensions can be expressed in terms of their
magnitudes as 1]
Tep = TepNcp = Tcp(— 0.302i — 0.905j + 0.302k)
Tar = TaeMge = Tpp(—0.382i — 0.841j — 0.382k)
Support reactions. Applying the force equilibrium equation,
we write

LF =0
A,i + A,j + A.kK + Tep(—0.302i — 0.905j + 0.302k)
+ Ty,(—0.382i — 0.841j — 0.382k) — 2j = 0
(A, — 0.302Tcp— 0.382Tg,)i + (Ay — 0.905Tcp
STO 84l 2 )f CA ot 0.3027 = 038275. )ki= 0
From this we obtain the three scalar force equations

ZF.=0 Au 03021 7 0.38217 = 0 (a)


SF, =0 A, — 0.905Tep — 0.841Tpp= 2 (b)
=F, — 0 fa, at 0.302T cp a 0.382T pr = 0 (c)

(continued )

205
206 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

EXAMPLE 4.13 (concluded)


“a
Next, applying the moment equilibrium equation,

2M, =90
15j X Tep(— 0.3021 — 0.905j + 0.302k)
+ 11j X Tpp(—0.382i — 0.841j — 0.382k)
+ (6i + 13j) X (—2j) =0
4.53Topk + 4.53T epi + 4.2Tgp-K — 4.2Tp-i — 12k = 0
(4.53Tcp — 4.2Tgp)i + (4.537ep + 4.2Tpp — 12)k = 0
From this we write the scalar moment equilibrium equations

2M, =0 4.53Tép — 4.2Tpp = 0 Tp = 1.08Tcp


~M,=0 4.53Tcp + 4.2T3, — 12 = 0

Substituting Tz, = 1.087


¢p into the above equation we obtain

Tcep = 1.32 kip and Tge = 1.08Tcp = 1.43 kip


Therefore, the tensions in cables CD and BE are

Tep = TepNcp = 1.32(—0.302i — 0.905j + 0.302k) kip


Tep = —0.4i — 1.19j + 0.4k (kip) <
Tepe = TeeMge = 1.43(—0.382i — 0.841j — 0.382k) kip
Tap = —0.55i — 1.2j — 0.55k (kip) <
Note that the third moment equilibrium equation, xM, = (lis
automatically satisfied. Substituting the values of T.p) and
Tp,¢ into Equations (a) through (c), we obtain the reactions at
the ball-and-socket supports at A. Thus,

Ay = 0.30215 038215 A, = 0.94 kip


A, = 0.94i (kip) <
A, = 0.905Tcp+ 0.841Tgp +2 Ay =4.4kip
A, = 4.4j (kip) <
A, = =0.302Tc, + 0.382Tse A, = 0.15 kip
A. = 0.15k (kip) <
| EXAMPLE 4.14

A steel trap door of 275-kg mass is supported by a hinge at A


and a cable BC. Determine the reactions at the supports.

Solution The free-body diagram of the door shows the two


known forces: the 3-kN applied force and the weight of the
door W = (275)(9.81)/1000 kN = 2.70 KN. In addition, the
diagram shows the five unknown reactions (three force
components and two couples) being exerted by the hinge on the
door, and one unknown cable tension.
The cable tension T is directed along line BC. The unit
vector along line BC is

SOlar oj OOK |
as ee wv), 2 2 2
(G9) amare Cle) geese ORV) Wag

= —0.873i + 0.436j + 0.218k


avich Akz T
B
ee We 2 /j(KN) 0.75 m
Therefore, the cable tension T can be written as x vl Bea 04s mn x

T = Tngc =T(— 0.8731 + 0.436j + 0.218k)


is Sion leSy son
Support reactions. We first apply the force equilibrium ' —3j (KN)
equation:

LF =0
A,i + A,j + A,k + T(—0.873i + 0.436j + 0.218k)
— 2.7) —3j = 0
(A, — 0.873T)i + (A, + 0.436T — 2.7 - 3)j
+ (A, + 0.218T)k = 0
Thus, the three scalar force equilibrium equations are

D0 A, — 0.873T =0 (a)
XF, =0 A, + 0.436T = 5.7 (b)
p= 0 A, + 0.218T =0 (c)
Next, we apply the moment equilibrium equation:

=M, = 0
Mala M ay 1 lot xX (— 2.79) + Gi 0:75k)
X (—3j) + Gi — 0.75k) X T(—0.873i + 0.436j + 0.218k)
=0
Maxi + Mayj — 4.05k — 9k + 2.251 + 1.317TKk — 0.657j
+ 0.65Tj + 0.33Ti = 0
(My, + 2.25 + 0.33T)i + (May — 0.65T + 0.657)j
+ (—4.05 — 9 + 1.31T)k = 0

(continued)

207
208 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

EXAMPLE 4.14 (concluded)

From this we obtain

M,. = 0 My 0330 ——— 2-25 (d)


2M, =0 My M,,=0 <
>M, =0 —13.05+ 1.31T=0 T=9.96kN
Therefore,

T = 9.96( —0.873i + 0.436j + 0.218k)


T= —8.7i + 4.34j + 2.17k (KN) <
By substituting the value of T into Equations (a) through (d),
we determine the remaining reactions as
A, = 0.873T A, = 8.70 kN
A, =8.70i (KN) <
A, = 5.7 — 0.4367 A, = 1.36 kN |
A, = 1.36j (KN) <
A, = —0.218T A, = —2.17KN
A, = —2.17k (KN) <
My, = —2.25-0.33T My, = —5.54kN-m
3 kN M,, = —5.54i(KN-m) <

EXAMPLE 4.15

The frame shown is supported by a ball-and-socket support at A,


a bearing at B that can exert only radial forces but no couples,
and a cable CD. Determine the tension in the cable CD.

Solution The tension T in cable CD can be directly determined


by summing the moments of the two known forces of
magnitudes 400 Ib and 500 Ib, and the cable tension T, about
the axis AB, and equating the sum to zero. Since line AB passes
se
through the supports at points A and B, all the reaction forces
4001b / |W
acting at those two supports will be eliminated from the
Bees) moment equilibrium equation, £M,, = 0, which will contain
only one unknown, 7. The unit vector n in the direction of line
AB is

= soe i) = abs
es a NII i (OI = OOO
V(—3.5)? + (1)? + (—3)?
(continued)
4.12 Summary 209

EXAMPLE 4.15 (concluded)

Recalling from Section 4.4 that the magnitude of the moment 4


of a force about an axis is given by the dot product of the unit —500j(1b) T
vector in the direction of the axis and the moment of the
force about any point on the axis, we can write

>M,, = n->M, = 0
in which }M, represents the sum of the moments of the two
known forces and the unknown cable tension about point A.
Then —400i(1b)

=M, = 1j X (— 400i) + (—3.5i — 6k) X (—500)) 1 ft ;


+ (—6k) X Tj
+ 400k + 1750k — 30001 + 67i ‘
= (6T — 3000)i + 2150k
Thus, we write the moment equilibrium equation as

Map =n°>M, = 0
(—0.742i + 0.212j — 0.636k)- {(6T — 3000)i + 2150k} = 0
— 0.742(6T — 3000) — 0.636(2150) = 0
—4.45T + 2226 — 1367.4 =0
T = 192.94 lb T = 192.94j (1b) <

4.12 SUMMARY
In this chapter, we have learned the following:

1. The moment of a force F about a point A on a three-dimensional body is


a vector defined as (Section 4.2)

i j k
M,=rxF=|(p—4x,) (yp—ya) Gp
—24) (4.6)
F. F, F,
2. The moment of a force about an axis is defined as the component along
the axis of the moment of the force about any point on the axis. The
moment of a force F about an axis ab can be expressed as (Section 4.4)

een,
M,=u-M,=n'(XFP)=|, 4 7 (4.22)
ie, ee

in which n is the unit vector along the axis ab.


210 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

Se The moment of a couple of equal, opposite, and parallel forces F and


—F is a vector defined as the cross product M = r X F (Section 4.5).
A system of forces acting on a rigid body may be reduced to an
equivalent one force—one couple system at a point by moving each force
to the point, by adding a couple with moment equal to the moment of
the force about the point, and then vectorially adding all the forces and
the couples, respectively, to obtain the resultant force and the resultant
couple (Section 4.6).
When the resultant force and the resultant couple of a system of forces
are mutually perpendicular (R - Mz = 0), they may be reduced to a
single resultant force by moving the resultant force to a point so that its
moment about the initial point, while acting at the new point, is equal
to the moment of the resultant couple. If the resultant force and the
resultant couple are not perpendicular, they cannot be reduced to
a single resultant force. They can, however, be reduced to a collinear
resultant force-couple system called the resultant wrench (Section 4.6).
A rigid body is in equilibrium if the resultant force-couple system is
zero (Section 4.7); that is,

R=SF=0 M,= >M=0 (4.33)


These equations are called the vector equations of equilibrium for rigid
bodies. In scalar form they can be expressed as

Ea LF, =0 S10
SM,=0 M,=0 %M.=0 Gee
The characteristics of common types of supports used to restrict or
prevent movements of bodies in three-dimensional space are summarized
in Table 4.1 (Section 4.8).
A body is considered statically determinate externally if all of its
reactions can be determined by solving the equations of equilibrium. If a
body is supported by 7 number of reactions, then if (Section 4.9)

r<6_ bodyis statically unstable externally


(partially constrained)
r=6_ bodyis statically determinate externally (4.35)
r>6_ bodyis statically indeterminate externally

Even when a body is supported by a sufficient number of reactions (r = 6), if


all reactions are parallel or intersect a common axis, they cannot maintain the
body in equilibrium under a general three-dimensional system of forces and
couples. Such a body is called geometrically unstable or improperly con-
strained (Section 4.9).
Problems 211

KEY TERMS

dot product 170 resultant force (three-dimensional


equations of equilibrium for three- force systems) 183
dimensional rigid bodies /72 resultant force-couple system (three-
mixed triple product 172 dimensional force systems) 183

moment of a force about an resultant wrench 185


axis 173
resultant couple (three-dimensional
force systems) 183
PROBLEMS

SECTIONS 4.2 AND 4.3


4.1 through 4.4 Determine the moment of the force about point A.

Figure P4.1

y
4 = —12i+9j —15K(N)
F = 91+ 5j + 6k(Ib)

eopitenleontt
& 300 mm
Figure P4.2 Figure P4.3 and P4.5 Figure P4.4 and P4.6
212 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

4.5 and 4.6 Determine the moment of the force about point B.

4.7 A transmission tower is supported by three cables as shown. The tension in


cable BF is 700 lb. Determine (a) the moment of the force exerted by the cable BF
at B about the base point A, and (b) the perpendicular distance from point A to
cable BF.

4.8 A 3-m boom is supported by two cables attached to the vertical wall as shown.
If the tension in cable CE is 500 N, determine the moment of the cable force
exerted at C about point A.

4.9 For the boom in Problem 4.8, if the tension in cable BD is 300 N, determine
the moment of the cable force exerted at B about point A.

4.10 In Problem 4.8, determine the perpendicular distance from point B to


cable CE.

4.11 A platform is supported by two cables attached to the vertical wall as shown.
Figure P4.7
The tensions in cables PQ and ST are 1500 Ib and 2000 Ib, respectively. Determine
the total moment due to both cable forces exerted on the platform about point D
if a = 65°.

Figure P4.8, P4.9, and P4.10 Figure P4.11 and P4.12

4.12 For the platform in Problem 4.11, determine the perpendicular distance from
point B to cable ST.
Problems 213

SECTION 4.4
4.13 A beam is subjected to two forces as shown. Determine the total moment due
to both forces about line AB.

4.14 A force is being applied to the handle of a wrench as shown. Determine the
moment of the force about line AB if F = 75 lb, 8 = 0°, anda = 90°.

4.15 Solve Problem 4.14, if 6 = 15° and a = 30°.

y
7kKN

D
Ae
A B 2m
3
3 kN lft
2m
C
4 m—»><——-4 mn»
Zz
Figure P4.13 Figure P4.14 and P4.15

4.16 The space frame shown is subjected to a force F = —3i — 2j + 4k (KN).


Determine the moment of the force about the x, y, and z axes.

4.17 For the frame in Problem 4.16, determine the moment of the force about line
CE.

Figure P4.16 and P4.17


214 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

4.18 A platform is supported by two cables attached to the vertical wall as shown.
The tensions in cables PQ and ST are 900 Ib and 1200 Ib, respectively. If a = 90°,
determine (a) the total moment due to both cable forces about line AC, and (b) the
perpendicular distance between the cable PQ and the line AC.

4.19 For the platform in Problem 4.18, determine the moment of the force in cable
PQ about line AD.

4.20 Solve Problem 4.19 if a = 50°.

4.21 A uniform plate of 51-kg mass is supported by a cable attached to the vertical
wall as shown. The tension in the cable produces the same moment about the
hinged axis AB as that caused by the weight of the plate, but with opposite sense.
Determine the tension in the cable if a = 900 mm.

4.22 If the maximum allowable tension in the cable supporting the plate in Problem
4.21 is 300 N, determine the smallest distance a so that the cable tension does not
exceed the allowable limit.

Figure P4.18, P4.19, and P4.20 Figure P4.21 and P4.22

SECTION 4.5
4.23 Determine the moment of the couple acting on the beam.

4.24 Determine the moment of the couple acting on the plate if 8 = 90°.

4.25 Solve Problem 4.24 if 8 = 135°.

4.26 and 4.27 Determine the resultant of the couples acting on the structure.
Problems 215

3ft 3ft
150 1b /
150lb Jo { :
2 ft Ez B

Figure P4.23

3 m—*
z

Figure P4.27

M, =131b-ft Dei, Sib, Shite

Figure P4.28

4.28 Determine the resultant of the three couples acting on the block. The moment
vectors of the three couples M,, M,, and M; are perpendicular to the planes
ABFE, ABCD, and DCGH, respectively.
216 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

SECTION 4.6
4.29 and 4.30 Replace the given force by an equivalent force-couple system at
point A.

Figure P4.29 Figure P4.30

4.31 A concrete column is subjected to a 50-KN axial load that is eccentric with
respect to its axis of symmetry through the center of gravity, G. Replace the force by
an equivalent force-couple system at G.

4.32 A force is being applied to the handle of a wrench as shown. Replace the
force by an equivalent force-couple system at point A. Use F = 60 lb, @ = 0°, and
a = 45°.

Concrete 4.33 Solve Problem 4.32 if 8 = 60°.


column

Figure P4.32 and P4.33


Figure P4.31

4.34 The space frame shown is subjected to a force F = 5i — 7j — 3k (KN).


Replace the force by an equivalent force-couple system at point A.
Problems 217

Figure P4.34

4.35 through 4.37 Determine the resultant force-couple system at point A.

Figure P4.35

8kN 12kN

Paes ls m leo im loom 15am


Zz
Figure P4.36 Figure P4.37
Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

4.38 A beam weighing W = 200 N is supported by cable BC, and is subjected to a


500-N force as shown. If the tension T in the cable is 400 N, and the distances a
and b are 200 mm and 300 mm, respectively, determine the resultant force-couple
system of the three forces at point A.

4.39 For the beam in Problem 4.38, determine the cable tension 7 and the distances
a and b so that the resultant of the three forces is a single couple. Determine the
moment of the resultant couple.

4.40 A rectangular plate is subjected to a system of parallel forces as shown. If


P = 0, determine the resultant force-couple system of the three forces at point A.

4.41 For the plate in Problem 4.40, determine the single resultant force of the three
parallel forces. Specify the point of application of the resultant force on the plate.

4.42 For the plate in Problem 4.40, determine the magnitude and the point of
application of the load P on the left edge of the plate, so that the single resultant
force of the four forces passes through the center of the plate. Determine the single
resultant force.

B
a
W
7-450 mm
| 300 mm
150mm Zz
Figure P4.38 and P4.39 Figure P4.40, P4.41, and P4.42

4.43 Solve Problem 4.42 if the plate is inclined at an angle a = 60° with respect to
the horizontal plane as shown.

Figure P4.43
Problems 219

4.44 through 4.47 Determine the resultant force-couple system at point A.


4.48 through 4.51 Can the given system of forces (including couples) be reduced to
a single resultant force? If yes, determine the resultant force; if no, determine the
resultant wrench. In either case, specify the point where the line of action of the
resultant intersects the xz plane.

Zz 500 lb-in.
Figure P4.44 and P4.48 Figure P4.45 and P4.49

Figure P4.46 and P4.50 Figure P4.47 and P4.51


220 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

SECTION 4.11
4.52 and 4.53 Determine the reactions at the fixed support A for the pipe.

4.54 through 4.56 Determine the reactions at the fixed support A for the cantilever
frame.

Figure P4.53

Figure P4.54 Figure P4.55 Figure P4.56

4.57 Determine the reactions at the supports for the beam. The beam is supported
by a ball-and-socket support at A and a roller on a guide at B as shown in the figure.
y
8kN 12kN

vA Mesyrnldsyirn
lasyiert svi) Hesyinn
ez,
Figure P4.57
Problems 221

4.58 A uniform boom of 30-kg mass is supported by a ball-and-socket support at A


and two cables BD and CE, attached to the vertical wall as shown. Determine the
tensions in the cables and the reactions at the ball-and-socket support at A.

Zz

Figure P4.58

4.59 Determine the reactions at support A for the piping system. Assume support A
as a fixed support.

500 Ib-in.
Figure P4.59
222 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

4.60 A uniform rectangular plate of 51-kg mass is supported by a hinge at C and a


cable DE as shown in the figure. Determine the tension in the cable, and the
reactions at the hinge support at C. Given: a = 900 mm.

" REEMA SE S sceppeemicsomaece tance SRV NTRS é 250 mm

f 1000 mm ———»

Figure P4.60

4.61 The plate is supported by a hinge at A and a vertical cable BC as shown in the
figure. Determine the tension in the cable, and the reactions at the hinge support at A.

Figure P4.61
Problems 223

4.62 A 6-ft-long uniform beam weighing 1800 lb is supported by a rough horizontal


surface at A, a smooth vertical wall at B, and a cable CD. Determine the tension in
the cable, and the reactions at supports A and B.

4.63 A uniform boom of 90-kg mass is carrying a load W = 3.5 KN. The boom is
supported by a ball-and-socket support at A and two cables BD and CE, attached
to the vertical wall as shown. If « = 90°, determine the tensions in the cables, and
the reactions at the ball-and-socket support at A.

4.64 Solve Problem 4.63 if a = 60°.

4.65 Solve Problem 4.63 if a = 115°.

Zz

Figure P4.62 q

3m

Figure P4.63, P4.64, and P4.65


224 Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

4.66 A uniform concrete slab weighing 40 KN is supported by three vertical cables


AD, BE, and CF as shown. If a = 1 m and b= 0, determine the tensions in the
cables.

4.67 For the concrete slab in Problem 4.66, determine the distances a and b so that
the tensions in the three cables are equal.

4.68 and 4.69 The beam shown in the figure is supported by a short bearing at A
which can exert only radial forces but no couples, a ball support at B, and a ball-and-
socket support at C. Determine the reactions at the supports.

5 So
C=— x

Figure P4.69
Problems 225

4.70 The frame shown in the figure is supported by three short bearings at A, B,
and C, which can exert only radial forces but no couples. If F = —2i + 5j — 3k
(kip) and M = 0, determine the reactions at the supports.

4.71 Solve Problem 4.70 if M = 10i — 20j + 15k (kip - ft).

4.72 A uniform platform weighing 4 kip is supported by two links CD and EF, and
two hinges at A and B that can exert only radial forces but cannot exert axial forces
and couples. If a = 90°, determine the forces in the two links and reactions at the
hinges at A and B.

4.73 Solve Problem 4.72 if a = 60°.

4.74 Solve Problem 4.72 if a ey

4.75 For the platform in Problem 4.72, assume that the link EF has been removed,
and that the hinge at A can exert axial force. Determine the force in the link CD
and the reactions at the hinges A and B. Given: a = 90°.

4.76 Solve Problem 4.75 if a = 60°.

4.77 For the platform in Problem 4.72, assume that the link CD and the hinge at B
have been removed, and that the hinge at A can exert axial force and couples about
the radial axes. Determine the force in link EF and the reactions at the hinge A.
Given: a = 90°.

4.78 Solve Problem 4.77 if a = 115°.

Figure P4.70 and P4.71 Figure P4.72, P4.73, P4.74, P4.75, P4.76, P4.77, and P4.78
Chapter 4 Equilibrium of Rigid Bodies in Three Dimensions

4.79 A space frame is supported by a ball-and-socket support at A, and by short


links at B and C as shown. If F = 0 and M = 9i — 13j — 7k (KN - m), determine
the reactions at the supports.

4.80 Solve Problem 4.79 if F = —2.5i + 1.5j — 2k (KN).

Figure P4.79 and P4.80


CHAPTER

CENTER OF GRAVITY,
CENTROID, AND
DISTRIBUTED FORCE

5.1 Introduction OUTLINE


5.2 Centroid—Definition

Two-Dimensional Problems—Areas and Lines

5.3. Centroid of an Area by Integration


5.4 Centroid of a Line by Integration
5.5 Centroid of Composite Areas
5.6 Centroid of Composite Lines
*5.7 Theorems of Pappus and Guldinus
*5.8 Distributed Load on Beams
*5.9 Hydrostatic Force on Submerged Surfaces

Three-Dimensional Problems—Volumes

5.10 Centroid by Integration


5.11 Centroid of Composite Volumes
5.12 Summary
Key Terms
Problems

227
228 Chapter 5 Center of Gravity, Centroid, and Distributed Force

5.1 INTRODUCTION
In order to understand the fundamental concept of center of gravity, we con-
sider a body of weight W as shown in Figure 5.1. This weight W is the sum
of the parallel gravity forces on all the elemental particles of the body, such
as A, B, .... Using integration, if the weight of the elemental particles at A,
B,..., is equal to dW,, dWz, ..., we can say that

W= |dw (5.1)
The center of gravity, G, of a body is defined as a point where the system of
parallel gravity forces on the elemental particles of the body can be replaced
by a single equivalent force W. The coordinates of the center of gravity of the
body shown in Figure 5.1 are x,y,z. The expressions for determining
x, y,andz can be obtained in the following manner. Let the position vector for
the point G be rg and, similarly, the position vectors for points A, B, ... be
r4, lg, ---, respectively. Now, taking the moment about O,

) X (—dWpj) to
+ rp
rg X (—Wj) =r, X (—dW,j
where rg =xi+t yj + zk
Figure 5.1
Tr, =X,i + yaj + 24k
Ip = Xpi + yaj + Z—k

i,j,k ll= unit vectors in the direction of x, y, and z axes


or we may write

rgW X (—j) =14dW, X (—j) + tgdWy X (—p ++


From this equation we observe that

roW =r,dW,+rgdW,+:::
thus

rgGW = [raw (5.2)

where r is a position vector for a typical elemental particle of the body and
r=xi+ yj + zk

Making proper substitutions for rg and r in Equation (5.2), we obtain

(xi + ¥j + ZK)W = I(xi + yj + zk) dW


The preceding equation shows that
5.1 Introduction 229

xW = |saw

From this we can see that the x coordinate of the center of gravity of a body
can be expressed as

a
ve W (5.3a)
5.3

In a similar manner, it can be shown that

[aw
yy= W 5.3b
(5.3b)

and

|zadWw
A (5.3c)

In the study of dynamics, which relates to problems concerning the motion


of bodies, it will be necessary to determine the coordinates of a point called
the center of mass. If the acceleration due to gravity (g) on all elemental
particles of the body is the same, then
W = gm and dW = gdm
where m= mass ofthe body
dm = mass of any elemental particle in the body

So, from Equations (5.3a), (5.3b), and (5.3c),

|seam |xam
xX = —— =
gm m

[» eam |vam
y = — = (5.4)
gm m

|seam |sam
Z = ——_ =
gm m
230 Chapter 5 Center of Gravity, Centroid, and Distributed Force

By comparing Equations (5.3) and (5.4), we can see that the center of mass
coincides with the center of gravity. It is important to note, however, that the
center of mass is independent of gravity.
The preceding equations and their derivations will be used later in this
chapter to solve various problems.

5.2 CENTROID—DEFINITION

Volume
Volume = V
Referring to Figure 5.2, if the body is made of a homogeneous material and
Specific Reece :
weight
= y_.< — “
the specific weight of the material is equal to y, then
Volume =dV
W=yV and dW = ydV
where V = volume ofthe body
dV = elemental volume
Substitution of the above expressions for W and dW in Equations (5.3)
yields the following:

Os}

The point coordinates x, y, and z as described by Equations (5.5) are defined


as the centroid of the volume. It is important to note that the centroid of the
Figure 5.2 volume of a body and the center of gravity of the same body are the same
provided the body is homogeneous. If the body is not homogeneous, the cen-
y
troid, being a geometrical property of the body, will in general not coincide
with the center of gravity.

Area

If a body is homogeneous—that is, has a constant value of specific weight


‘y—and has a small but constant thickness f, as shown in Figure 5.3, then it
can be modeled as an area A. Referring to Figure 5.3, for the elemental volume
shown,

dW = ytdA
Again, for the entire body

W=ytA

Substituting the above expressions into Equations (5.3a), (5.3b), and (5.3c),
Figure 5.3 we obtain
5.2 Centroid—Definition

yt |xa yt |yaa yt | aA
———— y = ———_ es
ytA ta yta

The constant specific weight y cancels out of the numerators and denominators,
and we obtain

(5.6)

These relationships define the positions for the coordinates of the centroid of
an area.
Most of the frequently encountered problems related to the determination
of the centroid of an area are two-dimensional problems. This means that the
area lies in one plane as shown in Figure 5.4. Referring to the area, which lies
in the xy plane in this case, we see that

Or7)

Figure 5.4

If an area lies in the xy plane and is symmetric about the y axis, then the
integration of f x dA will be equal to zero; thus, x= 0. This means that the
centroid of the area lies on the y axis. Similarly, if the area is symmetric about
the x axis, the integration of f y dA will be zero. Thus y will be equal to zero,
which means that the centroid will lie on the x axis. For any area with an axis
of symmetry, the centroid will lie along that axis. The areas shown in Figure
5.5 are symmetric about the x and y axes; therefore, x = O and y = 0. Figure 5.5
232 Chapter 5 Center of Gravity, Centroid, and Distributed Force

Vv
The areas shown in Figure 5.6 are symmetric about the y axis only. Thus for
these areas x = 0; however, y # 0.

Line
A.
L ss If a slender rod of length L is homogeneous (i.e., has a constant value of
specific weight yy) and has a constant area of cross section (A), then it can be
he approximated as a line of /ength L (Figure 5.7). For the segment dL

dW = yA dL
O 4 and, for the entire rod,

W = yAL
'G
Substitution of the above relationships for dW and W into Equations (5.3a),
Figure 5.6
(5.3b), and (5.3c) yields

yA |xa yA |va yA |sax


es y= a=
yAL yAL yAL
or

|saz ydL |sat


x= y= z= 5.8
aie ae es oo i oe)
For two-dimensional problems (that is, when the line lies in one plane such as
xy) as shown in Figure 5.8, then

Figure 5.7 (5.9)

Figure 5.8
5.3 Centroid of an Area by Integration 233

It is important to note that the centroid G for the line may not lie on the
line itself. As in the case of areas in the xy plane, if a line is symmetric about
the y axis then f x dL = 0, which results in x being equal to zero. Similarly,
if the line is symmetric about the x axis, then y will be equal to zero.
The remainder of the chapter is divided into two parts: The first considers
problems in two dimensions; the second, problems in three dimensions.

TWO-DIMENSIONAL PROBLEMS—AREAS AND LINES

5.3 CENTROID OF AN AREA BY INTEGRATION


Consider an area that lies on the xy plane and is bounded by an analytical
curve as shown in Figure 5.9a. The location of the centroid of the area can be
determined by considering an elemental area dA. For this elemental area,

dA = dx dy

So, using Equations (5.7)

|saa | vasdy |yaa [J vavay


ee eee
Proper choice of the differential element can simplify the calculation of a
problem. As we can see from the above relationships, double integration will
be required to determine x and y. However, in most cases, it is possible to
determine x and y by performing single integration. This is done by choosing
an elemental strip rather than an elemental area, as shown in Figure 5.9b. For
the vertical elemental strip shown, the centroid is located at C whose coordi-

(a) (b)
Figure 5.9
234 Chapter 5 Center of Gravity, Centroid, and Distributed Force

nates are x,, and y,,, where

Xe] =x Ya = 5 dA
= ydx
Thus, for the centroid of an area by integration for an elemental strip,
)
a a a e J U e )oeo SCG
__L
|vax (ee
Jaa
Similarly,

The location of the centroid could also have been determined by considering
a horizontal elemental strip as shown in Figure 5.9c. For this case

ia = 5 va dA = xdy
dA
Therefore, we find that
Be
IyA |yx dy iKdA a
y= = and x= =
Vo) . [aa [xa) [aa |xa

(a) :
: Figure 5.10a shows an area bounded by an analytical curve expressed in
terms of polar coordinates. For such a case one can use an elemental area dA
y that is triangular in shape. For the elemental area (Figure 5.10b),

1 1
dA = Vr a) = 57 40

= w = D, ‘
X= (= r) cos@ Jeo9
Xo, = (2: elas (2)sin 0

(Note: For the location of the centroid of a triangle, refer to Table 5.1, Section
5.5.) Substituting the preceding relationships into Equations (5.7), we obtain

pL 1 3
“ |yaa \(% sin 0)
(3-240) [Esin 6 dé

(b) j= = 1 : “
Figure 5.10
|aa [42a Yd
5.3. Centroid of an Area by Integration 235

and

Thus we find the location of the centroid, performing only one integration.

EXAMPLE 5.1

For the area shown, determine the location of the centroid. y

Solution The elemental strip of thickness dx is chosen, as


shown in the figure. For the strip we have that
* es y yee ss
dA = ydx Xe =X es 5 me

Using Equations (5.7) and performing the integration, we find


the coordinates of the centroid:
L zl L

fina Lo|(Berfe #f' oa

; Gl= af
.
Ey

_ att LLG\E][ (ee)


236 Chapter 5 Center of Gravity, Centroid, and Distributed Force

EXAMPLE 5.2

Determine the location of the centroid for the semicircular area


shown.

Solution An elemental horizontal strip with a thickness of dy


is chosen, as shown in the figure. For the elemental strip we
write

X.)dA
i= 0 y.=y dA = (2V R* — y”)dy r= A

Since x,, = 0, x =0 <


; R (R? Le, yy me

|5aaa iH}ee ae — (3/2) | AR


y=
A mR? aR? 30
2 2

y=

It can be seen that the y axis is an axis of symmetry for the


area. Therefore, the centroid should lie along that line; in other
words, x = 0.

EXAMPLE 5.3

Refer to the semicircular area of Example 5.2. Determine the


ordinate y for the centroid of the area by using polar coordinates.

Solution For the triangular elemental area chosen,


5.3 Centroid of an Area by Integration 237

EXAMPLE 5.4

For the area shown, determine the location of the centroid that is
x and y.

Solution At the outset, we need to determine the constants m


and k associated with Equations y, = mx and y, = kx’.
Note that at x = L, y= H. Thus, m = H/L andk = H/L’.
Consider the elemental strip as shown in the figure. We can
now write
238 Chapter 5 Center of Gravity, Centroid, and Distributed Force

5.4 CENTROID OF A LINE BY INTEGRATION


Consider a line as shown in Figure 5.11. From Equations (5.9), the coordinates
of the centroid are

|xdL |x dL
= and y=
idL |dL
From Figure 5.11 we can see that

(dL)” = (dx)? + (dy)


Therefore
ia Palit 2 2
O geen ea! db= /1+ (2) dx and also db = /1+ (2) dy
x Ly
Beet Hence, we can determine the centroidal coordinates of a line by

x /1+{—]
dy\* dx |x dx\*
/1+{—]d
J\ (2) Vi oy one
x= en
d
i h+(2) dx | h+(2) dy
dx dy (5.11)
dy “ | dx\*
[es eh Gee Peale he
mh |: a) ; (*) :
| age 5-4 5
dy I dx
Reel heey backs (ey eed |e)
J & @
In polar coordinates (Figure 5.12) the elemental length dL can be ex-
=
pressed as
2

dL = V (rdo)? + (dr? = |r24+ (*) de

We can see that x,, = r cos 8 and y,, = r sin 0; therefore

(5.12)
dr\?
Figure 5.12 |Foire |— sin ede

EXAMPLE 5.5

Locate the centroidal coordinates x andy of the semicircular arc


shown.

Solution Because of symmetry along the y axis,we know that


x = 0. From Equation (5.11), we can write
xe = R?
dy | | dy :
yo
(es 1+{(—] dx
7 Joy (2) 2 (2
a es
For thearc,
Re y=

dy dy ay
Dag = Dy\| SS | = SS =
. (2) dx
Therefore,

[. yee i ay i d
sa Rr? Nees Rr te y JA Oe
us TR TR TR wR 7
= on AK
Oana
TT

5.5 CENTROID OF COMPOSITE AREAS


Figure 5.13 shows a composite area ABCDE. A composite area consists of a
series of connected simple areas such as rectangles, circles, triangles and so
forth. To determine the centroid of a composite area, the following steps should
be followed.

1. Divide the composite area into several parts consisting of simple areas
girly 248, I, By Sh Shs oo a 6
2. Determine the areas of the parts. Let these be A,, A, Az, Ay,....
3. Determine the coordinates of the centroids of these parts. In Figure 5.13
(Soo DEER Nh Chas hOne wera ih oma
4. For the composite area, calculate O
Figure 5.13
ns Ax, a A>X> == A3X3 + ee Al
5 SS SSS 2S SS
Apt Aa + Ag to SAl (5.13)

y= Ay a AzY2 By A3y3 Tae mu LA


A, +A, +A;+--- DA;

The preceding equations are similar to Equations (5.7). For clarity when cal-
culating, the values can be arranged in tabular form.
The applications of the above principles are shown in the following example
problems. Table 5.1 provides the location of the centroid of some simple areas
that have been obtained from basic integration techniques.
239
Table 5.1 Location of centroid of common areas

Location of centroid
Shape Area,A
y
Rectangle ¥ |

bh h/2

a 2b = b’ h/3

i 4bh b/3 h/3

: —-
Circle y

: :
: ms
@Q-

Semicircle -’ =

mr? 4r

Quarter-circle
area
mr 4r 4r
ea 3m 31

Ellipse y

tab 0 0

240
Circular
segment

r2a—r? sinacos a 0 3 (__sin“a _)


3 \a —sinacosa

2 2rsin a
ay? are 0

Quarter-circle +
spandrel
UES) 2r (10—37)r
Tile 3(4—7) 3(4—7)

O
y
Trapezoid —)\ +4
h hi2b\ tbo)
77(0) +82) 3 (bi +b)
O x
= at 2S aa eS

Parabolic area y
+a a—>
e 4ah /3 0 3h/5
nh We
iy ¥
x
1%)
Semiparabolic 7
area Yee
i di 2ah/3 3a/8 3h/5
iE :
1

Parabolic »
spandrel
ah/3 3a/4 3h/10

General
spandrel

241
EXAMPLE 5.6

Determine the centroidal coordinates of the composite area


shown.
eal mm—>
Solution The composite area shown can be interpreted as
100mm
follows

Area ABCDEFG = Area ABCEFG — Area CDE


G oe————.._ o—_—_—_—4’

Radius =r =25 mm rectangle semicircle


E
{50mm Now we can prepare the following table.
F
x

A; a Yi A;Xx; Ai;

Area (mm’) (mm) (mm) (mm*) (mm?)


ABCEFG 30 x 10° 50S 125 30 + 100 = 130 SSO) SX 10? 3900 x 10%
CDE —0.98 x 10% 50 + 150 — 10.617 = 189.39 30 + 50 + 25 = 105 — 185.98 x 10° — 103.11 x 10°
SA, = 29.018 X 10° mm? 2AX, = DAY; =
3564.02 x 10? mm? 3796.89 x 10? mm4
4r 4 ™ 25
T— = = 10.61 mm (See Table 5.1.)
30 3%

_
9
LAX; 3564.02 x 10° SA,y; _ 3796.89 x 10°
LA; 29.018 x 10° A, 29.018 X 103
x= 12232 nam < y= 130.85mm <

EXAMPLE 5.7

A thin plate in a vertical plane has the shape of a quarter of a


circle and is hinged at O. The weight of the plate is 10.5 lb. A
force F acting at A keeps the orientation of the plate as shown.
Determine the magnitude of F.

F<«—
Solution The weight of the plate can be treated as a
concentrated force acting at the center of gravity G of the plate.
Referring to the figure and Table 5.1,

Taking the moment about O, for equilibrium


Hinge
O _ (10.5)(3.40)
(F)(8 in.) = (10.5 1b)(3.40 in.) F
8
F=446 lb <

242
5.6 Centroid of Composite Lines 243

5.6 CENTROID OF COMPOSITE LINES


A composite line consists of a series of connected lines having simple shapes,
such as straight lines and circular arcs, as shown in Figure 5.14. The centroidal
coordinates for the composite line can be determined by dividing the line into
a series of simple lines having lengths of L,, L5, L3, ... and determining the
centroidal coordinates of each simple line as (x,, y,), (%>, V2), (%3, Y3),----
The centroidal coordinates of the composite line can be calculated in a similar
manner to that for the case of composite areas (Section 5.5); hence

za + In, + LyX, +--+ xb;


L,+L1,+1,+-:- Dayby (5.14)
pene + Lny, + Lays ts: _ 2yiL;
Ly Loe La ee Da} O .
Figure 5.14
These equations are similar to those given in Equations (5.9).
Table 5.2 provides the location of the centroid for some lines having simple
shapes.

Table 5.2 Location of centroid of lines with common shapes


ood a
Location of centroid
Shape Length, L = — ~
y x
Straight
line
L

O
G ——
E 2 ;
SS (5

Semicircular
area
2r

@ if 0 =
iP
O
y Quarter of a
circular arc

Arc of a circle
244 Chapter 5 Center of Gravity, Centroid, and Distributed Force

EXAMPLE 5.8

y A thin homogeneous wire ABCD is shown. Determine its center


of gravity.

Solution This is the case of a thin homogeneous wire of


constant cross section. As stated in Section 5.2, such a wire can
be approximated as a line. The centroid of the line and the
center of gravity of the wire coincide. The wire ABCD can be
divided into three segments: AB, BC, and CD. The following
table can now be prepared to determine the center of gravity of
the wire ABCD.

Seg- Length, Xj Yj Lx; Liy;


ment _L, (in) (in) (in) (in?) (in’)
AB 4 2+2=4 2+4=4 16 16
BC 5 24+24+2=55 2+$=4 Qe 20
CD 5) Dee sia 2+4=6 PPS) 30
SL, = 14 in SLX, = 66 in? SL,y, = 66 in?
-_ 2Lx; _ 66 s=471in <
x <0 14 X= 4./1 m.

EPS iy 00 : ;
ee STi eae y=471in. <

*5.7 THEOREMS OF PAPPUS AND GULDINUS


A plane line, when rotated about a nonintersecting fixed axis, will generate a
surface of revolution. For example, if a straight line AB is rotated about a
fixed axis AC as shown in Flgure 5.15, it will generate the surface of a cone.
In a similar manner, if a plane area is rotated about a fixed axis it will generate
the volume of revolution of a body. For example, if a triangle ABD (Figure
5.16) is rotated about a fixed axis AC, it will generate the volume of a cone.
The area of the surface of revolution of a plane curve can be determined in
the following manner. Referring to Figure 5.17, let AB be a line of length L
whose centroid is located at G. Consider an elemental length dL on the curve.
If this length dL is rotated through 27 radians, the surface area dS generated
C will be equal to

Figure 5.15 dS = (271)(x)(dL)


5.7 Theorems of Pappus and Guldinus 245

E
Figure 5.16 Figure 5.17

Hence, the total surface area S generated by the rotation of the plane curve
AB is

s= |as= |2nvat = 20 |xa


However, we have seen from Equations (5.8) that

[xa = Fl,

where x is the distance between the AC axis and the centroid of the curve AB.
Therefore,

S = 2uxL (5.15)
Hence, we can make the following general statement.

The area of the surface of revolution generated by a plane curve about a STATEMENT
nonintersecting fixed axis is equal to the length of the curve times the
distance traveled by the centroid of the generating curve, or
S = OrL
where L = length of the curve
r = perpendicular distance between the fixed
axis and the centroid of the curve
6 = angle of revolution in radians
(Note: 86 =2m7)
246 Chapter 5 Center of Gravity, Centroid, and Distributed Force

Similarly, we consider the case of the volume generated by rotating a plane


area about a nonintersecting fixed axis by referring to Figure 5.18, in which
ABD is a plane of area A whose centroid is located at G. Consider an elemental
area dA. If this area is rotated through 27 radians, the volume dV generated
will be
dV = (27)(x)(dA)

Figure 5.18

The total volume V generated by rotating the area ABD about the fixed axis
AC can be given as
V= | 2axdA =2T xdA = 27xA (5.16)

where x is the distance between the AC axis and the centroid of the area ABD.
Based on the above derivation, we can now make the following general
statement.

STATEMENT The volume of revolution generated by a plane area about a nonintersecting


fixed axis is equal to the area A times the distance traveled by the centroid
of the generating area, or
V=06rA
where A = area to generate volume
r = perpendicular distance between the fixed
axis and the centroid of the area
6 = angle of revolution in radians
(Note: 8 S27)

The above two statements are usually referred to as the Pappus-Guldinus


theorems; they were originally formulated by the Greek geometer Pappus
(third century A.D.) and were reiterated by the Swiss mathematician Guldinus
(1577-1643).
5.7 Theorems of Pappus and Guldinus 247

EXAMPLE 5.9

The curve OA is a quarter of a circle. If the shaded area A is


revolved about the x axis through 360°, what will be the volume
of the body generated?

Solution The volume of the body generated will be

V = (2m) (y) (A)


———$$ —_
i t
y coordinate for area of the
the centroid of the composite
composite area area

The shaded area A shown is equal to A, — A, (where A, =


area of the quarter-circle spandrel and A, = area of the
semicircle). Thus,
2
AS aie Ag = E # 0) = a C= 350mm—"
y
=
050 Zee | ae
7350) 2
ay m(25)”
ry =25mm

= 26,288.73 — 981.75 = 25,306.98 mm?

For the composite area, using Equations (5.13) and subtracting


O
lL
Area=A,

i x

rather than adding for the missing semicircular area,

Ay, — A2y2

sae HEAHA)
Referring to Table 5.1 for a quarter-circle spandrel,

(4-73) (4 7)(3)
Similarly, for the semicircular area,

_ _ 47 _ (4)(25)

So

_ (26,288.73)(78.18) — (981.75)(10.61)
= 80.8 mm
" 25,306.98
Hence, substituting these values into the expression for the
volume, we find

=: 80.8 23,900.98. 5
V = (277)(y)(A) = (277) Mien: an OOSa.cm

V = 12,848 cm? <


248 Chapter 5 Center of Gravity, Centroid, and Distributed Force

*5.8 DISTRIBUTED LOAD ON BEAMS


The concept of centroid of areas can be conveniently used to determine the
reactions at the supports of beams that are subjected to distributed loads.
Figure 5.19a shows a simply supported beam of length L and width B. The
beam is subjected to a distributed load of p per unit area. The distributed load
p is a function of the distance x. Figure 5.19b shows a two-dimensional view
of the beam. The load per unit length of the beam is w, or w = pB. The
distribution of w along the length of the beam is shown in Figure 5.19c.

p(x)

Figure 5.19

The total load W on the beam can be determined by obtaining the area of
the load diagram:

is
w={, wdx (5.16)

The distributed load on the beam can be replaced by an equivalent single


concentrated load W (Figure 5.19d); this will facilitate the determination of
the reactions on the supports. In order to do that, we take the moments about
C; thus
L
wz= | w(x) dx (S17)
5.8 Distributed Load on Beams 249

or

L
ifw(x) dx
x= ere (5.18)

Comparing Equation (5.18) with Equations (5.7) shows that x in Equation


(5.18) is actually the distance of the centroid of the distributed load area from
C. Based on the above observations, we can make the following statements.

1. The magnitude of the total load W on the beam is equal to the area of
the distributed-load diagram.
2. The distributed load on the beam can be replaced by a single concentrated
load W, with its line of action passing through the centroid of the
distributed-load diagram.

EXAMPLE 5.10

A beam is subjected to a distributed load as shown. Determine


the reactions at its supports A and B.

Solution The distributed-load diagram is a triangle. The area


of the load diagram is

We (:)ases) =5.5kN G)
For centroidal location of the area of the load diagram,

The beam with the equivalent single concentrated load


W = 5.5 KN is shown.
Considering the free-body diagram of the beam, the
reactions at the supports are A, and A, at A, and B, at B.

@ 3F,=0 A,=0<
© =m, =0
(B,)(2.5) — (5.5)(0.833)=0 5.5)(0.833
B= a
B,=183kN@) <
@ =F =0
A,+B,=55kN A,=5.5—B,=5.5-1.83
A, =3.67kN(t) <
250 Chapter 5 Center of Gravity, Centroid, and Distributed Force

*5.9 HYDROSTATIC FORCE ON SUBMERGED SURFACES


The principles for determination of centroid can also be used to calculate the
hydrostatic force on surfaces of submerged bodies. In this section, the pro-
cedure for calculating hydrostatic force on submerged rectangular and curved
surfaces having constant width will be presented.
In order to understand the concepts, let us consider a flat rectangular plate
of length L and width B submerged in a liquid as shown in Figure 5.20a. The
width B is at a right angle to the cross section shown. Let the specific weight
of the liquid be equal to y. According to Pascal’s law, a fluid at rest creates
a pressure p ata point, that is the same in all directions. So at depth y, (Figure
5.20a) the fluid pressure (at point a) is equal to p; = yy,, and this pressure
acts normal to the surface of the plate. Similarly at point b, which is at a depth
Yy, the fluid pressure is equal to p, = ‘yy>. Hence it can be seen that the diagram

(a)

(c) (d)
Figure 5.20
5.9 Hydrostatic Force on Submerged Surfaces

of distribution of the hydrostatic pressure on the inclined plate will be a trun-


cated prism (Figure 5.20b). The variation of the hydrostatic force per unit
length of the plate can thus be given as P = yyB. This variation is shown in
Figure 5.20c. The total hydrostatic force R on the plate can be calculated as

L
R= i,IOBG Gils)

Thus, we can write the expression for the hydrostatic force R as

P,+P A
ne (AtPs), = BL (42) (5.20)

The line of action of the resultant hydrostatic force R can be obtained by taking
the moment of the force diagram about point O, or

|pas’
3 Te R 521
O21)

The relationship for x’ can be obtained by referring to Figure 5.20d, or

Rx'~ =R Pie oN a. Y1 sal ae2\ (See O27


i ite 2) )25 (2)( cos 0 )| aa
However, note that

ae vet +
2)
1 1
R= Pb =yy,BL BG NO) IIE (3)BE - 309

Substitution of the above relationships for R, R,, and R, into Equation (5.22)
yields the expression for the line of action of the resultant hydrostatic force:

ea el yo—y1\ (1 + 292 :
ase is a ‘ @ =) 3 )| ee

The calculation procedure for the magnitude and line of action of the re-
sultant hydrostatic force on a plate as presented in Figure 5.20 is somewhat
simpler than that for a submerged curved plate like the one in Figure 5.21a.
The curved plate shown has a constant width B. Figure 5.21b shows a two-
dimensional view of the plate along with the distribution of the hydrostatic
force per unit length (P = yyB). The resultant hydrostatic force R can be
calculated by considering a free-body diagram with a cross section abc (Figure
252 Chapter 5 Center of Gravity, Centroid, and Distributed Force

P,=7y2B

(a) (b)
Figure 5.21

5.21c) and a width B at right angle to the cross section shown. The magnitude
of the resultant force R can now be given as
R=V(R, + Wy + R2 (5.24)
where R, = resultant of the distributed hydrostatic force on the
horizontal surface ac
= (P,)(ac) = (p,B)(ac) = (yy,B)(ac)
R,-= resultant of the distributed hydrostatic force on the vertical
surface bc
— + = +
= G(P=P2)p = Teyrw( A222)
W = weight of the fluid = (Area abc)(y)(B)

The line of action of the resultant force R is shown in Figure 5—21a.


5.9 Hydrostatic Force on Submerged Surfaces 253

EXAMPLE 5.11

The cross section of a concrete dam is shown. Consider a 1-ft


length of the dam at a right angle to the cross section shown.
Determine the resultant hydrostatic force R exerted by water on
the face AB. The specific weight of water, y, is 62.4 lb/ft’.

Solution The magnitude of the hydrostatic force per unit


length at a depth y below the surface of the water is
P=vyyB

For this problem, y = 62.4 lb/ft? and B = 1 ft, so P = 62.4y


Ib/ft. Aty = 0,P = 0. Aty = 15 ft, P = (62.4)(15) =
936 lb/ft. The distribution of P along the face AB is shown,
from which we find that

R= (3}.36 V (15)? + S|

R = 7437.5 lb <

The resultant R is inclined at an angle a to the horizontal.


7
a= n-"(2) = 19.299 q@=19.29° <

EXAMPLE 5.12

Repeat Example 5.11 using the procedure described with refer- Water level
ence to Figure 5.21c.

Solution Consider the free-body diagram of the volume of


liquid BCD.

W = weight of water = (3)


(5.25 ft)(15 ft)(1 ft)(62.4 lb/ft?)

= 2457 lb
1
R,= (3)as X 62.4)(15) = 7020 Ib

So, the resultant is found to be

R= V(R3 + (W)? = V(7020)? + (2457)*


Re /437-91by <<

a= 19.29° <
254 Chapter 5 Center of Gravity, Centroid, and Distributed Force

EXAMPLE 5.13

Consider a 1-ft length of the concrete dam whose cross section


is shown in Example 5.11. Determine the reaction by the ground
on the base of the dam. The specific weight of concrete is 150
lb/ft’.

Solution Refer to the free-body diagram of the section


AEFGHBDC. The reaction by the ground on the base of the
dam is given by forces F, and F,, and moment M. From
Example 5.12, W = 2457 Ib and R, = 7020 lb. Also

W,=W;= (3)eft)(20 ft)(1 ft)(1501b/ft?) = 10,500 Ib

W, = (6 ft)(20 ft)(1 ft)(150 Ib/ft*) = 18,000 Ib


For equilbrium,

© XF, =0 F.-R,=0 F.=70200Ib©<


© =F, =0
F,-W-W,-W,—W,;=0
F, = 2457 + 10,500 + 18,000 + 10,500
F,=41,4571b@) <
@ =m; =0
15 Sp)
SM o(¥) + «w)(20= 528)

a5 wa(7 ata Olas 1)

= wra(7EE ‘)“2 wva(7x :)=0

M = —(7020)(5) + (2457)(18.25) + (10,500)(15.33)


+ (18,000)(10) + (10,500)(4.67)
M = 399,700 Ib: ft <
5.10 Centroid by Integration 255

THREE-DIMENSIONAL PROBLEMS—VOLUMES

5.10 CENTROID BY INTEGRATION


In Section 5.1 we saw that the coordinates for the center of gravity of a body
in three dimensions can be given as

O13)

We also saw that if a body is made of a homogeneous material, the coordinates


of the centroid of the volume of the body coincide with the coordinates of the
center of gravity (Section 5.2) and that

(5.5)

In Sections 5.10 and 5.11, we will assume that the bodies are made of
homogeneous materials. As in the case of areas (see Figures 5.5 and 5.6), if
the body is symmetrical about a given plane the centroid of the volume of the
body must lie on that plane. Referring to Figure 5.22, the body is symmetrical
about the yz plane and, hence, x = 0. Also it is symmetrical about the xy plane
and so z = O. In Figure 5.23, the body is symmetrical about the xy and xz
planes. This means that z = 0 and y = O. However, there is no symmetry
about the yz plane, thus x # 0.
For determination of the centroidal coordinates (Figure 5.22) one can con- 2

sider an elemental volume dV where dV = dx dy dz. If this relationship for Figure 5.22
dV is substituted into Equations (5.5), we obtain

a [[[xavay a:

ae
Mf fowa
Nese
25)

_Sffesoe
Wee Figure 5.23
256 Chapter 5 Center of Gravity, Centroid, and Distributed Force

From Equations (5.25) we see that triple integrations are required to obtain
x, y,andz. However, as in the case of areas (Figure 5.9), proper elemental
volumes can be cleverly chosen that will reduce the number of required inte-
grations. Some examples for doing this are described below with reference to
Figure 5.24.

Rod Element
Figure 5.24a shows a rod element. For this elemental volume,

dV = xdydz = Ldydz

The centroid of the elemental volume is C. Thus,

L oo
x el = 5 = constant Vela” a a4

From Equations (5.5) we can write

V ieee

By this method, the triple integration will be reduced to a double integration.

Disc Element
Figure 5.24b shows a disc element. The elemental volume
dV = mR? dx
where R = f(x) and x,, = x. From Equations (5.5), we write the expression
for the centroidal coordinate as

|xdV ix,,dV |(x)(R? dx)

iaR? dx
Since the volume is symmetrical about the xz and xy planes, y= O and
z=0.
5.10 Centroid by Integration 257

<=

(b) (c)
Figure 5.24

Cylindrical Shell Element


Figure 5.24c shows a shell element. The elemental volume
dV =2tydA

However, since

dA = (L — x)dy
258 Chapter 5 Center of Gravity, Centroid, and Distributed Force

then

dV = 27y(L — x) dy
= zs L=x L+x
XG =X —
el » 5)

Therefore, we find the necessary centroidal coordinate by

oe =
V V
{|27 y(L — x) dy
Note that x is a function of y. Since the volume is symmetrical about the xz
and xy planes, y = 0 and z = 0.

EXAMPLE 5.14

A homogeneous quarter cone is shown. Determine its centroidal


coordinate x.

Solution Consider the disc element as shown. The elemental


volume can be written as

dV = (=°)dx ad r= ()x
4 ji
Therefore,

Also, as X,, = x,
5.11 Centroid of Composite Volumes 259

EXAMPLE 5.15

Refer to the homogeneous quarter cone shown in Example 5.14.


Determine its centroidal coordinate y.

Solution From Example 5.14, the volume of the disc element


is

av = (=)(F)(xP ax f

G I
x |5.av ; NO

[ow L(G orl@)

5.11 CENTROID OF COMPOSITE VOLUMES


A composite body consists of a number of connected bodies of simple shapes
such as the one shown in Figure 5.25. In order to find the centroid of the
composite volume of the body, the following procedure can be used.

AN
(¥3,¥3,73)

Figure 5.25
260 Chapter 5 Center of Gravity, Centroid, and Distributed Force

1. Divide the body into several parts with common shapes such as 1, 2,
35 aie
Calculate the volume of each part as V;, V2, V3,....
3. Locate the coordinates of the centroid of each part as (X,, y),2Z)),
(X25 2522), (353,23), en
4. Calculate the coordinates of the centroid for the composite body as

wks
yy = x,V; + XV, + X3V; 420 0 KV,
a— s = ——_
V, + V,+V3t+--: 2V;
5 ate ele (5.26)
VV Va ee Vv,

So SS

Table 5.3 gives the location of the centroid of some bodies having common
shapes.

Table 5.3 Centroid of volumes of common shapes

Shape Volume [ Rin My


Hemisphere
x=0
23 7R p3 yas
yok

z=0

x=0
NY
3 LBH roeeris
ae

Zz =0

x=0

Ih
3 TRS 2 H mee
YA
z=0
5.11 Centroid of Composite Volumes 261

Half-cone

x=0
1 ae 7A
4 RH Ven
mR
z=
TT

¥=SL
I3 LBH asV=5JE

7=2B

Paraboloid of
revolution y
x=0
2
2) TR H
SSH
y 3

z=0

EXAMPLE 5.16

A homogeneous composite body is shown. Determine the coor-


dinates of the centroid.

Solution Due to symmetry we know that

10 and

Hemisphere

(continued)
262 Chapter 5 Center of Gravity, Centroid, and Distributed Force

EXAMPLE 5.16 (concluded)


In order to determine y, we divide the composite body into two
parts as shown. Now we can prepare the following table.

Number Ps de
of the Volume, V; Yi ViYi
body (in?) (in.) (in*)

1 (4)maoreo = 2094.4 tex 10,472

2 (2)maor = 2094.4 -(2)a0 = —3.75 —7,854

SV, = 4188.8 in? SV,y, = 2618 in?


y=
x N
SS
eA ie.e}
y = 0.625 in. <
aS
&1

EXAMPLE 5.17

y A homogeneous composite body is shown. Determine its cen-


troidal coordinates.

Radius of hole
in shape of a
half-cylinder = 2 in.
(continued)
5.12 Summary 263

EXAMPLE 5.17 (concluded)

Solution The given volume can be divided into three parts. 3 in.
Now the following table can be prepared.
1Sin.

Vol-
ume Volume nie 7 - = = =
num- V; x; Ji zi Vix; Vii Viz;
ber (in?) (in.) —(in.) (in) Cin’) (in*) (in*) O0in:
1 900 LS 10 18) 1350 9000 6750
2 G0 Sa 7S 72 eS) 9450 1800 6750
= 10.5
3] = 1N13}5I 9 (2) 75 —1017.9 —96.02 —848.25
3T
0.849

2V; = 2Vix,= 2Viy, = 2,2, =


1686.9 in? 9782.1 in* 10,703.98 12,651.75 18in. Radius= 2 in.
in* in*
_
x=
OA =
SEP x=58in. <
ZV, 1686.9
—_ 2V,y; _ 10,703.98 y= 6.35in. <
SV 116869
6 NG, IASei Orbs) z=7.5in, <
Z= = eo

XV; 1686.9

5.12 SUMMARY
In this chapter we have covered the principles of calculating the centroid of
homogeneous bodies, areas, and lines and have learned the following:
1. The centroidal coordinates of a volume can be given as (Section 5.2)

|vav |sav
i ee a 55
oy a TR o>)
The centroidal coordinates for areas and lines are as follows:

ixdA iydA |zdA


Areas xr= je yaaa
A Z
B= 7 (5.6)
5.6

|xdL |ydL |zdL


Lines a —— a (5.7)
264 Chapter 5 Center of Gravity, Centroid, and Distributed Force

2. The centroidal coordinates of a plane area in the xy plane can be


determined by single integration, if one chooses proper elemental strips
that are parallel to the x and y axes (Section 5.3).
3. The centroidal coordinates of a line in the xy plane can be determined
by single integration using Cartesian and polar coordinate systems
(Section 5.4).
4. For composite areas and lines in the xy plane (Sections 5.5 and 5.6),

— 2xA; _ -d7,A;
Areas: Se ee 5.1

Lines : pe
x= ee
y= (5.14)
:

5. The surface of revolution: of a plane curve about a nonintersecting fixed


axis and the volume of revolution of a plane area about a nonintersecting
fixed axis can be determined by using the theorems of Pappus and
Guldinus (Section 5.7).
6. The reactions at the supports of a beam carrying a distributed load can
be calculated using the concept of the centroid of an area (Section 5.8).
7. The hydrostatic force on a submerged surface can be calculated by using
the concept of the centroid of an area (Section 5.9).
8. Triple integration to determine the centroidal coordinates of a volume
can be reduced to double (or single) integration by properly choosing
elemental volumes (Section 5.10).
9. For composite volumes (Section 5.11),

7 2x,V; = oe LyiV; a ZV;

{SiS AA, Sy ae es

KEY TERMS

center of gravity 228 distributed load on a beam 248


center of mass 229 homogeneous body 230
centroid of an area 23/ hydrostatic force 250
centroid of the volume 230 rod element’ 256
composite area 239 shell element 257
composite line 243 surface of revolution 244
composite volume 259 volume of revolution 244
disc element 256
Problems 265

PROBLEMS

SECTION 5.3
5.1 through 5.20 Determine the centroidal coordinates of the shaded areas shown.
Use the integration method.
y

Figure P5.1 Figure P5.2, P5.67, and P5.118 Figure P5.3 and P5.119


Figure P5.4 Figure P5.5 Figure P5.6 and P5.120

4 ft

Figure P5.9
266 Chapter 5 Center of Gravity, Centroid, and Distributed Force

l
,-—a—|
Figure P5.10 Figure P5.11 Figure P5.12 and P5.122

O ii, Se. O 2in. 4in. ‘ O


Figure P5.13 Figure P5.14 and P5.123 Figure P5.15

a
Figure P5.16 Figure P5.17 Figure P5.18
Problems

VAX

= = x2

0 x O ——
3 a
Figure P5.19, P5.57, and P5.69 Figure P5.20

SECTION 5.4
5.21 through 5.25 Determine the centroidal coordinates for the lines shown. Use
the integration method.

Figure P5.21 Figure P5.22

Quarter of
pa circle

y=mxt+e

Y
O : +—2 iN O

Figure P5.23 and P5.70 Figure P5.24 Figure P5.25


268 Chapter 5 Center of Gravity, Centroid, and Distributed Force

SECTION 5.5
5.26 through 5.45 Determine the centroidal coordinates x and y for the composite
areas shown.

uy

40 mm

x2 +(y—R)?=R?
20mm vy )

O 20mm 40 mm O 5 O
Figure P5.26 and P5.58 Figure P5.27 Figure P5.28

O
Figure P5.30

15 in.

O}«— 200 mm—+ ‘ O 2 (i


Figure P5.31 Figure P5.32 and P5.59 Figure P5.33 and P5.60

-
Problems 269

Vv

O Le mm——+|

Figure P5.34 and P5.61 Figure P5.35 Figure P5.36

ae) errs

Figure P5.37 Figure P5.38 Figure P5.39

Quarter of
a circle
Quarter of
an ellipse

Figure P5.41 and P5.141 Figure P5.42


270 Chapter 5 Center of Gravity, Centroid, and Distributed Force

Parabolic area
Circle; Radius= 8in. Ma.

| 0
% x }+——400 mm ——+'++—— 400 mm —+|
Figure P5.43 Figure P5.44 and P5.142 Figure P5.45

5.46 A thin plate in the vertical plane has the shape of a sector of a circle and is
hinged at O. The weight of the plate is 18 Ib. Determine the magnitude of the force
P that will keep the orientation of the plate as shown.

5.47 The centroid of an area in the shape of a trapezoid shown is somewhere on the
dotted line. Assuming a is known, what is x in terms of a? Where on the dotted line
is the centroid?

5.48 A uniform thin strip of length L is folded over as shown. What is x in terms of
L if the center of gravity is at the junction A?

Hinge—g og [+ ——.x —+|+ —— a+ pee


Figure P5.46 Figure P5.47 Figure P5.48

5.49 A thin plate measuring 2L X L is hanging by a cable from the ceiling. The
plate has a hole of radius L/3. Determine the distance, a, so that the position of the
plate while hanging will be as shown.

Figure P5.49
Problems 271

5.50 A thin plate is hanging from the ceiling by a wire as shown. The plate remains
horizontal while hanging. Determine the angle 0.

SECTION 5.6
5.51 through 5.56 Determine the centroidal coordinates of the composite lines
shown.
GER OS)

|
|

6 <

Figure P5.50
y

100 mm

eS 100 mm
200 mm
x
O
Figure P5.51 Figure P5.52 Figure P5.53

aa

y
205

40mm
0 Be

O O
Figure P5.54 Figure P5.55 Figure P5.56
272 Chapter 5 Center of Gravity, Centroid, and Distributed Force

5.57 through 5.61. A thin homogeneous wire is bent to form the perimeter of the
area shown. Determine the centroidal coordinates.

5.57 See Figure P5.19:

5.58 See Figure P5.26.

5.59 See Figure P5.32.

5.60 See Figure P5.33.

5.61 See Figure P5.34.

5.62 A uniform rod in the shape of a semicircle has a weight of 0.65 lb/in. It is
hinged at O. Determine the horizontal reaction at the smooth surface A.

5.63 A thin uniform rod having the shape of a semicircle is hinged at A. Determine
the angle a for vertical equilibrium position.

Figure P5.62 Figure P5.63

5.64 ABC is a thin uniform rod. The section AB is in the form of a semicircle and

Aids
the rod is hinged at A. Determine the length L in terms of the radius R so that the
section of the rod BC is horizontal.

5.65 What is the value of H/L if the centroid of the triangular area is at the same
location as the centroid of the triangular thin wire?

pe pret
Figure P5.65
Problems 273

5.66 What is the value of L,/L, so that y for the triangular area is the same as the y uf a4
of the thin wire forming the perimeter of a square area?

SECTION 5.7*
Use a theorem of Pappus and Guldinus to solve Problems 5.67 through 5.85.

5.67 Determine the volume of solid generated by revolving the area shown in tl eave Le Wire
Figure P5.2 about (a) the x axis, and (5) the y axis.
O éx x
5.68 Determine the volume of solid generated by revolving the area shown in a camel
Figure P5.8 about (a) the x axis, and (b) the y axis.
Figure P5.66
5.69 Determine the volume of solid generated by revolving the area shown in
Figure P5.19 about (a) the x axis, and (b) the y axis.

- 5.70 Determine the surface area generated by revolving the line shown in Figure
P5.23 about the x axis.

5.71 Determine the surface area and the volume of the body.

5.72 Given the length of a line L, what is the angle @ if the volume of the cone of
revolution about the x axis is a maximum?

— Radius = 3 in.

Figure P5.71 Figure P5.72

5.73 Determine the volume enclosed by the shell.

Radius= 20 mm ie

Figure P5.73
274 Chapter 5 Center of Gravity, Centroid, and Distributed Force

5.74 Determine the weight of the water the vessel would hold. The specific weight
of water is 62.4 lb/ft*.

5.75 Determine the surface area generated by revolving the area about the x axis.

5.76 Determine the volume generated by revolving the area about the y axis.

50mm jh 50mm
oe -

aN
mi a Le 60° shading
j 20in.
=70i 250mm
A SY \

a 200 eet
100 mm 100 mm
Figure P5.74 Figure P5.75 and P5.76

5.77 Determine the volume generated by revolving the area about the x axis.

5.78 Determine the volume generated by revolving the area about the y axis.

5.79 Determine the volume of the solid body.

5.80 Determine the volume of the body generated by revolving the area about the
y axis.

5.81 What will be the surface area of the body described in Problem 5.80?

Conical hole

Radius= 1.6 in.

5 tae “ ag
200 mm = diameter fee eSan ts in ——+}.—4 in—+
Figure P5.77 and P5.78 Figure P5.79 Figure P5.80 and P5.81
Problems 275

5.82 The exterior of the brass pulley, the cross-section of which is shown, is to be
nickel plated. The thickness of the nickel plating is to be 0.02 in. How much
nickel will be required?

5.83 The ratio of the volume (V) generated by rotating the semicircular area
through an angle of 180° about the x axis to the surface area (S) generated by rotating
the semicircular arc through an angle of 180° about the x axis is 0.25 m (that is,
V/S = 0.25 m). Determine the ratio y,/y>.

py
2in.
py
2 in.

Gu eadius= lin.

3 in.
EY :
6in

O
Figure P5.82 Figure P5.83

5.84 Let the volume generated by rotating the area about the x and y axes be V, and
V,, respectively. If V;/V. = 0.5, determine the ratio of H/L.

5.85 A thin shell is shown. Find the surface area of the inside of the shell in terms
of a and L.

Oh 1 ——-+ Z

Figure P5.84 Figure P5.85


276 Chapter 5 Center of Gravity, Centroid, and Distributed Force

SECTION 5.8*
5.86 through 5.107 Determine the reactions at the beam support(s) for the loading
condition shown.

240 Ib/ft

- -:
Figure P5.86 Figure P5.87

a
Figure P5.89

Figure P5.91

24kN/m
16 kN/m 500 N/m

Figure P5.93 Figure P5.94

1 ft
Figure P5.96 Figure P5.97
Problems 277

Ww w(x)= w sin(2=*)
30lb/in w(x) =30 cos(#>) tb/in.

x x

+6in-}e——I8 in. -—2)2—$+ 22


Figure P5.98 Figure P5.99

1000 Ib/ft
500 lb/ft
e Parabola

Af B 300 Ib/ft
- 15 ft - -} ———14 ft
Figure P5.100 Figure P5.101

|+-— 6.5 ft—+}+— 6.5 ft—+|


00 1b/ft
200 lb/ft

{A ft tt ft foots

+3 ft-}-—6ft—+}-—4 ft
Figure P5.102 Figure P5.103

s 6 ft—+ 1000 lb/ft


aie w=w,(1 - =) Parabola
wl, :
° L2

x
we abo

A
A
+122 —__-|

Figure P5.104 Figure P5.105

10kN/m
1000 lb/ft
500 Ib/ft

A/ - B

ees itis
Figure P5.106 Figure P5.107
278 Chapter 5 Center of Gravity, Centroid, and Distributed Force

5.108 For the beam shown, write an expression for the reaction at B in terms of
w and L.

2w
f]

L\A AB
peinspo we ee OS L—|
0.252
Figure P5.108

SECTION 5.9*
5.109 AB is a gate in a dam. The gate is hinged at A and measures 4 ft X 6 ft.
Determine the reaction (lb/ft) of the sill at B on the gate, given y,,ater = 62.4 Ib/ft®.

5.110 A vertical wall is retaining oil and water. Consider | m of the wall at a right
angle to the cross section shown. Determine the resultant hydrostatic force on the
wall and also the location of the line of action of the resultant, given y,,ate. =
9.81 kKN/m? and y,;, = 6.5 KN/m?.

Figure P5.109 Figure P5.110

5.111 The inclined sliding gate measures 2m X 2.5 m. If the couple reactions at
A and B are negligible, determine the reactions of the grooves at A and B due to the
water pressure, given Yyarer = 9.81 KN/m?.
Problems 279

5.112 Determine the magnitude of the resultant hydrostatic force per foot length of
the concrete dam shown.

5.113 For the concrete dam shown, determine the minimum value of 5 that will
prevent overturning of the dam. Given: Yate, = 62.4 lb/ft? and Y.oncrete =
150 Ib/ft?.

5.114 The cross section of the metal trough shown has a shape of a quarter of an
ellipse. It is hinged to the wall at the bottom. The top of the trough is attached to the
wall by cables which are 5 ft apart. Determine the tension in each cable, given that
Gea i, b= 2.5 it, and >.<, = 62.4 lb/ft.

6 ft
rrr
Figure P5.112 Figure P5.113 and P5.117 Figure P5.114

5.115 The shaft of the gate will fail at a moment of 60 x 10° ft - lb. Determine
the maximum value of the water height 4, given the width of the gate as 5 ft and
Ywater = 62.4 Ib/ft*. The gate just touches the concrete structure at A.

5.116 The gate AB is 5 m wide and is hinged at A. Determine (a) the horizontal
and vertical reactions at A, and (5) the vertical reaction at B. Given: Y\ater =
9.81 kKN/m?.

Smooth
surface

cae este Behe Series

Figure P5.115 Figure P5.116


280 Chapter 5 Center of Gravity, Centroid, and Distributed Force

5.117 Refer to Figure P5.113. Given b = 12 ft. Consider 1 ft length of the dam.
Determine the reaction forces exerted by the ground at the toe, C, of the dam.

SECTION 5.10
5.118 through 5.123 By direct integration, locate the centroidal coordinate x for the
volume generated by revolving the following areas about the x axis through an angle
0 = 27 radians.

5.118 Area shown in Figure PS.2.

5.119 Area shown in Figure PS.3.

5.120 Area shown in Figure P5.6.

5.121 Area shown in Figure PS5.7.

5.122 Area shown in Figure P5.12.

5.123 Area shown in Figure P5.14.

5.124 through 5.129 Determine the centroidal coordinate y for the volumes shown.
Use the integration method.

Figure P5.124

Zz

Figure P5.125
Problems 281

Half cone

Zz Zz
Figure P5.126 Figure P5.127
y
Radius=R

Zz B
Figure P5.128 Figure P5.129
5.130 Determine the centroidal coordinate z for the volume.

5.131 Determine the centroidal coordinate y for the conical shell.

Figure P5.131
282 Chapter 5 Center of Gravity, Centroid, and Distributed Force

SECTION 5.11
5.132 through 5.140 Determine the centroidal coordinates x, y, z of the composite
bodies shown.

Diameter = 3 in. >“ 100mm

Figure P5.132 . Figure P5.133

Radius = 100 mm

ay.

50 mm diameter

600 mm

el b x

200 mm radius

Zz Fi
Figure P5.134 Figure P5.135
Problems 283

2 in. 10 in.

Hole through the block

11,5) ita,
ie
Leonie

Zz
Figure P5.136 Figure P5.137

if y
6 in. diameter

Z
Figure P5.139

6 in. diameter

Figure P5.138

Figure P5.140

es
284 Chapter 5 Center of Gravity, Centroid, and Distributed Force

5.141 Refer to Figure P5.41. Determine the centroidal coordinates of the volume
generated by the shaded area when revolved through 27 radians about the y axis.

5.142 Refer to Figure P5.44. Determine the centroidal coordinates of the volume
generated by the shaded area when revolved through 27 radians about the x axis.

5.143 Determine the centroidal coordinates of the volume of the hemispherical shell
for cells R, ~ R).

Figure P5.143

5.144 through 5.146 Determine the centroidal coordinates of the homogeneous wire
shown.

Figure P5.144

|
|
|
|
|
|
|
|
|
|
|
Radius = 50mm v
Figure P5.145 Figure P5.146
Problems 285

5.147 The assembly shown is made of two thin uniform plates (OABC and OAD)
and three thin uniform wires (BE, CE, and DE). Determine the centroidal coordinates
of the assembly. Assume the thickness of the plates and the radius of the wires are
1 mm each.

Figure P5.147
CHAPTER

ANALYSIS OF STATICALLY
DETERMINATE
STRUCTURES

OUTLINE 6.1 Introduction


6.2 Internal Forces at Connections
6.3. Trusses
6.4 Assumptions for the Analysis of Trusses
6.5 Arrangement of Members of Simple Plane Trusses
6.6 Statically Determinate Plane Trusses
6.7 Analysis of Plane Trusses by the Method of Joints
6.8 Analysis of Plane Trusses by the Method of Sections
*6.9 Space Trusses
6.10 Frames and Machines
6.11 Summary
Key Terms
Problems

286
6.2 Internal Forces at Connections 287

6.1 INTRODUCTION
In Chapters 3 and 4, we discussed how to determine the external forces and
reactions that are necessary to keep a single rigid body in equilibrium. Even
when the structure was made of two or more members (parts), we still treated
the entire structure as a single rigid body, and calculated forces that were
external to the structure.
In this chapter, we will consider the analysis of three common types of
engineering structures: trusses, frames, and machines. These structures are
generally composed of several members. In order to design these structures,
we need to know, in addition to the forces that are external to the entire struc-
ture, the forces that act on the individual members of the structure at the
connections and hold the entire assembly together. These forces are called
internal forces. The procedure for determining such internal forces is based
on a simple extension of the concept of equilibrium of rigid bodies presented
in Chapters 3 and 4. The procedure is based on the idea that if the structure
is in equilibrium, then the portions of the structure (containing one or more
members) are also in equilibrium; and on the application of Newton’s third
law, which states that for each action there is an equal and opposite reaction.
The objective of this chapter is to develop the analysis of statically deter-
minate trusses, frames, and machines, using the concept of equilibrium and
the application of Newton’s third law.

6.2 INTERNAL FORCES AT CONNECTIONS


Internal forces are simply defined as the forces exerted on a portion or a
member of a structure by the rest of the structure. In this chapter, we will focus
our attention on the internal forces that are exerted on a portion or a member
at the joints (connections). In Chapter 7, we will consider internal forces that
develop within the members.
Consider first the plane framed structure shown in Figure 6.1a. The frame
is composed of three members AB, BC, and CD. Members AB and BC are
connected by a frictionless hinged connection at joint B, and members BC and
CD are connected by a rigid (moment-resisting) connection at joint C. The
free-body diagram of the entire frame in Figure 6.15 shows the known external
forces P;, P;, and P3, as well as the unknown external reactions A, and A,,
and D, and D,, at supports A and D, respectively. Note that this free-body
diagram does not show the internal forces being transmitted through the con-
nections at joints B and C.
The free-body diagrams of the individual members and joints of the disas-
sembled frame are shown in Figures 6.1c through 6.1g. The diagram of mem-
ber AB (Figure 6.1c) shows, in addition to the known external forces P, and
P,, and support reactions A, and A,, the unknown internal forces B*® and
B*8, exerted upon member AB by the rest of the structure through the hinged
connection at joint B. (Recall from Section 3.16 that a hinge is capable of
288 Chapter 6 Analysis of Statically Determinate Structures

{?: Hinged connection

Rigid
connection

Figure 6.1

exerting two force components or a single force of unknown direction, but no


couple.) Note that without B4" and B??, member AB will not be in equilib-
rium. Member AB is, therefore, being supported at joint B. It is important to
realize that the term internal as used here means that the forces BY” and Bi)?
are internal as far as the entire assembled structure is concerned. As for mem-
ber AB, these forces are external to it, and must be shown on its free-body
diagram.
The free-body diagram of joint B (Figure 6.1d) shows the same internal
forces BY” and BY", but in opposite directions. This is in accordance with
Newton’s third law. While joint B is exerting forces B?? and B?? on member
AB, member AB is exerting the same forces, but in opposite directions, on
joint B. Also shown on joint B are the forces B?° and B?° being exerted on
it by member BC.
6.2 Internal Forces at Connections 289

Member BC of the structure is connected to two joints: the hinged joint B


on the left and the rigid joint C on the right; its free-body diagram, therefore,
shows the internal forces exerted on this member at ends B and C. Recall that
joint C, being a rigid joint, is capable of exerting two force components,
C®° and C®°, and a couple Mz-, on the member BC. These same forces and
couple arein turn being exerted by member BC on joint C with the senses
reversed (Figure 6.1). The free-body diagram of joint C also shows the known
external force P; acting on it. Note that on all of the free-body diagrams of
members and joints, the law of action and reaction between bodies in contact
has been followed. Of course, we generally do not know the correct senses of
the internal forces before the analysis, and must choose them arbitrarily.
Nevertheless, it is essential to show these forces in a consistent way (i.e., with
equal magnitudes but opposite directions) on the free-body diagrams involved.
As we will discuss in detail in Sections 6.6 to 6.10, we calculate the mag-
nitudes and directions of the internal forces by writing the equations of equi-
librium for each free-body diagram, and solving the equations for the unknown
internal forces. If an internal force turns out to be negative, then it acts in the
sense opposite to that originally assumed.
When a joint has only two members connected to it and no external loads
acting directly on it, the free-body diagram of that joint may be omitted in the
analysis, and the internal forces may be transmitted directly from one member
to the other with the directions reversed. This can be observed from Figure
6.ld. By applying the equations of equilibrium to joint B, we find that
BA® = BPC and BA? = B®; that is, the internal forces at B on member AB
are equal and opposite to the internal forces at B on member BC.
As another example, consider the three-member plane structure shown in
Figure 6.2a. The three members of the structure are connected together by
frictionless hinged connections at joints A, B, and C. The free-body diagram
of the entire structure, in Figure 6.2b, shows the known external force P acting
at joint B, and unknown external reactions A, and A,, at the hinged support A
and C,, at the roller support C. No internal forces are shown on this diagram
of the whole structure.
The free-body diagram of member AB (Figure 6.2c) shows the internal
forces A#? and A“", and B24? and B#%, exerted at its ends by the hinged joints
A and B, respectively. The two components of internal forces at each end of
the member can be replaced by a single resultant force, shown at member ends
A and B as F48 and F 3°, respectively, in Figure 6.2c. As no external loads or
other internal forces are acting on this member, it is in equilibrium under the
action of these two (internal) forces F4? and F4°. Recall from Section 3.15
that such a member is called a two-force member, and that the two forces
F48 and F4? must be equal, opposite, and collinear, with their lines of action
passing through the member ends A and B. Member AB is therefore subjected
to axial forces. The free-body diagrams of all the members and joints of the
dismembered structure are shown in Figures 6.2d through 6.2i. In Figure 6.2d,
for member AB, F4? and F#? are replaced by the equal and opposite forces
290 Chapter 6 Analysis of Statically Determinate Structures

B
m ©
Ave A
Ay eae
(db) (c)

V2
fap A ee
(g)

BK Fas Fae p

Figure 6.2

F,,. As shown in Figures 6.2f and 6.2h, members AC and BC are also two-
force members because they are connected to hinged joints at their two ends
and are not subjected to any external or other internal forces. Again, the correct
senses of the internal forces in two-force members are generally not known
before the analysis, and are arbitrarily assigned; but the two forces must be
equal, opposite, and collinear, with their lines of action passing through the
two points at which the member is connected.
6.2 Internal Forces at Connections 291

EXAMPLE 6.1

Identify all two-force members in the structure shown. Also, draw


the free-body diagram of the entire structure, and the free-body
diagrams of each member and each joint of the structure.

Solution

Two-force members. From the given figure we see that


member AC is connected at only two points—A and C—to
hinged joints, and is not subjected to any external load.
Similarly, member AD is connected at only two hinged joints,
A and D, and has no other load acting on it. Therefore,
members AC and AD are both two-force members.

Free-body diagrams. The free-body diagram of the entire


structure is shown. Note that it gives only the forces that are
external to the entire structure: the external loads P, M, and w,
and the external support reactions A,, A,, and B,. None of
the internal forces are shown on this diagram. —
The free-body diagrams of each member and joint of the
structure are shown. Since members AC and AD are two-force
members, each is subjected to only two forces, which are equal,
opposite, and collinear, with their lines of action passing
through the two points at which the member is connected. The
Cm P D@

GD

CyCD

(continued)
292 Chapter 6 Analysis of Statically Determinate Structures

EXAMPLE 6.1 (concluded)

remaining three members, AB, BD, and CD, are multiforce The diagrams of the joints show the same internal forces
members; their free-body diagrams show, in addition to acting on the members, but in opposite directions in conformance
the external loads and any support reactions, two internal force with Newton’s third law. The correct senses of the internal
components at each point where a member is connected by a forces and support reactions are not known and, therefore, have
hinged connection. been arbitrarily assumed on the diagrams.

EXAMPLE 6.2

Identify all two-force members in the structure shown. Also,


draw free-body diagrams of each member and each joint of the
structure.

Solution

Two-force members. All of the members of the structure can


be classified as two-force members, because the external
loads are acting only at the joints, and each member is
connected at its ends to only two hinged joints.

Free-body diagrams. The free-body diagrams of the four


joints and the five members of the structure are shown. Since
each member is a two-force member connected at its ends, it is
subjected to two equal, opposite, and collinear forces at its
ends. The free-body diagrams of the joints also show these
member forces, but in opposite directions.

P, Py

|Fao Fo Fp Foo D
aa I a arg rr 3

ai € Fan Fp
Fac
Fup
F'4c B Fep
G D

A
A B

Fae ar Fp

A> —E eee <-> P,


A Fy, Fz ae a
A B,
=

‘NZ
EXAMPLE 6.3
P, P,
Draw a free-body diagram of the portion ACE of the structure
shown. | Ee G
E ETT P,
Solution The free-body diagram shows all the forces acting on WA
the portion ACE of the structure: the external load P, at joint
E; two support reactions, A, and A,, at joint A; and the ™
three member forces, F4,, For, and Fp, being exerted by
members AB, CF, and FF at joints A, C, and E, respectively.
Note that these three member forces are acting in the directions
of the corresponding members. This is because members AB,
CF, and EF are all straight two-force members, each subjected
to two equal, opposite, and collinear forces at its ends. Also
note that the free-body diagram of portion ACE does not show
forces in member AC, AE, or CE, as these forces are internal to For
the portion ACE.

EXAMPLE 6.4

Draw free-body diagrams of the members and the joint of the


pliers shown.

Solution The externally applied loads, P,, cause the jaws of


the pliers to exert forces P, on the wire as shown on the
free-body diagram of the wire A. These wire reactions, P,, must
then be applied in opposite directions on the free-body
diagrams of the pliers’ jaws at points B and E as shown in the
figure.

293
294 Chapter 6 Analysis of Statically Determinate Structures

6.3 TRUSSES
A truss is formed by connecting straight bars (members) at their ends by
flexible connections to form a stable configuration—a rigid body. Trusses are
among the most commonly used structures; because of their light weight and
high strength, they are used in a variety of engineering applications ranging
from supports for bridges and roofs of buildings to support structures in space
stations.
The types of members used to construct modern trusses usually consist of
steel I-shapes, channels, and angles; steel or aluminum tubes; and wood struts;
they are joined by using bolted or welded connections. As the members are
generally slender, external loads and reactions are applied only at the joints.

Plane and Space Trusses .


If all the members of a truss, as well as the applied loads, lie in a single plane,
the truss is called a plane truss, and can be analyzed as a two-dimensional
structure. Plane trusses are commonly used for supporting bridges and roofs.
Most truss bridges consist of two plane trusses joined together by floor beams,
as shown in Figure 6.3, to form a three-dimensional structure. The weight of
the deck (usually a concrete slab for highway bridges), the stringers, the floor
beams, and the traffic is transmitted by the floor beams to the supporting trusses
at their joints. As this applied loading acts on each truss in its own plane, the
trusses can be treated as two-dimensional structures subjected to coplanar force
systems. Some common types of plane trusses used as bridge trusses are
shown in Figure 6.4.
A typical framing system for a roof is shown in Figure 6.5. In this case,
two or more plane trusses are connected at their joints by members, called
purlins, to form a three-dimensional structure. The roof is attached to the pur-
lins, which transfer the weight of the roof (and any other load due to snow,
etc.), as well as their own weight, to the supporting roof trusses at the joints.
This applied loading acts on each truss in its own plane which, therefore, can

Figure 6.3 Bridge truss


Parker t

nS \,
LA ZN LAD
Caen Cpa’ Cia ea

Baltimore truss
Cs OR
[\
OO

iB
Z K < > EIN
ZI.NA| AN
296 Chapter 6 Analysis of Statically Determinate Structures

Figure 6.5 Roof truss

be analyzed as a two-dimensional body subjected to a coplanar force system.


Some commonly used types of plane roof trusses are shown in Figure 6.6.
Although a great majority of three-dimensional structural systems can be
subdivided into plane or two-dimensional structures for the purposes of ana-
lyzing and designing, some truss systems, such as domed roof structures and
transmission towers, cannot be subdivided into planar components, due to their
shape, arrangement of members, or applied loading. These trusses—called
space trusses—are analyzed as three-dimensional structures subjected to three-
dimensional force systems.

6.4 ASSUMPTIONS FOR THE ANALYSIS OF TRUSSES


To simplify the analysis and design of trusses, it is commonly assumed that:

1. All members of the truss are connected only at their ends, by frictionless
hinged connections in plane trusses, and by ball-and-socket joints in
space trusses.
2. All external loads and support reactions are applied only at the joints of
the truss.
The effect of these simplifying assumptions is that all members of the truss
act as two-force members; thus, each member will be either in axial tension
(being pulled apart, as shown in Figure 6.7qa) or in axial compression (being
pushed in, as shown in Figure 6.75).
In real structures, joints are usually formed by connecting members to a
gusset plate with riveted, bolted, or welded connections. Also, some members
of a truss may be continuous at the joints. However, if the centroidal axes of
members are concurrent at a joint, as shown in Figure 6.8, the assumption of
hinged joints gives results that are generally satisfactory for design purposes.
The second assumption is valid for most trusses. As we discussed, the
external loads are usually transmitted to the trusses at joints by such means as
floor beams and purlins. The dead (self) weight of a well-designed truss is
usually small compared with the external load it is designed to support. This
6.4 Assumptions for the Analysis of Trusses 297

lax
ZN el
King post truss

iP G

li G
(a) Axial (b) Axial
tension compression
Figure 6.7 Two-force members

ee plate
Fink truss
Figure 6.6 Roof trusses

dead weight is either neglected in the analysis, or its effect approximated by


applying half of the weight of each member, as a concentrated load, to the
joint at each end of the member.
The analysis of plane trusses will be presented in Sections 6.5 through 6.8,
and that of space trusses in Section 6.9. The term analysis as used here means
the determination of the forces in the various members of a truss for a given Figure 6.8
298 Chapter 6 Analysis of Statically Determinate Structures

loading. The member forces thus determined can then be used to select ap-
propriate sizes for the members, and to design the connections of the truss.

6.5 ARRANGEMENT OF MEMBERS


OF SIMPLE PLANE TRUSSES
A plane truss is considered to be rigid if it does not change shape when sub-
jected to a general system of forces at its joints. The truss must maintain its
shape and remain a rigid body even when it is detached from the supports.
The simplest rigid plane truss can be constructed by connecting three mem-
bers at their ends by hinges to form a triangle as shown in Figure 6.9a. This
triangular truss is referred to as the basic truss element. Note that this triangular
truss is rigid in the sense that it does not change its shape under load. In
contrast, a rectangular truss constructed by connecting four members at their
ends by hinges, shown in Figure 6.9), is nonrigid because it may change its
shape, and collapse, when subjected to a general system of coplanar forces at
its joints.

(a)
Figure 6.9

The basic truss ABC of Figure 6.10a can be extended as shown in Figure
6.106 by attaching two new members, BD and CD, to the two existing joints,
B and C, and connecting them to form a new joint D. Provided that the new
joint D does not lie on the straight line passing through the two existing joints
B and C, the new truss will be rigid. The truss can be further extended by
repeating the same procedure (see Figure 6.10c) as many times as needed. A
truss constructed by this procedure is called a simple truss.
Some commonly used simple plane trusses are shown in Figure 6.11. The
reader is urged to examine each truss to verify it as a simple truss. The basic
truss element is identified as shaded on each truss.
Since a simple truss is formed from a basic truss element (which contains
three members and three joints) by adding two new members for each new
joint, the total number of members in a simple truss is m = 3 + 2(j — 3),
6.5 Arrangement of Members of Simple Plane Trusses 299

a G New member
D (New joint)

A @Bazimn
one ae =)B A
B

Figure 6.10

Figure 6.11

(continued)
300 Chapter 6 Analysis of Statically Determinate Structures

Figure 6.11 (continued)

or

m=2j-3 (6.1)
where j is the total number of joints, including those attached to the supports.

6.6 STATICALLY DETERMINATE PLANE TRUSSES


A plane truss is considered to be statically determinate if all the external
reactions, as well as the forces in all of its members, can be determined by
applying the equations of equilibrium.
Consider the plane truss of Figure 6.12a, composed of five members con-
nected together by four hinged joints. The free-body diagrams of the members
and the joints are shown in Figure 6.12b. As each member is a two-force
member, it is subjected to two equal-but-opposite collinear forces at its ends.
The free-body diagrams of each of the members show the member forces as
pulling on the members, indicating that all members have been assumed to be
in tension. Note that the free-body diagrams of the joints show the same mem-
ber forces but in opposite directions.
The analysis of the truss involves determining the magnitudes of five mem-
ber forces, Fz, Fec, Fan, Fgp, and Fep, and three external reactions, A,, A,,
and C,. (Note that the lines of action of the member forces and reactions are
known; only their magnitudes are unknown.) Thus, a total of eight unknown
quantities are to be determined. Since the whole truss is in equilibrium, each
joint must also be in equilibrium. We can see from the free-body diagrams of
the joints (Figure 6.12b) that at each joint, the external and internal forces
form a concurrent and coplanar force system, requiring that the two equilib-
rium equations {F, = 0 and &F, = 0 must be satisfied. As the truss has four
joints, the total number of equilibrium equations available is 2(4) = 8, which
6.6 Statically Determinate Plane Trusses 301

DEP,
Fp oes
AD, Fapk Fép
D D D

A B C
: F'aD oo CD
"AD Fp
A Fag Be IB BD F F ze N o
INS Re ae ek BC
{Fas A F4p P BEao B Cae BC {
y 2 Cy

(b)
Figure 6.12

can be solved to determine the eight unknowns. Therefore, the truss is statically
determinate.
It should be realized that three equations of equilibrium for the whole truss
could be used to determine the three unknown reaction components A,, A,,,
and C,. However, these three equations are not independent from the equilib-
rium equations of the joints, and do not provide any additional information.
Based on the foregoing discussion, we can develop the criterion for the
static determinacy of general plane trusses containing m members and j joints,
and supported by 7 number of external-reaction components. For the analysis
of the truss, we need to calculate m member forces and r reaction components;
thus, we determine a total of m + r unknown quantities. As the truss contains
j joints, and we can write two equilibrium equations, 2F, = 0 and 2F, = 0,
for each joint, the total number of equations of equilibrium available is 2). If
the number of unknowns for a truss is equal to the number of equilibrium
equations (m + r = 2j), then all of the unknowns can be determined by
solving the equilibrium equations and the truss is statically determinate.
If a truss has more unknowns than the available equations of equilibrium
(m + r > 2j), then all of the unknowns cannot be determined by solving the
available equilibrium equations. Such a truss is considered statically indeter-
minate. Statically indeterminate trusses have more members than the minimum
required for rigidity, and/or more reactions than the minimum required for
302 Chapter 6 Analysis of Statically Determinate Structures

equilibrium. The analysis of statically indeterminate structures is beyond the


scope of this textbook and will not be considered herein.
If a truss has less unknowns than the number of joint equilibrium equations
(m + r < 2j), then the truss is considered statically unstable. Static instability
may be due to a truss having fewer members than the minimum required for
rigidity, and/or fewer reactions than the minimum required for equilibrium.
The relationships given above for static determinacy, indeterminacy, and
instability of plane trusses are commonly written in the following form:

m<2j—r __ truss is statically unstable


m=2j —r_ truss is statically determinate (6.2)
m>2j—r truss is statically indeterminate

The first condition—for static instability—is both necessary and sufficient in


that if m < 27 — r, the truss is definitely statically unstable. However, the
remaining two conditions—for static determinacy and indeterminacy—are
necessary, but not sufficient, conditions. A truss may have a sufficient number
of members and reactions, but may still be unstable due to improper arrange-
ment of members and/or supports. In order for the criterion for static deter-
minacy as given by Equation (6.2) to be valid, the truss must act as a rigid
body in equilibrium under a general system of coplanar forces when attached
to the supports. Examples of statically determinate, indeterminate, and unstable
trusses are shown in Figure 6.13.

m=7; 7=5; r=3; m=2j7—r m=8; j=5; r=3; m>2j—r

(a) Statically determinate (b) Statically indeterminate

m=6; j=5; r=3; m< jr


(c) Statically unstable

Figure 6.13
6.7 Analysis of Plane Trusses by the Method of Joints 303

6.7 ANALYSIS OF PLANE TRUSSES


BY THE METHOD OF JOINTS
In the method of joints, the forces in the members of a truss are determined
by considering the equilibrium of its joints. As the whole truss is in equilib-
rium, each of its joints must also be in equilibrium. At each joint, the member
forces and any external loads and support reactions form a concurrent coplanar
force system. For each joint to be in equilibrium, the concurrent coplanar force
system acting on it must satisfy the two equations of equilibrium {F, = 0 and
DHif = 0. These two equations of equilibrium must be satisfied at each joint
of the truss. Since there are only two equations of equilibrium at a joint, they
can be used to determine a maximum of two unknown forces.
The method of joints consists of selecting a joint with no more than two
unknown, noncollinear forces acting on it, drawing its free-body diagram,
writing the two equilibrium equations YF, = 0 and {F, = 0, and solving
them to determine the unknown forces. This procedure is repeated until all the
desired forces have been determined. While all the unknown member forces
and all the external reactions can be determined by using the joint equilibrium
equations, in many trusses it may not be possible to find a joint with two or
less unknown forces to begin the method of joints, unless the external reactions
are known beforehand. For these, the external reactions are computed from
the equations of equilibrium of the whole truss before proceeding with the
method of joints to determine the member forces.
To illustrate the method of joints, consider the Warren truss shown in Figure
6.14a. The free-body diagram of the entire truss is shown in Figure 6.145, and
the free-body diagrams of all the members and joints are shown in Figure
6.14c. Since the member forces are not known at this stage of the analysis,
the sense of axial forces (tension or compression) in the various members have
been arbitrarily assumed. Members AB, BC, BD, BE, and CE are shown to be
in tension with axial forces tending to pull the members apart, whereas mem-
bers AD and DE are assumed to be in compression with axial forces pushing
into the members. The free-body diagrams of the joints show all the member
forces acting on the joints in directions opposite to their directions on the
member ends, in accordance with Newton’s third law. In the free-body diagram
of joint A, we observe that the tensile force F,, is pulling away from the joint,
whereas the compressive force F,, is pushing toward the joint. This effect—
members in tension pulling on the joint, and members in compression pushing
on the joint—can be seen on the free-body diagrams of all the joints. Since
the free-body diagrams of members are usually omitted in the analysis of
trusses by this method, and only those of the joints are drawn, it is important
to understand that a tensile member force is indicated on the free-body diagram
of a joint by an arrow pulling away from the joint, and a compressive member
force is indicated on the free-body diagram of a joint by an arrow pushing
toward the joint, provided that the arrow is located on the same side of the
joint as the member itself.
304 Chapter 6 Analysis of Statically Determinate Structures

To begin the analysis, we should select a joint with two or fewer noncol-
linear unknown forces acting on it. An examination of the free-body diagrams
of the joints in Figure 6.14c indicates that none of the joints satisfies these
criteria. Therefore, we determine the external reactions from the equilibrium
equations applied to the free-body diagram of the whole truss (Figure 6.145):

QF, = 0 ae. — 0

@=M,=0 -10(6) +C,(12) = Cy = 5 kip®


Obie 20 A,-10+5=0 A,=5kipQ)
E D E

4 ft
A
G may, &
10 kip Cy
me 3 ft+}-3 ft+h—3 ft ieft: ae ft Bs:
(a)

D Fg Fg Fg Foe E
<G——— 7

i. 7 ue e F F cg

ie Fe
E

r Fog
Fp

mms.“
At Ee, Fag |B Fac
Ay
10 kip

(c)

4 -
f eae aD | Fee
|
VS 53.13 Wd 531138
== SSS i ; Se 3
Faz 53135 53.18; =3. Fc
A, =5 kip FpAD = 6.25
5°: ki p F ep
(d) (e)
Figure 6.14

(continued)
6.7 Analysis of Plane Trusses by the Method of Joints
305

For = 6.25 kip »


| 53.13° A |
Pe 7.5 kip fee Seis
eee Beye Fac =3.75 kip
Re C, =5 kip
F pp = 6.25 kip (h)

(g)
D 7.5kip(C)

A C
{ 3.75 kip(T) LE 3.75 kip (T) t
5 kip 10 kip 5 kip
(i)
Figure 6.14 (continued)

Having computed the reactions, we can now begin the analysis at either
joint A, which now has two unknown forces Fy, and F,p, or joint C, which
also has two unknowns, Fx and Fog. We begin with joint A, whose free-body
diagram is shown in Figure 6.14d. From the dimensions of the truss, we find
that all inclined members are at an angle of 53.13° with the horizontal. The
unknown forces F4, and F4, can be determined from the two equilibrium
equations 2F, = 0 and }F, = 0, for the forces at joint A. Thus,

@Ox5, =U Qitsen eer 70


Sheeday Fay = 6:25 kip (C)
@QZF.=0 —6.25(cos 53.13°)+ Fig =0 Fag = 3.75 kip (T)
in which T denotes tension and C denotes compression. The equilibrium equa-
tion >F, = 0 was used before YF, = 0 because it contains only one unknown,
Fp. The value of Fy) = +6.25 kip is then used in the equation }F, = 0 to
obtain the answer for F,, = +3.75 kip. Note that the answers for both F,,
and Fz are positive; this indicates that the assumed senses of axial forces in
these members (tension in AB and compression in AD) were correct. If the
answer for a member was negative, its correct sense would have been the
reverse of that initially assumed. Before proceeding to the next joint, it is
common practice to write the values of the member forces, as they are cal-
culated, on a line diagram of the truss as shown in Figure 6.147.
Next, we draw a free-body diagram of joint D as shown in Figure 6.14e,
and determine Fz, and Fp, as follows:
®XF,=0 6.25(sin 53.13°) — Fgp(sin 53.13°) = 0
Fpp = 6.25 kip (T)
Q@SF,=0 6.25(cos 53.13°) + 6.25(cos 53.13°) — Fpg = 0
Fp -5 kip (©)
306 Chapter 6 Analysis of Statically Determinate Structures

Again, these latest member forces are recorded on the truss diagram in
Figure 6.147.
Using the free-body diagram ofjoint B (Figure 6.14f):

@ZF,=0 —10 + 6.25(sin 53.13°) + Fyz(sin 53.13°) = 0


Fp = 6.25 kip (T)
@SF.=0 3.75 — 6.25(cos 53.13°) + 6.25(cos 53.13°) + Fac = 0
Fac = 3.75 kip (T)
The free-body diagram of joint F is shown in Figure 6.14g. The only un-
known force acting at this joint, F-,, can be determined by using either of the
two equations of equilibrium:

QF, =0 7.5 — 6.25(cos 53.13°) + Fog(cos 53.13°) = 0


Fop = —6.25 kip
The negative answer indicates that the member CE is in compression rather
than tension as assumed originally. Therefore,

Fog = 6.25 kip (C)


Since we have determined all the member forces in the truss, the three
remaining joint equilibrium equations ({F, = 0 at joint E, 2F, = 0 and
LF, = 0 at joint C) can be used to check our calculations. Thus, at joint E

QF, =0 —6.25(sin 53.13°) + 6.25(sin 53.13°) = 0 checks

and finally, at joint C (Figure 6.14h),

QF. =0 —3.75 + 6.25(cos 53.13°)=0 checks


Qn S0 5 — 6.25(sin 53.13°9)=0 — checks

Zero-Force Members
Since trusses are often designed to support several different types of loads, it
is common to find members with zero forces when analyzing a truss subjected
to a particular loading condition. Zero-force members are also sometimes
added to trusses to prevent buckling of compression members, to brace slender
tensile members, and to maintain the rigidity of the truss.
The analysis of trusses can be considerably simplified if we can identify
the zero-force members by inspection (i.e., without drawing the joints’ free-
body diagrams and applying the equations of equilibrium). The two arrange-
ments that result in zero-force members are the following:
6.7 Analysis of Plane Trusses by the Method of Joints 307

When only two


noncollinear
members are
connected to a joint
and no external load
or support reaction
is applied to the
joint, then the force
in each member is
zero.

Figure 6.15a shows two noncollinear members connected at joint A, with no


applied load or support reaction. Summation of the forces in the y direction
yields

LF, =0 F,-(sin 8) = 0

Since the members are not collinear, 6 is not zero; therefore, F,- must be zero.
Summation of the forces in the x direction shows that F,, must also be zero.

(a)
When three
members are
connected to a joint,
two of which are —
collinear, and no
external load or
support reaction is
applied to the joint,
the force in the
third, noncollinear —
-member is zero. _

An arrangement of this type is shown in Figure 6.15). We can see that since
no external load is balancing the x component of F',,, the equilibrium equation (b)
>F. = 0 can be satisfied only if Fg is zero. Figure 6.15
The foregoing criteria are applied to identify zero-force members of the
two trusses shown in Figures 6.16a and 6.16b. The zero-force members are
indicated by red lines. Considering first the tower truss subjected to lateral
wind loading in Figure 6.16a, we see that two noncollinear members FH and
GH are connected at joint H to which no external load is applied; therefore,
members FH and GH are zero-force members.
308 Chapter 6 Analysis of Statically Determinate Structures

(a)

(b)
Figure 6.16

Next, consider the roof truss subjected to partial snow loading shown in
Figure 6.16b. We see that at joint K, three members—J/K, KL, and KO—are
connected; of these, JK and KL are collinear, whereas KO is not. Since no
external load is applied at joint K, member KO is a zero-force member. A
similar inspection at joint M identifies member EM as a zero-force member.
Next, consider joint O, where four members—DO, JO, KO, and LO—are
connected and no external load is applied. We have already identified KO as
a zero-force member. Of the three remaining members, DO and JO are colli-
near; therefore, LO must be a zero-force member. Similarly, at joint E, member
EL is identified as a zero-force member; similar reasoning is then used at joint
L to identify DL as a zero-force member.

Procedure for Analysis


The following step-by-step procedure can be used for analyzing most common
types of trusses by the method of joints.
6.7 Analysis of Plane Trusses by the Method of Joints 309

1. Determine by inspection any zero-force members of the truss.


2. Select a joint having no more than two unknown noncollinear forces acting
on it. If such a joint is found, go directly to Step 4. If not, go to Step 3 to
determine support reactions, then select a joint with not more than two un-
known forces, and then go to Step 4.
3. Determine support reactions. Draw the free-body diagram of the entire
truss showing all external loads and support reactions. (Do not show internal
member forces.) Write the three equations of equilibrium 2F, = 0, 2F, = 0,
and +M = 0, and solve them to determine support reactions.
4. Draw a free-body diagram of the selected joint (with two or less un-
knowns), showing all known and unknown forces acting on it. The sense of
the known forces must be correctly indicated on the free-body diagram. A
tensile member force is shown by an arrow pulling away from the joint,
whereas a compressive member force is indicated by an arrow pushing into
the joint. The unknown member forces are usually assumed to be tensile and
are therefore indicated on the free-body diagram by arrows pulling away from
the joint.
5. Write the two equilibrium equations }F, = 0 and 2F, = 0, and solve
them to determine the unknown forces. A positive answer for a member force
means that the assumed direction is correct, whereas a negative answer means
that it is incorrect.
6. If all the desired member forces and reactions have been determined, pro-
ceed to Step 7. If not, select another joint with two or less unknowns and go
back to Step 4.
7. If the support reactions were calculated in Step 3 from the equilibrium
equations of the entire truss, use the three remaining joint equilibrium equa-
tions to check the computations. If the reactions were determined from the
joint equilibrium equations, draw a free-body diagram of the entire truss and
apply the three equilibrium equations ({F, = 0, {F, = 0, and 2M = 0) to
check the computations. If the analysis has been carried out correctly, the
equilibrium equations (of the joints or of the entire truss as a rigid body) will
be satisfied.
310 Chapter 6 Analysis of Statically Determinate Structures

EXAMPLE 6.5

Determine the force in each member of the truss shown, by the


method of joints.

Solution

Zero-force members. At joint F, only two members—DF and


EF, which are not collinear—are connected, and no external
load is applied to this joint. Therefore, members DF and EF are
zero-force members. This result could also be obtained by
considering the equilibrium of joint F.

Joint F. The free-body diagram of this joint is shown. Both


20 ft unknown forces Fp, and Fz, are assumed to be tensile. The
10 ft 10 ft
angle of inclination of Fp, is 75.96° with respect to the x axis.
This angle, shown on the free-body diagram of the joint, is
obtained from the dimensions of the truss. The two unknowns
a

are determined from the equilibrium equations; thus


||
OxA, =0 — Frp(sin 75.96°) = 0 Fyp=0 <
Pep
LEdae a @zxF,=0 —Fep = 0 Fu =00
75.96°
Joint E. Joint E is analyzed next. Three member forces Foz,
Ppp Fpp, and Fry act on this joint; Fj, is already known. The
two unknowns can be obtained from

@sF.=0
10 — Fo,(cos 75.96°) + Fp,(cos 38.66°) = 0
Foe = 0.31 Fo, — 12.81
®-=F,=0
—5 — F,,(sin 75.96°) — Fp,(sin 38.66°) = 0
Using the expression for Fp, and simplifying, we obtain

3 — 1.16For = 0

For, = 2.58 kip (T) <


Using the answer for For,
Fyre = — 12.01 kip
Fp = 12.01 kip (C) <
Note that on the free-body diagram of joint E, both unknown
forces were assumed to be tensile. The positive answer for Fo;
indicates that Fc, is indeed tensile. The negative answer for Fp,
indicates that F,, is actually compressive.

(continued)
6.7 Analysis of Plane Trusses by the Method of Joints 311

EXAMPLE 6.5 (continued)

Joint D. This joint now has two unknown forces, Fg, and Fop, v
which we determine as follows:

Oxr, =0 re
F =

— 12.01(sin 38.66°) — F,,(sin 75.96°) = 0


Fap = —7.73 kip
Fp = 7.73 kip (C) <
@zF, =0
—Fep + 12.01(cos 38.66°) — 7.73(cos 75.96°) = 0
Fop = 7.50 kip (T) <

Joint C. Joint C now has only two unknowns, F,¢ and Fx, Fog = 2.58 kip
which we determine in the same way as before; thus

@rF,=0 oun,
29.74°
2047-5 1 2.98(Cos| 75.96-) — Fac(cos 75.96")
Fc
+ Fp(cos 29.74°) = 0
Fac = 3.58Fgc + 115.94
®=sF, =0
2.58(sin 75.96°) — Fyc(sin 75.96°) — Fyc(sin29.74°) = 0
Substituting the expression for F,. and simplifying, we obtain
Fac = — 27.71 kip Fpc = 27.71 kip(C) <«

and using the answer for Fac,


F,c = 16.75 kip (T) <

Joint B. The two unknowns—member force F,, and support iy


reaction B,—can be determined from the equilibrium equations Fan = 7.73 kip|

@2xF, =0 Fgc= 27.71 kip \ +14.04°


5 |
— Fy, + 27.71(cos 29.74°) + 7.73(cos 75.96°) = 0 29.74 }BNE: :
Fp = 25.94 kip (T) < Pap
®zF, =0 B,
B, — 27.71(sin 29.74°) — 7.73(sin 75.96°) = 0
B,=21.25kip@) <
Joint A. Since all member forces are now known, we y
determine the support reactions A, and A, by F ye= 6x7 Skip
. |
@ sF, =0 i 75.96°
A, + 25.94 + 16.75(cos 75.96°) = 0 Ay ——-x
Fup = 25.94ki
A, = —30.0 kip |
7 a i
A. =30kip© < (continued )
312 Chapter 6 Analysis of Statically Determinate Structures

EXAMPLE 6.5 (concluded)

Skip @zx, =0
:| A, + 16.75(sin 75.96°) = 0
10 kip F_A2.01 kip (C) A, = — 16.25 kip
2.58 kip (T) 7.5 kip (T) A, = 16.25 kip@) <
ha @ The forces in all the members and the support reactions
20 kip
27.71 kip (C)
have now been determined, and are shown along with the
16.75 kip (T) 7.73 kip (C) applied loading on the diagram of the truss.

30 kip Checking of computations. Finally, to check our computations,


25.94 kip(T) we consider the equilibrium of the entire truss; the free-body
diagram showing the applied loading and the computed support
16.25 kip 21.25 kip
reactions is shown. For equilibrium of the truss, the three
following equilibrium equations must be satisfied:

@=sF, =0 10+20-30=0 checks


©=r, =0 —5— 16.25+2125=0 checks
@s2M,=0 —30(40) + 16.25(30) + 20(20)
+ 5(20) + 21.25(10) = 0 checks

10 ft 10ft

EXAMPLE 6.6

Determine the force in each member of the Pratt roof truss shown,
by the method of joints.

(continued)
6.7 Analysis of Plane Trusses by the Method of Joints 313

EXAMPLE 6.6 (continued)

Solution

Symmetry. Since the geometry of the truss and the applied


loading are symmetrical about the center line of the truss, its
member forces and support reactions will also be symmetrical
with respect to the line of symmetry. It is thus sufficient to
analyze only one half of the truss. Here, we will analyze only
the left half; the forces in the right half will be obtained by
considering symmetry.
y Line of symmetry G
Zero-force members. At joint D, three members—CD, DE,
and DJ—are connected; of these, CD and DE are collinear. As
no external load is applied to the joint, DJ must be a zero-
force member.

Fy, =0 <

Support reactions. Since a joint with two or less noncollinear


unknowns cannot be found, we calculate the support reactions
from the free-body diagram of the entire truss.

@sF, =0 A,=0 <


Because of symmetry A, = G,, and

©=F, =0 2A,-5(12)=0 A,=30kN@) <


G, =30kN@) <
Joint A. We can now start the method of joints at joint A, with
only two unknown forces Fy, and F,,, acting on it. Applying
the equations of equilibrium, we obtain
©, =0 30 + Fy,(sin 26.57°) = 0 y peteeh:
Fy = —67.08 kN 26.57°
F4y, = 67.08 KN (C) < 30kN
@ZF.=0 Fy — 67.08(cos 26.57°) = 0
Fiz = 60 KN (T) <
Because of symmetry, the forces in members FG and GL
will be equal to the forces in members AB and AH, respectively.
Therefore,
Fo, = 67.08 KN (C) <
Frg = 60 kN (T) <
(continued)
314 Chapter 6 Analysis of Statically Determinate Structures

EXAMPLE 6.6 (continued)

y Joint H. Joint H now has two unknown forces, Fp,, and Fj,;.
| For equilibrium,

@SF, =0
F,,(cos 26.57°) + 67.08(cos 26.57°) = 0
Fy, = — 67.08 kN Fi, = 67.08 KN (C) <
F yj= 67.08 KN ©OxzF, =0
Eee — 12 — Fy, + 67.08(sin 26.57°) — 67.08(sin 26.57°) = 0
Foy = —12 KN Fay =12kN(C) <
Because of symmetry, we obtain

Fx, = 67.08 kN (C) <


Fr, =12kKN(C) <

a Joint B.

©=F,=0 ~12 + Fy,(sin 45°) = 0


Fay i kN F
a Fp, = 16.97 kN (T) <«

F'4p = 60 KN z - Feo——x | @3r= v


— 60 + 16.97(cos 45°) + Feo = 0
Fpc = 48 kN (T) <
From symmetry,
Fre = 16.97 kN (T) <
Frp =48 kN (T) <

% Joint I.
@s3F,=0
67.08(cos 26.57°) — 16.97(cos 45°) + F,;(cos 26.57°) = 0
Fry = —53.67 kN F,, = 53.67 kN (C) <
© F, =0
—12 — Fo, + 67.08(sin 26.57°)
— 53.67(sin 26.57°) — 16.97(sin 45°) = 0
Fe, = —18kN Fo, =18kN(C) <
(continued)
6.8 Analysis of Plane Trusses by the Method of Sections 315

EXAMPLE 6.6 (concluded)

From symmetry,

Fix = 53.67 kN (C) <


Fog = 18kKN(C) < For=18kN a
Joint C. fe ;
0 Steere (Sie) OF5 lis)—10) BC Pgs
Fe, = 21.63 KN (T) <
@sF, =0
— 48-172 1363(C0s) 50:31°) Fa, = 0
Fon =36kN(T) <
From symmetry,

Fey = 21.63 kN (T) <


Fp = 36 kN (T) <
All member forces and support reactions are now known.

Joint J. We will use the two equilibrium equations at joint J


to check our computations. |
@sr. ay, 12kN

53.67(cos 26.57°) — 21.63(cos 56.31°) 50:3 1) ee Oe


i oe

+ 21.63 (cos 56.31°) y)— 53.67 SIP) =) == 0


Bi6/(Cosi26.57 checks uN Wei F jp = 53.67 KN

@s>F,=0
~12 + 53.67(sin 26.57°) + 53.67(sin 26.57°) Fos= 21.63 KN Frs=21.63 kN

— 21.63(sin 56.31°) — 21.63(sin 56.31°) = 0.01 checks Fp, =0

6.8 ANALYSIS OF PLANE TRUSSES


BY THE METHOD OF SECTIONS
In the method of joints presented in the previous section, we start the calcu-
lations at a joint with no more than two unknown forces. We then proceed
from joint to joint, calculating two or fewer forces at each subsequent joint,
until all the unknown forces have been determined. This method is very effi-
cient when forces in all members of a truss are needed.
However, if we need to calculate forces in only certain members of a truss,
the method of joints is inefficient because, before reaching a joint that can be
analyzed for a desired member force, we may have to calculate forces in
several other members. This is especially true if the member of interest is far
away from the joint where the analysis can be started. For example, if we want
316 Chapter 6 Analysis of Statically Determinate Structures

to know the force in member CG only for the truss shown in Figure 6.17a,
after calculating the support reactions we must analyze joints A, F, and B, in
that order, before we can reach joint G and thus determine the desired force
in member CG.

Ay y

|-—6 ft—++—6 ft—+}— 6ft—-1-—6 ft—+|


(b)

Figure 6.17
6.8 Analysis of Plane Trusses by the Method of Sections 317

The method of sections allows us to determine forces in specific members


of the trusses directly, without first calculating many unnecessary member
forces as required in the method of joints.
In the method of sections, the truss is cut into two parts by passing an
imaginary section through the members whose forces are desired; the forces
are then determined by considering the equilibrium of one of the two portions
of the truss. Since the entire truss is in equilibrium, each of its portions must
also be in equilibrium. Each portion is treated as a rigid body in equilibrium
under the action of any external loads and support reactions applied to it, as
well as the forces in the members cut by the imaginary section. The unknown
member forces are determined by applying the three equations of equilibrium
LF, = 0, 2F, = 0, and 2M = 0 to the free-body diagram of one of the two
portions of the truss. Since there are only three equations of equilibrium for
the portion of the truss, they can be used to determine a maximum of three
unknown member forces. Therefore, sections are generally chosen that cut no
more than three members of the truss whose forces are unknown. In some
trusses, the arrangement of members is such that we can employ sections
passing through more than three members with unknown forces, and use the
equilibrium conditions of the corresponding free body to determine one, or
possibly two, of the unknown forces. Such sections, which can be used to
determine at most two unknown forces, are used only in the analysis of a few
certain types of trusses (the K-truss analyzed in Example 6.9 is one).
To illustrate the application of the method of sections, we will determine
forces in the members BC, CG, and GH of the truss shown in Figure 6.17a.
We begin the analysis by passing an imaginary section aa directly through the
three members BC, CG, and GH. Since this section does not cut more than
three members with unknown forces, we should be able to determine these
unknowns by considering the equilibrium of either of the two portions, ABGF
or CEIH. Both portions are attached to the external supports, so we need to
calculate the support reactions by applying the equations of equilibrium to the
free-body diagram of the entire truss (Figure 6.175).

@2F, =0 A,+15=0 A,=—15kip


A, = 1Skip©
@ZM,=0 —10(6) — 20(12) — 30(18) — 15(4) + E,(24) = 0
E, = 37.5 kip@)
©2F, =0 A, — 10 — 20 — 30 + 37.5 =0
A225 Kip)
The free-body diagrams of the two portions ABGF and CEIH are shown in
Figures 6.17c and 6.17d, respectively. In addition to the external loads and
support reactions, the diagrams also show the forces in the members BC, CG,
and GH that have been cut by section aa. (Note that the forces in the members
not cut by section aa have not been shown on the diagrams.) The member
318 Chapter 6 Analysis of Statically Determinate Structures

forces Fpc, Fog, and Fo, are necessary to keep each portion of the truss in
equilibrium by itself. From Figure 6.17c, it is obvious that without Fg-, Fog,
and F,,,, the left-hand portion ABGF cannot remain in equilibrium just under
the action of one load (10 kip) and two support reactions (15 kip and 22.5
kip). Similarly, we see from Figure 6.17d that the right-hand portion CEJH
will also not be in equilibrium without the member forces acting on it.
Since the senses of the member forces are not known at this stage of the
analysis, they are arbitrarily assumed. Member BC is assumed to be in tension;
therefore, force Fz, is indicated by an arrow pulling away at joint B of the
left-hand free body (Figure 6.17c), and by an arrow pulling away at joint C
of the right-hand free body (Figure 6.17d). The remaining two member forces,
Fog and FG,,, are both assumed to be compressive, indicated on the free-body
diagrams by arrows pushing toward the corresponding joints. A comparison
of Figures 6.17c and 6.17d shows that the member forces acting on the two
free-body diagrams conform to Newton’s third law: member forces acting on
one portion of the truss are equal but opposite to those acting on the other
portion.
Since we can use either portion to determine the unknown member forces,
and the left-hand portion has fewer loads applied to it, we will use its free-
body diagram (Figure 6.17c) to analyze member forces. From the dimensions
of the truss given in Figure 6.17a, we have that the inclined member CG rises
4 ft vertically over a horizontal distance of 3 ft; thus, its slope is 3:4. The
slopes of inclined members are sometimes depicted on free-body diagrams by
means of small right-angled triangles drawn on the arrows representing the
member forces, as shown in Figure 6.17c. Applying three equations of equi-
librium to the free-body diagram of portion ABGF, we obtain:

@Q3M,=0 —15(4) — 22.5(9) + 103) + Fy(4) = 0


Fac = 58.13 kip (T)
@2mM-=0 —22.5(12) + 10(6) + Fo,(4) = 0
Foy = 52.5 kip (C)
©ZF,=0 22.5-104+(4/5)Fog=0 Fog = —15.63 kip
Fog = 15.63 kip (1)
In these computations, to avoid solving simultaneous equations, the equi-
librium equations are applied in such a manner that each equation involves
only one unknown. We do this by using an alternate system of equilibrium
equations consisting of two moment summations, 1M, = 0 and 2M; = 0,
and one force summation, {F, = 0 (see Section 3.14). Note that both moment
summations are taken about the points of intersection of the lines of action of
all but one unknown force. In the first moment equation, the moments are
summed about point G, which is the point of intersection of the lines of action
of the two unknown forces Fog and F¢,,, both passing through point G; thus
their moment about G is zero and they do not appear in the resulting equation.
6.8 Analysis of Plane Trusses by the Method of Sections 319

Hence, the only unknown in the equilibrium equation is Fz, whose numerical
value (Fg- = +58.13 kip) is obtained directly. Similarly, in the second mo-
ment equation, »M. = 0, the moments are summed about point C, which is
the point of intersection of the lines of action of the unknown F,, and now
known Fg-, which eliminates both from the equation. The resulting equation
thus contains only one unknown, FG,,. It is important to realize that the point
about which the moments may be summed in applying a moment equilibrium
equation need not be on or within the boundaries of the free-body diagram; it
can be chosen anywhere in the plane of the truss.
The third equilibrium equation, }F, = 0, contains the only remaining un-
known F¢g. The other two member forces are horizontal and have no com-
ponent in the y direction. Note that the answer for F., is negative: our original
assumption of compression was incorrect, and Fy, is actually a tensile force.
Since the three equilibrium equations contain only one member force each,
they can be used in any order.
Finally, we check the computations by writing an alternate equilibrium
equation:
QF, =0 —15 + 58.13 — 52.5 + 15.63/5) = 0 checks

Procedure for Analysis


The following step-by-step procedure can be used for determining the member
forces of most common types of trusses by the method of sections.

1. Select a section that passes through as many members whose forces are
to be determined as possible (but not more than a total of three members with
unknown forces). The section should cut the truss into two portions.
2. If one of the two portions of the truss has no support reactions acting on
it, select this portion for the analysis of member forces and go to Step 3. If
both portions are attached to external supports, draw a free-body diagram of the
entire truss, and apply the three equations of equilibrium {F, = 0, XF, = 0,
and >M = 0 to calculate the support reactions. Next, for the analysis of
member forces, select that portion of the truss that is subjected to the least
number of external forces.
3. Draw a free-body diagram of the portion selected, showing all the external
loads and support reactions, and all the forces in the members cut by the
section. (Do not show internal forces in members not cut by the section.) The
unknown member forces are usually assumed to be tensile, indicated by arrows
pulling away from the joints on the free-body diagram.
4. Write three equations of equilibrium for the free-body diagram, and solve
them to determine the desired member forces. To avoid solving simultaneous
equations, try to write the equilibrium equations so that each may involve only
one unknown force.
5. When the forces in all members that were cut by the section have been
320 Chapter 6 Analysis of Statically Determinate Structures

determined, use an alternate, unutilized equilibrium equation to check the com-


putations. If the analysis has been carried out correctly, the alternate equation
will be satisfied.

EXAMPLE 6.7

Determine the forces in members EG, FG, and FH of the truss


shown, by the method of sections.

Solution Section aa is passed through the three members EG,


FG, and FH, whose forces are to be determined. This section
does not pass through more than three members with unknown
forces, and it cuts the truss into two portions, ABFE and GHJI.
10 kip The portion GHJI above section aa has no support
reactions acting on it. We will use this upper portion to
determine the member forces, thereby avoiding the calculation
of support reactions.
10 kip Free-body diagram. The free-body diagram of the upper
portion of the truss shows the three unknown member forces
Fg, Frg, and Fry, all assumed to be tensile, as indicated by the
arrows.
10 kip
Member forces. The three unknown member forces are
determined by applying the equations of equilibrium as follows.
Since Frg and Fg pass through point G, by summing
moments about G we obtain an equation containing only Fy,;:

@=smM, =0
— 5(20) — 15(20) — Fey4(20) = 0
Fey = —20 kip
Fry = 20 kip (C) <
The negative answer for F;,,, indicates that our assumption
of tension was incorrect; force F;,,, is actually compressive.
Similarly, since the lines of action of F;, and F;,, intersect
at point F, by summing moments about F an equation involving
20 ft
only one unknown, Fgg, is obtained. Thus,

@sm, =0
~10(20) — 5(40) + 15(20) + Fg(20) = 0
Fog = 5 kip (T) <
The remaining member force, F;,,, can be easily calculated
by summing forces in the horizontal direction. Note that Frg
-—20 tt—|
Section aa (continued)
6.8 Analysis of Plane Trusses by the Method of Sections 321

EXAMPLE 6.7 (concluded)

is the only member force with a horizontal component. Thus, Checking of computations. To check our computations, we
Qs E = 0 apply an alternate equation of equilibrium involving all three
a q member forces determined above.
LOR ie (COSI455)
10.
os kip
Frg = —21.21 OxF, =0
= = il) = Ss ZAG)
=F 20) = checks
Frg = 21.21 kip (C) <

EXAMPLE 6.8

Determine the forces in members DE, DJ, and DK of the Parker


bridge truss shown, by the method of sections.

Solution From the diagram of the truss, we observe that any


section passing through the three members DE, DJ, and DK
whose forces are to be determined, will also cut additional
members. For example, section aa passing through the three
members of interest also cuts fourth member //, thereby
releasing four unknown member forces: Fp, Fp;, Fox, and Fy.
The only unknown that we can find from this section is F;,,
by summing moments about point D. Since none of the three
required forces can be determined from the free-body diagrams
associated with section aa, we will first determine two of the
forces, Fy, and Fpx, by considering section bb. The third force,
Fp, will then be determined by analyzing section aa.

Section bb. This section passes through three members, DE,


DK, and JK, cutting the truss into two parts, ADJ and EGK.
Since the right-hand portion EGK has fewer applied loads, we
will use its free-body diagram for the analysis.

ISOKN ISOkKN 350kN

z m>+-8 m+
40m
(continued)
Section bb
322 Chapter 6 Analysis of Statically Determinate Structures

EXAMPLE 6.8 (concluded)

Before calculating member forces, we must determine


support reaction G,, by applying an equilibrium equation to the
free-body diagram of the entire truss.

@smM, =0
— 100(8) — 100(16) — 150(24) — 150(32) [440 kN] E
— 150(40) + G,(48) = 0
G, = 350 kN(t) 150 kN ISOkKN 350 kN

Free-body diagram. The diagram of portion EGK cut by ~ 8m>- 8m+} 8m—++8 m+
section bb is shown, with all three unknown member forces Section aa
Fog, Fpx, and F7x¢ assumed to be tensile. The angles of
inclination of members DK and JK with the horizontal are
calculated from the dimensions of the truss as 51.34° and Note that in the above equation, the terms F,,(cos 51.34°) and
14.04°, respectively. Fpx(sin 51.34°) are the components of Fp, in the x and y-
directions, respectively.
Forces in members DE and DK. Summing moments about K
(the point of intersection of the lines of action of Fp, and Section aa. We can now use section aa to calculate the force
Fix), we obtain in member DJ. The free-body diagram of the portion to the
right of section aa is shown.
@=s5m, =0
— Fp-(10) — 150(8) + 350(16) = 0 Force in member DJ.

For = 440 kN (T) < @©=mM, =0


Next, we calculate Fp, by summing moments about point — Fp ,(8) — 44010) + 64.03(sin 51.34°)(16) — 150(16)
O, which is the point of intersection of the lines of action — 150(24) + 350(32) = 0
of Fx and Fp,. The distance of O from E (see figure) can be
Fy,= 200 kN(T) <
calculated from
IA i
tan 14042 === At FO=<4
EO, BO Oa
Equilibrium of moments about O gives:

@©zm a=”
Fpx(cos 51.34°)(10) + Fp,(sin 51.34°)(40) + 150(40)
+ 150(32) — 350(24) = 0

Fox = 64.03 KN (C) <


6.8 Analysis of Plane Trusses by the Method of Sections 323

EXAMPLE 6.9

Determine the forces in members CD, DI, and IN of the K bridge


truss shown, by the method of sections.

Solution From the diagram of the truss, we observe that a


vertical section aa passing through the three members of
interest, CD, DI, and IN, also cuts a fourth member, MN,
thereby releasing four unknowns, a number we cannot
determine with three equations of equilibrium. Moreover, the
arrangement of members is such that it is not possible to obtain
forces in any of the members cut by the section—a point of SOKN SOkKN SOKN SOkN SOKN
concurrency of three unknown forces does not exist. 4m>k4 mk 4 me 4me 4m>+-4m
Trusses whose members are arranged in the shape of the
letter K are usually analyzed by a section curving around the
middle joint, like section bb shown in the figure. From the free-
body diagram of the portion of the truss on the left side of
section bb, we see that even though the section has cut four
members, CD, CI, IM, and MN, the forces in two members, CD
and MN, can be determined because there are two points of
A
concurrency, C and M, through each of which three unknown
forces pass. Since Fop, Fc,, and Fj, pass through point C,
the moment equilibrium equation about C will involve only 125kN 50kN 50 kN
Fyy- Similarly, by summing moments about point M, three he 4 mek 4 ms
unknowns—F,,, Fi,,, and Fy,,—can be eliminated from
Section bb
the equation, leaving only Fop. Thus, we will first determine
Fep by considering section bb. Section aa will then be analyzed
to calculate Fp, and Fy.
Since the truss and the loading are symmetric, the vertical
reaction at A is simply equal to one half of the total vertical
loading.
A, = @)(5)(50) = 125kKN A, = 125 kN)
also, by using

LF. =0 A, =90

Section bb. As discussed above, Fep can be determined from

@=sm,, =0
— 125(8) + 50(4) + Fop(6) = 0
Fop = 133.33 kKN(T) <«
(continued )
324 Chapter 6 Analysis of Statically Determinate Structures

EXAMPLE 6.9 (concluded)

L M Fun Section aa. The free-body diagram of the portion on the left
a side of section aa is shown. By summing moments about NV
AG Fin iEm (the point of intersection of the lines of action of Fj, and Fy),
i we obtain
SS iF 3m
A 2 @©=mM, =0
C 133.33 kN
— 125(12) + 50(8) + 50(4) + @Fp,(3) + @Fp,(4)
125 kN SOkKN 50 kN + 133.33(6) = 0
we 4m-+<+4 make 4m! Fp, = 20.84 kN (T) <
Section aa Finally, we calculate F,, by summing forces in the vertical
direction, thereby elminating Fy, from the equilibrium
equation. Thus,

@s5, =0
125 — 50 — 50 — (820.84 + @Fy =0

*6.9 SPACE TRUSSES


A space truss is constructed by connecting straight members at their ends by
flexible connections to form a stable or rigid three-dimensional configuration.
Space trusses, due to their shape and arrangement of members, cannot be
subdivided into plane trusses, but must be analyzed as three-dimensional bod-
ies subjected to three-dimensional force systems. As discussed in Section 6.4,
to simplify the analysis of space trusses, it is assumed that the members are
connected by frictionless ball-and-socket joints, and all external loads and
reactions are applied only at the joints. With these simplifying assumptions,
the members of a space truss can be treated as two-force members, either in
axial tension or in axial compression.
The simplest rigid space truss is constructed by connecting six members at
four joints to form a tetrahedron as shown in Figure 6.18a. This tetrahedron
is the basic space truss element—a three-dimensional rigid body that will not
change its shape when subjected to a general three-dimensional system of
forces applied to its joints. The basic element can be extended by attaching
three new members BE, CE, and DE to the three existing joints B, C, and D,
and connecting them to form a new joint E as shown in Figure 6.186. Provided
that the new joint F does not lie in the plane of the three existing joints, the
new truss will be rigid. The truss can be further extended by repeating the
same procedure. For example, the space truss shown in Figure 6.18c was
formed by adding joints F, G, and H to the truss of Figure 6.18b. Trusses
6.9 Space Trusses 325

E (New joint)

Figure 6.18

constructed by this procedure are called simple space trusses. The number of
members in a simple space truss is given by
m= 3j—-6

where m is the number of members, and / is the total number of joints, in-
cluding those attached to the supports.

Static Determinacy
To analyze a space truss containing m members and j joints, and supported by
r external reaction components, we need to calculate a total of m + r unknown
forces. At each joint, the external and internal forces form a concurrent three-
dimensional force system that must satisfy the three equilibrium equations
326 Chapter 6 Analysis of Statically Determinate Structures

LF. = 0, YF, = 0, and XF, = 0. As the truss contains j joints, the total
number of equations of equilibrium available is 3j. If m + r = 3), then all
the unknowns can be calculated by applying the equations of equilibrium, and
the truss is considered statically determinate.
If a space truss contains more unknowns than the available equilibrium
equations (i.e., m + r > 3), it is considered statically indeterminate; if a
space truss has fewer unknowns than the equilibrium equations (i.e., m + r
< 3)), it is considered statically unstable. It should be noted that the condition
m + r = 3j, while necessary for static determinacy, is not sufficient. Recall
from Section 4.9 that for a three-dimensional rigid body to be properly con-
strained, it must be supported by at least six reactions. Moreover, the arrange-
ment of the members and the support reactions must be such that the truss
maintains its shape, and does not move as a rigid body, when subjected to
external loads.

Analysis
The methods of joints and sections, presented in Sections 6.7 and 6.8 for the
analysis of plane trusses, can be extended to the analysis of space trusses. In
the method of joints, three equilibrium equations, }F, = 0, XF, = 0, and
>F, = 0, must now be satisfied at each joint of the space truss, and used to
calculate a maximum of three unknown forces at a joint. Thus, we start the
analysis at a joint with no more than three unknowns. Proceeding from joint
to joint, we calculate three or less unknown forces at each subsequent joint,
until all the unknown forces have been determined. Because it is difficult to
visualize the orientation of inclined members in space, it is convenient to
express member forces as vectors, with direction cosines computed from the
x, y, and z coordinates of the two end joints. The three scalar equations of
equilibrium of a joint may be expressed in vector notation as {F = 0.
The method of sections can be used to determine forces in specific members
of space trusses. As with plane trusses, an imaginary section is passed through
the members whose forces are desired, cutting the truss into two portions. The
desired forces are then determined by considering the equilibrium of one of
the two portions. For equilibrium of a portion of the space truss, the force
system acting on it must satisfy the following six scalar equations of equilib-
rium:

SF.=0 SF =0 &F =0
2M,=0 3M,=0 %M,=0
These six scalar equations may also be written as two vector equations,

XF =0 and =M= 0
6.9 Space Trusses 327

Since no more than six unknown forces can be determined from these equa-
tions, a section is generally selected that cuts no more than six members with
unknown forces.
Due to the considerable amount of computation involved in the analysis of
space trusses, digital computers are usually used, with computer programs
generally based on the method of joints. In these programs, the joint equilib-
rium equations are systematically established at all joints of the truss without
regard for the number of unknowns at any joint. The resulting system of si-
multaneous equations is then solved by an elimination technique to determine
the unknown member forces and reactions.

EXAMPLE 6.10

Determine the force in each member of the space truss shown.

Solution

Joint E. Since joint E has only three unknown forces, Frp,


Frc, and Fy, we can start the method of joints here. The free-
body diagram of this joint shows that the three unknown
forces are assumed to be tensile. The unit vectors in the
directions of the three unknown forces, obtained from the
dimensions of the truss, are

Neg = k Nec = —i Nep = —0.236i + 0.943j + 0.236k

Writing the three unknown forces as vectors,

Fp = Fepk Frc = —Fecl


Fep = Fep(—0.236i + 0.943j + 0.236k)

For equilibrium of joint £,

=F =0 Fig + Fac + Fep = 0


Frgk — Feci — 0.236 Fepi + 0.943 Fepj + 0.236 Fepk = 0

Setting the terms containing i, j, and k equal to zero, we get the


scalar equilibrium equations
>F,=0
— Fro — 0.236Fep =0 Of Fe = —0.236Fep
oe = 0 0.943Frp = 0 Frp =0 <
Fee = 0 <
ZF, =0 Fre + 0.236F ep = 0 Frp = 0 <

(continued )
328 Chapter 6 Analysis of Statically Determinate Structures

EXAMPLE 6.10 (continued)

Joint D. This joint now has three unknown forces—Fp,, Fp,


and F,-—the directions of which are given by the unit vectors

Np, = —0.236i — 0.943j + 0.236k


Appz = 0.236i — 0.943j + 0.236k
Npc = —0.236i — 0.943j — 0.236k
The equilibrium equation is
XF =0
Fy, + Fog + Foc — 10j
— 5k = 0
Fpa4(— 0.2361 — 0.943j + 0.236k)
+ Fpp(0.236i — 0.943) + 0.236k)
+ Fy
— 0.2361 — 0.943j — 0.236k) — 10j - 5k = 0
from which we obtain the three scalar equations

ZF, =0
~0.236Fy4
+ 0.236F pp — 0.236Fpc = 0
a0
~0.943Fp4 — 0.943 Fp — 0.943Fne — 10 = 0
>F=0
+ 0.236Fp, + 0.236Fpp — 0.236Fp¢ — 5= 0

Solving these equations, we obtain

Fp, = 10.6 kN Fy, = 10.6 kN(T) <


Fpp = —5.3kN Fpp =5.3kKN(C) <

Foc = —15.9kKN Fpc = 15.9kN(C) <


The negative answers for Fp, and Fp- indicate that our
assumption of tension was incorrect and they are actually
15.9npc (KN) compressive.

Joint C. Applying the vector equation of equilibrium, we


Fo, =0 obtain
er. SF=0
Foak + Fog(0.707i + 0.707k) + C,j
+ 15.9(—0.236i — 0.943j — 0.236k) = 0
This gives us the scalar equations

LF, =0 0.707F-, — 3.75 = 0 Fon = 5.31 kN (T) <«


(continued)
6.9 Space Trusses 329

EXAMPLE 6.10 (concluded)

SF = 0 C,-15=0 C, = 15kN@) A 5.31 ngc(KN)


5 .3npp (kN)
2FL=0 Fy, + 5.31(0.707) — 3.75 =0 Fx,=0 <

Joint B. Similarly, at joint B

SF =0 — Fo,i +Byj + Bk
+ 5.3(0.236i — 0.943] + 0.236k)
+ 5.31(—0.707i — 0.707k) = 0
SF. =0 =F), 2.5=0 ° F,,=25kNO©), <
YF, =0 B,-5=0 B,=5kN@)<
SF. =0 B.-2.5=0 B.=2.5kNY) <
Joint A. The three remaining unknown support reactions can y
be determined by considering the equilibrium of joint A. Thus, | 10.6M 4p (KN)
2SF=0 A,i+A,j+A,k — 2.5i |
+ 10.6(0.236i + 0.943j — 0.236k) = 0
|
2.5np, (KN)
LF, =0 A, —2.5+2.5 =0 A.=0< —— ——
P5

2F, =0 A,+ 10=0


A, = —10kN A, = 10kNQ) <

IF. =0 A—25=0 A =25kNY)<


330 Chapter 6 Analysis of Statically Determinate Structures

6.10 FRAMES AND MACHINES

Frames

Frames are constructed by connecting members to form rigid configurations.


Unlike trusses, which consist of only straight members, each connected at its
two ends, the members of frames need not be straight, and may be connected
at more than two locations. On trusses, the external loads are applied only at
the joints, which are considered to be hinged. The members of frames, how-
ever, may be connected together either by rigid (moment-resisting) connec-
tions, or by flexible connections (hinges or links). Also, the external loads on
frames may be applied on the members, as well as on the joints. Because of
these distinguishing features, the members of frames are, in general, not two-
force members, even though a frame may contain some two-force members.
A structure is classified as a-frame if it contains at least one multiforce (not
two-force) member.
Like trusses, frames are commonly used to support loads. Reinforced con-
crete and structural steel frames are used to support loads in high-rise-buildings,
bridges, and industrial plants. Frames are also used in aircraft and aerospace
vehicles, and in numerous other civil, mechanical and aerospace applications.
Although both rigid (moment-resisting) and flexible (hinged) joints are used
to connect the various members of a frame, rigid joints are more common.
Frames with rigid joints are usually statically indeterminate; thus, their analysis
is beyond the scope of this text. We will, however, consider the analysis of
statically determinate pin-connected frames.
As we discussed for trusses, a great majority of three-dimensional structural
systems can be divided into plane or two-dimensional structures for analysis
and design. In this section, the emphasis will be on the analysis of plane frames
subjected to coplanar force systems. By analysis we mean the determination
of the forces acting on the various members of the frame at connections, for
a given loading.

Machines

A machine may be defined as an assemblage of members connected together,


whose function is to transmit or modify forces. Members or parts of machines
are usually connected by hinged joints. Like frames, machines contain multi-
force members.
From the point of view of analysis, the main difference between frames and
machines is in the manner in which they are constrained. Frames are usually
properly constrained: Rigid frames, which do not change shape and remain as
rigid bodies when detached from their supports, must be supported by at least
three nonconcurrent and nonparallel reactions. Nonrigid frames, which change
shape and become nonrigid when detached from supports, must be supported
by more than three external reactions, to help them maintain their shapes as
6.10 Frames and Machines 331

rigid bodies when attached to the supports, and to prevent all possible rigid-
body movements under a general coplanar force system. Machines, unlike
frames, are seldom properly constrained; they usually contain movable parts
or members and, therefore, are stable or rigid only under certain specific load-
ings.

Analysis
The analysis of statically determinate frames and machines is based on the
concept that if a frame or a machine is in equilibrium, each of its members
and joints must also be in equilibrium. The forces acting on the members of
a structure are thus determined by considering the equilibrium of its members
and joints.
A frame or a machine is considered to be statically determinate if all the
external reactions, as well as the internal forces acting on all of its members,
can be determined by applying the equations of equilibrium.
Consider the plane frame shown in Figure 6.19a. The frame is composed
of three members, AE, BF, and CD, connected together by hinged connections
at joints C, D, and E. The free-body diagram of the entire frame is shown in
Figure 6.195, and those of the three members and joint F are shown in Figure
6.19c. As we discussed in Section 6.2, it is not necessary to draw free-body
diagrams of joints C and D, because no external loads are acting at these joints,
and only two members are connected to each of them. Internal forces at such
joints can be directly transmitted (with the directions reversed) from one mem-
ber to the other. Note that member CD is a two-force member: it is connected
at only two points, and is not subjected to any external loading. The other two
members—AE and BF—are multiforce members, because each of them is
connected at three locations, and is subjected to external loading. The correct
senses of the reactions and internal forces are not known at this stage, and are
arbitrarily assumed. On all of the diagrams, the law of action and reaction
between bodies in contact has been followed.
Analysis of the frame involves calculating the magnitudes of the five in-
ternal forces Fop, E24”, EA”, E8", and E®", and the three external reactions A,,
AS: and B,. Thus, the total number of unknown forces to be determined is
eight.
As the frame is in equilibrium, each of its members and joints must also be
in equilibrium. At each of the two multiforce members AE and BF, the internal
and external forces form a coplanar force system, requiring that the three
equilibrium equations 2F, = 0, 2F, = 0, and 2M = 0 be satisfied. Also, at
joint E, the external and internal forces form a concurrent and coplanar force
system, requiring that the two equilibrium equations 2F, = 0 and YF, = 0
be satisfied. Therefore, for the two multiforce members AE and BF, and one
joint EZ, the total number of equations available is 2(3) + 1(2) = 8. These
eight equilibrium equations can be solved to calculate the eight unknown
forces—the frame of Figure 6.19a is therefore statically determinate. It should
332 Chapter 6 Analysis of Statically Determinate Structures

kip/ft
0.5

<—10 ft 10ft

Figure 6.19

be noted that in counting the total number of equilibrium equations, the equa-
tions for the two-force member CD were not included. This is because the
only equilibrium equation available for this member {F, = 0, had already
been applied to the free-body diagram of the member. On the free-body dia-
6.10 Frames and Machines 333

gram of member CD (Figure 6.19c), the forces Fup acting at the two ends are
assumed to be equal, opposite, and collinear, thereby satisfying the equilibrium
condition for this member.
It should be realized that the three equations of equilibrium of the entire
frame as a rigid body could be used to determine the three unknown external
reactions A,, A,, and B,. However, these equations are not independent from
the equilibrium equations of the dismembered frame, and contain no additional
information.
In general, the analysis of a frame or a machine may begin at either a
multiforce member with three or less unknown forces, or at a joint with two
or less unknowns. The equations of equilibrium for the selected member or
joint are then written and solved to determine the unknown forces. Another
member or joint is then selected, with an equal or less number of unknown
forces than equilibrium equations, and the procedure is repeated until all the
desired forces are determined. If the structure is statically determinate, all
internal member forces as well as all external reactions can be determined by
this procedure. However, in many structures, it may not be possible to find a
member or joint with equal or less number of unknowns than the number of
equilibrium equations to start the analysis. For such structures, external reac-
tions are computed first from the equilibrium equations of the entire structure
before proceeding with the calculation of member forces.
To illustrate the method of analysis, consider again the frame shown in
Figure 6.19a. An examination of the free-body diagrams of the members and
the joint (Figure 6.19c) indicates that both multiforce members and joint E
have more unknown forces than the equilibrium equations. Therefore, we de-
termine the external reactions first, considering equilibrium for the entire frame
(Figure 6.19b):

@F, =0 —A, + 10 — 10(cos 30°) = 0


A, = 1.34 kip©
@=M,=0 —A,(20) — 10(10) + 10(sin 30°)(10)
+ 10(cos 30°)(20) — 15(5) = 0
A, = 2.41 kip@
©®2zF, =0 2.41 — 10(sin 30°) — 15 + B, =0
By = 17.59 kip1)
Both members AE and BF now have only three unknowns each. We will
use equilibrium for member AE to calculate the three unknown forces Fop,
EA", and 1B

@xF, =0 2.41 + BAF =0


EAF = —2.41 kip
@SM,=0 ~1.34(20) — 2.41(10) + 10(10) + Fop(10) = 0
334 Chapter 6 Analysis of Statically Determinate Structures

or
Fop = 4.91 kip (C)
@zF, =0 ~ 1,34+ 10 —4.91 + E4= =0
EAE = —3.75 kip
The negative answers for Fop, £4”, and E4” indicate that their actual senses
are opposite to those shown in Figure 6.19c.
Next, we consider the equilibrium ofjoint E.

@rF, =0 —(—3.75) — 10(cos 30°) + E28" = 0


E8F = +4.91 kip
OxrF, =0 —(—2.41) — 10(sin 30°) + E8" = 0
ESF = +2.59 kip
Since all unknown forces have been determined, the three equations of
equilibrium of member BF can be used to check our computations.

QBS = 09 OS 49 4. 90 checks
Ose =O 1750 259e 5 aC checks
@ZM,=0 —4.91(10) + 4.91(20) + 2.59(10) — 15(5)=0 checks
The results of the analysis are shown in Figure 6.19d.

Procedure for Analysis


The following step-by-step procedure can be used for the analysis of most
common types of frames and machines.

1. Identify by inspection any two-force members of the structure.


2. Draw free-body diagrams of all members, and of those joints having more
than two members connected to them and/or any external loads applied to
them. The senses of the internal forces on the free-body diagrams must be in
conformance with Newton’s third law.
3. Count the number of unknown forces and the number of equilibrium equa-
tions available. If the number of unknowns is greater, the structure is statically
indeterminate; end the analysis at this point. If the structure is statically de-
terminate, proceed to the next step.
4. Determine support reactions by applying the three equations of equilib-
rium to the free-body diagram of the entire structure. If the number of support
reactions is greater than three, as may be the case for nonrigid frames and
machines, it will not be possible to determine all the reactions. However, one
or two of the reactions can usually be calculated.
5. Select a multiforce member with three or less unknown forces, or a joint
with two or less unknowns.
6.10 Frames and Machines 335

6. Write the equilibrium equations ({F, = 0, 2F, = 0, and 2M = 0 in


case of a multiforce member, or 2F, = 0 and {F, = 0 in case of a joint),
and solve them to determine the unknown forces.
7. If all the desired internal forces and reactions have been determined, pro-
ceed to Step 8. Otherwise, return to Step 5.
8. Since the support reactions were calculated in Step 4 from the equilibrium
equations of the entire structure, there should be some equations remaining
unutilized, equal in number to the number of reactions computed in Step 4.
Use these to check the computations. If the analysis has been carried out
correctly, the equilibrium equations will be satisfied.

For some types of frames and machines, a member or a joint with equal or
less number of unknown forces than the number of equilibrium equations may
not be found to begin or continue the analysis. In such cases, it may be nec-
essary to write equilibrium equations in terms of unknowns for two or more
free-body diagrams, and solve the equations simultaneously to determine the
unknown forces.

EXAMPLE 6.11
1 kip/ft
Determine the forces acting on each member of the frame shown.

Solution

Two-force members. Members CF and DG are two-force


members.

Free-body diagrams. The free-body diagrams of the entire


frame, and of the five members of the frame—AE, BH, EH, CF,
and DG—are shown on the next page. Since no external load
is applied at any joint, and no more than two members are
connected at any joint, the diagrams of the joints are omitted;
the internal forces are directly transmitted from one member to
the other using the law of action and reaction.

Static determinacy. The total number of unknown forces


(including support reactions) acting on the five members of the
frame is nine: A,, A,, B,, E,, E,, H,, Hy, For, and Fpg. The
frame contains three multiforce members, AE, BH, and EH, and
there are three equilibrium equations for each multiforce
member; thus the total number of equations available is nine.
The frame is statically determinate.

(continued )
336 Chapter 6 Analysis of Statically Determinate Structures

EXAMPLE 6.11 (continued)

Support reactions. Applying the three equations of equilibrium Member BH. Member BH now has only three unknown
to the free-body diagram of the entire frame, we obtain forces, H,, H,, and Fpg, and there are three equilibrium
: equations for this member. Since two unknowns, H, and Fpo,
= —A,=0 =5k < 4 : y DG
@zr, c : 2 a; 'P © pass through point D, the third unknown, H,, can be calculated
@sm, =0 —5(6) + £(15)(7,5) =A 1S) ae by summing moments about D.

A,=55kip@) < ©2M,=0 H,(3) = 0 an Lge


©sF, =0 5.5 — 1(15)+ B, = 0 @zF, =0 —Fog{cos 45°)=9 Fog= 0%
B,=9.5kip@) < ®sr, =0 —H,+95=0 H,=95kip <
Member EH.
1 kip/ft
@s5m, =0—
eeere E,@) — 1(15)(4.5) + 9.5(12) = 0
3 ft E,= —15.5kip <
@ sr, =0
—(—15.5) — Fop(sin 45°) — 1(15) + 9.5 =0
9 ft For = 14.14 kip (T) <
@sF,=0 ~E, — 14.14(cos 45°) = 0
E,=—10kip
3
<

1 kip/ft
EXAMPLE 6.11 (concluded)

Opis =0 @Ozm, =0
—5+5+ 14.14(cos 45°) — 10 =0 checks —5(6) — 14.14(cos 45°)(9) — (— 10)(12) = 0.01
Ox, = () checks
5.5 + 14.14(sin 45°) — 15.5 = 0 checks The results of the analysis are shown in the figure.

1 kip/ft
9.5 kip

14.4 kip

14.14 kip
F

EXAMPLE 6.12

Determine the forces acting on each member of the gable frame


shown.

Solution The free-body diagrams of the entire frame, the two


members AC and BC, and joint C are shown. There are eight
unknown forces. The number of equilibrium equations is 3 X
(2 multiforce members) + 2 X (1 joint) = 8. The frame is
statically determinate.

5 kip

(continued)

337
338 Chapter 6 Analysis of Statically Determinate Structures

EXAMPLE 6.12 (continued)

Support reactions. The four support reactions cannot all be @Q xr ==) —2.86+ 10-—C*" =0
determined by applying the three equilibrium equations to the OC ana eee
free-body diagram of the entire structure. But the two vertical ys 3 ; P
reactions A, and B, can be calculated from the two equilibrium Ox y= 90 Jo Cs) Ce =05<
equations Taina

©rm,=0 @s=F,=0 714-8=0 C= +7.14kip <


—=A,(30)i= 10(15) + 5(30) + 107115) = 0 © F, =0 10 Co =0 CHE = 44 5\(9) kip <

A,=5kip@® <
©sr=0 5-5-10-5+B,=0 gee:
B,=15kip(t) < @zF, =0 _ =0
714-8, B.=7.14kip© <
Member AC. Member AC now has only three unknown forces,
An Cee sand cee which can be calculated from the three
equilibrium equations for this member.

@=M- =0
—A,(21) — 5(15) + 10(6) + 5(15) = 0
= 2.86 kip© <

1.0 kip

ay Brahe Cr
EXAMPLE 6.12 (concluded)

Entire frame. Since the equilibrium equation }F, = 0 forthe check of our results.
free-body
tee-body didiagram of f the
th entire
ire fiframe was not utilized
ili to Qs Fa
calculate support reactions, we can use it to provide a further Ee hes U8 haat gas
The results of the analysis are shown in the figure.

10 kip

Tip |714k

EXAMPLE 6.13

Determine the forces acting on each member of the frame shown.

Solution The free-body diagrams of the entire frame and of


the three members are shown. There are nine unknowns.
The number of equilibrium equations is 3 X (3 multiforce
members) = 9. The frame is statically determinate.

(continued )
eS ne
339
340 Chapter 6 Analysis of Statically Determinate Structures

EXAMPLE 6.13 (continued)

Support reactions. Using the free-body diagram of the entire @sr. =0


frame,
(a)
@srF, =0 A.=0 <
—A(4)+0.5(4(2)=0 4,=1kN@ <
Member AG.
@zM,=0
@sr=0 1-05(4)-1+B,=0 B,=2kN@)< xr, =0 1-05 —G,=0 G, =+05kN <
Oru,20" “=10) S05) 46.8) = 0
Member DE. All four unknown forces D,, D,, E,, and E,
acting on this member cannot be determined from the three
G,= +0.5kN <
equations of equilibrium. But we can determine the two vertical ew F=0 D,— 0.5 =0 D, =+05kN <
forces D, and E,, and establish a relationship between the two : :
horizontal forces Da and Be Using Equation (a) for member DE, we have

©3M,z=0 +0.5(4)(1) —1(1)=0


D2) E.=+05KN <

@sr=0 -(-05)—05(4)-1+5,=0

0.5 kKN/m

(continued)
EXAMPLE 6.13 (concluded)

Member BG. Equilibrium check: @=mM, =0 2.5(1) + 0.53) — 0.5(2) — 0.5(6) =O checks

Qs sr, =0 -05+05=0 checks The final results are shown in the figure.
@sF, =0 +2-25+05=0 checks

0.SkKN 2.5kKN 1 kN

EXAMPLE 6.14

Determine the forces acting on each member of the frame shown.

Solution The free-body diagrams of the entire frame, the two


members AC and CD, and the pulley B are shown. There are
nine unknown forces. The number of equilibrium equations is
3 X (2 multiforce members) + 3 (1 pulley) = 9. The frame is
statically determinate.

Bes5 ft->+_— 65 ft
(continued)

341
342 Chapter 6 Analysis of Statically Determinate Structures

EXAMPLE 6.14 (continued)

Support reactions. Of the four unknown reactions, only one, @©=zmM, =0 500(1.5) — D,(4) = 0
A,, can be determined from the equilibrium equations for
the entire structure.
D, = 187.5 1bCG) <
@yE=0 A, — 500 = 0 A, =500Ib(t) < @2xF, =0 C, — 500 + 187.5 =0
We can use the remaining two equilibrium equations to check
C, = +312.51bC) <
our computations after all the unknowns have been determined.
Member AC.
Pulley B.
@2zF, =0 A, + 500 — 312.5=0
2M, = 0 500(1.5)
— F(..5) = 0
A, = — 187.5 Ib
F = +500 lb(T) <
A, = 187.51lb©) <
@=sF =0 -B,+500=0 B,=+500lb <
M, — 500(5) = 0
®=F, =0 B,-500=0 B,=+500lb < res
M, = 2500 Ib: ft@) <
Member CD.

©, =0 C,=0 <

be iW: fee

(continued)
6.10 Frames and Machines 343

EXAMPLE 6.14 (concluded)

Entire frame. Equilibrium check:

G@SF=0 1875 +1875 =0 checks


@=M,=0 +2500 — 500(3.5) — 187.5(4)=0 checks
The answers are shown in the figure.

500 Ib

Sod) 1)

500 Ib
G

BES b S12251b
500 Ib

EXAMPLE 6.15

For the bolt cutter shown, determine the magnitude of the vertical
forces P that must be applied to generate a force of 1000 Ib on
the rod to be cut. Also, determine the forces on each member of
the machine under this loading condition.

Solution Member CG and rod H are two-force members. The


free-body diagrams of the five members and the rod H are
shown on the next page. Since the structure is symmetric, we
will determine only six unknown forces—?, B,, B,, D, D,, and
Fcog—acting on the upper-half of the machine. Because of
symmetry, F, = B, and F, = B,. The six unknowns can be
calculated from six equations of equilibrium for the two
multiforce members AD and BC; the structure is statically
determinate.
1in.4iM. 2 in,
(continued)
344 Chapter 6 Analysis of Statically Determinate Structures

EXAMPLE 6.15 (concluded)

Reactions. The two 1000-lb forces acting on the rod H may be


considered as reactions generated by the applied forces P,
which are unknown. The free-body diagram of the entire ae
machine does not provide any information except that the two <a = B G
forces P must be equal, opposite, and collinear. Note that while Lee f a8
this machine is in equilibrium under the given loading Le. Be ae 1000 Ib
condition, it is not constrained to prevent possible rigid-body 1000 lb
movements.
H
Member BC.

B,=0 | ar 1000 Ib
@zF, =0
ae Fy 4 $10001b
@=M- =0 1000(2) + B,(4)=0 B, = —500 Ib : fea te
Ox, =0 —(—500) — Fog + 1000 = 0 — ©
Fog = 1500 lb (T)
Member AD.

@=sF,=0
@©3M,=0 — + P(16) + (—500)(1) = 0
P= £3125 ib
@2F,=0 ~31.25
— 500 —D, =0
D, =—531.25 Ib
Because of symmetry

The answers are shown in the figure.


6.10 Frames and Machines 345

EXAMPLE 6.16

A force of 100 KN is applied to the handle of the pressing machine


as shown. Determine the vertical force exerted on the block G by
the machine. Also, determine the forces exerted on the members
of the machine at joints A, B, C, D, and F. The pressing plate F
slides vertically between the two smooth walls.

Solution Members AB, BC, and BF are two-force members.


Note that the horizontal force F, represents the reaction of
smooth walls on plate F. The force exerted by the machine on
the block is denoted by F,. The number of unknown forces is
nine: A,, Ay, D,, D,, ple
F,, F,, Tks, Fag, Fgc, and Fgp. Three equilibrium
equations are Eatlable for the multiforce member CE, and two
equations each for joints A, B, and F. Thus, the total number
of equilibrium equations is nine, and the machine is statically
determinate.

Support reactions. None of the six external reactions can be


determined from the three equilibrium equations of the
entire structure.

Member CE. (See figure on next page.)


@=sm, =0
Fac(150)(sin 45°) — 100(400) = 0
Fac = 377.12 KN (T) <
@=F,=0 D,— 100(sin 45°) = 0
D,=710.71kN@) <
@=sF.=0
—377.12 + D, + 100(cos 45°) = 0
D, = 306.41kNC) <
Joint B.

©=F, =0
F,,-(cos 15°) — F,,(cos 15°) = 0 Fer = Fap (a)
@=rF, =0
— F,,(sin 15°) — Fp,(sin 15°) + 377.12 = 0
Using Equation (a), we obtain
Fyp = 728.54 kN (C) <«
Fp = 728.54 kN (C) <
(continued)
346 Chapter 6 Analysis, of Statically Determinate Structures

EXAMPLE 6.16 (continued)

Joint A. Joint F.

@2xzF,=0 728.54(sin 15°) — A, = 0 @QsF, =0 728.54(sin 15°) — F, = 0


A, = 188.56kN@) < F. = 188.56kN@) <
© r, =0 728.54(cos 15°) — A, = 0 @sr,=0 F, — 728.54(cos 15°) = 0
A, = 703.72kNQ) < F, = 703.72kN() <

(continued)
6.10 Frames and Machines 347

EXAMPLE 6.16 (concluded)

Entire frame. Equilibrium check: @©=sm, =0


@ xr. =0 188.56(300 cos 15° + 150 sin 45°) + 703.72 x

— 188.56 + 306.41 + 100(cos 45°) — 188.56 = 0 (300 — 300 sin 15° + 150 cos 45°) — 100(400)
— 188.56(300 cos 15° — 150 sin 45°) — 703.72 x
checks
(300 — 300 sin 15° + 150 cos 45°) = —0.38 ~0
©, =0
checks
— 703.72 — 100(sin 45°) + 70.71 + 703.72 = 0
The results of the analysis are shown in the figure.
checks

703.72kKN

A #+— 188.56 kN

Abra KN
728.54 kN

B 100 kN

728.54 KN
728.54 kN E
$377.12kN B 377.12 kN
B ——___
377.12 kN 377.12 kN Cc
306.41 KN
728.54kN
728.54 kN
70.71 kN
B

F 2

728.54 kN
Vege: kN

F e+— 188.56 kN

703.72 kN
348 Chapter 6 Analysis of Statically Determinate Structures

EXAMPLE 6.17

For the lifting platform shown, establish a relationship between


the load W and the horizontal force P, in terms of the length L
and the height h of the platform. Also, determine the forces acting
on each member of the structure.

Solution The free-body diagrams of the entire platform and its


members are shown. There are nine unknowns. The number of
equilibrium equations is 3 X (3 multiforce members) = 9. The
structure is statically determinate.

Support reactions.

@s=sF, =0 A,-P=0
AL =P. _ (a)
@sM,=0 -A(L)+WL/2)=0 A,=w/20) <
© sr, =0 (W/2)-W+B,=0 B,=w/2@ <
Member DE.

ye, 36 D.=0 <


@sM,=0 -D(L)+WiL/2)=0 D,=+W/2 <
©sr, =0 (W/2)-W+E,=0 £,=+W/2 <
Member BD.
@=M- =0
(W/2)(L/2) + (W/2)(L/2) — P(h/2) = 0
P=WL/hC) <
@ sF, =0 CoAIWE/i=0) 0)C= 4h fee
©sr=0 (W/2)-C,-(W/2)=0 C,=0 <
From Equation (a), we obtain

A,=WL/hC) <
(continued)
6.10 Frames and Machines 349

EXAMPLE 6.17 (concluded)

Numerical values.

P =1.78kip© A, =1.78kip©
,=05kipQ) —-B, = 0.5kip)
=
(2
1ou=+178kip C,=0
D-=0 D,=+05kip E,= +0.5 kip
These numerical results are shown in the figure.

1 kip

0.5 kip 0.5 kip

0.Skip 0.5 kip


EXAMPLE 6.18

A mechanical system of two gears is supported by frictionless


bearings on a bracket as shown. If gear A is subjected to a clock-
wise torque M, = 1200 lb - in., determine the torque that must
be applied to gear B to maintain the system in equilibrium. Also,
determine the reactions at supports C and D of the bracket. Given:
rs = 6in. and rp = 4in.

Solution The free-body diagrams of the entire system and its


members are shown. There are nine unknowns. The number
of equilibrium equations is 3 X (3 multiforce members) = 9.
The structure is statically determinate.

Gear A.

@zF,=0 A,=0
@zm,=0
~ 1200+ F(6)=0 F=200Ib-in.
@©sr=0 A,+200=0 A,= —2001b
Gear B.

@s=F, =0 B,=0
@sr=0 -200+B,=0 B,=2001b
@sm,=0 200(4) - M, =0
M, = 800lb-in.Q) <
Bracket.

1200 Ib-in. © F, =0 C, + 200 — 200 = 0 C,=0 <


@©=sM, =0
D,(10) + 200(8.5) — 200(18.5) = 0
D,=2001IbC) <
@sr=0 C€,4+200=0 C,=—200b
C,=200IbQ <

(continued)

350
6.11 Summary 351

EXAMPLE 6.18 (concluded)

Entire system. Equilibrium check: 1200Ib-in. go Ibs;


hgh

@>rE =0 ~200 + 200 = 0 checks


©xzF, =0 0=0 checks
200 1b
@©2sm, =0 200(10) — 1200 — 800=0 checks
The answers are shown in the figure.

200 Ib

6.11 SUMMARY
In this chapter, we have learned the following:

i, Internal forces are the forces exerted on a portion or a member of a


structure by the rest of the structure (Section 6.2).
A truss is formed by connecting straight members at their ends by
hinged connections to form a rigid body (Section 6.3). The analysis of
trusses is based on two simplifying assumptions: first, that all members
are connected only at their ends—by hinged connections in plane
trusses, and by ball-and-socket joints in space trusses; second, that all
external loads and support reactions are applied only at the joints. With
these assumptions, all members of the truss can be treated as two-
force members (Section 6.4).
A truss is statically determinate if all support reactions and member
forces can be determined using the equations of equilibrium (Section
6.6).
The method of joints for the analysis of plane trusses consists of
selecting a joint with two or less unknown forces acting on it, drawing
the free-body diagram of the joint, and applying the two equations of
equilibrium 2F, = 0 and XF, = 0 to calculate the unknown forces.
This procedure is repeated until all desired forces are known (Section
6.7).
The method of sections for the analysis of plane trusses consists of
passing an imaginary section through the members whose forces are
desired, thereby cutting the truss into two parts, drawing the free-body
diagram of one of the two parts of the truss, and applying the three
equilibrium equations {F, = 0, 2F, = 0, and 2M = 0 to determine
the desired forces (Section 6.8).
The analysis of statically determinate frames and machines is based on
the concept that if the entire structure is in equilibrium, then each of
352 Chapter 6 Analysis of Statically Determinate Structures

its members and joints must also be in equilibrium. The analysis of a


plane frame or machine consists of selecting a multiforce member with
three or less unknown forces, or a joint with two or less unknowns,
and applying the equilibrium equations to determine the unknown
forces. Another member or joint is then selected, for which the number
of unknown forces is equal to or less than the number of equilibrium
equations, and the procedure is repeated until all the desired forces are
known (Section 6.10).

LIST OF KEY TERMS

axial force 289 roof truss 294


bridge truss 294 simple truss 298
compression 296 space truss 296
frame 330 statically determinate truss 300
internal force 257 statically indeterminate truss 30/
machine 330 statically unstable truss 302
method of joints 303 tension 296
method of sections 3/7 zero-force member 306
plane truss 294

PROBLEMS

SECTION 6.2
6.1 through 6.3 Identify all two-force members in each structure shown for the
given loading. Draw free-body diagrams of the entire structure, and of each member
and joint of the structure.

Figure P6.1 Figure P6.2 Figure P6.3


Problems 353

6.4 Identify all two-force members in the gripping device used to lift a rough
concrete block as shown. The device exerts vertical lifting forces and horizontal
gripping forces on the block at A and B. Draw free-body diagrams of the entire
assembly, and of each member and joint.

6.5 Draw free-body diagrams of the portions ACGF and DEH of the structure
shown.

Figure P6.4 Figure P6.5

SECTION 6.7
6.6 through 6.9 Identify all zero-force members in the truss shown for the given
loading.
P
H if J K L J

ZINZINZAN AN
_ B Cc D E Bae PAG: va Va’ TN FS
5 z
Figure P6.6 Figure P6.7

Figure P6.8 Figure P6.9


354 Chapter 6 Analysis of Statically Determinate Structures

6.10 and 6.11 Determine the force in each member of the truss shown by the
method of joints. Indicate whether each member is in tension or compression.

300 Ib

i Cc 500 Ib

4ft

A B

—3 ft—+|
Figure P6.10 Figure P6.11

6.12 and 6.13 Determine the force in each member of the truss shown, in terms of
load P, by the method of joints. Indicate whether each member is in tension or
compression.

Figure P6.12 Figure P6.13

6.14 through 6.18 Determine the force in each member of the truss shown by the
method of joints. Indicate whether each member is in tension or compression.

5 kip

Figure P6.14 Figure P6.15


Problems 355

| ip ft—}-4 ft

1 kip
Figure P6.16 Figure P6.17 Figure P6.18

6.19 and 6.20 Determine the force in each member of the truss shown, in terms of
load P and angle 6, by the method of joints.

Figure P6.19 Figure P6.20

6.21 through 6.29 Determine the force in each member of the truss shown by the
method of joints. Take advantage of symmetry if appropriate. Indicate whether each
member is in tension or compression.

ZAZN 3S

+3 1:
20 kN 20 kN 20 kN S
4m—}+—4 m——4 m—+—4 m
iP: as 3 L 3 L 3 aN 3
Sf
Figure P6.21 : eure P6.22
10kN

5 kN

Af 3)
on ‘- OO
eee

[es fobs ft-lbs totes ri 7 be mole2 m-ofe2 moo _

Figure P6.25 Figure P6.26

4kip

Figure P6.27 Figure P6.28

356
Problems 357

SECTION 6.8
6.30 Determine the forces in members DE, DK, and JK of the Pratt bridge truss
shown by the method of sections. Indicate whether each member is in tension
or compression.

6.31 Determine the forces in members CD, CI, and HI of the Pratt bridge truss
shown by the method of sections. Indicate whether each member is in tension
or compression.

6.32 Determine the forces in members FH, GH, and GI of the truss shown by the
method of sections. Indicate whether each member is in tension or compression. [ax
6.33 Determine the forces in members CE, CF, and DF of the truss shown by the k-3 m > 3m 3m ++-3 m—|
method of sections. Indicate whether each member is in tension or compression. :
Figure P6.29

H I ii K L
CRIES EGE RES ERS ES EE > ees

Skip 10kip 3kip 3 kip


Cpa G12 =72 +
Figure P6.30 and P6.31

Figure P6.32 and P6.33


358 Chapter 6 Analysis of Statically Determinate Structures

6.34 Determine the forces in members CD, DI, and IJ of the Howe roof truss
shown by the method of sections. Indicate whether each member is in tension or
compression.

6.35 Determine the forces in members DE, EK, and KL of the Howe roof truss
shown by the method of sections. Indicate whether each member is in tension
or compression.

6.36 The truss supports a sign weighing 200 lb/ft suspended from joints A and C as
shown. Determine the force in member CF by the method of sections. Indicate
whether the member is in tension or compression.

6.37 The truss supports a sign weighing 200 lb/ft suspended from joints A and C as
shown. Determine the forces in members BF and EF by the method of sections.
Indicate whether each member is in tension or compression.

uN : lad w= 200 1b/ft


Skip Skip
bee, panels @ 12 ft = 72 ft 3 panels @ 6 ft = 18 ft

Figure P6.34 and P6.35 Figure P6.36 and P6.37

6.38 Determine the forces in members CD, DK, and KL of the bascule bridge truss
in open position as shown. The bridge is held in equilibrium by support H, and a
cable attached at joint O. Use the method of sections, and indicate whether the
members are in tension or compression.

Figure P6.38
Problems 359

6.39 Determine the forces in members HJ, HR, and QR of the tower truss shown by
the method of sections. Indicate whether each member is in tension or compression.

6.40 Determine the forces in members EG, FG, and FH of the tower truss shown
by the method of sections. Indicate whether each member is in tension or compression.

3 panels @ 5 ft = 15 ft 3 panels @ 5 ft = 15 ft

1 kip! kip 1 kip1 kip 2 kip


2 kip 2 kip2 kip

ZV
Ay Sy Il

10 ft

10 ft

Figure P6.39 and P6.40

6.41 Determine the forces in members CD and HI of the truss shown by the
method of sections. Indicate whether each member is in tension or compression.

6.42 Determine the forces in members D/ and DJ of the truss shown by the method
of sections. Indicate whether each member is in tension or compression.

6 panels
@ 6 m= 36 m
SOkKN S5OkKN 60kN

Figure P6.41 and P6.42


360 Chapter 6 Analysis of Statically Determinate Structures

6.43 Determine the forces in members CD and DG of the truss shown by the
method of sections. Indicate whether each member is in tension or compression.

6.44 Determine the forces in members BD and DF of the truss shown by the
method of sections. Indicate whether each member is in tension or compression.

6.45 Determine the force in member CK of the Baltimore bridge truss shown.
Indicate whether the member is in tension or compression. (Hint: use section aa.)

6.46 Determine the forces in members CD, KN, and NO of the Baltimore bridge
truss shown, by the method of sections. Indicate whether each member is in tension
or compression.

100kKN 5O0kN 200kN 100kN


108 10 ft->b-10ft Lofty ee
Figure P6.43 and P6.44 Figure P6.45 and P6.46

6.47 Determine the forces in members EG and EI of the truss shown by the method
of sections. Indicate whether each member is in tension or compression.

Figure P6.47
Problems 361

*SECTION 6.9
6.48 through 6.52 Determine the force in each member of the space truss shown.
Indicate whether each member is in tension or compression.

va
\
LI 1.5m
et Ee eee Sa

1.5m Ball-and-socket
5 support

Ball-and-socket
support
Zz
Figure P6.48 Figure P6.49

| Ball-and-
socket
|| support

Vy Ball-and-socket
Zz support

Figure P6.50 Figure P6.51


362 Chapter 6 Analysis of Statically Determinate Structures

Figure P6.52

SECTION 6.10
6.53 through 6.55 Determine the forces acting on each member of the frame
shown.
C
7 kip

20 ft

A
1 aN
as 10 ft
Figure P6.53

5 ft—rhe-5 ft 7 ft —+|
Figure P6.54 Figure P6.55
Problems 363

6.56 and 6.57 Determine the forces acting on each member of the three-hinged
semicircular arch shown.

200 KN 150 kN

Figure P6.56 Figure P6.57

6.58 and 6.59 Determine the forces acting on each member of the frame shown.

100kN 200 kN
1 kip/ft

i 10 ft Orel ett
Figure P6.58 Figure P6.59

6.60 Determine the forces acting on each member of the gable frame subjected to
wind loading as shown.

Figure P6.60
364 Chapter 6 Analysis of Statically Determinate Structures

6.61 through 6.63 Determine the forces acting on each member of the frame
shown.

6.64 The single-axle semi-trailer, weighing 40 kip, is attached to the cab of the
truck at joint B, which can be assumed to be a hinged connection. The weight of the
cab is 8 kip as shown. Determine the vertical force acting on each of the three
axles A, C, and D.

10KN 0.25 m

30 kip 20 kip

60/ B C

x 5 re D

L-10 fto-1o tte 10 +fe-10 ft+10 ft

Figure P6.61 Figure P6.62

1.5 kip/ft

Figure P6.64
Problems 365

6.65 through 6.71 Determine the forces acting on each member of the frame
shown.

1 ft
b+}. —§4ft —+ F
BD EB

Pepsyitl |
500 Ib G =

,, ioe
5 ft ja 1000 ke 7)
a7 ee 2.5 ft
fsmotets ‘
Be ft 11 ft
ao 1.5m ree 2 ft
Figure P6.65 Figure P6.66 Figure P6.67

Figure P6.69

12 ft
1 kip/ft

8 ie a EW 2 ft Ginna A ST 2 CR
74 B |1000ib;, , = sor ah aa

5 ft'4ft' 5 ft '4 ft 6ft'6ft! 12ft ' 12ft (Qt Ott Gitt


Figure P6.70 Figure P6.71
366 Chapter 6 Analysis of Statically Determinate Structures

6.72 and 6.73 Determine the forces acting on the smooth rod A, and the members
of the pliers at pin B.

A Seah

icine
& 4.2in—|
Figure P6.72 Figure P6.73

6.74 through 6.76 The gripping tongs are used to lift a rough concrete block as
shown. Assuming no slippage: between the tongs and the block, determine the
horizontal (gripping) and the vertical (lifting) forces acting on the block at A and B.
Also, determine the forces acting on each member of the tongs.

300 300

3,ft 3 ff
300 300

10 ft 600 mm

200 400 200 400 ' 400 | 400


mm mm mm mm mm mm

Figure P6.74 Figure P6.75 Figure P6.76


Problems 367

6.77 Determine the forces acting on the smooth bolt A, and on each member of the
pliers shown.

6.78 The hydraulic crane is used to lift the weight as shown. The crane is attached
to the bed of the truck by a rigid connection at A. Determine the force exerted by the
hydraulic cylinder BD at D. Also, determine the forces acting on each member of
the crane. Neglect the weight of the crane.

6.79 Solve Problem 6.78 taking into consideration a weight of 15 lb/ft for each of
the three members of the crane.

50mm
35 mm
6mm 13mm 23 mm
Figure P6.77 Figure P6.78 and P6.79

6.80 Determine the relationship between the load P required to balance a weight W,
in terms of the distances a and L.

6.81 Solve Problem 6.80 if W = 1000 lb, a = 1 ft, and L = 3 ft. Also, determine
the forces acting on each member of the mechanism.

Figure P6.80 and P6.81


368 Chapter 6 Analysis of Statically Determinate Structures

6.82 A mechanical system of two gears is supported by frictionless bearings on a


cantilever beam as shown. If gear C is subjected to a counterclockwise torque
Mc = 65 N - m, determine the torque which must be applied to gear B to maintain
the system in equilibrium. Also, determine the reactions at support A of the beam.
Given: rg = 200 mm and rc = 100 mm.

6.83 For the mechanism shown, determine the magnitude of the couple M required
to maintain equilibrium, if P = 200 lb and 8 = 80°.
Figure P6.82
6.84 Determine the relationship between the force P and the couple M, in terms of
the angle 6, which is necessary for the mechanism to remain in equilibrium.

6.85 Determine the magnitude of the couple M required to maintain equilibrium of


the mechanism shown, if P = 100 N, L = 2m, and 0 = 30°.

6.86 Determine the relationship between the load P and the couple M in terms of
the angle 6, which is necessary for the structure to remain in equilibrium.

14 in.

7s
Figure P6.83 and P6.84
CHAPTER

DISTRIBUTED LOAD—
ANALYSIS OF BEAMS
AND CABLES

7.1 Introduction OUTLINE

Beams

7.2 Types of Beams


7.3 Internal Forces
7.4 Types of Loads on a Beam
7.5 Shear and Moment Diagrams
7.6 Relations between Distributed Load, Shear Force,
and Bending Moment

Cables

*7,7 Cable Carrying Concentrated Loads


*7.8 Cable Carrying Distributed Loads
*7.9 Cable Subjected to Its Own Weight
7.10 Summary
Key Terms
Problems

369
370 Chapter 7 Distributed Load—Analysis of Beams and Cables

7.1 INTRODUCTION
Structural members that offer resistance to bending due to applied loads are
called beams. Beams are the most common structural elements in use. They
are usually long prismatic bars that support transverse loads. In Chapter 5, we
solved some problems relating to the determination of the reactions at beam
supports. For design of a beam, one must know the relationships between the
applied loads and the internal reactions at any section of the beam. This is the
subject of our study in the first part of the present chapter.
Another important structural member is a flexible cable. Cables are mem-
bers that support only tension; they are used in the construction of suspension
bridges, electric power transmission lines, aerial tramways, and the like. For
proper design of a cable, it is important to know the relationships between the
applied load, the tension in the cable at any point, the span length, and the sag
at any point. This is the subject of our study in the second part of this chapter.

BEAMS

7.2 TYPES OF BEAMS


Beams are most often classified according to the nature in which they are
supported. Beam supports can generally be placed in one of three categories:

1. Simple support or pin connection.


2. Roller support (no friction).
3. Fixed support or clamp.

We discussed various types of support in Chapters 3 and 4. A summary of


these types is also given in Table 7.1. It is important to note that each support
and connection is an idealization of a real support. These idealizations are
necessary to describe the beam mathematically. For example, in Figure 7.1a,
a beam resting on a surface might be represented by a roller as shown in Figure
7.1b. This is an idealization because the beam is not actually supported at a
point, and it probably has some resistance to horizontal motion. The beam in
Figure 7.2a is bolted to a surface. This support might be idealized as a pin
connection as shown in Figure 7.2b. This pin support is an idealization be-
cause, again, the beam is not actually supported at a point, and it does not have
complete freedom to rotate.

(b) (a) (b)


Figure 7.1 Figure 7.2
7.2 Types of Beams 371

Table 7.1 Supports and connections for beams

Possible
Types of reactions at
support or Common support or
connection representations | connection Description

The beam at a pin connection or simple support can rotate


Simple support but cannot translate. The force at the pin can be in any
or pin connection direction, but usually the unknown reactions are given in
the Cartesian coordinate directions.

Roller support The force to the beam is always perpendicular to the surface
(no friction) on which the roller can roll, but it can be toward or away
from the beam.

Fixed support The beam at this connection cannot move; therefore, in


or clamp addition to a force, a moment may act at the fixed support.

Simply supported
- beam or simple beam ST ae Continuous beam

Cantilever beam (a)


en Broppedibeamn

(b)
Overhanging beam
| Fixed beam

(c)
Figure 7.3 Statically determinate beams Figure 7.4 Statically indeterminate beams

Beams supported in such a way that the reactions can be calculated by using
the methods of statics alone are called statically determinate beams. Beams
with more supports than are necessary to provide equilibrium are referred to
as statically indeterminate. Figure 7.3 shows examples of statically determi-
nate beams, and Figure 7.4 shows examples of statically indeterminate beams.
372 Chapter 7 Distributed Load—Analysis of Beams and Cables

7.3 INTERNAL FORCES


Figure 7.5a shows a simple beam supported at A and B, and subjected to a
force F. The support reactions for the beam can be determined by applying
the equations of equilibrium to the free-body diagram of the entire beam
(Figure 7.55). The equations of equilibrium are

Lh= 08 Omni 0
(a) The support reactions including the force(s) applied on the beam are referred
to as external forces. Now let us pass an imaginary section through the beam
at C. Figure 7.5c shows the free-body diagrams of the cut sections AC and
7: \ BC. The equilibrium of the beam sections AC and BC can be maintained if
4 |
lt ‘AS
the force components K. and V. and moment Mc_ are developed at the cut
section. These forces and moment are defined as internal forces and internal
moment. The internal force KF. is an axial force; V-, which 1s parallel to the
cut section, is referred to as the shear force. The moment Mc is referred to as
External
loading (b) the bending moment.
In this chapter, we will learn to determine the variation of the internal
loading (that is, shear force V and bending moment M) along the length of a
beam. In order to do so, we need to develop a sign convention and be careful
and consistent in using it, as distinct from the convention used for external
forces and moment. Following are the sign conventions generally used for
external and internal loadings.

Internal loading

(c) Sign Convention for External Loading


Figure 7.5
Force Considered positive if directed upward (‘ ).

Bending Moment Considered positive if directed clockwise ().

Sign Convention for Internal Loading


Force Considered positive if directed downward (| ) in the left-hand side of
the cut and directed upward ( 1 ) in the right-hand side of the cut (Figure 7.6).

Bending Moment Considered positive if counterclockwise (() in the left-


hand side of the cut and clockwise () in the right-hand side of the cut (Figure
Vv 7.6).
a Another way to remember the sign conventions for shear force V and bend-
ing moment M is to consider the external forces. The shear force V is positive
if the external forces are such that the beam tends to shear off at the cut section
V as shown in Figure 7.7a. Similarly, the bending moment M is positive if the
Figure 7.6 Positive sign conventions for internal V external forces are such that they tend to bend the beam at the cut section as
and M shown in Figure 7.75.
7.5 Shear and Moment Diagrams 373

7.4 TYPES OF LOADS ON A BEAM Vv


Loads acting on a beam may be of several types, and some of them can be
described with reference to Figure 7.8.

1. Concentrated load—such as force F acting at a point on the beam.


2. Uniformly distributed load—such as the load of intensity w,. The unit
for this type of load is expressed as force per unit length (for example, Vv
Ib/ft, KN/m) along the axis of the beam. (a)
3. Linearly varying distributed load—a distributed load that varies in
intensity, such as w, to w3, over a certain length of a beam. M M
4. Couple—such as couple moment M, acting at a point on the beam.

Distributed loads are not always uniform or linearly varying; there are other
types of distributed loads—for example, quadratic, cubic, and sinusoidal—that (b)
can be applied on the beam. Figure 7.7

Figure 7.8

7.5 SHEAR AND MOMENT DIAGRAMS


Often in the design of beams, it is necessary to draw the variation of the shear
force and bending moment along the length of the beam. Two methods are.
used to do so: the method of sections, and the geometrical method, using
known relations between load, shear, and bending moment. The geometrical
method will be discussed in detail in Section 7.6. Following is the procedure
for drawing the shear and bending-moment diagrams by the method of
sections.
1. Determine the reactions (forces and couples) at the beam supports. Re-
solve the forces into components in the directions parallel and perpendicular
to the beam axis.
2. Establish the origin of the x axis at the left end of the beam. Draw sections
in the regions located between the concentrated forces and couples, regions
where there is no discontinuity of the distributed loading.
An example of this procedure is shown in Figure 7.9. Section 1 is between
the vertical reaction at A and the concentrated force F applied at point a.
Section 2 is located between the concentrated force F at a and the uniformly
distributed load of w, lb/ft (or KN/m) which begins at b. Section 3 is in the — Figure 7.9
374 Chapter 7 Distributed Load—Analysis of Beams and Cables

region of the uniformly distributed load (i.e., b to c). At c there is a disconti-


nuity in the distributed load. Hence, section 4 is located between c and d.
Sections 5 and 6 are located, respectively, between the end of the distributed
load and the couple M, (i.e., d to e), and between the couple and the support
reaction at B.
3. Draw the free-body diagram of the entire left-hand side of the beam at
each section. The shear force V and the bending moment M at the cut section
should be shown in their positive senses.
4. Use the equations of equilibrium to find the shear force and the bending
moment.

OxF (vertical) = 0

© 2M a cut about the centroid of =U


the cross-sectional area)

5. Plot the variation of the internal shear force and bending moment with
distance x directly below the free-body diagram of the beam. If the shear force
is positive, it should be plotted above the x axis, and if negative, it should be
plotted below the x axis. The plot of the bending-moment diagram should also
be done in a similar manner.
The following example problems will help to illustrate the procedure for
developing the shear-force and the bending-moment diagrams.

EXAMPLE /7.1

For the simple beam shown, draw the shear-force and bending-
moment diagrams.

Solution The free-body diagram for the entire beam is shown.


The magnitude of the reactions at the left and right supports are

Ie Pa
R,= a and Rz = TA

In order to develop the shear-force and bending-moment


diagrams we write equations for V and M in terms of x for any
section between 0 = x = a and also for any section between
a =x =L. Consider the left-hand portion of the beam for
a section made between 0 = x S a (section 1). Note that, at the
cut, the shear force and the bending moment are shown in the
positive sense. For equilibrium, the sum of the vertical

(continued)
7.5 Shear and Moment Diagrams 375

EXAMPLE 7.1 (concluded)

forces, must be zero; thus,

OOS Fe siican ==) iF. a vy, =o

V Pb (a)
S SS a
ae
Again considering for moment equilibrium,

Ora section 1) = 0

Pb
(Foo PV oO) 4M, =0

Me Pbx
i= Sr (b)
Section 1
(O<x Sa)
Now, consider the left-hand portion of the beam for a
section made between a S x S L (section 2). It is important to
note that the figure shows the entire length of the beam
located to the left of section 2—not the length of the beam Section 2
between the force P and the cut. For equilibrium, then, (a<x <L)
Ry
Os ay 0

Pb
ee 0
ee (Eby Pa
= re kT (c)

Equation (c)
Ora section 2) an 0
|
Equation (d)
Pb
-(@)oy + P(x—a)+M,=0 |

Pa
M, = ae — x) (d)

The plots of the variation of V and M along the axis of the


beam are shown.

Important note: It is easy to see that the change in shear force


due to a concentrated load is equal to the magnitude of the
force itself. For example, at point C, the net change of shear
force is equal to

which is the magnitude of the concentrated load at that point.


376 Chapter 7 Distributed Load—Analysis of Beams and Cables

EXAMPLE 7.2

W = (4)(10) (1000) = 5000 Ib A simple beam subjected to a linearly increasing distributed load
is shown. Draw the shear-force and bending-moment diagrams
for the beam.

w = 1000 1b/ft Solution The magnitude of the reactions at the supports are

R, = 1667 |b and Rz = 3333 Ib


The free-body diagram of the beam located to the left of
Section 1 is shown, for which we find
Rp = 33331b
Ope a, a 0

. 1
Re (5)eoxi00n —V,=0

V, = R4 — 50x? = 1667 — 50x? (Ib) (a)


(4)(x) (100x) = 50x7Ib The distance x at which V, is equal to zero can be obtained by
5x calculating
y= (2289) (x) = 100x Ib/ft V, = 0 = 1667 — 50x?
1667
esr
i 50
Now for the moment, in a similar way we write
16671b P
Orme section 1) = 0

1 1
NI 3333 Ib =Npe8 3 |
(5)et000 |(3x)+M,=0

Equation (b) 50 50
M, = Ryx — Fx" = 1667 — x" (Ib - ft) (b)

The plots of V and M along the axis of the beam are shown.
7.5 Shear and Moment Diagrams 377

EXAMPLE 7.3

A cantilever beam is subjected to a couple and a concentrated


load. Draw the shear-force and the bending-moment diagrams for
the beam.

Solution The reactions at the fixed end of the beam are


shown. The free-body diagram of the left side of the beam
sectioned at 1 is shown, from which we see
Section |
Oo = 0 (0 <x <5 ft)
2-—V,=0 V, = 2 kip (a) 17 kip: ft x
2 kip
OrMa section 1) = 0
Section 2
7) = ese WE M, = -17+ 2x (b) (5 ft <x < 10 ft) (
17 kip:ft
Atx = 5 ft,M, = —7kip- ft.
2 kip
Now consider the free-body diagram of the left side of the
beam sectioned at 2. iG Equatio
2 Kip oe oT
Opie wee mg 0

V> = 2 kip
Osa section 2) =H

17 — 2x +34+M,=0
M, = —20 + 2x (kip ft) (d)
The complete shear-force and bending-moment diagrams are Equation (d)
shown. —10kip-ft
Equation
(b)
Important note: It is easy to see that the change in bending _17kip:ft
moment due to a concentrated couple is equal to the magnitude
of the couple. For example, at point B, the change in bending
moment is —7 — (—10) = 3, which is the magnitude of
the concentrated couple at that point.
378 Chapter 7 Distributed Load—Analysis of Beams and Cables

EXAMPLE 7.4

A cantilever beam is subjected to a point load and a linearly


M , =33,750|b-ft 1500 1b/ft
varying distributed load. Draw complete shear-force and bending-
( aoe 500 Ib (2)
moment diagrams for the beam.

¢ coeenae Seale Solution The magnitude of the reactions at the support, R,


4 rs
5 ft 5ft/_+— and M,, are shown. The magnitude of R, is 4250 lb, and
R ,=42501b | the magnitude of M, is 33,750 lb - ft. The free-body diagram
| of the beam located to the left of section 1 is shown
Section 1 | | (0 =x <5 ft). For equilibrium,
(0 Sx S55 ft) M, |
Fiala | | |
| Cosas = 0
| 4250-—V,=0. V, =42501b (a)
ae y | |
=e |
|
|
|
|| QO sMa section 1) = 0

33,750 — (4250)(x) + M, = 0
Section
At]
2 +[300(x |—5)](x —5) Ib
(Sft<x <10ft) Wee aye M, = 4250x — 33,750 (lb -ft) (b)
ihe Gas
SO0Ib 4 ae
The free-body diagram of the beam located to the left of
33,750 1b-ft }
section 2 is shown (5 ft = x = 10 ft).
1 ie
(Osean =)
4250 1b ‘
& w= 1500 (x| — 5) =300(x. — 5) ! Ib/ft
Vib) rat) 4250 — 500 — (3)
[(300)@ — 5)7] —V,=0
|
4250 _7 Equation (a) |
Equation (c) | V, = 3750 — 150(« — 5)? (Ib) (c)
|
I OsMa section 2) = 0
|
33,750 — 4250(x) + (500) — 5)

M (ib: ft)
+ 3100y¢« ~ sy +M,=0
M, = 3750x — 31,250 — 50(x — 5)? (Ib- ft) (d)
Equation (d) The plots of the shear force and bending moment along the
beam axis are shown.

7.6 RELATIONS BETWEEN DISTRIBUTED LOAD,


SHEAR FORCE, AND BENDING MOMENT
When a beam is subjected to a number of point loads, couples, and distributed
loads, the method of plotting shear-force and bending-moment diagrams as
outlined in Section 7.5 may become tedious. In this section we will derive two
differential equations that will greatly facilitate the plotting of shear-force and
7.6 Relations between Distributed Load, Shear Force, and Bending Moment 379

bending-moment diagrams for beams. The first differential equation will be a


relationship between shear force and load, and the second differential equation
will relate bending moment and shear force.
In order to do that, we consider a beam subjected to a distributed load w(x) w(x)

per unit length of the beam as shown in Figure 7.10a. Let CD be an elemental
segment of the beam having a length Ax. The free-body diagram of the seg-
ment CD is shown in Figure 7.10b. The shear force of magnitude V and bend-
ing moment of magnitude M along the C face will be assumed to be positive.
Along the D face, the magnitudes of the shear force and bending moments are
V + AV and M + AM, respectively.
DFoericay =0 V-WV+AV)+w@Ax=0 AV=w) Ax
Dividing both sides by Ax and taking the limit when Ax — 0, we obtain the
relationship between shear force and distributed load:

ape WU(x) (7.1)

In deriving Equation (7.1) we have assumed that distributed loads are positive
when acting upward and negative when acting downward. Now, taking the
moment about the centroid of the section at D, Figure 7.10

@©2M,=0 —M-—(V)(Ax) + (M + AM) — [wQd) Ax] (25)=0

AM=VAx+ (3)[w(x)](Ax)?

Again dividing both sides by Ax and taking the limit when Ax — 0, we obtain
the relationship between bending moment and shear force:

—==V Cie2)

The implications of Equations (7.1) and (7.2) are as follows:

Equation (7.1)
1. The slope of the shear curve dV/dx is equal to w(x).
2. Equation (7.1) is not valid at a point where a concentrated load is
applied since the shear curve becomes discontinuous at that point. (For
example, see the plot of shear curve below point C in Figure 7.11.)
3. Referring to Figure 7.10a and integrating Equation (7.1) between points
C and E gives Figure 7.11
380 Chapter 7 Distributed Load—Analysis of Beams and Cables

XE
Ve-Vo= I w(x) dx (7/23)

or

V; — Vc = area of the load diagram between Cand E (7.4)

4. Equations (7.3) and (7.4), which are derived from Equation (7.1), are
also not valid if a point load exists between points C and E.

Equation (7.2)

1. The slope dM/dx of the bénding-moment curve at a given point is equal


to the value of the shear force V.
2. Equation (7.2) is not valid at a point where a concentrated load is
applied. (For example, see the shear and bending-moment curves below
point C in Example Problem 7.1.)
3. Referring to Figure 7.10a and integrating Equation (7.2) between C and
E gives

XE

XG

or

M, — Mc = area of the shear diagram between C and E (7.6)

4. Equations (7.5) and (7.6), which are derived from Equation (7.2), are
not valid if a couple is applied at a point between C and E since the
bending-moment curve becomes discontinuous at that point. (See the
plot of the bending-moment curve at point C in Figure 7.12.)
5. Equations (7.5) and (7.6) are valid if point loads exist between points C
and E.
6. The absolute value of the bending moment is maximum where dM/dx
= O(ie., V = 0).

Figure 7.12
EXAMPLE 7.5

A cantilever beam is subjected to a uniformly distributed load.


Draw the shear-force and the bending-moment diagrams.

Solution The reactions at the fixed end A are shown.


2

R, = wh) M, =>

Based on the sign conventions shown in Figure 7.7, the shear


force at A is +wL. From Equation (7.1) we write

The negative sign on the right-hand side of the equation is due


to the fact that the distributed load is acting downward. For
the shear force, then,
V= —Jwdx= —wx+C,
where C, is a constant. Atx = 0,V = +wlL. Thus
wL = —(w)(0) + C,

C, = wl
Hence
VS We (a)
wl Equation (a)
The above equation for V is a straight line with a slope of — w.
The plot of the shear-force diagram is shown. Now, from
Equation (7.2), for the bending moment we write

dM
—=V= —-wx+ wl
dx M
M = JV dx = J(—wx + wL) dx
2
M = —"— + whx + Cy (b)

where C, is a constant. Atx = 0,M = —(wL’)/2 (see sign


convention in Figure 7.7). Thus ss Equation (d)
iL?
_*= = —2C + wl) +,
2 2
wL?
2
CC 5 (c)

Combining Equations (b) and (c) gives

wx? wL?
M SS ae ap HLS S 5 (d)
d

Equation (d) shows that atx = 0, M = —(wL’)/2, and at x =


L, M = 0. The plot of the moment diagram is shown. The
magnitude of the maximum moment is wL?/2.

381
382 Chapter 7 Distributed Load—Analysis of Beams and Cables

EXAMPLE 7.6

A cantilever beam has a uniformly distributed load on one half


of the beam. Draw the shear-force and bending-moment
diagrams.

Solution The support reactions are shown. Based on the sign


conventions shown in Figure 7.7,
B
af wL
— mM
3wL?
V Equation (b) || 2 - 8
Equation (e) | For the segment AB, from Equation (7.1) we can write, for the
|
relationship between shear force and distributed load,

V=—-wx+C,

ATA es — Olandive—Wig/2aSo
L
Va —wr + (a)

Between A and B, w = O. Therefore,

vL,
= 7 (constant) (b)
3wL? :
IT Equation (c) For the segment AB, from Equation (7.2) we can write the
relationship between shear force and bending moment as
wh
Wl = (E)o ap (C5

AtA, x = Oand M = —(3wL’)/8. So


3wL?
C,=-
8

Thus

wL 3wL?
Mi= (=)(63 = 8 (c)

From Equation (c), note that, at x = L/2

ma (E\(L) _ 30?
_ wl’
Ny) 8
Now for the segment BC, Equation (7.1) gives

V= -wx+C, (d)

(continued)
7.6 Relations between Distributed Load, Shear Force, and Bending Moment 383

EXAMPLE 7.6 (concluded)

where C; is a constant. At x = L/2, V = wL/2. Therefore At x = L/2,M = —(wL’)/8. So we can compute


wL? W \i fle L
Ee
pi 5
(en) fet 3 —-— = —(—](—)+@b|(=)+c,
8 2/\4 2,

C3 — wl
pumas
Thus
2
Therefore,
V = —wx + wl (e)
wx? wL?
From Equation (e) we can see that the variation of V with x is a M = -—+wlx-—
ye vd 5 f
(f)
straight line with a slope of —w. At x = L the value of V is
zero. From Equation (7.2), for segment BC From Equation (f) it can be seen that, atx = L, M = 0. The
dM shear and bending-moment diagrams are plotted.
— =V= —-wx+ wh
dx et ee

Equation (e)

wx?
M= ere ACs

EXAMPLE /7.7

For the simply supported beam shown, draw the shear-force and
8kip 9kip 4kip
Be . .

bending-moment diagrams.

Solution The support reactions at A and B are shown as R, = c D E


11.5 kip and Rz = 9.5 kip. t~-5 ft +t-5 ft ft
We can take sections in the regions AC, CD, DE, and EB ! | R,=9.5 kip
1 |

and draw the shear-force diagram as shown. This also could


|
H|
||
have been done by using Equation (7.1). For example, for |
|
region AC |
|
|
|
dV
i
V=wxt+C,

Using the sign convention shown in Figure 7.7, at x = 0,


V = +11.5 kip. In the region AC where there is no distributed
load, w = O. Thus we have

V=11.5=(0)@)+C,
V =C, = 11.5 kip (constant) (continued)
384 Chapter 7 Distributed Load—Analysis of Beams and Cables

EXAMPLE 7.7 (concluded)

Similarly, for the region CD and B are zero. With this in mind and referring to Equation
(7.6),
V=wx+t+C,
Mc — M, = area under the shear force curve between C and A
Atx = 5 ft, V = 11.5 — 8 = 3.5 kip (based on the sign
convention of Figure 7.7). For this region w = 0. So = (11.5 kip)(5 ft) = +57.5 kip-ft

V=3.5=(0)@)+C, Since M, = 0,

V = C, = 3.5 kip (constant) Mc = +575 kip -ft

Similar logic follows for regions DE and EB. We now plot the Similarly
bending-moment diagram. We know that for any given segment
Mp — Mc = Mp — 57.5
(that is AC, CD, DE, and EB), the shear force is constant.
From Equation (7.2), = area under the shear force curve betweenD and C

dM = 3.5)G)
— = V = constant = C
dx Mp = +57.5 + (3.5)(5) = 75 kip:ft
MeV — Veto) (©)
M=Cx+C'
M,, = 75 — 27.5 = 47.5 kip:ft
(f
M, — M; = M, — 47.5 = +[(—9.5)(5)] = —47.S kip
-ft
constant
Thus we prove that M, = 0. The plot of the bending-moment
This means that the variation of moment will be a straight line diagram along the x axis is shown.
when plotted against x. We also know that the moments at A

CABLES

Cables are used in suspension bridges, electric transmission lines, aerial tram-
ways, and many other structures. The analysis of cables can be divided into
three major categories:
1. Cables carrying concentrated loads.
2. Cables carrying distributed loads.
3. Cables subjected to their own weight.
In the following sections we will discuss each of the above types of cable.

*7.7 CABLE CARRYING CONCENTRATED LOADS


Figure 7.13a shows a cable ABCD carrying two concentrated loads of mag-
nitudes P, and P,. Let the magnitudes of the loads and the distances L,, L,,
and L3, and / be known. In order to determine the tensions in the cable seg-
ments, let one of the sags (such as y,) be known. It will be assumed that (a)
the cable is flexible, and (b) the weight of the cable is negligible. Thus, any
portion of the cable between two concentrated loads may be considered as a
7.7 Cable Carrying Concentrated Loads 385

Figure 7.13

two-force member. Figure 7.13b shows the free-body diagram of the entire
cable. Let A, and A, and D, and D, be the scalar components of the reactions
at A and D, respectively. Assuming that L,, L,, L3, /, y,, P;, and P, are known,
the following procedure may be used to determine the tensions in the cable
segments, and the sag y,. Considering equilibrium, from Figure 7.13b we write

@2F, =0 —A,+D,=0
AnD, (7.7)
@=F, =0 A, + D, — P, — P, =0
A, + D, =P, + Py (7.8)
@©2>M, =0
+A) — AZ, + L, + Ly) + Pi, + Ls) + P(L3) = 0 C59)
386 Chapter 7 Distributed Load—Analysis of Beams and Cables

Since there are four unknowns and three equations [that is, Equations (7.7),
(7.8), and (7.9)], one more equation is needed to solve for A,, Ay, D,, and DS
In order to develop that equation, consider the free-body diagram of the cable
segment AB shown in Figure 7.13c. The magnitude of tension in the cable
segment BC is Tz.
@©>M, =0 +A,y, — AyL, = 0 (7.10)
Using Equations (7.7), (7.8), (7.9), and (7.10), we can now solve for A,, Ay
D,, and D,.
It is essential to realize that segment AB is a two-force member, as are BC
and CD. Figure 7.13d shows the forces that constitute the equilibrium of point
A. Thus

@zF,=0 TiniCOS, Ope AO


A, = Typ COS Op

Ty3 = Ay as 7/03)
COS O42
Since A, and 6,4, = tan~! (y,/L,) are known, T,, can be calculated. Referring
to Figure 7.13c again,

@ =F. =0 Tyo COS Apc — A, = 0


A,. = Tgc COS Og¢ (7.12)

which we may write as

From Equations (7.12) and (7.13) the magnitudes of 7, and 8, can be cal-
culated since A,, A,, and P, are known. Thus we can calculate the sag y> as
yY2 = y, + Ly tan Oe¢ (7.14)

Figure 7.13e shows the free-body diagram of the cable segment ABC. Tp is
the magnitude of the tension in the cable segment CD. Also,

For equilibrium,

@ IF, =0 Ten COS 0a =A,


A
Boner,
Tcp = = (7.15)
af

It is important to note that the horizontal component of the cable tension is


the same for all segments. For instance,
7.7 Cable Carrying Concentrated Loads 387

A,= Typ cosO4n = TgcC0SO8g¢ = Tep cos8cp

Equation(7.11) Equation(7.12) Equation(7.15)

From the preceding relationship it is obvious that the maximum tension will
occur in the section of the cable that has the largest value of 0 (i.e., smallest
value of cos 0). Following is a numerical example to illustrate the above
concept.

EXAMPLE 7.8

ABCDE is a cable subjected to three concentrated loads. Deter- (Note: Figure is not to scale.)
mine the tensions in the cable, T4,, Tgc, Tcp, and Tp,p.

Solution The free-body diagram of the whole body is shown.


For equilibrium,

OF, —0 =k, (a)


®28, =0 A, + £, = 300 + 500 + 250 = 1050 (b)
@sm, =0
—(A,)(20) — (A,)(4) + (300)(17) + (500)(11) + (250)(5) = 0
20A, + 4A, = 11,850 (c)
The free-body diagram for the segment ABC of the cable is
shown. From this
@sM- =0
— (Ay)(9) + (A,)(0.7) + (300)(6) = 0
or
OA en UJAp a leOU) (d)
Solving Equations (c) and (d), we obtain

A, = 1413 b©)
A, = 309.9 Ib)
Now we can compute the tension in segment AB:
Tyg = V Az + AZ = V (1413)? + (309.9)7
Typ = 1446.58 lb <
Also, note that the angle 8,, with respect to the horizontal can
be found to be
3001b S001
Oap = wn-1(4) = in (222)
1413
= 12.37° es (tans Ot =i 6 els
x,

(continued)
388 Chapter 7 Distributed Load—Analysis of Beams and Cables

EXAMPLE 7.8 (concluded)

Next, consider the system of forces for equilibrium at B.


@ sF, =0
Tgc COS Ogc = Typ COS 94g = A, = 1413 Ib (e)
@®zF, =0
Tgc sin 9gc + 300 = Tyg sin O42 = A, = 309.9 Ib
Tgc Sin Ogc = 9.9 Ib (f)
So, we can solve for the tension in segment BC with
500 Ib
3 ft >b— 6ft—>| Tgc = 1413 Ib <«

Also
T 4p= 1446.581b
Osc = tan! (22,) = 0.4°

Considering the system of forces for equilibrium at C,

@>F, =0
Tcp COS 8cp = Tgc COS Ogc = A, = 1413 lb (g)

300 lb =F =0
Tep Sin 8cp + Tac Sin gc = 500
Substituting T,. sin 8,- = 9.9 Ib into the above equation gives

Tcp Sin 8¢p = 490.1 Ib (h)


Dep =.V (Len 60s Ueno ta op Sin Oana
= V(1413)?
+(490.1?
500 1b Top = 1495.58 lb <
Consider the system of forces for equilibrium at E. From
Equations (a) and (b),

E,. = A, = 1413 Ib
E, = 1050'— A, = 1050 — 309.9 = 740.1 Ib

Tp = VEZ + EP= V(1413)? + (740.1)?


Tpr = 1595 lb <
7.8 Cable Carrying Distributed Loads 389

*7.8 CABLE CARRYING DISTRIBUTED LOADS


Figure 7.14a shows a cable ABC which is flexible and has negligible self
weight. It is subjected to a distributed load. The point B is the lowest point in
the cable and it is taken as the origin. In order to find the magnitude of the
cable tension at any point D, we consider the segment BD (Figure 7.14b). Let
Tg and Tp be the magnitudes of the cable tension at B and D, respectively.
The magnitude of the resultant of the distributed load between points B and
D is equal to W. For equilibrium to be satisfied, we find that

@2F, =0 Tp cos = To (7.16)

OF, =0 Tp sin
0= W (7.17)
and

@-m, =0 ~Toy + WL —x) =0 (7.18)


If W, L, x, and y are known, the magnitude of T, can be calculated from
Equation (7.18) as
meV x)
ie (7.19)
y
Again, from Equations (7.16) and (7.17), we can write the equation for the
magnitude of the tension at any point D as

Ty = VT2, + W? (7.20)

Figure 7.14
390 Chapter 7 Distributed Load—Analysis of Beams and Cables

Also, the angle of the cable with respect to the horizontal at any point D can
be expressed as

tan 0 al (7.21)
To ,

We can make an important observation from Equation (7.20). If point D co-


incides with B, then W = O and Tp = To. This means that the tension is
minimum at the lowest point in the cable. Also, we can see from Equation
(7.16) that the magnitude of the horizontal component of the tension at any
point is constant.

Parabolic Cable
A special case of the above analysis is the condition where the cable carries a
horizontally uniformly distributed load (Figure 7.15a). This is a close approx-
imation to the condition of a suspension bridge. Figure 7.155 shows the free-
body diagram of the section BD with point B being the lowest point of the
cable. For this case W = wx and x = x/2. Hence, from Equation (7.19) we
find that the magnitude of tension at the origin can be written

facets eaten Fe A (7.22)


Ms y 2y
Force/unit length = w/unit length
From Equations (7.20) and (7.21) we have the equations for the magnitude of
(a) the tension at point D, and the angle at D with the horizontal, as

Tp = V Te + (wx)? (E23)

and

wx
tan 6) = —— (7.24)
To

Figure 7.15 y=— (725)

The preceding relation is an equation for a parabola with its vertex at B;


therefore, this is referred to as a parabolic cable.
7.8 Cable Carrying Distributed Loads 391

From Equation (7.23), we can see that the magnitude of the tension in the
cable will be maximum at a point where the magnitude of x is maximum.
Referring to Figure 7.15a, the magnitude of x is maximum at point A. Thus,
for maximum tension, we substitute x = L, in Equation (7.23) to obtain T,,,,:

ihm ee VT WL)?

Again, at point A, x = L, and y = H,. Substituting these values for x and y


in Equation (7.22), we obtain

wl},
O- 3H (7.26)
A

Hence we determine the magnitude of maximum tension as

Sef pia \: wl, ib Se (4)


beNg G2)
Maes Te (24) a (wL 4)* =

The length of the cable S, (that is, from B to A) can be determined thus:

S, =JSds

Referring to Figure 7.15a, we can write

(dS = (dx + (ayy? oor 2SdS = 4 + (2) dx


dx

Therefore,

La D
d
s,= |as= | My a PN ae
9 dx

However, from Equation (7.24)

This gives

La D
St =[as=| Mace (=) Je (7.28)

From binomial expansion, we can write

n(n — ay fi(he Oe J =
Cis) =| tert 1 31
392 Chapter 7 Distributed Load—Analysis of Beams and Cables

Substituting n = 4 and (wx/T,) for x in the above expression, we obtain


2 PP? 44
wx wx wx
oy) =14+—5-S
+: 7.29
es (=) Me” ee Ca
Finally, combining Equations (7.26), (7.28), and (7.29), we obtain

ol.
2 4
2H. 2a,
Si Lt S| ee eee 7.30

Similarly, the cable length S, (that is, between B and C) can be given as

sels 4 2(Hey 2 Gem


: 2 4

a ae BL ONES

The series in Equations (7.30) and (7.31) converge for values of


H,/L, < 0.5 and H-/Le < 0.5. It will thus be sufficiently accurate to take
only the first two terms for our calculations. Hence

Pa

S,= Lf)ufz(t) (7.32)

and

S> = Le||+ 2(He)|


2H (7.33)

The total length of the cable, then, is

2 25
2 (H 2 (H
S=S,+S,=Ly tLe + Zh (4) + =Le (2) (7.34)
‘A
EXAMPLE 7.9

A cable supports a uniformly distributed load of 100 Ib per foot


of the horizontal length. The cable is suspended from points A
and B which are at the same elevation and 200 ft apart. Determine
(a) the minimum tension in the cable, (b) the maximum tension
in the cable, and (c) the total length of the cable. The lowest point
on the cable is 50 ft below A and B.

Solution w= 100 1b/ft


Part a. Equation (7.25) gives
wx? wx?
yy - or nS
2To

For this problem, at x = L/2 = 200/2 = 100 ft, the


magnitude of y is 50 ft. Thus
2

al
- (;) _ wl? _ (100)(200)
Q 2H 8H (850)
To= 10,0001b <
Part b. Refer to Equation (7.27). For this problem T7,,,, will
occur at both A and B since they are at the same elevation. So

2 2,

iter |aia
2H,
ee |2Hp
We know that w = 100 lb/ft, L, = Lz = 100 ft = L/2, and
H, = Hz = 50 ft = H. Substituting these values, then, gives

wL TaN
Tmax
= | 1 + |—
2 4H
pen) 200 ]
Sc cele
(4)(50)
Tmax = 14,142 Ib <
Part c. Referring to Equations (7.32) and (7.33), the total
cable length S$ will be

een
eA BDa

Again, L, = L/2 and H, = H. Substituting these gives

g-(2\f14
Rae
(2)(2)
BND
|afi 2(#)
Dyien| ae 3\L

sam +(3)(30) |
Ph
8)\/ 50
= + {/-— —_—

S=233ft <

393
394 Chapter 7 Distributed Load—Analysis of Beams and Cables

*7.9 CABLE SUBJECTED TO ITS OWN WEIGHT


When the weight of the cable becomes important, the analysis must be done
assuming a uniformly distributed load along the length of the cable. Figure
7.16a shows a cable with its lowest point at B. Let the cable weight be w per
unit length. Figure 7.16b shows the free-body diagram of the segment BD of
the cable. Let T, and T be the magnitudes of the tensions at B and D, respec-
tively. Also note that the length of the cable between B and D is equal to S.
w/Unit length The magnitude of the resultant of the distributed load between B and D, then,
of cable is

W = Sw (7.35)
(a) For equilibrium, we write

@Qryr=0 T cos 0 = Ty (7.36)


Ose T sin 0 = W “(130
So
T=V iW Vee Sus
Let us assume that

To = wp (7.38)

where p = aconstant (a specific distance—see Figures 7.16a and 7.16b). Thus,


the magnitude of the tension at a point in a cable subjected to its own weight
can be described as
(b)
Figure 7.16

T = Viwpy? + (Sw)? = wVp? + S? (7.39)


The value of S can be calculated if we can develop an equation for it in
terms of x and y, such as will follow. Referring to Figure 7.16b, we write

Sas
iS cos

and from Equation (7.36),

To
Ot
T cos A)

Therefore

cadets
d ip eres ffleek ie ds ied.
dS th TF wVp2 +82 Vp? +8?
7.9 Cable Subjected to Its Own Weight 395

If we choose the origin as shown in Figure 7.16a, then*


[a a Ss pds

0 oa 0 Vp? + S?

thus we can write the equation for the coordinate x of a point as

S
x=p sinh*(8) \ (7.40)
Pp

Again from Figure 7.16b,


dy
ae
— = tan 0

Combining Equations (7.36) and (7.37), we obtain

W
(Ang) ==
To
Therefore,

De Ves
ax. I, wp
s
dy = (5)ax (7.41)
P
From Equation (7.40)
S
a sinh(#) (7.42)
Pp Pp

Combining Equations (7.41) and (7.42) gives

[+= [el
a = ele

y—p=pcosh| —} — p
Dp

58
y= peoss(*) (7.43)

*The integration can be found in any standard table for integrals.


tSee Appendix D (Sections C, D, and F) for mathematical relationships of hyperbolic
functions.
396 Chapter 7 Distributed Load—Analysis of Beams and Cables

This is the equation of a catenary with a vertical axis. For that reason, cables
hanging under their own weight are referred to as catenary cables.
Equations (7.42) and (7.43) yield, for the value of S,

=~ -p? (7.44)
Combining Equations (7.39) and (7.44), we obtain

T = wy (7.45)

EXAMPLE 7.10

|. L=200ft Refer to Example 7.9. Replace the uniformly distributed load of


100 lb/ft of horizontal length by 100 Ib/unit length of the cable.
Determine (a) the minimum tension in the cable, (b) the maxi-
mum tension in the cable, and (c) the length of the cable.

Solution For point A,


1000 1b/Unit x = 100 ft
length of cable (a)
y =(p + 50) ft
From Equation (7.43),
1
y = pcosh (
*)= pcosh (2) (b)
p Pp
Combining Equations (a) and (b),

100 5
p +.50 =p cosh (2) or ae a = cosn(22)
p P P
By trial and error, we find that p = 107.43 ft. Thus,
Yq = p + 50 = 107.43 + 50 = 157.43 ft
Part a. From Equation (7.38),
T min = To = wp = (100)(107.43)
Trin = 10,743 Ib <
Part b. From Equation (7.45),

Tmax = [4 = Wy, = (100)(157.43)


Tmax = 15,743 lb <
Part c. From Equation (7.44),
S? = y? = p?
Total length is

2S = 2V y2 — p* = 2V 157.43)" — (107.43)
2S = 230.2 ft <
7.10 Summary 397

7.10 SUMMARY
In this chapter, we have covered the following:

1. The classification of various types of beams based on their end supports


(Section 7.2).
2. The definition of internal forces and moments in a beam and their sign
conventions (Section 7.3).
3. The definition of various types of loading on a beam (Section 7.4).
The procedure to draw shear-force and bending-moment diagrams for
beams (Section 7.5).
5. The relationship between distributed load, shear, and moment as
expressed by the following equations (Section 7.6):
dV
re = wr) (7.1)

dM _
=m =V GE)

where x = distance
V = internal shear force
M = internal moment

These relations can be used to draw shear-force and bending-moment


diagrams.
6. The procedure to determine the tension and sag in various segments of a
flexible cable of negligible self weight carrying concentrated loads
(Section 7.7).
7. The procedure to determine the tension at various points of a flexible
cable of negligible self weight carrying a distributed load. If the load is
a uniformly distributed vertical load, it is referred to as a parabolic
cable (Section 7.8). For this,
“4 wx?
Minimum tension = 7, = bs (7.22)
y
Tension at any point = T = VT2Z, + (wx)* (7.23)
If the ends of the cable are at the same elevation, then

wh?
°C 8H
wL TN
ecm led ee
2 4H

Total cable length = S = L} 1 + 3\7


8(H\
whereL = horizontal distance between the end supports
H = maximum sag
398 Chapter 7 Distributed Load—Analysis of Beams and Cables

8. The procedure to determine the tension at various points of a flexible


cable subjected to its own weight (Section 7.9). For these cables,
referred to as catenary cables,

Minimum tension = Ty = wp (7.38)


Tension at any point = T = wy (7.45)

The coordinates of a point on the cables are

sae bs)
x = psinh = (7.40)
Pp
85
y = pcosh (
*) (7.43)
Pp

KEY TERMS
SSS SSS a

beam 370 parabolic cable 390


bending moment 372 shear force 372
catenary cable 396 statically determinate beam 37/
concentrated load 373 statically indeterminate beam 37/
flexible cable 370 uniformly distributed load 373
linearly varying distributed load 373

PROBLEMS
a a = a SS eS SSCSCSOSOSO™O™—OSSSC™

SECTION 7.5
For Problems 7.1 through 7.28, determine the shear force and bending moment at the
sections indicated.

7.1 (a) Immediately to the right of the 5-kip point load, and (b) immediately to the
left of the 5-kip point load.

7.2 At the middle of the beam.

7.3 Ata distance L/3 from the left support.

5 kip 8 kip

/-—4 ft —+|«— 4 ft —+}«— 4 ft—+


Figure P7.1 and P7.2 Figure P7.3
Problems 399

7.4 Ata distance L/6 from the right support.

7.5 At the center of the beam.

7.6 (a) Immediately to the left of the concentrated couple, and (b) immediately to
the right of the couple.

7.7 At the middle of the beam.

Mi wi 2 ta ea a
Figure P7.4 and P7.5 Figure P7.6 and P7.7

7.8 Ata distance L/4 from the right end of the beam.

7.9 At the middle of the beam (that is, at a distance of 3L/4 from the left support).

7.10 At the middle of the beam.

7.11 At the middle of the beam.

7.12 Ata distance of 4 ft from the left support.

7.13 Ata distance of 1 ft from the right end of the beam.

7.14 At the middle of the beam.

ph
Figure P7.8 and P7.9 Figure P7.10

2000 Ib aes

1000 lb/ft

8 kip:ft
pase tit= 2 ft
Figure P7.11 Figure P7.12, P7.13, ai\d P7.14
400 Chapter 7 Distributed Load—Analysis of Beams and Cables

7.15 Immediately to the left of the right support.

7.16 At the middle of the beam.

7.17 Ata distance of 0.5 m from the fixed end.

7.18 Ata distance of 2 m from the fixed end.

7.19 Ata distance of 5 ft from the left end of the beam.

7.20 Ata distance of 2 ft from the right end of the beam.

3kN 500 lb
1000 lb/ft

4 ot L__j,_L_,| 1.5 mt 5m 10
Figure P7.15 and P7.16 Figure P7.17 and P7.18 Figure P7.19 and P7.20

7.21 Ata distance of L/3 from the fixed end.

7.22 Ata distance of L/3 from the fixed end.

7.23 Ata distance of 4 ft from the fixed end.

Wo
7
A B
kK L >| k L - “ 12 ft ~
Figure P7.21 Figure P7.22 Figure P7.23

7.24 At the middle of the beam.

7.25 At the middle of the beam.

7.26 At the middle of the beam.

Figure P7.24 Figure P7.25 Figure P7.26


Problems 401

7.27 Ata distance of 6 ft from the left end of the beam immediately to the right
of C.

7.28 Ata distance of L/3 from the fixed end.

300 lb/ft 400 lb/ft

san A B ~Zoo1-ft mht


ett
2 ft apy +
L3 “ar
L3 :3
aap" “3

Figure P7.27 Figure P7.28

7.29 through 7.47 Draw the shear-force and bending-moment diagrams for the
beams shown.

5 kip 8 kip

as | | =e

e 4 ft—+}+— 4 ft —+}.— 4 ft Me]


Figure P7.29 Figure P7.30

p— 4p bof
Figure P7.31 Figure P7.32

pe ae 3
ees
Figure P7.33 Figure P7.34
402 Chapter 7 Distributed Load—Analysis of Beams and Cables

2000 Ib
3 kip

1000 Ib/ft

D
A Ge A 5)
B /\ B AC
8 kip:ft
1+—4 ft—+l. 8 ft ——+| Le—=2 ft —»|+ — 9 ft_--—2
ft
Figure P7.35 Figure P7.36

3kN 500 lb
w 1000 lb/ft
| | | | |, 0.9 kN/m
A
L\B _ SC B GC A B é
L
kK 4+}. £ a £ | - 1.5 m—+}—1.5 m—> +10 ft ————+}+— 4 ft

Figure P7.37 Figure P7.38 Figure P7.39

Wo

A B
ee }+¥——___—.]2 = west

Figure P7.40 Figure P7.41 Figure P7.42

WD)

A B A B
it C A
meeB r Gi
—— : ee iD eS L D L
ee ee eae 3 = > =<
Figure P7.43 Figure P7.44 Figure P7.45

300 lb/ft S00 is


E
54 B ~€001b-ft maim
pees re na a
a
Figure P7.46 Figure P7.47
Problems 403

7.48 A hook is welded to the beam AC at B. The hook carries a load of 2000 lb.
Draw the shear-force and bending-moment diagrams for the beam.

7.49 A person weighing 140 Ib is standing at the edge of a diving board AB. Draw
the shear-force and bending-moment diagrams for the diving board.

E 10ft seo ft>|

‘K =e
Swimming pool
2000 lb
Figure P7.48 Figure P7.49

7.50 Draw the shear-force and bending-moment diagrams for the beam ACB.

7.51 A curved bar ABC is subjected to a horizontal force P. Express the magnitude
of the shear force V and the bending moment M as a function of P, R, and @ at
Section B.

7.52 Draw the shear-force and bending-moment diagrams for the two beams AB
and BC.

na
A
C P P

M TV
WwW

B C
ne Q A B
Be is
ea L
ee Ub,
ae
Ie

Figure P7.50 Figure P7.51 Figure P7.52


404 Chapter 7 Distributed Load—Analysis of Beams and Cables

7.53 A boxcar rolls on a beam. The weight of the boxcar is W. Determine the
distance x at which the maximum moment occurs. Also determine the magnitude of
the maximum bending moment.

7.54 ABC is a beam with a distributed load w per unit length. Determine the
relationship between L, and L, for which the magnitude of the largest bending
moment will be at B.

,
oe
a
lk i |
Figure P7.53 Figure P7.54

SECTION 7.6
7.55 through 7.69 Draw the shear-force and bending-moment diagrams for the
beam shown.

1000 lb/ft

Figure P7.55 Figure P7.56 Figure P7.57

Wo

A
B +

iL4 _
-}—4—-—
-— 2 m—>}+— 2 m—>e— 2 m—>
Figure P7.58 Figure P7.59 Figure P7.60
Problems 405

M,
A B
Cc 4

L1+ |
L Se Sas
Figure P7.61 Figure P7.62 Figure P7.63

2 kip/ft Cable
10 kip:ft Wy Cable
500 1b 750 1b
A
Lees B G
8 ft >{<—§| L 2L A B G D Te,
a 2 ft aoe zs be-5 ftehe5ft-mp-S ft>p<4 fty
Figure P7.64 Figure P7.65 Figure P7.66

1000 lb/ft 1000 Ib/ft 1200 lb/ft 1200 Ib/ft


Wo /2 Wo

Es
A D D
B G E 1500 lb-ft :
_ Coe L L ib }+——7 ft—>}~+>}+-+}——6 ft —>
ee | sas oe 1.5 ft 1.5 ft
Figure P7.67 Figure P7.68 Figure P7.69

7.70 through 7.74 Shown are shear-force diagrams for simply supported beams.
Draw the load and the bending-moment diagrams. There are no couples acting on the
beam.

+30 kN
+5200 Ib

A A B G

—4800 lb
—30kN
—3 ft 3 ft+}+—4
ft—> es 5 |
Figure P7.70 Figure P7.71
406 Chapter 7 Distributed Load—Analysis of Beams and Cables

+21.93 kN

+12.93 kN

CD E
A B 1.07 kN

‘ —10.07 kN

Figure P7.72 Figure P7.73

}+—4 ft —>}«—4 ft—>+}+—4 ft >


Figure P7.74

SECTION 7.7
7.75 through 7.77 Find the tension in the cable segments AB, BC, and CD.

12001b 1800 Ib
|+—— 6 ft —+}+—4 ft +}«—_5 ft -—2 m—+}+—2 m>}e— 2 m—+| 3 ft>}-— 4 ft >1«—— 6 ft ——
Figure P7.75 Figure P7.76 Figure P7.77

7.78 ABCD is a cable. What should be the height H so that the segment BC will
remain horizontal?

7.79 ABCD is a cable. A horizontal force F holds the cable in equilibrium.


Determine the magnitude of the force.

7.80 ABCD is a cable subjected to two forces shown. Determine the distance y.
Problems 407

120 1b 200 Ib
—— On em | sin ay§ lke eae
1.5m hae Wai Ai Maite a
Figure P7.78 Figure P7.79 Figure P7.80

SECTION 7.8
7.81 The center span of a suspension bridge is 1000 ft long. The ends of the two
cables supporting the center span are at the same elevation. Each cable supports
a uniformly distributed vertical load of 3000 lb/ft. The sag of the cable at the center
of the span is 120 ft. Determine the maximum and minimum tensions in the cable.

7.82 Refer to Problem 7.81. Determine the length of each cable.

7.83 A 130-ft length of a pipeline is supported by a cable. The pipeline weighs


830 lb/ft. Determine the maximum and minimum tensions in the cable. Assume no
support from soil.

130 ete S|
A

Figure P7.83 and P7.84 A

7.84 Determine the length of the cable described in Problem 7.83. 20 m

7.85 The cable AB supports a load of 1200 N. The cable is horizontal at B.


Determine the magnitude of the average value of tension in the cable [i.e., B
(To + Tmax)/2] and the point where it occurs.

7.86 A cable having a length of 45 m was used to span a horizontal distance of 1200 N
40 m. The elevation of both ends of the cable was the same. Determine the maximum ae 0
sag of the cable. Assume this to be a parabolic cable. Figure P7.85
408 Chapter 7 Distributed Loada—Analysis of Beams and Cables

7.87 The cable AB shown supports a distributed load of w = w, cos (1x/L). For
the shape of the cable, show that

w,L? TX
= COS
y WT, in
7.88 The cable AB is horizontal at B. It is subjected to a nonuniform distributed
vertical load w = 100x Ib. Express the deflection curve y as a function of x, H,
and L.

7.89 For the cable in Problem 7.88, determine the maximum tension T.,.max as a
function of L and H.

iy OSES w = 100x Ib
Figure P7.87 Figure P7.88 and P7.89

7.90 For the cable, show that

dy _ w(s)
dx O
where T, = tension at B.

w(x)

Figure P7.90
Problems 409

7.91 AB and BC are two parabolic cables. If the resultant force at B is zero,
determine the maximum sag H of the cable BC. Assume w to be the same for both
cables.

7.92 ABC is a cable that passes over a frictionless pulley at B. Neglecting the
weight of the cable, determine the weight W.

7.93 The cable AB carries a uniform vertical load of 3 KN/m. Determine the
maximum tension in the cable.

7.94 For the cable described in Problem 7.93, derive an equation for the deflection
curve.

i 200 ft >+-—100 ft—


Ty rs < G
20 ft
Ke
@:AV
SEV
ANY
AX

ee ii

Figure P7.91

+ SOft a
Se 20
A B__F@ rictionless y r a Z4
pulley
10°
Lic B . bua
W
100 lb/ft 3 kN/m
Figure P7.92 Figure P7.93 and P7.94

SECTION 7.9
7.95 The cable AB weighs 40 lb per foot length. Determine the maximum and
minimum tensions in the cable.

pe — 300 ft >|
B A
Za
46.25 ft

Figure P7.95
410 Chapter 7 Distributed Load—Analysis of Beams and Cables

7.96 The cable is 180 ft long and weighs 10 Ib per foot length. Given the maximum
sag H = 20 ft, determine the span length L.

7.97 The 50-m-long cable weighs 200 N per meter length. The maximum tension in
the cable is 12 KN. Determine the span length L.

7.98 The cable ACB has a weight of w per unit length. Show that

I,
y= Zocosn(=)= 1
w To
where T, = tension at C.
y

: L = B A
B 7 A

H
x
C
Figure P7.96, P7.97, and P7.101 Figure P7.98

7.99 An electric tramway cable weights 4.2 Ib/ft. Its ends are supported at the same
elevation 650 ft apart. If its maximum sag is 122 ft, determine (a) the length of the
cable, and (b) the maximum tension.

7.100 A cable is 70 m in length. It is suspended between two points 65 m apart at


the same elevation. The maximum tension in the cable is 350 KN. Calculate (a)
the maximum sag of the cable, and (b) the weight of the cable per meter length.
7.101 For the cable shown, if the total weight is equal to the maximum tension,
determine the ratio L/H as a function of S;/H where S; is the total length of
the cable.

7.102 The cable shown weights 3 lb per foot. Determine (a) the deflection curve
equation for the cable, (b) the height H, and (c) the maximum tension in the cable.
7.103 A balloon has an upward buoyant force of 200 lb. The cable AB is 200 ft
xq
long and weighs 0.8 Ib/ft. Determine the height H of the balloon.

~Frictionless
pulley
180 Ib
Figure P7.102

Figure P7.103
CHAPTER

FRICTION

8.1 Introduction OUTLINE


8.2 Mechanics of Dry Friction
8.3. Analysis of Some Dry Friction Problems
8.4 Wedges
8.5 Square-Threaded Screws
8.6 Belt Friction
*8.7 Frictional Resistance on Thrust Bearings—Disc Friction
*8.8 Journal Bearings
*8.9 Rolling Resistance
8.10 Summary
Key Terms
Problems

411
412 Chapter 8 Friction

8.1 INTRODUCTION
In the preceding chapters we have assumed that the surfaces of contact between
two bodies were perfectly smooth. As a result of this assumption, the forces
of interaction always acted normal to the surfaces at points of contact. How-
ever, in all practical problems the surfaces are rough to some degree and there
may exist small or large frictional force components. This frictional force acts
tangentially to the surfaces of contact, and it opposes the force(s) tending to
cause the relative motion. Friction is an important subject for study. In some
applications we wish to minimize its effects, but in others, these effects are
essential. Without friction forces cars, trains, and bicycles would not work. It
would not be possible to walk without friction.
In general, friction can be divided into two major categories:

1. Dry friction. This is the friction encountered along the unlubricated


surfaces of contact of two solids under a condition of sliding or having a
tendency to slide. Dry friction is also referred to as Coulomb friction,
since its fundamentals were investigated by Coulomb in 1781. .
2. Fluid friction. This type of friction is developed in fluids—gas or
liquid—when adjacent layers move at different velocities. The differential
movement gives rise to friction forces between the fluid elements.

In this chapter, we will discuss the various aspects of dry friction only.
Fluid friction is covered in the study of fluid mechanics.

8.2 MECHANICS OF DRY FRICTION


In order to understand the mechanics of dry friction, consider a box having a
weight W resting on a horizontal surface (Figure 8.1a). Since the weight W
will act vertically in a downward direction, the reaction N at the surface will
be equal in magnitude to W and will act in a vertically upward direction as
shown in the free-body diagram of the box (Figure 8.1). If a horizontal force
P is applied to the box as shown in Figure 8.1c and the magnitude of P is
small, the box will not move. Therefore, there must be a horizontal force F
which is equal in magnitude and opposite in sense to balance the applied force
P. This force F is the static-friction force as shown in Figure 8.1d.
It is also important to note that the position of the force N, which is the
resultant of a distributed force, has changed in Figure 8.1d as compared to
Figure 8.1b. The reason for this change in the position of N is due to the fact
that, when the force P is applied, the pressure on the surface of contact at the
right side of the box increases, accompanied by a decrease of pressure on the
left side of the box. The line of action of N is now located in such a manner
that 2M, = 0, or
Nx = FH

This prevents overturning of the box. However, if x is equal to or greater than


half the width of the box, then tipping will occur.
8.2 Mechanics of Dry Friction 413

=<+—~9 =<—9S

(c) (d)
Figure 8.1

As the force P is gradually increased, the magnitude of F will also increase,


accompanied by an increase in the magnitude of x which defines the location Motion-impending
of the normal reaction resultant N. However, there will be a limiting value of condition
P = Pirax = Fmax at which the box will start to slide. Once the box starts
sliding, the magnitude of F drops from F,,,, to F, and the box continues
moving. Figure 8.2 shows the phenomenon described above, which is an ideal-
ization of the real-life situation. Based on Figure 8.2, we can make the follow-
|
ing statements.
force,
Frictional
F
|
Range of |
Range of
kinetic
1. The region of P = P,,,,, is the range of static friction (also see Figure static | friction
8.3a). In this region friction !

P=F and |
tae 8 aoe Applied force,P
Figure 8.2
2. The instant at which P becomes equal to P,,,,, may be called the point
of impending motion (also see Figure 8.35).
3. After the impending-motion state is reached, the magnitude of F
decreases from F,,,,, to F, and motion continues. This is the range of
414 Chapter 8 Friction

kinetic friction. F, may thus be called the kinetic-friction force (also see
Figure 8.3c).
Experiments have shown that F,,, x is almost proportional to the normal
force N, or
Ww
———— EF
Fnax = bs (8.1)
\

Op, where w, is a constant called the coefficient of static friction. The phrase
almost proportional is used above since F,,,,, 18 slightly affected by the area
(Note: F<F,,,,
36 <@,; no motion) of contact and other properties.
Similarly, we can write that
(a)
F, = pN (8.2)
where 1, is the coefficient of kinetic friction.
The magnitudes of , and y, depend on the materials in contact and also

Td Ww
on the
values
roughness of the particular surfaces under consideration. Approximate
of , and y, between various materials are given in Table 8.1. Note
that these coefficients of friction are dimensionless quantities, and can thus be
AS Frnax
used with both SI units and U.S. Customary units. Care should be taken in
using these values since they are only approximate.
N ds he

(Note: Motion impending to right) Table 8.1 Approximate values of w, and pi,
for various materials in contact

(b) Materials in contact B, By


Steel on steel 0.75 0.55
Cast iron on cast iron 1.05 0.81
Cast iron on wood 0.4—-0.6 0.33-0.45
Rubber on wood 0.4 0.3
Wood on wood 0.25-0.5 0.18—0.38
Wood on leather 0.25-—0.5 0.17-0.4

Angle of Friction
We will again refer to Figure 8.3. From Figure 8.3a we can see that the re-
sultant reaction
(Note: Motion to right)
R=N+F
(c) The magnitude of R is equal to VN? + F?. The resultant reaction R makes
Figure 8.3
an angle @ with the normal drawn to the surface of contact. This angle is
determined by
8.3. Analysis of Some Dry Friction Problems 415

The magnitude of ¢ increases with the increase of the magnitude of F (since


the magnitude of N is constant) and, at the point of impending motion (Figure
8.3b),

© =, = tan * (Fu = tan” “(j1))

Thus we can state the relationship between the coefficient of static friction and
the angle between R and the normal N as

L, = tan d, (8.3)

The angle ¢, is defined as the angle of static friction.


In a similar manner, when the body is in motion (Figure 8.3c)

p= 0, — in 1(21)Satan (Wp)

or

pH, = tan d, (8.4)

The angle ¢, is defined as the angle of kinetic friction.

8.3 ANALYSIS OF SOME DRY FRICTION PROBLEMS


All problems involving dry friction can be solved by drawing proper free-body
diagrams and paying attention to Equations (8.3) and (8.4). We will now use
some dry friction problems to illustrate that. Figure 8.4a shows a block having
a weight W placed on an inclined plane in a state of equilibrium. The angle
of inclination of the plane with the horizontal is equal to 8. Referring to the
free-body diagram shown in Figure 8.45, the normal reaction N = W cos 8
and the friction force F = W sin 0. Thus,

F_ Wsin®
t =>-—-— =>

ace N Wcos8

or & = 9. This means that the resultant reaction, which has a magnitude of
R = W, makes an angle ¢ with the normal to the surface of contact, and the
angle ¢ is equal to the angle of inclination 6 of the plane with the horizontal.
If the angle of inclination 8 of the plane is gradually increased, we will
reach a point when 6 will be equal to the angle of static friction , as shown
in Figure 8.4c. At this time
F = F,,,x = W sin 8 = W sin o,
N =W cos 0 = W cos o,

—ma
N — tan@
Janek
an = tan d. == ps
no,
416 Chapter 8 Friction

(Note: 6<@,; R= W;no motion)

(a) (b)

&
(Note: 06=¢,; R=W; motion impending) (Note: @ >¢,; downward motion)

(c) (d)
Figure 8,4

(Note: 0<¢,) This is the condition of impending motion, and the block is ready to slide down
the plane.
(a) If the inclination 6 of the plane is further increased (that is, 8 > o,) the
block will be in motion (Figure 8.4d). For this case

F=F,<W
sin 0
N=Wecos 0

For another example, consider a block of weight W resting on an inclined


plane. Let the inclination @ of the plane with respect to the horizontal be less
than ,. Let it be required to determine the magnitude of the force P (Figure
8.5a) that will bring the block to an impending-motion condition. At the point
of impending motion as shown in Figure 8.55, we consider equilibrium:
@®2F, =0 N =Wcos 0
Also
(b)
Figure 8.5 @ IF, =0 P — Wsin 0 = F = Fyn
8.3. Analysis of Some Dry Friction Problems 417

However

F P —Wsin0
= tan db, = * =
Bs ®s N W cos 6

If the magnitudes of &,, W, and 6 are known, the magnitude of P can be


calculated. It is important to note that the resultant reaction R makes an angle
&, with the normal to the plane. Hence the resultant makes an angle of
6 + &, with the vertical.
Figure 8.6a shows a block of weight W resting on an inclined plane. For (Note: 6<¢,)
this case, the inclination of the plane, 0, is less than ,. If a force P acts on
the block, the free-body diagram for the impending-motion condition will be (a)
as shown in Figure 8.65. For this case, considering equilibrium gives

@®XF, =0 N =Wcos 0
Also

@3F, =0 —P—Wsin@+F=0
or
P+W sin 0 =F = Fx
Therefore,

Foe FEW sind


et Daas
oir kaa @
Note that the resultant reaction R will make an angle of ¢, with the normal to Figure 8.6

the plane. However, R will make an angle of @, — 9 with the vertical (compare
this to Figure 8.5).
418 Chapter 8 Friction

EXAMPLE 8.1

A box weighing 250 Ib rests on a horizontal surface. An inclined


force P = 90 Ib acts on the box. Given ., = 0.35, determine
whether or not the box will remain in equilibrium.

Solution The free-body diagram for the box is shown. There


are three unknowns—F, N, and x—which can be determined
from the three equations of equilibrium for the box. For
equilibrium

@rF, =0
P cos 30° — F=0
F =P cos 30° = (90)(cos 30°) = 77.94 lb
@zspF, =0
—P sin 30°-W+N=0
N=W + P sin 30° = 250 + (90)(sin 30°) = 295 Ib
Now we can use Equation (8.1) to write

Foray = egN = (0.35)(295) = 103.25 Ib


Considering the third equilibrium equation,

@M, =0
—(P cos 30°)(1) + (P sin 30°)(1.5) + (N)(x) = 0
_ Pcos 30° — 1.5P sin 30°
0 oe ua TS @ i ya a oe
he _ (90)(cos 30°) — (1.5)(90)(sin 30°)
" 295
F 71.94 — 67.5
+1 =e sarees 0.0888
B=
5)=
55, 'N
(Noe eho uppee nF Since x < B/2 (that is, 1.5 ft) the box will not tip. For tipping,
are)
the free-body diagram will be as shown in the figure. Also,
since F < F,,,,, there will be no sliding. Thus, the box will be
in a State of equilibrium.
8.3 Analysis of Some Dry Friction Problems 419

EXAMPLE 8.2

A block has a weight of 0.75 KN and is resting on an inclined


plane as shown. Determine the direction and the smallest value
of the force P that will just start moving the block up the plane.
Let b, = 34°.

Solution The free-body diagram for the block for the


impending-motion condition is shown. The resultant reaction R
will make an angle of , with the normal drawn to the plane.
In Chapters 2 and 3 it has been shown that force triangles
can be conveniently used to solve equilibrium problems. This is
also true in problems relating to friction. The corresponding
force triangle for this problem is shown, from which it can be
seen that since the direction of R is fixed, the smallest value of W=0.75 kN
P is obtained when the angle between P and R is 90°. Or
P=W sin(0 + ,) = (0.75)sin(20° + 34°)
P=0.607kN <

To determine the direction of P, let B be the angle of P with


respect to the inclined plane. Then

B = 54° — 20° p=34<

This angle
must be 90°
for P to be
minimum

6 +9,= 20° + 34°=54°


420 Chapter 8 Friction

EXAMPLE 8.3

Frictionless pulley A block having a weight of W is placed on an inclined plane. The


block is attached to a cable that passes over a frictionless pulley.
The other end of the cable is attached to a weight W,. Determine
the range of values of W, in terms of W for which the block will
remain in equilibrium. Let 1, = 0.25.

Solution

Maximum value of W,. The maximum value of W, will occur


when upward motion of W is impending. The free-body
diagram for that condition is shown. For equilibrium

@=F, =0 W sind —-W, + p,N =0 (a)


@2XF, =0 N —Wcos6=0
Witaan) N = Weos 45° = 0.707W (b)
From Equations (a) and (b),
F nax =HsN 7
\ W, = W sin 0 + p,N = W sin 45° + (0.25)(0.707W) = 0.884W
Witaan) N
Minimum value of W,. The minimum value of W, will occur
when downward motion of W is impending. The free-body
diagram for that condition is shown.

® sr, =0 W sin 0 — p,N — W, =0 ()


@=F,=0 N—Wcos6=0
So

N = Wcos 0 = W
cos 45° = 0.707W (d)

From Equations (c) and (d),

W, =Wsin9—-p.N
= W sin 45° — (0.25)(0.707W) = 0.53W_ (e)
Thus, the range for equilibrium is
0.53W=W, =0.884W <
8.3 Analysis of Some Dry Friction Problems 421

EXAMPLE 8.4

A uniform rod of weight W is supported at A and B. Point A is


A
the top edge of the wall. Assuming that the coefficient of static
friction 1, at A and B is the same, determine the maximum value
of a for which no slippage will occur.

Solution The free-body diagram for the rod is shown. L

F,= pa and Fz = wp
For equilibrium
@sF,=0 N, sina
— F, cos a — Fz = 0 (a)
@=r, =0 _
—-W+N,+N, cosa + F, sina =0 (b)
@53M, =0
-n,(§) + na(écos «|= ro(
Ein«|=0 (c)

From Equation (a) we can write

N, Sin a — p, N, cos a — p,Nz = 0


N,(sin & — p, Cos a) = p,Np

Ne
Nz = — (sin a — pcos @) (d)

From Equation (c) we can write

ib, L Le
= IN, 3 =P Ns Be = [DIN poe =0

—N,
+ Np cos
a — p,Nz sina
=0
—N, + N,(cos a — p, sin a) = 0 (e)

Combining Equations (d) and (e) gives

Nie :
—N, + —(sin a — p, Cosa)(cos a — p, sin a) = 0
S
or

—p, + (sin a — pt, Cos a)(cos a — p, sin a) = 0


(1 + p2)sin a cos a = p, + w,(sin2a + cos*a)

Using double-angle formula,

sin 2a =
422 Chapter 8 Friction

EXAMPLE 8.5
A cylinder w:ighs 120 lb and has a diameter of 20 in. It is pulled
P
to the left by a horizontal force P, which is applied by a rope tied
to the cylinder. Determine the height of the step H for which the
cylinder can be pulled up the step and, at the same time, there
will be impending slippage at A. Let p, = 0.25.

Solution The free-body diagram is shown. Note that, at point


B, there will be no friction force since the cylinder will just
be lifted. Also, since the force P tends to give a counterclockwise
rotational tendency to the cylinder, the frictional force at A
must oppose the motion. For equilibrium,

© Mp =0.
P(10) + N(O) + (120)(0) — p,N(10) = 0
Therefore,
P= wp,N = 0.25N
@=sF, =0
(Note: F=u,N) —-P+Nsina—p,N
cos a=0
—0.25N + N sin a — 0.25N cos a = 0
sina = 0.25 + 0.25 cos a
or
(0.25 + 0.25 cos a)? = sin? a = 1 — cos?a
0.0625 + 0.0625 cos? a + 0.125 cos a = 1 — cos*a
1.0625 cos? a + 0.125 cos a — 0.9375 = 0
cos*a + 0.1176 cos a — 0.882 = 0
Solving the above equation, we find that a = 28°. Thus,
H = 10 — 10 cos a = 10 — (10)(cos 28°)
H=1.17in. <
EXAMPLE 8.6

Three blocks A, B, and C are shown. DE is a cable attached to


the wall at D and to block A at E. Determine the maximum value
of the force P before any slippage takes place. Let W, = 150 lb,
W, = 250 lb, Wc = 300 Ib.

Solution There are two possible alternatives for slippage, and


they are:

1. Block B will slip and block C will stay in place.


2. Blocks B and C will slip together.

Alternative 1. The free-body diagrams for blocks A and B are


shown. From the free-body diagram for block A,
@=F, =0 F,-T=0
Fy = p)N,4 = 0.25N4
Therefore,
T = 0.25N, (a)
sr, =0 Ny = 150 Ib (b)
From Equations (a) and (b) we compute

T = (0.25)(150) = 37.5 Ib

From the free-body diagram for block B we find, for equilibrium,

©sr, =0 N50Ne 0
We know that V, = 150 lb. Thus

Nz = 4001b
@)=F, =0
P=F, + Fp, =0.25N, + 0.30N, (c)
= (0.25)(150) + (0.30)(400) = 37.5 + 120 = 157.5 N

Alternative 2. The free-body diagram for blocks B and C


together is shown.
@=sF,=0 N,-250-300-N,=0
Ne = 250 + 300 + 150 = 700 N
@ 2F,=0
P—F,—F.=0
Fy = WN, = 0.25N4 = (0.25)(150) = 37.5 lb
Fo = pagNc = (0.25)(700) = 175 N Since, in Alternative 1, P = 157.5 N, which is less than that in
Alternative 2 where P = 212.5 N, the maximum force P
Therefore, before any slippage occurs is
P=F,
+ Fo = 37.5 + 175 = 212.5N ISTS5N <

423
424 Chapter 8 Friction

8.4 WEDGES
Wedges are simple machines often used for one of two purposes: to transform
an applied force into a larger force, or to raise or lower heavy pieces of ma-
chinery through a small distance. To demonstrate the principle of wedges, we
consider a heavy block A of weight W as shown in Figure 8.7a. Block B is a
wedge to which a force P is applied to lift block A. Let it be required to find
the minimum value of P. Figures 8.7b and 8.7c show the free-body diagrams

(d) (e)
Figure 8.7
8.4 Wedges 425

for the block A and the wedge B. In the development of the following analysis,
it is very important to realize that the relationship F = w,N gives the maximum
possible frictional force along any surface of contact. From Figure 8.7b, for
equilibrium,
@2F, =0 N, —N> sina — F, cosa = 0
We can write that F, = p,.N,; thus

N, —N, sina —p.N, cos a = 0 (8.5)


Again considering equilibrium,

er =0 —F, -W+N, cos a — F, sina =0


Substituting F; = p,,N, and Fz = p,.N, into the preceding equation, we
obtain
— pV, —W +N, cos a — pyN, sina = 0 (8.6)
From Equations (8.5) and (8.6), if the magnitudes of 1,;, bo, W, and a are
known, the magnitudes of N, and N, can be determined.
Now refer to Figure 8.7c. Assume that the weight of wedge B is negligible
when compared to that of block A.

xr = 0 N; + F,sin a — Nj cosa =0
t
soo
So we may write
N3 = Nz COS a — PN, sina (8.7)
Again,

QF, =0 F, —-P + F,cosa + N,sina = 0


t t
B33 BN
Thus we may write
P = p3N3 + N2 cos a + No sin a (8.8)
Since the magnitudes of N, and N, are already known, the magnitude of N;
can be calculated from Equation (8.7). Equation (8.8) can then be used to
obtain the required minimum magnitude of P.
The magnitude of force P can also be obtained by drawing force triangles.
For example, Figure 8.7d shows the force triangle for the block A, the free-
body diagram of which is shown in Figure 8.7b. Using the law of sines,

Nie APE ce ee a
sin(90+ ,,) sin(Q0 — $,,; — a — $,))
426 Chapte. 8 Friction

or
ell W sin(90 + o,,)
(8.9)
> sin(90 — $,, — a — b,9)
Again the force triangle for the wedge B is shown in Figure 8.7e. The free-
body diagram for the wedge was shown in Figure 8.7c.

P ~' R,
sin(a a 00 a ,3) Z sin(90 ic 6,3)

or
_ Ry sina + db. + $,3)
(8.10)
sin(90 — o,3)
Example 8.7 shows the use of the principle of wedges.

EXAMPLE 8.7

Determine the minimum force P required to move block A, which


weighs 10 KN. Assume p., = 0.3 for all surfaces.

Solution The free-body diagrams of block A and wedge B are


shown. Referring to the free-body diagram of block A,

Qre=0 | NS F,=0
N, = F; = wN3
= 0.3N3 (a)
Or, =0 N; —F,-W=0
We may also write

F, = p,N2 = p5N3 = (0.3)°(N3) = 0.09N,


N; — 0.09N, — 10 =0
10
N; = —— = 10.99 kN
Pai
Now, referring to the free-body diagram of the wedge B,
@=F,=0 N, cos 20° — F, cos70° — N = 0 (b)
We know that
F, = 0.3N, and N, = 0.3N3 = (0.3)(10.99) = 3.297 KN
So, from Equation (b),

0.940N, — (0.3)(0.342)N, — 3.297 = 0


BPA)
N, == = 3.937 kN
0.940 — (0.3)(0.342)
(continued)
8.4 Wedges 427

EXAMPLE 8.7 (concluded)

Again for wedge B,


@=F, =0
Seriya 1 Sil Oe
SeING Sin AN Oks) 0)
P =F, + F, sin 70° + N, sin 20° — 0.5
= 0.3N, + 0.3N, sin 70° + N, sin 20° — 0.5
= (0.3)(3.297) + (0.3)(3.937)(sin 70°)
+ (3.937)(sin 20°) — 0.5
= 0.989 + 1.109 + 1.347 — 0.5 = 2.945 kN
P=2.945KN <
Alternate Solution. Since w, = 0.3, from Equation (8.3) we
write
, = tan. (.) — tan. ‘(0:3)’ = 16.7°
Referring to the free-body diagram of block A, the force
triangle is shown. From this we can compute

R; W
sin 16.7° sin 56.6°
_ 10 sin OIE
= = 3.44 kN
2 sin 56.6°
Referring to the free-body diagram of the wedge B, the force
triangle is shown.
P+05 3.44
sin 53.4° sin 53.3°
sin 53.4°
P8444 GaKN
(==)
P=2.944kN <

asc0' = 6-73.35

90 —$, = 73.3°

902d 20 =538°
428 Chapter 8 Friction

8.5 SQUARE-THREADED SCREWS


Square-threaded screws are used in jacks, vises, and presses, for the purpose
of transmitting power from one part of the machine to another. Figure 8.8a
shows a jack with square threads. The purpose of the jack is to raise or lower
an object having a weight W by applying a force P to the handle. The force
P is applied at a radial distance d. In order to raise or lower the object, the
force P must be transformed into a tangential force S acting at an average
radial distance r as shown in Figure 8.8b. The moment due to the force P
should be equal to the moment due to the force S. Thus, Sr = Pd, or

S=— (8.11)

(a) (b) (c)

Figure 8.8
8.5 Square—Threaded Screws 429

A single thread traces a helix as it is wrapped around the shaft as shown in


Figure 8.8c. We see from this figure that the pitch p is the distance between
the same point on successive threads. In Figure 8.8d the thread of the base has
been ‘‘unwrapped’’ and is shown as a straight line. Note that a = lead angle,
and the lead L = p = vertical distance through which the screw advances in
one turn for a single-threaded screw.
For raising an object, the screw may be represented by a block of weight
W as shown in Figure 8.8d. Considering equilibrium for the impending-motion
condition,

@2F.=0 S—Nsina—Fcosa=0
With F = w,N, we can write

S = Msin a + p, cos a) (8.12)

In the y direction,

Orr 0 —W+Ncosa—F
sina =0
Again, with F = w,N,

N= ee ae (8.13)
cos a — b, Sina

Combining Equations (8.12) and (8.13) gives

_ W(sin a + pb, Cos a) (8.14)


(COS: it sill)

However, from Equation (8.11)

Hence, for raising the weight, the force needed at the end of the handle can
be given by

Wr| wp, + tana


Pare |ee 8.15
d Lae sie,

Figure 8.8e shows the free-body diagram of the block (which has been
substituted for the screw) for the downward impending-motion condition as-
suming the lead angle « to be less than &, (i.e., a < &,). This assumption is
essential. Otherwise, the load will be lowered even when § = 0. Now

or. =0 —-S+Fcosa—Nsina=0
430 Chapter 8 Friction

With F = p,N,
S = N(p, cos a — sin a) (8.16)

Qrr=0 —-W+Fsina+Ncosa=0
t
F=,p,N

W = N(p, Sin a + COs a)

or

N= a (8.17)
(wu, SIN a + COS a)

Combining Equations (8.16).and (8.17) gives


_ Wp, — tan a)
= (8.18)
(1 + p, tana)

So

Sr Wr(w, — tan a)
ae mer eR ET Ad 8.1
d di+u,
tan a) Ce

Equation (8.19) has been derived with the assumption that 6, > a, which
means that the screw will remain in place under the load when P = 0. This
is referred to as the self-locking condition.
Equations (8.15) and (8.19) are valid for a single-threaded screw, for which

L= p

t t
lead pitch

For a double, triple, or n-threaded screw,

L=np

From Figure 8.8d, it can be seen that


; iL
ANC, = ==
2ur

For an n-threaded screw

n
tan (8.20)
2ur

Substituting Equation (8.20) into Equations (8.15) and (8.19), for an


n-threaded screw,
8.6 Belt Friction

pa ee + np
(for raising the object) (8.21)
ad 2ur — ny,p

p= Wr|Mt]
2
7 earZam ==)
=e (for lowering the object) (8.22)

EXAMPLE 8.8

A board is clamped between the jaws of a C-clamp. The single Board


square-threaded screw has a mean diameter of 0.55 in. and a lead (.Miz 40tnelb
of 0.15 in. The C-clamp is tightened by a moment of 40 Ib - in.
Determine the clamping force on the board. Let pw, = 0.3.

Solution Refer to Equation (8.21)

M = Pd = 40 |b - in L=0.15
in. =p
0.55
r = — = 0.275 in. pw, = 0.3
2
n = 1 (since it is a single-threaded screw)

Now we can write the equation

Qtr, + |
Pd= wr
2ur — nw.p

(2)(a1)(0.275)(0.3) + (1)(0.15)
40 = wyo275)| 2]= 0.1092W
(2)(a)(0.275) — (1)(0.3)(0.1
W = clamping force = 366.3lb <

8.6 BELT FRICTION


Belts are used to turn pulleys and drums. The frictional force developed be-
tween the belts and the pulleys and drums has various engineering applications,
such as in belt drives and brakes. In this section we will derive relationships
for the tension in the two sides of a belt passing over a circular drum.
432 Chapter 8 Friction

Flat Belt

Figure 8.9a shows a flat belt passing over a drum. The free-body diagram of
a small segment CD of the belt at the impending-slip condition is shown in
Figure 8.9b. For equilibrium,

A6 A
QF, =0 —Tcos (=) + (TF + AT) cos (2°)—p, AN=0 (8.23)

Again

@®zF,=0 -Tsin (2) — (T + AT) sin (2) + AN=0 (8.24)


or

AN =T sin (=) + (T + AT) sin (=) (8.25)

Direction of
impending
slippage

Figure 8.9
8.6 Belt Friction 433

Substituting Equation (8.25) into Equation (8.23), we obtain

—T°Cos (=) + (Fe aryeos( 22)


D 2

= aC sin (=) al AT)sin (=) =0

(a8) -n(-2) (8)


or

Dividing the above equation by A6/2 gives

sin
Le |Oe tee |ae
AG ight) eal 2 (2°) 7

SA | SS

(32) | ie
2 AQ aT
COS |=a) ae | >] =

Hence

LDA es= 2,8


In (2) (8.26)

Finally, the relation between the tensions in the two sides of the flat belt can
be given as

T = T,et* (8.27)
If the belt is actually slipping, then the coefficient of static friction should
be replaced by the coefficient of kinetic friction; thus,

Ty = Tier (8.28)
434 Chapter 8 Friction

The total contact angle B in Equations (8.26), (8.27), and (8.28) is given in
radians. Also note that T, is always larger than T, in order to cause slippage
in the desired direction. It is important to keep in mind that Equation (8.27)
applies to the condition of impending slip and Equation (8.28) applies for the
case of slip.

V-Belt

In many cases of belt drives, V-belts (Figure 8.10a) are needed. In order to
derive a relationship between T, and T>, let us consider a small segment CD.
Referring to Figures 8.10b and 8.10c, it can also be shown that, for impending
slip,

BB
Tia Te (8.29)

Direction of
impending
slippage

T,

(a)

T+AT
Pe a ey
aor 2 AN sin GS)

(b) (c)
Figure 8.10
8.6 Belt Friction 435

EXAMPLE 8.9

A belt is passing over a fixed drum. If the magnitude of the force


F is 2.5 kN, determine the angle 6 at which the 4-kN load will
just begin to slip (the impending slip of the belt is in the clockwise
direction). Let up, = 0.25.

Solution We use Equation (8.26) to solve this problem. We


know that

T, =4kN T, =F =2.5 kN p, = 0.25

Substituting these values into Equation (8.26) gives

1 4
B= (5) In (+) = 1.88 radians = 107.7°

Therefore,

5 = 90 — 17.7 = 1/230 <<

EXAMPLE 8.10

To keep a boat from drifting, it must be moored by a rope


wrapped around a mooring bit. If the magnitude of the force F
from the ship is 4.2 KN, what should be the minimum magnitude
of T to prevent the rope from slipping? Let it be given that p, =
0.3, and that the rope has 2'4 turns around the mooring bit.

Solution We can solve this problem using Equation (8.27).


Given that T, = F = 4.2 kN, T, = T, p, = 0.3, and
B = (27)(2.25), we can write
4.2 = Te0-22m(2.25)

4.2
T =
69.488
——_ =
0.0604kKN
IS
T=604N
=
<
436 Chapter 8 Friction

*8.7 FRICTIONAL RESISTANCE ON THRUST BEARINGS—


DISC FRICTION
Thrust bearings such as end bearings (Figure 8.11a), collar bearings (Figure
8.115), and disc-type clutches are used to give thrust or axial support to rotating
shafts. Frictional moments are developed in thrust bearings. The frictional
resistance is developed due to the normal pressure exerted by one of the flat
circular areas on another. Friction between circular areas of this type is called
disc friction.
We will now develop an expression for the moment M required to produce
(a)
the impending rotation of a hollow shaft (inside radius = R,; outside
radius = R,) on an end bearing. Referring to Figure 8.12, on the surface of
contact consider an elemental area dA where
dA =rd®@dr (8.30)

Figure 8.11

Figure 8.12

The normal force dN on this elemental area can be given as

lie
(Wise (dA) (r d® dr) (8.31)
a(R2 — R?) ~ a(R? — R?)
where P is the normal force on the shaft with uniform normal pressure over
the area. Thus, the maximum friction resistance is found as

ie
= api apn 0a) (8.32)

The maximum moment of the frictional force on the elemental area about the
axis of the shaft is

(r* d® dr)
8.7 Frictional Resistance on Thrust Bearings—Disc Friction 437

The total maximum frictional moment can then be given as


6=2t 3 3

M = JdM = | (eae d) = P| =<]


dle,
=R; Sara
a R2— R?

or

2u.P |R2+R,R; + R?
sree ae Sig
If the shaft is solid, then R; = 0; and

2u5PR
= ace (8.34)

EXAMPLE 8.11

A flat disc is placed on a table. The disc has a radius of 10 in.


and is subjected to an axial force of 500 lb. Determine the mag-
nitude of the moment M that will cause the disc to slip. Given
p, = 0.28.

Solution As we are concerned with the slipping of the disc,


which is solid, we use Equation (8.34). We know that
p, = 0.28, P = 500 lb, and R, = 10 in. Thus we can simply
compute
(2)(0.28)(500)(10)
Vie
3
M = 933.3 lb: in. <
438 Chapter 8 Friction

*8.8 JOURNAL BEARINGS


Journal bearings are used to provide lateral support to rotating shafts (Figure
Journal
8.13a). Under most circumstances, journal bearings are lubricated, which
bearing
means that the frictional resistance between the shaft and the bearing will
primarily depend on (a) the viscosity of the lubricant, and (b) the speed of
rotation. However, in this section, we will determine the frictional resistance
that develops between the shaft and the bearing when the bearing is not lu-
bricated. Figure 8.13b shows the end view of the shaft and one of the bearings.
Let W be the load on the bearing transmitted by the shaft. (Note: This means
that the magnitude of the total weight of the shaft and the wheel is equal to
2W.) Also, let 2M be the couple required to keep the wheel rotating. Thus, a
couple M will be required at each bearing. The reaction of the bearing is equal
to R. The reaction will be directed vertically upward and will be equal in
magnitude to the weight W.. However, the point of application of R will be
such that the line of action will not pass through the center of the cross section
of the shaft. In fact, it will be located in such a way that the resisting moment
will be equal to — M. Thus the magnitude of M is
M=Ra

where a=r sin >,


r = radius of the shaft
, = angle of kinetic friction

Hence

M = Pr sin 6, (8.35)
Figure 8.13 For a small value of o,,

p, = tan b, ~ sin o,
Therefore

M = Rrp, (8.36)
EXAMPLE 8.12

A pulley loosely fits a shaft. The radius of the pulley is 150 mm,
and the radius of the shaft (7) is 20 mm. Assuming that the weight
of the pulley can be neglected and the coefficient of static friction
between the pulley and the shaft (w,) is 0.25, determine the mag-
nitude of the minimum force F, required to be applied to the belt
passing over the pulley to raise the weight W = 8 KN. (Note: The
belt is not slipping.)

Solution The free-body diagram of the pulley is shown. If


p., = 0.25 = tan d,, then d, = 14.04°.

r sin b, = (20)(sin 14.04°) = 4.85 mm

Taking the moment about B,

@=mM, =0
—W(150 + 4.85) + F,(150 — 4.85) = 0
_ W(154.85) _ (8)U54.85)
45 15 145.15

F W=8kN

EXAMPLE 8.13

Refer to Example 8.12. Determine the magnitude of the maxi-


mum force F, for which the weight of 8 KN will just begin to be
lowered.

Solution For the case under consideration, the free-body


diagram of the pulley is shown. As in Example 8.12,
r sin d, = 4.85 mm. Taking the moment about C,

@©=sm, =0
—W(150 — 4.85) + F,(150 + 4.85) = 0
_ W(145.15) _ (8)(145.15)
Re S085 154.85
F,=7.5kN <

439
440 Chapter 8 Friction

*8.9 ROLLING RESISTANCE


Figure 8.14a shows a wheel of weight W moving at a constant velocity on a
perfectly rigid surface. If there is no source of resistance (such as air), the
motion will continue indefinitely. Unfortunately, this is not truly the case. In
reality, a deformation in the supporting surface is introduced at the point of
contact of a rolling wheel as shown in Figure 8.14. This deformation results
in anormal pressure acting on the wheel. Let the resultant forces of the pressure
distributions in the front and the rear of wheel be N, and N,, respectively.
Figure 8.14c shows the resultant of N, and N, to be equal to N, or
N=N,+N,

(a)

(c) (d)
Figure 8.14
8.10 Summary 441

The magnitude of the horizontal component of N; is larger than the magnitude


of the horizontal component of N,. Therefore, a horizontal force P (Figure
8.14b) is required to keep the wheel moving. Figure 8.14d shows the free-
body diagram of the wheel. The forces involved are W, P, and N, and they
are concurrent. Note that the resultant N acts at A so that b = r sin 0.
In order to initiate and maintain the motion of the wheel at a constant
velocity, the moment of all forces about A must be equal to zero, or
Wb = Pr cos 8

For small values of 6, cos 8~ 1. Thus, Pr = Wb, or

pP=— (8.37)

The distance b is commonly referred to as the coefficient of rolling resistance.


However, the quantity b is not a dimensionless quantity; instead, it has a di-
mension of length and varies between 0.01 in. for a steel wheel rolling over a
steel surface to about 2 in. for a steel wheel rolling over soft ground.

EXAMPLE 8.14

A circular steel wheel having a radius of 200 mm rolls down an Constant velocity
ene
inclined plane with a constant velocity. Determine the coefficient
of rolling resistance.

Solution Referring to the figure, we see that the inclined


surface is tangent to the wheel at A. The normal reaction R will Ve
oe

act at B. The coefficient of rolling resistance is then equal to i —7z=


j 0)200 mm
b=r sin2° = 200 sin2° = 6.98
b=698mm <

8.10 SUMMARY
In this chapter, we have covered the following:

1. Dry friction is encountered between the unlubricated surfaces of contact


of two solids (Section 8.1).
2. The coefficient of static friction and the angle of static friction are
related as [Equations (8.1) and (8.3)]

max
=f =
Bs an o, N
442 Chapter 8 Friction

Similarly, the coefficient of kinetic friction and the angle of kinetic


friction are related as [Equations (8.2) and (8.4)]

py, = tan b, = -

The above relationships can be used to solve several types of dry


friction problems (Sections 8.2 and 8.3).
Wedges are simple machines that can be used to move heavy machinery
through small distances (Section 8.4).
For jacks with square n-threaded screws, the force F applied for raising
or lowering a weight W can be given as (Section 8.5)

Wr | 2 iF
For raising: foe ne [et] (8.21)
Ge \e2aiiaen Le
|
Wr | 2 =
Forlowering: P=— |—~be—” (8.22)
a> | 2ur-+ np

To determine impending slip, the tensions on two sides of a flat belt


passing over a drum can be given by the relation (Section 8.6)
Ty = T,enr (8.27)

Similarly, to determine impending slip for V-belts,

BB
Ty = Tyesin(a/2) (8.29)
Note that T, > T).
The moment M required to overcome the frictional resistance of a thrust
bearing is given by the relation (Section 8.7)

_ 2p,P E HOR Ry “|
M (8.33)
3 R, +R;
where P = normal force on the shaft
R,, R; = outside and inside radii of the shaft
Journal bearings provide lateral support to rotating shafts. The magnitude
of the moment M required at each bearing to keep the shaft turning is
(Section 8.8)

M = Rrp, (8.36)

The horizontal force P needed to keep a wheel rolling can be given as


(Section 8.9)

P= (8.37)
Problems 443

where b is the coefficient of rolling resistance, which has a dimension of


length.

KEY TERMS
a a a

angle of kinetic friction 4/5 impending motion 4/3


angle of static friction 4/5 impending-slip condition 432
coefficient of kinetic friction 4/4 journal bearing 438
coefficient of rolling resistance 44/ kinetic friction 4/4
coefficient of static friction 4/4 lead angle 429
Coulomb friction 4/2 pitch 429
disc friction 436 self-locking condition 430
dry friction 4/2 static friction 4/3
fluid friction 4/2 wedge 424

PROBLEMS
ES
s.rES ES SS at Po a

SECTION 8.3
8.1 A box is placed on an inclined plane. The weight of the box is 3.25 KN.
Determine the magnitude of the force P to just start the box up the plane.
Let p, = 0.3.

8.2 The box on the inclined plane weighs 250 lb. Determine the magnitude of the
force P needed (a) to just start the box up the plane, and (5) for which the box
will just be prevented from moving down the plane. Let wp, = 0.3.

8.3 The box A on the inclined plane is attached to a 100-lb weight by a cable
passing over a frictionless pulley. Given 1, = 0.25, determine the magnitude of
force P, which will just start the box up the plane.

Figure P8.1 Figure P8.2 Figure P8.3


444 Chapter 8 Friction

8.4 Determine the magnitude of the force P needed to start the 700-N block moving
to the left.

8.5 Blocks A and B weigh 150 lb and 200 Ib, respectively, and are resting on an
inclined plane. The coefficient of static friction 1, for all surfaces of contact is 0.25.
Determine the angle 8 for which block B will start sliding down.

8.6 Blocks A and B weigh 100 lb and 180 lb, respectively. Determine the magnitude
of the largest instantaneous force P for which block A will start sliding down. Let
a = 0°.

8.7. Solve Problem 8.6 with a = 10°.

Cable

u, =0.35
Figure P8.4 Figure P8.5 Figure P8.6 and P8.7

8.8 Blocks A, B, and C weigh 200 N, 400 N, and 600 N, respectively. Determine
the magnitude of the smallest force P required to move block C.

8.9 A person is pulling the 600-N block with a cable. Determine the tension in the
cable required to pull the block up the plane. Assume 1, = 0.28 for all surfaces
of contact.

8.10 A bucket of concrete weighs 50 kN. Determine the tension T in the cable
required to raise the bucket of concrete.
Frictionless

Figure P8.8

pulley

Figure P8.9 Figure P8.10


Problems 445

8.11 A ladder of length L is supported at the wall and the floor. Determine the
greatest angle a so that the ladder does not slip.

8.12 A cabinet is being pushed by a force P. If the force P is increased, determine


whether the body will first slide or tip, and the magnitude of the force at that instant.

8.13 A tank weighing 3.4 KN rests between a wall and an inclined plane. In order
to gain access to a port on the side of the tank, it needs to be rotated in a clockwise
direction. Determine the force P required for motion-impending condition. Assume
the wall to be frictionless.

b<— 2.5 ft —>|

W=1200 Ib

u.=0.32 Ht, = 0.42


Figure P8.11 Figure P8.12 Figure P8.13

8.14 A 25-ft-long ladder is supported by a frictionless wall at the top, and the floor.
The weight of the ladder is 65 Ib. A person weighing 125 Ib starts climbing up the
ladder. How high (#) can the person climb before the ladder starts slipping?

8.15 A cable is tied to blocks A and B as shown. Determine the minimum value of
the weight W of the block B to maintain equilibrium.

8.16 A cylinder is resting between a wall and the floor. The weight of the cylinder
is 600 N. It is being pulled by a rope. Determine the magnitude of the force P to
be applied to the rope for the motion-impending condition.

Frictionless W
pulley

er 2 oe =O
Figure P8.14 Figure P8.15 Figure P8.16
446 Chapter 8 Friction

8.17 Determine the minimum value of 1, at C for which the system will be in
equilibrium. Neglect the weight of the bars AB and BC.

8.18 A uniform rod is 20 ft long and weighs 120 lb. The magnitude of the force P
is 50 lb. Determine the minimum value of 1, at B for which static equilibrium is
possible.

8.19 For the uniform rod, given W = 200 lb. At B, 1, = 0.28. Determine the
minimum value of P for which static equilibrium is possible.

8.20 A cylinder having a radius of 0.5 m and weight of 3.2 KN rests between a wall
and a triangular wooden block having a weight of 10 KN. Given p, = 0.4 at A and
B, and », = 0.35 along CD, determine if the system is in equilibrium.

8.21 A door weighs 240 Ib. Assume p,, at A and B to be 0.26 and 0.36, respectively.
Figure P8.17
Determine the magnitude of the force P that will just cause the door to slide to the
left.

20-ft-long rod

.
O
0.5 m radius

Frictionless
surface é
Figure P8.18 and P8.19 Figure P8.20 Figure P8.21

Cable 8.22 The slotted cylinder having a weight W is resting on an inclined plane. Show
gh that slippage between the cylinder and the plane will occur when

8.23 Blocks A and B weigh 900 N and 1.5 kN, respectively. The link rod weighs
400 N. Calculate the magnitude of the force P that will start to move the system
to the right. For all surfaces of contact, assume p, = 0.3.

8.24 The cylinder has a weight W and a radius R. The static coefficient of friction
at A and B is equal to .,. Show that, for impending slippage of the cylinder, the
magnitude of the moment is

y= WR(p, RO + Hs)
p2
Figure P8.22 1+ ps;
Problems 44

; B
Figure P8.23 Figure P8.24

8.25 The cylinder has a weight W = 300 Ib and a radius R = 2 ft. The coefficient
of static friction at A and B is 0.3. Determine the magnitude of the moment M for
impending slippage.

Figure P8.25

8.26 A weightless cylinder of radius R is gripped between two plates that are
pinned at A. If the coefficient of friction at B and C is equal to 1, show that the
maximum angle a at which the cylinder can be gripped can be given as

a =2 tan“'p,

Figure P8.26
448 Chapter 8 Friction

8.27 AB and BC are two uniform rods, pin-connected at B. Determine the minimum
value of 2, at C so that the assembly will be in equilibrium. Note that the weight of
each rod is equal to W.

8.28 Uniform rods AB and BC weigh 150 N and 250 N, respectively. Let , at A
and C be equal to 0.45. Determine the greatest value of @ for which the system will
be in equilibrium.

Figure P8.27 Figure P8.28

8.29 A signpost has to be erected near a construction project. During erection,


a woman tries to hold the post AB in equilibrium by a cable attached at B. Determine
the angle 6 for which the post will just start slipping. Let L = 22 ft, H = 13 ft,
and p, = 0.25. Neglect the weight of the signboard.

8.30 AB and BC are two uniform rods, each of which weighs 25 Ib. A horizontal
force P is applied at the middle of rod AB. When the magnitude of P becomes equal
to 35 lb, slippage occurs at C. Determine p,, at C.

8.31 The uniform rod AB weighs 60 N. It is supported by a wall at A and a


frictionless roller at C. Determine the smallest angle 8 for which equilibrium will
exist. Let uw, at A = 0.6.

ae

Frictionless

Figure P8.29 Figure P8.30 Figure P8.31


Problems 449

8.32 A uniform rod AB is placed between two inclined planes. The length and
weight of the rod are L and W, respectively. For impending-slippage condition, show
that the normal reactions at A and B are as follows:

W 1 bs
DIR rts) |kCOS CuESINKCY

Ni = Sa
W 1 7
bes
2 a ane) |RCOS CcmmmS IIL

8.33 Refer to Problem 8.32. For impending-slippage condition, show that the
maximum value of 0 is
2
_, |Bs tan® a + ps
6 = tan ad
(Win = Walaa!

8.34 Three identical cylinders are shown. Each has a radius R and length L.
Assume that the value of «1, is the same at all points of contact and determine its
minimum value for equilibrium.

8.35 A cylinder of weight W and radius R is placed between two plates. Express the
angle a in terms of , and @ for the condition of impending downward movement.

(Length=L;
Weight = W) Ms

Figure P8.32 and P8.33 Figure P8.34 Figure P8.35

8.36 A uniform beam ABC is 5 ft long and weighs 35 Ib. It is pinned at A and rests
on a wheel at B. Determine the magnitude of the maximum force P that can be
applied in equilibrium to the cord wrapped around the axle. Only consider the
friction at B. Given: p, = 0.35.

Figure P8.36
456 Chapter 8 Friction

8.37 Two identical drums, each weighing 4.25 kN, are pushed up a slope by a
truck. The coefficient of friction, ,, at A is zero and at B, C, and D is 0.25.
Determine the normal force P from the truck and the lower drum which will just
start the motion. Note: The direction of the force P is parallel to the slope.

8.38 A man is pulling a uniform rod with a cable. The weight of the rod is 42 Ib.
When the rod just slips, determine the magnitude of jt, at A. For impending-motion
condition, let 9 = 40° anda = 25°.

8.39 Refer to Problem 8.21. Determine the magnitude of — P to move the door to
the right.

8.40 For the system shown, determine the greatest distance L for which the 210-Ib
block will not slide to the left.

Frictionless
Pulley
pulley
Cable

EOS)
Figure P8.37 Figure P8.38 Figure P8.40

8.41 The solid half-cylinder shown has a weight of W. It is being pulled by a cord
AB. For impending slippage at C, show that

@ = sin + iss
4 — 3m,

Cable

Figure P8.41
Problems 451

8.42 A uniform bar ABC rests on a disc with a radius of 3 ft. Given the weight of
the bar = 500 lb, the weight of the disc = 300 lb, p, at B and D = 0.3, determine
the minimum magnitude of the force F for which the cylinder will move.

8.43 A cylinder having a radius of 2 ft and a block rest on an inclined plane. The
weight of the cylinder is 1.4 kip. Determine the minimum value of the weight of the
block for which the system will be in equilibrium. For all surfaces, 1, = 0.56.

8.44 Blocks A and B rest on an inclined plane, making an angle @ with the
horizontal. The weights of the blocks are W, and W,, respectively. Note that 4) <
tan @ < 1.) Derive an expression for 9 for impending downward motion in terms
of W,, Wo, Way, aNd pg):
Figure P8.42
8.45 Refer to Problem 8.44. Let puc4) = 0.25, pg) = 0.42, W, = 500 N, and
W, = 900 N. Determine the maximum value of @ for which equilibrium will exist.

Figure P8.43 Figure P8.44 and P8.45

8.46 A disc weighing 75 lb is placed between a rod ABC and a rough surface. The
weight of the rod is 100 lb. The coefficient of static friction ., at B and D is the
same. Determine the minimum value of w, for which the cylinder will be in
equilibrium.
8.47 The beam AB is carrying a distributed load. Determine the magnitude of the
force F needed to pull the column supporting the beam.

Figure P8.46 Figure P8.47


452 Chapter 8 Friction

SECTION 8.4
8.48 Determine the force P required for bringing the block A to the vergeof
moving upward. Let W, = 300 lb, W, = 120 Ib, 8 = 30°, and p,, = py =
B3 = 0.3.

8.49 Refer to Problem 8.48. If the direction of the force P acting on block B is
reversed, determine its minimum value for which block B will be on the verge of
moving downward.

8.50 Repeat Problem 8.48 with the following: W, = 980 N, Wz = 250N, 0 =


20°, Psy = yo = 0.5, and p35 = 0.35.

8.51 Refer to Problem 8.50. If the direction of the force P acting on the block B is
reversed, determine its minimum value for which block A will be on the verge of
moving downward.

8.52 Repeat Problem 8.48 with the following: W, = 850 lb, W, = 180 lb, @ =
25°, wy, = 0.4, yp = 0.25, and p,3 = 0.2.
8.53 Refer to Problem 8.52. If the direction of the force P acting on block B is
reversed, determine its minimum value for which block A will be on the verge of
moving downward.

8.54 Determine the magnitude of the minimum force P needed to be applied to the
wedge A to just start moving block B. Neglect the weight of the wedge. Let W =
300 Ib, 8 = 10°, and p,, = py = By = 0.6.

8.55 Repeat Problem 8.54 with the following: W = 18.6 KN, 8 = 5°, p,) = bo =
0.4, and p,3 = 0.3.

8.56 A wedge is being driven into a log of wood. Determine the smallest value
of 2, in terms of the wedge angle a for which the wedge will not pop out after being
driven into the log by the hammer. Neglect the weight of the wedge.

Figure P8.48, P8.49, P8.50, P8.51, P8.52, and P8.53 ‘Figure P8.54 and P8.55 Figure P8.56
Problems 453

8.57 Determine the minimum value of the force P needed to move the block C.
Let W = 4000 lb, 8 = 10°, and ,,, at all surfaces of contact = 0.45. Neglect
the weight of the wedges.

8.58 Repeat Problem 8.57 with the following: W = 26 kN, @ = 5°, and p, = 0.6.

8.59 A cylinder of radius R = 0.3 m and weight = 3.5 KN is supported between


an inclined wall and a wedge. Determine the minimum value of the force P
needed to just start the cylinder moving upwards. Assume yp, = 0.5 for all surfaces
of contact, and neglect the weight of the wedge.

8.60 The weight of the cylinder A is 15 Ib. The cylinder has a radius of 1 ft.
Determine the maximum weight of the wedge B for which the system will be in
equilibrium.

Figure P8.57 and P8.58 Figure P8.59 Figure P8.60

8.61 AB is a beam. C is a wedge. Determine the minimum value of the force P


required to move the wedge to the right. Assume ., = 0.25 at all surfaces of
contact.
8.62 Ifthe direction of the force P is reversed in Problem 8.61, determine its
minimum value to move the wedge to the left.

200 lb 600 Ib

Figure P8.61 and P8.62


454 Chapter 8 Friction

SECTION 8.5
8.63 A C-clamp holds two pieces of board together with a clamping force of
410 lb. The clamp has single square thread with a pitch of 0.08 in. and a mean
diameter of 0.65 in. Determine the torque required to tighten the clamp further.
Let p, = 0.35.

8.64 Refer to Problem 8.63. All values remaining the same, determine the torque
required to loosen the clamp.

8.65 Redo Problem 8.63 assuming that the screw is double threaded.

8.66 Two rods (A and B) have right-handed square single threads at their ends. The
threads have a mean diameter of 0.7 in. and a pitch of 0.12 in. C is a coupling.
Determine the magnitude of the moment of the couple (M) required to be applied to
the coupling to rotate it. Let p, = 0.28.

8.67 Refer to Problem 8.66. Assuming that rod A has right-handed threads and rod
B has left-handed threads, determine the magnitude of the moment of the couple to
tighten the system.

Board

Coupling C
Rod A |) Rod B

800 Ib 800 Ib
Figure P8.63, P8.64, and P8.65 Figure P8.66 and P8.67

8.68 A force P = 500 N is applied to the single-threaded jack screw. Given that p,
between the screw and the base of the jack = 0.55, thread diameter = 52 mm, and
pitch = 8 mm, determine the weight W that can be lifted.

8.69 Refer to Problem 8.68. If the weight W is 4.8 KN, determine the force P
required to raise it.

8.70 A scissors jack is used to raise a car for changing a flat tire. The jack screw
has single square thread with a mean diameter of 15 mm and a pitch of 3.5 mm.
Determine the magnitude of the moment of the couple M required to raise the car
higher. Let p, = 0.35.

8.71 Refer to Problem 8.70. Determine the magnitude of the moment of the couple
required to lower the car.

8.72 B is a wedge operated by a single-threaded screw to raise the block A. Given


the weight of the block A = 2.2 KN, mean radius of the screw = 10 mm, pitch
of the screw = 5 mm, and yp, for the screw = 0.2, determine the magnitude of the
moment of the couple M that must be applied to the handle of the screw. Neglect
the weight of the wedge and the friction at the contact surface of the screw and
the wedge.
Problems 455

WwW
}.-a00mm-+>

Figure P8.68 and P8.69 Figure P8.70 and P8.71 Figure P8.72

SECTION 8.6
8.73 A and B are two blocks connected by a rope passing over a fixed drum. Given
W, = 70 lb, w,,; = 0.55, and ,, = 0.3, determine the weight of block B for
impending upward motion of block A.

8.74 Refer to Problem 8.73. Determine the weight of block B for impending
downward motion of the block A. Use w,, = 0.26 and p,. = 0.3.

8.75 What should be the minimum coefficient of friction between the rope and the
metal shaft so that the two blocks (A and B) will remain static?

8.76 A box weighing 800 N is tied to a rope that has been wrapped around a
horizontal pole. Determine the magnitude of the maximum tension T for which the
box will not be lifted up. Let uw, = 0.25.

A 800 N
Figure P8.73 and P8.74 Figure P8.75 Figure P8.76
456 Chapter 8 Friction

8.77 Each of the two fixed bars has a diameter of 2 in. The center-to-center spacing
of the bars is 12 in. Determine the magnitude of the horizontal force P needed to
lift the block, which weighs 100 Ib. Given: j1, between the rope and the bars = 0.3.

8.78 Ina handbrake, a flat belt ABC passes over a wheel. The wheel is subjected to
a counterclockwise moment M with a magnitude of 125 Ib - in. Given p,, between
the belt and the wheel = 0.25, determine the magnitude of the minimum force
P required to prevent the wheel from rotating.

8.79 A flat belt passes over the flywheel. The coefficient of static friction between
the belt and the flywheel is 0.32. If the flywheel is subjected to a counterclockwise
moment M = 60 N - m, what should be the minimum magnitude of the force P
to prevent it from rotating?

- 35 in. ——+| bo |
6in, 14in. 400mm 500mm
Figure P8.78 Figure P8.79

8.80 A and B are two fixed drums, and C is a pulley that can rotate freely. A
cylinder of weight W has to be lifted by a rope. Given , (between the rope and the
drums A and B) = 0.35, maximum allowable tension in the rope = 150 Ib, and 0 =
60°, determine the largest weight W that can be lifted.

8.81 Refer to Problem 8.80. Let 1, = 0.5, 8 = 45°, and W = 800 N. Determine
the maximum tension in the rope for equilibrium.

WFj

Figure P8.80 and P8.81


Problems 457

8.82 A flat belt is laid over a table and then it passes over a fixed drum. Determine
the minimum magnitude of weight W of the box that can be placed over the belt that
will prevent the belt from slipping over the drum. Use p., = 0.4 for all surfaces of
contact.

8.83 A flat belt passes over two pulleys. A motor can apply a couple Mz, with
a maximum magnitude of 200 Ib - in. to the pulley B without causing slippage over
either pulley. Determine the tension in the belt on both sides of pulley A.
Let p, = 0.3.

8.84 A is acylinder with a weight of 2.6 KN and a diameter of 0.4 m. A steel rod
ab is rigidly attached to the cylinder. A flat belt bcd passes over a fixed drum B.
What is the maximum weight W that can be supported at d so that equilibrium of the
system as shown will not be disturbed? Let 1, = 0.35 on all surfaces of contact.

Figure P8.82 Figure P8.83 Figure P8.84

SECTION 8.7
8.85 The end bearing of a solid shaft having a diameter of 3 in. is shown. The axial
force on the shaft is 8000 lb and we will assume that the pressure distribution on
the bearing is uniform. Given p, = 0.32, determine the magnitude of the minimum
moment M on the shaft necessary to keep it rotating at a constant speed.

8.86 Redo Problem 8.85 assuming that the shaft is hollow, and the inside diameter
of the shaft is 1.2 in.

3-in. diameter (|P= 8000 Ib

Figure P8.85 and P8.86


458 Chapter 8 Friction

8.87 Determine the magnitude of the moment M necessary to start the shaft
rotating. Let p, = 0.4.

8.88 Determine the magnitude of the minimum moment M necessary to start the
shaft rotating. Given: static coefficient of friction = «,, normal pressure on the
surface of contact is uniformly distributed.

Figure P8.87 Figure P8.88 and P8.93

8.89 Repeat Problem 8.88 for the shaft shown.

8.90 Determine the magnitude of the minimum moment M necessary to start the
shaft rotating. Assume that the normal pressure varies linearly from zero at x = R,
to a maximum at x = R).

8.91 Repeat Problem 8.90 assuming that the normal pressure varies as Cx where
C = constant.

Figure P8.89 and P8.94 Figure P8.90, P8.91, and P8.92


Problems 459

8.92 Repeat Problem 8.91 assuming that the normal pressure varies as C/x? where
C = constant.

8.93 Solve Problem 8.88 with the following values: R = 4 in., 8 = 20°,
pw, = 0.45, and P = 300 lb.

8.94 Solve Problem 8.89 with the following values: R, = 4 in., R, = 1 in,
@ = 25°, uw, = 0.35, and P = 400 lb.

8.95 For the thrust bearing shown, determine the magnitude of the minimum
moment M required to turn the shaft against friction. Assume that the pressure on
the constant surface can be given as p = C sin 0, where C = a constant.

1?

Figure P8.95

SECTION 8.8
8.96 A pulley of radius R, fits loosely to a shaft of radius R,. Assuming no slip
between the rope and the pulley, show that the magnitude of the minimum force F
required to raise the load W can be expressed as

Pee od
fay | ——
Ea p(R,/R>)

8.97 Refer to Problem 8.96. Show that the magnitude of the force F required to
lower the load W is

ie wf}= aad
12 w(R,/R>)

Figure P8.96 and P8.97


460 Chapter 8 Friction

8.98 For the pulley and shaft assembly shown, let R,; = 3 in., R, = 9 in.,
ww, = 0.2, and W = 400 lb. Determine the magnitude of the smallest horizontal
force F needed for impending motion to raise the load.

8.99 A lever loosely fits into a shaft of radius R. W, and W, are two loads hanging
at both ends of the lever. For impending clockwise rotation of the lever, W, =
500 N, W, = 350 N, R = 200 mm, and L = 400 mm, determine jz, between the
shaft and the lever.

Figure P8.98 Figure P8.99 and P8.100

8.100 Refer to Problem 8.99. Given wp, = 0.35, W, = 300 lb, R = 4 in., and
L = 14 in., determine the magnitude of the load W, for impending counterclockwise
motion.

8.101 The bell crank fits loosely into a shaft of diameter 150 mm. Let F, = 300 N
and , = 0.25. Determine the magnitude of the force F, for impending clockwise
rotation of the bell crank.

8.102 Determine the magnitude of the force F, for impending counterclockwise


rotation. Let F, = 300 N and p, = 0.25.

250mm

Fy

|+—300 mm—+
Figure P8.101 and P8.102
Problems 461

8.103 A load of 500 lb is being raised by two pulleys. Let p, = 0.28. Also,
For pulley A: diameter = 4 in.
diameter of bearing iL tants
For pulley B: diameter = 6 in.
diameter of bearing = 1.5 in.
Determine the tension in each portion of the cable (i.e., T, T;, T>) if the load is being
slowly raised.

SECTION 8.9
8.104 A circular disc has a diameter of 28 in. If it is rolling down a sloping ground
with a slope of | vertical to 10 horizontal with a constant velocity, what will be the
Figure P8.103
coefficient of rolling resistance?

8.105 Repeat Problem 8.104 with the following: the radius of the disc = 300 mm,
and the sloping ground makes an angle of 5° with the horizontal.

8.106 A wheel has a radius of 18 in. The weight of the wheel is 50 lb. Determine
the magnitude of the horizontal force P required to keep the wheel moving along
a level surface if the coefficient of rolling resistance is 0.1 in.

8.107 Repeat Problem 8.106 with the following changes: radius of the wheel =
250 mm, weight of the wheel = 550 N, and coefficient of rolling resistance =
3.2 mm.

8.108 A wheelbarrow has two wheels. Each of the wheels has a radius of 8 in.
The wheelbarrow carries a total load W. Assuming that the coefficient of rolling
resistance is 0.5 in., determine the maximum force P that must be applied to
the handle to keep the wheelbarrow moving. Let W = 680 lb.

8.109 The wheelbarrow has two wheels and weighs 300 lb. The coefficient of
rolling resistance = 0.5 in.; the magnitude of minimum force P needed to keep the
wheel moving = 30 lb. Determine the radius of the wheels.

8.110 A freight car wheel carries a load of 30 KN. The diameter of the wheel is
0.8 m and the diameter of the axle of the wheel is 76 mm. Determine the magnitude
of the horizontal force P applied to the axle that will be needed to keep the wheel
moving. Let p, = 0.28, and the coefficient of rolling resistance = 0.4 mm.

Figure P8.110
CHAPTER

MOMENT OF INERTIA

OUTLINE 9.1 Introduction


9.2 Moment of Inertia of an Area
9.3 Radius of Gyration of an Area
9.4 Polar Moment of Inertia and Radius of Gyration
9.5 Parallel Axis Theorem for Moment of Inertia of an Area
9.6 Moment of Inertia of Composite Areas
*9.7 Product of Inertia of an Area
*9.8 Product of Inrtia of Composite Areas
*9.9 Principal Moment of Inertia of an Area
*9.10 Mohr’s Circle for Moment of Inertia of an Area
9.11 Moment of Inertia and Radius of Gyration of Masses
9.12 Mass Moment of Inertia by Integration
9.13 Mass Moment of Inertia of Composite Bodies
9.14 Summary
Key Terms
Problems

462
9.1 Introduction

9.1 INTRODUCTION
In this chapter we will discuss the principles related to the determination of
the moment of inertia of areas and the moment of inertia of masses. In order
to introduce the concept of the moment of inertia of areas, let us consider a
simple example of a container having inside dimensions of L X B X H (Figure
9.1a). The container is filled to the top with water. Let the specific weight of
water be equal to y. As discussed in Chapter 5, the distribution of the hydro-
static pressure on the wall abcd of the container varies linearly with depth.
Thus, at a depth y the pressure p = yy. Note that the y axis is taken as positive
downward for convenience, since the pressure increases downward. This vari-
ation of pressure with depth is shown in Figure 9.1b. The force dF on an
elemental area dA of the wall located at a depth y can be expressed as
dF =pdA=vyydA

(a) : (b)
Figure 9.1

If we need to determine the moment of the hydrostatic force M exerted on


the wall face abcd about the z axis, we can calculate this as

dM = y dF

moment arm

and

m= [am=[yar=| (paay=y |y*a4


The integral f y* dA is referred to as the moment of inertia of the area of
the wall face abcd about the z axis. The need for the moment of inertia of an
464 Chapter 9 Moment of Inertia

area is frequently encountered in problems where the calculation of the mo-


ment of a linearly varying distributed load at a distance from the moment axis
is required.
Another example where it is necessary to determine the moment of inertia
of an area is shown in Figure 9.2. Figure 9.2a shows a beam in pure bending.
The bending will introduce an internal distributed force along any section,
such as abcd, of the beam. Let the internal force per unit area of the cross
section—called stress—be o. The variation of o along the cross section abcd
is shown in Figure 9.2b. Note that the x and y axes are the centroidal axes of
the cross-sectional area. The internal resisting moment can be calculated as

m=|am=| yar=| yoaa


A A

Figure 9.2

From Figure 9.2b, we can write that o = ky, where k = a constant of pro-
portionality. Therefore,

M=| youa={ (ndyaa=e| yaa

= (k)(moment of inertia of the cross section about its


A = Area centroidal x axis)

In general, the moment of inertia of an area / about an axis in the same plane
can be defined as (Figure 9.3)

1=| Pa
A

The moment of inertia of an area is sometimes referred to as the second moment


of area,’ which is probably more appropriate. However, for uniformity, in this

¢ Figure 9.3 'frdA = first moment of area and, thus, f r> dA = second moment of area.
9.2 Moment of Inertia of an Area 465

chapter we will use the more common term moment of inertia, so named for
its similarity to the integrals for the moments of inertial forces of rotating
bodies studied in dynamics. The moment of inertia of an area is a measure of
how much area is located how far away from an axis.
The moment of inertia of a mass is a quantity that defines the resistance
of a body to angular acceleration. In general terms the mass moment of inertia
of a body / about a given axis can be defined as (Figure 9.4) *\| Mass=7m

p= [Pam

For a given rigid body, the mass moment of inertia is a measure of the distri-
bution of its mass with reference to a given axis. Some applications for the
mass moment of inertia will be shown in the dynamics problems.

9.2 MOMENT OF INERTIA OF AN AREA


Figure 9.4
Figure 9.5 shows an area in the xy plane in which dA is an elemental area. As
expfained in the preceding section, the moment of inertia of the elemental area
about the x axis can be expressed as

dl, = y’ dA
The moment of inertia of the entire area about the x axis, /,, can be obtained
by integration of the preceding relationship, or

I=BES | d=]
=
y 2 dA (9.1)

In a similar manner, the moment of inertia about the y axis, I,, can be calculated O
as dl, = x’ dA, and for the entire area, Figure 9.5

l= i,dl, = i,x? dA (9.2)

Moment of inertia is always a positive quantity. The dimension of moment


of inertia is (length)*; so its unit is in*, ft*, m*, and so on.
In Equations (9.1) and (9.2), the term dA is equal to dx dy. Thus double
integration may be performed to obtain /, and /,. However, as in the case of
determining the centroids of areas (Chapter 5),we can choose an elemental
area dA that has a differential width in only one direction. In that case only a
single integration has to be performed to obtain /, and /,. This principle is
demonstrated in the following examples.
466 Chapter 9 Moment of Inertia

EXAMPLE 9.1

Determine the moment of inertia with respect to the y axis, that


is, /,, of the rectangular area shown. Use an elemental area whose
length is parallel to the y axis.

Solution The elemental area shown in the figure has a


differential thickness of dx; therefore,

dA = h dx

Using Equation (9.2),


b 37? 3
Mg | 2da=[f
1,= rey Ged bebenas
rien arg ey |ea LU
if] 3

Important note: Direct application of 1, = f x? dA is possible


in the above solution since all points of the elemental area
b— —+ are at a distance x from the y axis. However, if we had chosen
an elemental area whose length was parallel to the x axis as
shown, we could not have used J, = Jfx? dA because all points
of the elemental area are not the same distance from the axis.
However, we could use the following principle.
According to the above solution,

Moment of inertia for arectangular area about a given


axis coinciding with the edge of the area (9.3)
= 3[dimension of the area parallel to the axis X
(dimension of area perpendicular to the axis) ]

Thus, for the elemental area shown in the figure,

sa 3
dit Ze )

Using this principle, we find that


9.2 Moment of Inertia of an Area 467

EXAMPLE 9.2

For the rectangular area shown, determine the moment of inertia Vy


with respect to the x axis (/,) by taking (a) an elemental area
, ; id b
whose length is parallel to the x axis, and (b) an elemental area : |
whose length is parallel to the y axis.

Solution

Part a. For the elemental area shown, we write

dA
= bdy

Now, using Equation (9.1) gives


h 3
n= fya=[ orabh 3

Part b. For the elemental area shown, according to Equation


(9.3),

dl, = ;(dx)(h’)

Therefore,

‘hal bh? 3
=| ai,=+|
ma: (pe 5 gee
yar
468 Chapter 9 Moment of Inertia

EXAMPLE 9.3

Determine the moment of inertia of the rectangular area about its


centroidal axes—that is, /,, and /,,. Note that O' is the centroid
of the réctangular area.

Solution

Determination of I: We choose an elemental horizontal strip


that is parallel to the x axis. For this strip,
dA = bdy'
+h/2 "3 +h/2
— 1\2 = 12 = a
= oytaa= [i 0a) = 0/2]

-(2)-(2))2
Determination of I: We choose the vertical elemental strip,
parallel to the y axis. For this strip
dA =hdx'
b/2 8
hb
1 == |@)IN2 aa=|_,
=
ht’)1y2 dx! a=~,
hb?
Iy = 12 <
9.2 Moment of Inertia of an Area 469

EXAMPLE 9.4

For the shaded area shown, determine (a) /,, and (5) /,. Be

Solution

Part a. The equation for the curve OA is y = kx*. We need to


determine the value of the constant k. At x = L, y is equal to
H. Thus,

H
le de k=
e
Now let us consider the horizontal elemental strip.

y yL? )dy=(L-—=vy
L
dA = cn(t — /-)=(L—
2) ( |=Hay ( Ar 5) »
ytal LE

1.=
fe)| a= |
Se, 2) (L—-—=Vy]a
Wa)

=
H
Diet beod
H
Eel SORE Rye, Touma 3
aa EL ye
kas Ug ela
|,wre Ne oe alee
_uP21
ee ‘
Ty
Part b. We choose the vertical elemental strip as shown. For
this strip, dA = y dx, and »

l= [2a Ss [ 20ay

However,

Therefore
470 Chapter 9 Moment of Inertia

EXAMPLE 9.5

y
For the shaded area shown, determine the moment of inertia with
respect to the y axis.

Solution At the outset, we need to determine the constants m


and k. Note that at x = L, y = H. Thus,

H re
y=mx m aye
ORT.
Le y=kx?

mx — kx?
and
x Y
Ae
yak sl ae,
eit)

Now we consider a vertical elemental strip as shown. For this


strip

H
dA = (mx — kx*) dx = (2 - | dx
L
H H
l= [2a =| (#.- He) ea

Lp ei Hi]
J (? to) a a 2
= —x~ — —;* =—]|]— —

9.3 RADIUS OF GYRATION OF AN AREA


The radius of gyration is a quantity with unit of length, measuring distribution
of the area from an axis; it is often used in the design of structural members
in compression. Consider an area A as shown in Figure 9.6a. Let the moments
of inertia of this area about the x and y axes be /, and /,, respectively. If a thin
strip having the same area A is placed parallel to the x axis at a distance k, as

5s

(a)
Figure 9.6
9.3 Radius of Gyration of an Area 471

shown in Figure 9.65 such that /, = Ak?, then for the area A the parameter k,
is referred to as the radius of gyration with respect to the x axis, or

(9.4)

Similarly, if a thin strip of area A is placed at a distance k, from the y axis


such that J, = Ak,, then the parameter k, will be the radius of gyration with
respect to the y axis for the area A (Figure 9.6c), or

(925)

EXAMPLE 9.6

Refer to Example 9.1 and determine the radius of gyration of the The area of the rectangle is A = bh, and thus the radius of
rectangular area with respect to the y axis. gyration is

Solution In Example 9.1, we determined that the moment of


inertia about the y axis was

_ hb?
Pe:

EXAMPLE 9.7

Refer to Example 9.4. Determine the radii of gyration of the area Integration gives us
with respect to the x and y axes.

Solution In Example 9.4, we found the moments of inertia to


be

Knowing the area and the moments of inertia, we find the radii
of gyration of the area:
In order to determine the radii of gyration k, and k,, we need to
determine the area A of the shaded portion OAB (refer to the
solution of Part a, Example-9.4). For the horizontal elemental
strip,

a= (1-ev5) 0
472 Chapter 9 Moment of Inertia

9.4 POLAR MOMENT OF INERTIA


AND RADIUS OF GYRATION
Figure 9.7a shows a circular shaft of radius R subjected to a twist or torsional
moment, T. This torsional moment will be resisted at each cross section (such
as abcd) by a distribution of tangential stress, tT—that is, tangential force per
unit area, as shown in Figure 9.7b. The distribution of t is shown in Figure
9.7c. Note that O is the centroid of the cross-sectional area of the section, and

t=kr

where k = aconstant of proportionality and r = radial distance. The resisting


moment of the section about the centroidal axis—that is, the z axis, which is
perpendicular to the cross section—can then be calculated as

M= lhdM = [,r dF = r(t dA) = i(r)(kr)(dA) = k ir? dA

(c)
Figure 9.7
9.4 Polar Moment of Inertia and Radius of Gyration 473

The integral f 7? dA is referred to as the polar moment of inertia. In general,


referring to Figure 9.8, the polar mement of inertia of the shaded area (which
ia Be TS
is in the xy plane) about the z axis can be expressed as 1
|
|
|
|
|
Jos idA (9.6) |
|
|
\ |
where J, = polar moment of inertia about the pole O. It is important to note
Fay |
the differences between the integral f y? dA as given in Equation (9.1) [or
J x? dA as given in Equation (9.2)] and the integral fr? dA given above. The
oats i|

differences are:

1. The area in reference and the moment axis in the development of


Equation (9.1) are in the same plane. However, the area in the derivation Figure 9.8
of Equation (9.6) is normal to the moment axis.
2. In Equations (9.1) and (9.2), y and x are rectangular coordinates,
respectively. However, r in the derivation of Equation (9.6) is the radial
coordinate.

From Figure 9.8, it can be seen that


ie lle y’

Thus ay

Thin strip; Area=A


Jo=Jor+Praa=|ras| rane 7
This relationship is very useful in solving practical problems.
The polar radius of gyration, kp, can be defined as

ko = VJo/A (9.8)
The polar radius of gyration can be visualized by placing a thin strip of area
A at a distance k, from the pole O so that its moment of inertia about the z axis
is equal to Jp, as shown in Figure 9.9. Figure 9.9
474 Chapter 9 Moment of Inertia

EXAMPLE 9.8

AR
Determine the centroidal polar moment of inertia of the circular
area shown. Also determine the polar radius of gyration.

Solution We consider an elemental annular area as shown. For


this area,

dA = 2ur dr

=
Therefore, using Equation (9.6),
R 4
Jo = i? dA = i (P)\(Qnr dr) = =R

<

The centroidal polar radius of gyration is found from Equation


(9.8)as

pee ee
Oy A) y tar?) NS ko

EXAMPLE 9.9

y A circular sector is shown. Calculate (a) I, (b) I,, (c) Jo, and
(d) k,, k,, and Ko.

Solution

Part a. For the differential area, dA = (r da) dr. Then

1,= |yaa =|{ (r sina)*(r


da dr)
a=0.873 rad pr=12in.

= [ [ r? sin?a dr da
a=0 r=0

r4 12in. 9.873 rad


Fe yes
=}|— [ sin? a da
(Note: @=50°=0.873 rad) 4], 0
0.873 rad
=sisa [ sin? a da
7 0.873 rad
i= 5184 Is~ ze] = 986.5 in*
0

1, = 986.5in? <
(continued)
9.5 Parallel Axis Theorem for Moment of Inertia of an Area 475

EXAMPLE 9.9 (concluded)

Part b. Part d. First computing the area of the sector,


a=0.873 rad pr=12in.
OR 2 0.873)(12 2)
l= [x d= [ay i (r cosa)?(r da dr) At is WOO “ 13 62.86 in?
0.873 rad

—a sin2a
= 5184] E 4 | = 3539.0 in; in* ie
1
hi
i
986.5
62.86 3.96 in k,i = 3.96 in. <
I, = 3539.0in*t <
y h 3539 in <
Part c. bin ==Waar! oo gguae ost
|= k, = 7.50 in.

Jot, + 1e= 980.5 +°3539.0


Re
= \|
fe)
Jp =4525:5int ik

9.5 PARALLEL AXIS THEOREM FOR MOMENT


OF INERTIA OF AN AREA
If the moment of inertia of an area about a given axis passing through its
centroid is known, then the moment of inertia of the same area about an axis
parallel to the centroidal axis can be obtained by using the parallel axis theo-
rem, or vice versa. The principles of the parallel axis theorem can be explained
by referring to Figure 9.10. Let x’ and y’ be the centroidal axes. Also let the
moment of inertia of the shaded area about the x’ and y’ axes be /,, and /,,,
respectively. The x axis is an axis which is parallel to the centroidal x’ axis.
Now

1,= | 2aa=| (a,+y92aa =a |aa +24, |y'aa+ |y2a0

O
Figure 9.10
476 Chapter 9 Moment of Inertia

Note that, in the above equation, the term fy’dA = 0. Also the term
fy’? dA = I,’. Thus the moment of inertia of an area about the x axis parallel
to the centroidal x’ axis is given by

=a+.Ad (9.9)

In a similar manner we can show the moment of inertia with respect to the y
axis as

1,=1,+ Ad (9.10)

and that for the polar moment of inertia,

Jo =Jo + Ak, (9.11)


where dg = dz + d>.
In general terms, the parallel axis theorem can be stated in the following
manner’

Moment of inertia of an area about a given axis, /

©
Moment of inertia of the area about a centroidal axis parallel to
the given axis, log

The product of the area (A) and the square of the perpendicular
distance (d”) between the given axis and the centroidal axis

or

1 =Icg + Ad? (9.12)

* Important note: The parallel axis theorem applies only from (or to) a centroidal axis, and does
not apply between any other two parallel axes.
9.5 Parallel Axis Theorem for Moment of Inertia of an Area 477

EXAMPLE 9.10

The circular area discussed in Example 9.8 is shown. Use the y


results of Example 9.8 and the parallel axis theorem to determine
I, and I,.

Solution From Example 9.8, we know that

jue2
And from Equation (9.7),

Joc= it Ly
Due to symmetry, we see that
= Ty O

Therefore,

Jo Ra

Based on the parallel axis theorem given in Equation (9.9),


I, = I, + Adz. For the circle, d, = R. Thus we have

= " 2 Dir
5aR* paca
if 4 (7R )(R) 4 x 4

Similarly,
478 Chapter 9 Moment of Inertia

EXAMPLE 9.11
y y A circular sector is shown. Determine /,, and /,, (that is, the mo-
ments of inertia about the centroidal axes).

Solution We consider an elemental area as shown, for which


dAv— 77,40 ar
Gace r=R
a [» dA=|_ ie (r sin 8)?(r d@ dr)
Sel

: RO 2 eek |Oeming
ie Ta a a5. mame
R*

= sp — sin a COs a)

The parallel axis theorem gives /,, = J, — Ad. We know that


A = aR’ and d, = 0. Therefore,
R*

le= 7 — sin a cosa) — (aR*)(0)


4

l= re —sinacosa) <

Note: This result should have been obvious, since Ox and Ox’
coincide.
6=+a sr=R

l= [2 CIN i (rcos 0)*(r dé dr)

ORG
= cos“ a
8 d8 = —(a e
+ sin a cos a)
AW =<. 4

Again, the parallel axis theorem gives J, = I, — Ad;. We also


QR si (see Table 9.1). Thus,
know that A= aR? and d, = =a Qa
2
ie 2R si
Ll Ad = a0 + sin a cos a) — (aR?) (Ax2)
3a
R* 4R* sin?
[= (Gt SI COSI!) — eee <
¢ 4 9a
9.6 Moment of Inertia of Composite Areas 479

9.6 MOMENT OF INERTIA OF COMPOSITE AREAS


Figure 9.11 shows a composite area, ABCDE. A composite area consists of a
series of connected simple areas such as rectangles, circles, and triangles. In
order to determine the moments of inertia of a composite area about given
axes (i.e., x and y axes), the following steps should be followed.

1. Divide the composite area into several parts consisting of simple areas
SuenBS th, Wh SZ soa a
Determine the areas of the parts. Let these be A,, A>, A3, Ay, ....
3. Determine the coordinates of the centroids of these parts with respect to
the x and y axes. In Figure 9.11, these are (X,, y,), (Xo, Yo), (%3, Y3),
(COS RACE
4. Calculate the moments of inertia of the parts about their centroidal axes
parallel to the x and y axes. Let these be /,, and /,, for area 1,2 /,,
Xo, and
Figure 9.11
I,, for area 2,
2
ands
x3 ¥3
ea tOmatedrs
wien andsn folmaled 4a
x4 y4

5. Calculate the moment of inertia of each part about the x and y axes by
using the parallel axis theorem, or

bi: a Le a Ay)"

G = My, + AG)
For area | ss

ee Te AS (y,)-
Forarea2) 72. ees
fa= Tyg + Age)”
Continue in this manner for all parts.
6. Calculate the moments of inertia of the composite area as

eve ee eta Ge (9713)


Bel hht Lot Lond.

The application of the above procedure is shown in the following example


problems. Table 9.1 presents the moments of inertia of some simple areas
obtained by the integration method. Many of these are widely available in
engineering handbooks.
Rolled steel of various shapes is used in the construction of beams, columns,
and other structural members. The cross-sectional dimensions of some com-
mon wide-flange shapes (W shapes), American standard shapes (S shapes),
and American standard channels (C shapes) are given in Tables 9.2 through
9.4, along with the moments of inertia and radii of gyration about their cen-
troidal axes.
Table 9.1 Moment of inertia of some common areas

Coordinates of Movement of inertia


Shape
| Om
Rectangle
y y [p= Gon
Lye hb

I, = bh?
I, = 3hb?
b3n3
O
Toy" = 6(52
+h2)
Triangle
bh3
eee a isce
a
: 1
Iy = 22(6? — be +c?)
7
Vi a
3 bh3
bea hoe

t bh3
te 3G
beh
Ly = 36
_ bh3
mae
_ beh
I, =e

pecs!
cd
sad!
eG.

I oT een |
yale 64

om _ 5ar4+_ snD4
Lie ilar

Semicircular area

yoy?
Ud
(912 — 64)r4

480
Table 9.1 Moment of inertia of some common areas (continued)

Quarter-circle area
!
y y _ (9n2— 64)r4_
| een sds oe
|
pes
SmamITG
_ 4
——-x lp Te

Ellipse

oie
mba?
aa

Circular sector y!
— 2rsi r4 :
ees I.x =>(a@—sinacos
4
a) =/,:x
3a
—— 4
‘if .
a I, =q (at sin a cos a)
4 Armd)
if : 4 r*sin-a
ly 4 (a+
.=__—_— +
sin @ cosa) x sy a
5 )

x=0
>) rt Lae 3:
pa 4 xX if.
) = 7 (a+2sin Q cos &— sina cos @)

sinta r4 ee 4
ED. fy = 79 (3a — 2 sin acosa—3sinacos a) =/,)

ri h3 (bi + 4byb +b5)


e -36(b, +03)

Note: Point O' as shown in the figures in the left-hand column is the location of the centroid.
481
Table 9.2 Properties of some common rolled steel wide-flange shapes (W shapes)

Flange "axis y —y axis


Uy ' .

Area Depth, Web, by ty I ky. Iy kk.


Ye

d Le
Shape Designation (in.2) (in.) (in.) (in.) (in.) (in4) (in.) (in.4 ) (in.)

y! W 24 X 55 16.2 2395700395 O0See OF50SmetsSOmeoe0) 1.34


|
|
W 18 X 50 14.7 17.99 0.355 7.495 0.570 800 7.40

W 16 X 36 10.6 15.86 0.295 6.985 0.430 448 6.51

W 14 X 26 Ted 1391025 0250.4 20) 245 5.65

W 10 X 22 6.5 LOO? 40S a5 0060 118 4.27

W 8X 18 533, Sal 402s 0a Om OS a0 62 3.43

Note: OQ’ is the location of the centroid.


I, and J,,, are moments of inertia about centroidal axes.
Kk, and ky are radii of gyration about the x'—x’ axis and y’—y’ axis, respectively.
W 24 X 55 means nominal depth is 24 in., and weight is 55 Ib per ft.

Table 9.3. Properties of some common rolled steel American standard shapes (S shapes)

Flange 36/50! gyais


Area ee Web, by ty I ies
ie
Shape Designation (in.) (in.) (in.) (in.) (ins) (in) en) (in.)

S 20 X 66 19242 OO OS 05a 625 Din O19 SelOO ESS IAs)

Suis X50 14.7 15.0 0.550 5.640 0.622 486 5.75 1.03

SS WAOK 35) NOs 12.0 0.428 5.078 0.544 229 4.72 0.98

SLOP ess) 10.3 10.0 0.594 4.944 0.491 147 3.78 8.4 0.90

SP Sexes 6.7 8.0 0.441 4.171 0.426 65 3.10 4.3 0.80

Note: O’ is the location of the centroid.


7, and J, are moments of inertia about centroidal axes.
k,, and k,, are radii of gyration about the x'—x’ and y'—y’ axes respectively.
S 20 X 66 means nominal depth is 20 in., and weight is 66 lb per ft.

482
Table 9.4 Properties of some common rolled steel American standard channels (C shapes)

Flange X—x axis y'—y’' axis


% Area Depth, Web, by Fe! ioe aoe Ls ky
ts
Designation Gna ue Gn) “Gn). Gn.) Gn.) Gn4) Gn Gn.*) Gn.)

CIS eX 0 0.798 14.7 15.0 0.716 3.716 0.650 404 5.24 11.0 0.876

G12 x30 0.674 8.8 20 O50 S170 OSO Ue coos Sell Were}

CRON 20 ORG OGRE WOO OS) Do WAS TW) 35 Ds Wyss

C Bx 12 0.571 3.4 SrOMmOn 20 ee 260 O SSO MSS sell 30629

' Cl 6x8 0.511 2.4 CO Ol20Oe Po? 00343 meee See 34 ee Ol 60.5317

Note: O' is the location of the centroid.


"7, and /,, are moments of inertia about centroidal axes.
k,, and k,, are radii of gyration about the x'—x' axis and y'—y' axis respectively.
C 15 X 50 means nominal depth is 15 in., and weight is 50 lb per ft.

EXAMPLE 9.12

For the composite area shown, determine the moments of inertia


I, and I, about the x and y axes, respectively. (Note: The x and y
axes are also the centroidal axes for the composite area.)

Solution The composite area can be divided into three


rectangles as shown. We can now prepare the following table.

Area Area, A; “MG Law = =5


No. (in?) in* (in*) ix; Aiy;
1 (24)(2) = 48 i9(24)(2)° i9(2)(24)° 0
2 = (10)(2) = 20 72(2)(10)° : 72(10)(2)? 720 2426.67
3 _(10)(2) = 20 72(2)(10)° . 7(10)(2)° 720 886.67 2426.67
I, = 1789.34 in* I, = 7157.34 int <
rem oo 2
Ley = Leg + AYO Ty = Iw + Axe
It is important to note that the moments of inertia can be
added only if each has been calculated about the same axis.

483
484 Chapter 9 Moment of Inertia

EXAMPLE 9.13

yp y Determine the centroidal moment of inertia /,, for the composite


| area shown. Note that O’x’ and O’y’ are centroidal axes of the
composite area.

Solution

Determination of d,. The composite area can be seen to be

A A = A, a A,

t {
Area of the square Area of the quarter circle

d See — Any2
A, — A,

. ar? oth 4r
A,=h ae re

£ Therefore,

y
x
wo) - (F)(=s)
D)

a ula
4 30 2 3h Ls Ir?

— 3\ 4h? - ar?
4

Determination of I, and I,;._ For Area 1, the centroidal


moment of inertia (see Table 9.1) is

1
La he
Howl?
The centroidal moment of inertia for Area 2 (see Table 9.1) is

_ On? — 64)r*
* 1447
Determination of [,,. For the composite area,

Ty = Uy + Ard, — Wy)? + Leg + An(, = Fa)"


a as Fs

Therefore,

es ee * , Or = 64y4
mms Camis (AWTS) 144
2
“ ne 2 3h? — 2r° 4r
4 )|3\4r = oP
9.7 Product of Inertia of an Area 485

EXAMPLE 9.14

Two rolled steel C15 X 50 channel sections are welded as shown.


Determine /, and /,.

Solution

Determination of I... We write the equation for the moment of


inertia of the composite area:

T= |e+ Al =
“Ne ap lke ae All|=
Na =2|1.+ Al =
aN
é ; 2 : D : »
—— SY

For Section | For Section 2

Now using Table 9.4,


2)
1
l= 2|
404+ 04n(¥) I, = 2461.75in.4 <

Determination of I,.

Le= [+ AG?) + Uy + Alby + (6, — x) 1°}

For Section | For Section 2

Again using Table 9.4,


I, = (11.0 + (14.7)(0.798)7]
+ {11.0 + (14.7)[G.716) + (3.716 — 0.798)]?}
I, = 678in.* <

*9.7 PRODUCT OF INERTIA OF AN AREA i ak


In Section 9.1, we mentioned that the moment of inertia of an area is a measure
of how much area is located how far away from an axis. However, the moment dA dA
of inertia is not the only parameter for characterization of the distribution of
area about an axis. This can be explained with reference to Figure 9.12. Every
elemental area dA in the first quadrant has as its mirror image an elemental O
area in the second quadrant. Thus it can be seen that the magnitude of /, for
the area A in the first quadrant will be the same as the magnitude of /, of the
area in the second quadrant. The moment of inertia /, of both areas will be the
same. However, the location of an area in different coordinate quadrants can Figure 9.12
486 Chapter 9 Moment of Inertia

be differentiated by defining a quantity called the product of inertia of an


area (Figure 9.13), which we may write as

es if
xy dA (9.14)

where /,,, is the product of inertia of the area with respect to the x and y axes.
The product of inertia of an area with respect to certain rectangular axes
may be a positive or negative quantity, or may be zero. This can be explained
with reference to Figure 9.14. As shown in Figure 9.14a, if the area is located

O
| in the first quadrant, both x and y coordinates for every elemental area dA are

Figure 9.13

2nd quadrant lst quadrant

3rd quadrant 4th quadrant

(a)

(b)
Figure 9.14
9.7 Product of Inertia of an Area 487

positive and, hence, /,., is positive. If the area is located in the second quadrant,
the x coordinate for every elemental area will be negative while the y coor-
dinate will be positive, which will result in a negative value of /,,. Similar
considerations will show that, in the third and fourth quadrants, the values of
I, will be positive and negative, respectively. We can also see from Figure
9.146 that if the area is symmetrical about one of the axes, the product of
inertia + xy dA of each elemental area dA to the right of the y axis, there will
be an elemental area to the left of the axis whose product of inertia will be
—xy dA. This will result in a net value of zero for the product of inertia for
the entire area, J...
As in Section 9.5, we can also derive a parallel axis theorem for obtaining
the value of /,., for an area with respect to given rectangular axes x and y, if
the value of the product of inertia of the same area with respect to a set of
centroidal axes parallel to the x and y axes is known. This can be accomplished
in the following manner. Referring to Figure 9.15, x’ and y’ are centroidal
axes for an area A. The product of inertia of the area with respect to the x and
y axes is

I= |eo) da = | a. +204, + yA
=| wyyar+a.].y aA +d,| x dA + dd, |dA O
Figure 9.15
As explained in Section 9.5, f y’ dA = O and f x’ dA = 0. Thus, the parallel
axis theorem for the product of inertia of areas can be written:

1, = [,x'y' dA +d.d, \dA =Iyy + Add, (9.15)

where /,,,, is the product of inertia of the area with respect to the centroidal
x' and y’ axes.
The dimension of the product of inertia will be (length)* and, hence, its
units are cm*, m‘*, in*, and so forth. The use of product of inertia will be shown
in Section 9.9. As in the case of obtaining the moment of inertia of an area,
we can use elemental strips to avoid double integration.
EXAMPLE 9.15
y Determine the product of inertia /,,, of the shaded area shown.

Solution We consider a vertical elemental strip having an area


dA = y dx. The coordinates of the centroid of this elemental
strip are x,, = x and y,, = y/2. For the elemental strip, let
x' and y’ be the centroidal axes. From Equation (9.15),

dl, F dl, a (d, dy) dA

For this case, d, = X,, = x and d, = y,, = y/2. So,

Gl = a Po. dx):
(2)o

Due to symmetry, d/,,,, = 0. Hence,

1 ]
Ly = [a, =4 [2 xar—4e [ear
_lp xen + Dd ee xen +L) _

Da | 2a el G 4m +1) |,

Now, we need to determine the value of k in terms of L and H.


Since y = kx’,
(ahaa or Ka

Hence,
Gs 2
(n+ 1) 2(n+ 1) H?L2
jet ee = A); pee l= <
9 4n+1)|, \2") [4m4+D}] 7% 4@+4+1)

EXAMPLE 9.16

Determine the product of inertia /,, of the rectangular area shown.

Solution We consider a vertical elemental strip for which


dA = hdx. As in Example 9.15, for this elemental strip
dl, = dl, + (d, d,) dA
However, dl... = 0, d, = x. = x, and d, = y,. = h/2.
Therefore,

h h?
diy, = (d,d,) da = x5 (h dx) = x dx

488
9.8 Product of Inertia of Composite Areas 489

*9.8 PRODUCT OF INERTIA OF COMPOSITE AREAS


The product of inertia of composite areas can be calculated by using the par-
allel axis theorem for product of inertia of areas in practically the same manner
as for the case of the moment of inertia of areas described in Section 9.6. The
procedure is described below in a step-by-step manner with reference to Fig-
ure 9.16.

1. Divide the composite area into a number of simple areas such as 1, 2, 3,


“al oe
Determine the areas of the simple divisions as A,, A>, A3, Ay, ...-
3. Locate the centroids of the simple areas as (X,, y,), (X2, ¥2), (X3, ¥3),
(re Valeo ok «
4. Calculate the product of inertia of the simple areas with reference to O 43
their centroidal axes as L111), Leyiays
x
Letyiays Lety(ayy oe
Figure 9.16
5. Use Equation (9.15) to obtain the product of inertia of each area with
respect to the x and y axes as

Caan = POVeby im A\x1y\

Tyay = Letyay + ArXay2


Ty) = Letyay + AsXaY3
Lo ey Te Agx4y4
Continue in a similar manner for all simple areas.
It is important to note that areas such as 1 and 4 in Figure 9.16 will
have /,.,, equal to zero since they are symmetric about at least one
centroidal axis.
6. Obtain the product of inertia of the entire area as /,,, = 1xy ay LEO
fe) + taxa To 7*-
The products of inertia with respect to the centroidal areas for some com-
mon areas are given in Table 9.5.
Table 9.5 Product of inertia of some common areas with respect to centroidal axes (x’ and y’)

Area I yt

Triangle y’

= bh?(Qe—b)

_— at b2h2
72

on — 3
( aaa yr

490
EXAMPLE 9.17

For the area shown, determine the product of inertia /,,.

Solution We can divide the area into two rectangles, that is,
1 and 2, as shown. From Equation (9.15), for each area,

Ixy a Tey ar Axy

However, for each area /,,,, = 0; so, for each area, /,,, = Axy.
The following table can now be prepared.

Area, A; Xj Ji A;X,Y;
Area No. (in?) (in.) (in.) (in*)
30 5 IES) DDS)
24 1 9 216

M> XY; = ly = 441in* < be 10 in.

EXAMPLE 9.18

Determine the product of inertia of the composite area with re-


spect to the x and y axes shown.

Solution The area can be divided into three triangles as


shown. The following table can now be prepared.

-—150 mm—

Ley A AiX;Y; I = Ley + A;X,Y;

(mm4) (mm?) (mm‘) (mm‘*)


—=(120)7240) = =—11.52 x 10° 14,400 46.08 x 10° 34.56 x 10°

+= (12071507 = AG se 1° 9000 —18 x 10° —13.5 x 10°

1
+55 (90)"(150)" = 255i X< 110° 6750 2025 <a0? 22.781 Xx 10°

Ly = 43.841 X 10°mm* <

491
492 Chapter 9 Moment of Inertia

*9.9 PRINCIPAL MOMENT OF INERTIA OF AN AREA


ag The moment of inertia of an area varies with the rotation of the coordinate
axes. There is one orientation of the coordinate system that will produce max-
imum and minimum moments of inertia. These /,,,, and /,,;, are called the
principal moments of inertia. We will now develop a general relationship for
determining the moments of inertia of an area about an inclined axis. Figure
9.17 shows an area whose moment of inertia about the x and y axes are,
respectively, /, and /,. Let it be required to determine the moments of inertia
of the same area about the m and n axes shown in the figure, and also the
product of inertia with respect to the m and n axes. By definition,

I= |n2aA 1, = |maa Iyy = |mn dA

Examining Figure 9.17, we can see that

m=xcosat+ysina n=ycosa-—xsina

Figure 9.17 Substituting the preceding relationships for m and n into the expressions for
I» 1, and I,,,, we obtain the following. About the m axis,

[ae |creos a — x sin a)*dA

- |» cos?
a dA + [2 sin?
a dA —

[2» sin a cos a dA

I, = 1 cos: oI, sin a — 1. sin 200 (9.16)

About the n axis,

Le Jo cos a + y sina)? dA

= [2 cos? a dA +» sin?
a dA +

[20 sin a cos a dA

I, = I, cos’ a + I, sin? a + I, sin 20 (9.17)

And, for the product of inertia,

Lian = |ccos a+ ysina)(y cos a — x sina) dA


: : 5 ip (9.18)
= 1, sin a cos a — J, sin a cos a + T,(Cos” a — sin* a)
9.9 Principal Moment of Inertia of an Area 493

Recall the following trigonometric identities:


sin 2a = 2 sina cosa
y 1 + cos2a
COS Oren
2
es 1 —cos 2a
sin Largreny = =

Using the preceding relationships, Equations (9.16), (9.17), and (9.18) can be
expressed in more convenient terms as

pe Lee ee fe i
mee 5 ap cos 2a — I sin 2a (9.19)

| cee reed &


l= eo To cos 2a + I,, sin 20 (9.20)

yee Te bap be
mn — 5 sin 2a + I. COS 2a (9.21)

Adding Equations (9.19) and (9.20), we obtain the first invariant, which is

In +1, =1, + 1, (9.22)


It is apparent from Equation (9.19) [or (9.20)] that the magnitude of /,, (or
[,) varies with a. Now we may ask ourselves the following questions:

1. What are the magnitudes of the maximum moment of inertia /,,, and the
minimum moment of inertia /,,, as the inclination angle « is varied?
These maximum and minimum moments of inertia are referred to as the
principal moments of inertia.
2. What are the values of a = a, (i.e., the direction of the axes) that
correspond to the maximum and minimum moments of inertia as stated
above? These axes are referred to as the principal axes.

In order to answer these questions, we can use the principles of maxima and
minima from calculus; thus

die
= =
da
el,
—2 sin preRe] Be cos 2a= 0
»
494 Chapter 9 Moment of Inertia

Or, using a = @,,

tan 2a,, = ————— (9.23)


bal t/Lytle
2

From Equation (9.23), we can also write that (also see Figure 9.18)

lees
2
COS 20 fie (9.24)
2
if i! L, 2

9) ay

i, =a
hr
and
axe ge

sin 2a, = (9.25)


—)
mee watle
Figure 9.18 If Equations (9.24) and (9.25) are substituted into Equation (9.19), we
obtain
N Leal LHP \2
ip =r
P2
sha 2 (*») :)+7 ay (9.26)
Note that Equation (9.23) will result in two values of a, that are 90 degrees
apart. One of these a, values will correspond to an axis about which the
moment of inertia /,, = Le and the other value of a, will correspond to an
axis about which /,, = [,.; that is, the maximum and minimum moments of
inertia.
Following is a step-by-step procedure to calculate the principal moments
of inertia /,, and/,,, and also the orientation of the principal axes.

1. Obtain /,, /,, and /,,.


2. Use Equation (9.23) to determine the two values of a, that are 90
degrees apart.
3. Use Equation (9.26) to obtain i and /,,.

4. In order to identify which value of a, (as calculated in Step 2)


corresponds to which principal moment of inertia (as calculated in Step
3), use one of the a, values determined in Step 3 in Equation (9.19) and
calculate /,,, or

at
x y edt x | y
I cos 2a, _ Le sin 2a,
de 2 2
9.9 Principal Moment of Inertia of an Area 495

This value of /,, will coincide with one of the values calculated in Step
3; the other value of a, will correspond to the remaining /,,.

EXAMPLE 9.19

Refer to Example 9.12. For the area determine the principal mo-
ments of inertia /,,, and /,, and the orientation of the correspond-
ing principal axes passing through the centroid.

Solution The area considered in Example 9.12 has been drawn


again. Note that O is the centroid of the composite area. From
Example 9.12, /, = 1789.34 in* and J, = 7157.34 in*. We
need to determine the product of inertia /,,. In doing so, we can
prepare the following table:

Area A; Ty(i) Xx; Yi Tey + Axi;


No. (in?) (in*) (in.) (in.) (in*)
1 48 0 0 0 0
2 20 0 =i +6 — 1320
3 20 0 ap ili =6 — 1320
Ly = — 2640 int

Now we can determine the principal moments of inertia. From


Equation (9.26), we write

IL ae dl a TENZ
teed x 4y
typ, 7 + ( :)ate is
2 D
_ 1789.34 + 7157.34
2
x (es 7157.34
+ (— 2640)?
2 2
= 4473.34 + 3764.76 (in*) Now, let us substitute a, = — 22.27° into Equation (9.19):
1, = 8238.1in* < i
fects ovteaBn Br
i 5 aF cos 2a, — I,, sin2a,
I, = 708.58 int <
To determine the orientation of the principal axes, from _ (1789.34 + 7157.34 be 1789.34 — 7157.34
equation (9.23) we write is 2 2
Ai — (—2640) X cos (— 44.54°) — (— 2640) sin(—44.54°)
7
tan 20, = ——— = _______ = — 0.984
I,—I, (1789.34 — 7157.34 = 708.73 in*
2 s
2a, = —44.53° and 135.47", so Therefore,

a, = —22.27° and 67.73° i 67.73" (Onaxis) and a, = —22.27° (Om axis) «


496 Chapter 9 Moment of Inertia

*9.10 MOHR’S CIRCLE FOR MOMENT OF INERTIA


OF AN AREA
A graphical procedure can be conveniently used to obtain the principal mo-
ments of inertia and the orientations of the principal axes for an area A. In this
section we will discuss this graphical procedure. From Equations (9.19) and
(9.21) we can write

R+h\) [(h-4 aati


ae 5 = 5 —} cos 2a — J, sin2a (9.27)

and
I,-1, ‘
(mae aE, sin 2a + I, cos 2a (9.28)

By adding the left-hand and right-hand sides of Equations (9.27) and (9.28)
we obtain

I,+1,\] I,-1,\7
a oe Bee ees trl (9.29)

This equation is of the form (x — a)* + y? = R?, which is the equation for
a circle, for which the abscissa (x) is the moment of inertia (/) axis, and the
ordinate ( y) is the product of inertia axis. This circle, referred to as the Mohr’s
circle, is named after the German engineer Otto Mohr (1835-1918). The cen-
ter of the Mohr’s circle is located at a point that has coordinates of a, 0. Note
that

a = (9.30)
9.30

The radius of the Mohr’s circle is equal to R, or

Ly ( My )

Use of Mohr’s Circle


In order to use the Mohr’s circle to obtain Los ES a, and ane the following
step-by-step procedure may be adopted (with reference to Figure 9.19).
1. Obtain the area A as well as /,, I,, and [,, with reference to the x and y
coordinate axes (Figure 9.19a). For our example, let /, > /,,and I,,, be positive;
however, in general this is not necessary.
9.10 Mohr’s Circle for Moment of Inertia of an Area 497

Product of
) inertia
n
Wa
Je
Ge

a
MOL
x
het,
Ne)]
)

Qa 3)
|
i p(2) IP 2X | IP
a x O 2. a C = : Moment
NS . a .
Op (1) Ke of inertia
x x v
SS 2
\m CO is
Tet,

(a) 2 a

(b)
Figure 9.19

2. Adopt a set of coordinate axes with the moment of inertia being the ab-
scissa and the product of inertia being the ordinate (Figure 9.195).
3. Plot point X with coordinates /,, /,,,, and plot point Y with coordinates /,,
pe Figure 9.19D).
4. Join points X and Y by a straight line. Mark point C which is the inter-
section of line XY with the moment of inertia axis (Figure 9.195).
5. With C as the center and CX = CY = R as the radius, draw a circle
(Figure 9.196). This is the Mohr’s circle for the problem under consideration.
It is important to note that [Equation (9.30)]

oat Lady
OG a ;

and the radius of the circle is [Equation (9.31)]


wl valine 2

CX =CY=R= (3 ) + 72

Also,

XD I
tan 2a, == = 24
CD (Ce vb,
2)
or

XD i
2a, _= tan ih (2)
[Pes Wy= tan =i) /f(cae (9732)

Z
498 Chapter 9 Moment of Inertia

The magnitude of a, obtained from the preceding relationship is the same as


defined by Equation (9.23).
6. Determine the principal moments of inertia (Figure 9.19b) as

2 BD 7

and

aed ged i

eee Leis i"

2 D AES

Note that P, and P, are the points on the Mohr’s circle whose abscissas yield
the maximum and minimum values of the moment of inertia /.
7. In order to determine the orientation of the principal axes, calculate [from
Equation (9.32)]

1 Ty
lo,,| = = tan~ || -—-
To) ge)
2
In Figure 9.195, note that if we start from X to come to P,; we must move
through an angle of 2a, in the clockwise direction. Thus, we need to move
through half of that angle (that is, 3 of 2a, = a,) in the same direction—
clockwise in this case—from the x axis in Figure 9.19a to reach the orientation
of the axis m about which the moment of inertia of the area A is equal to /, .
The n axis perpendicular to the m axis can then be easily determined. The
moment of inertia of the area about the n axis will then be equal to /,.
9.10 Mohr’s Circle for Moment of Inertia of an Area 499

EXAMPLE 9.20

Solve Example 9.19 using Mohr’s circle. : Product


BS of
inertia (in‘*)

Solution From Example 9.19, we know that J, = 1789.34 in#,


If: = 7157.34 in*, and I — 2640 in*. With these values, (7157.34, 2640)
the Mohr’s circle is shown. The coordinates for points X and Y
are as follows:
3764.76 =R~_
Point X: J, = 1789.34 in*, I, = —2640in*
Point Y: I, = 7157.34 in*, —I,, = 2640 in* ot Moment
From the figure we en

Toe si SE aie on oe
3 a (1789.34, —2640)
|CD| = |CE| = || = 2684 in’

|Cx| = |CY| = |R| = V (CD)? + (Dx)?


= V (2684)? + (2640)? = 3764.76 in*
I,, = OP, = OC + CP, = OC + R = 4473.34 + 3764.76
I,, = 8238.1 int <
I,, = OP, = OC — CP, = OC — R = 4473.34 — 3764.76
I,, = 708.58in* <
2640
2a,, = ZXCP, = tan™' ( 35 AT
2684
a,, = 67.73° <
Hence, if we move 67.73° in a counterclockwise direction from
the x axis we will reach the orientation of the axis about
which the moment of inertia of the area is i 298238, Lain,
The orientation of the principal axes are shown in the figure of
Example 9.19.
500 Chapter 9 Moment of Inertia

9.11 MOMENT OF INERTIA AND RADIUS


OF GYRATION OF MASSES
Definition

We briefly discussed the moment of inertia of a mass about a given axis in


Section 9.1 as a measure of the distribution of a mass with respect to a given
axis; thus,

[= [v dm (9.33)
where dm = elemental mass in a body, and r = perpendicular distance from
the elemental mass to the axis. The term fr? dm appears in the study of
dynamics, in the formulas related to the kinetic energy of rotation, torque
acceleration, and angular momentum. With this basic definition we can write
the moment of inertia of a mass about the y axis (as shown in Figure 9.20) as

Noe ir*dm= |(x? + 27) dm (9.34)


Figure 9.20

Similarly, the moments of inertia with respect to the x and z axes can be
expressed as

ies [v dm = i(y? + 2*) dm (9.35)

(i ir2>dm= |(x? + y*) dm (9.36)

The moment of inertia of a mass is always a positive quantity. The unit for
the moment of inertia of a mass in SI units is kg - m*, and in U.S. Customary
units it is slug - ft.

Radius of Gyration
As in the case of plane areas (Section 9.3), the radius of gyration of a body
can be expressed as
jt
k= |-
m
where k = radius of gyration, and m = mass of the body. The radius of
gyration k is defined as the perpendicular distance from the axis to a point
where the entire mass m must be concentrated such that k*m will be equal to
9.11 Moment of Inertia and Radius of Gyration of Masses 501

I (Figure 9.21). Thus the radius of gyration of a mass about the x, y, and z m

axes will be

fl I I
k=
: fe
/—
y fo
k= |2
k, EA fe
ez
(9.37)

Parallel Axis Theorem


If the mass moment of inertia of a body about an axis passing through its Figure 9.21

center of gravity is known, then the moment of inertia of the body about
another parallel axis can be determined by using the parallel axis theorem.
This is similar to the procedure described in Section 9.5 for the case of the
moment of inertia of the area. Figure 9.22 shows a body whose center of
gravity is located at O’. Let x’, y’, and z’ be the coordinate axes passing through
its center of gravity. The coordinate axes x, y, and z are parallel, respectively,
to the x’, y’, and z’ axes. Let the coordinates of O’ in the x-y-z coordinate
system be x, y, and z. If we consider an elemental mass dm in the body, then
from Equations (9.34), (9.35), and (9.36) the mass moments of inertia can be
given as

[= |(y? + 27) dm (9.38)

Ub |(x? + 2”) dm (9.39)

= (x? + y*) dm (9.40)

Figure 9.22
502 Chapter 9 Moment of Inertia

However, the coordinates are related as

pee ge le 5 (9.41)
y=y+y (9.42)
Z=2 +7 (9.43)

Substituting Equations (9.42) and (9.43) into Equation (9.38), we obtain

Le |[(y' +9? + @ +27] dm


= [2+ 2%) dm + |GP +P) dm-+ 25 Sy! dm + 2 2 dm
Note that

|(y'? + 2'*) dm=I,


= moment of inertia with respect to the x’ axis

[62+ 2 am=( +2 | dm=mge+%) = md?


= m(square of the perpendicular distance between the x
and x’ axes)

iy'dm=0

z dm=0

Therefore,

l= f+ made (9.44)

In a similar manner, we can obtain

1, = 1, + md? (9.45)

I, = 1, + md? (9.46)
where d, = perpendicular distance between the y and the y’ axes
d, = perpendicular distance between the z and the z’ axes
So, in general, we can state that the mass moment of inertia I about a given
axis is equal to the sum of the mass moment of inertia of a parallel axis passing
through the center of gravity of the body (Icg), and the product of the mass
(m) and the perpendicular distance between the axes under consideration (d),
9.12 Mass Moment of Inertia by Integration 503

or

1 = Icg + md? (9.47)


This is similar in form to Equation (9.12).

9.12 MASS MOMENT OF INERTIA BY INTEGRATION


In this section we will consider only symmetrical bodies generated by revolv-
ing acurve around an axis. For example, let us consider a right circular cylinder
that is symmetrical about the y axis (Figure 9.23). The cylinder has a height
H and an outside radius R. In this case, a shell element may be used for
determining the mass moment of inertia of the body. By doing so, a single
integration will be required.
Consider an elemental cylindrical shell having a thickness dr as shown in
Figure 9.23. For this shell element
dm = (2tr dr)H p

where p represents the density of the material. Considering the density to be


constant, we obtain
R
R4 HpR 4
1, = Ir2 dm = I (2ar? Hp) dr = 20H (“) == - (9.48)
However, the mass of the cylinder is m = wR7Hp; therefore,
Figure 9.23
2 R? mR?
if = (TR HOS = 5 (9.49)

In some cases a disk element may be considered for determination of the


mass moment of inertia of a body. This can be explained with reference to
Figure 9.24 in which a cone is shown. If it is required to determine the mass
moment of inertia of the cone about the y axis by single integration, we can
consider an elemental disk having a thickness of dy. The moment of inertia of
this elemental disk about the y axis can be obtained from the results of the
moment of inertia for a cylinder derived above by substituting dy for the height
of the cylinder in Equation (9.48). Thus,
4
Tpr
a 5 (dy)

However, for the cone, r = (R/h)y; therefore,

Figure 9.24
504 Chapter 9 Moment of Inertia

The mass of the cone is m = 3m7pR7h; therefore,

[ies ae
58)

EXAMPLE 9.21

Determine the mass moment of inertia of a thin circular plate of


radius R about the x, y, and z axes.

Solution For the moment of inertia, /,, consider an elemental


volume dV having a mass of dm with a cross-sectional area
of dA and a thickness t as shown.
x
I= [2 dm
dV =Volume
dm = pt dA

1. |pre?dA = pt |2 dA
The term f z? dA is actually the moment of inertia of the
circular area about the x axis. From Table 9.1, we can obtain

: aR*
Zab = ray a dees

Thus,
a7R*
Ie = ptl (area) faa pt VAs.

However, the mass of the disk is m = TR? tp; therefore,

[=
x 4 =<

For the moment of inertia /,, as in the case of J, we consider


the elemental volume shown.
R*
= [x dm = pr |x2 dA = o(=)

mR*
pt Focaieay — ae <

In a similar manner, we obtain

1,=] + 2°) dm= pr |«2 +27)dA

= n([ dA + [2 is)= pt(l, + 1.)


mR2

ie = oa <
9.12 Mass Moment of Inertia by Integration 505

EXAMPLE 9.22

A rectangular prism is shown. Determine the mass moment of y


inertia of the prism about the x axis. Note that the origin O is the
center of gravity of the body.

Solution Let us consider a thin disk having a thickness of dx.


As shown in the preceding example, the moment of inertia of
this disk about the x axis can be written as
dl, = p dx Ix(area)

i t
Density Thickness Moment of inertia
of the of the of the area of the disk
material plate about the x axis

It is important to note that J\(area) is actually the polar moment


of inertia of the area of the disk about the axis passing through
its centroid, and

1
Ix(area) = —bd(b*
19a + d?)

So we may write
ser I
x |dl,={
1,= OA eu
—bdp(b? + d*)SALI
dx = —bdlp(b* + d? )

The mass of the rectangular prism is m = bdlp; therefore,

1
=n
Bal bt) )

Note: In a similar manner we can determine

1 1
l= nb
Ks Ts ) and /, = ne
—m(d> ==
+ (>,
17
506 Chapter 9 Moment of Inertia

EXAMPLE 9.23

y y Using the results of Example 9.22, determine the moment of in-


ertia of the rectangular prism about the y axis shown.

Solution Consider the elemental disk of thickness dx as


shown. For this plate,
x,Xx dl, = (p)(dx)[1,(areay] + (dm)(x?)
1
= oxan( 5a) + [(p)(b)(d)(dx)]x?

However, the mass of the prism is m = pbdi; therefore,

ie fF
I= m\ = = aX
: yr 3}

Note: This problem can also be solved by using the parallel


axis theorem. Equation (9.45) gives J, = I, + md;, where J,
is the centroidal mass moment of inertia. From Example 9.22,

ly =m? + 2
y aig )
Also, d, = 1/2. Therefore,
2
1 I
= pine +177) + m(s)
, 9.12 Mass Moment of Inertia by Integration 507

EXAMPLE 9.24

A solid is formed by revolving the shaded area OAB shown about y


the x axis. Determine the mass moment of inertia of the solid with
respect to the x axis.

Solution We consider an elemental disk of thickness dx and


radius r. The mass of the elemental disk is written

dm = mr?dx p
However, r = Vx; therefore,

dm = w(V x)? p dx = mxp dx


The moment of inertia of the elemental disk about the x axis
(see Example 9.21) is

dl, = (3)amr? = Sp dV)? = S29 de


Thus,
L L
1 ae \ ee mLp
L=/d =| —ax7p dx =| — —} =
; : 9 Ns i (0)]3], 6

The mass of the solid is written

Therefore,
508 Chapter 9 Moment of Inertia

9.13 MASS MOMENT OF INERTIA OF COMPOSITE BODIES


A composite body consists of a number of connected bodies of simple shapes
such as the one shown in Figure 9.25. In order to determine the mass moments
of inertia /,, /,, and /, of a composite body with respect to the x, y, and z axes,
the following procedure can be used.
1. Divide the composite body into several parts with common shapes, such
as; [N25 Oe
Calculate the mass of each part as m,, m), m3,....
3. Locate the coordinates for the center of gravity of each part as
(%Yi> 21), Xa»
YasZ0)s (30 Y323)s +++
4. Calculate the moments of inertia of each part about the axes passing
through their centers of gravity that are parallel to the x, y, and z axes.
Figure 9.25
Let these be /,,, 1,,, 1,; for part 1; I,,, 1y., 1,, for part 2; [,;, 1, I, for
Pali oseee
5. Calculate the moments of inertia of each part with respect to the x, y,
and z axes by using the parallel axis theorem as

= 2
Lay
x = Lew + Mida
De et md) (9.50)

Lai)= Law tM; ide)

where dno= (Vy) (2)


do == (x;) sis C4 ie

di) == (x;i a Oy)

p= Nee Ss oc

6. Calculate the moments of inertia of the composite body as

L= 2h
L, = 2lyy (9.51)

Table 9.6 gives the mass moments of inertia of some bodies having simple
geometric shapes.
Table 9.6 Mass moment of inertia of some bodies with simple geometric
shapes (with respect to the axes passing through their centers of gravity)

Shape Mass moment of inertia


Uniform slender rod
ip pies Je 2
y 12 mL

y=
re 4 mL mL?
12

Circular cylinder

id
y IG = x mR?

I =I = 4 mGR? +L?)

Thin plate

509
510 Chapter 9 Moment of Inertia

EXAMPLE 9.25

Two cylinders are attached to a disk. The disk and the cylinders
are made of aluminum having a specific weight of 170 Ib/ft*.
Given: R = 12 in.,t = 3 in.,d = 3 in., andL = 6 in. Calculate
Meox? ths,Ay? Brae Ly

Solution

Step 1. The body has been divided into three parts as shown.

Step 2. The masses of the parts are as follows. Note that g =


32.2 ft/sec”, which is acceleration due to gravity.

Part I (disk):

- a( LI0\ = (m)(12)°(3) LO Ve Re
m, = (1R o(2 = jeu 30.2) ~ oy sec?/ft

Part 2 (cylinder):

_ (md*L\(170\ _ |(m3) | (170) _ ;


ior ( 4 )(g 7 2p le ck
Part 3 (cylinder): m, = 0.13 lb - sec?/ft

Step 3. The coordinates of the center of gravity of the parts


are as follows:

x; Ji Z
Part (in.) (in.) (in.)
1 0 0 0
2; 4.5 10.5 0
3 —4.5 — 10.5 0

Step 4. Calculation of I.¢i). Ly, and 1,


Part 1: Consider this as a thin disk. Referring to Example 9.21,
2
1 1 12
I. =—-m,R2
ni =5 =({-
(3)« i9(2) =.2.075 lb -- ft
? —| ft -
- sec sec2

1
I,1 = I,,ay = 4
—m,R? = 1.038 Ib -ft - sec?

(continued)
9.14 Summary 511

EXAMPLE 9.25 (concluded)

Part 2: I, =I, + mG)? + (%)?]


2 2 2 2
1
eee
x |d |
vm($)
1
\o13 (= a
(3)
3
EO0.001
O0lIbe ho -sec sec?
lb - ft = 0.0032 + 0.1
4.5
—— |
0

= 0.02148 lb - ft - sec?

2 2 Lee lbpes ma[(X)? + (¥2)"]


1 3 6
013)
(4) f«(= 5)
== | (3) |
2 2

= 0.0032 + os| (42) + (223) |


= 0.0032
lb -ft -sec?
= 0.12101 Ib - ft - sec?
Part 3:
Part 3: Same as Part 2,
[= 0.001 lb - ft - sec?
I, = 0.10053 Ib - ft - sec?
1, = I, = 0.0032 Ib - ft - sec?
I,, = 0.02148 Ib - ft - sec*
Step 5. Calculation of 1.;), Lyi, and I,¢)
I,, = 0.12101 Ib - ft - sec®
Part I: Step 6.
I, = 2.075 Ib -ft - sec? 1, = 1, + Ie, + Ie, = 2.075 + 0.10053 + 0.10053
I,, = 1, = 1.038 Ib - ft - sec? I, = 2.2761b - ft. sec? <
Part 2: I, =1,, + 1,, +1), = 1.038 + 0.02148 + 0.02148
Le = ‘tp oF m[(V2)” ar (Z,)"]
I, = 1.081 Ib - ft- sec? <

10.5
2
0
2 I, =1,, + 1, + 1, = 1.038 + 0.12101 + 0.12101
= 0.001+ 0.13] (——) +(— L, = 1.280Ib- ft-sec? <
= 0.10053Ib - ft - sec?

9.14 SUMMARY
In this chapter we have developed the principles for calculating the moments
of inertia of areas and masses, and have outlined the following:

1. Moment of inertia of an area with respect to the x and y axes are


expressed as (Section 9.2)

I= |y? dA (o.1)

=
=] 2 dA (9.2)
512 Chapter 9 Moment of Inertia

Nw The radius of gyration of an area with respect to the x and y axes are
given as (Section 9.3)
I.
k= |= 9.4

Kgy == |a
A (9.5
: )

The polar moment of inertia and the radius of gyration for areas are of
the form (Section 9.4)

Jo= ir2 dA (9.6)

ko = ie (9.8)
According to the parallel axis theorem (Section 9.5), for the moment of
inertia of areas ;
I =Ic¢g + Ad? (9.12)
The moment of inertia of a composite area with respect to a given axis
can be determined by summing the moments of inertia of component
parts about the same axis (Section 9.6).
The product of inertia of an area with respect to the x and y axes can
be expressed as (Section 9.7)

i xy dA (9.14)
The product of inertia of a composite area with respect to a given set
of rectangular axes can be obtained by summing the products of inertia
of the component parts about the same axis (Section 9.8).
The principal moments of inertia of an area can be expressed as
(Section 9.9)

y Lae, Lp?
rn= ~ + (* :)+12 (9.26)

The principal moments of inertia and the orientation of the principal


axes can be determined by using Mohr’s circle (Section 9.10).
10. The mass moments of inertia and the radii of gyration of a body with
respect to the x, y, and z axes are as follows (Section 9.11):

With respect to the x axis,

I= |(y? + 22) dm (9.35)


1
k= |2 (9.37)
m
Key Terms 513

With respect to the y axis,

ee |(x? + 2°) dm (9.34)

a—
y
_
m
(9337)

With respect to the z axis,

=| 2+ y?)dm (9.36)

l,
|
ST ee (O31)
m
ine According to the parallel axis theorem, the mass moment of inertia of a
body about a given axis is (Section 9.11)
1 =Icg + md? (9.47)
WP, The mass moment of inertia of a composite body about a given axis
can be calculated by summing the moments of inertia of the individual
parts of the body about the same axis.

KEY TERMS

Mohr’s circle 496 polar moment of inertia 473


moment of inertia of a mass 465 polar radius of gyration 473
moment of inertia of an area 463 principle moment of inertia
parallel axis theorem for moment of an area 492
of inertia of a mass 50/ product of inertia of an area 456
parallel axis theorem for moment radius of gyration of a mass 500
of inertia of an area 475 radius of gyration of an area 470
514 Chapter 9 Moment of Inertia

PROBLEMS
LL a a a ac

SECTION 9.2
9.1 through 9.12 Determine the moment of inertia of the shaded area with respect
to the x axis. Use the integration method.

9.13 through 9.24 Determine the moment of inertia of the shaded area with respect
to the y axis. Use the integration method.

SECTION 9.3
9.25 through 9.36 Determine the radii of gyration about the x and y axes for the
shaded area shown.

Figure P9.1, P9.13, and P9.25 Figure P9.2, P9.14, and P9.26 Figure P9.3, P9.15, and P9.27

1
Figure P9.4, P9.16, and P9.28 Figure P9.5, P9.17, and P9.29 Figure P9.6, P9.18, and P9.30
Problems 515

0 a
Ss Opp
Figure P9.7, P9.19, and P9.31 Figure P9.8, P9.20, and P9.32 Figure P9.9, P9.21, and P9.33

— O 3 in. Sin. ; O Dien 4 in.


Figure P9.10, P9.22, and P9.34 Figure P9.11, P9.23, and P9.35 Figure P9.12, P9.24, and P9.36

SECTION 9.4

9.37 through 9.41 For the shaded area shown, determine (a) the polar moment of
inertia with respect to the origin O, and (b) the polar radius of gyration, kp. Use the
integration method.

2 ft

0.04m
™y = 2x?

0.02 m

O O 0.02m 0.04m ss O -
Figure P9.37 { Figure P9.38 Figure P9.39
516 Chapter 9 Moment of Inertia

O
Figure P9.40 Figure P9.41

SECTION 9.5
9.42 through 9.53 Using the parallel axis theorem, determine the centroidal
moments of inertia /,, and Ly of the shaded area shown.

p=
Figure P9.42 Figure P9.43 Figure P9.44

Figure P9.45 Figure P9.46 Figure P9.47


Problems 517

y=(k—V
x)"

_—_— = x
O
Figure P9.48 Figure P9.49

ee

Figure P9.50 Figure P9.51


518 Chapter 9 Moment of Inertia

SECTION 9.6
9.54 through 9.57 For the shaded composite area shown, determine the moment of
inertia and the radius of gyration with respect to the x axis (use Table 9.1).

9.58 through 9.61 For the shaded composite area shown, determine the moment of
inertia and the radius of gyration with respect to the y axis (use Table 9.1).

be 100m

O AR
Brine
Figure P9.54 and P9.58 Figure P9.55 and P9.59

O O1+—200 mm —+
Figure P9.56 and P9.60 Figure P9.57 and P9.61

9.62 through 9.67 For the shaded area shown, determine the moment of inertia with
respect to the x’ axis (use Table 9.1). Note that O’ is the centroid of the composite
area.

9.68 through 9.73 For the shaded area shown, determine the moment of inertia with
respect to the y’ axis (use Table 9.1). Note that O’ is the centroid of the composite
area.
Problems 519

iW<= 200'mm—
25 mm

280mm 200 mm

| | 25 mm -—8 in ¥ +160 mm——> ¥


Figure P9.62 and P9.68 Figure P9.63 and P9.69 Figure P9.64 and P9.70

: 6 in.
}<— 8 in.——>| {1
lin

1X
Spine

lin.
+ 10 in ———+ } en
Figure P9.65 and P9.71 Figure P9.66 and P9.72 Figure P9.67 and P9.73

9.74 For the shaded area shown, given A = 25.2 in, Jy,y,= 820 in*, d, = 6 in.,
and d, = 10 in., determine Jy. Note that O’ is the centroid of the area.
U
y

Figure P9.74
520 Chapter 9 Moment of Inertia

9.75 The moment of inertia of the triangle with respect to the x axis is 512 in*.
Determine its moment of inertia with respect to the xx, axis.

9.76 through 9.79 The cross section of a beam is shown. Determine the moments
of inertia /, and J, of the composite section of the beam. Note that O is the centroid
of the composite beam section.

ve
|

111 in.—+ |
| In.
x x}
Figure P9.75

Figure P9.76 Figure P9.77

Figure P9.78 Figure P9.79

SECTION 9.7

9.80 through 9.85 Determine the product of inertia /,,, of the shaded area shown.
Use the integration method. Use the parallel axis theorem to determine the product
of inertia /,.,,, with respect to the centroidal axes x’, y’.

Figure P9.80 Figure P9.81


Problems 521

O SS tn
Figure P9.82 Figure P9.83

a
Figure P9.84 Figure P9.85

SECTION 9.8
9.86 through 9.91 Use the parallel axis theorem to determine the product of inertia
Ty: of the area shown. Note that x’ and y’ are the centroidal axes.

SECTION 9.9
9.92 through 9.97 For the section shown, determine the principal moments of
inertia and the orientation of the corresponding principal axes passing through the
centroid O'.

20 mm

Figure P9.86 and P9.92 Figure P9.87 and P9.93 Figure P9.88 and P9.94
522 Chapter 9 Moment of Inertia

}-— 6 in.—+

3 in.
ae

}+———200 mm———+ pea 18 in }~+—_______——


18 in.—_____———_+
Figure P9.89 and P9.95 Figure P9.90 and P9.96 : Figure P9.91 and P9.97

SECTION 9.10
9.98 through 9.103 Using Mohr’s circle, determine the principal moments of inertia
and the orientation of the corresponding principal axes passing through the centroid
O' of this section.
!
Vy
i 20 mm
foes 200 mm—+
25 mm

— ee =!

280 mm

>| }-25 mm in. }+—— 160 mm——+>


Figure P9.98 Figure P9.99 Figure P9.100

6 in.

— 1x’

15 in.

—— a

Figure P9.101 Figure P9.102 Figure P9.103


Problems 523

SECTIONS 9.11 and 9.12


9.104 through 9.108 A solid is formed by revolving the shaded area shown about
the x axis. Determine the mass moment of inertia of the solid with respect to the
x axis in terms of mass m of the solid. Use the direct integration method.

9.109 through 9.113 A solid is formed by revolving the shaded area shown about
the x axis. Determine the mass moment of inertia of the solid with respect to the
y axis (in terms of mass m of the solid).

Figure P9.104 and P9.109 Figure P9.105 and P9.110 Figure P9.106 and P9.111

Figure P9.107 and P9.112 Figure P9.108 and P9.113


524 Chapter 9 Moment of Inertia

SECTION 9.13
9.114 Determine the mass moment of inertia of the assembly with respect to
(a) x axis, (b) y axis, and (c) z axis. The assembly is made of steel, and the specific
weight of steel = 480 lb/ft’.
9.115 For the hollow cylinder shown, determine the mass moment of inertia with
respect to the x axis. Note: The hollow cylinder is made of brass, whose specific
weight is 530 lb/ft’.

Diameter= 2 in.

N
1S in.

Diameter= 3 in.

Figure P9.114 Figure P9.115

9.116 Determine the mass moment of inertia J, of the cone in terms of its mass m.

9.117 The frustum of a cone has a cylindrical hole. Determine the mass moment of
inertia /,. The density of the material is 3000 kg/m°.

Radius = 100 mm

50-mm diameter

200-mm radius

Figure P9.116 Figure P9.117


Problems 525

9.118 The assembly shown is made of steel (specific weight = 480 Ib/ft?).
Determine the mass moment of inertia of the assembly with respect to the y axis.

9.119 A rectangular prism has a cylindrical hole drilled through it. The diameter
of the hole is 2 in. Determine /, and /,. Given: specific weight of the material =
250 Ib/ft?.

9.120 For the rectangular prism described in Problem 9.119, determine the mass
moment of inertia about the vertical y axis.

4
6-in. diameter

6-in. diameter

Figure P9.118 Figure P9.119 and P9.120

9.121 For the assembly shown, determine the mass moment of inertia about the x
axis. Specific weight of the material is 300 Ib/ft’.

#12
in. —>

20 in.

Figure P9.121
526 Chapter 9 Moment of Inertia

9.122 A wheel consists of a thin ring and six spokes. The mass of the ring is ™,,
and the mass of each spoke is m,. Determine the mass moment of inertia of the
wheel about an axis passing through O and perpendicular to the cross section shown.

9.123 Determine the mass moment of inertia of the wheel about an axis passing
through point O and perpendicular to the cross section shown. Specific weight of the
material is 400 Ib/ft?.

Figure P9.122 Figure P9.123


CHAPTER

WORK AND ENERGY*

10.1 Introduction OUTLINE


10.2 Basic Concept of Work
10.3 The Principle of Virtual Work
10.4 Procedure for Analysis
10.5 Potential Energy
10.6 Equilibrium—Principle of Stationary Potential Energy
10.7 Stability of Equilibrium—Principle of Minimum Potential Energy
10.8 Summary
Key Terms
Problems

527
528 Chapter 10 Work and Energy

10.1 INTRODUCTION
In the previous chapters, we developed the methods for analysis of equilibrium
of particles, rigid bodies, and structures composed of rigid bodies, based on
the equations of equilibrium. In this chapter, we will introduce the concept of
work, and develop the analysis of equilibrium based on the principle of virtual
work. While this alternate method of analysis can be used to determine the
unknown forces and reactions acting on particles and single rigid bodies, it is
especially useful for the analysis of structural and mechanical systems com-
posed of multiple rigid bodies that are partially constrained and/or nonrigid
when detached from their supports. By applying the principle of virtual work,
the unknown external forces and reactions and the equilibrium configuration
of such structures can be determined directly, without disassembling the struc-
ture as is necessary when the analysis is based on the equations of equilibrium.
Another important concept, potential energy, will be introduced in this
chapter, and the principle of virtual work will be restated in terms of the
potential energy of the system. This potential-energy-based approach offers an
additional advantage: it not only enables us to determine the configurations in
which a structural/mechanical system may be in equilibrium under a given
system of forces, but also whether an equilibrium configuration is stable or not.
Our discussion will be confined, in this chapter, to those systems in which
the friction forces are sufficiently small to be neglected in the analysis. The
formulations presented here can be generalized to include frictional effects,
but in order to include the effect of internal friction at joints or connections,
for example, it is usually necessary to dismember the structure to determine
the friction forces. Therefore, for these structures, the concepts of work and
energy may not offer any advantage over the conventional analysis using the
equilibrium equations.
The objective of this chapter is to develop the analysis of the equilibrium
of bodies based on the concepts of work and potential energy.

10.2 BASIC CONCEPT OF WORK


Work of a Force

The work done by a force on a body is simply defined as the force times the
displacement of the body in the direction of the force. Consider a force F of
constant magnitude and direction acting on a body as it moves from position
A to B on a smooth surface as shown in Figure 10.1. The movement, or change
in position, of the body is called the displacement and is shown as §S in the
figure. The work done by the force F on the body is given by

S cos dey Smooth
surface U = F(S cos 8) (10.1)
U=FS cos 6 where U is the work done, and @ is the angle between the lines of action of F
Figure 10.1 and S.
10.2 Basic Concept of Work 529

The term S cos 0 is the magnitude of the component of the displacement S


in the direction of the force. From Equation (10.1), we can see that the work
of a force may also be considered as the product of the component of the force
in the direction of the displacement (i.e., F cos 0) and the displacement S.
Work is a scalar quantity. It is considered positive when the force, and the
displacement component in the direction of the force, have the same sense;
and negative when the force and the displacement component have the op-
posite sense. From Figure 10.1, we can see that if the angle between the force
and the displacement is acute (8 < 90°), the work is positive; whereas, if the
angle is obtuse (8 > 90°), the work is negative.
Three special cases of work are shown in Figure 10.2. As shown in Figure
10.2a, when the force F and the displacement S have the same direction
(8 = 0°), the work is positive and has the magnitude equal to the product of
the magnitudes of the force and the displacement, or U = FS. When the force
F is perpendicular to the displacement S (8 = 90°), as shown in Figure 10.2b,
the work is zero (U = Q). In the third case, as shown in Figure 10.2c, when
the force F is in the direction opposite to the displacement S (8 = 180°), the
work is negative with the magnitude equal to the product of the magnitudes
of FandS (U = —FS5S).
It is important to realize that the definition of work does not require that
the displacement used to determine the work of a force be necessarily caused
by the force itself. The displacement S may be due to the force F or it may be
caused by any other action independent of F. The only requirement is that F
must act on the body while it undergoes the displacement S.
In our discussion of work thus far, we have only considered the case in
which the magnitude and direction of the force, as well as the direction of the
displacement, remain constant. To derive the general expression of work for
the case when the force and/or displacement may vary, consider a force F
acting on a particle that moves along an arbitrary path as shown in Figure
4
10.3. The work dU that the force F performs as the particle undergoes an
infinitesimal displacement, dS, from a position A to a neighboring position B,
can be written as

dU = F (dS)(cos 8) (10.2) Figure 10.2

By using the definition of the dot product of two vectors (Section 4.3), we can
rewrite Equation (10.2) as

dU =F -dS (10.3)

However,

F=Fi+ Fj + Fk (Pe
Ce
;
7
~

=
S
ies &
Fafa
tl~
N

dS = dxi + dyj + dzk Figure 10.3


530 Chapter 10 Work and Energy

Substitution of the above expressions into Equation (10.3) yields

dU = (Fi + Fj + Fk): (dui + dyj + dzk)


Using the distributive property of dot products, and the dot products of unit
vectors (Section 4.3), we obtain

dU = F,.dx + F,dy + F,dz (10.4)


As the particle undergoes a finite displacement from position C to position
D (Figure 10.3), the total work U performed by the force F is obtained by
integrating the expression of dU along the path of the particle between C and
D, or
D

U= | dU
Cc

By using Equations (10.2) to (10.4), the total work can be expressed as


D
U= i.F cos @dS OS)

D
U= I.F-dsS (10.6)

D
U= I.(Fo dx FF, dy F, dz) (10.7)

The dimensions of the work are length times force. The commonly used
units of work are foot-pound (ft - 1b) or inch-pound (in : Ib) in the U.S. Cus-
tomary system, and N - m or joule (J = N - m) in the SI system.

Work of Internal Forces

The internal forces acting on the particles of rigid bodies, as well as those
acting at the connections of structural or mechanical systems, always occur in
pairs of equal but opposite collinear forces, in accordance with Newton’s third
law. Also, a rigid body does not deform or change dimensions: the distances
between the particles of a rigid body remain the same. Therefore, during any
displacement, the work done by one of the pair of internal forces is canceled
by the work of the other force. Thus, the total work of the internal forces
is zero.

Work of a Couple
Consider a couple of equal and opposite forces, — F and F acting at points A
and B, respectively, on a rigid body as shown in Figure 10.4a. The perpen-
10.2 Basic Concept of Work

Total displacement e) Translation (+) Rotation


(a) (b) (c)
Figure 10.4

dicular distance between their parallel lines of action is a. As the body under-
goes an infinitesimal displacement in the plane of the forces, points A and B
displace into positions A, and B,, respectively, as shown in the figure. This
general displacement may be considered to occur in two distinct steps. In the
first step, both points A and B (as well as the whole body) undergo the same
displacement dS, thereby bringing the point A into its final displaced position
A,, and moving point B into an intermediate position B’ (Figure 10.4). In the
second step, the body is rotated about A,, until the point B moves from its
intermediate position B’ into its final position B, as shown in Figure 10.4c.
The first type of movement in which all the points of the body displace equally
is called the translation. The work done by a couple during a translation is
always zero. It can be seen from Figure 10.46 that since both points A and B
undergo equal displacements, dS, the work of the force F is canceled by the
work of the force —F.
During the second part of the displacement, the force — F does not perform
any work as its point of application A, does not displace. The work done by
the force F, as its point of application displaces from B’ to By, is
dU = F (dSp,) = Fa(d8)
where d@ = the infinitesimal angle of rotation (Figure 10.4c).
However, Fa = M = magnitude of the couple. So,
dU =M d6 (10.8)
The total work performed by a couple on a body, as the body undergoes a
finite rotation from position 8, to position 8, in the plane of the couple, can
be obtained by integrating Equation (10.8) between positions 8, and 9,, or
82
U= F M a0 (10.9)
1

The work of a couple is considered positive if the moment of the couple and
the rotation have the same sense, and negative when the moment and the
532 Chapter 10 Work and Energy

rotation have the opposite sense. Since the rotation is measured in radians, the
dimensions of the work of a couple are the same as those of the work of a
force, that is, length times force.

10.3 THE PRINCIPLE OF VIRTUAL WORK


The principle of virtual work, initially formulated by Johann Bernoulli
(1667-1748), can be stated as follows:

STATEMENT | If a particle, a rigid body, or a structural system composed of multiple rigid


bodies is in equilibrium under the action of a system of forces and couples,
and is subjected to any small virtual displacement, the total virtual work
done by the external forces and couples is zero.

In the above statement, the term virtual displacement implies imaginary (as-
sumed) displacement that does not really take place.
We will now apply the above principle to the case of a particle in equilib-
rium at a position A under the action of an arbitrary system of forces as shown
in Figure 10.5. Imagine that the particle is given a small virtual displacement
5S from its equilibrium position A to another position B. Note that the small
virtual displacement is denoted by 8S to distinguish it from the real displace-
ment dS. As the particle undergoes the virtual displacement, each of the forces
acting on it does work, which is called the virtual work. The total virtual work
Figure 10.5 dU of all the forces acting on the particle is

d8U = X(F-8S) = XF-5S =0 (10.10)


In this equation, since the virtual displacement 8S is not zero, the total virtual
work can be zero only if }F = 0; that is, if the particle is in equilibrium.
Equation (10.10) can be rewritten as

8U = (SF,i+ TF,j + SF.K):(8xi + 8yj + 82k) =0

8U = XF, 8x + SF, by + TF 8z = 0 (10.11)


where dx, dy, and 8z are magnitudes of the components of the virtual displace-
ment dS in the x, y, and z directions, respectively.
Equations (10.10) and (19.11), which express the principle of virtual work
for particles, represent the necessary and sufficient conditions for the equilib-
rium of particles. It can be seen from Equation (10.11) that the condition of
zero virtual work can only be satisfied for nonzero virtual displacement com-
ponents if the equations of equilibrium of the particle are satisfied (XF,, = 0,
XF, = 0, and &F, = 0).
10.3. The Principle of Virtual Work 533

Now let us consider the case of a rigid body in equilibrium under the action
of a coplanar system of forces and couples as shown in Figure 10.6. Since any
small virtual displacement of the body can be decomposed into a translation
5S and a rotation 58, the total virtual work 6U done by all the forces and the
couples acting on the body can be expressed as the sum of the virtual work
done during translation, 5U,, and the virtual work done during rotation, 5U,.
Thus, 5U = 6U, + SU.. (As stated previously, the total virtual work of the
internal forces 1s zero.)
During a virtual translation 5S = 6xi + dyj of the body, the virtual work
done by all the forces is given by

8U, = X(F-8S) = XF,8x + ZF,8y = 0 (10.12)


Figure 10.6
Recall that the couples do not perform any work during translation. As the
body is in equilibrium, {F, = 0 and {F, = 0. Hence, 6U, = 0.
The virtual work done by the force F, during a small virtual rotation 50 of
the body about any arbitrary point C can be expressed as (see Figure 10.6)

6U,, = F, Sr, = (F,)(6r,) cos a

where 67, is the displacement of the point of application A of F,, due to rota-
tion 60.
As 66 is infinitesimal, 5r, can be considered to be tangent to the circular
path at A, in the direction perpendicular to r, (from C to A). Substituting
dr, = 7,(58) into the expression for 6U,.,, we obtain

oU,, = F\r,80 cos a = (Fir, cos a) 60 = M,50

where M, = Fr, cos a is the moment of F, about the center of rotation C.


The total virtual work done by all the forces and the couples acting on the
body during a virtual rotation 60 is therefore given by

8U,
= &(M 80) = M50 = 0 (10.13)

where >M is the sum of the moments of all the forces about C, plus the
moments of all the couples acting on the body.
As the body is in equilibrium, }M = 0 regardless of where point C is
located in the plane of the body, and therefore, 5U, = 0. Combining Equations
(10.12) and (10.13), we can express the principle of virtual work for rigid
bodies as

5U = 8U,+8U,=0 (10.14)
The principle of virtual work as expressed by Equation (10.14) remains
valid for systems composed of multiple rigid bodies in equilibrium provided
that the connections are frictionless, that is, the work done by internal forces
at the connections is zero.
534 Chapter 10 Work and Energy

10.4 PROCEDURE FOR ANALYSIS


The following step-by-step procedure can be used to determine the reactions
at the supports, and/or the equilibrium configurations, of rigid bodies by ap-
plying the principle of virtual work.
1. Draw a free-body diagram of the structure showing all the external (ap-
plied and reaction) forces and couples acting on it.
2. Apply a virtual displacement or rotation so that only one of the unknown
forces and couples may perform virtual work. The virtual displacements of the
remaining unknown forces (and hence, their virtual work) must be zero. Draw
the displaced configuration of the structure on its free-body diagram using
dashed lines. As an example, consider the beam in Figure 10.7a, composed of
y two rigid bodies AB and BD connected together by a frictionless hinge at B.
The patterns of virtual displacement that can be used to determine the reactions
A,, C,, and D, are shown in Figures 10.7b to 10.7d, respectively. Note that
for each virtual-displacement pattern, only one of the unknown forces under-
goes virtual displacement, and hence performs virtual work. The virtual dis-
placements (and therefore the virtual work) of the remaining unknowns are
zero.
3. Write the equation of virtual work (6U = 0) consisting of the total virtual
work done by all the known forces and couples as well as the unknown force
or couple.
4. Express each virtual displacement or rotation in the virtual work equation
in terms of a common virtual displacement or rotation. In some structures, this
may be possible directly from the geometry of the displaced structure. In oth-
ers, it may be necessary to first relate the positions of the forces that perform
work to a common variable, and then differentiate these relationships to ex-
press the virtual displacements of the forces in terms of the virtual movement
of the common variable. This is illustrated in Examples 10.2 and 10.3.
5. As the virtual-work equation now contains only one common virtual dis-
placement, which can be factored out, it can be solved for the desired unknown.
(d) Virtual displacement for determining Dy,
Figure 10.7
10.4 Procedure for Analysis 535

EXAMPLE 10.1

Determine the reaction at the support B for the simply supported


beam shown, using the principle of virtual work.

Solution The free-body diagram of the beam is shown, in


which the three unknown reaction components are A,, A,,
and B,,.

Virtual displacement. We choose the virtual displaced


configuration of the beam so that only the desired unknown
B, performs the work, while the work of the remaining
two unknowns, A, and A,, is zero. As shown in the figure, this
is accomplished by giving point B a small vertical displacement, Displaced
configuration C’
while point A is kept in its original position. The point of i

application of B, displaces by a distance dy, from B to a new


location B’. Similarly, the point of application of the 12-kN ‘Original G
force displaces by a distance dyc from C to C’. Note that for Ay configuration B,,
the two remaining unknowns, the point of application A
ae 6m 4 m—+|
does not undergo any virtual displacement, and hence, their
virtual work is zero.

Equation of virtual work. We can write the equation of virtual


work as
6U = —128y¢ + B, dyz = 0
Note that since dy, has the same sense as B, (vertically
upward), the virtual work of B, is positive, whereas dyc has the
sense opposite to that of the 12-kKN force and, therefore, the
12-kKN force performs negative work.
Next, we express the two virtual displacements dy, and dyc
in terms of a common virtual displacement. From the geometry
of the displaced beam, we observe that the triangles ABB’ and
ACC’ are similar. Thus, The positive sign for B, indicates that its sense is in the vertical
upward direction as initially assumed on the free-body
WE
6
OF
10 dy,Yo = 0.66 Ye diagram. Therefore,
B,=72kN() <
By substituting this relationship into the equation of virtual
work, we obtain Note: In the above computations, dy, was used as the common
virtual displacement. We could have instead used dy, (with
dU = —12 (0.6) dy, + B, dyz = 0 dyz = 1.67 dy-) or 68 (with dyz = 1056 and dy, = 688) as
(7:2 + B,) oy, = 0 the common virtual displacements.
Equation (a) obtained herein by applying the principle of
As dyz # 0, we have virtual work could have been obtained directly by simply using
12+ B,=0 (a) the moment equilibrium equation about A, 2M, = 0. As this
example indicates, the principle of virtual work offers no
or
computational advantage over the conventional analysis using
B, = 7.2KN the equilibrium equations for the single rigid bodies.
EXAMPLE 10.2

For the lifting platform shown, establish a relationship between


the weight Q and the moment of the couple M required to main-
tain equilibrium. Use the principle of virtual work. Also, deter-
mine the angle of inclination a. Let L = 12 ft, Q = 5 kip, and
M = 38 kip - ft.

Solution The free-body diagram of the frame shows the


known force Q and the four unknown reactions, Fp, B,, By,
and M.

Virtual displacement. The virtual displaced configuration of


the frame is obtained by giving it a small counterclockwise
virtual rotation 5a about points A and B. As the lengths of the
members AD and BE are equal, member DE remains horizontal
and translates vertically upward. Note that only the weight
Q and the couple M perform virtual work. As the virtual
displacements of the three remaining unknowns are zero, they
perform no virtual work.

Equation of virtual work. We can write

6U = —Q(8y,) + M(da) = 0

In order to relate dy, to da, we first express the coordinate y.


of point C, which is the point of application of Q, as a function
of the angle a, and then differentiate this expression to obtain
the relationship between dy, and da. We see from the figure
that
yco=Lsina
Differentiating this expression, we obtain

dyc = L cos a 6a
By substituting the above relationships into the equation of
virtual work, we obtain
6U = —Q(L cos a a) + M(Sa) = 0
(—QL cosa +M)da=0
Since
6a ¥ 0,
—-QLcosa+M=0 M=QLcosa <
and

a=cos ! a
OL

For
L = 12 ft, Q = 5 kip, and M = 38 kip - ft,

a=cos ! = = 5) foe
1x5 pat:
(continued)

536
EXAMPLE 10.2 (concluded)
SS,

Note: By applying the principle of virtual work, we were able example problem using the equations of equilibrium. As this
to obtain the solution for this problem directly; the solution example illustrates, the principle of virtual work offers
by means of the equilibrium equations would require considerable computational advantage over the equilibrium
dismembering the structure, drawing the free-body diagrams of equations approach when analyzing structural and mechanical
its individual members, and calculating the internal forces at systems composed of multiple rigid bodies.
the connections. The student is encouraged to solve this

EXAMPLE 10.3

Determine the expression for the horizontal reaction at the sup-


port B for the truss shown, using the principle of virtual work.

Solution The free-body diagram of the truss shows the two


known forces P and Q, and the four unknown reaction
components A,, A,, B,, and lee

Virtual displacement. Point B is displaced a small distance in


the horizontal direction, while keeping support A in its
original position. Note that the virtual displacements of B and C
in the horizontal direction are —dx, and — 8x, respectively;
the minus signs indicate that they are directed opposite to
the positive x axis.

Equation of virtual work. This we can write

OU = BY(—8xz) — POyc) — O(— 8x—) = 0


In order to relate the three virtual displacements to the virtual
rotation da, we write the expressions for the coordinates of the
points of application of P, Q, and B, in terms of the angle a.
Thus,

Xp = 2L cos a Xc =L cos a yo =Lsina

Differentiating these expressions, we obtain


dx, = —2Lsinasa 8x, = —Lsinada
dyc = Leos a da
Substituting the above relationships into the virtual work
equation, we obtain
6U = B,(2L sin a) 8a — P(L cos a) 8a — Q(L sin a) 8a = 0
(2B,L sin a — PL cos a — QL sin a)da = 0
As 6a # 0,

2B,L sina — PL cosa — QL sina =0

ye P wees on! ; iP g
2sina 2 e tannow2

537
538 Chapter 10 Work and Energy

10.5 POTENTIAL ENERGY


The energy of a mechanical or structural system is simply defined as its ca-
pacity for doing work. The term potential energy, initially introduced by Wil-
liam J. M. Rankine in 1853, describes the energy a system has because of its
position with respect to a reference position. Before we can relate work to the
change in potential energy of a system, it is necessary to understand the idea
of conservative systems.

Conservative Systems
A mechanical or structural system is considered to be conservative if the work
done by all the forces acting on it depends only upon the initial and final
positions of the system, regardless of the path followed between the initial and
the final positions. In other words, the work done by the forces is independent
of the path of the system, and depends only on its terminal positions.
Consider, for example, a body of weight Q shown in Figure 10.8. Using
Equations (10.6) and (10.7), the work performed by the weight of the body as
it moves from position B to A can be written as

u=| -op-ai+an=| oa
‘A y1

y S)

U = Oy7 = OV Oy, 97) (10.15)


where y, and y, are the vertical coordinates of points A and B, respectively.
As Equation (10.15) indicates, the work done by the weight depends only
Vi upon the vertical displacement (y, — y,) of the center of gravity of the body,
and it remains the same regardless of the path followed by the body from B
O to A. Therefore, the system shown in Figure 10.8 is conservative.
Figure 10.8 Mechanical and structural systems containing appreciable friction forces
cannot be considered to be conservative. This is because the work done by the
friction forces depends on the length of the path along which the body moves
from one position to another; thus, for longer paths, the work of the friction
forces will be greater than for the shorter paths. Such systems are called non-
conservative. Our treatment of potential energy in this chapter is restricted to
conservative systems.

Potential Energy of External Forces


Consider a body of weight Q in position A, at an elevation y, above a reference
plane as shown in Figure 10.9. Since the weight Q has the potential of per-
forming work equal to + Qy, if allowed to move from elevation y, to the
reference plane (y = 0), its potential energy is Vj, = Qy,. Similarly, the
potential energy of Q in position B (at elevation yg) would be V;z = Qyz, and
in position C (at a distance of y. below the reference plane) it would be
Vic = —Qyc. Note that the potential energy of Q, when it is located at the
reference plane, is zero.
10.5 Potential Energy 539

VB aA

Vra= Ov, x 4 VB

Q VA

V = O0—————__—_ L_ y=0
ye Reference
C plane
Veo=-OVe s :

Figure 10.9

The dimensions of potential energy are the same as those of work: length
times force. The commonly used units are ft - lb or in - lb in the U.S. Custom-
ary system, and joules in the SI system.
Now, let us consider the change in potential energy of Q as the body dis-
places from position A to position B. The change is positive because the po-
tential energy increases as the body moves upward, and is equal to

Ven — Ven = +OCyp — Ya) (10.16)


However, the work done by Q as the body undergoes this displacement is
negative, and is given by
Ua ee Yaya) (10.17)
By comparing Equations (10.16) and (10.17), we see that the change in po-
tential energy is equal to the negative of the work done by the body as it
displaces from position A to B, or

Equation (10.18) can be rewritten as


OT Ven er (10.19)
Considering the downward displacement of the body from position B to
position A, we observe that the change in potential energy is negative, and is

Vea = Vip OV eaOyn = Oyn — Ya)


while the corresponding work done is positive and is equal to

Opa Oni Ya)


Again, we observe that the change in potential energy is the negative of the
work done; thus,
Vea = Vip as Up
540 Chapter 10 Work and Energy

or

Una, — Vip i Vea (10.20)

In general, for any displacement of the body from an elevation | to another


elevation 2, the relationship between the work done by the weight and its
change in potential energy can be expressed as

It is important to realize that the changes in potential energy do not depend


upon the location of the reference plane from which the elevations are mea-
sured, but remain the same regardless of where the reference plane is located.
Since we will only be concerned with the changes in potential energy, rather
than its total values, we may locate the reference plane at any convenient level.
Although Equation (10.21) was derived for the case of a weight or a gravity
force, it can be used for any other type of conservative external force simply
by denoting the negative of the work done by the force as the change in its
potential energy. In the case of a conservative system subjected to more than
one external force, Equation (10.21) remains valid, with U,_,, now represent-
ing the total work of all the external forces during the displacement of the
system from position 1 to position 2, and V,, and V;, now representing the
sum of the potential energies of all the external forces in positions 1 and 2,
respectively, with respect to an arbitrary reference position.

Elastic Potential Energy


So far in this chapter, we have focused attention on rigid bodies, and systems
composed of rigid bodies. When a system also contains elastic springs, the
potential energy of the springs, referred to as the elastic potential energy, must
be included in the total potential energy of the system.
Figure 10.10 shows a linear spring whose undeformed length is L. In Sec-
tion 2.10, the relationship between a force F and the corresponding deforma-
tion x of a linear elastic spring was given as

F=ke (10.22)
where k = stiffness of the spring (or spring constant).
The work done by the force F, which the spring exerts on the body A
attached to it, during an infinitesimal displacement dx can be written as

dU = —F dx = —kx dx (10.23)
The work performed by the spring force F on the body as it undergoes a finite
displacement from the undeformed state of the spring (x = 0) is obtained by
integrating the expression of dU. Thus,

u=| -Fax=|, —kx dx = —4k? (10.24)


10.5 Potential Energy

k |
|
A
|
The force F exerted
Undeformed length Z by the spring on

lp p>~aQ@A
Y= +kx 2
x

Figure 10.10

In general, the work of F as the body displaces from a position x, to another


position x, is obtained by integrating Equation (10.23) from x, to x:
X92 X72

Ua5 = | —Fdx= | —kx dx = $k(x2 — x3) (10.25)


xj x]

These expressions of work remain the same for either elongation or compres-
sion of the spring.
Since a spring with a deformation x has the potential of performing positive
work of magnitude $kx” on the body if allowed to return to its undeformed
length, its potential energy is given by

V, = +3kx? (10.26)

As the deformation of the spring increases from x, to x,, the change in its
elastic potential energy is equal to (Figure 10.10)

Von — Ver = 3h — ako = 3k — x7) (10.27)


Comparing Equations (10.25) and (10.27), we observe that the change in elas-
tic potential energy is equal to the negative of the work done by the spring
force on the body as it displaces from position | to 2, or
Veo mal Vey — U,_.2 (10.28)

which can alternately be written as

Uy = Va Va (10.29)
542 Chapter 10 Work and Energy

The above relationships for a linear spring can easily be modified for a
torsional spring. The moment-rotation relationship of such a spring can be
expressed as
M=k0 (10.30)
where k = stiffness of the torsional spring (units of Ib - ft/rad or N - m/rad).
The potential energy of a torsional spring can be given by

V, = 5h 67 (10.31)

where 8 is the rotation in radians measured from the undeformed position of


the spring.

Total Potential Energy


The total potential energy V of a mechanical or structural system composed
of rigid bodies and springs can be expressed as

V=V,+V, (10.32)
where V; is the sum of the potential energies of all the external forces acting
on the system, and V, is the sum of the elastic potential energies of all the
springs in the system.
The relationship between the change in total potential energy of a system
as it displaces from a position 1 to another position 2, and the work done by
the external and spring forces during the displacement is

Cee eae (10.33)


The above relationship indicates that the change in total potential energy is
negative of the work done by the forces.

10.6 EQUILIBRIUM—PRINCIPLE OF STATIONARY


POTENTIAL ENERGY
The principle of virtual work presented in Section 10.3 can now be restated
in terms of the total potential energy of the system. For a small virtual dis-
placement, Equation (10.33) can be written as

dU =V, — (V, + 8V)


dU = —8V (10.34)
where 5V = the change in the total potential energy of the system due to the
virtual displacement. Since 5U = O for a system in equilibrium, Equation
(10.34) yields
10.6 Equilibrium—Principle of Stationary Potential Energy 543

dV =0 (10.35)

as a condition for the equilibrium of the system. Equation (10.35) is a restate-


ment of the principle of virtual work in terms of the total potential energy of
the system. It states:

_ If a system is in equilibrium, the change in its total potential energy due to STATEMENT Ii
| any small virtual displacement is zero.

When the deformed configuration of a mechanical or structural system can be


defined in terms of a single variable, s, the total potential energy of the system
can be expressed as a function of s, or V = V(s). The equilibrium condition
dV = Ocan then be expressed as

oV= (=)ds = 0 (10.36)


ds
Since 5s represents the nonzero virtual displacement, the equilibrium condition
becomes

=0 (10.37)

Equation (10.37) represents the principle of stationary potential energy, which


states:

In an equilibrium configuration, the total potential energy of the system is STATEMENT Iil
_ Stationary; that is, the first derivative of the total potential energy is zero.

The application of this principle considerably simplifies the analysis of


equilibrium for mechanical and structural systems composed of multiple rigid
bodies and springs.
EXAMPLE 10.4

Determine the equilibrium configurations in terms of the angle a


of the two-link mechanism shown, using the principle of station-
ary potential energy. The spring is undeformed when a = 0°. Let
= 48 in., k = 20 lb/in., and P = 288 Ib.

Solution The free-body diagrams of the mechanism in the


undeformed (a = 0°) and the deformed configurations are
shown.
We can see that the work done by the force P during the
displacement of the structure from its undeformed state
(a = 0°) to a deformed state defined by the angle a is equal to
ROE a2 Ecos 0)\e—s 225 le COSia)s
kL sin a Therefore, the potential energy of P, which is negative of
the work done, is
V-= —2PL(1— cosa)

The elastic potential energy of the spring, which undergoes


a deformation equal to L sin a, is

Ae , V, = 3k(L sin a)? = $kL’ sin? a


Peas); cos @~@————>|__
So, the total potential energy of the system is
- 2L V=V,+V, = —2PL(1 — cos a) + 3kL? sin? a
We can now determine the equilibrium configurations by
setting the first derivative of V equal to zero. Thus,
dV a
ae 2PL(sin a) + $kL7(2 sin a cos a) = 0
Q

—2PL sina + kL? sina cosa = 0


L sin a(—2P + kL cosa) = 0

This equation is satisfied if


sina = 0 and —2P + kL. cosa
=0

These solutions indicate that there are equilibrium configurations


corresponding to
OP.
P= 288 lb Q — OF an d Q — COs a (

Substituting the given numerical values, we obtain a for the


second and third equilibrium configurations as

a=cos ! pata SS SES. 116©


203048) ae ae
Therefore,
On—n0n and CUS Sassy <<

The first two equilibrium configurations are shown. Note that


P= 288 |b since the cosine cannot be greater than unity, the second
and third equilibrium configurations do not exist when
Equilibrium configurations Pek.

544
10.7 Stability of Equilibrium—Principle of Minimum Potential Energy 545

10.7 STABILITY OF EQUILIBRIUM—PRINCIPLE


OF MINIMUM POTENTIAL ENERGY
A mechanical or structural system is considered to be in stable equilibrium if,
when subjected to a small displacement from its equilibrium configuration, it
returns to that configuration. Conversely, if a small displacement causes a
system to move farther away from its equilibrium configuration, it is consid-
ered to be in unstable equilibrium. If a system, when given a small displace-
ment, remains in that displaced configuration, neither returning toward nor
moving farther away from the equilibrium configuration, it is said to be in
neutral equilibrium.
To illustrate these ideas, consider a bead of weight Q that slides on a fric-
tionless wire as shown in Figure 10.11a. The total potential energy of the bead
in a position x is equal to the potential energy of its weight Q, and can be
written as

V(x) = Oy)
A plot of the total potential energy function V(x) is shown in Figure 10.110.
Now, let us consider the three positions of the bead denoted as A, B, and C in
Figures 10.11a and 10.115. Note that these positions are equilibrium positions
because the slope of the total potential energy function is zero (dV/dx = 0)
at these points. When the bead is in position A, it is in stable equilibrium
because if subjected to a small displacement, it will return to its equilibrium
position A, at which the total potential energy is minimum. In position B, where
the total potential energy is maximum, the bead, if given a small displacement,
will move farther away from the equilibrium position B; thus, the bead is in
unstable equilibrium. In the third position C, the bead is in neutral equilibrium,
because if subjected to a small displacement, it will neither return to nor move

yx) V(x) ay <0

| Reference plane

(a) (b)
Figure 10.11
546 Chapter 10 Work and Energy

farther away from C, but will remain in the displaced position. The total
potential energy is constant in the neighborhood of the neutral equilibrium
position C.
In general, if a system is in stable equilibrium, its total potential energy is
a relative minimum. This is known as the principle of minimum potential
energy. Also, as just discussed, the equilibrium of a system is unstable if the
total potential energy is a relative maximum, and neutral if the total potential
energy is constant in the neighborhood of the equilibrium configuration.
When the total potential energy of a system can be expressed as a function
of a single variable s, then the stability, instability, or neutrality of equilibrium
can be determined by examining the second derivatives of the function V(s)
at the equilibrium positions, where the first derivative of V(s) is zero. If the
second derivative of the function V(s) is positive at an equilibrium position,
then V(s) is minimum, and the system is in stable equilibrium. Conversely, if
the second derivative is negative at an equilibrium position, then V(s) is max-
imum, and the system is in unstable equilibrium. Thus, the criteria for the
stability, instability, and neutrality of equilibrium of a system can be expressed
as follows:

dV d’V aig
—=0 and ia) Stable equilibrium (10.38)
ds ds

dV av whee
arent) and <0 Unstable equilibrium = (10.39)
ds ds?

V
As" =\() On = 2.3.00) Neutral equilibrium (10.40)
s

If the second derivative of V is zero, then its higher derivatives must be ex-
amined. If the order of the lowest nonzero derivative is even and the derivative
is positive, then the equilibrium is stable. Otherwise, it is unstable. As given
by Equation (10.40), for neutral equilibrium, all derivatives of V must be zero.
In this chapter, we have considered only those systems whose deformed
configurations (and the total potential energy) can be defined in terms of a
single variable. Such systems are called the single degree of freedom systems.
Analysis of the multi degree of freedom systems, where more than one inde-
pendent variable is needed to define the deformed configuration of a system,
is beyond the scope of this text.
10.7 Stability of Equilibrium—Principle of Minimum Potential Energy 547

EXAMPLE 10.5

Examine the stability of the equilibrium configurations of the


two-link mechanism that we determined in Example 10.4.

Solution In Example 10.4, we obtained three equilibrium


configurations of the system corresponding to a = 0° anda =
+ 53.13°. It was also shown that the total potential energy and
its first derivative are given by

V = —2PL(1 — cos a) + 3kL? sin?a

— = —2PL sina + kL’ sinacosa


da

By differentiating the above expression with respect to a, we


obtain the second derivative of V. Thus,

a°v
Eines 2
—2PL cos a + kL*(—sin~ ED, a + cos* 2 a)
(o)

Substituting the numerical values L = 48 in., k = 20 lb/in.,


and P = 288 lb,

a’V
ae — 27,648 cos a + 46,080 (—sin*a + cos? a)
‘a
Next, we determine the numerical values of the second
derivative fora = 0° anda = +53.13°. Fora = 0°,

d°v
aries 27,648 + 46,080 = + 18,432 >0

As the second derivative is positive,

the system is in stable equilibrium ata = 0° <

hOnmOm—me) Salou

mh = —27,648 cos +53.13°

+ 46,080(—sin? +53.13° + cos” +53.13°)


= —29,491 <0

Therefore,

the system is in unstable equilibrium ata = +53.13° <


548 Chapter 10 Work and Energy

10.8 SUMMARY
In this chapter, we have learned the following:

1. The work done by a force F on a body is equal to the magnitude of the


force times the displacement of the body in the direction of the force
(Section 10.2). The work performed by F as its point of application on
the body undergoes a finite displacement from position C to D is
given by
D D D

u=|(G F-as = |
C
Fos 6 ds = |
GS
(F, dx + F, dy + F, dz)
(10.5) — (10.7)
Similarly, the work of a couple due to a finite rotation of the body from
8, to 8, is given by
82
U= M do - C093
8;

The principle of virtual work can be expressed as (Section 10.3)

oU = 8U, + 6U,.=0 (10.14)

Potential energy is the energy a system has because of its position with
respectto a reference position. The potential energy of a weight Q at
an elevation y above a reference plane is equal to Qy. The potential
energy of any other type of conservative external force is denoted by the
negative of the work done by the force as the body displaces from a
reference position to the displaced position. The elastic potential energy
of a linear spring due to a deformation x is equal to 3kx” (Section 10.5).
The principle of virtual work can be restated in terms of the total
potential energy of the system as

dsV=0 (10.35)

When the deformed configuration of a mechanical or structural system


can be defined in terms of a single variable s, the equilibrium condition
dV = Ocan be expressed as
QV
aa (10.37)

which is the principle of stationary potential energy (Section 10.6).


A mechanical or structural system is considered to be in stable
equilibrium if, when subjected to a small displacement from its
equilibrium configuration, it returns to that configuration. The principle
of minimum potential energy states that if a system is in stable
equilibrium, its total potential energy is a relative minimum
(Section 10.7).
Problems 549

KEY TERMS

conservative system 538 principle of virtual work 532


neutral equilibrium 545 stable equilibrium 545
potential energy 535 unstable equilibrium 545
principle of minimum potential virtual displacement 532
energy 546 work 528
principle of stationary potential
energy 543

PROBLEMS

SECTION 10.4
10.1 Determine the horizontal reaction at the support A using the principle of
virtual work.

10.2 Determine the horizontal reaction at the support C of the two-member frame
shown using the principle of virtual work.

7 kip

bee} —v} —_+ :


Sits ft. 10 ft LS ftebe5ft--}<—7 ft—+|
Figure P10.1 Figure P10.2

10.3 The two-member frame is subjected to two vertical loads as shown. Determine
the horizontal reaction at the support A if a = 90°. Use the principle of virtual work.

10.4 Solve Problem 10.3 if a = 60°.

100kN 200 kN
550 Chapter 10 Work and Energy

10.5 Determine the relationship between the force P and the couple M necessary for
the mechanism to remain in equilibrium. Use the principle of virtual work.

10.6 For the mechanism shown, determine the angle of inclination a if P = 5 kip,
L = 12 ft, and M = 38 kip- ft. Use the principle of virtual work.

10.7 Determine the relationship between the forces P and the couple M necessary
for the mechanism to remain in equilibrium. Use the principle of virtual work.

10.8 Determine the magnitude of the couple M required to maintain equilibrium of


the mechanism shown if P = 100 N, L = 2 m, and 0 = 30°. Use the principle of
virtual work.

10.9 Determine the relationship between the force P and the couple M necessary for
the mechanism to remain in equilibrium. Use the principle of virtual work.

10.10 For the mechanism shown, determine the magnitude of the couple M required
to maintain equilibrium, if P = 200 Ib and @ = 80°. Use the principle of virtual
work.

Figure P10.5 and P10.6 Figure P10.7 and P10.8 Figure P10.9 and P10.10

10.11 For the lifting platform shown, establish a relationship between the weight Q
and the horizontal force P required to maintain equilibrium. Use the principle of
virtual work.

10.12 For the lifting platform shown, determine the height h if Q = 2 kip, P = 1.5
kip, and L = 15 ft. Use the principle of virtual work.

10.13 For the crankshaft mechanism shown, determine the force P necessary for the
mechanism to remain in equilibrium if a couple of moment 65 N - m is applied and
L, = 150 mm, L, = 75 mm, and a = 45°. Use the principle of virtual work.

10.14 Solve Problem 10.13 if a = 110°.

10.15 The hydraulic crane is used to lift the weight as shown. Determine the force
exerted by the hydraulic cylinder BD at D using the principle of virtual work.
Neglect the weight of the crane.
Problems 551

= 10

4ft
)B “a

5 ft
HA
Figure P10.11 and P10.12 Figure P10.13 and P10.14 Figure P10.15

10.16 and 10.17 Determine the force exerted by the pliers on the smooth rod A
using the principle of virtual work.

|Ib

a
} ee

Nob

12 1b

4.2 in.
ep :
Figure P10.16 Figure P10.17

10.18 The single-axle semi-trailer, weighing 40 kip, is attached to the cab of the
truck at joint B, which can be assumed to be a hinged connection. The weight of the
cab is 8 kip. Determine the vertical force acting on the axle D using the principle
of virtual work.

Figure P10.18
Chapter 10 Work and Energy

10.19 For the beam shown, determine the vertical reaction at the support A using
the principle of virtual work.

10.20 For the beam shown, determine the vertical reaction at the support D using
the principle of virtual work.

10.21 For the beam shown, determine the vertical reaction at the support C using
the principle of virtual work.

10.22 For the beam shown, determine the vertical reaction at the support A using
the principle of virtual work.

10.23 For the beam shown, determine the vertical reaction at the support C using
the principle of virtual work.

10.24 For the beam shown, determine the reaction couple at the fixed support A
using the principle of virtual work.

4m4m 8m 8m 8m4m4m 8m [+10ftet-10 ft-fo10 ftete10 ftfel 0 ftef


Figure P10.19 and P10.23, Figure P10.20 and P10.22

SOKN 100 kN 175 KN

[3 mte-3 m-of+3 mobe-3 m-ob—3 ma


Figure P10.21 and P10.24

10.25 Determine the reaction in the x-direction at support B of the space truss
shown by using the principle of virtual work.

10.26 Determine the reaction in the z-direction at support B of the space truss
shown by using the principle of virtual work.

SECTION 10.6
10.27 Solve Problem 10.1 using the principle of stationary potential energy.

10.28 Solve Problem 10.5 using the principle of stationary potential energy.
Ball-and-socket
Imvzv 1m
Ve support 10.29 Solve Problem 10.7 using the principle of stationary potential energy.
Zz
Figure P10.25 and P10.26 10.30 Solve Problem 10.12 using the principle of stationary potential energy.
Problems 553

SECTION 10.7
10.31 Determine the equilibrium configuration(s) of the bar shown, in the range of
—90° < a S 90°, using the principle of stationary potential energy. Examine the
stability of the equilibrium configuration(s). The spring is undeformed when a = 0°.
Let k = 60 lb/in., P = 200 lb, L = 2 ft, anda = L.

10.32 Solve Problem 10.31 if a = L/2.

10.33 Solve Problem 10.31 if the spring is undeformed when a = 5°.

10.34 Solve Problem 10.31 if the spring is undeformed when a = 5° anda = L/2.
Figure P10.31, P10.32, P10.33, and P10.34
10.35 Determine the equilibrium configuration(s) of the bar shown, in the range of
—90° < a < 90°, using the principle of stationary potential energy. Examine the
stability of equilibrium configuration(s). The spring is undeformed when a = 0°. Let
k = 50 N-m/rad, Q = 150N, and L = 300 mm.

10.36 Solve Problem 10.35 if the spring is undeformed when a = 5°.

10.37 Solve Problem 10.35 if k = 30 N- m/rad.

Torsional
spring

Figure P10.35, P10.36, and P10.37

10.38 Determine the minimum value of the spring constant k so that the system will
be in stable equilibrium for a = 0°. The springs are undeformed when a = 0°.

Figure P10.38
554 Chapter 10 Work and Energy

10.39 and 10.40 Determine the minimum value of the spring constant k so that the
system will be in stable equilibrium for a = 0°. The spring is undeformed when
a = 0°.

10.41 The handle OA of the mechanism, which is rigidly attached to the circular
disk, is subjected to a force P as shown. Determine the equilibrium configuration(s)
of the system using the principle of stationary potential energy. Examine the stability
of equilibrium configuration(s). The spring is undeformed when a = 90°. Let k =
30 Ib/in., a = 12 in., r = 4 in., and P = 100 lb.

10.42 Solve Problem 10.41 if the spring is undeformed when a = 80°.

Figure P10.40 Figure P10.41 and P10.42

10.43 Determine the equilibrium configuration(s) of the pantograph shown using


the principle of stationary potential energy. Examine the stability of equilibrium
configuration(s). The spring is undeformed when a = 70°. Let a = 0.5 m, k = 1800
N/m, and P = 150N.

10.44 Determine the minimum value of the spring constant k, so that the frame will
be in stable equilibrium for a = 0°. The springs are undeformed when a = 0°.
Use P = 10 kip and L = 8 ft.

no, SS

ia
Figure P10.43 Figure P10.44
APPENDIX A

ST PREFIXES

Prefix Factor Symbol


atto 10~'8 a
femto 1Ome f
pico Ome Pp
nano 10~? n
micro 10~° pe
milli Ome m
kilo 10° k
mega 10° M
giga 10° G
tera 10!2 T

555
APPENDIX B

CONVERSION FACTORS

Table B.1 Conversion factors from US Customary units to SI units Table B.2_ Conversion factors from SI units to US Customary units
Length Speed Length Speed
1 ft = 0.3048 m 1 ft/sec = 0.3048 m/sec femmes 28ileft 1 m/sec = 3.2808 ft/sec
1 ft = 304.8 mm 1 mile/hr = 0.44704 m/sec 1mm = 3.281 x 10-7 ft 1 m/sec = 2.23694 mile/hr
1 in. = 0.0254 m 1 mile/hr = 1.6093 km/hr 1m = 39.37 in. 1 km/hr = 0.6214 mile/hr
1 in. = 25.4 mm Acceleration 1 mm = 0.03937 in. Acceleration
Area 1 ft/sec? = 0.3048 m/sec?” Area 1 m/sec? = 3.2808 ft/sec?
1 ft? = 0.0929 m2 1 in./sec? = 0.0254 m/sec? 1 m* = 10.764 ft? 1 m/sec? = 39.3701 in./sec”
1 ft? = 929 x 10? mm? Specific weight 1 mm? = 10.764 x 107° ft? Specific weight
deine) = 6.452) <1068-m- 1 Ib/ft? = 0.1572 kN/m? 1 m? = 1550 in? 1 KN/m? = 6.361 lb/ft?
1 in? = 645.16 mm? 1 Ib/in? = 271.66 kKN/m? 1 mm? = 0.00155 in? 1 kKN/m? = 0.00368 Ib/in?
Volume Moment Volume Moment
ieftr—228:317e nl Omome 1 lb-ft = 1.3558 N-m ‘lim? = shay) Re 1 N-m = 0.7375 lb -ft
1 ft? = 28.317 x 10° mm? 1 Ib- in. = 0.11298 N-m 1 cm? = 35.32 x 1074 ft? 1N-m = 8.851 Ib- in.
1 in? = 16.387 xX 10-° m? 1 m? = 61,023.4 in? Energy
Energy
1 in? = 16.387 x 10° mm? 1 ft-lb = 1.3558 J Goules) ra een 1 J Goule) = 0.7375 ft - Ib
Force
Moment of inertia of area 1kN = 0)4.8 Ib Moment of inertia of area
1 Ib = 4.448 N
1 in* = 0.4162 x 10° mm* 1 m* = 2.402 x 10° in*
1
1
lb =
kip =
4.448 x 107-2 kN
4.448 kN
1 in* = 0.4162 x 10~° m4 eee
1 N/m = 0.0685 lb/ft
1 mm* = 2.402 x 1076 in*
1 lb/ft = 14.593 N/m

556
APPENDIX C

SPECIFIC WEIGHT OF
COMMON MATERIALS
(AVERAGE VALUES)

___Specific
weight
Material (lb/ft?) (KN/m?)
Water 62.4 9.81
Wrought iron 480 75.46
Structural steel 490 77.03
Stainless steel 495 77.82
Cast aluminum 172 27.04
Magnesium 115 18.08
Brass 545 85.68
Concrete 150 23.58
Red oak (air-dry wood) 44 6.92
Douglas fir (air-dry wood) 38 5.97
APPENDIX D

MATHEMATICAL
EXPRESSIONS

A. Quadratic equation 15. sin? x + cos?x = 1


If ax? + bx +c =0 16. sec?x = 1 + tan?x
—b + Vb* — 4ac 17. csc?x = 1 + cot?x
aa 2a 18. sin@x + y) = sinx cos y + cos x sin y
19. sin — y) = sinx cos y — cos x sin y
B. Trigonometric relationships 20. cos(x + y) = cos x cos y — sinx sin y
Z\e cos(x — y) = cos x cos y + sin x sin y
sinx =
csc x tan x + tan y
22: tan + =
1 ») 1 — tan x tan y
cos xX =
sec Xx
1 _ sinx
23: erg yy)2 Ea t =i!
1 + tan
x tan y
tan
x =
cotx cosx cot tyigel
24. Cot ety) Ee
sin x = cos(90 — x) cot x + coty
cos x = sin(90 — x) wie = 9) _ cot
t x cot y + 1
tanx = cot(90 — x) 25:
cot y — cot x
cot x = tan(90 — x)
sin 2x = 2 sin x cos x C. Hyperbolic functions
cos 2x == cos’2 x Be RRP)
sine x =? p.
= 2 cos* x —-l= 1
Lt sinh x = ————
1 — 2 sin?x cosech x
2 tanx 1
tan 2x = cosh x =
1 — tan?x sech x
cot?x — 1 1
cot 2x = tanh x =
2 cotx coth x
sin 3x = 3 sinx — 4sin?x sinh x= + Vcosh?
x—1
cos 3x = 4cos*>x — 3 cosx cosh x = + V1 + sinh? x
3 tan x — tan’x
tan 3x
1 — 3 tan?x
558
Mathematical Expressions 559

DD series 4
d
ii nyo
du
aa | iat
oe 7
dx ) . dx
1 SNe Sor Pe eee ore = =)
a Me {ph ly le
ae e se : dx u" yr! dx

K <) 6
, =d
dx
( Bey
& = ih n= py”)
nv U du

dx
de
du
ea =
d )
3}. Pe
D
os oer
! 5! = ea)
CA Sea eedx
(for
x? < ©) memes Non dk
Gan @ x2 ae 3
4. COIR Sy ae Sere Serena
8 eea (SIilec) ) a COSEY
dx
+++ (for x? < «)
Q, oF
= (cos x) = —SIisin x
5. (tay =1+m+ = oe
n(n
— 1)\n — 2 IQ, =d (@in we) = Geer 2 x
MOO bo (for x? <1) dx

= ||) eed (CSCEG Em COUNICSCEX


6. (x + y)" =x? 4 eee) a n(n n-2y2 dx
2! d
nin — 1)\(n — 2 121 (SCCE) a—ntallenisecnG
aL MO ray 2h -++ (for dx
d (COPS) a— se CSCary
D
ee) 2 ee
dx
ee B 4
Oho.
14. —(Sinh x) = coshx
dx
real values of x)
d :
15. — (cosh x) = sinh x
E. Logarithms dx
1. Ify = a‘ thenx = log,y
2. Ify = e* thenx = log, y = Iny L6) —d (tanh x) ="sech- 2) x
dx
3. loga" =nloga
1 fe » — (osu) ere
= ) aes
4. log (2)= —logn dx Ba Ba u dx
n
18 d a ) 1 du
ek (OCF) ae
5. log (ab) = loga + logb deere Me ied
log 4]= loga — logb G. Integrals
1 iiadx ax
F. Derivatives xitl

' d Ga du 2s [© dx = (except forn = —1)


(ll) = iQ ar ll
dx dx


ODTiby,te Vv) =te eraDaca
a) ge

du du
Ue at

|
Ss 5 = Q.=
560 Appendix D Mathematical Expressions

x dx 1
| dx ele b 18. [ AA = Sta + mx - ane + bo)

day eG a heore 1
19. [2% = —In(a + bx’)
| dx 1 , (xVab aly ak bx?, ~ 2b
__—— ——_— [. — SS

a+ bx* Vab Si a C 20. xdx opeey 2


1 Vxe = a
9 I f2 #4 a? = — Vee a ra x d
2 21; ee eee
In (x + Vx? + a’)] V an x
Pete O28) {sin dx=—— COse
10. | a — x° dx 232) fecOs dx esis
1 - 2A. sii taniadx, =-In (sec 4)
75 xVa— x? 4 at sin lal 25. f cotxdx
= In (sin x)
in 2x
11 [x > a wy LF oon 1 (a x 3 XS. |sin?x ax =S . = 7

3 :
2 _ sin 2x
12, |x2Va? = x?dx = -2 Vie = a + 7. Jootede=
3+
2 , esi
we

© |va=e + asin (*)| 28. |sin.xcos x dx = 5

3 al cos x
13. [ eve? a) 1 (»7 2)
3 2 5 29. isin”nie x dx ae in? 38)
(@ ae Sir

Via — 2 30. |cos? x dx = a (2 + cos? x)


14. | Va ¥ bxdx = 22 Via + be :
Bik J sinh x dx ~= cosh x
3b 32. f{cosh x dx\= sinhx

1 iE PA SIODS UiAS 2 eae ir He eh aah ee


Via + bx) 35. fxcosxdx = cosx + xsinx
dx 2a + bx 36. fx? sinx dx = 2x sinx — x* cosx + 2cosx
16. NRPETS. ; 37. {x cosxdx = 2x cosx + x7 sinx — 2 sine

i
APPENDIX E

PROPERTIES OF AREAS
AND HOMOGENEOUS
BODIES
Table E.1 Properties of areas

Moment of inertia
Shape Area PA (or second moment of area)
Rectangle y y’
bh x=3
=_b
I, = xbn3
—_h = Lye
= fy = 3hb

Ie = bh}
ly onl
= hb

Triangle
bh3
1 —_bte Ik, See
xbh 2,5 3 x 1

= bh (2 2
y=3 Ty = 770 Es CS)
bh3
136s
pe bha (p2 —betc~)2

1 =D _ bh}
Dear: zbh x=3 Lae,
Bey): _ beh
Ya ly =
_ bh?
Fy! = 36
a 0eh
hy =

Circle 4
mr? x=0 ae oe
yer yep Le L.
Ip = Ly =e

Semicircle ' 7
iis
2 4 = tee
Ty = Ly = Ty =e
_ (902
— 64\ 4
‘toa 720 )r

Note: Location of centroid is O'; Ox and Oy are Cartesian coordinate axes; O'x' and O'y’ are centroidal coordinate axes; x and y are centroidal coordinates with respect to
the Ox and Oy axes.

562
Table E.1 Properties of areas (continued)
Moment of inertia
Shape (or second moment of area)

Quarter-circle
mrt
ik, Se ee
Be y 1

Ellipse yyy
I, = 1 =nab?

elie Te

Peae

Circular sector y'


4
ig °
= 1, => (a— sina cosa)
4
if A
h = 7 (at sin a cos a)

4
r 5 4 r4sin2a
Ty = 7 (a+ sin a cos a) — 9° a

Circular segment yy’ 4


r : :
1s = 7 (a+2 sin? a cosa — sin a cos &)
4
ie . 5
fe = [y= (3a — 2 sin? a cos @— 3 sin a cos a)

Trapezoid y y "
h(2b, + b>)
y= h3
3(by + b>)
If. = 75 (3b, + by)

: h3 (b? + 4b,by + b3)


h — x!
25 36(b; + b>)

| nel

563
Table E.1 Properties of areas (concluded)
Moment of inertia
Shape (or second moment of area)

Quarter-circle spandrel
y 2r
X= 3(4—m)
= _ (10 —-3m) 1 TT
341m). L=(5-% r4

3a
8
“on
Ie = 7g
1
(ab3)
me5 [aaa
3

ab 3
=e 254
3 L, =

_ bat
fae

564
Table E.2 Properties of homogeneous bodies
Shape Volume Xi, yale Mass moment ofinertia

Uniform slender rod


AL 0,0,0
Lp=y ll = ml?

Circular cylinder oy
nR2L 0, 0, 0 ifce yi 2
=5mR

fae l,=[)=
y z 4 m(3R? + L?)

0, 0, 0

Hemisphere; '
Radius=R ae

Note: Location of center of gravity is O’; Ox, Oy, Oz are Cartesian coordinate axes; O’x’, O'y’, O'z' are centroidal coordinate axes; m = mass; X, y, Z are
centroidal coordinates with respect to the Ox, Oy, and Oz axes.

565
Table E.2 Properties of homogeneous bodies (continued)
Shape Mass moment of inertia _
Rectangular prism I

ly alle
= =z m(L? Disp)
+ B?)

Ty = ygm(L?
+H?)
IT = B2+k L?
ye oe Be
er DX cngll 2
if. = 5 mH +zmL

Iy =1y
= mH?
Ly =
Zrtll
7yzmL
2

I = bm(L?
+ H?)
I, =3mL?
y = mL

1,Zoe
= 4 mH?
+4g ittmL?

I, Sy, =I, eo Sap


= 35 mR°+ eu
75 mH ee

1,y =1y,=2>mR?
== 7
Ty = Ty = 2mR?
+ 2 mi?

coh fed Et 2
Ty = (oz mR

Ty Spey
= 39 mR* pah
+ then
35 mH as)

566
Table E.2 Properties of homogeneous bo dies (concluded)
Volume Mass moment of inertia
Shape
Elliptic cylinder
5 ie =I, = 5m(a? + b?)
Y
pe 1,Yer =i4 mb2+4m
3
L?
z=0 ieee i
I, =qma + mL*

TyPie
=4mb2+ mL?
12

Lpalan
= ama’ wore
+ mL 2

Tetrahedron '
ee
aT ay ah cag Nie Ge)20)
Tg
—_b 1
at I, =ygm@*+c’)
= ee
ea. ie = Tym(a? +b?)

I x = 803 m(b24+ c?)

Xm(a?+
I,vy = 80 ce?)

1,Zz =>80 m(a?+b2)

567
ANSWERS TO
ODD-NUMBERED
PROBLEMS

CHAPTER1 2.35 F, = 278.54i + 111.41j (N)


1.1 1868.18 KN 2.37 F, = 416.78i — 238.11j (N)
1.3 24.587 m/s 2.39 F = 149.43i + 13.07] (Ib)
Bo eeS:85.Ns 2.41 129.73. N, Ps 49.06°
1.7 23.559 N/m’ 2.43 228.22 Ib, LS 29.66°
ee ~ 2.45 78.50 Ib, WY 76.76°
1.11 a = mv? (kg/s’) 2.47 679.75 Ib
CHAPTER 2 2.49 Ts = 373.36 Ib, Tac = 243.69 Ib
21 363.01b, A 854° 2.51 W = 428.98 N, Ts, = 377.23N
23 225.3N, % 341° 2.53 Tp, = 116.67 lb, Tgc = 89.37 Ib
27 304.0 1b, % 22.6° 2.55 P = 76.69 lb, a = 32.3°
29 F, = 5221b,R = 102.8 1b 2.57 Tag = 187.08 Ib, Toc = 99.54 Ib, Top = 108.0 Ib, a, = 60°
2.11 3773.94N, % 80.79° at EO
2.13 F, = 684N,a = 40° 261i
245 12001b 2.63 237.7N
217 2282 1b % 150.34° 2.65 F , = 0, Fry = 224.98 N, Fy, = 268.12 N
2.19 F, = 385.6N, resultant direction 60° A. 2.67 F, = 56.29 lb, F, = 54.54 Ib, F, = ~32.5 Ib
2.21 Along Am 89.07 Ib, along An 47.39 Ib 2.69 + (2) 16:79°,0) 95-25 Xe) 124.6 Ib
2.23 Along Am 321.39 N, along An 163.17 N OTN = C1486 Pit Uiay at as
2.25 180.4 Ib, 94° 2.73 a = 71.25°,B = 20°, y = 96.72°
A ie 2.75 a = 70.62°,B = 53.52°,y = 43.11°
299 %60N 277 o = 197.89°B = 5421 yng SI
2.31 F, — 150i (N), F, = 129.91 + 75) (N), Fy = 150) (N), 2.79 F = 123.7iNpo 295,97 p = 12839 y — 20 ds
F, = — 129.91 + 75j (N), 2.83 F = 163.48 Ib, « = 66.55°, B = 137.23°, -y = 123.44°
Fs = —150i (N), Fs = — 129.91 — 75j (N), 2.85 n = 0.3251 — 0.697j + 0.639k (Ib)
F, = —150j (N), Fs = 129.91 — 75j (N) 2.87 (a) F= 119.87 lb, y = 66.17°
2.33 F, = 122.871 + 86.03j (N), F, = —68.4i + 187.93j (N), (b) n= 0.7661 + 0.5j + 0.404k (Ib)
F; = 50i — 86.6j (N)
568
Answers to Odd-Numbered Problems 569

2.89 (a) 111.4° 3.35 20 Ib © at 12.5 in. from A


(b) n = 0.342i — 0.365j + 0.866k (N) 3.37 8 kN ©, Me, = 32 kN-m@
2.91 F, = 131.671 + 59.54j — 68.7k (Ib) 3.39 = 0, Mg, = 30kN-m©)
a= 34.62°, B = 68.15°, y = 115.43° 3.41 = 0M = 0
2.93 Ty, = —81.97i + 104.32j — 44.71k (N), a = 125.84°, 3.43 = 9.34 KN pa 52.7° at 332 mm
7nr
arma from A on AB,
B = 41.83°, y = 108.62° and 876 mm from B on BD
2.95 F = —45j (Ib) 3.45 R = 934kN ZV 52.7°, Mpg = 3kKNm@
2.97 T>, = 4.917 KN, T,, = —7.865 KN, T, = 4.917 kN 3.47 R = —86.2i + 342.3] (Ib) at 27.5 in. to the left and 3.5 in.
a = 62.08°, B = 138.52°, y = 62.08° below point A
2.99 26m 3.49 R = 60.03 kN 7.65° at 1.06 m from A on AC and 0.39 m
2.101 T = 16.5i + 39.6j — 61.58k (Ib), a = 77.3°, B = 58.15°, from B on BD
y = 145.14°
3.51 Yes; statically determinate
2.103 (a) R = —135i + 10k (N), (b) 135.37 N,
3.53 No; statically indeterminate
(c) a = 175.76°, B = 90°, y = 85.76°
3.55 Yes; partially constrained
2.105 (a) R = 64.191 + 154.8] + 80.36k (Ib),
3.57 Yes; statically determinate
(b) 185.85 lb,
(c) a = 69.79°, B = 33.6°, y = 64.23° 3.59 A, = 50 KN ©, C, = 50kKN
©, C, = 30kN®
2.107 R = 236.381 — 295.48j + 96.53k (Ib) 3.61 e=tS) 52KN V7 207 De 3112KN ©,
2.109 a = 126.42°,8 = 75.26°, y = 139.76°, F, = 589.49 N
D, = 19.44 kN @)
2.111 F; = 586.73 N, a = 124.98°, B = 143.78°, y = 98.23°
3.63 No; improperly constrained

2.113 733.63 lb
3.65 A, = 15kN©,A, = SKNQO,T= 49.5
kN LA 45°
2.115 Fp, = 231.43 N, Fog = 180.29 N, Foc = 180.31 N
3.67 A, = 5KN@®,B, = 5KIN@,T= 20kIN©
2.117 Fy = 125 lb, W = 96 lb, Fy = 192 Ib 3.69 Bs. = 0,A, = 0, M, = 18kN-m@)
3.71 = 0,A, = 1.75 kip ®, B, = 3.25 kip @
2.119 Fy, = 10.76 lb, Fyg = 10.76 lb, Foc = 26.83 Ib Is

2.121 Fp, = 156.59 N, Fy, = 156.59 N, Fo, = 0 3.73 A,


ll 0, A, Il= 10.53 kip @, B, = 0.47 kip ©
2.123 Fyz = Fac = 151.52 Ib,F= 108.19 Ib 3.75 Ae = 15.11 kip ©, A, = 7.67 kip ©, Fz = 3.54 kip ‘X
3.77 = 43.79 kN 7, B, = 41.54 KN ©, B, = 26.15kN@
CHAPTER 3 Ji
A

3.79 A,
= 86.2 lb ©, A, = 342.3 lb ©, My = 9715 lb-in @
3.1 M, = 1200 N-m ©)
3.81 Ay = 18KN.GiA, =25kN@)f.. = 22.5kN&
3.3. M, = 600 lb-ft @ 3.83 F = 348/d
3.5 M, = 768N-m@Q) 3.85 T = 13.02 kN LN 10°, A, = 42.36 kN/Y 10°,
3.7. M, = 600 lb-ft @ B, = 31.5kN// 10°
3.9 M, = 571.5 lb-ftQ@, Mz = 450 lb-ft @ 3.87 A, = 0.9 kip ©, B, = 6 kip ©, B, = 8.9 kip ©
3.11 (a) F= 258.8 Ib 45° LN, (6) My = 500 lb-ft @ 3.89 T = 657.2 lb 39° LN, B, = 510.7 lb ©, B, = 129.2 lb ©
3.13 (a) Mo = 2.546 (—cos®@ + sind) N-m ©), 30 KN ©, A, = 0, A, = 12kN@
(b) 0 = —45°, Mo = 3.6N-m Q, = 49 kN (@),.a = 13m;A, = 0
(c) @ = 135°, Mo = 3.6N-m ©
8
Sh| = 30kip
Pal ©, E, = 50 kip , A, = 30 kip©
3.15 M, = 73.71NmQ
7. T= 600.7 N 2
F, = Wcos 8 bY, F = w( 2 PCOS o)
by,
3.19 M = 18kN-mO, F, = 9kNO, Fp = 9KIN® COs
Fy = 2W tan?©
3.21 M = —1229k (Ib-ft), F, = —189i + 283.6j (1b),
F, = 189i —283.6j (Ib) 3.99) sae fo |
3.23 F, = 10.19 kip ©, F, = 10.19 kip © tan @ + tana

meoe Ff,= 11:11 1b ©, F, = 11.11 1b ©


3.27. F=15kip &%,M = 45 kip-ft@ CHAPTER 4
3.29 F = 15kip XQ, M = 90 kip-ft @ 4.1 M, = —5000i — 3000k (lb-ft)

3.31 F = 13kip@,M = 78 kip-in@ 4.3 M, = 3i + 9j — 12k (lb-ft)

3.33 F = 2.5kip 53° at 2.988 ft fromA 45 My, = 18i — 9j — 19.5k (b-ft)


570 Answers to Odd-Numbered Problems

4.7 (a) M4 = 2940i — 7042k (lb-ft), (b) d = 10.9 ft 4.65 T, = 5.85 (KN), Tc = 4.4 (KN), A, = 5.09i (KN),
4.9 M, = —145.2j + 145.2k (N-m) A, = —3.85j (KN), A, = —0.46k (KN)
4.11 Mp = —6540i — 14424j (lb-ft) 467 a=1m,b=0
4.13 Mus = — 8i (kN-m) 4.69 A, = 0.88i (KN), A, = —1.27j (KN),
B, = 7.74j (KN), C, = —2.61i (KN),
4.15 Mag = — 159.5k (1b-ft)
C, = —3.01j (KN), C, = k (KN)
4.17 Mer = —3.55i — 5.33j + 5.33K (kKN-m)
4.71 A, = 2.9i (kip), A, = — 1.3 (kip),
4.19 Map = —1220i + 2035k (lb-ft) B, = —0.9i (kip), B, = 6.5k (kip),
4.21 T = 380.2 N C, = —3.7j (kip), C, = —3.5k (ap)
4.23 M = 300i + 450k (lb-ft) 4.73 Fo = 13.91 kip (1), F, = 2.72 kip (T),
4.25 M = —8.5j — 14.5k (N-m) A, = 651i (kip), A, = —4.71j (kip),
4.27 M = —12i + 31.97j — 25.97k (KN-m) B, = 1.38i (kip), B, = 11.28j (kip)
4.29 F = —400j (N), M = 200i — 400k (N-m) 4.75 Fo = 18 kip (T), A, = 9i (kip),
A, = —11.9j (kip), A, = —k (kip),
4.31 F = —50j (kN), M = —20i + 15k (kN-m)
B, = —2i (kip), B, = 13.9j (kip)
4.33 F = 42.43i — 42.43] (Ib),
4.77 F, = 18 kip (T), A, = 7i (kip),
M = 42.43i + 42.43j — 174k (lb-ft)
A, = 2j (kip), A, = 11k (kip),
4.35 R = 150i (Ib), Mpg, = 800j — 1800k (lb-ft) M,, = —29i (kip-ft), M4, = —220j (kip-ft)
4.37 R = —0.75i + 2.6j + 1.3k (kip),
4.79 A, = 0, A, = —1.75j (KN), A, = 3.25k (KN),
Mra, = 27.81 — 12.57j + 9.98k (kip-ft)
B, = 4j (KN), = —2.25j (KN),
4.39 T = 538.37 N, a = 346 mm, b = 277 mm (CoS SOs) k (KN)
Meg = —75.2j + 29.9k (N-m)
CHAPTER 5
4.41 R = 1]j (kip),x = 7.27 ft, z = 1.91 ft
4.43 = 9j (kip), a = 4.33 ft, R = 20j (kip) 5.1 Fe = BYP, yy = 18

4.45 R = 234.91 + 64.5j + 159.8k (Ib), + +1


Mer, = —3516i + 3835j — 2048k (lb-in.) fre eri | fin lca,
n+2 , 2(2n +1)
4.47 R = —8.86i + 15.86j + 4.17k (Ib), Beeline
Mer, = 55.141 + 75.12j — 43.98k (Ib-in.) 5.5 P= ; iy = ale
4.49 Resultant wrench:
5.7m = SL /9 = Sha?
R = 234.91 + 64.5j + 159.8k (Ib),
Me = —2507i —688j — 1708k (Ib-in.) at x = —5.27 in. and 5.9 x = 2/15 ft, y = 3 ft
z = 15.64 in. 5.11 x =L/2,y = 7/8
4.51 Resultant wrench: Sal de = 424 ine ye — 9 iin
R = —8.86i + 15.86j + 4.17k (lb), 5.15 <x = 2.661 ft, y = 4.565 ft
=< R = —13.24i + 23.73j + 6.25k (lb-in) Bal = IDL PSY = ee
at x = —3.17 in. and
z = —4.31 in.
4.53 A, = 0, A, = 400j (N), A, = 0, M4, = —200i (N-m), 5.19 ee en ee
My, = 9, My. = 400k (N-m) 3a 3a
4.55 ee= —87.5i (Ib), A, = 303.1j (1b), 5.21 x = (2R/6) sin 0/2, y = 0
A, = —151.6k (Ib), M4, = —2121i (1b-ft), D: 23920 yal 64
Ma, = —350j (lb-ft), M4, = 525k (Ib-ft) D205 = — inves Din
4.57 A, = 0, A, = 11.2j (KN), A, = —7k (kN),
Die) mci SS lett nye wlrOSontt
B, = 8.8j (kN), B, = 2k (KN)
ys) eS WW, 9)= SSO ain,
4.59 A, = —234.9i (Ib), A, = —64.5j (Ib), A, = — 159.8k (Ib),
M,, = 3516i (Ib-in.), D.o1 x = 215 iniyy — 20am,
My, = —3835j (Ib-in.), M4, = 2048k (Ib-in.) 5.33 x = 5.67 in., y = 9.08 in.
4.61 Tz = 5 KN,A, = Si (KN),A, = 0, 5.35 x =0,y = 203.54 mm
A, = —8.66k (KN), My, = 8i (KN-m), SF he SS OS i, yy = My sin,
My, = 4.39j (KN-m)
5.39 x = —0.135in., y= —0.101 in.
4.63 Tz Il= 5.04 KN, Tc = 3.79 KN, A, = 5.63i (KN),
= 4 SS
A, |= —1.47j (KN), A, = —0.46k (KN) 5:41 3a? + b),y = 4b/30
Answers to Odd-Numbered Problems 571

5.43 =O) ye=0405.in: 5.125 (R sin? 6/2)/0


5.45 = 0, ¥= 121.03 mm 5.127 R/a
5.47 = 0.577a, y = 0.691a 5.129 2H/ 3
5.49 T16L 5.131 H/3
5.51 = 8.291 in., y = 6.531 in. 5.133 x = 50 mm, y = 165.02 mm, z = 100 mm
5.53 = 162.5 mm, y = 37.5 mm Guilkk) i = 0, y = 233.93 mm, z = 0
5.55 OO =) 3.17 in, y= 5.71 in:
ORT
MI
HE
HL
EM CHG, ap = 0.837 in., y = —9.41 in., Il oS

Ril + cosa + 2 sin a) 5.139 x = 112.5 mm, y = 35.0 mm, NI


NI 190.0 mm
5.57
macae ’
5.141 x= Z = 0, y = 3b/8
R(sin a — 2 cosa + 2) 5.143 x= z=0,y = R/2
<I |
, 2a + 2) §.145 x= — 35.63 mm, Zz = 59.87 mm
5.59 = 164.51 mm, y = 75.91 mm §.147 x= 48.24 mm, y = 136.24 mm, z = 71.31 mm
5.61 |ct
ad = 0, y = 381.75 mm
5.63 32.48° CHAPTER6

5.65 V3/2 6.1 BCi s a two-force member


5.67 (a) tH?L/7, (b) 20HL?/5 6.3 CD and EF are two-force members
5.69 (a) 2mR7(1 — cos a)/3, (b) 27R? sin a/3 6.7 BH, CH, CI, and FL are zero-force members
5.71 V = 84.84 in®, S = 96.53 in? 6.9 CK , DK, EO, FL, GL, GM, and HM are zero-force members

5.73 33,727.6 mm? 6.11 Fis 500 lb (T), Fyc = 366.7 lb (T), Fac = 833.3 Ib (C)
5.75 1.001 m? 6.13 Fy; = 2.414P (C), Fic = 1.414P (©), Fec = 2.613P (T)
5.77 0.464 m3 6.15 Fis = 100 KN (C), Fop = 0, Fac = Fep = 50 KN (©),
5.79 0.0178 m? Fap = 111.8 kN (T)
5.81 945.22 in? 6.17 Fyc = 242.3 kN (T), Fec = 323.08 kN (T),
= Fgp = 288.1 KN (C), Fep = 210 KN (T)
5.83 1/2R Fan

5.85 (ma*/6)[(1 + 4L7/a?)?/? — 1] 6.19 Fyc = Fen = SLE ote


2 sin? 6 tan 0
5.87 Ry = Wol @, My = wol?/2 0)
5.89 Ry = WoL/2®, My = WoL?/6 @ Fi, = z 50
Fp
> tan :
5.91 Re=twel/2OoM, =w,L7/3 ©)
5.93 R, = 14kN@),Rg = 34kN@® Fgc ~ sinmt ~ tan?6
5.95 R, = 80 lb @, M, = 480 lb-ft © 6.21 Fas = For = 40 KN (T), Fac= Fep = 53.33 KN (T),
5.97 R, = 1499.6 lb ®, Rg = 3000.4 Ib @ Fg = Foy = 40 KN (©), Fur = Fey = 50 KN (OC),
5.99 Ry, = WoL/t ®, My = 3WoL?/4a0 @ Fpr = Fp, = 30 KN (T), Fag = Fog = 16.67 kN (C),
5.101 R, = 4899.85 lb ®, M, = 13,063 lb-ft © Fog = 20 kN (T)
6.23 Fyc = 66 kip (T), Fog = 18 kip (T), Fer = Frg = 6 kip (C),
5.103 Ry = 4woL/t? ©, Rp = Um — 2)WoL/T? O
Fep = 48 kip (C), Fog = 12 kip (C), Fug = 30 kip (T),
5.105 R, = 8000 lb ©, M, = 48,000 lb-ft © Fgc = 42.43 kip (C), Fop = 25.46 kip (T),
5.107 R, = —2.33 kN ©, Rp= 26.33 kN ® For = 25.46 kip (C), Fag = 8.49 kip (T)
5.109 1081.8 Ib ©) 6.25 9 Fyc = Fp, = 6.71 kip (©), Fap = 2.12 kip (T),
5.111 R, = 94.3 kN, Rg = 108.48 kN Fep = 2.12 kip (©), Fup = 4.5 kip (T), Fop = 6 kip (O),
5.113 3.757 ft For = 3 kip (C)
6.27 Faz = Feo = Fep = Fog = 30 kip (T),
5.115 18.05 ft
Pap = Frg = 35 kip (C), Fry = Foy = 23.34 kip (C),
5.117 C, = 12,480 lb/ft, C, = 30,000 Ib/ft, Mc = 228,000 lb-ft
For = Fog = 0, For = Fog = 11.66 kip (C),
5.119 [(2N + 1)/(2N + 2)]L Fou = 12 kip (T)
3 6.29) Fuc = Fppn = 72.5 KN (1), Fag = 78.7 KN (OC),
5.121 ae
Fr = 64.6 KN (C), Fer = 92.9 KN (C), Frg = 78.8 KN (©),
5.123 2.728 in. Fep = 36.8 KN (T), Fog = Fog = 44.9 KN (T),
= For = 14.14 KN (C)
Answers to Odd-Numbered Problems

es)S I= 15 kip (T), Fo, = 5 kip (T), Fy, = 15 kip (C) Member CE: C, = 1.33 kip ©, C, = 1.43 kip ©,
oo i) = 90 KN (C), For = 87.5 KN (C), For = 142.5 KN (T)
| D, = 2.05 kip ©, D, = 1.43 kip ®, E, = 3.38 kip ©
Foe = Fe 0; Fee kip 6.69 Member AC: A, = C, = 0, A, = 13.75 kN ©,
For = 0, Fer = 2683.3 lb (T) B, = 38.75 INQ, C, = 25kN@
Member CDi 0 = D, = 0,C, = Dy = 25kN®
Fry = 5.1 kip (C), Fug = 3.2 Kip (©); For = 5.25 kip (T) Member DF: D, = 0, D, = 25 KN©, E, = 20kN@,
Fop = 162.9KN (C), Fy, = 182.1 KN (T) F, = 40 kN)
Fop = 11 kip (©), Fog = 1.03 kip (C) 6.71 Member AB: A, = B, = B, = 0, A, = 12 kip©
Fox = 35.36 KN (C) Member BD: B, = B, = D, = D, = 0,C, = 24 kip©
Frg = 36 kip (C), Fe, = 36 kip (T) Member DF: D, = D, = 0, E, = 22.5 kip @),
Fis = 1.22 ip (©), Fy Fac = OI kip: F, = 7.5 kip©
= 4.52 kip (T), Fyn = 7.98 kip (C), Fep = 1.73 kip (C) 6.73 Rod A: F, = 166.7N
Member BC: A, = 83.4N ©, A, = 1444NQ,
Fan = For = Fou = For = Fer = Freq — 0,
Ba 83.4N ©, B, = 194.4.N Q)
es Q i fe!=
= Fog = 2 kip (T), Fan = Fog = 3 kip (©),
Member BD: A, = §3.4NO, A, = 1444N0,
Fao = Fog = 3.54 kip (C); Fan = Fee = 0 54 kip Ce
Foe = Fcp= 0.71 kip (T), Fey = 6 kip (C) B, = 83.4N©, B, = 194.4 N @
6.53 Member AC: A, = 0.88 kip ©, Ay = 5.25 kip®, 6.75 Member AF: A, = 10kNO, A, = 2kNQ,
Cc. = 0.88 kip ©, (es sem) CG = 13kNO,C, = (0,/2, = 3iNO,F, = 2kN®
Member BC: Fac = 1.96 kip (C)
Member BE: B, ='10kKNO,B = 2 kN(D,
= 13 KN©, D, = 0,E, = 3kKNO,E, = 2kIN@®
6.55 Member AC: A, = 22.5 kN ©, A, = 15 kN ©, Member CD: Feop = 13 KN (T)
Co = 75kN©,C, = 15kN@). Members EG and FG: Frg = Frg = 3.6 KN (T)
Member BC: B, ST GWOu 15kN ®),
C, = 7.5KN ©, C, = 1SKN 6.77 BoltA:F, = 2697N
Member AG: A, = 0, A, = 2697 N ©, B, = 3446N ©,
6.57 Member AC: A, = 174.1 KN ©, A, = 274.1 KN ®, B== 4067 NQ, poe= 3446 NO, E, = ‘378N@
C, = 174.1 KN ©, C, = 74.1kN© Member CF: a 3446 N ©, cy = 1370NQ,
Member BC: B, = 25.9
KN ©, B, = 75.9
kN ©, = 3446 N ©, D, = 1378NO
C, = 174.1
KN ©, C, = 74.1 kN ®
seabetDE: Fpg = 3712 N(©)
MemberlG4) =00045 iu = 5.9kip ©,
6.79 Member AC: A, = 0, A, = 2495 Ib ®, M, = 32111 lb-ft ©,
C, = 9.55
kip ©, C, = 15.9 kip ()
B, = 8028 lb ©, B, = 9692 Ib ©, C, = 8028 lb ©,
Member BC: B, = 9.55 kip ©, B, = 31.9 kip ©, C, = 7333 Ib ©
G = 9.55 kip ©, C, = 15.9 kip @
Member BD: B, = 8028 lb ©, B, = 9692 lb ,
6.61 Member AC: A, = 15 kip ©, A, = 8 kip®, De 3028 Ib ©,D, = 9575 Ib
By = 28 kip, C, = 0,C, = 10 kip @ Member CE: C, = 8028 Ib ©, C, = 7333 Ib ©,
Member CD: C, = 0,C, = 10 kip ®, D, = 10 kip© bh = 8028 Ib ©, D, = 9575 lb @
6.63 Member AC: A, SuctOul = 48 kip ©, 6.81 Member AC: B, = C, = 0, B, = 375 lb @, C, = 250 lb ©,
C, = 0,C, = 48 kip ® = 125 lb®
Member CE: C, = 0,C, = 48 kip©, Member CE: C, = E, = 0,C, = 250 lb ®, D, = 750 Ib ©,
= 1125 kip Q,E, =105kn@ E, = 500 lb ©
6.65 Monee = B, = 0,A, = 500 lb @, Member EG: E, = 0, E, = 500 Ib ©, F, = 1500 lb @
M, = 2500 lb-ft ©,B, = 1000 Ib ©, D, = 1000 lb ©, 6.83 M = 3618 lbin.@
D, = 500 Ib @
6.85 M = 200NmQ
Member BC: Fac= 1000 Ib (C)
Mente CE: C, = 1000 lb ©, C, = 0, D, = 1000 Ib ©, CHAPTER7
D, = 500 Ib @
7.1 (a) V = 1 kip,M = 24 kip-ft
6.67 eats: A, = 3.38 kip©, A, = 3 kip ©, (b) V = 6 kip, M = 24 kip-ft
B, = 0.95 kip ©,B, = 1.43 kip @, C, = 4.33 kip ©,
7.3. V=wL/6,M = wL2/9
C, = 1.57 kip ©
Me B, = 0.95 kip©, B, = 1.43 kip ®, 7.5 V=0,M = 3wL?/32
= 2.05 kip ©, D, = 1.43 kip ©, F. = 3 kipO, 7.7 V= —P/6,M = 5PL/12
rate 7.9 V = —wL/8,M = —3wL?/32
Pulley C: C, = 3kip ©, C, = 3 kip ® 7.11. V = —667 lb,M = 22,000 lb-ft
Answers to Odd-Numbered Problems

7.13 V =0,M = —8kipft 7.93 427.86kN


7.15 V = -wL/4,M = —wL?/32 7.95 Tmax = 11,850 1D, Trin = 10,000 Ib
7.17 V = 3.9 kN, M = —7.95 kN-m 7.97 484m
7.19 = —5500 lb, M = —15,000 lb-ft 7.99 (a) 707.6 ft, (b) 2972 Ib
7.21 V = 4/9 wo, M = —14/81 woL? Sr; he 1
7.23 V = 8000 lb, M = —32,000 lb-ft 7.101
.101 L/H = 2—-— 7sinh (;
1 {| 5 =)
—___

7.25 V=0,M=0 7.103 28.3 ft


7.27 = —156 lb, M = 2064 lb-ft
CHAPTER 8
7.29 V, = 6 kip, M. = 28 kip-ft
7.31 Vio = 0, Mz = 3wL?/32 8.1 2.89 kN
7.33 V, = —wL/8,M, = —wL2/8 8.3 45.54 Ib
7.35 V, = 7333 Ib, Ms 333» = 22,220 Ib-ft 8.5 ayF
7.37 My = Mc = —wl2/2 8.7 106.06 Ib
7.39 V, = —500 Ib, M, = —55,000 lb-ft 8.9 486.36 N
7.41 V = Wox?/2L — Wox + WoL/2,M = Wox?/6L — wox?/2 8.11 55.17°
+ WoLx/2 — WoL?/2 (x = 0 at A) 8.13 1.5 kN
2 3 8.15 740.2 lb
7.43 V, = Sat, - -(2e + a 8.17 0.125

7.45 = WoL/4, Mp = —WoL?/24 8.19 49.7 Ib


8.21 80 1b
7.47 ooha5wL/4,M, = —SwL2/12
7.49 Me = —1400 lb-ft, V; = —140 Ib 8.23. 2.177 KN

7.51 V = Psin ®@,M = PRcos 8


8.25 199.81 Ib-ft
7.53 x = L/2 — L'/4, Max = W/4L(L — L'/2)? 8.27 0.5 tan (6/2)
7.55 8.29 68°
Vp = WoL/6, Mg = —WoL?/27
8.31 38°
7.57 Atx = 4.9 ft, V = 0,M = 6532 lb-ft (x = 0 at A)
7.59 8.35 a = tan! (p,/sin @)
Atx = L/2, M = woL?/n2, V = 0
8.37 2.93 kN
7.61 V = 2Lwo/m — (2LWo/7) sin (11x/2L), M = 2Lwox/m +
(4L?wo/17) cos (1x/2L) — 2L2wo/a 8.39 69.53 Ib

7.63 ee SMS, te 8.41 6 = sin7 (aaa


37,
=a)
Tay Peay Pets Ohee 1
8.43 3.07 kip
Weide

(n+ In + DL" 8.45 19.75°


7.65 Vz = w,L/3, Mg = —4w,L?/27 8.47 210N
7.67 Vz = 0, Mz = 2666 lb-ft 8.49 53.81 Ib
7.69 Vo = —581 lb, Mc = 25,963 lb-ft 8.51 600N
7.71 Mz, = 75 kN-m 8.53 49.96 lb
7.73 Mc = 31.39 kKN-m 8.55 6.81 kN
7.75 Pp 1968)1 Tne 2177 5.101b, Tn — 1830. 11b 8.57 2864.8 Ib
7.77 Typ = 4015.35 Ib, Tye = 3464.9 Ib, Top = 3586.12 Ib 8.59 6.24 kN
7.79 485.7 Ib 8.61 185.16 lb
7.81 Poets AT eave T i= 5. 13 101d 8.63 52.58 Ib-in
7.83 Tmax = 107,480 Ib, Tin = 84,054 Ib 8.65 58.69 Ib-in
7.85 1200 N, x= 19.84 m 8.67 190.4 Ib-in.
7.89 (100L? VL? + 9H?)/6H 8.69 0.192 kN
7.91 5 ft 8.71 12.49N-m
574 Answers to Odd-Numbered Problems

8.73 106.95 lb 5 5
9.29 = /(=-U4k= [=
8.75 0.711 a 21 eee 13 i
8.77. — 177.15 lb 48 /16
8.79 136.85 N 9.31 k=
&: /—ft,.k,
5 7 = /—ft
567
8.81 8439.6N
8.83 1, = 371b,T>, = 77 lb 9.33 k, = V2/3,k, = L/V2[V1 — (4/n?)]
8.85 2560 Ib-in. 9.35 k, = 97.55 in. k, = 4.27 in.
8.87 3733.3 Ib-in. 9.37 (a) 13.822 ft*, (b) 5.3 ft
Ree 9.39 (ene (b) 1.384 ft
aso 7 He (Ai = ; 105 kee
3 sin 0 \R? — R3
9.41 (a) 17.12 X 1077 m*, (b) 0.0565 m

8.91 3pbs (# = “| 9.43 1, = BH?/36, 1, = B°H/36


4 \Ri — R3 9.45 1, = 19LH?/480, I, = 8HL3/175
8.93 1052.6 Ib-in. 9.47 1, = 0.1098R4,/,, = 1R*/8
8.95 w.Pr 11
8.99 0.353 Lee Di SI oer a
8.101 327.73 N 9.51 1, = ab? (m/16 — 4/9m), I, = a@b(m/16 — 4/97)
8.103 T = 266.78 lb, T, = 233.12 lb, T, = 266.88 Ib 9.53 1,, = 0.507 in*, J, = 0.0533 in*
8.105 0.0261 m 9.55 I, = 38.57 X 10° mm‘, k, = 69.43 mm
8.107 7.04N 9.57 I, = 329.69 x 10° mm4, k, = 96.08 mm
8.109 6.36 in. 9.59 1, = 36.667 X 10° mm‘, k, = 67.7 mm
CHAPTER 9 9.61 1, = 1334.23 x 10° mm‘, k, = 193.3 mm
aie 9.63 110.95 in*
9.1 5GaeD 9.65 636.43 in*
9.3 BH3/12 9.67 4.785 x 10° mm*
; 9.69 87.42 in*
AE ee 9.71 149.33 in®
9.7 256/45 ft? 9.73 2.66 X 10° mm*

99 4L/9n 9.75 3778.8 in*


9.11 258.84 x 104 in’ 9.77 1, =91321.23in4 1, = 13292
i HL? 9.79 1, == 1376 in’,
14
I, =
an in
+44

Zales 9.81 1, = 197 /5, Igy = —- Lb?


HB 135
OS 5 Ceo 9.83 /,, = 112.5 x 10°mm*4,/,,, = 12.5 x 10° mm*
9.17 < 13/4 9.85 1, = H7/12)17,, = = HL
a ae ’ 8 a= 21.6 X 10° 6 mm 4

9.21 9 L3/m — 4L3/n3 9.91 0


14,896 . 9.93 J, = 57.61 x 10° mm‘, @,, = 61.48"
22 ie I,, = 6.13 X 10° mm, a,, = —28.53°
05 On 9.95 J, = 91.66 x 10° mm* a, = 0°
9.25 k= nfLt *) Jk = (+) I,, = 38.54 X 10° mm‘, a,, = 90°
ees e gS I = 2160 in’, a, = 90°
9.97 1,
9.27 k=
NE te | = OBOE 2 i,; 1584 int, a,, = 0
st is o

: 6 9.99 I, 142.805 in*, Oe aah EWP


Il 55.562 in#, Baw PS Mic
I,
Answers to Odd-Numbered Problems 575

9.101 = 636.43 in’, a, = 0° 10.5 M = PLsina


= 149.33 in’, c= 90° 10.7 M = 2PL sin @
9.103 = ORIEBS OS 10° mm, a, = 0 _ 448P(14 + 12 cos 6)
10.9 M=
v
= 2.65 X 10? mm’%, a, = 90° 340 + 336 cos 8

9.105 10.11 P = QL/h


10.13 P = 889.5NQ)
9.107 10.15 Fep = 11713 Ib LS 50.2°
WAIN
Mine
el
I>
10.17 R, = 166.7N[{S 30°
lie Ie? 10.19 A, = 25kN@®
9.109
eee bm 2 ny ;) 10.21 C, = 297.9 kN @
9.111 m(H?/6 + L*/2) 10.23 C, = 92.85 kN @)
9.113
4037
aaeu.” 10.25 Bo=0
10.27 A, = 0.88 kip ©)
9.115 56.11 X 107? lb-ft-s? 10.29 M = 2PL sin 0
9.117 1.754 kg-m? 10.31 a = 0°, stable; a = +82.02°, unstable
9.119 I, = 112.9 X 1073 slug-ft?, J, = 115.2 X 107? slug-ft? 10.33 a = 5.81°, stable; a = 81.24°, unstable
9.121 9.992 slug-ft? 10.35 a = 0°, stable
9.123 59.04 slug-ft? 10.37 a = 0°, unstable; a = +85.7°, stable

CHAPTER 10 10.39 Konin = QOL/2

10.1 A, = 0.88 kip © 10.41 a = 90°, unstable; a = 211.78°, stable; a = —31.78°, stable

A, = 45kN © 10.43 a = 0°, unstable; a = 75°, stable


10.3
INDEX

Acceleration Body; see also Applied forces; External forces; Centroid—Cont.


of particle, 4—5S Homogeneous bodies; Partially of two-dimensional area, 233—37
with particle equilibrium, 33 constrained bodies; Rigid body of two-dimensional volume, 230
Angle of kinetic friction, 415 determination of unknown forces in, 35-36 Coefficient of kinetic friction, 414
Angle of static friction, 415 force of, 3 Coefficient of rolling resistance, 441
Applied forces or loads, 80 homogeneous, 230 Coefficient of static friction, 414, 421
Areas mass of, 4 ; Collinear vectors, 7
of a body, 230-32 radius of gyration of, 5|00—501 Commutative law of vector addition, 20, 52
centroid of, 233-35 two-dimensional, 82—83 Components
centroid of composite, 239-41 Bridge truss, 294—95 of force vector, 26—27
Mohr’s circle, 496—98 rectangular, 27
moment of inertia of, 463-66 Cables; see also Catenary cables; Flexible Composite area, 239-42
moment of inertia of composite, 479-83 cables; Parabolic cable Composite line, 243-44
principal moment of inertia, 492—95 applications and types, 384—96 Composite volume, 259-63
product of inertia of, 485—87 as connections, 38—39, 53 Compression, axial, 296, 324
product of inertia of composite, 489—90 subjected to own weight, 394—96 Concentrated load (of cable), 384-88
second moment of, 464 Cartesian unit vectors, 30, 32, 33, 52 Concurrent forces; see also Varignon’s theorem
Associative law of vector addition, 24, 52 cross products of, 88-90 in force-couple system, 111—12
Axial compression, 296, 324 dot products of, 170—71 resultant of, 49, 53
Axial forces moments in, 92—93 in three dimensions, 40—41
of beam, 372 Catenary cables, 396, 398 in three-force member, 123
at connections, 289 Center of gravity, 228-30 in two dimensions, 19—21
Axis; see Moment; Parallel axis theorem Center of mass, 229-30 using Varignon’s theorem, 93-94
Centroid Concurrent force systems, 133
of area, 231 Concurrent vectors, 7
Basic truss element, 298; see also Simple space of composite lines, 243-44 Connections; see also Supports and connections
trusses; Simple truss of composite two-dimensional areas, axial forces at, 289
Beams; see also Bending moment; Shear force 239-42, 248 cables as, 38-39, 53
definition and types, 370—71, 397 of generating area, 246 internal forces at, 287-93
distributed loads on, 248-49 of generating curve, 245 Conservative system, 538—40; see also
loads on, 373 of line, 238-39 Nonconservative system
Bearings; see Journal bearings; Thrust bearings of three-dimensional composite volumes, Constraint forces, 36; see also Improperly
Belt friction, 431-35 259-63 constrained bodies; Partially constrained
Bending moment, 372, 378-80 of three-dimensional volume, 255-59 bodies

576
Index 577

Conversion of units, 13-14 Equal vectors, 7 Force systems—Cont.


Coplanar force, 19 Equations of equilibrium coplanar, 113, 110-18, 303, 331
Coplanar force systems for three-dimensional rigid bodies, 194 parallel, 113
forms of, 113 for two-dimensional rigid bodies, 119-22, three-dimensional, 164, 181—93, 210
of frames, 331 125, 138-47 two-dimensional, 82—83, 87, 91, 111-13
nonconcurrent, 110—18 Equilibrium; see also Neutral equilibrium; zero resultant, 113
of plane truss method of joints, 303 Reaction; Stable equilibrium; Unstable Force triangle, 20, 22, 23
Coplanar vectors, 7, 8 equilibrium Force vectors
Cosines; see Direction cosines; Law of cosines of particle, 33-35, 53 components replace resultant, 26
Coulomb friction, 412 of particle in three dimensions, 51 direction cosines of, 42—44
Couples; see also Resultant couple of plane truss, 300-301 replaced by resultant, 26
components and resultants of, 177 of rigid body, 119, 147 representation of, 47—48
defined, 100, 146-47 stability of, 545 as sliding vectors, 164
equivalent, 102—5 total potential energy in, 542—43 Frames
moment of, 100-102, 106, 147 External forces analysis of, 331-50, 352
work of, 530-32 applied on beam, 372 defined, 330
Cross products; see also Mixed triple product defined, 80 forces in, 331—34
distributive property of, 88, 89-90 function of, 287 Free-body diagram
moment of a couple as, 101 potential energy of, 538-40 of particle, 35-38, 53
of vectors, 87—90 External loading (sign convention), 372 of three-dimensional rigid body, 201
of two-dimensional rigid body, 134-37,
138-46
Deformation Fixed vectors, 6, 81 Free vector, 6
of rigid bodies, 4, 79-80 Flexible cables, 370, 397 Friction; see also Belt friction; Disc friction;
with rolling resistance, 440 Fluid friction, 412 Dry friction; Fluid friction; Kinetic
Direction cosines, 42—44 Force; see also Axial forces; Components; friction; Static friction; Wedges
Disc element, 256—57 Concurrent forces; Constraint forces; angle of static, 414-15
Disc friction, 436—37 Coplanar force; Couple; Free-body of belts, 431-34
Displacement; see also Potential energy; diagram; Hydrostatic force; Resultant mechanics of dry, 412—15
Translation; Virtual displacement force; Three-force member; Two-force problems of dry, 415-17
of a body, 528-29 member; Work Frictional resistance
work of couple in, 531 collinear, 123 of journal bearings, 438
Distributed loads; see also Linearly varying components and rectangular components, of thrust bearings, 436—37
distributed load; Uniformly distributed 26-29 Frictionless pulley, 39, 420
load concentrated, 3
on beams, 248-49 concurrent and nonconcurrent, 123
Geometric instability, 132—33
on cables, 389-92 defined, 3
Geometric method, 378-84
relation to shear force and bending moment, external and internal, 80
Gravitation; see Newton’s law of gravitation;
378-80 in free-body diagram, 35—38
Universal constant of gravitation
Dot product or scalar product; see also Mixed moments of, 83
Gravitational force
triple product reactive, 34, 36
acting on a body, 9
applications of, 171-72 resultant, 19, 111
between two particles, 5—6
distributive and commutative properties of, resultant of concurrent, 49-50
Gravity (center of), 228
170, 172 in truss members, 303-9
Gyration; see Polar radius of gyration; Radius
Dry friction; see also Coulomb friction; unit vectors representing, 45—46
of gyration
Impending motion as vector quantity, 19
defined, 412 Force-couple system, 106-7, 147; see also
mechanics of, 412 Resultant force-couple system Homogeneous bodies, 230, 263
problems of, 415-17 reduction in three-dimensional system of Hydrostatic force, 250—52
forces, 182-83
reduction of a force system to, 110—12
Elastic potential energy; see Potential energy reduction to single force, 107-10 Impending motion
Energy, 538; see also Conservative system; Force systems condition of, 416—17, 419
Potential energy concurrent, 133 point of, 413, 415
578 Index

Impending-slip condition, 432, 434—35 Loads—Cont. Moment vectors—Cont.


Improperly constrained bodies on a beam, 373, 397, 464 couples with same, 102—5
three-dimensional, 200-201, 210 on cables, 384—92 of resultant couples, 178
two-dimensional, 133 deformation under, 80 in three-dimensional force systems, 164
Indeterminacy, static, 130 frames as supports for, 330 Motion; see Impending motion; Newton’s laws
Instability; see also Improperly constrained sign conventions for internal and external, of motion
bodies; Partially constrained bodies B72 Multiforce members (in frames and machines),
geometric (of three-dimensional bodies), on trusses, 294 330
199-201, 210
geometric (of two-dimensional bodies), Machines; see also Wedges
Negative vector, 7
132-33 analysis of, 331-50, 352
Neutral equilibrium, 545—46
static (of plane truss), 302 defined, 330
Newton’s law of gravitation, 5—6
static (of space trusses), 326 Mass
Newton’s laws of motion, 4—5, 80, 287
static (of three-dimensional bodies), defined, 4
Nonconcurrent forces, 123
198-99, 210 moment of, 465, 500
Nonconservative system, 538
static (of two-dimensional bodies), 130—34 moment of inertia of, 465, 500
Integration unit of, 13
to locate centroid of a two-dimensional Mass moment of inertia, 500 Pappus-Guldinus theorems, 244—46
area, 233-37 of composite bodies, 508-11 Parabolic cable, 390—92, 397
to locate centroid of two-dimensional line, by integration, 503-7 Parallel axis theorem
238-39 Mechanics for moment of inertia of a mass, 501-3
Internal forces; see also Bending moment; defined, 2 for moment of inertia of an area, 475-76,
Shear force of deformable bodies, 79 479-83, 487, 489
in a beam, 372, 397 of dry friction, 412-15 Parallel force systems, 113
at connections, 287—93 historical background, 2—3 Parallelogram law of addition, 8, 20—21
defined, 80, 287, 351 Method of joints Partially constrained bodies
of frame, 331 for plane truss analysis, 303, 351 three-dimensional, 198—99
work of, 530 for space truss analysis, 326—27 two-dimensional, 130—33
Internal loading (sign convention), 372 Method of sections Particle; see also Free-body diagram
Internal moment (of beam), 372 to draw shear-force and bending-moment defined, 3
diagrams, 373-78 in equilibrium, 33-35, 51
Joints; see also Method of joints for plane truss analysis, 315-24, 351 subjected to force, 79
in frames, 330, 331 for space truss analysis, 326—27 Pascal’s law, 250
of machines, 330 Mixed triple product, 172—75 Pitch, 429
method of (for trusses), 303—9, 326-27 Mohr’s circle, 496-99 Planar body, 82
Journal bearings, 438-39 Moment; see also Internal moment; Torque Plane area, 244
of a couple, 100—102, 106, 147 Plane curve, 244-45
defined, 83 Plane line, 244
Kinetic friction, 414; see also Angle of kinetic
determinants of, 94—95 Plane truss; see also Basic truss element
friction; Coefficient of kinetic friction
of a force about an axis, 173, 209 analysis by method of joints, 303—15
magnitude and sense of, 83-86 analysis by method of sections, 315—24
Law of cosines, 8, 21, 22, 25 units of, 83 defined, 294
Law of sines, 8, 21, 22, 23, 25 Moment arm or lever arm, 83 nonrigid rectangular, 298
Lead angle, 429 Moment of a force, 173, 209 rigid triangular, 298
Linear elastic spring, 39, 540-41 Moment of inertia; see also Mass moment of statically determinate and indeterminate,
Linearly varying distributed load, 373, 464 inertia; Parallel axis theorem; Polar 300-302
Lines moment of inertia; Principal moment of unstable, 302
centroid of, 238 inertia (of an area); Product of inertia Polar moment of inertia, 472—74
centroid of composite, 243—44 of an area, 463-70, 475-77, 485 Polar radius of gyration, 473
two-dimensional, 232—33 of an area using Mohr’s circle, 496-99 Polygon rule, 24
Loads; see also Applied forces; Concentrated of composite areas, 479-85 Potential energy, 538—42, 548; see also
load; Distributed load; Linearly varying of a mass, 465, 500 Principle of minimum potential energy;
distributed load; Uniformly distributed Moment vectors, 90—93 Principle of stationary potential energy;
load of couples, 177 Principle of virtual work
Index 579

Potential energy—Cont. Resultants—Cont. Shell element, 257—58


elastic, 540 of nonconcurrent coplanar force systems, Sign conventions, 372
total, 542-43, 545-46, 548 110-18 Simple space trusses, 325
Principal moment of inertia (of an area), of nonconcurrent three-dimensional force Simple truss, 298-300
492-95 systems, 181-93 Single degree of freedom system, 546
Principle of minimum potential energy, from rectangular components, 31-32 SI units, 9-12
545-46, 548 of two forces, 19-23 Sliding vectors; see also Fixed vectors
Principle of stationary potential energy, using triangle rule, 24, 25 external forces as, 80—82
543-44, 548 Resultant wrench, 185, 210 moment and force vectors as, 164
Principle of virtual work Revolution; see Surface of revolution; Volume moment vectors as, 91—92
statement and function of, 532—37, 548 Right-hand rule use of, 7
in terms of total potential energy, 542—43 in three-dimensional force systems, 164 Space, 3
Product of inertia in two-dimensional force system, 87, 91 Space trusses, 296, 324; see also Simple space
of an area, 485-88 Rigid body; see also Improperly constrained trusses
of composite areas, 489-91 bodies; Partially constrained bodies; Spring constant, 39—40
Pulley; see Frictionless pulley Three-force member; Two-force Springs, 540; see also Linear elastic spring;
Pythagorean theorem, 28, 32 member Potential energy
concept of, 80 Square-threaded screws, 428-31
defined, 4 Stable equilibrium, 545-46
Radius of gyration; see also Polar radius of equations of equilibrium for, 119-22 Static determinacy
gyration in equilibrium, 119, 123-24, 147, 194 in analysis of reactions, 138-46
of an area, 470-71 external forces as sliding vectors, 80-82 of beams, 371
of a body, 500-501 forces acting on, 80 of frames and machines, 331, 351—52
Reactions or reaction forces statically determinate and indeterminate, of plane truss, 300—302
analysis of two-dimensional rigid body, 127-34 of space trusses, 325-26
138-46 translation and rotation of, 79 for three-dimensional bodies, 197—201, 210
defined, 34, 80 trusses as, 294 for two-dimensional bodies, 127—30, 135,
determining three-dimensional rigid body, two-dimensional, 146, 147 136, 137
201 Rigid-body mechanics, 2 Static friction, 412, 413; see also Angle of
exerted by supports and connections, 125 Rod element, 256 static friction; Coefficient of static
in free-body diagram, 36, 82 Rolling resistance, 440—41; see also Coefficient friction
by supports and connections, 194-97 of rolling resistance Static indeterminacy
Rectangular components Roof truss, 294, 296, 308 of beams, 371
of a force, 27-29, 52, 191-92 Rotation of plane truss, 301-2
of a force in space, 40-42 of couples, 100 of space trusses, 326
force vector in terms of, 47 of rigid body, 79 for three-dimensional bodies, 197-201, 210
resultant from, 31—32 support prevention for three-dimensional for two-dimensional bodies, 130—34
Resistance; see Frictional resistance; Rolling bodies, 195-97 Structures, 197—201; see also Frames; Internal
resistance support prevention for two-dimensional forces; Machines; Trusses
Resultant couple bodies, 125-27 Supports and connections
in three-dimensional force system, 183 virtual, 533 for beams, 370-71, 373-74
in two-dimensional force system, 111—12 work of couple in, 531—32 for three-dimensional rigid bodies, 194—97
Resultant force Rules; see Polygon rule; Right-hand rule; for two-dimensional rigid bodies, 125—27
in three-dimensional force system, 183, 210 Triangle rule Surface of revolution, 244
in two-dimensional force system, 111-13 Systems; see Conservative system; Force-
Resultant force-couple system, 112 Scalar product; see Dot product or scalar couple system; Force system;
reduction to resultant wrench, 185—86 product Nonconservative system; Resultant
reduction to single resultant force, 112-13, Scalar quantities, 6, 29, 529 force-couple system; Single degree of
183-85, 189-90 Screws; see Square-threaded screws freedom system
with rigid body in equilibrium, 194 Sections; see Method of sections
in three-dimensional force system, 183, Self-locking condition, 430
187-88 Shear force Tension
Resultants of beam, 372, 378-80 axial, 296, 324
of concurrent forces, 49—50, 53 sign conventions for, 372 on cables, 38-39
580 Index

Tension—Cont. Uniformly distributed load, 373, 390, 394 Vectors—Cont.


of flexible cable, 398 U. S. Customary units, 12 collinear, 7
supported by cables, 370 Units; see also SI units; U. S. Customary units concurrent, 7
Three-force member, 123 base, 9 coplanar, 8
Thrust bearings, 436 conversion of, 13-14 cross product of, 87—90
Time, 3 derived, 9 equal, 7
Torque, 83 of mass, 13 fixed, 6, 81
Translation; see also Supports and connections of measurement, 8—13 free, 6
of particle, 79 of moment, 83 moment vectors, 90—93
of rigid body, 79 supplementary, 9 negative, 7
support prevention for three-dimensional Unit vectors; see also Cartesian unit vectors sliding, 7, 80-82
bodies, 195-97 described, 29-30 Virtual displacement, 532—37
support prevention for two-dimensional representing a force, 45-46 Virtual work; see Principle of virtual work
bodies, 125-27 Universal constant of gravitation, 5 Volume
virtual, 533 Unstable equilibrium, 545—46 centroid of composite, 259-61
work of couple in, 531 centroid of elemental, 256—58
Transmissibility principle, 80-82, 106, 146 centroid of volume of body, 255
Triangle rule; see also Force triangle Varignon’s theorem, 93-94, 146 of revolution, 244—46
to determine resultant, 24 Vector addition; see Associative law of vector
in vector addition, 20 addition; Commutative law of vector
Wedges, 424—27
Trusses; see also Basic truss element; Bridge addition; Law of cosines; Law of sines;
Weight, 6
truss; Method of joints; Method of Parallelogram law of addition; Triangle
cable’s own, 394—96
sections; Plane truss; Roof truss; Simple rule
in free-body diagram, 36
truss; Space truss; Zero-force members Vector product or cross product, 87—90 Work; see also Principle of virtual work
analysis of, 296—98, 351 Vector quantities, 6, 19; see also Parallelogram
of couples, 530—32
defined, 294 law of addition
of a force, 528—29
formation of, 351 Vectors; see also Cartesian unit vectors;
of internal forces, 530
joints in, 303-9, 326-27 Collinear vectors; Concurrent vectors;
positive and negative, 529, 531-32
sections of, 315-20, 326—27 Dot product or scalar product; Equal
Two-force members, 123, 289—92 vectors; Fixed vectors; Free vector;
plane truss members as, 296 Negative vector; Right-hand rule; Unit Zero-force members, 306—8, 309
space truss members as, 324 vector Zero resultant force systems, 113
U.S. Customary Units and Their SI Equivalents
Quantity U.S. Customary Unit SI Equivalent

Acceleration ft/s? 0.3048 m/s?

Angular acceleration rad/s? 1.0 rad/s?

Angular velocity rad/s 1.0 rad/s

Area ls 0.0929 m?

Energy 1.356 J

Force 4.448 kN
4.448 N
Impulse 4.448N-s

Length 0.3048 m
25.4 mm

Mass lb mass 0.4536 kg


oz mass 28.35 g
slug (Ib-s?/ft) 14.594 kg
ton (2000lbs) 907.18 kg

Moment Ib +ft 1.356 N-m

Moment of inertia of in* 0.4162 x 10®mm/‘*


an area = 0.4162 x 10°m*

Moment of inertia of Ib - ft - s? 1.356 kg - m?


amass

Power 1.356 W

Pressure 47.88 Pa
6.895 kPa

Velocity 0.3048 m/s

Volume (liquid) 3.785 L

Volume (solid) 0.02832 m*


16.39 cm?

Work 1.3558 J
3.60 x 10°J
SI Units and Their U.S. Customary Equivalents
Quantity SI Unit U.S. Customary Unit

Acceleration m/s? 3.281 ft/s?

Angular acceleration rad/s? rad/s?

Angular velocity rad/s rad/s

Area m? 10.764 ft?

Energy J (Joule) 0.7375 Ib - ft

Force N 0.22482 Ib
kN 0.22482 kip

Impulse N-s 0.22482 Ib-s

Length m 3.281 ft
km 0.6215 mi

Mass kg 0.06854 slug

Moment N-m 0.7375 lb - ft

Moment of inertia of m* 8.63 x 10° ft*


an area

Moment of inertia of kg +m? 0.7375 Ib - ft- s?


amass

Power W 0.7375 lb - ft/s

Pressure Pa 1.45 x 10“lb/in?

Velocity m/s 3.281 ft/s |

Volume (liquid) i ~ 0.2642 gal


Volume (solid) m° Sorolutts
cm? 0.061 in®

Work J 0.7375 lb - ft
ISBN O-25b-hi45e-0
|7 90000

9"780256114522

45-346b61-01

You might also like