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

tensors+1:2 kinematics

The document is a textbook titled 'Continuum Mechanics of Solids' authored by Lallit Anand and Sanjay Govindjee, aimed at first-year graduate engineering students. It covers fundamental concepts in Solid Mechanics, including kinematics, balance laws, and constitutive equations, while also addressing modern multi-physics responses of materials. The book serves as a pedagogical resource for classroom teaching and self-study, providing a comprehensive yet pragmatic overview of Solid Mechanics.

Uploaded by

joffin2012
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)
16 views

tensors+1:2 kinematics

The document is a textbook titled 'Continuum Mechanics of Solids' authored by Lallit Anand and Sanjay Govindjee, aimed at first-year graduate engineering students. It covers fundamental concepts in Solid Mechanics, including kinematics, balance laws, and constitutive equations, while also addressing modern multi-physics responses of materials. The book serves as a pedagogical resource for classroom teaching and self-study, providing a comprehensive yet pragmatic overview of Solid Mechanics.

Uploaded by

joffin2012
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/ 73

OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

Continuum Mechanics of Solids


OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

Continuum
Mechanics
of Solids

Lallit Anand
Department of Mechanical Engineering, Massachusetts Institute of Technology,
Cambridge, MA 02139, USA

Sanjay Govindjee
Department of Civil and Environmental Engineering, University of California,
Berkeley, Berkeley, CA 94720, USA

1
3
Great Clarendon Street, Oxford, OX2 6DP,
United Kingdom
Oxford University Press is a department of the University of Oxford.
It furthers the University’s objective of excellence in research, scholarship,
and education by publishing worldwide. Oxford is a registered trade mark of
Oxford University Press in the UK and in certain other countries
© Lallit Anand and Sanjay Govindjee 2020
The moral rights of the authors have been asserted
First Edition published in 2020
Reprinted with corrections in 2023

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, without the
prior permission in writing of Oxford University Press, or as expressly permitted
by law, by licence or under terms agreed with the appropriate reprographics
rights organization. Enquiries concerning reproduction outside the scope of the
above should be sent to the Rights Department, Oxford University Press, at the
address above
You must not circulate this work in any other form
and you must impose this same condition on any acquirer
Published in the United States of America by Oxford University Press
198 Madison Avenue, New York, NY 10016, United States of America
British Library Cataloguing in Publication Data
Data available
Library of Congress Control Number: 2020938849
ISBN 978–0–19–886472–1
DOI: 10.1093/oso/9780198864721.001.0001
Printed and bound by
CPI Group (UK) Ltd, Croydon, CR0 4YY
Links to third party websites are provided by Oxford in good faith and
for information only. Oxford disclaims any responsibility for the materials
contained in any third party website referenced in this work.
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

Preface

Continuum Mechanics is a useful model of materials which views a material as being contin-
uously divisible, making no reference to its discrete structure at microscopic length scales
well below those of the application or phenomenon of interest. Solid Mechanics and Fluid
Mechanics are subsets of Continuum Mechanics. The focus of this book is Solid Mechanics,
which is concerned with the deformation, flow, and fracture of solid materials. A study of
Solid Mechanics is of use to anyone who seeks to understand these natural phenomena, and
to anyone who seeks to apply such knowledge to improve the living conditions of society. This
is the central aim of people working in many fields of engineering, e.g., mechanical, materials,
civil, aerospace, nuclear, and bio-engineering. This book, which is aimed at first-year graduate
students in engineering, offers a unified presentation of the major concepts in Solid Mechanics.
In addressing any problem in Solid Mechanics, we need to bring together the following major
considerations:
• Kinematics or the geometry of deformation, in particular the expressions of
stretch/strain and rotation in terms of the gradient of the deformation field.
• Balance of mass.
• The system of forces acting on the body, and the concept of stress.
• Balance of forces and moments in situations where inertial forces can be neglected,
and correspondingly balance of linear momentum and angular momentum when they
cannot.
• Balance of energy and an entropy imbalance, which represent the first two laws of
thermodynamics.
• Constitutive equations which relate the stress to the strain, strain-rate, and temperature.
• The governing partial differential equations, together with suitable boundary conditions
and initial conditions.
While the first five considerations are the same for all continuous bodies, no matter what mate-
rial they are made from, the constitutive relations are characteristic of the material in question,
the stress level, the temperature, and the timescale of the problem under consideration.
The major topics regarding the mechanical response of solids that are covered in this
book include: Elasticity, Viscoelasticity, Plasticity, Fracture, and Fatigue. Because the consti-
tutive theories of Solid Mechanics are complex when presented within the context of finite
deformations, in most of the book we will restrict our attention to situations in which the
deformations may be adequately presumed to be small—a situation commonly encountered
in most structural applications. However, in the last section of the book on Finite Elasticity of
elastomeric materials, we will consider the complete large deformation theory, because for such
materials the deformations in typical applications are usually quite large.
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

vi PREFACE

In addition to these standard topics in Solid Mechanics, because of the growing need for
engineering students to have a knowledge of the coupled multi-physics response of materials
in modern technologies related to the environment and energy, the book also includes
chapters on Thermoelasticity, Chemoelasticity, Poroelasticity, and Piezoelectricity—all within
the (restricted) framework of small deformations.
For pedagogical reasons, we have also prepared a companion book with many fully
solved example problems, Example Problems for Continuum Mechanics of Solids (Anand and
Govindjee, 2020).
The present book is an outgrowth of lecture notes by the first author for first-year graduate
students in engineering at MIT taking 2.071 Mechanics of Solid Materials, and the material
taught at UC Berkeley in CE 231/MSE 211 Introduction to Solid Mechanics by the second
author. Although the book has grown out of lecture notes for a one-semester course, it contains
enough material for a two-semester sequence of subjects, especially if the material on coupled
theories is to be taught. The different topics covered in the book are presented in essentially
self-contained “modules”, and instructors may pick-and-choose the topics that they wish to
focus on in their own classes.
We wish to emphasize that this book is intended as a textbook for classroom teaching or self-
study for first-year graduate students in engineering. It is not intended to be a comprehensive
treatise on all topics of importance in Solid Mechanics. It is a “pragmatic” book designed to
provide a general preparation for first-year graduate students who will pursue further study or
conduct research in specialized sub-fields of Solid Mechanics, or related disciplines, during the
course of their graduate studies.

Attributions and historical issues


Our emphasis is on basic concepts and central results in Solid Mechanics, not on the history of
the subject. We have attempted to cite the contributions most central to our presentation, and
we apologize in advance if we have not done so faultlessly.

Lallit Anand
Cambridge, MA
Sanjay Govindjee
Berkeley, CA
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

Acknowledgments

Lallit Anand is indebted to his teacher Morton Gurtin of CMU, who first taught him continuum
mechanics in 1975–76, and once again during the period 2004–2010, when he worked with
Gurtin on writing the book, The Mechanics and Thermodynamics of Continua (Gurtin et al.,
2010). He is also grateful to his colleague David Parks at MIT—who has used a prelimi-
nary version of this book in teaching his classes—for his comments, helpful criticisms, and
his contributions to the chapters on Fracture and Fatigue. Discussions with Mary Boyce,
now at Columbia University, regarding the subject matter of the book are also gratefully
acknowledged.
Sanjay Govindjee is indebted to his many instructors and in particular to his professors at
the Massachusetts Institute of Technology and at Stanford University, who never shied away
from advanced concepts and philosophical discussions. Special gratitude is extended to the late
Juan Carlos Simo, Jerome Sackman, and Egor Popov. Additional thanks are given to current
colleague Robert L. Taylor and countless students whose penetrating questions have formed
his thinking about solid mechanics.
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

Contents

I Vectors and tensors 1

1 Vectors and tensors: Algebra 3


1.1 Cartesian coordinate frames. Kronecker delta. Alternating symbol 3
1.1.1 Summation convention 4
1.2 Tensors 6
1.2.1 What is a tensor? 6
1.2.2 Zero and identity tensors 7
1.2.3 Tensor product 7
1.2.4 Components of a tensor 7
1.2.5 Transpose of a tensor 8
1.2.6 Symmetric and skew tensors 9
1.2.7 Axial vector of a skew tensor 9
1.2.8 Product of tensors 10
1.2.9 Trace of a tensor. Deviatoric tensors 11
1.2.10 Positive definite tensors 12
1.2.11 Inner product of tensors. Magnitude of a tensor 12
1.2.12 Matrix representation of tensors and vectors 13
1.2.13 Determinant of a tensor 14
1.2.14 Invertible tensors 15
1.2.15 Cofactor of a tensor 15
1.2.16 Orthogonal tensors 15
1.2.17 Transformation relations for components of a vector and a tensor
under a change in basis 15
Transformation relation for vectors 16
Transformation relation for tensors 17
1.2.18 Eigenvalues and eigenvectors of a tensor. Spectral theorem 17
Eigenvalues of symmetric tensors 19
Other invariants and eigenvalues 20
1.2.19 Fourth-order tensors 20
Transformation rules for fourth-order tensors 22
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

x CONTENTS

2 Vectors and tensors: Analysis 23


2.1 Directional derivatives and gradients 23
2.2 Component expressions for differential operators 24
2.2.1 Divergence, curl, and Laplacian 25
2.3 Generalized derivatives 27
2.4 Integral theorems 29
2.4.1 Localization theorem 29
2.4.2 Divergence theorem 30
2.4.3 Stokes theorem 31

II Kinematics 33

3 Kinematics 35
3.1 Motion: Displacement, velocity, and acceleration 35
3.1.1 Example motions 37
Simple shear 37
Elementary elongation 37
Rigid motion 37
Bending deformation 38
3.2 Deformation and displacement gradients 39
3.2.1 Transformation of material line elements 40
3.2.2 Transformation of material volume elements 41
3.2.3 Transformation of material area elements 42
3.3 Stretch and rotation 43
3.3.1 Polar decomposition theorem 44
3.3.2 Properties of the tensors U and C 45
Fiber stretch and strain 46
Angle change and engineering shear strain 47
3.4 Strain 48
3.4.1 Biot strain 48
3.4.2 Green finite strain tensor 48
3.4.3 Hencky’s logarithmic strain tensors 49
3.5 Infinitesimal deformation 49
3.5.1 Infinitesimal strain tensor 49
3.5.2 Infinitesimal rotation tensor 51
3.6 Example infinitesimal homogeneous strain states 52
3.6.1 Uniaxial compression 52
3.6.2 Simple shear 53
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

CONTENTS xi

3.6.3 Pure shear 54


3.6.4 Uniform compaction (dilatation) 55
3.6.5 Infinitesimal rigid displacement 56
3.7 Volumetric deviatoric split 56
3.7.1 Volume changes 56
3.7.2 Strain deviator 57
3.8 Summary of major kinematical concepts related to infinitesimal strains 58
APPENDICES 59
3.A Linearization 59
3.B Compatibility conditions 60
3.B.1 Proof of the compatibility theorem 63

III Balance laws 67

4 Balance laws for mass, forces, and moments 69


4.1 Balance of mass 69
4.2 Balance of forces and moments. Stress tensor. Equation of motion 71
4.3 Local balance of forces and moments 73
4.3.1 Pillbox construction 73
4.3.2 Cauchy’s tetrahedron construction 75
4.3.3 Cauchy’s local balance of forces and moments 76
4.3.4 Cauchy’s theorem 77
4.4 Physical interpretation of the components of stress 79
4.5 Some simple states of stress 80
4.5.1 Pure tension or compression 80
4.5.2 Pure shear stress 81
4.5.3 Hydrostatic stress state 82
4.5.4 Stress deviator 82
4.6 Summary of major concepts related to stress 83

5 Balance of energy and entropy imbalance 85


5.1 Balance of energy. First law of thermodynamics 85
5.1.1 Local form of balance of energy 87
5.2 Entropy imbalance. Second law of thermodynamics 88
5.2.1 Local form of entropy imbalance 89
5.3 Free-energy imbalance. Dissipation 89
5.3.1 Free-energy imbalance and dissipation inequality for mechanical
theories 90
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

xii CONTENTS

6 Balance laws for small deformations 91


6.1 Basic assumptions 91
6.2 Local balance and imbalance laws for small deformations 92

IV Linear elasticity 95

7 Constitutive equations for linear elasticity 97


7.1 Free-energy imbalance (second law): Elasticity 97
7.1.1 Voigt single index notation 100
7.2 Material symmetry 102
7.3 Forms of the elasticity and compliance tensors for some anisotropic linear
elastic materials 103
7.3.1 Monoclinic symmetry 104
7.3.2 Orthorhombic or orthotropic symmetry 105
7.3.3 Tetragonal symmetry 106
7.3.4 Transversely isotropic symmetry. Hexagonal symmetry 108
7.3.5 Cubic symmetry 110
Inter-relations between stiffnesses and compliances: Cubic
materials 111
7.4 Directional elastic modulus 114
7.4.1 Directional elastic modulus: Cubic materials 114
Anisotropy ratio for a cubic single crystal 116
7.5 Constitutive equations for isotropic linear elastic materials 116
7.5.1 Restrictions on the moduli μ and λ 118
7.5.2 Inverted form of the stress-strain relation 119
7.5.3 Physical interpretation of the elastic constants in terms
of local strain and stress states 120
Limiting value of Poisson’s ratio for incompressible materials 122
7.5.4 Stress-strain relations in terms of E and ν 122
7.5.5 Relations between various elastic moduli 123
7.6 Isotropic linear thermoelastic constitutive relations 124

8 Linear elastostatics 128


8.1 Basic equations of linear elasticity 128
8.2 Boundary conditions 129
8.3 Mixed, displacement, and traction problems of elastostatics 130
8.4 Uniqueness 130
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

CONTENTS xiii

8.5 Superposition 131


8.6 Saint-Venant’s principle 132
8.7 Displacement formulation. The Navier equations 133
8.8 Stress formulation. The Beltrami—Michell equations 134
8.9 Two-dimensional problems 137
8.9.1 Plane strain 137
Navier equation in plane strain 139
Equation of compatibility in plane strain 139
8.9.2 Plane stress 140
Navier equation in plane stress 142
Compatibility equation in terms of stress 142
8.9.3 Similarity of plane strain and plane stress equations 143
8.9.4 Succinct form of the governing equations for plane
strain and plane stress 144

9 Solutions to some classical problems in linear elastostatics 145


9.1 Spherical pressure vessel 145
9.2 Cylindrical pressure vessel 149
9.2.1 Axial strain 0 for different end conditions 153
9.2.2 Stress concentration in a thick-walled cylinder under internal
pressure 154
9.3 Bending and torsion 155
9.3.1 Bending of a bar 155
9.3.2 Torsion of a circular cylinder 159
9.3.3 Torsion of a cylinder of arbitrary cross-section 162
Displacement-based formulation 164
Stress-based formulation. Prandtl stress function 165
9.4 Airy stress function and plane traction problems 173
Displacements in terms of the Airy stress function 174
9.4.1 Airy stress function and plane traction problems in polar
coordinates 175
Displacements in terms of the Airy stress function in polar
coordinates 176
9.4.2 Some biharmonic functions in polar coordinates 177
9.4.3 Infinite medium with a hole under uniform far-field loading in
tension 177
9.4.4 Half-space under concentrated surface force 182
9.4.5 A detailed example of Saint-Venant’s principle 194
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

xiv CONTENTS

9.5 Asymptotic crack-tip stress fields. Stress intensity factors 196


9.5.1 Asymptotic stress fields in Mode I and Mode II 197
Asymptotic Mode I field 199
Asymptotic Mode II field 202
9.5.2 Asymptotic stress field in Mode III 205
Solution for uz (r, θ) in the vicinity of the crack 207
9.6 Cartesian component expressions at crack-tip 209
9.6.1 Crack-opening displacements 211

V Variational formulations 213

10 Variational formulation of boundary-value problems 215


10.1 Principle of virtual power 215
10.1.1 Virtual power 216
10.1.2 Consequences of the principle of virtual power 217
10.1.3 Application of the principle of virtual power
to boundary-value problems 219
10.2 Strong and weak forms of the displacement problem
of linear elastostatics 221

11 Introduction to the finite element method for linear elastostatics 223

12 Principles of minimum potential energy


and complementary energy 228
12.1 A brief digression on calculus of variations 228
12.1.1 Application of the necessary condition δI = 0 230
Euler–Lagrange equation 230
Essential and natural boundary conditions 230
12.1.2 Treating δ as a mathematical operator. The process of taking
“variations” 232
The process of taking “variations” 233
12.2 Principle of minimum potential energy 235
12.3 Principle of minimum potential energy and the weak form of the
displacement problem of elastostatics 238
12.4 Principle of minimum complementary energy 239
12.5 Application of the principles of minimum potential energy and
complementary energy to structural problems 244
12.5.1 Castigliano’s first theorem 245
12.5.2 Castigliano’s second theorem 245
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

CONTENTS xv

VI Elastodynamics. Sinusoidal progressive waves 249

13 Elastodynamics. Sinusoidal progressive waves 251


13.1 Mixed problem of elastodynamics 251
13.2 Sinusoidal progressive waves 252

VII Coupled theories 255

14 Linear thermoelasticity 257


14.1 Introduction 257
14.2 Basic equations 258
14.3 Constitutive equations 258
14.3.1 Maxwell and Gibbs relations 260
14.3.2 Specific heat 261
14.3.3 Elasticity tensor. Stress-temperature modulus. Evolution
equation for temperature 261
14.3.4 Entropy as independent variable 263
14.3.5 Isentropic and isothermal inter-relations 264
14.4 Linear thermoelasticity 266
14.5 Basic field equations of linear thermoelasticity 268
14.6 Isotropic linear thermoelasticity 269
14.6.1 Homogeneous and isotropic linear thermoelasticity 270
14.6.2 Uncoupled theory of isotropic linear elasticity 271
14.7 Boundary and initial conditions 272
14.8 Example problems 274

15 Small deformation theory of species diffusion coupled


to elasticity 286
15.1 Introduction 286
15.2 Kinematics and force and moment balances 286
15.3 Balance law for the diffusing species 287
15.4 Free-energy imbalance 288
15.5 Constitutive equations 288
15.5.1 Thermodynamic restrictions 289
15.5.2 Consequences of the thermodynamic restrictions 289
15.5.3 Chemical potential as independent variable 291
15.5.4 Drained and undrained inter-relations 293
15.6 Linear chemoelasticity 294
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

xvi CONTENTS

15.6.1 Isotropic linear chemoelasticity 295


15.6.2 Governing partial differential equations 297
Boundary and initial conditions 298

16 Linear poroelasticity 299


16.1 Introduction 299
16.2 Biot’s theory 300
16.2.1 Fluid flux 302
16.2.2 Material properties for poroelastic materials 303
16.3 Summary of the theory of linear isotropic poroelasticity 304
16.3.1 Constitutive equations 304
16.3.2 Governing partial differential equations 305
Alternate forms of the governing partial differential equations 305
16.3.3 Boundary conditions 306
16.4 Microstructural considerations 311
16.4.1 Constitutive equations of poroelasticity in terms of the
change in porosity 313
u
16.4.2 Estimates for the material parameters α, K , and M 314
16.5 Linear poroelasticity of polymer gels: Incompressible materials 315
16.5.1 Governing partial differential equations 317
Boundary and initial conditions 317
16.5.2 Alternate forms for the governing partial differential equations 318

17 A small deformation chemoelasticity theory for energy


storage materials 320
17.1 Introduction 320
17.2 Basic equations 320
17.3 Specialization of the constitutive equations 321
17.3.1 Strain decomposition 321
17.3.2 Free energy 322
17.3.3 Stress. Chemical potential 323
17.4 Isotropic idealization 324
17.4.1 Chemical strain 324
17.4.2 Elasticity tensor 324
17.4.3 Species flux 324
17.4.4 Stress. Chemical potential. Species flux 325
17.5 Governing partial differential equations for the isotropic theory 325
Partial differential equations 325
Boundary and initial conditions 326
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

CONTENTS xvii

18 Linear piezoelectricity 327


18.1 Introduction 327
18.2 Kinematics and force and moment balances 328
18.3 Basic equations of electrostatics 329
18.4 External and internal expenditures of power. Power balance 330
18.5 Dielectric materials 331
18.6 Free-energy imbalance for dielectric materials 332
18.7 Constitutive theory 332
18.7.1 Electric field as independent variable 333
18.8 Piezoelectricity tensor, permittivity tensor, elasticity tensor 334
18.9 Linear piezoelectricity 336
18.10 Governing partial differential equations 337
Boundary conditions 337
18.11 Alternate form of the constitutive equations for linear piezoelectricity
with σ and e as independent variables 338
18.12 Piezocrystals and poled piezoceramics 339
18.12.1 Voigt notation 339
18.12.2 Properties of piezoelectric crystals 340
18.12.3 Properties of poled piezoceramics 340
APPENDICES 350
18.A Electromagnetics 350
18.A.1 Charge conservation 350
18.A.2 Quasi-static limit 351
18.A.3 Energy transport in the quasi-static limit 351
18.B Third-order tensors 352
18.B.1 Components of a third-order tensor 352
18.B.2 Action of a third-order tensor on a second-order tensor 353
18.B.3 Transpose of a third-order tensor 354
18.B.4 Summary of a third-order tensor and its action on vectors and
tensors 354

VIII Limits to elastic response. Yielding and plasticity 355

19 Limits to elastic response. Yielding and failure 357


19.1 Introduction 357
19.2 Failure criterion for brittle materials in tension 358
19.3 Yield criterion for ductile isotropic materials 359
19.3.1 Mises yield condition 361
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

xviii CONTENTS

19.3.2 Tresca yield condition 362


19.3.3 Coulomb–Mohr yield criterion for cohesive granular
materials in compression 364
19.3.4 The Drucker–Prager yield criterion 365

20 One-dimensional plasticity 369


20.1 Some phenomenological aspects of the elastic-plastic stress-strain
response of polycrystalline metals 369
20.1.1 Isotropic strain-hardening 371
20.1.2 Kinematic strain-hardening 372
20.1.3 Strain-rate and temperature dependence of plastic flow 373
20.2 One-dimensional theory of rate-independent plasticity
with isotropic hardening 374
20.2.1 Kinematics 375
20.2.2 Power 375
20.2.3 Free-energy imbalance 376
20.2.4 Constitutive equation for elastic response 376
20.2.5 Constitutive equation for the plastic response 377
Plastic flow direction 377
Flow strength 378
Yield condition. No-flow conditions. Consistency condition 378
Conditions describing loading and unloading. The plastic
strain-rate ˙p in terms of ˙ and σ 380
20.2.6 Summary of the rate-independent theory with isotropic
hardening 381
20.2.7 Material parameters in the rate-independent theory 382
Some useful phenomenological forms for the strain-hardening
function 383
20.3 One-dimensional rate-dependent plasticity with isotropic hardening 385
20.3.1 Power-law rate dependence 386
20.3.2 Power-law creep 388
20.3.3 A rate-dependent theory with a yield threshold 388
20.3.4 Summary of the rate-dependent theory with isotropic hardening 390
20.4 One-dimensional rate-independent theory with combined isotropic
and kinematic hardening 391
20.4.1 Constitutive equations 391
Defect energy 392
Plastic flow direction 393
Flow resistance 394
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

CONTENTS xix

Yield function. No-flow conditions. Consistency condition 394


The flow rule 395
20.4.2 Summary of the rate-independent theory with combined
isotropic and kinematic hardening 396
20.5 One-dimensional rate-dependent power-law plasticity with combined
isotropic and kinematic hardening 404

21 Three-dimensional plasticity with isotropic hardening 406


21.1 Introduction 406
21.2 Basic equations 406
21.3 Kinematical assumptions 407
21.4 Separability hypothesis 407
21.5 Constitutive characterization of elastic response 408
21.6 Constitutive equations for plastic response 408
21.6.1 Flow strength 409
21.6.2 Mises yield condition. No-flow conditions. Consistency
condition 410
21.6.3 The flow rule 412
21.7 Summary 413
21.8 Three-dimensional rate-dependent plasticity with isotropic hardening 416
21.8.1 Power-law rate dependence 417
21.9 Summary 418
21.9.1 Power-law creep form for high homologous temperatures 419

22 Three-dimensional plasticity with kinematic and isotropic


hardening 421
22.1 Introduction 421
22.2 Rate-independent constitutive theory 421
22.2.1 Yield function. Elastic range. Yield set 423
22.2.2 The flow rule 424
22.3 Summary of the rate-independent theory with combined isotropic
and kinematic hardening 425
22.4 Summary of a rate-dependent theory with combined isotropic and
kinematic hardening 427

23 Small deformation rate-independent plasticity based on a


postulate of maximum dissipation 429
23.1 Elastic domain and yield set 429
23.2 Postulate of maximum dissipation and its consequences 429
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

xx CONTENTS

23.3 Application to a Mises material 432


APPENDICES 433
23.A Kuhn–Tucker optimality conditions 433

24 Some classical problems in rate-independent plasticity 434


24.1 Elastic-plastic torsion of a cylindrical bar 434
24.1.1 Kinematics 434
24.1.2 Elastic constitutive equation 435
24.1.3 Equilibrium 435
24.1.4 Resultant torque: Elastic case 436
Shear stress in terms of applied torque and geometry 437
24.1.5 Total twist of a shaft 437
24.1.6 Elastic-plastic torsion 437
Onset of yield 438
Torque-twist relation 439
Spring-back 442
Residual stress 442
24.2 Spherical pressure vessel 444
24.2.1 Elastic analysis 444
24.2.2 Elastic-plastic analysis 447
Onset of yield 447
Partly plastic spherical pressure vessel 448
Fully plastic spherical pressure vessel 451
Residual stresses upon unloading 451
24.3 Cylindrical pressure vessel 452
24.3.1 Elastic analysis 452
24.3.2 Elastic-plastic analysis 455
Onset of yield 455
Partly plastic cylindrical pressure vessel 455
Fully plastic cylindrical pressure vessel 459

25 Rigid-perfectly-plastic materials. Two extremal principles 460


25.1 Mixed boundary-value problem for a rigid-perfectly-plastic solid 460
25.2 Two extremal principles for rigid-perfectly-plastic materials 461
25.2.1 Extremal principle for power expended by tractions and body
forces 461
25.2.2 Extremal principle for velocity 463
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

CONTENTS xxi

25.3 Limit analysis 465


25.4 Lower bound theorem 467
25.5 Upper bound theorem 468
25.5.1 Upper bound estimates obtained by using block-sliding
velocity fields 469
25.6 Example problems 471
25.6.1 Bounds to the limit load for a notched plate 471
Lower bound 471
Upper bound 472
25.6.2 Hodograph 474
25.6.3 Plane strain frictionless extrusion 475
25.6.4 Plane strain indentation of a semi-infinite solid with a flat punch 478

IX Fracture and fatigue 483

26 Linear elastic fracture mechanics 485


26.1 Introduction 485
26.2 Asymptotic crack-tip stress fields. Stress intensity factors 486
26.3 Configuration correction factors 489
26.4 Stress intensity factors for combined loading by superposition 490
26.5 Limits to applicability of KI -solutions 491
26.5.1 Limit to applicability of KI -solutions because of the
asymptotic nature of the KI -stress fields 491
26.5.2 Limit to applicability of KI -solutions because of local inelastic
deformation in the vicinity of the crack-tip 493
26.5.3 Small-scale yielding (ssy) 495
26.6 Criterion for initiation of crack extension 496
26.7 Fracture toughness testing 497
26.8 Plane strain fracture toughness data 503

27 Energy-based approach to fracture 506


27.1 Introduction 506
27.2 Energy release rate 507
27.2.1 Preliminaries 507
27.2.2 Definition of the energy release rate 508
27.3 Griffith’s fracture criterion 509
27.3.1 An example 510
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

xxii CONTENTS

27.4 Energy release rate and Eshelby’s tensor 514


27.4.1 A conservation integral I 515
27.5 J -integral 516
27.5.1 J -integral and the energy release rate G 518
27.6 Use of the J -integral to calculate the energy release rate G 521
27.7 Relationship between G , J , and KI , KII , and KIII 522
27.8 Use of J as a fracture parameter for elastic-plastic fracture mechanics 523
APPENDICES 525
27.A Equivalence of the two main relations for G(a) 525

28 Fatigue 529
28.1 Introduction 529
28.2 Defect-free approach 532
28.2.1 S-N curves 532
28.2.2 Strain-life approach to design against fatigue failure 534
High-cycle fatigue. Basquin’s relation 535
Low-cycle fatigue. Coffin–Manson relation 536
Strain-life equation for both high-cycle and low-cycle fatigue 537
Mean stress effects on fatigue 538
Cumulative fatigue damage. Miner’s rule 538
28.3 Defect-tolerant approach 539
28.3.1 Fatigue crack growth 539
28.3.2 Engineering approximation of a fatigue crack growth curve 542
28.3.3 Integration of crack growth equation 543

X Linear viscoelasticity 547

29 Linear viscoelasticity 549


29.1 Introduction 549
29.2 Stress-relaxation and creep 550
29.2.1 Stress-relaxation 551
29.2.2 Creep 553
29.2.3 Linear viscoelasticity 554
29.2.4 Superposition. Creep-integral and stress-relaxation-integral
forms of stress-strain relations 554
29.3 A simple rheological model for linear viscoelastic response 556
29.3.1 Stress-relaxation 557
29.3.2 Creep 559
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

CONTENTS xxiii

29.4 Power-law relaxation functions 561


29.5 Correspondence principle 562
29.5.1 Correspondence principle in one dimension 563
29.5.2 Connection between Er (t) and Jc (t) in Laplace transform space 564
29.6 Correspondence principles for structural applications 565
29.6.1 Bending of beams made from linear viscoelastic materials 566
First correspondence principle for bending of beams made
from a linear viscoelastic material 566
Second correspondence principle for bending of beams made
from a linear viscoelastic material 568
29.6.2 Torsion of shafts made from linear viscoelastic materials 570
29.7 Linear viscoelastic response under oscillatory strain and stress 572
29.7.1 Oscillatory loads 572
29.7.2 Storage compliance. Loss compliance. Complex compliance 573
29.7.3 Storage modulus. Loss modulus. Complex modulus 574
29.8 Formulation for oscillatory response using complex numbers 575
29.8.1 Energy dissipation under oscillatory conditions 576
29.9 More on complex variable representation of linear viscoelasticity 580
 
29.9.1 E (ω), E (ω), and tan δ(ω) for the standard linear solid 581
29.10 Time-integration procedure for one-dimensional constitutive
equations for linear viscoelasticity based on the generalized Maxwell
model 588
29.11 Three-dimensional constitutive equation for linear viscoelasticity 592
29.11.1 Boundary-value problem for isotropic linear viscoelasticity 594
29.11.2 Correspondence principle in three dimensions 595
29.12 Temperature dependence. Dynamic mechanical analysis (DMA) 597
29.12.1 Representative DMA results for amorphous polymers 598
29.12.2 DMA plots for the semi-crystalline polymers 600
29.13 Effect of temperature on Er (t) and Jc (t). Time-temperature equivalence 601
29.13.1 Shift factor. Williams–Landel–Ferry (WLF) equation 607

XI Finite elasticity 609

30 Finite elasticity 611


30.1 Brief review of kinematical relations 611
30.2 Balance of linear and angular momentum expressed spatially 613
30.3 First Piola stress. Force and moment balances expressed referentially 614
30.4 Change of observer. Change of frame 617
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

xxiv CONTENTS

30.4.1 Transformation rules for kinematic fields 617


30.4.2 Transformation rules for the Cauchy stress and the Piola stress 618
30.4.3 Form invariance of constitutive equations. Material-frame-
indifference 619
30.5 Free-energy imbalance under isothermal conditions 619
30.5.1 The second Piola stress 620
30.6 Constitutive theory 621
30.6.1 Thermodynamic restrictions 621
30.7 Summary of basic equations. Initial/boundary-value problems 622
30.7.1 Basic field equations 622
30.7.2 A typical initial/boundary-value problem 623
30.8 Material symmetry 623
30.9 Isotropy 625
30.10 Isotropic scalar functions 626
30.11 Free energy for isotropic materials expressed in terms of invariants 627
30.12 Free energy for isotropic materials expressed in terms of principal
stretches 628
30.13 Incompressibility 629
30.14 Incompressible elastic materials 630
30.14.1 Free energy for incompressible isotropic elastic bodies 631
30.14.2 Simple shear of a homogeneous, isotropic, incompressible
elastic body 633
APPENDICES 635
30.A Derivatives of the invariants of a tensor 635

31 Finite elasticity of elastomeric materials 637


31.1 Introduction 637
31.2 Neo-Hookean free-energy function 639
31.3 Mooney–Rivlin free-energy function 640
31.3.1 The Mooney–Rivlin free energy expressed in an alternate fashion 643
31.4 Ogden free-energy function 643
31.5 Arruda–Boyce free-energy function 644
31.6 Gent free-energy function 647
31.6.1 I2 -dependent extension of the Gent free-energy function 650
31.7 Slightly compressible isotropic elastomeric materials 651
31.7.1 Models for the volumetric energy 652
31.7.2 Slightly compressible form of the Arruda–Boyce free energy 654
31.7.3 Slightly compressible form of the Gent free energy 654
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi

CONTENTS xxv

XII Appendices 655

A Cylindrical and spherical coordinate systems 657


A.1 Introduction 657
A.2 Cylindrical coordinates 657
A.3 Spherical coordinates 664

B Stress intensity factors for some crack configurations 671

Bibliography 679
Index 689
OUP CORRECTED PROOF – FINAL, 15/6/2020, SPi
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

PA R T I

Vectors and tensors


OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

1 Vectors and tensors: Algebra

The theories of Solid Mechanics are mathematical in nature. Thus in this chapter and the next,
we begin with a discussion of the essential mathematical definitions and operations regarding
vectors and tensors needed to discuss the major topics covered in this book.1
The approach taken is a pragmatic one, without an overly formal emphasis upon mathemati-
cal proofs. The end-goal is to develop, in a reasonably systematic way, the essential mathematics
of vectors and tensors needed to discuss the major aspects of mechanics of solids. The physical
quantities that we will encounter in this text will be of three basic types: scalars, vectors, and
tensors. Scalars are quantities that can be described simply by numbers, for example things like
temperature and density. Vectors are quantities like velocity and displacement—quantities that
have both magnitude and direction. Tensors will be used to describe more complex quantities
like stress and strain, as well as material properties. In this chapter we deal with the algebra of
vectors and tensors. In the next chapter we deal with their calculus.

1.1 Cartesian coordinate frames. Kronecker delta.


Alternating symbol
Throughout this text lower case Latin subscripts range over the integers
{1, 2, 3}.
A Cartesian coordinate frame for the Euclidean point space E consists of a reference point
o called the origin together with a positively oriented orthonormal basis {e1 , e2 , e3 } for the
associated vector space V . Being positively oriented and orthonormal, the basis vectors obey
ei · ej = δij , ei · (ej × ek ) = eijk . (1.1.1)
Here δij , the Kronecker delta, is defined by

⎨1, if i = j,
δij = (1.1.2)
⎩0, if i = j,

1
The mathematical preliminaries in this chapter and the next are largely based on the clear expositions by Gurtin
in Gurtin (1981) and Gurtin et al. (2010).

Continuum Mechanics of Solids. Lallit Anand and Sanjay Govindjee, Oxford University Press (2020).
© Lallit Anand and Sanjay Govindjee, 2020.
DOI: 10.1093/oso/9780198864721.001.0001
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

4 VECTORS AND TENSORS: ALGEBRA

while eijk , the alternating symbol,2 is defined by




⎪ 1, if {i, j, k} = {1, 2, 3}, {2, 3, 1}, or {3, 1, 2},


eijk = −1, if {i, j, k} = {2, 1, 3}, {1, 3, 2}, or {3, 2, 1}, (1.1.3)



⎩ 0, if an index is repeated,

and hence has the value +1, −1, or 0 when {i, j, k} is an even permutation, an odd permuta-
tion, or not a permutation of {1, 2, 3}, respectively. Where convenient, we will use the compact
notation
def
{ei } = {e1 , e2 , e3 },
to denote our positively oriented orthonormal basis.

1.1.1 Summation convention


For convenience, we employ the Einstein summation convention according to which summa-
tion over the range 1, 2, 3 is implied for any index that is repeated twice in any term, so that,
for instance,
ui v i = u1 v 1 + u2 v 2 + u3 v 3 ,

Sij uj = Si1 u1 + Si2 u2 + Si3 u3 ,

Sik Tkj = Si1 T1j + Si2 T2j + Si3 T3j .


In the expression Sij uj the subscript i is free, because it is not summed over, while j is a dummy
subscript, since
Sij uj = Sik uk = Sim um .
As a rule when using the Einstein summation convention, one can not repeat an index more
than twice. Expressions with such indices indicate that a computational error has been made.
In the rare case where three (or more) repeated indices must appear in a single term, an explicit
3
summation symbol, for example i=1 , must be used.
Note that the Kronecker delta δij modifies (or contracts) the subscripts in the coefficients of
an expression in which it appears:
ai δij = aj , ai bj δij = ai bi = aj bj , δij δik = δjk , δik Skj = Sij .
This property of the Kronecker delta is sometimes termed the substitution property.
Further, when an expression in which an index is repeated twice but summation is not to be
performed we state so explicitly. For example,
ui v i (no sum)
signifies that the subscript i is not to be summed over.

2
The alternating symbol is sometimes called the permutation symbol.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

CARTESIAN COORDINATE FRAMES. KRONECKER DELTA. ALTERNATING SYMBOL 5

Because {ei } is a basis, every vector u admits the unique expansion


u = u j ej ; (1.1.4)
the scalars ui are called the (Cartesian) components of u (relative to this basis). If we take the
inner product of (1.1.4) with ei we find that, since ei · ej = δij ,
u i = u · ei . (1.1.5)
Guided by this relation, we define the coordinates of a point x with respect to the origin o by
xi = (x − o) · ei . (1.1.6)
In view of (1.1.4), the inner product of vectors u and v may be expressed as
u · v = (ui ei ) · (vj ej )
= ui vj δij

= ui v i = u1 v 1 + u2 v 2 + u3 v 3 . (1.1.7)
It should also be recalled that this relation is equivalent to u · v = |u| |v| cos θuv , where | · |
denotes the length of a vector (its magnitude or norm) and θuv is the angle between vectors
u and v. Equation (1.1.4) further implies that the cross product of two vectors
u × v = (uj ej ) × (vk ek )
= u j v k (ej × ek )

= eijk uj vk ei . (1.1.8)
In particular, (1.1.8) implies that the vector u × v has the component form
(u × v)i = eijk uj vk . (1.1.9)
This definition of the cross product is fully compatible with the relation |u × v| = |u| |v| sin θuv ,
which indicates that the magnitude of the cross product of two vectors is equal to the area of
the parallelogram whose edges are defined by u and v.
It can be shown that the permutation symbol is related to the Kronecker delta by
⎡ ⎤ ⎡ ⎤
δi1 δi2 δi3 δi1 δj1 δk1
eijk = det ⎣ δj1 δj2 δj3 ⎦ = det ⎣δi2 δj2 δk2 ⎦ , (1.1.10)
δk1 δk2 δk3 δi3 δj3 δk3
where det[·] represents the determinant of a 3 × 3 matrix. A consequence of (1.1.10) is that
⎡ ⎤
δip δiq δir
eijk epqr = det ⎣ δjp δjq δjr ⎦ ,
δkp δkq δkr (1.1.11)

= δip (δjq δkr − δjr δkq ) − δiq (δjp δkr − δjr δkp ) + δir (δjp δkq − δjq δkp ).
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

6 VECTORS AND TENSORS: ALGEBRA

Some useful identities which follow from successive contractions of (1.1.11), the substitution
property of the Kronecker delta and the identity δii = 3 are

eijk eipq = δjp δkq − δjq δkp , (1.1.12)

and hence also that


eijk eijl = 2δkl ,
(1.1.13)
eijk eijk = 6.
Also useful is the identity
ei = 12 eijk ej × ek . (1.1.14)
Let u, v, and w be vectors. Useful relations involving the inner and cross products then
include,
u · (v × w) = v · (w × u) = w · (u × v),

u · (u × v) = 0,
(1.1.15)
u × v = −v × u,

u × ( v × w) = ( u · w) v − ( u · v ) w.
The operation
def
[u, v, w] = u · (v × w), (1.1.16)
is known as the scalar triple product and is physically equivalent to the (signed) volume of the
parallelepiped whose edges are defined by any three linearly independent vectors u, v, and w.

1.2 Tensors
1.2.1 What is a tensor?
In mathematics the term tensor is a synonym for the phrase “a linear transformation which
maps a vector into a vector.” A tensor S is therefore a linear mapping that assigns to each vector
u a vector
v = Su. (1.2.1)
One might think of a tensor S as a machine with an input and an output: if a vector u is the
input, then the vector v = Su is the output. Linearity of a tensor S is the requirement that
S(u + v) = Su + Sv for all vectors u and v,
(1.2.2)
S(αu) = αSu for all vectors u and scalars α.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

TENSORS 7

1.2.2 Zero and identity tensors


Two basic tensors are the zero tensor 0 and the identity tensor 1:
0v = 0, 1v = v
for all vectors v.

1.2.3 Tensor product


Another example of a tensor is the tensor product3 u ⊗ v, of two vectors u and v, defined by
(u ⊗ v)w = (v · w)u (1.2.3)
for all w; the tensor u ⊗ v maps any vector w onto a scalar multiple of u.

EXAMPLE 1.1 Projection tensor

As a concrete example of the tensor product consider the tensor


P = 1 − n ⊗ n,
where n is a vector of unit length. The action of P on an arbitrary vector u is then given by
Pu = (1 − n ⊗ n)u,
= 1u − (n ⊗ n)u,
= u − ( u · n) n . (1.2.4)
Equation (1.2.4) can be interpreted to be the vector u minus the projection of u in the
direction of n. Hence Pu is equal to the projection of the vector u into the plane orthogonal
to n. Note that in this example we have exploited the fact that the dot product of a vector
with a unit vector results in its projection onto the direction defined by the unit vector.
Further, we have exploited the additive algebraic structure of tensors, whereby for tensors
S and T, (S + T)u = Su + Tu for all vectors u.

1.2.4 Components of a tensor


Given a tensor S, the quantity Sej is a vector. The components of a tensor S with respect to the
basis {e1 , e2 , e3 } are defined by
def
Sij = ei · Sej . (1.2.5)

3
Other common terminologies for a tensor product are tensor outer product or simply outer product, as well as,
dyadic product.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

8 VECTORS AND TENSORS: ALGEBRA

With this definition of the components of a tensor, the Cartesian component representation of
the relation v = Su is
vi = ei · v = ei · Su
= ei · S(uj ej )
= (ei · Sej )uj
= Sij uj . (1.2.6)
It should be observed that (1.2.6) indicates that each component of v is a linear combination
of the components of u, where the weights are given by the components of S.
Further, a tensor S has the representation
S = Sij ei ⊗ ej , (1.2.7)
in terms of its components Sij and the basis tensors ei ⊗ ej .

EXAMPLE 1.2 Identity tensor components

Using (1.2.5), the components of the identity tensor are given by


(1)ij = ei · 1ej = ei · ej = δij .
Thus the identity tensor can be expressed as 1 = δij ei ⊗ ej , equivalently as 1 = ei ⊗ ei .

EXAMPLE 1.3 Zero tensor components

Using (1.2.5), the components of the zero tensor are given by


(0)ij = ei · 0ej = ei · 0 = 0.
3 3
Thus the zero tensor can be expressed as 0 = i=1 j=1 0 ei ⊗ ej .

1.2.5 Transpose of a tensor


The transpose S of a tensor S is the unique tensor with the property that
u · Sv = v · S u (1.2.8)
for all vectors u and v. The components of the transpose are determined via (1.2.5) as
ei · S ej = ej · Sei = Sji ;
thus the components of S are
(S )ij = Sji . (1.2.9)
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

TENSORS 9

1.2.6 Symmetric and skew tensors


A tensor S is symmetric if
S = S , Sij = Sji , (1.2.10)
and skew if
S = −S , Sij = −Sji . (1.2.11)
Clearly,
Sij = 1
2 Sij + Sji + 1
2 Sij − Sji .
Thus, every tensor S admits the decomposition
S = sym S + skw S (1.2.12)
into a symmetric part and a skew part, where
sym S = 12 (S + S ),
(1.2.13)
skw S = 12 (S − S ),
with components
(sym S)ij = 12 (Sij + Sji ),
(1.2.14)
(skw S)ij = 12 (Sij − Sji ).
Note that a symmetric tensor has at most six independent components, and a skew tensor has
at most three independent components; the latter point follows since (skw S)ii = 0 (no sum).

1.2.7 Axial vector of a skew tensor


Since a skew tensor has only three independent components, it is possible to define its action
on vectors by another vector. Given any skew tensor Ω, there is a unique vector ω —called the
axial vector of Ω—such that
Ωu = ω × u (1.2.15)
for all vectors u. Since
(Ωu)i = Ωij uj and (ω × u)i = eijk ωj uk = eikj ωk uj ,
we have that
Ωij = eikj ωk ≡ −eijl ωl . (1.2.16)
Further, operating on both sides of (1.2.16) with eijk and using the epsilon-delta identity
(1.1.13)1 we obtain
eijk Ωij = −eijk eijl ωl = −2δkl ωl = −2ωk ,
and hence in terms of the three independent components of the skew tensor Ω, the three
components of its axial vector ω are given by
ωi = − 12 eijk Ωjk , (1.2.17)
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

10 VECTORS AND TENSORS: ALGEBRA

which in expanded form is


ω1 = Ω32 = −Ω23 , ω2 = Ω13 = −Ω31 , ω3 = Ω21 = −Ω12 . (1.2.18)

1.2.8 Product of tensors


Given tensors S and T, the product ST is defined by composition; that is, ST is defined by
(ST)v = S(Tv) (1.2.19)

for all vectors v. By (1.2.5), Tej = Tlj el and by (1.2.9), S ei = Sik ek ; thus,
ei · STej = S ei · Tej
= (Sik ek ) · (Tlj el )

= Sik Tlj (ek · el )


 
δkl
= Sik Tkj ,
and, hence,
(ST)ij = Sik Tkj . (1.2.20)
Generally, ST = TS. When ST = TS, the tensors S and T are said to commute.

EXAMPLE 1.4 Product with the identity tensor

The product of the identity tensor 1 with any arbitrary tensor T, is T itself:
(1T)v = 1(Tv) = Tv,
and since this holds for all vectors v we have
1T = T.
Alternatively, in components
(1T)ij = ei · 1Tej
= 1 ei · Tej
= ei · Tej
= Tij .
The components of 1T are equal to the components of T. The two are equal to each other.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

TENSORS 11

1.2.9 Trace of a tensor. Deviatoric tensors


The trace is the linear operation that assigns to each tensor S a scalar tr S and satisfies
tr(u ⊗ v) = u · v (1.2.21)
for any vectors u and v. Linearity is the requirement that (cf. (1.2.2))
tr(αS + β T) = α tr(S) + β tr(T)
for all tensors S and T and all scalars α and β . Thus, by (1.2.7),
tr S = tr(Sij ei ⊗ ej )

= Sij tr(ei ⊗ ej )

= Sij (ei · ej )

= Sii , (1.2.22)
and the trace is well-defined for all tensors. Some useful properties of the trace are
tr(Su ⊗ v) = v · Su,

tr(S ) = tr S,
(1.2.23)
tr(ST) = tr(TS),

tr 1 = 3.
As a consequence of (1.2.23)2 ,
tr S = 0 whenever S is skew. (1.2.24)
A tensor S is deviatoric (or traceless) if
tr S = 0, (1.2.25)
and we refer to
S ≡ dev S
def
= S − 13 (tr S)1 (1.2.26)
as the deviatoric part of S,4 and to
1
3 (tr S)1 (1.2.27)
as the spherical part of S. Trivially,
S = S − 13 (tr S)1 + 13 (tr S)1,
   
S s1

4
The notation “dev” is useful for denoting the deviatoric part of the product of many tensors; e.g., dev(ST · · · M).
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

12 VECTORS AND TENSORS: ALGEBRA

where s = 13 tr S, and it is noted that tr dev S = 0. Thus every tensor S admits the decomposition
S = S + s1 (1.2.28)
into a deviatoric tensor and a spherical tensor.

EXAMPLE 1.5 Deviatoric and spherical parts of a tensor

The deviatoric part of a tensor


T = 3e1 ⊗ e1 + 2e1 ⊗ e3
is determined as
T = T − 13 (tr T)1
= (3e1 ⊗ e1 + 2e1 ⊗ e3 ) − 1
3 · 31
= 2e1 ⊗ e1 + 2e1 ⊗ e3 − 1e2 ⊗ e2 − 1e3 ⊗ e3 .
1
The spherical part of T is given by 3 tr T1 = 1.

1.2.10 Positive definite tensors


A tensor C is positive definite if and only if
u · Cu > 0, ui Cij uj > 0 (1.2.29)
for all vectors u = 0.

1.2.11 Inner product of tensors. Magnitude of a tensor


The inner product of two vectors u and v is defined by
u · v = v · u = ui v i .
Analogously, the inner product of two tensors S and T is defined by
S : T = T : S = Sij Tij .
The symbol : is known as the double contraction symbol.
By analogy to the notion of the magnitude (or norm) of a vector u,
√ √
|u| = u · u = u i u i ,
the magnitude (or norm) |S| of a tensor S is defined by
√ 
|S| = S : S = Sij Sij .
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

TENSORS 13

EXAMPLE 1.6 Contraction of a skew tensor with a symmetric tensor

Consider a symmetric tensor S (Sij = Sji ) and a skew tensor Ω (Ωij = −Ωji ). Their
double contraction S : Ω = Sij Ωij will always be zero:
Sij Ωij = −Sij Ωji = −Sji Ωji = −Sij Ωij .
Since the only scalar with the property that it is equal to its negative is zero, we conclude
that Sij Ωij = 0 whenever S is symmetric and Ω is skew.

1.2.12 Matrix representation of tensors and vectors


Tensors and vectors of the type presented have related matrix representations. We write [u] and
[S] for the matrix representations of a vector u and a tensor S with respect to the basis {ei }:
⎡ ⎤ ⎡ ⎤
u1 S11 S12 S13
[u] = ⎣ u 2 ⎦ , [S] = ⎣ S21 S22 S23 ⎦ .
u3 S31 S32 S33

• All the operations and operators we have introduced for vectors and tensors are in one-
to-one correspondence to the same operations and operators for matrices.

For example,
⎡ ⎤⎡ ⎤
S11 S12 S13 u1
[S][u] = ⎣ S21 S22 S23 ⎦ ⎣ u2 ⎦
S31 S32 S33 u3
⎡ ⎤
S11 u1 + S12 u2 + S13 u3
= ⎣ S21 u1 + S22 u2 + S23 u3 ⎦
S31 u1 + S32 u2 + S33 u3
⎡  ⎤
S u
i 1i i
=⎣ S u ⎦
i 2i i
i S3i ui

= [Su],
so that the action of a tensor on a vector is consistent with that of a 3 × 3 matrix on a 3 × 1
matrix. Further, the matrix [S ] of the transpose S of S is identical to the transposition of the
matrix [S]:
⎡ ⎤
S11 S21 S31
[S ] ≡ [S] = ⎣ S12 S22 S32 ⎦ .
S13 S23 S33
14 VECTORS AND TENSORS: ALGEBRA

Similarly, the trace of a tensor S is equivalent to the conventional definition of this quantity
from matrix algebra
tr S ≡ tr [S] = S11 + S22 + S33 = Skk .

The inner product of two vectors


⎡ ⎤
  v1
u · v ≡ [u] [v] = u1 u2 u3 ⎣ v 2 ⎦ = u1 v 1 + u2 v 2 + u3 v 3 ≡ ui v i ,
v3
and the tensor product

⎤ ⎡ ⎤
u1   u1 v 1 u1 v 2 u1 v 3
[u ⊗ v] ≡ [u][v] = ⎣ u2 ⎦ v1 v2 v3 = ⎣ u2 v 1 u2 v 2 u2 v 3 ⎦ .
u3 u3 v 1 u3 v 2 u3 v 3
If one chooses to perform computations using matrix representations of vectors and tensors,
care must be taken to use a single basis in all calculations, as the matrix representations are
basis dependent.

1.2.13 Determinant of a tensor


Tensors, like square matrices, have determinants. The general definition of the determinant is
an operation that assigns to each tensor S a scalar det S defined by
Su · (Sv × Sw)
det S = (1.2.30)
u · ( v × w)
for any three non-coplanar vectors {u, v, w}. Thus, det S is the ratio of the volume of the
parallelepiped defined by the vectors Su, Sv, and Sw to the volume of the parallelepiped defined
by the vectors u, v, and w.
Definition (1.2.30) is fully equivalent to the conventional one given for square matrices; in
particular
 
 S11 S12 S13 
 
det S ≡ det[S] =  S21 S22 S23 
 S S33 
31 S32

= S11 (S22 S33 − S23 S32 ) − S12 (S21 S33 − S23 S31 ) + S13 (S21 S32 − S22 S31 )

= eijk Si1 Sj2 Sk3 = 16 eijk epqr Sip Sjq Skr . (1.2.31)
It is useful also to observe that
det(ST) = det S det T, (1.2.32)
and that
det S = det S. (1.2.33)
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

TENSORS 15

1.2.14 Invertible tensors


A tensor S is invertible if there is a tensor S−1 , called the inverse of S, such that
SS−1 = S−1 S = 1. (1.2.34)
Further,
S is invertible if and only if det S = 0. (1.2.35)
Note also that det S−1 = 1/ det S.

1.2.15 Cofactor of a tensor


Let S be an invertible tensor, then the tensor
cof S = (det S)S− ,
def
(1.2.36)
is called the cofactor of S. A straight forward but slightly involved calculation shows that for
all linearly independent vectors u and v,
cof S(u × v) = Su × Sv. (1.2.37)
That is, cof S transforms the area vector u × v of the parallelogram defined by u and v into the
area vector Su × Sv of the parallelogram defined by Su and Sv.

1.2.16 Orthogonal tensors


A tensor Q is orthogonal if and only if
Q Q = QQ = 1. (1.2.38)
If Q is orthogonal, then
det Q = ±1.
An orthogonal tensor is a rotation (or a proper rotation) if det Q = 1, and a reflection (or an
improper rotation) if det Q = −1.

1.2.17 Transformation relations for components of a vector and a tensor


under a change in basis
Given a Cartesian coordinate system with a right-handed orthonormal basis {ei }, a vector v
and a tensor S may be represented by their components
vi = ei · v and Sij = ei · Sej ,
respectively. In another coordinate system, cf. Fig. 1.1, with a right-handed orthonormal basis
{e∗i }, v and S have the component representations
vi∗ = e∗i · v and ∗
Sij = e∗i · Se∗j .
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

16 VECTORS AND TENSORS: ALGEBRA

e2
e∗2

e1
o
e∗3

e∗1
e3

Fig. 1.1 Change in basis.

We now consider how the two representations (vi , Sij ) and (vi∗ , Sij

) in the two coordinate
systems are related to each other.
Let Q be the rotation tensor defined by

Q = e1 ⊗ e∗1 + e2 ⊗ e∗2 + e3 ⊗ e∗3 ≡ ek ⊗ e∗k ,


def
(1.2.39)
so that
e∗i = Q ei i = 1, 2, 3. (1.2.40)
That Q is a rotation follows from QQ = (ei ⊗ e∗i )(e∗j ⊗ ej ) = δij ei ⊗ ej = 1 and similarly for
Q Q, along with the right-handedness of the bases to ensure det Q = +1. The components of
Q with respect to {ei } are given by
Qij = ei · (ek ⊗ e∗k )ej = δik e∗k · ej = e∗i · ej .
Thus, a base vector e∗i may be expressed in terms of the basis {ej }, with Qij serving as
components of the vector e∗i :
e∗i = Qij ej , (1.2.41)
and the matrix representation of the components of Q with respect to the basis {ei } is
⎡ ⎤
Q11 Q12 Q13
[Q] = ⎣Q21 Q22 Q23 ⎦ .
Q31 Q32 Q33

Transformation relation for vectors


For a vector v the components with respect to the basis {e∗i } are
vi∗ = e∗i · v = (Qij ej ) · v = Qij (ej · v) = Qij vj .
Thus, the transformation relation for the Cartesian components of a vector v under a change
of basis are:
vi∗ = Qij vj . (1.2.42)
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

TENSORS 17

These relations may be expressed in matrix form as follows:


⎡ ∗⎤ ⎡ ⎤⎡ ⎤
v1 Q11 Q12 Q13 v1
⎣v2∗ ⎦ = ⎣Q21 Q22 Q23 ⎦ ⎣v2 ⎦ ,
v3∗ Q31 Q32 Q33 v3

where the components of Q are given in the {ei } basis.

Transformation relation for tensors


For a tensor S the components with respect to the basis {e∗i } are
     

Sij = e∗i · Se∗j = Qik ek · S Qjl el = Qik Qjl ek · Sel = Qik Qjl Skl .

Thus, the transformation relation for the Cartesian components of a tensor S under a change
of basis are:

Sij = Qik Qjl Skl . (1.2.43)

These relations may be expressed in matrix form as follows:


⎡ ∗ ∗ ∗
⎤ ⎡ ⎤⎡ ⎤⎡ ⎤
S11 S12 S13 Q11 Q12 Q13 S11 S12 S13 Q11 Q12 Q13
⎣S21
∗ ∗
S22 ∗ ⎦
S23 = ⎣Q21 Q22 Q23 ⎦ ⎣S21 S22 S23 ⎦ ⎣Q21 Q22 Q23 ⎦ ,
∗ ∗ ∗
S31 S32 S33 Q31 Q32 Q33 S31 S32 S33 Q31 Q32 Q33
where the components of Q are given in the {ei } basis.

1.2.18 Eigenvalues and eigenvectors of a tensor. Spectral theorem


A scalar ω is an eigenvalue of a tensor S, if there is a unit vector μ such that
Sμ = ωμ, (1.2.44)
in which case μ is an eigenvector of S corresponding to the eigenvalue ω . Physically, eigenvec-
tors of S are those vectors which remain parallel to themselves when mapped by the tensor S,
Fig. 1.2.
If ω and μ are an eigenvalue and corresponding eigenvector for a tensor S, then (1.2.44)
implies that
(S − ω1)μ = 0. (1.2.45)

µ Sµ

Fig. 1.2 Eigenvector μ of a tensor S.


OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

18 VECTORS AND TENSORS: ALGEBRA

For (1.2.45) to possess a non-trivial (μ = 0) solution, the tensor (S −ω1) can not be invertible,
so that, by (1.2.35),
det(S − ω1) = 0.
Thus, the determinant of the 3 × 3 matrix [S] − ω[1] must vanish,
det [S] − ω[1] = 0,
which is simply the classic requirement for a system of homogeneous equations to have a non-
trivial solution. Each eigenvalue ω of a tensor S is, therefore, a solution of a polynomial equation
of the form ω 3 − a1 ω 2 + a2 ω − a3 = 0, where the coefficients are functions of S. A tedious
computation shows that this characteristic equation5 has the explicit form
ω 3 − I1 (S)ω 2 + I2 (S)ω − I3 (S) = 0, (1.2.47)
where I1 (S), I2 (S), and I3 (S), called the principal invariants6 of S, are given by

I1 (S) = tr S, ⎪



 2 
I2 (S) = 2 tr(S) − tr(S ) ,
1 2 (1.2.48)




I3 (S) = det S.

Ik (S) are called invariants because of the way they transform under the group of orthogonal
tensors:
Ik (QSQ ) = Ik (S) for any orthogonal tensor Q. (1.2.49)

The solutions of the characteristic equation (1.2.47), cubic in ω , are the eigenvalues ωi ,
i = 1, 2, 3. Since the principal invariants Ik (S) are always real, the theory of polynomials tells
us that

• The characteristic equation has (i) either three real roots (not necessarily distinct), or (ii)
one real and two complex conjugate roots.

Once the eigenvalues have been determined, the eigenvectors corresponding to each eigenvalue
can be found by substituting back into (1.2.45).

5
The Cayley–Hamilton theorem also tells us that a tensor satisfies its characteristic equation; that is
S3 − I1 (S)S2 + I2 (S)S − I3 (S)1 = 0. (1.2.46)
6
The principal invariants can also be defined in terms of any collection of three vectors a, b, c, with [a, b, c] = 0,
via the relations
 
I1 (S) = [Sa, b, c] + [a, Sb, c] + [a, b, Sc] /[a, b, c],
 
I2 (S) = [Sa, Sb, c] + [a, Sb, Sc] + [Sa, b, Sc] /[a, b, c],
 
I3 (S) = [Sa, Sb, Sc] /[a, b, c].
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

TENSORS 19

Eigenvalues of symmetric tensors


It turns out that for a symmetric tensor the eigenvalues are all real, and the corresponding
eigenvectors {μi } are mutually orthogonal. Suppose that S is symmetric and that ω1 = ω2 are
real eigenvalues of S with corresponding eigenvectors μ1 and μ2 . Then, since S is symmetric,
0 = (S μ 1 − ω1 μ 1 )
0 = μ2 · (Sμ1 − ω1 μ1 )
= Sμ2 · μ1 − ω1 μ2 · μ1
= ω2 μ 2 · μ 1 − ω1 μ 2 · μ 1
= (ω2 − ω1 )μ2 · μ1
Since ω1 = ω2 , we must have
μ2 · μ1 = 0.
Thus, for a symmetric tensor, eigenvectors corresponding to distinct eigenvalues are orthogo-
nal. When the eigenvalues are repeated, the situation is a bit more complex and is treated below.
Notwithstanding, if S is symmetric, with distinct eigenvalues {ω1 > ω2 > ω3 }, then they are
real, and there exists an orthonormal set of corresponding eigenvectors {μ1 , μ2 , μ3 } such
that Sμi = ωi μi (no sum).

Spectral Theorem Let S be symmetric with distinct eigenvalues ωi . Then there is a corre-
sponding orthonormal set {μi } of eigenvectors of S and, what is most important, one uniquely
has that
 3
S= ωi μ i ⊗ μ i . (1.2.50)
i=1
The relation (1.2.50), which is called a spectral decomposition of S, gives S as a linear combi-
nation of projections, with each μi ⊗ μi (no sum) a projection tensor onto the eigenvector μi .
Since the eigenvectors {μi } are orthonormal, they can also function as a basis, and in this
basis the matrix representation S is diagonal:
⎡ ⎤
ω1 0 0
[S] = ⎣ 0 ω2 0 ⎦ .
0 0 ω3
When the eigenvalues of S are repeated, the spectral theorem needs to be modified. The
remaining relevant cases are when two of the eigenvalues are the same and when all three
eigenvalues are the same. Respectively, in these cases one has
(a) If ω1 = ω2 = ω3 , then S admits the representation
S = ω1 μ1 ⊗ μ1 + ω2 (1 − μ1 ⊗ μ1 ), (1.2.51)
which indicates that μ1 , as well as any vector orthogonal to μ1 , is an eigenvector of S.
(b) If ω ≡ ω1 = ω2 = ω3 , then S admits the representation
S = ω1, (1.2.52)
which indicates that any vector will qualify as an eigenvector of S and that S is spherical.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

20 VECTORS AND TENSORS: ALGEBRA

Other invariants and eigenvalues


The principal invariants (1.2.48) of a tensor are completely characterized by the eigenvalues
{ω1 , ω2 , ω3 }:

I1 (S) = ω1 + ω2 + ω3 , ⎪


I2 (S) = ω1 ω2 + ω2 ω3 + ω3 ω1 , (1.2.53)



I3 (S) = ω1 ω2 ω3 .
Invariants play an important role in continuum mechanics. As is clear from (1.2.53) for a tensor
S the eigenvalues {ω1 , ω2 , ω3 } are the basic invariants in the sense that any invariant of S can
be expressed in terms of them. In many applications, as an alternate to (1.2.53) it is often more
convenient to choose as invariants the following three symmetric functions of {ω1 , ω2 , ω3 }:

tr S = ω1 + ω2 + ω3 , ⎪


tr S2 = ω12 + ω22 + ω32 , (1.2.54)



tr S3 = ω13 + ω23 + ω33 .
These three quantities are clearly invariant and they are independent in the sense that no one
of them can be expressed in terms of the other two.
As a further alternative set of invariants recall that a tensor S may always be decomposed
into a deviatoric and spherical part as
1
S= S
 + (tr S)1 , with tr S = 0. (1.2.55)
3  
deviatoric part
spherical part

The decomposition (1.2.55) leads to the possibility of using the invariant (tr S) and the
invariants of S as an alternative set of invariants. Since S has only two independent non-zero
invariants  2 
tr S , tr S3 , (1.2.56)
it is sometimes convenient to adopt the list
 
tr S, tr S2 , tr S3 (1.2.57)
as a set of alternative invariants of a tensor S.

1.2.19 Fourth-order tensors


In our discussions we will also need the concept of a fourth-order tensor. A fourth-
order tensor is defined as a linear transformation that maps a second-order tensor to a
second-order tensor. A fourth-order tensor C is therefore a linear mapping that assigns to
each second-order tensor A a second-order tensor

B = CA. (1.2.58)
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

TENSORS 21

Linearity of C is the requirement that


C(A + B) = CA + CB for all tensors A and B,
C(αA) = αCA for all tensors A and scalars α.
It should also be noted that common alternate notations for CS are
C[S] and C : S.
The latter form, C : S, emphasizes the action of C via its component form, which is discussed
next.
Recall that the components Tij of a second-order tensor are defined as
Tij = ei · Tej .
For discussing the component form of fourth-order tensors it is convenient to introduce the
basis tensors
def
Eij = ei ⊗ ej , (1.2.59)
with the orthonormality property
Eij : Ekl = δik δjl .
Using this notation, the components Cijkl of C are defined as
Cijkl = Eij : CEkl . (1.2.60)
This allows one to also express a fourth-order tensor as
C = Cijkl Eij ⊗ Ekl ≡ Cijkl ei ⊗ ej ⊗ ek ⊗ el . (1.2.61)
From these expressions, one can also infer the component-wise action of a fourth-order tensor
on a second-order tensor as
S = CT, Sij = Cijkl Tkl . (1.2.62)

EXAMPLE 1.7 Fourth-order identity tensor

The fourth-order tensor


I = δik δjl Eij ⊗ Ekl
is the identity tensor. Verify that I as defined possesses the property of an identity operator,
viz. IT = T for all tensors T.
IT ≡ I : T = (δik δjl Eij ⊗ Ekl ) : (Tpq Epq )
= δik δjl Tpq (Ekl : Epq )Eij
= δik δjl Tpq δkp δlq Eij
= Tij Eij = T.
Since this expression was derived for an arbitrary tensor T, I must be the fourth-order
identity tensor.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

22 VECTORS AND TENSORS: ALGEBRA

EXAMPLE 1.8 Symmetric fourth-order identity tensor

Isym = 12 (δik δjl + δil δjk )Eij ⊗ Ekl


is the symmetric identity tensor. It maps any second-order tensor to its symmetric part.
Check:
Isym T ≡ Isym : T = 1
2 (δik δjl + δil δjk )Eij ⊗ Ekl : (Tpq Epq )
1
= 2 (δik δjl + δil δjk )Tpq (Ekl : Epq )Eij
1
= 2 (δik δjl + δil δjk )Tpq δkp δlq Eij
1
= 2 (δip δjq + δiq δjp )Tpq Eij
1
= 2 (Tij + Tji )Eij
= sym T.
If T is already symmetric, then Isym functions as an alternate identity tensor.

Transformation rules for fourth-order tensors


The transformation rule for the components of a fourth-order tensor under a change in basis
from {ei } to {e∗i } with e∗i = Qij ej can be determined in the same way as for second-order
tensors by first noting that
E∗ij = e∗i ⊗ e∗j = Qip Qjq ep ⊗ eq = Qip Qjq Epq .
Then we obtain

Cijkl = E∗ij : CE∗kl ,
= Qip Qjq Qkr Qls Epq : CErs ,
or

Cijkl = Qip Qjq Qkr Qls Cpqrs . (1.2.63)
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

2 Vectors and tensors: Analysis

In continuum mechanics, the scalars, vectors, and tensors whose properties that we discussed
in Chapter 1 are used to represent physical quantities such as temperature, displacement, stress,
and the like. These quantities are typically functions of position within a body. For example,
the temperature at the left end of a rod ϑ(xleft ) could differ from the temperature at the right
end of the rod ϑ(xright ), in which case physical experience teaches us that there will be a
flow of energy from the hotter end to the cooler end, and the flow rate is dependent on the
gradient (rate of change with respect to position) of the temperature. In the general case, the
temperature is what we term a scalar field, a function of position, ϑ(x), defined over the domain
of the material body, Ω, of interest. We will also encounter vector fields and tensor fields in what
follows. An example of a vector field would be the displacement field in a deformed body, u(x).
An example of a tensor field would be the stresses in a body under load, σ(x). In each of these
cases we will be interested in computing rates of change (from position to position) and in
integral manipulations of the fields.

2.1 Directional derivatives and gradients


In order to compute rates of change with position, consider the domain Ω shown in Fig. 2.1
and an arbitrary scalar field ϕ(x) defined over Ω. If we wish to discuss the rate of change of ϕ
at a point x, then we first need to specify along which direction in space we are interested in
finding the rate of change. We will define this direction by the vector h.
The rate of change of ϕ at the point x in the direction h is known as the directional derivative,
and defined by,
  
ϕ(x + αh) − ϕ(x) d 
ϕ(x + αh)
def
Dϕ(x)[h] = lim = . (2.1.1)
α→0 α dα α=0

The notation Dϕ(x)[h] stands for an operator Dϕ(x) operating linearly on h. The linear
operator Dϕ(x) is the gradient of ϕ at x. Other common notations for the gradient are
grad ϕ(x) and ∇ϕ(x). It should be noted that the computed rate of change will be per unit h.
Thus it is common to assume |h| = 1, so that the directional derivative provides rates of change
of ϕ per unit x. The expression after the last equality in (2.1.1) is often the most convenient way
of computing the directional derivative.

Continuum Mechanics of Solids. Lallit Anand and Sanjay Govindjee, Oxford University Press (2020).
© Lallit Anand and Sanjay Govindjee, 2020.
DOI: 10.1093/oso/9780198864721.001.0001
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

24 VECTORS AND TENSORS: ANALYSIS

Slope

h
x
Ω

Fig. 2.1 Directional derivative construction for a generic scalar field.

EXAMPLE 2.1 Gradient of a scalar field

The gradient of ϕ(x) = x · a + x · x, where a is a constant vector, is given by first finding


the directional derivative as

d 
grad ϕ(x)[h] = ϕ(x + αh)
dα α=0

d 
= [(x + αh) · a + (x + αh) · (x + αh)] = h · a + 2x · h
dα α=0
= (a + 2x) · h.
Since (a + 2x) · h is linear in h, the gradient of ϕ(x) is,
grad ϕ(x) = a + 2x,
which is a vector field.

The definition given also applies to vector and tensor fields. Thus for a vector field v(x), the
directional derivative at a point x in the direction h is given by
  
v(x + αh) − v(x) d 
Dv(x)[h] = lim = v(x + αh) , (2.1.2)
α→0 α dα α=0

and similarly for tensor fields. It should be noted that in the case of a vector field, the
directional derivative is also a vector each of whose components gives the rate of change of
the corresponding component of v in the direction of h. The gradient in this case will be a
tensor field (that when applied to h gives the directional derivative of v in the direction of h).

2.2 Component expressions for differential operators


Let ϕ(x) and v(x) be scalar and vector fields. The gradients of these fields
grad ϕ(x) and grad v(x),
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

COMPONENT EXPRESSIONS FOR DIFFERENTIAL OPERATORS 25

are, respectively, vector and tensor fields, whose components are given by
  ∂ϕ(x)
ei · grad ϕ(x) = grad ϕ(x) i = ,
∂xi
(2.2.1)
  ∂vi (x)
ei · grad v(x)ej = grad v(x) ij
= .
∂xj
The latter equalities for each case show that the components of the gradient expressions are
simply the derivatives of the components in the coordinate directions.

EXAMPLE 2.2 Components of the gradient of a scalar field

To show the correctness of the component expression (2.2.1)1 expand Definition (2.1.1):

d  ∂ϕ ∂ϕ ∂ϕ
ϕ(x1 + αh1 , x2 + αh2 , x3 + αh3 ) = h1 + h2 + h3
dα α=0 ∂x 1 ∂x 2 ∂x 3
∂ϕ
= hi .
∂xi

(grad ϕ)i

2.2.1 Divergence, curl, and Laplacian


Three other important differential operators are the divergence, curl, and Laplacian. The
divergence and curl of a vector field v and a tensor field T may be defined as follows:
∂vi
div v = tr[grad v] = ,
∂xi
∂vk
(curl v)i = eijk ,
∂xj
(2.2.2)
∂Tij
(div T)i = ,
∂xj
∂Tjq
(curl T)ij = eipq ;
∂xp
thus div v is a scalar field, curl v and div T are vector fields, and curl T is a tensor field.1

1
It also common to see the alternate notations ∇· v = div v, ∇× v = curl v, ∇· T = div T, and ∇× T = curl T
for the divergence and curl of vectors and tensors. Also, the definition of the curl of a tensor field is not consistent in the
literature, and one needs to be aware of the definition being used by a particular author. We will employ the definition
given in (2.2.2), but it is common to also see the alternate definition (curl T)ij = eipq Tqj,p .
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

26 VECTORS AND TENSORS: ANALYSIS

The Laplacian or Laplace operator  for scalar fields ϕ and vector fields v are defined as
follows:
∂2ϕ
ϕ = div grad ϕ, ϕ = ,
∂xi ∂xi
(2.2.3)
∂ 2 vi
v = div grad v, vi = ;
∂xj ∂xj
thus, ϕ is a scalar field and v is a vector field. Further, for a tensor field T, T is the tensor
field defined such that
(T)a = (Ta) for every constant vector a, (2.2.4)
or equivalently, using components,
∂Tij
Tij = . (2.2.5)
∂xk ∂xk
Notation:
In indicial notation it is customary to employ a comma to denote partial differentiation with
respect to a spatial coordinate, as follows:
∂ϕ ∂vi ∂Tij
= ϕ,i , = vi,j , = Tij,k , etc.
∂xi ∂xj ∂xk
Thus, the Laplacians of ϕ and v may be written as
Δϕ = ϕ,ii and Δvi = vi,jj ,
and so on. We will use this shorthand notation whenever convenient.

EXAMPLE 2.3 Gradient and divergence of a vector field

The gradient of the vector field


v(x) = (x · x)x,
can be computed by first noting that vi = xk xk xi , then the components of the gradient
are given by
vi,j = (xk xk xi ),j
= δkj xk xi + xk δkj xi + xk xk δij
= 2xi xj + xk xk δij .
In the computation we have taken advantage of the fact that ∂xi /∂xj = δij , since the
coordinate directions are independent of each other. Thus,
grad v = 2x ⊗ x + (x · x)1.
The divergence follows immediately as,
div v = tr[grad v] = 2x · x + (x · x)3 = 5x · x.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

GENERALIZED DERIVATIVES 27

EXAMPLE 2.4 Divergence of a tensor field

The divergence of the tensor field


T = βx2 (e1 ⊗ e3 + e3 ⊗ e1 ) − βx1 (e2 ⊗ e3 + e3 ⊗ e2 ),
where β is a constant scalar, can be found by applying the definition for its components
Tij,j : ⎡ ⎤
∂x2
⎢ β ⎥ ⎡ ⎤
⎢ ∂x3 ⎥
⎢ ⎥ 0
∂x1
[div T] = ⎢⎢ −β ⎥ = ⎣ 0 ⎦.

⎢ ∂x3 ⎥
⎣ ∂x2 0
∂x1 ⎦
β −β
∂x1 ∂x2

2.3 Generalized derivatives


It is also possible to define derivatives of functions with respect to arguments that are not simply
position. For example, if one has a function of the displacement vector, it is possible to define the
derivative of the function with respect to the displacement vector. The meaning is, as expected,
the rate of change of the function with respect to changes in the displacement vector.
Consider a function f (ϕ, v, T) whose value is possibly scalar, vector, or tensor valued, and
where ϕ, v, and T are scalar, vector, and tensor arguments. The derivatives of f with respect to
its arguments are defined as follows:

• The derivative ∂f /∂ϕ is the expression such that



d  ∂f
f (ϕ + αh, v, T) = h
dα α=0 ∂ϕ
for all scalars h.

• The derivative ∂f /∂ v is the expression such that



d  ∂f
f (ϕ, v + αh, T) = ·h
dα α=0 ∂v
for all vectors h.

• The derivative ∂f /∂ T is the expression such that



d  ∂f
f (ϕ, v, T + αH) = :H
dα α=0 ∂ T
for all tensors H.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

28 VECTORS AND TENSORS: ANALYSIS

EXAMPLE 2.5 Derivative of a scalar valued function with respect to a vector

The derivative of the scalar valued function


n(v) = |v|,
that is the norm of v, is determined as follows:

n(v + αh) = |v + αh| = (v + αh) · (v + αh)
 
d 
 d  
n(v + αh) = (v + αh) · (v + αh)
dα dα
α=0
 α=0

1 1
=  (h · (v + αh) + (v + αh) · h)
2 (v + αh) · (v + αh)
α=0
1 1
= (h · v + v · h)
2 |v |
v
= · h.
|v|
Thus,
∂n v
= .
∂v |v|

EXAMPLE 2.6 Derivative of a vector valued function with respect to a tensor

The derivative of the vector valued function


w(T) = TTa,
where a is a constant vector, is found as follows. Since,
w(T + αH) = (T + αH)(T + αH)a, and

d 
Dw(T)[H] = w(T + αH) = [H(T + αH)a + (T + αH)Ha]α=0 = HTa + THa,
dα α=0

using components we obtain,


 
Dw(T) ijk Hjk = Hip Tpq aq + Tis Hst at ,
 
= δij δpk Tpq aq + Tis δsj δtk at Hjk ,
 
= δij Tkq aq + Tij ak Hjk ,
which gives,  
D w( T ) ijk
= δij Tkq aq + Tij ak .
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

INTEGRAL THEOREMS 29

Thus in direct notation,


∂w
D w( T ) ≡ = 1 ⊗ Ta + T ⊗ a,
∂T
which is a third-order tensor.

EXAMPLE 2.7 Derivative of a tensor with itself

The derivative of a tensor with itself can be found by considering the function
G(T) = T,
so that
∂T
DG(T)[H] = H.
∂T
Expanding gives

d 
(T + αH) = H = IH.
dα α=0

Thus
∂T ∂Tij
= I, = δik δjl .
∂T ∂Tkl
In the case that T is symmetric this result does not possess the correct (minor) symmetries
upon exchange of i ↔ j or k ↔ l. For the symmetric case, with some effort, it can be
shown that the appropriate expression should be ∂ T/∂ T = Isym , ∂Tij /∂Tkl = 12 (δik δjl +
δil δjk ).

2.4 Integral theorems


In addition to the differential operations introduced above, we will also have need for integral
operations on various fields. There are three basic theorems that we will need: (i) the local-
ization theorem, (ii) the divergence theorem, and (iii) the Stokes theorem. Each of these is
discussed below.

2.4.1 Localization theorem


Localization Theorem Let f (x) be an arbitrary scalar, vector, or tensor field, continuous at
all points x in a domain Ω. If R f (x) dv = 0 for all subdomains R ⊂ Ω, then f (x) = 0 for all
x ∈ Ω.
This theorem is central to deriving the local equations governing the behavior of bodies from
global (laboratory level) observations of physical phenomena.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

30 VECTORS AND TENSORS: ANALYSIS

2.4.2 Divergence theorem


Divergence Theorem Let R be a bounded region with boundary ∂R. Assume we are given a
scalar field ϕ, a vector field v, and a tensor field T, over the domain R. Let n denote the outward
unit normal field on the boundary ∂R of R. Then
 
ϕ n da = grad ϕ dv,
∂R R
 
v · n da = div v dv, (2.4.1)
∂R R
 
Tn da = div T dv.
∂R R

The identities (2.4.1) have the component forms:


 
∂ϕ
ϕni da = dv,
∂xi
∂R R
 
∂vi
vi ni da = dv, (2.4.2)
∂xi
∂R R
 
∂Tij
Tij nj da = dv.
∂xj
∂R R

Note that these identities follow a general rule: the ni in the surface integral results in the partial
derivative ∂/∂xi in the volume integral; thus, for a tensor of any order,
 
∂Tij···k
Tij···k nr da = dv. (2.4.3)
∂xr
∂R R

The central utility of the divergence theorem is that it allows one to convert surface expres-
sions/information to volume expressions/information. Later we will see that the divergence
theorem together with the localization theorem will allow us to derive important governing
equations for many physical phenomena.

EXAMPLE 2.8 Application of the divergence theorem

Considering a domain Ω with volume vol(Ω) and a constant tensor ε, find the value of
the surface integral 
1
εij xj ni da.
vol(Ω)
∂Ω
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

INTEGRAL THEOREMS 31

This integral can be computed using (2.4.3) as


 
1 1
εij xj ni da = (εij xj ),i dv
vol(Ω) vol(Ω)
∂Ω Ω

1
= εij xj,i dv
vol(Ω)
Ω

1
= εij δji dv
vol(Ω)
Ω

1
= εii dv
vol(Ω)
Ω
vol(Ω)
= εii = εii .
vol(Ω)

2.4.3 Stokes theorem


Stokes Theorem Let ϕ, v, and T be scalar, vector, and tensor fields with common domain R.
Then given any positively oriented surface S , with boundary C a closed curve, in R,
 
ϕ dx = n × grad ϕ da
C S
 
v · dx = n · curl v da (2.4.4)
C S
 
Tdx = (curl T) n da.
C S

The identities (2.4.4) have the component forms:


 
ϕdxi = eijk nj ϕ,k da,
C S
 
∂vk
vi dxi = ni eijk da, (2.4.5)
∂xj
C S
 
∂Tiq
Tij dxj = ejpq nj da.
∂xp
C S

The Stokes theorem plays a very important role in fluid mechanics and theories of elec-
tromagnetic phenomena, to name a few. In the mechanics of solid materials it plays a less
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

32 VECTORS AND TENSORS: ANALYSIS

important role, but we will see one important application of the theorem in the context of
small deformation mechanics.

EXAMPLE 2.9 Green’s theorem in a plane

Applying the Stokes theorem (2.4.5)2 to a planar surface S oriented perpendicular to e3


and for
v = f e1 + g e2
gives Green’s theorem:
  
∂g ∂f 
− da = (f dx1 + gdx2 ). (2.4.6)
∂x1 ∂x2
S C

Further, the special choices f = 0 and g = 0, respectively, imply


     
∂g ∂f
da = gdx2 = gn1 ds, da = − f dx1 = f n2 ds,
∂x1 ∂x2
S C C S C C
(2.4.7)
where we have used the geometric relations
dx2 = n1 ds, dx1 = −n2 ds, (2.4.8)
with ds an elemental arc length of the bounding curve C .
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

PA R T II

Kinematics
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

3 Kinematics

Continuum mechanics, is for the most part, a study of deforming bodies, and kinematics is
the mathematical description of the possible deformations that a body may undergo—it is
the first step in understanding the mechanics of solids. In this chapter we will develop the
necessary tools for describing general deformations and for analyzing them. We will start with
descriptions of deformations of arbitrary magnitude, which leads to non-linear measures of
strain. Following this, we will specialize our results to the important case of small deformations,
which results in linear measures of strain.

3.1 Motion: Displacement, velocity, and acceleration


The shape of a solid body changes with time during a deformation process, and will occupy dif-
ferent regions of space, called configurations, at different times. To characterize a deformation
process, we first adopt any convenient configuration of the body as reference, and set our clocks
to measure time from zero at the moment when the body exists in this reference configuration.
Typically, the reference configuration is taken as an unstressed state. We denote our chosen
reference configuration of the body by B (cf. Fig. 3.1). Points X in B are called material points,
and identified by their position vectors (X −o) which joins them to the origin o of a rectangular
Cartesian coordinate system. The coordinates of a material point are then given by the scalar
product of the vector (X − o) with each of the three mutually orthogonal unit base vectors
{ei | i = 1, 2, 3} of the coordinate system:
Xi = ei · (X − o).
At some later time t the body is in the configuration Bt , and the material point which was
at X would have moved to some position x in Bt . We call Bt the deformed configuration of
the body at time t. The coordinates of the new position vector (x − o) of the material point are
given by
xi = ei · (x − o).
The deformation of the body at each time t is described mathematically by a mapping, the
deformation map,
x = χ (X, t) , xi = χi (X1 , X2 , X3 , t) , (3.1.1)
which gives the place occupied by the material point X at time t. The deformation map
χ describes the motion of the body; it maps points in the reference configuration B to the
deformed configuration Bt (Fig. 3.1). Clearly, χ (X, 0) = X. We assume that every x ∈ Bt

Continuum Mechanics of Solids. Lallit Anand and Sanjay Govindjee, Oxford University Press (2020).
© Lallit Anand and Sanjay Govindjee, 2020.
DOI: 10.1093/oso/9780198864721.001.0001
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

36 KINEMATICS

Reference body

Deformed body
X

Bt
u=x−X
e3
(X−o) x

o (x−o)
e2

e1

Fig. 3.1 Reference configuration and a current deformed configuration of a body.

is the unique image of some X ∈ B , that is the mapping is one-to-one and has a unique
inverse
X = χ−1 (x, t), Xi = χ−1
i (x1 , x2 , x3 , t), (3.1.2)
for every point in the body.
The vector
u(X, t) = χ(X, t) − X, ui (X1 , X2 , X3 , t) = χi (X1 , X2 , X3 , t) − Xi , (3.1.3)
represents the displacement of X at time t.
The partial derivatives


u̇(X, t) = χ̇(X, t) = χ(X, t),
∂t
(3.1.4)
∂2
ü(X, t) = χ̈(X, t) = 2 χ(X, t),
∂t
represent the velocity and acceleration (vectors) of material points, respectively. Using the
inverse map (3.1.2) we can alternatively describe the velocity and acceleration as functions
of (x, t):
 
v(x, t) ≡ χ̇ χ−1 (x, t), t , (3.1.5)
 −1 
v̇(x, t) ≡ χ̈ χ (x, t), t . (3.1.6)
These then provide the velocity and acceleration of the material particle currently occupying
the position x in space.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

MOTION: DISPLACEMENT, VELOCITY, AND ACCELERATION 37

3.1.1 Example motions


There are a number of simple motions that are useful to keep in mind as we develop the tools
needed to analyze arbitrary deformations of solid bodies.

Simple shear
Simple shear is a deformation that is similar to the deformation that occurs when one slides a
stack of papers or a deck of cards. The motion involves only displacement in the e1 -direction
and that motion is proportional to a parameter γ that controls the magnitude of the motion;
cf. Fig. 3.2. Expressed mathematically, one has
χ(X1 , X2 , X3 , t) = (X1 + γ(t)X2 )e1 + X2 e2 + X3 e3 ,
x1 = X1 + γ(t)X2 , x 2 = X2 , x 3 = X3 . (3.1.7)

Elementary elongation
Elementary elongation is a motion which involves a displacement only in the e1 -direction and
that motion is proportional to a parameter α that controls the magnitude of the motion; cf. Fig.
3.3. Mathematically, one has
χ(X1 , X2 , X3 , t) = α(t)X1 e1 + X2 e2 + X3 e3 , x1 = α(t)X1 , x 2 = X2 , x 3 = X3 .
(3.1.8)

Rigid motion
A rigid motion is a motion in which the distance between any two points in the body remains
constant at all times. Such a motion can be described by a translation vector c(t), a rotation
tensor Q(t), and a fixed point, say, o, as
χ(X, t) = c(t) + Q(t)(X − o). (3.1.9)

e2

1
γ

e1

Fig. 3.2 Simple shear deformation. Dashed outline indicates the deformed body.

e2
αL

L
e1

Fig. 3.3 Elementary elongation deformation. Dashed outline indicates the deformed body.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

38 KINEMATICS

EXAMPLE 3.1 Distance between two points in a rigid motion

Verify that the distance between any two points in a rigid body remains constant during
a rigid motion. Consider any two material points Xa and Xb . At any time t the distance
between these points is given by
|xa − xb | = |c(t) + Q(t)(Xa − o) − c(t) − Q(t)(Xb − o)|
= |Q(t)(Xa − Xb )|

= Q(t)(Xa − Xb ) · Q(t)(Xa − Xb )

= (Xa − Xb ) · Q(t) Q(t)(Xa − Xb )

= (Xa − Xb ) · (Xa − Xb )
= |Xa − Xb |,
and is hence a constant for all time.

Bending deformation
The first three example deformations are relatively simple and good for testing one’s intuition.
This last example motion is a bit more complex, but at the same time also represents an intuitive
motion. Consider a beam whose axis displaces according to a given function v(X1 , t) and
whose cross-sections rotate (independently of v(X1 , t)) according to a second given function
θ(X1 , t); cf. Fig. 3.4. The resulting motion can be expressed as
χ(X1 , X2 , X3 , t) = [X1 − X2 sin θ(X1 , t)] e1 + [X2 + v(X1 , t) − X2 (1 − cos θ(X1 , t))] e2
+ X3 e3
x1 = X1 − X2 sin θ(X1 , t),
x2 = X2 + v(X1 , t) − X2 (1 − cos θ(X1 , t)) , (3.1.10)
x 3 = X3 .

θ(X1)

e2

v(X1)

e1

Fig. 3.4 Kinematics of a shear deformable beam. Dashed outline indicates the deformed body.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

DEFORMATION AND DISPLACEMENT GRADIENTS 39

3.2 Deformation and displacement gradients


The gradient

∂ ∂
F(X, t) = ∇χ = χ(X, t), Fij = χi (X1 , X2 , X3 , t) , (3.2.1)
∂X ∂Xj

is called the deformation gradient (tensor).


The displacement gradient (tensor) is defined by

∂ ∂
H(X, t) = ∇u = u(X, t), Hij (X1 , X2 , X3 , t) = ui (X1 , X2 , X3 , t).
∂X ∂Xj
(3.2.2)
Thus, since
u(X, t) = x − X = χ(X, t) − X,
the displacement gradient and the deformation gradient tensors are related by

H(X, t) = F(X, t) − 1, Hij = Fij − δij . (3.2.3)

• If at a given time the deformation gradient F, and hence the displacement gradient H, is
independent of X, then the deformation is said to be homogeneous at that time.

EXAMPLE 3.2 Deformation gradient in simple shear

To find deformation gradient that corresponds to simple shear, start with (3.1.7) and apply
(3.2.1):
⎡ ⎤
∂χ1 /∂X1 ∂χ1 /∂X2 ∂χ1 /∂X3 ⎡ ⎤
⎢ ⎥ 1 γ 0
⎢ ∂χ /∂X ∂χ /∂X ∂χ /∂X ⎥
[F] = ⎢

2 1 2 2 2 3 ⎥ = ⎣ 0 1 0 ⎦.

⎣ ⎦ 0 0 1
∂χ /∂X ∂χ /∂X ∂χ /∂X
3 1 3 2 3 3

With these basic definitions, it is now possible to discuss how lines of material particles,
volumes of material, and areas of material in a body transform during deformation. These
relations will then allow us to deduce sensible definitions for stretch and strain in bodies
undergoing arbitrary deformations.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

40 KINEMATICS

3.2.1 Transformation of material line elements


From the definition of the deformation gradient tensor (3.2.1) it follows that
dx = FdX, dxi = Fij dXj . (3.2.4)

The vector dX represents an infinitesimal segment of material at X in B , and the vector dx


represents the deformed image of dX at x in Bt ; cf. Fig. 3.5.
It is important to note that the physical interpretation of F mapping a vector of material
points from the reference configuration to the deformed configuration holds in the limit as
|dX| → 0. This can be appreciated by observing
dx = χ(X(2) , t) − χ(X(1) , t).
If we now replace χ(X(2) , t) = χ(X(1) + dX, t) by its Taylor series expansion in its first
argument about X(1) , then

dx = χ(X(1) , t) + F(X(1) , t)dX + o(|dX|) − χ(X(1) , t)

= F(X(1) , t)dX + o(|dX|)

≈ F(X(1) , t)dX,
where in the last step we have ignored the higher order terms in the Taylor series expansion.
Thus in summary, the tensor F maps short material vectors (line elements) from the reference
configuration to the deformed configuration.
Further, since at each time t the mapping x = χ(X, t) is one-to-one, the determinant of F

∂χ
J ≡ det = det F = 0. (3.2.5)
∂X

Reference body

dX (2)
X Deformed body

X(1)
Bt

dx
x(2)

x(1)

Fig. 3.5 Schematic showing how F maps dX in B to dx in Bt , where dX = X(2) − X(1) and dX = X(2) − X(1) ,
as |dX| → 0.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

DEFORMATION AND DISPLACEMENT GRADIENTS 41

Since F is not singular, the inverse F−1 exists, and hence the following also holds true:
dX = F−1 dx. (3.2.6)

EXAMPLE 3.3 Relative change in length in simple shear

Vertical segments of material in simple shear have a second-order change in relative


length.
In simple shear, F = 1 + γ e1 ⊗ e2 . Consider a short material vector in the reference
body of the form dX = ζ e2 , (ζ 1), its change in relative length is then given by
|FdX| − ζ |F(ζ e2 )| − ζ |ζ(γ e1 + e2 )| − ζ 
= = = = 1 + γ 2 − 1.
ζ ζ ζ
The quantity just computed is the change in length of the material line segment divided
by its original length and is hence the (normal) strain in the vertical direction.
The quantity
def

λ = |Fζ e2 |/ζ = 1 + γ 2 = 1 + ,
defines the stretch of the line segment.
From this computation, we note that stretch and strain at a point depends on the
direction of interest.

3.2.2 Transformation of material volume elements


Consider now a small volume of material at X in the reference body B in the shape of a
parallelepiped whose edges are defined by three short, non-colinear vectors a, b, and c, and
denote this infinitesimal volume of material as
dvR = [a, b, c].
According to (3.2.4) each edge of the parallelepiped will map to a new vector when the body is
deformed, thus defining a new deformed parallelepiped with volume
dv = [Fa, Fb, Fc]
in the deformed body Bt ; cf. Fig. 3.6.
Recalling from (1.2.30) that [Fa, Fb, Fc]/[a, b, c] is equal to J ≡ det F, we find that
dv = J dv R . (3.2.7)
Since an infinitesimal volume element dvR cannot be deformed to zero volume or negative
volume, we require that
J > 0. (3.2.8)
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

42 KINEMATICS

Reference body

nR
B Deformed body
daR
Bt n
dvR

da
dv

Fig. 3.6 Transformation of infinitesimal volume and area elements.

The quantity J , which was introduced in (3.2.5) as a shorthand notation for the determinant
of F, is called the volumetric Jacobian, or simply the Jacobian of the deformation. It is
the appropriate mathematical quantity for describing how infinitesimal volumes of material
transform from the reference to the deformed configuration. One also observes that J is the
appropriate measure of volumetric stretch (ratio of current to original volume) at a point, and
e = J − 1 is an appropriate measure of volumetric strain.

EXAMPLE 3.4 Volume stretch during rigid motion

The ratio of the volume of an infinitesimal volume element of material during rigid motion
to its volume in the reference configuration is unity. To see this compute the Jacobian of
the motion given in (3.1.9).
dv
J= = det F = det Q = 1,
dvR
where the last equality follows from the property that the determinant of a rotation is +1;
see Sec. 1.2.16.

3.2.3 Transformation of material area elements


The construction of an appropriate expression for the transformation of infinitesimal area
elements follows a similar pattern. Consider the oriented area element daR nR at X in the
reference configuration B ; cf. Fig. 3.6. Assuming this area element to be a parallelogram, it
can be defined by two infinitesimal edge vectors a and b, such that
nR daR = a × b.
Upon deformation this area element will map to a new oriented area element
nda = Fa × Fb.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

STRETCH AND ROTATION 43

Exploiting expression (1.2.37) for the cofactor of an invertible tensor, gives Nanson’s For-
mula:
nda = J F− nR daR . (3.2.9)
Nanson’s formula tells us that the appropriate mathematical quantity for describing how area
elements transform from the reference to the deformed configuration is the cofactor of the
deformation gradient. From this one can also define area stretch at a point, λarea = da/daR =
|J F− nR |, and area strain earea = λarea −1. Note that these quantities depend on the orientation
of the area element at the point of interest.

EXAMPLE 3.5 Area stretch in simple shear

To find the area stretch for an area element with normal e1 in three-dimensional simple
shear we need to compute the inverse transpose of the deformation gradient and its
Jacobian:
⎡ ⎤ ⎡ ⎤
1 γ 0  −1  1 −γ 0
[ F] = ⎣ 0 1 0 ⎦ , F =⎣ 0 1 0 ⎦, and
0 0 1 0 0 1
⎡ ⎤
 −  1 0 0
F = ⎣ −γ 1 0 ⎦ .
0 0 1
Since J = 1 in simple shear, it follows that

λarea = |F− e1 | = 1 + γ2.

3.3 Stretch and rotation


As is evident from the tools developed for describing various aspects of a general motion, the
mapping of material vectors by the deformation gradient is a basic concept. If we consider that
the deformation gradient is simply a tensor, then we are led to the observation that its affect on
a vector can be broken down into two steps: the stretching of the vector and then the rotation
of the vector, or alternately the rotation of the vector and then the stretching of the vector.
This interpretation of the action of the deformation gradient is the central content of the polar
decomposition theorem.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

44 KINEMATICS

3.3.1 Polar decomposition theorem


Let F be a tensor with det F > 0. Then there are unique, symmetric, positive definite tensors U
and V and a rotation R such that
F = RU = VR. (3.3.1)
We refer to F = RU and F = VR, respectively, as the right and left polar decompositions of
F. Granted these decompositions, F determines U and V through the relations1


U = F F,
 (3.3.2)

V= FF ,

whereby
V = RUR , (3.3.3)
and
R = FU−1 = V−1 F. (3.3.4)

• U and V are known as the right and left stretch tensors, respectively.

• The left and right stretch tensors are useful in theoretical discussions but are often
problematic to apply because of the need to compute the square root of a tensor. For
that reason, we introduce the right and left Cauchy–Green (deformation) tensors C
and B defined by:

∂χk ∂χk
C = U2 = F F, Cij = Fki Fkj = ,
∂Xi ∂Xj
(3.3.5)
2 
∂χi ∂χj
B = V = FF , Bij = Fik Fjk = .
∂Xk ∂Xk

• Note that the stretch tensors, U and V, as well as the Cauchy–Green tensors, C and B,
are symmetric and positive definite.

• Being symmetric and positive definite, U and V admit spectral representations of the
form
3
U= λ i ri ⊗ ri ,
i=1
(3.3.6)

3
V= λ i li ⊗ li ,
i=1

1
The square root of a symmetric tensor is defined via its spectral decomposition (see Sec. 1.2.18). A symmetric
 √ def  √
tensor S admits the spectral representation S = i si vi ⊗ vi , and its square root is defined as S = i si vi ⊗ vi .
The square root is always taken as the positive square root in this context.
OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

STRETCH AND ROTATION 45

where

– λ1 , λ2 , λ3 > 0, the principal stretches, are the eigenvalues of U and, by (3.3.3),


also of V;
– r1 , r2 , and r3 , the right principal directions, are the eigenvectors of U
Uri = λi ri (no sum on i). (3.3.7)

– l1 , l2 , and l3 , the left principal directions, are the eigenvectors of V:


Vli = λi li (no sum on i). (3.3.8)

• From (3.3.3) and (3.3.6) it follows that


li = Rri , i = 1, 2, 3, and also that R = li ⊗ ri . (3.3.9)

• The tensors C and B have the following forms when expressed in terms of principal
stretches and directions:
3
C= λ2i ri ⊗ ri ,
i=1
(3.3.10)

3
B= λ2i li ⊗ li .
i=1

• Further, since F = RU,



3
F= λ i li ⊗ ri . (3.3.11)
i=1

3.3.2 Properties of the tensors U and C


Consider now infinitesimal undeformed line elements dX(1) and dX(2) and corresponding
deformed line elements, cf. Fig. 3.7,
dx(1) = FdX(1) and dx(2) = FdX(2) . (3.3.12)
Since R R = 1, U = U , and C = U2 , it follows that
dx(1) · dx(2) = (RUdX(1) ) · (RUdX(2) ),

= UdX(1) · UdX(2) , (3.3.13)

= dX(1) · U2 dX(2) ,

= dX(1) · CdX(2) . (3.3.14)


OUP CORRECTED PROOF – FINAL, 13/6/2020, SPi

46 KINEMATICS

Reference body

Deformed body
(2)
dX(1) dX
X
Bt

dx(1)

dx(2)
x

Fig. 3.7 Schematic showing how F maps two non-colinear, infinitesimal line elements at X.

Fiber stretch and strain


A consequence of (3.3.13) is that
|dx(1) | = |UdX(1) |; (3.3.15)
the right stretch tensor U therefore characterizes the deformed length of infinitesimal fibers. Let
dX
dS = |dX| and e= ,
|dX|
denote the magnitude and direction of an infinitesimal undeformed element, and
dx
ds = |dx| and ẽ = ,
|dx|
the magnitude and direction of the corresponding deformed element. Then, by (3.3.15)
ds
def
λ(e) = = |Ue| (3.3.16)
dS
represents the stretch in the direction e at X. Bearing this in mind, we refer to fibers at X in
the direction e as stretched or unstretched according as λ = 1 or λ = 1. Further, one also notes
that
λ2 = e · C(X)e. (3.3.17)
In summary:

• The stretch λ at X relative to any given material direction e is determined by the right stretch
tensor U(X) through the relation λ(e) = |U(X)e|, or equivalently by λ2 = e · Ce.

• The stretch λ at X takes its extremal values when e is one of the right principal directions;
thus the principal stretches are the extremal values. This result follows because the
necessary equations for maximizing (3.3.16) subject to |e| = 1 are given by the relations
Ue = λe, the eigenvalue equations for U.

You might also like