Você está na página 1de 695

Notes

on Continuum Mechanics
Eduardo W.V. Chaves
Lecture Notes
on Numerical Methods
in Engineering and Sciences
Notes on Continuum Mechanics

Lecture Notes on Numerical Methods in
Engineering and Sciences


Aims and Scope of the Series
This series publishes text books on topics of general interest in the field of computational
engineering sciences.
The books will focus on subjects in which numerical methods play a fundamental role for
solving problems in engineering and applied sciences. Advances in finite element, finite
volume, finite differences, discrete and particle methods and their applications are examples
of the topics covered by the series.
The main intended audience is the first year graduate student. Some books define the
current state of a field to a highly specialised readership; others are accessible to final year
undergraduates, but essentially the emphasis is on accessibility and clarity.
The books will be also useful for practising engineers and scientists interested in state of
the art information on the theory and application of numerical methods.


Series Editor
Eugenio Oate
International Center for Numerical Methods in Engineering (CIMNE)
School of Civil Engineering, Technical University of Catalonia (UPC), Barcelona, Spain

Editorial Board
Francisco Chinesta, Ecole Nationale Suprieure d'Arts et Mtiers, Paris, France
Charbel Farhat, Stanford University, Stanford, USA
Carlos Felippa, University of Colorado at Boulder, Colorado, USA
Antonio Huerta, Technical University of Catalonia (UPC), Barcelona, Spain
Thomas J.R. Hughes, The University of Texas at Austin, Austin, USA
Sergio R. Idelsohn, CIMNE-ICREA, Barcelona, Spain
Pierre Ladeveze, ENS de Cachan-LMT-Cachan, France
Wing Kam Liu, Northwestern University, Evanston, USA
Xavier Oliver, Technical University of Catalonia (UPC), Barcelona, Spain
Manolis Papadrakakis, National Technical University of Athens, Greece
Jacques Priaux, CIMNE-UPC Barcelona, Spain & Univ. of Jyvskyl, Finland
Bernhard Schrefler, Universit degli Studi di Padova, Padova, Italy
Genki Yagawa, Tokyo University, Tokyo, Japan
Mingwu Yuan, Peking University, China



Titles:

1. E. Oate, Structural Analysis with the Finite Element Method.
Linear Statics. Volume 1. Basis and Solids, 2009
2. K. Winiewski, Finite Rotation Shells. Basic Equations and
Finite Elements for Reissner Kinematics, 2010
3. E. Oate, Structural Analysis with the Finite Element Method.
Linear Statics. Volume 2. Beams, Plates and Shells, 2013
4. E.W.V. Chaves. Notes on Continuum Mechanics. 2013


Notes on Continuum Mechanics






Eduardo W.V. Chaves
School of Civil Engineering
University of Castilla-La Mancha
Ciudad Real, Spain











ISBN: 978-94-007-5985-5 (HB)
ISBN: 978-94-007-5986-2 (e-book)

Depsito legal: B-29347-2012


A C.I.P. Catalogue record for this book is available from the Library of Congress




Lecture Notes Series Manager: M Jess Samper, CIMNE, Barcelona, Spain

Cover page: Pall Disseny i Comunicaci, www.pallidisseny.com

Printed by: Artes Grficas Torres S.A.,
Morales 17, 08029 Barcelona, Espaa
www.agraficastorres.es




Printed on elemental chlorine-free paper












Notes on Continuum Mechanics
Eduardo W.V. Chaves


First edition, 2013

International Center for Numerical Methods in Engineering (CIMNE), 2013
Gran Capitn s/n, 08034 Barcelona, Spain
www.cimne.com


No part of this work may be reproduced, stored in a retrieval system, or transmitted in any form
or by any means, electronic, mechanical, photocopying, microfilming, recording or otherwise,
without written permission from the Publisher, with the exception of any material supplied
specifically for the purpose of being entered and executed on a computer system, for exclusive
use by the purchaser of the work.







To my Parents


Contents












PREFACE....................................................................................................................................................... XIX
ABBREVIATIONS.......................................................................................................................................... XXI
OPERATORS AND SYMBOLS....................................................................................................................XXIII
SI-UNITS ..................................................................................................................................................... XX

INTRODUCTION................................................................................................................. 1
1 MECHANICS............................................................................................................................................... 1
2 WHAT IS CONTINUUM MECHANICS?.................................................................................................... 1
2.1 Hypothesis of Continuum Mechanics ......................................................................................... 1
2.2 The Continuum............................................................................................................................... 2
3 SCALES OF MATERIAL STUDIES............................................................................................................. 3
3.1 Scale Study of Continuum Mechanics .........................................................................................3
4 THE INITIAL BOUNDARY VALUE PROBLEM (IBVP) ......................................................................... 6
4.1 Solving the IBVP............................................................................................................................. 6
4.2 Simplifying the IBVP...................................................................................................................... 7

1 TENSORS............................................................................................................................. 9
1.1 INTRODUCTION..................................................................................................................................... 9
1.2 ALGEBRAIC OPERATIONS WITH VECTORS ....................................................................................10
1.3 COORDINATE SYSTEMS .....................................................................................................................16
1.3.1 Cartesian Coordinate System....................................................................................................16
1.3.2 Vector Representation in the Cartesian Coordinate System...............................................17
1.3.3 Einstein Summation Convention (Einstein Notation) ........................................................20
1.4 INDICIAL NOTATION.........................................................................................................................20
1.4.1 Some Operators..........................................................................................................................22
1.4.1.1 Kronecker Delta.............................................................................................................22
1.4.1.2 Permutation Symbol ......................................................................................................23
1.5 ALGEBRAIC OPERATIONS WITH TENSORS.....................................................................................28
1.5.1 Dyadic ..........................................................................................................................................28
1.5.1.1 Component Representation of a Second-Order Tensor in the Cartesian
Basis...................................................................................................................................32
1.5.2 Properties of Tensors ................................................................................................................34
1.5.2.1 Tensor Transpose ..........................................................................................................34
1.5.2.2 Symmetry and Antisymmetry.......................................................................................36
1.5.2.3 Cofactor Tensor. Adjugate of a Tensor .....................................................................42
1.5.2.4 Tensor Trace...................................................................................................................42
1.5.2.5 Particular Tensors ..........................................................................................................44
1.5.2.6 Determinant of a Tensor ..............................................................................................45
1.5.2.7 Inverse of a Tensor........................................................................................................48
1.5.2.8 Orthogonal Tensors ......................................................................................................51

Contents
VII
V
NOTES ON CONTINUUM MECHANICS


V
1.5.2.9 Positive Definite Tensor, Negative Definite Tensor and Semi-Definite
Tensors............................................................................................................................. 52
1.5.2.10 Additive Decomposition of Tensors........................................................................ 53
1.5.3 Transformation Law of the Tensor Components................................................................ 54
1.5.3.1 Component Transformation Law in Two Dimensions (2D) ................................. 61
1.5.4 Eigenvalue and Eigenvector Problem.................................................................................... 65
1.5.4.1 The Orthogonality of the Eigenvectors..................................................................... 67
1.5.4.2 Solution of the Cubic Equation................................................................................... 69
1.5.5 Spectral Representation of Tensors........................................................................................ 72
1.5.6 Cayley-Hamilton Theorem....................................................................................................... 76
1.5.7 Norms of Tensors ..................................................................................................................... 78
1.5.8 Isotropic and Anisotropic Tensor........................................................................................... 79
1.5.9 Coaxial Tensors.......................................................................................................................... 80
1.5.10 Polar Decomposition.............................................................................................................. 81
1.5.11 Partial Derivative with Tensors............................................................................................. 83
1.5.11.1 Partial Derivative of Invariants ................................................................................. 85
1.5.11.2 Time Derivative of Tensors....................................................................................... 86
1.5.12 Spherical and Deviatoric Tensors ......................................................................................... 86
1.5.12.1 First Invariant of the Deviatoric Tensor.................................................................. 87
1.5.12.2 Second Invariant of the Deviatoric Tensor............................................................. 87
1.5.12.3 Third Invariant of Deviatoric Tensor ...................................................................... 89
1.6 THE TENSOR-VALUED TENSOR FUNCTION................................................................................. 91
1.6.1 The Tensor Series ...................................................................................................................... 91
1.6.2 The Tensor-Valued Isotropic Tensor Function ................................................................... 92
1.6.3 The Derivative of the Tensor-Valued Tensor Function ..................................................... 94
1.7 THE VOIGT NOTATION.................................................................................................................... 96
1.7.1 The Unit Tensors in Voigt Notation...................................................................................... 97
1.7.2 The Scalar Product in Voigt Notation.................................................................................... 98
1.7.3 The Component Transformation Law in Voigt Notation.................................................. 99
1.7.4 Spectral Representation in Voigt Notation.......................................................................... 100
1.7.5 Deviatoric Tensor Components in Voigt Notation........................................................... 101
1.8 TENSOR FIELDS ................................................................................................................................ 105
1.8.1 Scalar Fields .............................................................................................................................. 106
1.8.2 Gradient..................................................................................................................................... 106
1.8.3 Divergence................................................................................................................................ 111
1.8.4 The Curl .................................................................................................................................... 113
1.8.5 The Conservative Field........................................................................................................... 115
1.9 THEOREMS INVOLVING INTEGRALS ............................................................................................ 117
1.9.1 Integration by Parts ................................................................................................................. 117
1.9.2 The Divergence Theorem...................................................................................................... 117
1.9.3 Independence of Path............................................................................................................. 120
1.9.4 The Kelvin-Stokes Theorem................................................................................................. 121
1.9.5 Greens Identities..................................................................................................................... 122

Appendix A: A GRAPHICAL REPRESENTATION OF A SECOND-ORDER TENSOR..... 125
A.1 PROJECTING A SECOND-ORDER TENSOR ONTO A PARTICULAR DIRECTION..................... 125
A.1.1 Normal and Tangential Components.................................................................................. 125
A.1.2 The Maximum and Minimum Normal Components........................................................ 127
A.1.3 The Maximum and Minimum Tangential Component .................................................... 128
A.2 GRAPHICAL REPRESENTATION OF AN ARBITRARY SECOND-ORDER TENSOR................... 130
A.2.1 Graphical Representation of a Symmetric Second-Order Tensor (Mohrs Circle)...... 134
A.3 THE TENSOR ELLIPSOID................................................................................................................ 138
A.4 GRAPHICAL REPRESENTATION OF THE SPHERICAL AND DEVIATORIC PARTS.................. 139
A.4.1 The Octahedral Vector .......................................................................................................... 139


III
CONTENTS

2 CONTINUUM KINEMATICS........................................................................................145
2.1 INTRODUCTION.................................................................................................................................145
2.2 THE CONTINUOUS MEDIUM..........................................................................................................146
2.2.1 Kinds of Motion.......................................................................................................................147
2.2.1.1 Rigid Body Motion ......................................................................................................147
2.2.2 Types of Configurations .........................................................................................................149
2.2.2.1 Mass Density.................................................................................................................150
2.3 DESCRIPTION OF MOTION .............................................................................................................151
2.3.1 Material and Spatial Coordinates ...........................................................................................151
2.3.2 The Displacement Vector.......................................................................................................152
2.3.3 The Velocity Vector.................................................................................................................152
2.3.4 The Acceleration Vector .........................................................................................................152
2.3.5 Lagrangian and Eulerian Descriptions..................................................................................152
2.3.5.1 Lagrangian Description of Motion............................................................................152
2.3.5.2 Eulerian Description of Motion................................................................................153
2.3.5.3 Lagrangian and Eulerian Variables............................................................................153
2.4 THE MATERIAL TIME DERIVATIVE ..............................................................................................156
2.4.1 Velocity and Acceleration in Eulerian Description............................................................158
2.4.2 Stationary Fields .......................................................................................................................159
2.4.3 Streamlines ................................................................................................................................161
2.5 THE DEFORMATION GRADIENT...................................................................................................163
2.5.1 Introduction..............................................................................................................................163
2.5.2 Stretch and Unit Extension....................................................................................................163
2.5.3 The Material and Spatial Deformation Gradient ................................................................165
2.5.4 Displacement Gradient Tensors (Material and Spatial) .....................................................168
2.5.5 Material Time Derivative of the Deformation Gradient. Material Time Derivative
of the Jacobian Determinant................................................................................................171
2.5.5.1 Material Time Derivative of F . The Spatial Velocity Gradient ..........................171
2.5.5.2 Rate-of-Deformation and Spin Tensors...................................................................172
2.5.5.3 The Material Time Derivative of
1 -
F ......................................................................174
2.5.5.4 The Material Time Derivative of the Jacobian Determinant ................................174
2.6 FINITE STRAIN TENSORS.................................................................................................................176
2.6.1 The Material Finite Strain Tensor .........................................................................................177
2.6.2 The Spatial Finite Strain Tensor (The Almansi Strain Tensor) ........................................181
2.6.3 The Material Time Derivative of Strain Tensors ................................................................183
2.6.3.1 The Material Time Derivative of the Right Cauchy-Green Deformation
Tensor .............................................................................................................................183
2.6.3.2 The Material Time Derivative of the Green-Lagrange Strain Tensor .................183
2.6.3.3 The Material Time Derivative of
1 -
C ......................................................................184
2.6.3.4 Material Time Derivative of the Left Cauchy-Green Deformation Tensor.......184
2.6.3.5 The Material Time Derivative of the Almansi Strain Tensor ...............................185
2.6.4 Interpreting Deformation/Strain Tensors...........................................................................186
2.6.4.1 The Relationship between the Strain and Stretch Tensors ...................................187
2.6.4.2 Change of Angle...........................................................................................................188
2.6.4.3 The Physical Interpretation of the Deformation/Strain Tensor
Components. The Right Stretch Tensor ...................................................................189
2.7 PARTICULAR CASES OF MOTION...................................................................................................191
2.7.1 Homogeneous Deformation..................................................................................................191
2.7.2 Rigid Body Motion...................................................................................................................192
2.8 POLAR DECOMPOSITION OF F .....................................................................................................195
2.8.1 Spectral Representation of Kinematic Tensors...................................................................197
2.8.2 Evolution of the Polar Decomposition................................................................................203
2.8.2.1 The Alternative Way to Express the Rate of Kinematic Tensors........................208
2.9 AREA AND VOLUME ELEMENTS DEFORMATION......................................................................215
2.9.1 Area Element Deformation....................................................................................................215
IX
NOTES ON CONTINUUM MECHANICS


2.9.1.1 The Material Time Derivative of the Area Element .............................................. 217
2.9.2 The Volume Element Deformation ..................................................................................... 218
2.9.2.1 The Material Time Derivative of the Volume Element ........................................ 219
2.9.2.2 Dilatation....................................................................................................................... 220
2.9.2.3 Isochoric Motion. Incompressibility ........................................................................ 220
2.10 MATERIAL AND CONTROL DOMAINS ........................................................................................ 220
2.10.1 The Material Domain............................................................................................................ 220
2.10.2 The Control Domain ............................................................................................................ 221
2.11 TRANSPORT EQUATIONS .............................................................................................................. 222
2.12 CIRCULATION AND VORTICITY................................................................................................... 224
2.13 MOTION DECOMPOSITION: VOLUMETRIC AND ISOCHORIC MOTIONS.............................. 225
2.13.1 The Principal Invariants ....................................................................................................... 227
2.14 THE SMALL DEFORMATION REGIME......................................................................................... 228
2.14.1 Introduction............................................................................................................................ 228
2.14.2 Infinitesimal Strain and Spin Tensors ................................................................................ 229
2.14.3 Stretch and Unit Extension.................................................................................................. 231
2.14.4 Change of Angle .................................................................................................................... 232
2.14.5 The Physical Interpretation of the Infinitesimal Strain Tensor ..................................... 232
2.14.5.1 Engineering Strain..................................................................................................... 233
2.14.6 The Volume Ratio (Dilatation)............................................................................................ 235
2.14.7 The Plane Strain..................................................................................................................... 236
2.15 OTHER WAYS TO DEFINE STRAIN.............................................................................................. 239
2.15.1 Motivation............................................................................................................................... 239
2.15.2 The Logarithmic Strain Tensor ........................................................................................... 241
2.15.3 The Biot Strain Tensor ......................................................................................................... 242
2.15.4 Unifying the Strain Tensors ................................................................................................. 242
2.15.5 One Dimensional Measurements of Strain (1D).............................................................. 243
2.15.5.1 Cauchys strain or Engineering strain or the Linear strain.................................. 243
2.15.5.2 The Logarithmic or True strain............................................................................... 243
2.15.5.3 The Green-Lagrange strain...................................................................................... 243
2.15.5.4 The Almansi strain ....................................................................................................243
2.15.5.5 The Swaiger strain ..................................................................................................... 244
2.15.5.6 The Kuhn strain......................................................................................................... 244


3 STRESS .............................................................................................................................245
3.1 INTRODUCTION................................................................................................................................ 245
3.2 FORCES ............................................................................................................................................... 245
3.2.1 Surface Forces (Traction) ....................................................................................................... 245
3.2.2 Gravitational Force (Body Force) ......................................................................................... 246
3.3 STRESS TENSORS...............................................................................................................................247
3.3.1 The Cauchy Stress Tensor...................................................................................................... 248
3.3.1.1 The Traction Vector....................................................................................................248
3.3.1.2 Cauchys Fundamental Postulate .............................................................................. 248
3.3.2 The Relationship between the Traction and the Cauchy Stress Tensor ......................... 252
3.3.3 Other Measures of Stress ....................................................................................................... 260
3.3.3.1 The First Piola-Kirchhoff Stress Tensor ................................................................. 260
3.3.3.2 The Kirchhoff Stress Tensor ..................................................................................... 262
3.3.3.3 The Second Piola-Kirchhoff Stress Tensor............................................................. 262
3.3.3.4 The Biot Stress Tensor ............................................................................................... 264
3.3.3.5 The Mandel Stress Tensor.......................................................................................... 264
3.3.4 Spectral Representation of the Stress Tensors.................................................................... 265


4 OBJECTIVITY OF TENSORS ........................................................................................269
4.1 INTRODUCTION................................................................................................................................ 269
X
CONTENTS

4.2 THE OBJECTIVITY OF TENSORS.....................................................................................................270
4.2.1 The Deformation Gradient ....................................................................................................272
4.2.2 Kinematic Tensors...................................................................................................................273
4.2.3 Stress Tensors...........................................................................................................................275
4.3 TENSOR RATES..................................................................................................................................277
4.3.1 Objective Rates.........................................................................................................................278
4.3.1.1 The Convective Rate ...................................................................................................279
4.3.1.2 The Oldroyd Rate ........................................................................................................279
4.3.1.3 The Cotter-Rivlin Rate................................................................................................280
4.3.1.4 The Jaumann-Zaremba Rate ......................................................................................280
4.3.1.5 The Green-Naghdi Rate (Polar Rate) .......................................................................282
4.3.2 The Objective Rate of Stress Tensors ..................................................................................282


5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS................. 285
5.1 INTRODUCTION.................................................................................................................................285
5.2 DENSITY.............................................................................................................................................285
5.2.1 Mass Density.............................................................................................................................286
5.3 FLUX....................................................................................................................................................286
5.4 THE REYNOLDS TRANSPORT THEOREM......................................................................................287
5.4.1 Reynolds Transport Theorem for Volumes with Discontinuities...................................288
5.5 CONSERVATION LAW.......................................................................................................................291
5.6 THE PRINCIPLE OF CONSERVATION OF MASS. THE MASS CONTINUITY EQUATION........291
5.6.1 The Mass Continuity Equation in Lagrangian Description...............................................293
5.6.2 Incompressibility......................................................................................................................295
5.6.3 The Mass Continuity Equation for Volume with Discontinuities ...................................295
5.7 THE PRINCIPLE OF CONSERVATION OF LINEAR MOMENTUM. THE EQUATIONS OF
MOTION...........................................................................................................................................297
5.7.1 Linear Momentum...................................................................................................................297
5.7.2 The Principle of Conservation of Linear Momentum.......................................................297
5.7.2.1 The Equilibrium Equations........................................................................................298
5.7.3 The Equations of Motion with Discontinuities ..................................................................301
5.8 THE PRINCIPLE OF CONSERVATION OF ANGULAR MOMENTUM. SYMMETRY OF THE
CAUCHY STRESS TENSOR..............................................................................................................302
5.8.1 Angular Momentum................................................................................................................302
5.8.2 The Principle of Conservation of Angular Momentum....................................................303
5.9 THE PRINCIPLE OF CONSERVATION OF ENERGY. THE ENERGY EQUATION.....................307
5.9.1 Kinetic Energy..........................................................................................................................307
5.9.2 External and Internal Mechanical Power .............................................................................307
5.9.3 The Balance of Mechanical Energy.......................................................................................310
5.9.4 The Internal Energy.................................................................................................................312
5.9.5 Thermal Power .........................................................................................................................313
5.9.6 The First Law of Thermodynamics. The Energy Equation..............................................314
5.9.6.1 The Energy Equation in Lagrangian Description...................................................315
5.9.7 The Energy Equation with Discontinuity............................................................................316
5.10 THE PRINCIPLE OF IRREVERSIBILITY. ENTROPY INEQUALITY.............................................318
5.10.1 The Second Law of Thermodynamics................................................................................318
5.10.2 The Clausius-Duhem Inequality..........................................................................................320
5.10.3 The Clausius-Planck Inequality............................................................................................321
5.10.4 The Alternative Form to Express the Clausius-Duhem Inequality ...............................321
5.10.5 The Alternative Form of the Clausius-Planck Inequality................................................323
5.10.6 Reversible Process .................................................................................................................323
5.10.7 Entropy Inequality for a Domain with Discontinuity......................................................324
5.11 FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS..................................................326
5.11.1 Particular Cases ......................................................................................................................327
5.11.1.1 Rigid Body Motion ....................................................................................................327
XI
NOTES ON CONTINUUM MECHANICS



5.11.1.2 Flux Problems ............................................................................................................ 327
5.12 FLUX PROBLEMS............................................................................................................................. 328
5.12.1 Heat Transfer ......................................................................................................................... 328
5.12.1.1 Thermal Conduction................................................................................................. 328
5.12.1.2 Thermal Convection Transfer ................................................................................. 330
5.12.1.3 Thermal Radiation..................................................................................................... 330
5.12.1.4 The Heat Flux Equation........................................................................................... 330
5.13 FLUID FLOW IN POROUS MEDIA (FILTRATION) ...................................................................... 334
5.14 THE CONVECTION-DIFFUSION EQUATION............................................................................. 335
5.14.1 The Generalization of the Flux Problem........................................................................... 338
5.15 INITIAL BOUNDARY VALUE PROBLEM (IBVP) AND COMPUTATIONAL MECHANICS ...... 338


6 INTRODUCTION TO CONSTITUTIVE EQUATIONS.............................................. 341
6.1 INTRODUCTION................................................................................................................................ 341
6.2 THE CONSTITUTIVE PRINCIPLES................................................................................................... 343
6.2.1 The Principle of Determinism............................................................................................... 344
6.2.2 The Principle of Local Action ............................................................................................... 344
6.2.3 The Principle of Equipresence .............................................................................................. 344
6.2.4 The Principle of Objectivity................................................................................................... 344
6.2.5 The Principle of Dissipation.................................................................................................. 344
6.3 CHARACTERIZATION OF CONSTITUTIVE EQUATIONS FOR SIMPLE THERMOELASTIC
MATERIALS...................................................................................................................................... 345
6.4 CHARACTERIZATION OF THE CONSTITUTIVE EQUATIONS FOR A THERMO-
VISCOELASTIC MATERIAL............................................................................................................. 351
6.4.1 Constitutive Equations with Internal Variables.................................................................. 355
6.5 SOME EXPERIMENTAL EVIDENCE................................................................................................ 360
6.5.1 Behavior of Solids.................................................................................................................... 360
6.5.1.1 Temperature Effect ..................................................................................................... 362
6.5.1.2 Some Mechanical Properties of Solids ..................................................................... 362
6.5.2 Behavior of Fluids ................................................................................................................... 369
6.5.2.1 Viscosity ........................................................................................................................ 370
6.5.3 Behavior of Viscoelastic Materials ........................................................................................ 371
6.5.4 Rheological Models ................................................................................................................. 372


7 LINEAR ELASTICITY ....................................................................................................375
7.1 INTRODUCTION................................................................................................................................ 375
7.2 INITIAL BOUNDARY VALUE PROBLEM OF LINEAR ELASTICITY............................................. 376
7.2.1 Governing Equations.............................................................................................................. 376
7.2.2 Initial and Boundary Conditions ........................................................................................... 377
7.3 GENERALIZED HOOKES LAW...................................................................................................... 377
7.3.1 The Generalized Hookes Law in Voigt Notation............................................................. 378
7.3.2 The Component Transformation Law for the Generalized Hookes Law .................... 379
7.3.2.1 The Matrix Transformation for Stress and Strain Components.......................... 380
7.3.2.2 The Transformation Matrix of the Elasticity Tensor Components .................... 381
7.4 THE ELASTICITY TENSOR............................................................................................................... 381
7.4.1 Anisotropy and Isotropy......................................................................................................... 381
7.4.2 Types of Elasticity Tensor Symmetry................................................................................... 382
7.4.2.1 Triclinic Materials ........................................................................................................ 382
7.4.2.2 Monoclinic Symmetry (One Plane of Symmetry)................................................... 383
7.4.2.3 Orthotropic Symmetry (Two Planes of Symmetry) ............................................... 384
7.4.2.4 Tetragonal Symmetry .................................................................................................. 384
7.4.2.5 Transversely Isotropic Symmetry (Hexagonal Symmetry).................................... 386
7.4.2.6 Cubic Symmetry........................................................................................................... 388
7.4.2.7 Symmetry in All Directions (Isotropy)..................................................................... 390
XII
CONTENTS

7.5 ISOTROPIC MATERIALS....................................................................................................................392
7.5.1 Constitutive Equations............................................................................................................392
7.5.2 Experimental Determination of Elastic Constants.............................................................393
7.5.2.1 Youngs Modulus and Poissons Ratio.....................................................................393
7.5.2.2 The Shear and Bulk Moduli........................................................................................394
7.5.3 Restrictions on Elastic Mechanical Properties ....................................................................398
7.6 STRAIN ENERGY DENSITY..............................................................................................................399
7.6.1 Decoupling Strain Energy Density........................................................................................402
7.7 THE CONSTITUTIVE LAW FOR ORTHOTROPIC MATERIAL.......................................................404
7.8 TRANSVERSELY ISOTROPIC MATERIALS ......................................................................................405
7.9 THE SAINT-VENANTS AND SUPERPOSITION PRINCIPLES .......................................................406
7.10 INITIAL STRESS/STRAIN................................................................................................................408
7.10.1 Thermal Deformation...........................................................................................................408
7.11 THE NAVIER-LAM EQUATIONS.................................................................................................410
7.12 TWO-DIMENSIONAL ELASTICITY................................................................................................410
7.12.1 The State of Plane Stress ......................................................................................................411
7.12.1.1 The Initial Strain.........................................................................................................412
7.12.2 The State of Plane Strain ......................................................................................................413
7.12.2.1 Thermal Strain............................................................................................................415
7.12.3 Axisymmetric Solids ..............................................................................................................417
7.13 THE UNIDIMENSIONAL APPROACH............................................................................................418
7.13.1 Beam Structural Elements ....................................................................................................418
7.13.1.1 The Internal Normal Force and the Bending Moments .....................................420
7.13.1.2 The Shear Forces and the Torsional Moment ......................................................421
7.13.1.3 The Strain Energy......................................................................................................422


8 HYPERELASTICITY...................................................................................................... 423
8.1 INTRODUCTION.................................................................................................................................423
8.2 CONSTITUTIVE EQUATIONS...........................................................................................................424
8.2.1 Elastic Tangent Stiffness Tensors .........................................................................................427
8.2.1.1 The Material Elastic Tangent Stiffness Tensor .......................................................427
8.2.1.2 The Spatial Elastic Tangent Stiffness Tensor..........................................................428
8.2.1.3 The Instantaneous Elastic Tangent Stiffness Tensor.............................................430
8.2.1.4 The Elastic Tangent Stiffness Pseudo-Tensor ........................................................431
8.3 ISOTROPIC HYPERELASTIC MATERIALS .......................................................................................432
8.3.1 The Constitutive Equation in terms of Invariants..............................................................434
8.3.1.1 The Constitutive Equation in terms of and ........................................................434
8.3.1.2 The Constitutive Equation in terms of ..................................................................436
8.3.2 Series Expansion of the Energy Function ...........................................................................436
8.3.3 Constitutive Equations in terms of the Principal Stretches ..............................................437
8.4 COMPRESSIBLE MATERIALS............................................................................................................440
8.4.1 The Stress Tensors...................................................................................................................442
8.4.2 Compressible Isotropic Materials..........................................................................................445
8.4.2.1 Compressible Isotropic Material in terms of the Invariants .................................446
8.5 INCOMPRESSIBLE MATERIALS........................................................................................................447
8.5.1 Geometrical Interpretation.....................................................................................................449
8.5.2 Isotropic Incompressible Hyperelastic Materials................................................................450
8.5.2.1 Series Expansion of the Energy Function for an Isotropic Incompressible
Hyperelastic Materials ..................................................................................................451
8.6 EXAMPLES OF HYPERELASTIC MODELS ......................................................................................451
8.6.1 The Neo-Hookean Material Model.......................................................................................452
8.6.2 The Ogden Material Model ....................................................................................................452
8.6.2.1 The Incompressible Ogden Material Model............................................................452
8.6.2.2 The Hadamard Material Model..................................................................................453
8.6.3 The Mooney-Rivlin Material Model......................................................................................453
XIII
NOTES ON CONTINUUM MECHANICS


XIV
8.6.3.1 Strain Energy Density ................................................................................................. 453
8.6.3.2 The Stress Tensor ........................................................................................................ 454
8.6.4 The Yeoh Material Model ...................................................................................................... 454
8.6.4.1 Strain Energy Density ................................................................................................. 454
8.6.4.2 The Stress Tensor ........................................................................................................ 454
8.6.5 The Arruda-Boyce Material Model ....................................................................................... 454
8.6.6 The Blatz-Ko Hyperelastic Model ........................................................................................ 455
8.6.7 The Saint Venant-Kirchhoff Model ..................................................................................... 455
8.6.7.1 Strain Energy Density ................................................................................................. 455
8.6.7.2 The Stress Tensor ........................................................................................................ 456
8.6.7.3 The Elastic Tangent Stiffness Tensor ...................................................................... 456
8.6.8 The Compressible Neo-Hookean Material Model ............................................................. 457
8.6.8.1 Strain Energy Density ................................................................................................. 457
8.6.8.2 The Stress Tensor ........................................................................................................ 457
8.6.8.3 The Elastic Tangent Stiffness Tensor ...................................................................... 458
8.6.9 The Gent Model ...................................................................................................................... 460
8.6.10 The Statistical Model............................................................................................................. 460
8.6.11 The Eight-Parameter Model ................................................................................................ 461
8.7 ANISOTROPIC HYPERELASTICITY ................................................................................................. 462
8.7.1 Transversely Isotropic Material ............................................................................................. 463


9 PLASTICITY..................................................................................................................... 465
9.1 INTRODUCTION................................................................................................................................ 465
9.2 THE YIELD CRITERION................................................................................................................... 467
9.2.1 The Yield Surface for Anisotropic Materials....................................................................... 468
9.2.1.1 The Yield Surface Gradient........................................................................................ 468
9.2.2 The Yield Surface for Isotropic Materials............................................................................ 468
9.2.3 The Yield Surface for Materials Independent of Pressure................................................ 471
9.2.3.1 The von Mises Yield Criterion .................................................................................. 471
9.2.3.2 The Tresca Yield Criterion......................................................................................... 475
9.2.4 The Yield Criteria for Pressure-Dependent Materials ....................................................... 477
9.2.4.1 The Mohr-Coulomb Criterion................................................................................... 478
9.2.4.2 The Drucker-Prager Yield Criterion......................................................................... 482
9.2.4.3 The Rankine Yield Criterion...................................................................................... 486
9.2.5 Evolution of the Yield Surface.............................................................................................. 489
9.3 PLASTICITY MODELS IN SMALL DEFORMATION REGIME (UNIAXIAL CASES) ..................... 491
9.3.1 Rate-Independent Plasticity Models (Uniaxial Case) ......................................................... 491
9.3.1.1 Perfect Elastoplastic Behavior................................................................................... 491
9.3.1.2 Isotropic Hardening Elastoplastic Behavior ........................................................... 495
9.3.1.3 Kinematic Hardening Elastoplastic Behavior ......................................................... 500
9.3.1.4 Isotropic-Kinematic Elastoplastic Behavior............................................................ 502
9.4 PLASTICITY IN SMALL DEFORMATION REGIME (THE CLASSICAL PLASTICITY
THEORY).......................................................................................................................................... 503
9.4.1 The Infinitesimal Strain Tensor and Constitutive Equation............................................. 504
9.4.2 Helmholtz Free Energy .......................................................................................................... 505
9.4.3 Internal Energy Dissipation and the Evolution of the Internal Variables ..................... 505
9.4.4 The Elastoplastic Tangent Stiffness Tensor........................................................................ 507
9.4.5 The Classical Flow Theory................................................................................................... 512
9.4.5.1 Perfect Plasticity........................................................................................................... 512
9.4.5.2 Isotropic-Kinematic Hardening Plasticity ............................................................... 513
9.5 PLASTIC POTENTIAL THEORY ....................................................................................................... 515
9.6 PLASTICITY IN LARGE DEFORMATION REGIME........................................................................ 518
9.7 LARGE-DEFORMATION PLASTICITY BASED ON THE MULTIPLICATIVE
DECOMPOSITION OF THE DEFORMATION GRADIENT.......................................................... 518
9.7.1 Kinematic Tensors................................................................................................................... 518
CONTENTS

9.7.1.1 Deformation and Strain Tensors...............................................................................520
9.7.1.2 Area and Volume Elements Deformation...............................................................524
9.7.1.3 The Spatial Velocity Gradient ....................................................................................525
9.7.1.4 The Oldroyd Rate ........................................................................................................528
9.7.1.5 The Cotter-Rivlin Rate................................................................................................529
9.7.2 The Stress Tensors...................................................................................................................531
9.7.2.1 Stress Tensor Rates......................................................................................................532
9.7.3 The Helmholtz Free Energy...................................................................................................533
9.7.3.1 Decoupling the Helmholtz Free Energy..................................................................533
9.7.3.2 The Objectivity Principle for the Helmholtz Free Energy....................................533
9.7.3.3 The Isotropic Helmholtz Free Energy .....................................................................534
9.7.3.4 The Rate of Change of the Isotropic Helmholtz Free Energy.............................534
9.7.4 The Plastic Potential and the Yield Criterion......................................................................536
9.7.5 The Dissipation and the Constitutive Equation..................................................................537
9.7.6 Evolution of the Internal Variables.......................................................................................538
9.7.7 The Elastoplastic Tangent Stiffness Tensors.......................................................................539
9.7.7.1 The Elastoplastic Tangent Stiffness Tensor ............................................................540
9.7.8 The Hyperelastoplastic Model with von Mises Yield Criterion........................................542
9.7.8.1 The Helmholtz Free Energy ......................................................................................542
9.7.8.2 The Stress Tensor ........................................................................................................543
9.7.8.3 Formulation Considering the Transformation as an Isochoric
Transformation..............................................................................................................544
9.7.8.4 The Rate of Change of the Helmholtz Free Energy ..............................................545
9.7.8.5 Yield Criterion and Evolution of the Internal Variables .......................................546


10 THERMOELASTICITY................................................................................................ 547
10.1 THERMODYNAMIC POTENTIALS .................................................................................................547
10.1.1 The Specific Internal Energy ...............................................................................................548
10.1.2 The Specific Helmholtz Free Energy .................................................................................548
10.1.3 The Specific Gibbs Free Energy .........................................................................................549
10.1.4 The Specific Enthalpy ...........................................................................................................550
10.2 THERMOMECHANICAL PARAMETERS .........................................................................................552
10.2.1 Isothermal and Isentropic Processes ..................................................................................552
10.2.2 Specific Heats and Latent Heat Tensors............................................................................553
10.3 LINEAR THERMOELASTICITY.......................................................................................................556
10.3.1 Linearization of the Constitutive Equations......................................................................556
10.3.1.1 The Linearized Piola-Kirchhoff Stress Tensor .....................................................557
10.3.1.2 The Linearized Heat Flux Vector............................................................................558
10.3.1.3 Linearized Entropy....................................................................................................560
10.3.1.4 The Helmholtz Free Energy Approach .................................................................560
10.3.1.5 Linearization of the Constitutive Equations..........................................................561
10.3.1.6 Linear Thermoelasticity in a Small Deformation Regime ...................................561
10.3.1.7 Linear Thermoelasticity in a Small Deformation Regime ...................................562
10.4 THE DECOUPLED THERMO-MECHANICAL PROBLEM IN A SMALL DEFORMATION
REGIME............................................................................................................................................565
10.4.1 The Purely Thermal Problem...............................................................................................567
10.4.2 The Purely Mechanical Problem..........................................................................................568
10.5 THE CLASSICAL THEORY OF THERMOELASTICITY IN FINITE STRAIN (LARGE
DEFORMATION REGIME) .............................................................................................................569
10.5.1 The Coupled Heat Flux Equation.......................................................................................570
10.5.2 The Specific Helmholtz Free Energy .................................................................................572
10.6 THERMOELASTICITY BASED ON THE MULTIPLICATIVE DECOMPOSITION OF THE
DEFORMATION GRADIENT..........................................................................................................573
10.6.1 Kinematic Tensors.................................................................................................................574
10.6.2 The Stress Tensor ..................................................................................................................576
XV
NOTES ON CONTINUUM MECHANICS



10.6.3 Area and Volume Elements................................................................................................. 576
10.6.4 Isotropic Materials................................................................................................................. 578
10.6.5 The Constitutive Equations ................................................................................................. 579
10.6.5.1 The Constitutive Equation for Energy .................................................................. 579
10.6.5.2 The Constitutive Equations for Stress ................................................................... 580
10.6.5.3 The Constitutive Equation for Entropy ................................................................ 582
10.7 THERMOPLASTICITY IN A SMALL DEFORMATION REGIME................................................... 583
10.7.1 The Specific Helmholtz Free Energy................................................................................. 583
10.7.2 Internal Energy Dissipation................................................................................................. 584


11 DAMAGE MECHANICS ................................................................................................587
11.1 INTRODUCTION.............................................................................................................................. 587
11.2 THE ISOTROPIC DAMAGE MODEL IN A SMALL DEFORMATION REGIME.......................... 588
11.2.1 Description of the Isotropic Damage Model in Uniaxial Cases .................................... 588
11.2.1.1 The Constitutive Equation....................................................................................... 589
11.2.2 The Three-Dimensional Isotropic Damage Model.......................................................... 590
11.2.2.1 Helmholtz Free Energy ............................................................................................ 590
11.2.2.2 Internal Energy Dissipation and the Constitutive Equations ............................ 591
11.2.2.3 Ingredients of the Damage Model...................................................................... 593
11.2.2.4 The Hardening/Softening Law............................................................................... 599
11.2.3 The Elastic-Damage Tangent Stiffness Tensor ................................................................ 601
11.2.4 The Energy Norms................................................................................................................ 602
11.2.4.1 The Symmetrical Damage Model (Tension-Compression) Model I.............. 602
11.2.4.2 The Tension-Only Damage Model Model II .................................................... 603
11.2.4.3 The Non-Symmetrical Damage Model Model III ............................................ 604
11.3 THE GENERALIZED ISOTROPIC DAMAGE MODEL................................................................. 605
11.3.1 The Strain Energy Function................................................................................................. 606
11.3.2 Spherical and Deviatoric Effective Stress.......................................................................... 607
11.3.3 Thermodynamic Considerations ......................................................................................... 607
11.3.4 The Elastic-Damage Tangent Stiffness Tensor ................................................................ 608
11.4 THE ELASTOPLASTIC-DAMAGE MODEL IN A SMALL DEFORMATION REGIME................ 609
11.4.1 The Elasto-Plastic Damage Model by Sim&Ju (1987) in a Small Deformation
Regime..................................................................................................................................... 610
11.4.1.1 Helmholtz Free Energy ............................................................................................ 610
11.4.1.2 Internal Energy Dissipation. Constitutive Equations. Thermodynamic
Considerations............................................................................................................... 611
11.4.1.3 Damage Characterization......................................................................................... 612
11.4.1.4 The Elastic-Damage Tangent Stiffness Tensor .................................................... 612
11.4.1.5 Characterization of the Plastic Response. The Elastoplastic-Damage
Tangent Stiffness Tensor....................................................................................... 613
11.5 THE TENSILE-COMPRESSIVE PLASTIC-DAMAGE MODEL...................................................... 615
11.5.1 Helmholtz Free Energy ........................................................................................................ 616
11.5.2 Damage Characterization..................................................................................................... 617
11.5.3 Evolution of the Damage Parameters................................................................................ 618
11.5.4 Evolution of the Plastic Strain Tensor............................................................................... 619
11.5.5 Internal Energy Dissipation................................................................................................. 619
11.6 DAMAGE IN A LARGE DEFORMATION REGIME ...................................................................... 621
11.6.1 Gurtin & Francis One-Dimensional Model ..................................................................... 622
11.6.2 The Rate Independent 3D Elastic-Damage Model.......................................................... 622
11.6.3 The Damage Variable. Damage Evolution........................................................................ 623
11.6.4 The Plastic-Damage Model by Sim & Ju (1989) ............................................................ 624
11.6.4.1 Specific Helmholtz Free Energy ............................................................................. 624
11.6.4.2 Internal Energy Dissipation. Constitutive Equations. Thermodynamic
Considerations............................................................................................................... 624
11.6.4.3 Damage Characterization......................................................................................... 626
XVI
CONTENTS

XVII
11.6.4.4 The Hyperelastic-Damage Tangent Stiffness Tensor ..........................................626
11.6.4.5 Characterization of the Plastic Response. The Effective Elastoplastic-
Damage Tangent Stiffness Tensor .............................................................................627
11.6.4.6 The Elastoplastic-Damage Tangent Stiffness Tensor..........................................628
11.6.5 The Plastic-Damage Model by Ju(1989).............................................................................628
11.6.5.1 Helmholtz Free Energy.............................................................................................629
11.6.5.2 Internal Energy Dissipation. Constitutive Equation. Thermodynamic
Considerations...............................................................................................................629
11.6.5.3 Characterization of Damage. The Tangent Damage Hyperelasticity Tensor ..630
11.6.5.4 The Elastic-Damage Tangent Stiffness Tensor ....................................................630
11.6.5.5 Characterization of Plastic Response. The elastoplastic Tangent Stiffness
Tensor. ............................................................................................................................631
11.6.5.6 The Elastoplastic-Damage Tangent Stiffness Tensor..........................................632


12 INTRODUCTION TO FLUIDS.................................................................................... 635
12.1 INTRODUCTION..............................................................................................................................635
12.2 FLUIDS AT REST AND IN MOTION...............................................................................................636
12.2.1 Fluids at Rest ..........................................................................................................................636
12.2.2 Fluids in Motion.....................................................................................................................637
12.3 VISCOUS AND NON-VISCOUS FLUIDS.........................................................................................637
12.3.1 Non-Viscous Fluids (Perfect Fluids) ..................................................................................638
12.3.2 Viscous Fluids.........................................................................................................................638
12.4 LAMINAR TURBULENT FLOW.......................................................................................................639
12.5 PARTICULAR CASES ........................................................................................................................640
12.5.1 Incompressible Fluids ...........................................................................................................640
12.5.2 Irrotational Flow....................................................................................................................641
12.5.3 Steady Flow.............................................................................................................................641
12.6 NEWTONIAN FLUIDS .....................................................................................................................642
12.6.1 The Stokes Condition...........................................................................................................645
12.7 STRESS, DISSIPATED AND RECOVERABLE POWERS.................................................................645
12.8 THE FUNDAMENTAL EQUATIONS FOR NEWTONIAN FLUIDS...............................................647
12.8.1 The Navier-Stokes-Duhem Equations of Motion............................................................648
12.8.1.1 Alternative Form of the Fundamental Equations for Newtonian Fluids.........648
12.8.1.2 The Fundamental Equations for Incompressible Newtonian Fluid..................649
12.8.2 The Navier-Stokes Equations of Motion...........................................................................650
12.8.3 The Euler Equations of Motion..........................................................................................650
12.8.3.1 Non-Viscous and Incompressible Fluids...............................................................651
12.8.3.2 Bernoullis Equation..................................................................................................652
12.8.4 The Equation of Vorticity ....................................................................................................653


BIBLIOGRAPHY .................................................................................................................... 659
INDEX.................................................................................................................................. 667













Preface










The Continuum Mechanics is a key subject to several degrees based on physical science,
such as: Civil Engineering, Industrial Engineering, Meteorology, Magnetism,
Oceanography, Aerodynamics, Hydrodynamics, Marine Engineering, etc.
This book grew out of notes for the course Introduction to Continuum Mechanics of the
career of Civil Engineering of the University of Castilla-La Mancha (Spain), and is intended
for students who are initiating a university degree based on physical science, and is also
intended for PhD students as well researchers.
In order to provide greater clarity for students, this book presents a thorough detail at the
time of the demonstration of the equations. At the time of writing the book, the author has
had a big concern for trying to unify the existing nomenclature, and to this end has
consulted numerous articles and books on the subject. With respect to the notation, the
developments of the equations are indiscriminately presented in tensorial, inditial and Voigt
notations. Another aspect is that the book is self-contained, so that the concepts used are
defined in the text.
Finally, I would like to express my gratitude to: Houzeaux (Guillaume), Vzquez (Mariano),
Gallego (Inmaculada), Pulido (Loli), Bentez (Jos Mara), Casati (Mara Jesus), Vlez
(Eduardo), Solares (Cristina), Olivares (Miguel ngel), Escobedo (Fernando), Simarro
(Gonzalo), Sanz (Ana), for aid to the revision of the first edition in Spanish. I would also
like to thank Toby Wakely for reviewing the English.
I would also to thank two Professors who marked my teaching and research career: Prof.
Xavier Oliver and Prof. Wilson Venturini (in memoriam).


Eduardo W. V. Chaves
Ciudad Real-Spain, October 2012





Preface
XIX




Abbreviations










IBVP Initial Boundary Value Problem
BVP Boundary Value Problem
FEM Finite Element Method
BEM Boundary Element Method
FDM Finite Difference Method




Latin

i.e. id est that is
et al. et alii and the others
e.g. exempli gratia for example
etc. et cetera and so on
v., vs. versus versus
viz. vidilicet namely











Abbreviations
XXI




Operators and Symbols










2
- - -
(-)
Macaulay bracket
- Euclidian norm of -
) (- Tr trace of ) (-
T
) (-
transpose of ) (-
1
) (
-
-
inverse of ) (-
T -
-) (
inverse of the transpose of ) (-
sym
) (-
symmetric part of ) (-
skew
) (-
antisymmetric (skew-symmetric) part of ) (-
sph
) (-
spherical part of ) (-
dev
) (-
deviatoric part of ) (-
- module of -
[ ] [ ] - jump of -

scalar product
( ) - = - det determinant of ( ) -
- =
-
`
Dt
D
material time derivative of ( ) -
) (- cof cofactor of -;
( ) - Adj adjugate of ( ) -
( ) - Tr trace of ( ) -
: double scalar product (or double contraction or double dot product)
2
V
Scalar differential operator
tensorial product
) (- = - grad V gradient of -
) (- = - div V divergence of -
r
vector product (or cross product)








Operators and Symbols
XXIII




SI-Units









length m - metro energy, work, heat Nm J - Joules
mass
kg - kilogram
power W
s
J
= watt
time s - second
temperature K - Kelvin permeability
2
m
velocity
s
m
dynamic viscosity s Pa
acceleration
2
s
m
mass flux
s m
kg
2

energy Nm J - Joules energy flux
s m
J
2

force N - Newton thermal conductivity
mK
W

pressure, stress
2
m
N
Pa = - Pascal mass density
3
m
kg




Prefix Symbol
n
10
Prefix Symbol
n
10
pico
p 12
10
-

kilo k
3
10
nano
q 9
10
-

Mega
M
6
10
micro
6
10
-

Giga G
9
10
mili
m
3
10
-

Tera
T
12
10
centi c
2
10
-


deci d 10










SI-Units
XXV


)
`

Continuum Mechanics
Introduction





1 Mechanics
Broadly speaking, Mechanics is the branch of physics that studies the behavior of a body
when it is subjected to forces, (e.g. deformation) and how it evolves over time. In general,
Mechanics can be classified into:
Theoretical Mechanics;
Applied Mechanics;
Computational Mechanics.
Theoretical Mechanics establishes the laws that govern a particular physical problem based on
fundamental principles.
Applied Mechanics transfers theoretical knowledge to use it in scientific and engineering
problems.
Computational Mechanics solves problems by simulation with numerical tools implemented in
the computer.
In this book we focus our attention to the Theoretical and Applied Mechanics.
2 What is Continuum Mechanics?
Broadly speaking, Continuum Mechanics is the branch of Mechanics that studies motion
(deformation) of a medium that consists of matter subjected to forces. For example, how
would a wooden and a concrete beam deform when the same force is applied to them?
Another example we can look at is fluids, e.g. for a given pressure, how does water (or oil)
flow in a pipeline?
2.1 Hypothesis of Continuum Mechanics
As we know, a physical body consists of small molecules (an agglomeration of two or more
atoms). Then, by means of sophisticated experiments, we can observe that these
constituents are not distributed homogeneously, that is, there are gaps (voids) between
them. However, within the scope of Continuum Mechanics these phenomenological

Introduction
1 , Notes on Continuum Mechanics, Lecture Notes on Numerical
Methods in Engineering and Sciences 4, DOI 10.1007/978-94-007-5986-2_1,
International Center for Numerical Methods in Engineering (CIMNE), 2013
E.W.V. Chaves
NOTES ON CONTINUUM MECHANICS

2
characteristics are ignored. For example, if we are dealing with a fluid in Continuum
Mechanics, the properties: mass density, pressure and velocity are assumed to be
continuous function. Treating a system of molecules as a continuous medium is valid if we
compare the mean free path of molecules ( A ) (average distance particles travel before
colliding with each other) with the characteristic physical length scale (
C
). For example,
for solids and liquids we have cm
7
10
-
= A and for gases cm
6
10
-
= A , Chung (1996). Then
the ratio
C

A
is known as the Knudsen number ( Kn ). If this number is much smaller than
unity, the domain can be treated as a continuum; otherwise we must use statistical
mechanics to obtain the governing equations of the problem whereby we can establish that:
approach c microscopi 1
approach c macroscopi 1
>
A

<<
A

Kn
Kn

The fundamental hypothesis in Continuum Mechanics is that the matter of which the
medium is made up is continuously distributed and that the variables involved in the
problem (e.g. velocity, acceleration, pressure, mass density, etc.) are continuous functions.
Then, by means of approximations or additional equations to that initially proposed for the
problem, we can characterize a continuum with discontinuous variables associated with the
problem, e.g. fracture problem and shock waves among others.
2.2 The Continuum
In general, when we apply force to solids they are able to recover their original states when
said force is removed. However, this is not the case with fluids, i.e. solids and fluids
apparently act very differently. Therefore, traditionally, continuum mechanics has been
divided into two groups: solids and fluids (liquids and gases). As we will see throughout this
book, the fundamental equations of Continuum Mechanics are the same for both of these.
For many decades, solid and fluid mechanics have been treated independently from each
other. However, nowadays, it is not advisable to work like this. Firstly, it is necessary to
simulate more complex materials, e.g. materials that have characteristics of solids and fluids
simultaneously. These materials, besides presenting elastic properties, (obeying the
constitutive law for solids), also exhibit characteristics of fluids due to their viscosity, for
example: viscoelastic materials. Secondly, the need to simulate the problem of fluid-solid
interaction has improved the relationship between fluids and solids.
Recently, a third branch of continuum mechanics has emerged, which is related to
multiphysics problems, characterized by phase change, e.g. from solid to liquid phase or
vice versa, and which includes mechanical systems that transcend classical mechanical
boundary of solids and fluids. Then, traditionally, continuum mechanics can be divided
into:

cs Multiphysi
Gases
Liquids
Fluids
Solids
Continuum The
INTRODUCTION

3
3 Scales of Material Studies
According to Willam(2000), materials science can be studied on different scales, (see Figure
1), namely:
Metric level
At this level, we include most problems posed in Civil, Mechanical, Aerospace
Engineering.
Millimeter level
At this level, it may enroll the specimen used to measure the material mechanical
properties in the laboratory.
Micrometer level
Micro-structural characteristic, such as micro-defects and cement hydration products,
are observed at this scale.
Nanometer level
At this level, we contemplate atomic and molecular processes.

















Figure 1: Multiscale in Material Mechanics, Willam(2000).

3.1 Scale Study of Continuum Mechanics
The continuum mechanics is raised at a macroscopic level. That is, the variables of the
problem at a macroscopic level are considered as being the average of these variables at a

Nano Mechanics
Meso Mechanics
Macro Mechanics
Structural Mechanics
Micro Mechanics
m
0
10 1
m
3
10 1
-

m
6
10 1
-

m
9
10 1
-


NOTES ON CONTINUUM MECHANICS

4
mesoscale level. Let us take, for example, blood, which can be treated in different ways,
depending on the scale under consideration. At a m
6
10
-
scale, we consider blood flows
around a blood cell where the deformation of the cell walls is taken into consideration.
Then, at a m
4
10
-
scale, we can consider the fluid flow through a set of blood cells, which
thus allows us to observe the fluid effects on cells. Next, at a m
3
10
-
scale (macroscopic
level), we can consider the fluid flow through arteries or veins (ignoring the individual cells)
as being a fluid with certain macroscopic properties (e.g. velocity, pressure, etc.), (see Figure
2).



















Figure 2: Scale levels in blood.
Another example we can use is a material made up of a mixture of materials such as
concrete, which is fundamentally formed by mixing cement, aggregates, and water. At the
m
9
10
-
scale, we can distinguish the atomic structure of the cement and aggregates. Then,
at a m
6
10
-
scale it is possible to identify individual cement grains before hydration and
grains of calcium silicate and calcium hydroxide can be appreciated, upon hydration.
Finally, at the m
3
10
-
millimeter scale, we can distinguish individually each of the aggregates
and pores (gaps). Note, at this level, the interaction between parts of cement and aggregates
is important.
On the m 10 metric scale and on the m 1 laboratory scale, the concrete internal structure
can be examined to ensure that its properties are identical in all directions and at all its
points, which is what characterizes a homogenous and isotropic material.
Another example for understanding in which scale continuum mechanics is raised is by
measuring mass density ( j ), which is a macroscopic variable for continuum mechanics.
Blood flow in an artery
(macro scale 10
-3
m)
Meso scale - 10
-4
m
Micro scale - 10
-5
m
INTRODUCTION

5
We can determine the mass density of a cube (with sides a ) by dividing its total mass by its
volume. So, let us consider a new cube (with sides a ) whose volume is less than the first
one. In Figure 3(b) we can observe that, depending on the position of the new cube, we
can obtain different values for mass density, as different position contain different amounts
of matter and voids.
That is, if we can vary the a -dimension from a very small size, we will notice that the mass
density value will oscillate, (see Figure 4). However, there will be a a -dimension region in
which the mass density value maintains constant. The continuum mechanics is raised into
this interval.












Figure 3: Mass density measurement.
It is possible to extend the continuum mechanics to other scales by adding certain
hypothesis, such as the so-called scale effect, but this is not a subject covered in this book.









Figure 4: Mass density.
) (a log
j
Scale of
Continuum
Mechanics
a
a
a
a
cube with less matter
(more empty)
cube with more matter
(less empty)
(a) (b)
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
NOTES ON CONTINUUM MECHANICS

6
4 The Initial Boundary Value Problem (IBVP)
Continuum mechanics, based on certain principles, attempts to formulate the equations
that govern given physical problems by means of partial differential equations. To these we
must add the boundary and initial conditions in order to guarantee the uniqueness of the
problem. This set of partial differential equations and the boundary initial conditions make
up the Initial Boundary Value Problem (IBVP), (see Figure 5).
With a static or quasi-static problem the IBVP becomes a Boundary Value Problem (BVP)
where the initial conditions are redundant.

















Figure 5: Statement and solution of the problem.
4.1 Solving the IBVP
Once the physical problem is stated, it can be solved and the IBVP solution can be
analytical (exact solution), or numerical (approximated solution), (see Figure 5).
In practice, obtaining the analytical solution of the IBVP is very difficult or even
impossible because of the problem complexity (e.g. due to its geometry, forces, or
boundaries), hence we must resort to using IBVP numerical solution. However, obtaining
the analytical solution for simple problems is very important since it serves as a reference
to indicate the degree of accuracy (precision) of the numerical technique used.
Among the most widely used numerical techniques for the IBVP solution we can list,
among others:
The Finite Differences Method - FDM;
The Finite Element Method - FEM;
Set of Partial Differential Equations
Boundary conditions
Initial conditions
Physical problem
IBVP
SOLUTION
Analytical (exact)
Numerical
C

u
,

j
INTRODUCTION

7
The Boundary Element Method - BEM;
The Finite Volume Method - FVM;
The Meshless Method.
We cannot state that any one of the techniques mentioned above is the best. First we must
ask what type of problem we want to solve and, depending on this, one technique or
another, or even a combination of different techniques can be used to optimize the
solution.
In general, all techniques transform the continuous problem into a discrete system of
equations.
The finite difference method (FDM) is based on discretizing the domain by points in
which the governing equations are valid. The FDM was the first numerical method to
emerge and today it is still in use on problems where stabilization problems occur and is
also used to dicretize the time domain.
The finite element method (FEM) is based on discretizing the domain into subdomains
called finite elements, in which the governing equations are valid. Moreover, it has proved
to be more accurate in solving problems that FDM. Today the FEM technique is the most
used and widespread in the solid mechanics field.
Conversely, in the Boundary Element Method (BEM), only the domain boundary is
discretized by elements. From a viewpoint of the solution accuracy, the BEM provides
more accurate solutions than FEM for elastic problems and it is a better method for
working with semi-infinite or infinite problem domains. Nevertheless, the BEM has its
downside in nonlinear problems, where we need to dicretize the domain by cells.
Generally, the IBVP contains both spatial variables (displacement, pressure, etc.) and
temporary variables (rates of change of spatial variables), so we need to have both spatial
and time dicretization to obtain the numerical solutions. For example, for spatial
discretization we can use the FEM and for time discretization we can use another
technique, such as the FDM.
4.2 Simplifying the IBVP
There are cases where the problem (IBVP) includes certain features which allow it to be
simplified whereby its complexity can be drastically reduced and with which even the
analytical problem solution can be obtained. These simplifications will be pointed out and
elaborated on in detail throughout this book, but the engineers will have to decide for
themselves when these simplifications can be used for a given problem and for this a sound
grounding in the general theory is needed.









1 Tensors











1.1 Introduction
As seen previously in the introductory chapter, the goal of continuum mechanics is to
establish a set of equations that governs a physical problem from a macroscopic
perspective. The physical variables featuring in a problem are represented by tensor fields,
in other words, physical phenomena can be shown mathematically by means of tensors
whereas tensor fields indicate how tensor values vary in space and time. In these equations
one main condition for these physical quantities is they must be independent of the
reference system, i.e. they must be the same for different observers. However, for matters
of convenience, when solving problems, we need to express the tensor in a given
coordinate system, hence we have the concept of tensor components, but while tensors are
independent of the coordinate system, their components are not and change as the system
change.
In this chapter we will learn the language of TENSORS to help us interpret physical phenomena.
These tensors can be classified according to the following order:
Zeroth-Order Tensors (Scalars): Among some of the quantities that have magnitude but
not direction are e.g.: mass density, temperature, and pressure.
First-Order Tensors (Vectors): Quantities that have both magnitude and direction, e.g.:
velocity, force. The first-order tensor is symbolized with a boldface letter and by an arrow
at the top part of the vector, i.e.: -
,
.
Second-Order Tensors: Quantities that have magnitude and two directions, e.g. stress and
strain. The second-order and higher-order tensors are symbolized with a boldface letter.
In the first part of this chapter we will study several tools to manage tensors (scalars,
vectors, second-order tensors, and higher-order tensors) without heeding their dependence

Tensors
1
9 , Notes on Continuum Mechanics, Lecture Notes on Numerical
Methods in Engineering and Sciences 4, DOI 10.1007/978-94-007-5986-2_2,
International Center for Numerical Methods in Engineering (CIMNE), 2013
E.W.V. Chaves
NOTES ON CONTINUUM MECHANICS

10
on space and time. At the end of the chapter we will introduce tensor fields and some field
operators which can be used to interpret these fields.
In this textbook we will work indiscriminately with the following notations: tensorial,
indicial, and matricial. Additionally, when the tensors are symmetrical, it is also possible to
represent their components using the Voigt notation.
1.2 Algebraic Operations with Vectors
There now follows a brief review of vectors, in the Euclidean vector space ( ) E , so that we
may become acquainted with the nomenclature used in this textbook.
Addition: Let a
,
, b
,
be arbitrary vectors, we can show the sum of adding them, (see Figure
1.1 (a)), with a new vector ( c
,
) thus defined as:
a b b a c
,
, ,
, ,
- -
(1.1)






Figure 1.1: Addition and subtraction of vectors.
Subtraction: The subtraction between two arbitrary vectors ( a
,
, b
,
), (see Figure 1.1 (b)), is given
as follows:
b a d
,
,
,
-
(1.2)
Considering three vectors a
,
, b
,
and c
,
the following properties are satisfied:
c b a c b a c b a
,
,
, ,
,
, ,
,
,
- - - - - - ) ( ) (
(1.3)
Scalar multiplication: Let a
,
be a vector, we can define the scalar multiplication with a
,
/ .
The product of this operation is another vector with the same direction of a
,
, and whose
length and orientation is defined with the scalar / as shown in Figure 1.2.






Figure 1.2: Scalar multiplication.
c
,

a
,

b
,

c
,

a
,

b
,

b
,
-
d
,

a) b)
a
,

1 > / 0 < / 1 /
a
,
/
a
,
/
a
,
/
1 0 < / <
a
,

a
,
a
,

1 TENSORS

11
Scalar Product: The Scalar Product (also known as the dot product or inner product) of two
vectors a
,
, b
,
, denoted by b a
,
,
, is defined as follows:
0 cos b a b a
,
,
,
,
(1.4)
where 0 is the angle between the two vectors, (see Figure 1.3(a)), and - represents the
Euclidean norm (or magnitude) of -. The result of the operation (1.4) is a scalar.
Moreover, we can conclude that a b b a
,
, ,
,
. The expression (1.4) is also true when b a
,
,
,
therefore:
a a a a a a a a a a a
, , , , , , , , , , ,
0
0
2
0
cos (1.5)
Hence, the norm of a vector is a a a
, , ,
.
Unit Vector: A unit vector, associated with the a
,
-direction, is shown with a a

, which has
the same direction and orientation of a
,
. In this textbook, the hat symbol (
-
) denotes a
unit vector. Thus, the unit vector, a

, codirectional with a
,
, is defined as:
a
a
a
,
,


(1.6)
where a
,
represents the norm (magnitude) of a
,
. If a

is the unit vector, then the


following must be true:
1

a
(1.7)
Zero Vector (or Null Vector): The zero vector is represented by a:
,
0 (1.8)
Projection Vector: The projection vector of a
,
onto b
,
, (see Figure 1.3(b)), is defined as:
b a a proj
b
b

, ,
,
,
proj
Projection vector of a
,
onto b
,

(1.9)
where a proj
b
,
,
is the projection of a
,
onto b
,
, and b

is the unit vector associated with the


b
,
-direction. The magnitude of a proj
b
,
,
is obtained by means of the scalar product:
b a a proj
b


, ,
,
Projection of a
,
onto b
,

(1.10)
So, taking into account the definition of the unit vector, we obtain:
b
b a
a proj
b
,
,
,
,
,


(1.11)
Then, the projection vector, a proj
b
,
,
, can be calculated by:

b
b
b a
b
b
b
b a
b
b
b a
a proj
b
,
,
,
,
,
,
,
,
,
,
,
,
,
,

2
scalar


(1.12)

NOTES ON CONTINUUM MECHANICS

12








Figure 1.3: Scalar product and projection vector.
Orthogonality between vectors: Two vectors a
,
and b
,
are orthogonal if the scalar product
between them is zero, i.e.:
0 b a
,
,

(1.13)
Vector Product (or Cross Product): The vector product of two vectors, a
,
, b
,
, results in
another vector c
,
, which is perpendicular to the plane defined by the two input vectors,
(see Figure 1.4). The vector product has the following characteristics:
Representation:
a b b a c
,
, ,
, ,
r - r
(1.14)
The vector c
,
is orthogonal to the vectors a
,
and b
,
, thus:
0 c b c a
,
,
, ,

(1.15)
The magnitude of c
,
is defined by the formula:
0 sin b a c
,
, ,
(1.16)
where 0 measures the smallest angle between a
,
and b
,
, (see Figure 1.4).
The magnitude of the vector product b a
,
,
r is geometrically expressed as the area of the
parallelogram defined by the two vectors, (see Figure 1.4):
b a
,
,
r A (1.17)
Therefore, the triangle area defined by the points OCD, (see Figure 1.4 (a)), is:
b a
,
,
r
2
1
T
A (1.18)
If a
,
and b
,
are linearly dependent, i.e. b a
,
,
c with c denoting a scalar, the vector product
of two linearly dependent vectors becomes a zero vector, 0 b b b a
, , , ,
,
r r c .
Scalar Triple Product (or Mixed Product): Let a
,
, b
,
, c
,
be arbitrary vectors, we can
define the scalar triple product as:
( ) ( ) ( )
( ) ( ) ( ) a b c c a b b c a
b a c a c b c b a
,
,
, , ,
, ,
, ,
,
, , , ,
,
,
,
,
r - r - r -
r r r


V
V
(1.19)
a) Scalar product b) Projection vector
0
a
,

b
,

0
a
,

b
,

.
a
b
,
,
proj
r s 0 s 0
0 cos b a b a
,
,
,
,

b


1 TENSORS

13
where the scalar V represents the volume of the parallelepiped defined by c b a
,
,
,
, , , (see
Figure 1.5).








Figure 1.4: Vector product.
If two vectors are linearly dependent then, the scalar triple product is zero, i.e.:
( ) 0 a b a
,
,
,
,
r
(1.20)
Let a
,
, b
,
, c
,
, d
,
, be vectors and c, , be scalars, the following property is satisfied:
) ( ) ( ) ( ) ( d c b d c a d c b a
,
,
, ,
, ,
,
,
,
,
r - r r - , c , c
(1.21)
NOTE: Some authors represent the scalar triple product as, ( ) c b a c b a
,
,
, ,
,
,
r = ] , , [ ,
( ) a c b a c b
, ,
,
, ,
,
r = ] , , [ , ( ) b a c b a c
,
, ,
,
, ,
r = ] , , [ .








Figure 1.5: Scalar triple product.
Vector Triple Product: Let a
,
, b
,
, c
,
be vectors, we can define the vector triple product as
( ) c b a w
,
,
, ,
r r . Then, we can demonstrate that the following relationships to be true:
( ) ( ) ( )
( ) ( )c b a b c a
a b c b a c c b a w
,
,
,
,
, ,
,
,
,
,
, , ,
,
, ,
-
r r r r - r r

(1.22)
whereby it is clear that the result of the vector triple product is another vector w
,
, belonging to
the plane
1
H formed by the vectors b
,
and c
,
, (see Figure 1.6).
a
,

b
,

0
a
,
r b
,

.
.
V
c
,

= V Scalar triple product
b a c
,
, ,
r
a
,

b
,

0
a b c
,
,
,
r -
.
.
a
,

b
,

0
.
.
A

T
A
O
C
D
a) b)
c
,

O
C
D
NOTES ON CONTINUUM MECHANICS

14













Figure 1.6: Vector triple product.
Problem 1.1: Let a
,
and b
,
be arbitrary vectors. Prove that the following relationship is
true:
( ) ( ) ( )( ) ( )
2
b a b b a a b a b a
,
,
, ,
, ,
,
,
,
,
- r r
Solution:
( ) ( )
( )
( )
( )
( )
( )( ) ( )
2
2
2
2
2 2
2
2
2
2
2
2
2
2
2
2
2
2
2
2

cos
cos
cos 1
sin
sin
b a b b a a
b a b a
b a b a
b a b a
b a
b a
b a
b a b a b a
,
,
, ,
, ,
,
,
,
,
,
,
,
,
,
,
,
,
,
,
,
,
,
,
,
,
,
,
,
,

-
-
0 -
0 -
0 -
0
0
r r r

Linear Transformation
Let u
,
and v
,
be arbitrary vectors, and c be a scalar, we can state F is a linear
transformation if the following is true:
) ( ) ( ) ( v u v u
, , , ,
F F F - -
) ( ) ( u u
, ,
F F c c

1
H

2
H
a
,

b
,

c
,

c b
,
,
r

1
H - plane defined by b
,
, c
,


2
H - plane defined by a
,
, c b
,
,
r
w
,

w
,
belonging to the plane
1
H
1 TENSORS

15
Problem 1.2: Given the following functions r r o E ) ( and
2
2
1
) ( r r E , demonstrate
whether these functions show a linear transformation or not.
Solution:
[ ] ) ( ) ( ) (
2 1 2 1 2 1 2 1
r o - r o r - r r - r r - r o E E E (linear transformation)














The function
2
2
1
) ( r r E does not show a linear transformation because the condition
) ( ) ( ) (
2 1 2 1
r - r r - r has not been satisfied:
[ ] [ ]
) ( ) ( ) ( ) (
2
2
1
2
1
2
1
2
2
1
2
1
) (
2 1 2 1 2 1
2 1
2
2
2
1
2
2 2 1
2
1
2
2 1 2 1
r - r r r - r - r
r r - r - r r - r r - r r - r r - r

E
E E E E E


















) (r o
r
2 1
r - r
2
r
1
r
) (
2
r o
) (
1
r o
) ( ) ( ) (
2 1 2 1
r o - r o r - r o
r
2 1
r - r 1
r
2
r
) (r
) (
2 1
r - r
) (
2
r
) (
1
r
) ( ) (
2 1
r - r
NOTES ON CONTINUUM MECHANICS

16
1.3 Coordinate Systems
A tensor, which has physical meanings, must be independent of the adopted coordinate
system. Sometimes for reasons of convenience, we need to represent a tensor in a specific
coordinate system, hence, we have the concept of tensor components, (see Figure 1.7).










Figure 1.7: Tensor components.
Let a
,
be a first-order tensor (vector) as shown in Figure 1.8 (a), the tensor representation
in a general coordinate system, defined as
3 2 1
, , , is made up of its components
(
3 2 1
a a a , , ), (see Figure 1.8 (b)). Some examples of coordinate system are: the Cartesian
coordinate system; the cylindrical coordinate system; and the spherical coordinate system.






Figure 1.8: Vector representation in a general coordinate system.
1.3.1 Cartesian Coordinate System
The Cartesian coordinate system is defined by three unit vectors: i

, j

, k

, denoted by the
Cartesian basis, which make up an orthonormal basis. The orthonormal basis has the
following properties:
1. The vectors that make up this basis are unit vectors:
1

k j i
(1.23)
or:
1

k k j j i i (1.24)
a
,

a)
a
,


1


2

3

b)
a
,
(
3 2 1
a a a , , )
TENSORS
Mathematical representation of the physical
quantities
(Independent of the coordinate system)
COMPONENTS
Tensor Representation in a
Coordinate System
1 TENSORS

17
2. The unit vectors ( i

, j

, k

) are mutually orthogonal, i.e.:


0

i k k j j i (1.25)
3. The vector product between the vectors ( i

, j

, k

) is the following:
j i k i k j k j i

;

;

r r r (1.26)
The direction and orientation of the orthonormal basis can be obtained using the right-
hand rule as shown in Figure 1.9.
k j i

r i k j

r j i k

r (1.27)



Figure 1.9: The right-hand rule.
1.3.2 Vector Representation in the Cartesian Coordinate
System
The vector a
,
, (see Figure 1.10), in the Cartesian coordinate system, is represented by its
different components (
x
a ,
y
a ,
z
a ) and by the Cartesian bases ( i

, j

, k

) as:
k j i

z y x
a a a - - a
,
(1.28)











Figure 1.10: Cartesian coordinate system.
Let a
,
, b
,
, c
,
be arbitrary vectors, we can describe some vector operations in the Cartesian
coordinate system, as follows:
i


x
y
z
a
,

i



x
a

y
a

z
a
NOTES ON CONTINUUM MECHANICS

18
The scalar product b a
,
,
becomes a scalar, which is defined in the Cartesian
system as:
) ( )

( )

(
z z y y x x z y x z y x
b a b a b a b b b a a a - - - - - - k j i k j i b a
,
,
(1.29)
Thus, it is true that
2
2 2 2
a a a
, , ,
- - - -
z y x z z y y x x
a a a a a a a a a .
NOTE: The projection of a vector onto a given direction was established in the
equation (1.10), thus defining the component concept. For example, if we want to
know the vector component along the y -direction, all we need to do is calculate:
y z y x
a a a a - - )

( )

(

j k j i j a
,
.
The norm of a
,
is:
2 2 2
z y x
a a a - - a
,
(1.30)
Then, the unit vector codirectional with a
,
is:
k j i

2 2 2 2 2 2 2 2 2
z y x
z
z y x
y
z y x
x
a a a
a
a a a
a
a a a
a
- -
-
- -
-
- -

a
a
a
,
,

(1.31)
The zero vector is:

0 0 0 - - i j k
,
0 (1.32)
Addition: The vector sum of a
,
and b
,
is represented by:
( ) ( ) ( )k j i k j i k j i

)

( )

(
z z y y x x z y x z y x
b a b a b a b b b a a a - - - - - - - - - - - b a
,
,
(1.33)
Subtraction: The difference between a
,
and b
,
is:
( ) ( ) ( )k j i k j i k j i

)

( )

(
z z y y x x z y x z y x
b a b a b a b b b a a a - - - - - - - - - - - b a
,
,
(1.34)
Scalar multiplication: The resulting vector defined by a
,
/ is:
k j i

z y x
a a a / - / - / /a
,
(1.35)
The vector product ( b a
,
,
r ) is evaluated as:
k j i
k j i
k j i

) (

) (

) (


x y y x x z z x y z z y
y x
y x
z x
z x
z y
z y
z y x
z y x
b a b a b a b a b a b a
b b
a a
b b
a a
b b
a a
b b b
a a a
- - - - -
- - r b a c
,
, ,

(1.36)
where the symbol ) (- = - det denotes the matrix determinant.
The scalar triple product ] c b a [
,
,
,
, , is the determinant of the 3 by 3 matrix, defined
as:

1 TENSORS

19
( ) ( ) ( )
( ) ( ) ( )
x y y x z x z z x y y z z y x
y x
y x
z
z x
z x
y
z y
z y
x
z y x
z y x
z y x
c b c b a c b c b a c b c b a
c c
b b
a
c c
b b
a
c c
b b
a
c c c
b b b
a a a
V
- - - - -
- -
r r r b a c a c b c b a c b a
,
, , , ,
,
,
,
, ,
,
,
) , , (

(1.37)
The vector triple product made up of the vectors ( c b a
,
,
,
, , ) is obtained, in the
Cartesian coordinate system, as:
( ) ( ) ( )
( ) ( ) ( )k j i


2 1 2 1 2 1 z z y y x x
c b c b c b / - / - / - / - / - /
- r r c b a b c a c b a
,
,
,
,
, , ,
,
,

(1.38)
where
z z y y x x
c a c a c a - - / c a
, ,
1
, and
z z y y x x
b a b a b a - - / b a
,
,
2
.
Problem 1.3: Consider the points: ( ) 1 , 3 , 1 A , ( ) 1 , 1 , 2 - B , ( ) 3 , 1 , 0 C and ( ) 4 , 2 , 1 D , defined in the
Cartesian coordinate system.
1) Find the parallelogram area defined by

AB and

AC ; 2) Find the volume of the


parallelepiped defined by

AB,

AC and

AD ; 3) Find the projection vector of

AB onto

BC .
Solution:
1) Firstly we calculate the vectors

AB and

AC :
( ) ( ) k j i k j i k j i

2 - - - - - - - -

OA OB AB a
,

( ) ( ) k j i k j i k j i

0 - - - - - - - - -

OA OC AC b
,

With reference to the equation (1.36) we can evaluate the vector product as follows:
k j i
k j i

) 6 (

) 8 (
2 2 1
0 4 1

- - - -
- -
- r b a
,
,

Then, the parallelogram area can be obtained using definition (1.19), thus:
104 ) 6 ( ) 2 ( ) 8 (
2 2 2
- - - - - r b a
,
,
A
2) Next, we can evaluate the vector

AD as:
( ) ( ) k j i k j i k j i

1 - - - - - - - -

OA OD AD c
,

and using the equation (1.37) we can obtain the volume of the parallelepiped:
( ) ( ) ( )
16 18 2 0

0 ) , , (
- -
- - - - - r k j i k j i b a c c b a
,
, , ,
,
,
V

3) The

BC vector can be calculated as:


( ) ( ) k j i k j i k j i

0 - - - - - - - - -

OB OC BC
NOTES ON CONTINUUM MECHANICS

20
Hence, it is possible to evaluate the projection vector of

AB onto

BC , (see equation
(1.12)), as:
( ) ( )
( ) ( )
( )
( )
( )
( ) k j i k j i
k j i
k j i k j i
k j i k j i

3
5

3
5

3
5

2
4 4 4
0 8 2

2
2
- - - - -
- -
- - -

- - -
- - - - - -
- - - - -

BC
BC BC
AB BC
AB
BC
BC
_
proj

1.3.3 Einstein Summation Convention (Einstein Notation)
As we saw in equation (1.28) a
,
in the Cartesian coordinate system was defined as:
k j i

z y x
a a a - - a
,
(1.39)
Said expression can be rewritten as:
3 3 2 2 1 1

e e e a a a a - -
,
(1.40)
where we have considered that:
x
a a =
1
,
y
a a =
2
,
z
a a =
3
, i

1
= e , j

2
= e , k

3
= e , (see
Figure 1.11). In this way we can express equation (1.40) by means of the summation
symbol as:

- -
3
1
3 3 2 2 1 1

i
i i
e e e e a a a a a
,

(1.41)
Then, we introduce the summation convention, according to which the repeated indices
indicate summation. So, equation (1.41) can be represented as follows:
) 3 , 2 , 1 (

3 3 2 2 1 1
- - i
i i
e e e e a a a a a
,

) 3 , 2 , 1 (

i
i i
e a a
,

(1.42)
NOTE: The summation notation was introduced by Albert Einstein in 1916, which led to
the indicial notation.
1.4 Indicial Notation
Using indicial notation, the three axes of the coordinate system are designated by the letter
x with a subscript. So,
i
x is not a single value but i values, i.e.
1
x ,
2
x ,
3
x (if 3 , 2 , 1 i )
where these values
1
x ,
2
x ,
3
x correspond to the axes x , y , z , respectively.
Let a
,
be a vector represented in the Cartesian coordinate system as:
3 3 2 2 1 1

e e e a a a a - -
,
(1.43)
1 TENSORS

21
where the orthonormal basis is represented by
1

e ,
2

e ,
3

e , (see Figure 1.11), and


1
a ,
2
a ,
3
a are the vector components. In indicial notation the vector components are represented
by
i
a . If the range of the subscript is not indicated, we assume that 1,2,3 show these
values. Therefore, the vector components are represented as:
(
(
(


3
2
1
) (
a
a
a
a
i i
a
,

(1.44)












Figure 1.11: Vector representation in the Cartesian coordinate system.
Unit vector components: Let a
,
be a vector, the normalized vector a

is defined as:
1

with

a
a
a
a
,
,

(1.45)
whose components are:
) 3 , 2 , 1 , , (

2
3
2
2
2
1

- -
k j i
k k
i
j j
i i
i
a a
a
a a
a
a a a
a
a
(1.46)
In light of the previous equation we can emphasize two types of indices:
The free index (live index) is that which only appears once in a term of the expression. In the
above equation the free index is the ( i ). The number of the free index indicates
the tensor order.
The dummy index (summation index) is that which is repeated only twice in a term of the
expression, and indicates summation. In the above equation (1.46) the
dummy index is the ( j ), or the ( k ) index.
OBS.: An index in a term of an expression can only appear once or twice. If it
appears more times, then a large error has occurred.

2
x y =

1
x x =

3
x z =
a
,


1

e = i
2

e = j

3

e = k

1
a a =
x


2
a a =
y


3
a a =
z

NOTES ON CONTINUUM MECHANICS

22
Scalar product: Using definitions (1.4) and (1.29), we can express the scalar product
( b a
,
,
) as follows:
) 3 , 2 , 1 , ( cos
3 3 2 2 1 1
- - 0 j i
j j i i
b a b a b a b a b a b a b a
,
,
,
,
(1.47)


Problem 1.4: Rewrite the following equations using indicial notation:

1)
3 3 3 3 2 2 3 1 1
x x a x x a x x a - -
Solution: ) 3 , 2 , 1 (
3
i x x a
i i

2)
2 2 1 1
x x x x -
Solution: ) 2 , 1 ( i x x
i i

3)

- -
- -
- -
z
y
x
b z a y a x a
b z a y a x a
b z a y a x a
33 32 31
23 22 21
13 12 11

Solution:

- -
- -
- -
3 3 33 2 32 1 31
2 3 23 2 22 1 21
1 3 13 2 12 1 11
b x a x a x a
b x a x a x a
b x a x a x a

j index
dummy

3 3
2 2
1 1
b x a
b x a
b x a
j j
j j
j j

i index
free


i j ij
b x a
As we can appreciate in this problem, the use of the indicial notation means that the
equation becomes very concise. In many cases, if algebraic operation do not use indicial or
tensorial notation they become almost impossible to deal with due to the large number of
terms involved.
Problem 1.5: Expand the equation: ) 3 , 2 , 1 , ( j i x x A
j i ij

Solution: The indices j i, are dummy indices, and indicate index summation and there is no
free index in the expression
j i ij
x x A , therefore the result is a scalar. So, we expand first the
dummy index i and later the index j to obtain:
_ _ _
3 3 33
2 3 32
1 3 31
3 3
3 2 23
2 2 22
1 2 21
2 2
3 1 13
2 1 12
1 1 11
1 1
x x A
x x A
x x A
x x A
x x A
x x A
x x A
x x A
x x A
x x A
x x A
x x A x x A
j j j j j j
i expanding
j i ij
-
-
-
-
-
-
-
-


Rearranging the terms we obtain:
3 3 33 2 3 32 1 3 31 3 2 23
2 2 22 1 2 21 3 1 13 2 1 12 1 1 11
x x A x x A x x A x x A
x x A x x A x x A x x A x x A x x A
j i ij
- - -
- - - - -

1.4.1 Some Operators
1.4.1.1 Kronecker Delta
The Kronecker delta
ij
c is defined as follows:
e
x
p
a
n
d
i
n
g


j

1 TENSORS

23

j i iff
j i iff
ij
0
1
c
(1.48)
Also note that the scalar product of the orthonormal basis
j i
e e

is equal to 1 if j i and
equal to 0 if j i . Hence,
j i
e e

can be expressed in matrix form as:
ij j i
c
(
(
(

(
(
(

1 0 0
0 1 0
0 0 1




3 3 2 3 1 3
3 2 2 2 1 2
3 1 2 1 1 1
e e e e e e
e e e e e e
e e e e e e
e e (1.49)
An interesting property of the Kronecker delta is shown in the following example. Let
i
V
be the components of the vector V
,
, therefore:
3 3 2 2 1 1
V V V V
j j j i ij
c c c c - -
(1.50)
As ) 3 , 2 , 1 ( j is a free index, we have three values to be calculated, namely:
j i ij
i ij
i ij
i ij
V V
V V V V V j
V V V V V j
V V V V V j

- -
- -
- -
c
c c c c
c c c c
c c c c
3 3 33 2 23 1 13
2 3 32 2 22 1 12
1 3 31 2 21 1 11
3
2
1

(1.51)
That is, in the presence of the Kronecker delta symbol we replace the repeated index as
follows:
j i j i
V V

c
(1.52)
For this reason, the Kronecker delta is often called the substitution operator.
Other examples using the Kronecker delta are presented below:
jk ik ij
A A c , 3
33 22 11
- - c c c c c c c
jj ii ji ij
,
33 22 11
a a a a a
ii ji ji
- - c
(1.53)
To obtain the components of the vector a
,
in the coordinate system represented by
i
e

, it
is sufficient to obtain the scalar product with a
,
and
i
e

, i.e.
i pi p i p p i
a a a c e e e a

,
.
With that, it is also possible to represent the vector as:
i i i i
e e a e a


, ,
a (1.54)
Problem 1.6: Solve the following equations:
1)
jj ii
c c
Solution: ( )( ) 9 3 3
33 22 11 33 22 11
- - - - c c c c c c c c
jj ii

2)
1 1 o o
c c c
Solution: 1
11 1 1 1 1

o o
c c c c c c
NOTE: Note that the following algebraic operation is incorrect 1 3
11 1 1


c c c c ,
since what must be replaced is the repeated index, not the number
1.4.1.2 Permutation Symbol
The permutation symbol
ijk
(also known as Levi-Civita symbol or alternating symbol) is defined as:
NOTES ON CONTINUUM MECHANICS

24
{
{


= -
= -

) ( ) ( ) (
) 3 , 1 , 2 ( ), 1 , 2 , 3 ( ), 2 , 3 , 1 ( ) , , ( 1
) 2 , 1 , 3 ( ), 1 , 3 , 2 ( ), 3 , 2 , 1 ( ) , , ( 1
0 k i or k j or j i if
k j i if
k j i if
: i.e. cases remaining the for
ijk

(1.55)
NOTE:
ijk
are the components of the Levi-Civita pseudo-tensor, which will be introduced
later on.
The values of
ijk
can be easily memorized using the mnemonic device shown in Figure
1.12(a), in which if the index values are arranged in a clockwise direction, the value of
ijk

is equal to 1, if not it has the value of 1 - . In the same way we can use this mnemonic
device to switch indices, (see Figure 1.12(b)).








Figure 1.12: Mnemonic device for the permutation symbol.
Another way to express the permutation symbol is by means of its indices:
) )( )( (
2
1
i k k j j i
ijk
- - - (1.56)
Using both the definition seen in (1.55) and Figure 1.12 (b), it is possible to verify that the
following relations are valid:
kji jik ikj ijk
kij jki ijk


- - -

(1.57)
Using the Kronecker delta property, we can state that:
( ) ( ) ( )
j i j i k k i k i j k j k j i
k j i k j i k j i k j i k j i k j i
nk mj li lmn ijk
2 3 3 2 1 2 3 3 2 1 2 3 3 2 1
1 2 3 1 3 2 2 1 3 3 1 2 2 3 1 3 2 1
c c c c c c c c c c c c c c c
c c c c c c c c c c c c c c c c c c
c c c
- - - - -
- - - - -


(1.58)
The above equation can be represented by means of the following determinant:
k k k
j j j
i i i
k j i
k j i
k j i
ijk
3 2 1
3 2 1
3 2 1
3 3 3
2 2 2
1 1 1
c c c
c c c
c c c
c c c
c c c
c c c
(1.59)
After which, the term
pqr ijk
can be evaluated as follows:
r q p
r q p
r q p
k k k
j j j
i i i
pqr ijk
3 3 3
2 2 2
1 1 1
3 2 1
3 2 1
3 2 1
c c c
c c c
c c c
c c c
c c c
c c c
(1.60)
1
2
3
1
ijk


1 -
ijk


a)
i
j
k
b)

kij jki ijk


ikj ijk
-


jik
kji
ikj ijk


-
-
-

1 TENSORS

25
Taking into account that ) ( ) ( ) ( B A AB det det det , where - = -) ( det is the determinant
of the matrix -, the equation (1.60) can be rewritten as:

(
(
(

(
(
(


3 3 3
2 2 2
1 1 1
3 2 1
3 2 1
3 2 1
r q p
r q p
r q p
k k k
j j j
i i i
pqr ijk
c c c
c c c
c c c
c c c
c c c
c c c

kr kq kp
jr jq jp
ir iq ip
pqr ijk
c c c
c c c
c c c
(1.61)
The term
ip
c was obtained by means of the operation
mp mi p i p i p i
c c c c c c c c - -
3 3 2 2 1 1

and
ip mp mi
c c c , the other terms were obtained in a similar fashion.
For the special exception when k r , the equation (1.61) is reduced to:

3

kq kp
jk jq jp
ik iq ip
pqr ijk
c c
c c c
c c c

3 , 2 , 1 , , , , - q p k j i
jp iq jq ip pqk ijk
c c c c
(1.62)
Problem 1.7: a) Prove the following is true
ip pjk ijk
c 2 and 6
ijk ijk
. b) Obtain the
numerical value of
i k j ijk 1 3 2
c c c .
Solution: a) Using the equation in (1.62), i.e.
jp iq jq ip pqk ijk
c c c c - , and by substituting q
for j , we obtain:
ip ip ip jp ij jj ip pjk ijk
c c c c c c c 2 3 - -
Based on the above result, it is straight forward to check that:
6 2
ii ijk ijk
c
b) 1
123 1 3 2

i k j ijk
c c c
The vector product of two vectors ( b a
,
,
r ) leads to a new vector c
,
, defined in
(1.36), and the components of c
,
, in Cartesian system, are given by:
3
3
1 2 2 1 2
2
3 1 1 3 1
1
2 3 3 2
3 2 1
3 2 1
3 2 1

) (

) (

) (

e e e
e e e
b a c
_ _ _
,
, ,
c c c
b a b a b a b a b a b a
b b b
a a a
- - - - -
r
(1.63)
Using the definition of the permutation symbol
ijk
, defined in (1.55), we can express the
components of c
,
as follows:
k j ijk i
k j jk
k j jk
k j jk
b a c
b a b a b a c
b a b a b a c
b a b a b a c

-
-
-
3 1 2 321 2 1 312 3
2 3 1 213 1 3 231 2
1 2 3 132 3 2 123 1

(1.64)
Then, the vector product ( b a
,
,
r ) can be represented by means of the permutation symbol
as:
i jki k j i ijk k j k j k j
i ijk k j k k j j
i k j ijk
e e e e
e e e
e b a

)

(

b a b a b a
b a b a
b a
r
r
r
,
,
(1.65)
NOTES ON CONTINUUM MECHANICS

26
Therefore, we can also conclude that the following relationship is valid:
i ijk k j
e e e

)

( r
(1.66)
The permutation symbol and the orthonormal basis can be interrelated using the triple
scalar product as follows:
( ) ( )
ijk k j i mk ijm k m ijm k j i m ijm j i
r r r e e e e e e e e e e e

c
(1.67)
The triple scalar product made up of the vectors c b a
,
,
,
, , is expressed by:
( ) ( ) ( )
k j i ijk k j i k j i k k j j i i
c b a c b a c b a r r r / e e e e e e c b a

,
,
,
(1.68)
( ) ) 3 , 2 , 1 , , ( r / k j i
k j i ijk
c b a c b a
,
,
,

(1.69)
or
( ) ( ) ( )
3 2 1
3 2 1
3 2 1
c c c
b b b
a a a
r r r / b a c a c b c b a
,
, , , ,
,
,
,
,

(1.70)
Starting from the equation (1.69) we can prove the following are true:
( ) ( ) ( ) b a c a c b c b a
,
, , , ,
,
,
,
,
r r r :
( )
( )
( )
( )
( )
( ) ] , , [
] , , [
] , , [
] , , [
] , , [
] , , [
a b c a b c
c a b c a b
b c a b c a
b a c b a c
a c b a c b
c b a c b a
,
,
, ,
,
,
, ,
,
, ,
,
,
, ,
,
, ,
,
, ,
,
, ,
, ,
,
, ,
,
,
,
, ,
,
,
- = r - -
- = r - -
- = r - -
= r
= r
r =

k j i kji
k j i jik
k j i ikj
k j i kij
k j i jki
k j i ijk
c b a
c b a
c b a
c b a
c b a
c b a

(1.71)
where we take into account the property of the permutation symbol as given in (1.57).
Problem 1.8: Rewrite the expression ( ) ( ) d c b a
,
,
,
,
r r without using the vector product
symbol.
Solution: The vector product ( ) b a
,
,
r can be expressed as
( )
i k j ijk k k j j
e e e b a

b a b a r r
,
,
. Likewise, it is possible to express ( ) d c
,
,
r as
( )
n m l nlm
e d c

d c r
,
,
, thus:
( ) ( )
m l k j ilm ijk in m l k j nlm ijk
n i m l k j nlm ijk n m l nlm i k j ijk
d c b a d c b a
d c b a d c b a



r r
c
e e e e d c b a

)

( )

,
,
,
,

Taking into account that
lmi jki ilm ijk
(see equation (1.57)) and by applying the
equation (1.62), i.e.:
ilm jki kl jm km jl lmi jki
- c c c c , we obtain:
( )
m l l m m l m l m l k j kl jm km jl m l k j ilm ijk
d c b a d c b a d c b a d c b a - - c c c c
Since ( ) c a
, ,

l l
c a and ( ) d b
, ,

m m
d b holds true, we can conclude that:
( ) ( ) ( )( ) ( )( ) c b d a d b c a d c b a
,
, ,
,
, ,
, ,
,
,
,
,
- r r
Therefore, it is also valid when c a
, ,
and d b
, ,
, thus:
1 TENSORS

27
( ) ( ) ( )( ) ( )( ) ( )
2
2
2
2
b a b a a b b a b b a a b a b a b a
,
,
,
, ,
, ,
,
, ,
, ,
,
,
,
,
,
,
- - r r r
which is the same equation obtained in Problem 1.1.
Problem 1.9: Prove that ( ) ( ) [ ] [ ] ) ( ) ( b a c d b a d c d c b a
,
, ,
, ,
,
,
,
,
,
,
,
r - r r r r
Solution: Expressing the correct equality term in indicial notation we obtain:
[ ] [ ] ( ) [ ] ( ) [ ]
k j ijk i p k j ijk i p
p
b a c d b a d c ) ( ) ( -
)
`

r - r b a c d b a d c
,
, ,
, ,
,
,
,

( )
p i i p
k
j ijk p i k j ijk i p k j ijk
d c d c b a d c b a d c b a - -
Using the Kronecker delta the above equation becomes:
( ) ( ) ( )
np im ni pm n m k j ijk np n m im ni n m pm
k
j ijk
c c c c c c c c - - d c b a d c d c b a
and by applying the equation
mnl pil np im ni pm
- c c c c , (see eq. (1.62)), the above equation
can be rewritten as follows:
( ) ( ) ( )( ) [ ]
n m mnl k j ijk pil mnl pil n m k j ijk
d c b a d c b a
Since
k j ijk
b a and
n m mnl
d c represent the components of ( ) b a
,
,
r and ( ) d c
,
,
r ,
respectively, we can conclude that:
( )( ) [ ] ( ) ( ) [ ]
p n m mnl k j ijk pil
d c b a
,
,
,
,
r r r d c b a
Problem 1.10: Let a
,
, b
,
, c
,
be linearly independent vectors, and v
,
be a vector,
demonstrate that:
0 c b a v
,
,
,
, ,
- - , c
where the scalars c , , , are given by:
r q p pqr
k j i ijk
r q p pqr
k j i ijk
r q p pqr
k j i ijk
c b a
v b a
c b a
c v a
c b a
c b v

, c ; ;
Solution: The scalar product made up of v
,
and ( c b
,
,
r ) becomes:
_
,
,
,
_
,
, ,
,
,
, ,
,
,
0 0
) ( ) ( ) ( ) (

r - r - r r c b c c b b c b a c b v , c
) (
) (
c b a
c b v
,
,
,
,
,
,
r
r

c
which is the same as:
r q p pqr
k j i ijk
c b a
c b v
c b a
c b a
c b a
c b v
c b v
c b v
c c c
b b b
a a a
c c c
b b b
v v v


3 3 3
2 2 2
1 1 1
3 3 3
2 2 2
1 1 1
3 2 1
3 2 1
3 2 1
3 2 1
3 2 1
3 2 1

c

One can obtain the parameters , and in a similar fashion.
Problem 1.11: Prove the relationship given in (1.38) is valid, i.e.:
( ) ( ) ( )c b a b c a c b a
,
,
,
,
, , ,
,
,
- r r .
Solution: Taking into account that ( )
k j ijk i i
c b r c b d
,
, ,
) ( and that ( )
k j qjk q
c b r d a
,
,
, we
obtain:
NOTES ON CONTINUUM MECHANICS

28
( ) [ ]
( )
( ) ( ) b a c a
c b a
,
, , ,
,
,
,
- -
- -
r r
q q q j j k q k
k j s sj qk k j s sk qj k j s sj qk sk qj
k j s jki qsi k j s ijk qsi k j ijk s qsi q
c b c b a c b a
c b a c b a c b a
c b a c b a c b a
c c c c c c c c
) (

( ) [ ] ( ) ( ) [ ]
q q
b a c c a b c b a
,
, , , ,
,
,
,
,
- r r
1.5 Algebraic Operations with Tensors
1.5.1 Dyadic
The tensor product, made up of two vectors v
,
and u
,
, becomes a dyad, which is a particular
case of a second-order tensor. The dyad is represented by:
A v u v u =
, , , ,

(1.72)
where the operator denotes the tensor product. Then, we define a dyadic as a linear
combination of dyads. Furthermore, as we will see later, any tensor can be represented by
means of a linear combination of dyads, (see Holzapfel (2000)).
The tensor product has the following properties:
1. ) ( ) ( ) ( x v u x v u x v u
, , , , , , , , ,
=
(1.73)
2. w u v u w v u
, , , , , , ,
- - , c , c ) ( (1.74)
3.
[ ] [ ] ) ( ) (
) ( ) ( ) (
x r w x u v
x r w x u v x r w u v
, , , , , ,
, , , , , , , , , , ,


-
- -
, c
, c , c
(1.75)
where c and , are scalars. By definition, the dyad does not contain the commutative
property, i.e., u v v u
, , , ,
.
The equation (1.72) can also be expressed in the Cartesian system as:
)

(
)

(
)

( )

(
j i ij
j i j i
j j i i
e e
e e
e e v u A



A
v u
v u
, ,
) 3 , 2 , 1 , ( j i (1.76)

_
basis
j i
components
ij
Tensor
e e A

A
) 3 , 2 , 1 , ( j i (1.77)
In this textbook, the components of a second-order tensor can be represented in different
ways, namely:
ij j i ij ij
components
A v u

!
!
) ( ) ( v u A
v u A
, ,
_
, ,
(1.78)
These components are explicitly expressed in matrix form as:
1 TENSORS

29
(
(
(


33 32 31
23 22 21
13 12 11
) (
A A A
A A A
A A A
A A
ij ij
A
(1.79)
It is easy to identify the tensor order by the number of free indices in the tensor
components, i.e.:
Second-order tensor
Third-order tensor
Fourth-order tensor l k j i ijkl
k j i ijk
j i ij
e e e e
e e e T
e e U






I I
T
U

) 3 , 2 , 1 , , , ( l k j i (1.80)

Problem 1.12: Define the order of the tensors represented by their Cartesian components:
i
v ,
ijk
d ,
ijj
F ,
ij
r ,
ijkl
C ,
ij
o . Determine the number of components in tensor C .
Solution: The order of the tensor is given by the number of free indices, so it follows that:
First-order tensor (vector): v
,
, F
,
; Second-order tensor: , ; Third-order tensor: d;
Fourth-order tensor: C
The number of tensor components is given by the maximum index range value, i.e.
3 , 2 , 1 , , , l k j i , to the power of the number of free indices which is equal to 4 in the case of
ijkl
C . Thus, the number of independent components in C is given by:
( ) ( ) ( ) ( ) 81 3 3 3 3 3
4
l k j i
The fourth-order tensor
ijkl
C has 81 components.
Let A and B be second-order tensors, we can then define some algebraic operations
including:
Addition: The sum of two tensors of same order is a new tensor defined as follows:
A B B A C - - (1.81)
The components of C are represented by:
ij ij
) ( ) ( B A C - or
ij ij ij
B A C -
(1.82)
or, in matrix notation as:
B A C - (1.83)
Multiplication of a tensor by a scalar: The multiplication of a second-order tensor
( A ) by a scalar ( / ) is defined by a new tensor D , so that:
ij ij
components in
) ( ) (

A D A D / / (1.84)
or, in matrix form:
OBS.: The tensor order is given by the number of free indices in its components.
OBS.: The number of tensor components is given by
n
a , where the base a is the
maximum value in the index range, and the exponent n is the number of the free
index.
NOTES ON CONTINUUM MECHANICS

30
(
(
(

/ / /
/ / /
/ / /
/
(
(
(

33 32 31
23 22 21
13 12 11
33 32 31
23 22 21
13 12 11
A A A
A A A
A A A
A A A
A A A
A A A
A A
(1.85)
It is also true that:
) ( ) ( v A v A
, ,
/ / (1.86)
for any vector v
,
.
Scalar Product (or Dot Product): The scalar product (also known as single
contraction) between a second-order tensor A and a vector x
,
is another vector
(first-order tensor) y
,
, defined as:

j j
j
j
k jk
j kl l jk
l l k j jk
e
e
e
e e e
x A y

( )

(
y
x A
x A
x A
y

_
, ,
c

(1.87)
The scalar product between two second-order tensors A and B is another second-order
tensor, that verifies: A B B A :


l i il
l i kl ik
l i jk kl ij
l k kl j i ij
e e
e e
e e
e e e e B A C



)

( )

(




C
B A
B A
B A
_
AB
c

l i il
l i kl ik
l i jk kl ij
l k kl j i ij
e e
e e
e e
e e e e A B D



)

( )

(




D
A B
A B
A B
_
BA
c
(1.88)
It also satisfies the following properties:
C A B A C B A - - ) ( ; C B A C B A ) ( ) (
(1.89)
The powers of second-order tensors
The scalar product allows us to define the power of second-order tensors, as seen below:
A A A A A A A A A 1 A
3 2 1 0
; ; ; , and so on, (1.90)
where 1 is the second-order unit tensor (also called the identity tensor).

Double Scalar Product (or Double contraction)
Consider two dyads, d c A
,
,
and v u B
, ,
. The double contraction between them is
defined in different ways, namely: B A
:
and B A .
Double contraction : ) (
( ) ( ) ( )( ) u d v c v u d c
,
,
, , , ,
,
,

(1.91)
In components

kl
c

jk
c
jk
c
1 TENSORS

31



) (
)

( )

(
scalar
ji ij
il jk kl ij
l k kl j i ij

c c


B A
B A
B A e e e e B A

(1.92)
The double contraction ) ( is commutative, i.e. A B B A .
Double contraction (: ):
( ) ( ) ( )( ) v d u c v u d c B A
,
,
, , , ,
,
,
:
:

(1.93)
The double contraction (: ) is commutative, so:
( ) ( ) ( )( ) ( )( ) B A v d u c d v c u d c v u A B
:
:
:

,
,
, ,
,
, , ,
,
, , ,

(1.94)
The breakdown into its components appears like this:


) (
)

( )

(
scalar
ij ij
jl ik kl ij
l k kl j i ij
/


B A
B A
B A
c c
e e e e B A
: :

(1.95)
In general, B A B A
:
, however, they are equal if at least one of them is symmetric, i.e.
B A B A
sym sym
:
or
sym sym
B A B A
:
, so
sym sym sym sym
B A B A
:
.
The double contraction with a third-order tensor ( S ) and a second-order tensor ( B )
becomes:
( ) ( ) ( )( )
( ) ( ) ( )( )a d v c u a d c v u S B
c u d v a v u a d c B S
,
,
, , , ,
,
, , ,
, ,
,
, , , , ,
,
,




:
:
:
:
(1.96)
As we can verify the result is a vector. In symbolic notation, the double contraction ( S B
:
)
is represented by:

i jk ijk i kq jp pq ijk q p pq k j i ijk
e e e e e e e

B S B S B S c c
:

(1.97)

The double contraction of a fourth-order tensor ( C ) with a second-order tensor ( ) is
defined as:

j i ij
j i kl ijkl j i lq kp pq ijkl q p pq l k j i ijkl
e e
e e e e e e e e e e


o
r r r C C C c c
:

(1.98)
where
ij
o are the components of
:
C .

ik
c

jl
c

il
c

jk
c
NOTES ON CONTINUUM MECHANICS

32
Next, we express some properties of the double contraction (: ):
( )
( ) ( ) ( ) B A B A B A
C A B A C B A
A B B A
/ / /
- -

: : :
: : :
: :
)
)
)
c
b
a

(1.99)
where C B A , , are second-order tensors, and / is a scalar.
Via the definition of the double scalar product, it is possible to obtain the components of
the second-order tensor A in the Cartesian system, i.e.:
ij lj ki kl j l k kl i j i l k kl ij
A A A A c c e e e e e e e e A

)

(

)

( )

( ) (
:

(1.100)
If we consider any two vectors a
,
, b
,
, and an arbitrary second-order tensor, A , we can
demonstrate that:
) (
) (

b a A
e e e e b A a
,
,
,
,


:
j i ij j ij i jr pi r ij p r r j i ij p p
b a A b A a b A a b A a c c
(1.101)
Vector product
The vector product between a second-order tensor A and a vector x
,
is a second-order
tensor given by:
l i k ij ljk k k j i ij
e e e e e x A

)

( )

( r r x A x A
,

(1.102)
where we have used the definition (1.67), i.e.
l ljk k j
e e e

r . In Problem 1.11, we have
shown that the relation ( ) ( ) ( )c b a b c a c b a
,
,
,
,
, , ,
,
,
- r r holds, which is also represented by
means of dyads as:
( ) [ ] ( ) [ ]
j k k j k j j k k j k k j
a b c c b c b a
,
,
, ,
,
,
,
,
- - - r r a b c c b c b a b c a ) ( ) ( ) (
(1.103)
In the particular case when c a
, ,
we obtain:
( ) [ ]
[ ] [ ]
[ ] {
j
p j p jp k k p j kp k jp k k
j kp p k jp p k k j k k j k k j
b a a 1 a a
a b a
,
, , , ,
,
,
,
-
- -
- - r r
) (
) ( ) ( ) (
) ( ) ( ) ( ) (
b a a a a b a a a a
a b a b a a a b a b a a
c c c
c c
(1.104)
Thus, the following relationships are valid:
( ) ( ) ( )
[ ] b a a 1 a a a b a
a b c c b c b a b c a c b a
,
, , , , ,
,
,
,
,
, ,
,
,
,
,
,
, , ,
,
,


- r r
- - r r
) ( ) (
) (
(1.105)
1.5.1.1 Component Representation of a Second-Order Tensor in the
Cartesian Basis
As seen before, a vector which has 3 independent components is represented in a
Cartesian space as shown in Figure 1.11. An arbitrary second-order tensor has 9
independent components, so we would need a hyperspace to represent all its components.
Afterwards, a device is introduced to represent the second-order tensor components in the
Cartesian basis.
An arbitrary second-order tensor T is represented in the Cartesian basis by:
1 TENSORS

33
3 3 33 2 3 32 1 3 31
3 2 23 2 2 22 1 2 21
3 1 13 2 1 12 1 1 11
3 3 2 2 1 1




e e e e e e
e e e e e e
e e e e e e
e e e e e e e e T
- - -
- - - -
- - -
- -
T T T
T T T
T T T
T T T T
i i i i i i j i ij

(1.106)
Next, we calculate the projection of T onto
k
e

:
3 3 2 2 1 1

e e e e e e e e e T
k k k i ik jk i ij k j i ij k
T T T T T T - - c
(1.107)
thereby defining three vectors, namely:

- -
- -
- -

)

(
3 33 2 23 1 13 3
)

(
3 32 2 22 1 12 2
)

(
3 31 2 21 1 11 1
3
2
1

3

2

1

e
e
e
t e e e e
t e e e e
t e e e e
e e T
,
,
,
T T T T
T T T T
T T T T
T
i i
i i
i i
i ik k
k
k
k

(1.108)
Graphical representation of these three vectors
)

(
1
e
t
,
,
)

(
2
e
t
,
,
)

(
3
e
t
,
, in the Cartesian basis, is
shown in Figure 1.13. Note also that
)

(
1
e
t
,
is the projection of T onto
1

e , [ ] 0 , 0 , 1

) 1 (

i
n ,
which can be verified by:
)

(
31
21
11
33 32 31
23 22 21
13 12 11
1
0
0
1
)

(
e
n T
i i
t
T
T
T
T T T
T T T
T T T

(
(
(

(
(
(

(
(
(


(1.109)









Figure 1.13: The projection of T in the Cartesian basis.
The same result obtained in (1.109) could have been evaluated by the scalar product of T ,
given in (1.106), with the basis
1

e , i.e.:
[
]
)

(
3 31 2 21 1 11
1 3 3 33 2 3 32 1 3 31
3 2 23 2 2 22 1 2 21
3 1 13 2 1 12 1 1 11 1
1




e
t e e e
e e e e e e e
e e e e e e
e e e e e e e T
,
- -
- - -
- - - -
- - -

T T T
T T T
T T T
T T T

(1.110)
where we have used the orthogonality property of the basis, i.e. 1

1 1
e e , 0

1 2
e e ,
0

1 3
e e . Taking into account the components are represented in matrix form, (see
Figure 1.14), we can establish that, the diagonal terms (
11
T ,
22
T ,
33
T ) are normal to the
plane defined by the unit vectors (
1

e ,
2

e ,
3

e ), hence they will be referred to as normal



1
x

2
x

3
x

)

(
2
e
t
,


)

(
3
e
t
,


)

(
1
e
t
,


2

e

3

e

1

e
NOTES ON CONTINUUM MECHANICS

34
components. The components displayed tangentially to the plane are called tangential
components, and correspond to the off-diagonal terms of
ij
T .














Figure 1.14: Representation of the second-order tensor components in the Cartesian
coordinate system.
NOTE: Throughout the textbook, we will use the following notations:


)

(
)

(
)

( )

(
l i jl ij
l i jk kl ij
l k kl j i ij
e e
e e
e e e e B A



B A
B A
B A
c
(1.111)

Note that the index is not repeated more than twice either in symbolic notation or in
indicial notation. Also note that the indicial notation is equivalent to the tensor notation
only when dealing with scalars, e.g. /
ij ij
B A B A: , or
i i
b a b a
,
,
.
1.5.2 Properties of Tensors
1.5.2.1 Tensor Transpose
Let A be a second-order tensor, the transpose of A is defined as:
)

( )

(
i j ij j i ji
T
e e e e A A A (1.112)
If
ij
A are the components of A , it follows that the components of the transpose are:
Tensorial notation
Symbolic notation
Cartesian basis
Indicial notation

1
x

2
x

3
x

1 11

e T

2 21

e T
3 31

e T

1 12

e T

3 32

e T

2 22

e T

3 33

e T

1 13

e T
2 23

e T

)

(
1
e
t
,


)

(
2
e
t
,


)

(
3
e
t
,


(
(
(

33 32 31
23 22 21
13 12 11
T T T
T T T
T T T
T
ij

1 TENSORS

35
ji ij
T
A ) (A (1.113)
If v u A
, ,
, the transpose of the dyad A is given by u v A
, ,

T
:
( )
( )
( )
( )
j i ji i j ij
T
j i ij
j i i j i j j i
T
j i j i
j j i i i i j j
T
j j i i
T
T
e e e e e e
e e e e e e
e e e e e e
u v v u A







A A A
v u v u v u
u v u v v u
, , , ,

(1.114)
Let A and B be second-order tensors and c, , be scalars, and the following
relationships are valid:
A A
T T
) ( ;
T T T
A B A B , c , c - - ) ( ;
T T T
B A A B ) ( (1.115)
( ) ( )
( ) ( ) B A e e e e B A
B A e e e e B A




ji ij il jk kl ij l k kl i j ij
T
ji ij jk il kl ij k l kl j i ij
T
B A B A B A
B A B A B A
c c
c c


: :
: :

(1.116)
The transpose of the matrix A is formed by changing rows for columns and vice versa,
i.e.:
(
(
(

(
(
(


(
(
(

33 23 13
32 22 12
31 21 11
33 32 31
23 22 21
13 12 11
33 32 31
23 22 21
13 12 11
A A A
A A A
A A A
A A A
A A A
A A A
A A A
A A A
A A A
T
T transpose
A A (1.117)
Problem 1.13: Let A , B and C be arbitrary second-order tensors. Demonstrate that:
( ) ( ) ( ) B C A C A B C B A : : :
T T

Solution: Expressing the term ( ) C B A : in indicial notation we obtain:
( ) ( )
( )
kj ik ij jq il kp pq lk ij
q l kp j i pq lk ij
q p pq k l lk j i ij
C B A C B A
C B A
C B A



c c c
c e e e e
e e e e e e C B A


:
: :

Note that, when we are dealing with indicial notation the position of the terms does not
matter, i.e.:
ik kj ij kj ij ik kj ik ij
B C A C A B C B A
We can now observe that the algebraic operation
ij ik
A B is equivalent to the components of
the second-order tensor
kj
T
) ( A B , thus,
( ) C A B A B :
T
kj kj
T
kj ij ik
C C A B ) ( .
Likewise, we can state that ( ) B C A :
T
ik kj ij
B C A .
Problem 1.14: Let u
,
, v
,
be vectors and A be a second-order tensor. Show that the
following relationship holds:
u A v v A u
, , , ,

T

Solution:
l jl j j jl l
il i jl kj k jk k il jl i
i i l j jl k k k k j l jl i i
T
u A v v A u
u A v v A u
u A v v A u





c c c c
e e e e e e e e
u A v v A u

, , , ,

NOTES ON CONTINUUM MECHANICS

36
1.5.2.2 Symmetry and Antisymmetry
1.5.2.2.1 Symmetric tensor
A second-order tensor A is symmetric, i.e.:
sym
A A = , if the tensor is equal to its transpose:
symmetric is if

A A A =
ji ij
components in T
A A (1.118)
in matrix form:
(
(
(

=
33 23 13
23 22 12
13 12 11
A A A
A A A
A A A
sym T
A A A A
(1.119)
From the above it is clear that a symmetric second-order tensor has 6 independent
components, namely:
11
A ,
22
A ,
33
A ,
12
A ,
23
A ,
13
A .
According to equation (1.118), a symmetric tensor can be represented by:
) (
2
1
) (
2
1
2
T
ji ij ij
ji ij ij
ji ij ij ij
ji ij
A A A - -
-
- -

A A A
A A A
A A A A
A A

(1.120)
A fourth-order tensor C , whose components are
ijkl
C , may have the following types of
symmetries:
Minor symmetry:
jilk ijlk jikl ijkl
C C C C
(1.121)
Major symmetry:
klij ijkl
C C
(1.122)
A fourth-order tensor that does not exhibit any kind of symmetry has 81 independent
components. If the tensor C has only minor symmetry, i.e. symmetry in ) 6 ( ji ij , and
symmetry in ) 6 ( lk kl , the tensor features 36 independent components. If besides
presenting minor symmetry it also provides major symmetry, the tensor features 21
independent components.
1.5.2.2.2 Antisymmetric tensor
A tensor A is antisymmetric (also called skew-symmetric tensor or skew tensor), i.e.:
skew
A A = :
ric antisymmet is if A A A = - -
ji ij
components in T
A A (1.123)
which broken down into its components is the same as:
(
(
(

- -
- -
0
0
0
23 13
23 12
13 12
A A
A A
A A
skew T
A A A
(1.124)
Therefore, an antisymmetric second-order tensor has 3 independent components, namely:
12
A ,
23
A ,
13
A .
1 TENSORS

37
Under the conditions expressed in (1.123), an antisymmetric tensor can be represented by:
) (
2
1
) (
2
1
2
T
ji ij ij
ji ij ij
ji ij ij ij
A A A - -
-
- -
A A A
A A A
A A A A

(1.125)
Let us consider an antisymmetric second-order tensor denoted by W, then satisfy the
above relationship (1.125):
) (
2
1
) (
2
1
) (
2
1
il jk jl ik kl il jk kl jl ik kl ji ij ij
c c c c c c c c - - - W W W W W W (1.126)
Using the relation between the Kronecker delta and the permutation symbol given by
(1.62), i.e.
lkr ijr il jk jl ik
- - c c c c , the equation (1.126) is rewritten as:
lkr ijr kl ij
W W
2
1
- (1.127)
Expanding the term
lkr kl
W , for the dummy indices ( k , l ), we can obtain the following
nonzero terms:
r r r r r r lkr kl 23 32 13 31 32 23 12 21 31 13 21 12
W W W W W W W - - - - - (1.128)
thus,
r lkr kl
lkr kl
lkr kl
lkr kl
w
w r
w r
w r
2
2 2 3
2 2 2
2 2 1
3 12 21 12
2 13 31 13
1 23 32 23

- - -
-
- - -

W
W W W W
W W W W
W W W W
(1.129)
In which we assume the following variables have changed:
(
(
(

-
-
-

(
(
(

- -
-
(
(
(

0
0
0
0
0
0
0
0
0
1 2
1 3
2 3
23 13
23 12
13 12
32 31
23 21
13 12
w w
w w
w w
ij
W W
W W
W W
W W
W W
W W
W
(1.130)
Hence, we introduce the axial vector w
,
associated with the antisymmetric tensor, W, as:
3 3 2 2 1 1

e e e w w w - - w
,
(1.131)
The magnitude of the axial vector w
,
is given by:
2
12
2
13
2
23
2
3
2
2
2
1
2
2
W W W - - - - w w w w w w
, , ,
. (1.132)
Substituting (1.129) into (1.127) and by considering that
rij ijr
we obtain:
rij r ij
w - W
(1.133)
Multiplying both sides of the equation (1.133) by
kij
we can obtain:
k rk r kij rij r ij kij
w w w 2 2 - - - c W (1.134)
where we have applied the relation
rk kij rij
c 2 , which was evaluated in Problem 1.7,
thus we can conclude that:
NOTES ON CONTINUUM MECHANICS

38
ij kij k
w W
2
1
-
(1.135)
Graphical representation of the antisymmetric tensor components and its corresponding
axial vector, in the Cartesian system, is shown in Figure 1.15.












Figure 1.15: Antisymmetric tensor components and the axial vector.
Let a
,
and b
,
be arbitrary vectors and W be an antisymmetric tensor, it follows that:
b W a b W a a W b
,
,
,
, ,
,
-
T

(1.136)
when b a
,
,
, it holds that:
0 ) ( a a W a W a
, , , ,
: (1.137)
NOTE: Note that ) ( a a
, ,
is a symmetric second-order tensor. Later on we will show that
the result of the double contraction between a symmetric tensor and an antisymmetric
tensor equals zero.
Let us consider an antisymmetric tensor W and an arbitrary vector a
,
. The components of
the scalar product a W
,
are given by:
3 33 2 32 1 31
3 23 2 22 1 21
3 13 2 12 1 11
3 3 2 2 1 1
3
2
1
a W a W a W
a W a W a W
a W a W a W
a W a W a W a W
- -
- -
- -
- -
i
i
i
i i i j ij

(1.138)
Bearing in mind that the normal components are equal to zero for an antisymmetric tensor,
i.e., 0
11
W , 0
22
W , 0
33
W , the scalar product (1.138) becomes:
( )

-
-
-

2 32 1 31
3 23 1 21
3 13 2 12
3
2
1
a W a W
a W a W
a W a W
i
i
i
i
a W
,

(1.139)
The above components are the same as the result of the algebraic operation a
,
,
r w :

23 1
W - w

1
x

2
x

3
x

12
W

12
W

23
W

13
W

13
W

13 2
W w

12 3
W - w

3 3 2 2 1 1

e e e w w w - - w
,


23
W

(
(
(

- -
-
0
0
0
23 13
23 12
13 12
W W
W W
W W
W
ij

1 TENSORS

39
( ) ( ) ( )
( ) ( ) ( )
3 2 32 1 31 2 3 23 1 21 1 3 13 2 12
3 2 1 1 2 2 3 1 1 3 1 3 2 2 3
3 2 1
3 2 1
3 2 1



e e e
e e e
e e e
a
a W a W a W a W a W a W
a a a a a a
a a a
- - - - -
- - - - - - -
r
w w w w w w
w w w
,
,
w

(1.140)
where
32 23 1
W W - w ,
31 13 2
W W - w ,
21 12 3
W W - w . Then, given an antisymmetric
tensor W and the axial vector w
,
associated with W, it holds that:
a a W
,
,
,
r w
(1.141)
for any vector a
,
. The property (1.141) could easily have been obtained by taking into
account the components of W given by (1.133), i.e.:
( ) ( )
i k j ijk k jik j k ik i
w w a a W
,
,
,
r - w a a a W
(1.142)
The vector w
,
can be represented by its magnitude, . w
,
, and by the unit vector
codirectional with w
,
, i.e.
*
1

e . w
,
. Then, the equation (1.141) can still be expressed as:
a e a a W
, ,
,
,
r r
*
1

. w (1.143)
Additionally, we can choose two unit vectors
*
2

e ,
*
3

e , which make up an orthonormal basis


with the unit vector
*
1

e , (see Figure 1.16), so that:


*
2
*
1
*
3
*
1
*
3
*
2
*
3
*
2
*
1

;

;

e e e e e e e e e r r r (1.144)






Figure 1.16: Orthonormal basis defined by the axial vector.
By representing the vector a
,
in this new basis,
*
3
*
3
*
2
*
2
*
1
*
1

e e e a a a a - -
,
, the relationship
shown in (1.143) obtains the form below:
[ ] a e e e e
e e e e e e e e
e e e e a e a W
e e
0
,
_ _ _
, ,

-
- r - r - r
- - r r
-

)

(
)

( )

(
)

(

*
3
*
2
*
2
*
3
*
2
*
3
*
3
*
2
*
2
*
3
*
1
*
3
*
3
*
2
*
1
*
2
*
1
*
1
*
1
*
3
*
3
*
2
*
2
*
1
*
1
*
1
*
1

.
. .
. .
a a a a a
a a a

(1.145)
Thus, an antisymmetric tensor can be represented, in the space defined by the axial vector,
as follows:
)

(
*
3
*
2
*
2
*
3
e e e e W - . (1.146)
Note that by using the antisymmetric tensor representation shown in (1.146), the
projections of the tensor W according to directions
*
1

e ,
*
2

e and
*
3

e are respectively:

*
3

e

*
1

e
3

e

2

e

1

e

*
2

e

*
1

e . w
,

NOTES ON CONTINUUM MECHANICS

40
*
2
*
3
*
3
*
2
*
1

;

;

e e W e e W 0 e W . . -
,

(1.147)
We can also verify that:
[ ]
[ ] . .
. .
- -
-


*
3
*
3
*
2
*
2
*
3
*
2
*
3
*
2
*
2
*
3
*
2
*
2
*
3
*
3
*
2
*
3

)

(

)

(

e e e e e e e W e
e e e e e e e W e
(1.148)
Then, the tensor components of W in the basis formed by the orthonormal basis
*
1

e ,
*
2

e ,
*
3

e , are given by:


(
(
(

-
0 0
0 0
0 0 0
*
.
.
ij
W
(1.149)
In Figure 1.17 we can see these components and the axial vector representation. Note that
if we take any basis that is formed just by rotation along the
*
1

e -axis, the components of


W in this new basis will be the same as those provided in (1.149).










Figure 1.17: Antisymmetric tensor components in the space defined by the axial vector.
1.5.2.2.3 Additive decomposition. Symmetric and antisymmetric part
Any arbitrary second-order tensor A can be split additively into a symmetric and an
antisymmetric part:
skew sym T T
skew sym
A A A A A A A
A A
- - - -
_ _
) (
2
1
) (
2
1

(1.150)
or, into its components:
) (
2
1
) (
2
1
ji ij
skew
ij ji ij
sym
ij
and A A A A A A - - (1.151)
If A and B are arbitrary second-order tensors, it holds that:
( ) ( ) ( ) [ ]
[ ] A B A A B B A
A B A A B A A B A A B A A B A


-
-
(

-
sym T T T
T T T
T
T T
sym
T
2
1
2
1
2
1

(1.152)

*
3

e

*
1

e
.

*
2

e

1
x

2
x
3
x
.

*
1

e . w
,


(
(
(

-
0 0
0 0
0 0 0
*
.
.
ij
W
1 TENSORS

41
Problem 1.15: Show that 0 W : is always true when is a symmetric second-order
tensor and W is an antisymmetric second-order tensor.
Solution:
(scalar) )

( )

(
ij ij jk il lk ij k l lk j i ij
W W W o o o c c e e e e W : :
Thus,
_ _ _
33 33
32 32
31 31
3 3
23 23
22 22
21 21
2 2
13 13
12 12
11 11
1 1
W
W
W
W
W
W
W
W
W
W
W
W W
o
-
o
-
o
o -
o
-
o
-
o
o -
o
-
o
-
o
o o
j j j j j j ij ij

Taking into account the characteristics of a symmetric and an antisymmetric tensor, i.e.
21 12
o o ,
13 31
o o ,
23 32
o o , and 0
33 22 11
W W W ,
12 21
W W - ,
13 31
W W - ,
23 32
W W - , the equation above becomes:
0 W :
Problem 1.16: Show that a) M Q M M Q M
, , , ,

sym
; b)
skew skew sym sym
B A B A B A : : : -

where M
,
is a vector, and Q, A , B are arbitrary second-order tensors.
Solution:
a) ( ) M Q M M Q M M Q Q M M Q M
, , , , , , , ,
- -
skew sym skew sym

Since the relation ( ) 0


_
, , , ,
tensor symmetric
skew skew
M M Q M Q M : holds, it follows that:
M Q M M Q M
, , , ,

sym

b)
skew skew sym sym
skew skew sym skew skew sym sym sym
skew sym skew sym
B A B A
B A B A B A B A
B B A A B A
: :
: : : :
: :
-
- - -
- -

_ _
0 0
) ( ) (

Then, it is also valid that:
skew skew skew sym sym sym
B A B A B A B A : : : : ;
Problem 1.17: Let T be an arbitrary second-order tensor, and n
,
be a vector. Check if the
relationship n T T n
, ,
is valid.

Solution:
l l l l
l kl k
l ik kl i
l k kl i i
e
e
e
e e e T n

) (

)

(

3 3 2 2 1 1
T n T n T n
T n
T n
T n
- -


c
,
and
l l l l
l lk k
l ki lk i
i i k l lk
e
e
e
e e e n T

) (

)

(
3 3 2 2 1 1
T n T n T n
T n
T n
n T
- -


c
,

With the above we can prove that
lk k kl k
T n T n , then:
n T T n
, ,

NOTES ON CONTINUUM MECHANICS

42
If T is a symmetric tensor, it follows that the relationship n T T n
, ,

sym sym
holds.
Problem 1.18: Obtain the axial vector w
,
associated with the antisymmetric tensor
skew
) ( a x
, ,
.
Solution: Let z
,
be an arbitrary vector, it then holds that:
z w z a x
, , , , ,
r
skew
) (
where w
,
is the axial vector associated with
skew
) ( a x
, ,
. Using the definition of an
antisymmetric tensor:
[ ] [ ] x a a x a x a x a x
, , , , , , , , , ,
- -
2
1
) ( ) (
2
1
) (
T skew

and by replacing it with z w z a x
, , , , ,
r
skew
) ( , we obtain:
[ ] [ ] z w z x a a x z w z x a a x
, , , , , , , , , , , , , ,
r - r - 2
2
1

By using the equation [ ] ) ( a x z z x a a x
, , , , , , , ,
r r - , (see Eq. (1.105)), the above equation
becomes:
[ ] z w z x a a x z z x a a x
, , , , , , , , , , , , ,
r r r r r - 2 ) ( ) (
with the above we can conclude that:
skew
with associated vector axial the is ) ( ) (
2
1
a x x a w
, , , , ,
r
1.5.2.3 Cofactor Tensor. Adjugate of a Tensor
Let A be a second-order tensor and a
,
, b
,
be arbitrary vectors then there is then a unique
tensor ) cof(A , known as the cofactor of A , as we can see below:
) ( ) ( ) ( b A a A b a A
,
,
,
,
r r ) cof(
(1.153)
We can also define the adjugate of A as:
[ ]
T
) (A A cof ) adj(

(1.154)
which satisfies the following condition:
[ ] ) adj( ) adj(
T T
A A (1.155)
The components of ) cof(A are obtained by expressing the equation (1.153) in terms of its
components, i.e.:
[ ] [ ]
kr jp ijk tpr it r kr p jp ijk r p tpr it
A A ) cof( b A a A b a ) cof( A A
(1.156)
By multiplying both sides of the equation by
qpr
and by also considering that
tq qpr tpr
c 2 , we can conclude that:
[ ] [ ]
[ ]
kr jp qpr ijk iq
kr jp qpr ijk
tq
qpr tpr it kr jp ijk tpr it
A A ) cof(
A A ) cof( A A ) cof(


2
1
2

A
A A
_
c

(1.157)
1.5.2.4 Tensor Trace
Lets start by defining the trace of the basis )

(
j i
e e :
1 TENSORS

43
ij j i j i
c e e e e

)

( Tr
(1.158)
Thus, we can define the trace of a second-order tensor A as follows:
33 22 11
)

( )

( )

( ) (
A A A
A A A Tr A A Tr Tr
- -

ii ij ij j i ij j i ij j i ij
c e e e e e e A
(1.159)
And, the trace of the dyad ) ( v u
, ,
can be evaluated as:
v u
e e e e v u v u
, ,
, , , ,


- -

3 3 2 2 1 1
)

( )

( ) ( ) (
v u v u v u
v u v u v u Tr v u Tr Tr
i i ij j i j i j i j i j i
c
(1.160)
NOTE: As we will show later, the tensor trace is an invariant, i.e. it is independent of the
coordinate system.
Let A , B be arbitrary tensors, then:
The transposed tensor trace is equal to the tensor trace:
( ) ( ) A A Tr Tr
T
(1.161)
The trace of ( ) B A - is the sum of traces:
( ) ( ) ( ) B A B A Tr Tr Tr - - (1.162)
Proving this is very simple:
( ) ( ) ( )
( ) ( ) ( ) [ ] ( ) ( )
33 22 11 33 22 11 33 33 22 22 11 11
B B B A A A B A B A B A
Tr Tr Tr
- - - - - - - - - -
- - B A B A

(1.163)
The scalar product trace ) ( B A becomes:
( ) [ ]
[ ]
( ) A B B A
e e
e e e e B A






)

(
)

( )

(
Tr B A
Tr B A
B A Tr Tr
li il
im
m i jl lm ij
m l lm j i ij
_
c
c
(1.164)
and, the double scalar product ( : ) can be expressed in trace terms as:
( ) ( )
) ( ) (
) (
) (
B A B A
B A B A
B A
B A
B A





T T
kk
T
kk
T
kl
kl
T
il ik kl
kl
T
lj kj
jl jk il ik il ik lj kj
ij ij
Tr Tr
B A B A
B A B A
B A
c c
c c c c
_
_
:
(1.165)
Similarly, it is possible to show that:
( ) ( ) ( )
ki jk ij
C B A Tr Tr Tr B A C A C B C B A
(1.166)
ii
A Tr ) (A
[ ]
jj ii
A A Tr Tr Tr ) ( ) ( ) (
2
A A A
( ) ( )
li il
A A Tr Tr
2
A A A ; ( ) ( )
ki jk ij
A A A Tr Tr
3
A A A A
(1.167)

NOTES ON CONTINUUM MECHANICS

44
Problem 1.19: Let T be a second-order tensor. Show that:
( ) ( ) ( ) ( )
m
m
T
m
T
T
m
and T T T T Tr Tr .
Solution:
( ) ( ) ( )
m
T T T T T
T
m
T T T T T T T T
For the second demonstration we can use the trace property ( ) ( ) T T Tr Tr
T
, thus:
( ) ( ) ( )
m
T
m
m
T
T T T Tr Tr Tr
1.5.2.5 Particular Tensors
1.5.2.5.1 Unit Tensors
The second-order unit tensor, also called the identity tensor, is defined as:
j i i i j i ij
e e e e e e 1

1 c
(1.168)
where 1 is the matrix with the components of tensor 1.
ij
c is the Kronecker delta symbol
defined in (1.48).
Fourth-order unit tensors can be defined as follows:

e e e e e e e e 1 1


k j i ijk k j i j ik
I c c I (1.169)

e e e e e e e e 1 1


k j i ijk k j i jk i
I c c I
(1.170)

e e e e e e e e 1 1


k j i ijk k j i k ij
I c c I (1.171)
Taking into account the fourth-order unit tensors defined above, it holds that:
( ) ( ) ( )
( ) ( )
A
e e e e
e e e e e e e e A



j i ij j i k j ik
j i q kp pq j ik q p pq k j i j ik


A A
A A


c c
c c c c c c : : I

(1.172)
and
( ) ( ) ( )
( ) ( )
T
j i ji j i k jk i
j i q kp pq jk i q p pq k j i jk i
A
e e e e
e e e e e e e e A





A A
A A


c c
c c c c c c : : I

(1.173)
and
( ) ( ) ( )
( ) ( )
1 A
e e e e
e e e e e e e e A
) (


Tr
A A
A A



j i ij kk j i k k ij
j i q kp pq k ij q p pq k j i k ij
c c c
c c c c c c


: : I

(1.174)
The symmetric part of the fourth-order unit tensor I is defined as:
( ) ( )
jk i j ik ijk
components in sym
c c c c

- - =
2
1
2
1
I 1 1 1 1 I I (1.175)
The property of the tensor product is presented below. Consider a second-order unit
tensor,
j i ij
e e 1

c . Then, the tensor product can be defined as:
1 TENSORS

45

( ) ( ) ( )

e e e e e e e e 1 1


j k i k ij k k j i ij
c c c c
(1.176)
which is the same as:
( )

e e e e 1 1


k j i j ik
c c I (1.177)
and the tensor product is defined as:
( ) ( ) ( )
j k i k ij k k j i ij
e e e e e e e e 1 1



c c c c
(1.178)
or
( )

e e e e 1 1


k j i jk i
c c I
(1.179)
The antisymmetric part of I is defined as:
( ) ( )
jk i j ik
skew
ijk
components in skew
c c c c

- -
2
1
2
1
I 1 1 1 1 I (1.180)
With a second-order tensor A and a vector b
,
, the following relationships are valid:
b 1 b
, ,

sym sym
A A A A : : I I ;
ii
A Tr ) (A 1 A:
( ) ( )
li il
A A Tr Tr A A A 1 A
2 2
:
( ) ( )
ki jk ij
A A A Tr Tr A A A A 1 A
3 3
:
(1.181)
Problem 1.20: Show that ) (T 1 T Tr : , where T is an arbitrary second-order tensor.
Solution:
) (

T
e e e e 1 T
Tr
T T T
T
T


jj ii ij ij
jl ik kl ij
l k kl j i ij
c
c c c
c : :

1.5.2.5.2 Levi-Civita Pseudo-Tensor
The Levi-Civita Pseudo-Tensor, also known as the Permutation Tensor, is a third-order pseudo-
tensor and is denoted by:
k j i ijk
e e e


(1.182)
which is not a true tensor in the strict meaning of the word, and whose components
ijk

were defined in (1.55), the permutation symbol.
1.5.2.6 Determinant of a Tensor
The determinant of a tensor is a scalar and is expressed as:
_
T
k j i ijk k j i ijk
A
A A
3 2 1 3 2 1
) ( A A A A A A det =
(1.183)
NOTES ON CONTINUUM MECHANICS

46
It is also an invariant (independent of the adopted system). Demonstrating the equation
above (1.183) can be done starting directly from the determinant:
3 2 1
3 2 31 3 3 2 21 2 3 2 11 1
3 2 3 31 3 2 2 21 3 2 1 11
22 13 23 12 31 32 13 33 12 21 32 23 33 22 11
33 32 31
23 22 21
13 12 11
) ( ) ( ) (
) ( ) ( ) (
) (
k j i ijk
k j jk k j jk k j jk
k j jk k j jk k j jk
A A A
A A A A A A A A A
A A A A A A A A A
A A A A A A A A A A A A A A A
A A A
A A A
A A A
det

- -
- - -
- - - - -
A A

(1.184)
Some determinant properties with second-order tensors are described below:
1 ) ( 1 det
(1.185)
We can conclude from (1.183) that:
) ( ) ( A A det det
T

(1.186)
We can also show that:
) ( ) ( ) ( B A B A det det det , ) ( ) (
3
A A det det c c where c is a scalar (1.187)
A tensor ) (A is said to be singular if 0 ) ( A det .
If you exchange two rows or columns, the determinant sign changes.
If all elements of a row or column equal zero, the determinant is also zero.
If you multiply all the elements of a row or column by a constant c (scalar), the
determinant is A c .
If two or more rows (or column) are linearly dependent, the determinant is zero.
Problem 1.21: Show that
kq jp rt rjk tpq
A A A A .
Solution:
We start with the following definition:
3 2 1 3 2 1 k j r tpq rjk tpq k j r rjk
A A A A A A A A
(1.188)
and also taking into account that the term
tpq rjk
can be replaced by (1.61):
rp jt kq rt kp jq kt jp rq kp jt rq kt jq rp kq jp rt
kq kp kt
jq jp jt
rq rp rt
tpq rjk
c c c c c c c c c c c c c c c c c c
c c c
c c c
c c c
- - - - -

(1.189)
Then, by substituting (1.189) into (1.188) we can obtain:
( ) ( ) ( )
qk pj tr rjk kq jp rt rjk
qk pj jk t qk pj jk t qk pj jk t
q t p p q t t p q p t q t q p q p t tpq
A A A A A A
A A A A A A A A A
A A A A A A A A A A A A A A A A A A


- -
- - - - -
3 3 2 2 1 1
3 2 1 3 2 1 3 2 1 3 2 1 3 2 1 3 2 1
A

Problem 1.22: Show that
kq jp rt tpq rjk
A A A
6
1
A .
Solution:
1 TENSORS

47
Starting with the definition
kq jp rt rjk tpq
A A A A , (see Problem 1.21), and by multiplying
both sides of the equation by
tpq
, we obtain:
kq jp rt tpq rjk tpq tpq
A A A A
(1.190)
Using the property defined in expression (1.62), we obtain
6 - -
tt pp tt tp tp pp tt tpq tpq
c c c c c c c . Then, the relationship (1.190) becomes:
kq jp rt tpq rjk
A A A
6
1
A
Problem 1.23: Let a
,
, b
,
be arbitrary vectors and c be a scalar. Show that:
( ) b a b a 1
,
,
,
,
- -
2 3
c j j c j det

(1.191)
Solution: The determinant of A is given by
3 2 1 k j i ijk
A A A A . If we denote by
j i ij ij
b a A c jc - , thus,
1 1 1
b a A
i i i
c jc - ,
3 3 3
b a A
k k k
c jc - ,
3 3 3
b a A
k k k
c c - , then
the equation in (1.191) can be rewritten as:
( ) ( )( )( )
3 3 2 2 1 1
b a b a b a det
k k j j i i ijk
c jc c jc c jc c j - - - - b a 1
,
,
(1.192)
By developing the equation (1.192), we obtain:
( ) [
]
3 2 1
3
3 2 1
2
2 3 1
2
1 3 2
2
3 2 1
2
3 1 2
2
2 1 3
2
3 2 1
3
b b b a a a b b a a b b a a b a b a
b a b a b a det
k j i k j i j k i i k j
k j i k i j j i k k j i ijk
c c jc c jc c jc
c c c j c c c j c c c j c c c j c j
- - - -
- - - - - b a 1
,
,

Note that:
3
123
3
3 2 1
3
j j c c c j
k j i ijk
,
0
0
) ( ) ( ) (
) (
2 1 1 2 213 2 1 2 1 123 2 1 3 3 2 1
3 1 1 3 3 1 3 1 3 1 2 2 3 1
2
1 1 2 2 3 3
2
1 23 2 3 1 3 12
2
3 2 1 3 1 2 2 1 3
2

-
-
- - - -
- -

b b a a b b a a b b a a b b a a
b b a a b b a a b b a a b b a a
b a b a b a b a b a b a
b a b a b a




j i ij k j i ijk
k i k i j k i ijk
i i j j k k
k j i ijk k i j ijk j i k ijk
c
c
c j c j j
c c c c c c c j
c b a
,
,

0
3 2 1
b b b a a a
k j i ijk

Notice that, there was no need to expand the terms
2 3 1 j k i ijk
c b b a a ,
3 2 1 k j i ijk
c b b a a , and
3 2 1
b b b a a a
k j i ijk
to realize that these terms equal zero, since
0 ) (
2 3 1 2 3 1
r
j j j k i ijk
c c b b b b a a a a
, ,
, similarly for other terms.
Taking into account the above considerations we can prove that:
( ) b a b a 1
,
,
,
,
- -
2 3
c j j c j det
For the particular case when 1 j the above equation becomes:
( ) b a b a 1
,
,
,
,
- - 1 c c det

Then, it is simple to prove that ( ) 0 b a
,
,
c det , since
( ) [ ] 0 ) (
3 2 1
3
3 2 1
3
r a a a b a
, , ,
,
,
b b b b b b a a a det c c c
k j i ijk


The following relations are satisfied:
[ ] [ ] ) )( ( ) ( ) ( ) ( 1 ) ( ) (
2
b b a a b a b a b a a b b a 1
, ,
, ,
,
,
,
,
,
, ,
, ,
,
- - - - - - c, , c , c det
(1.193)
where c, , are scalars. If 0 , we can regain the equation ( ) b a b a 1
,
,
,
,
- - c c 1 det ,
(see Problem 1.23). If , c we obtain:
NOTES ON CONTINUUM MECHANICS

48
( ) ( ) ( ) ( ) ( )( )
( ) ( )
(

r - -
(

- - - - - -


2
2
2
2 1
1
b a b a
b b a a b a b a b a a b b a 1
,
,
,
,
, ,
, ,
,
,
,
,
,
, ,
, ,
,
c c
c c c c c det

(1.194)
where we have used the property ( ) ( )( ) ( )
2 2
b a b b a a b a
,
,
, ,
, ,
,
,
r - - , (see Problem 1.1).
It is also true that:
( ) [ ] [ ] ) ( ) ( ) ( ) (
3 2 2 3
B B A A B A B A det adj Tr adj Tr det det , c, , c c , c - - - - (1.195)
Moreover, in the particular case when 1 c , 1 A , b a B
,
,
, and bearing in mind that
( ) 0 b a
,
,
det , and ( ) 0 b a
,
,
cof , we can conclude that:
( ) ( ) [ ] [ ] b a 1 b a 1 b a 1
,
,
,
,
,
,
- - - - , , , , 1 1
j i
b a Tr Tr det det
(1.196)
which has already been demonstrated in Problem 1.23.
We next show that the following property is valid:
[ ] [ ] ) ( ) ( ) ( ) ( ) ( c b a A c A b A a A
,
,
, ,
,
,
r r det

(1.197)
To achieve this goal we start with the definition of the scalar triple product given in (1.69),
i.e. ( )
k j i ijk
c b a r c b a
,
,
,
, and by multiplying both sides of this equation by the
determinant of A we obtain:
( ) A A c b a
k j i ijk
c b a r
,
,
,
(1.198)
It was proven in Problem 1.21 that
rk qj pi pqr ijk
A A A A , thus:
( )
[ ] [ ] ) ( ) ( ) ( ) ( ) ( ) (
) )( )( (
a A c A b A c A b A a A
A A c b a
, ,
,
,
,
,
,
,
,

r r
r
k rk j qj i pi pqr k j i rk qj pi pqr k j i ijk
c A b A a A c b a A A A c b a

(1.199)
1.5.2.7 Inverse of a Tensor
We use the notation
1 -
A to denote the inverse of A , which is defined as follows:
if 1 A A A A A A
- - -
=
1 1 1
0 (1.200)
Or in indicial notation:
if
ij kj ik kj ik ij
c
- - -
= A A A A A
1 1 1
0 A (1.201)
To obtain the inverse of a tensor we start from the definition of the adjugate tensor given
in (1.153), i.e. ) ( ) ( ) ( b A a A b a A
,
,
,
,
r r ) adj(
T
. Then by applying the dot product
between an arbitrary vector d
,
and this equation we obtain:
[ ] { [ ] [ ]
[ ]
_
, ,
,
, ,
,
, ,
,
, ,
,
,
c
d A A b A a A
d 1 b A a A d b A a A d b a A



-
r
r r r
1
) ( ) (
) ( ) ( ) ( ) ( ) (
T
) adj(

(1.202)
In equation (1.199) we demonstrated that ( ) [ ] ) ( ) ( ) ( c A b A a A A b a c
,
,
,
,
, ,
r r thus,
1 TENSORS

49
[ ] { [ ] ( ) d A b a A d A A b A a A d b a A
, ,
,
, ,
,
, ,
,

- -
r r r
1 1
) ( ) ( ) (
T
) adj( (1.203)
Denoted by ) ( b a p
,
, ,
r , the above equation (1.203) can be rearranged as follows:
[ ] {
[ ]
[ ] [ ] [ ] d b a A A d b a A
A A
A A
, ,
,
, ,
,
r r


-
-
-
) ( ) (
1
1
1
: : ) adj(
d p A d p ) adj(
d A p d p ) adj(
i k ki i k ki
i ki k i k ki
(1.204)
Thus, we can conclude that:
[ ]
-1
A A A) adj(
[ ] [ ]
T
) cof( ) adj( A
A
A
A
A
1 1
1

-

(1.205)
Let A and B be invertible tensors, the following properties hold:

( )
( )
[ ]
1 1
1 1
1
1
1 1 1
) ( ) (
1
) (
- -
- -
-
-
- - -


A A
A A
A A
A B B A
det det
,
,

(1.206)
The following notation will be used to represent the inverse transpose:
1 1
) ( ) (
- - -
= =
T T T
A A A (1.207)
Next, we prove the relation ) adj( ) adj( ) adj( A B B A holds. To do this, we take the
definition of the inverse of a tensor given in (1.205) as a starting point:
[ ] [ ]
[ ] [ ]
( ) [ ] [ ]
[ ]
[ ] [ ]
[ ] [ ] ) adj( ) adj( ) adj(
) adj( ) adj(
) adj(
) adj( ) adj(
) adj( ) adj(
) adj( ) adj(
A B B A
A B
B A
B A
B A A B B A B A
A B A B B A
A
A
B
B
A B







-
- - - -
1
1 1 1 1

(1.208)
where we have used the property B A B A . Similarly, it is possible to show that
[ ] [ ] ) cof( ) cof( ) cof( B A B A .
Procedure for obtaining the inverse of the matrix A
1) Obtain the cofactor matrix: ) (A cof as follows:
Consider the matrix A as:
(
(
(

33 32 31
23 22 21
13 12 11
A A A
A A A
A A A
A
(1.209)
We define the matrix M, where the component
ij
M is the determinant of the resulting
matrix by removing the
th
i row and the
th
j column, i.e.:
NOTES ON CONTINUUM MECHANICS

50
(
(
(
(
(
(
(
(

22 21
12 11
23 21
13 11
23 22
13 12
32 31
12 11
33 31
13 11
33 32
13 12
32 31
22 21
33 31
23 21
33 32
23 22
A A
A A
A A
A A
A A
A A
A A
A A
A A
A A
A A
A A
A A
A A
A A
A A
A A
A A
M (1.210)
Thus, we define the cofactor matrix as:
ij
j i
M A
-
- ) 1 ( ) ( cof (1.211)
2) Obtain the adjugate matrix, ) adj(A , as follows:
[ ]
T
) (A A cof ) adj(
(1.212)
3) Obtain the inverse matrix:
A
A
A
) adj(

-1

(1.213)
So, the relation [ ] 1 A A A ) adj( holds, where 1 is the identity matrix.
Taking into account (1.64), we can express the components of the first, second, and third
row of the cofactor matrix, (1.211), as:
k j ijk i 3 2 1
A A M ,
k j ijk i 3 1 2
A A M ,
k j ijk i 2 1 3
A A M ,
respectively.
Problem 1.24: Let A be an arbitrary second-order tensor. Show that there is a nonzero
vector 0 n
,
,
so that 0 n A
,
,
if and only if 0 ) ( A det , Chadwick (1976).
Solution: Firstly, we show that, if 0 n A A
,
,
= 0 ) ( det . Secondly, we show that, if
0 ) ( = A A 0 n det
,
,
.
We assume that 0 ) ( = A A det , and we choose an arbitrary basis } , , { h g f
,
,
,
(linearly
independent). We apply these vectors in the definition seen in (1.197):
( ) [ ] ) ( ) ( ) ( h A g A f A A h g f
,
,
, ,
,
,
r r
Due to the fact that 0 ) ( = A A det , the implication is that:
[ ] 0 ) ( ) ( ) ( r h A g A f A
,
,
,

Thus, we can conclude that the vectors ) ( f A
,
, ) ( g A
,
, ) ( h A
,
, are linearly dependent.
This implies that there are nonzero scalars c, , , so that:
( ) 0 n A 0 h g f A 0 h A g A f A
,
,
, ,
,
, , ,
,
,
- - - - , c , c ) ( ) ( ) (
where 0 h g f n
, ,
,
,
,
- - , c since } , , { h g f
,
,
,
is linearly independent, (see Problem 1.10).
Now we choose two vectors k
,
, m
,
, which are linearly independent to n
,
. We apply
definition (1.199) once more:
( ) [ ] ) ( ) ( ) ( n A m A k A A n m k
, ,
,
, ,
,
r r
Considering that 0 n A
,
,
, and ( ) 0 r n m k
, ,
,
owing to the fact that k
,
, m
,
, n
,
are linearly
independent, we can conclude that:
1 TENSORS

51
( ) 0 0
0
r

A A n m k
_
, ,
,

1.5.2.8 Orthogonal Tensors
Orthogonal tensors play an important role in continuum mechanics. A second-order tensor
( Q) is said to be orthogonal when the transpose (
T
Q ) is equal to the inverse (
1 -
Q ), i.e.
1 -
Q Q
T
. Then, it follows that:
ij kj ki jk ik
T T
c Q Q Q Q ; 1 Q Q Q Q (1.214)
A proper orthogonal tensor has the following properties:
The inverse of Q is equal to the transpose (orthogonality):
T
Q Q
-1
(1.215)
The tensor Q is a proper, rotation tensor, if:
1 ) ( - = Q Q det (1.216)
If 1 - Q , the orthogonal tensor is said to be improper (a reflection tensor).
If A and B are orthogonal tensors, the resulting tensor C B A is also an orthogonal
tensor, i.e.:
T T T T
C B A A B A B B A C
- - - -
) ( ) (
1 1 1 1
(1.217)
Consider two arbitrary vectors a
,
and b
,
. An orthogonal transformation applied to these
vectors becomes:
b Q b a Q a
,
,
,
,

~
;
~

(1.218)
And the dot product between these new vectors ( a
,
~
) and ( b
,
~
) is given by:
k k j
kj
ij ik k j ij k ik i i
T
b a b Q Q a b Q a Q b a

_
,
,
,
_
,
,
,
,
,
c
) )( (
~
~
) ( ) (
~
~
b a b Q Q a b Q a Q b a
1

(1.219)
So, it is also true when b a
,
,
~
~
, thus
2
2
~ ~ ~
a a a a a a
, , ,
, , ,
. Therefore, we can conclude
that in an orthogonal transformation, the magnitude vectors and the angle between them
are maintained, (see Figure 1.18).





NOTES ON CONTINUUM MECHANICS

52







Figure 1.18: Orthogonal transformation applied to vectors.
1.5.2.9 Positive Definite Tensor, Negative Definite Tensor and Semi-
Definite Tensors
A tensor is said to be positive definite when the following notations hold:
Tensorial notation Indicial notation Matrix notation
0

> x T x

0 >
j ij i
x T x

0 > x x T
T

(1.220)
for all vectors 0 x
,

. Conversely, a tensor is said to be negative definite when these notations


hold:
Tensorial notation Indicial notation Matrix notation
0

< x T x

0 <
j ij i
x T x

0 < x x T
T

(1.221)
for all vectors 0 x
,
,
.
A tensor is said to be semi-positive definite if 0

x T x for all vectors 0 x
,

. Similarly, we
define a semi-negative definite tensor when the following holds: 0

s x T x .
If
j i ij
x x T )

(

x x T x T x : c , then the derivative of c with respect to x
,
is given
by:
( )
i ik ki i ik j kj jk i ij j ik ij
k
j
i ij j
k
i
ij
k
x T T x T x T x T x T
x
x
x T x
x
x
T
x
- - -
o
o
-
o
o

o
o
c c
c
(1.222)
Thus, we can conclude that:
sym sym
T
x x
x T
x
2

o o
o

o
o

c c

(1.223)
Remember that it is also true that x T x x T x


sym
, therefore if the symmetric part of
a tensor is positive definite the tensor is too.
NOTE: As we will demonstrate later, the eigenvalues must be positive for T to be
positive definite. The proof is in the subsection Spectral Representation of Tensors.
Problem 1.25: Let F be an arbitrary second-order tensor. Show that the resulting tensors
F F C
T
and
T
F F b are symmetric tensors and semi-positive definite tensors. Also check in
what condition are C and b positive definite tensors.
Solution: Symmetry:
Q
a
,

b
,

0
0
a
,
~

b
,
~


b a b a
b b
a a
,
,
,
,
,
,
,
,

~
~
~
~

1 TENSORS

53
b F F F F F F b
C F F F F F F C




T T T T T T T
T T T T T T T
) ( ) (
) ( ) (

Thus, we have shown that F F C
T
and
T
F F b are symmetric tensors.
To prove that the tensors F F C
T
and
T
F F b are semi-positive definite tensors, we
start with the definition of a semi-positive definite tensor, i.e., a tensor A is semi-positive
definite if 0

x A x holds, for all 0 x
,

. Thus:

0

( )

( )

( )

(

) (

) (

2 2






x x
x x x x
x x x x x x x x
T
T T
T T T
F F
F F F F
F F F F F F F F

Or in indicial notation:
0 0
) )( ( ) )( (
) ( ) (
2 2



i ik i ki
j jk i ik j kj i ki
j jk ik i j ij i j kj ki i j ij i
F F
F F F F
F F b F F C
x x
x x x x
x x x x x x x x

Thus, we proved that F F C
T
and
T
F F b are semi-positive definite tensors. Note
that
2

x x x F C equals zero, when 0 x
,

, if 0 x
,

F . Furthermore, by definition
0 x
,

F with 0 x
,

if and only if 0 ) ( F det , (see Problem 1.24). Then, the tensors


F F C
T
and
T
F F b are positive definite if and only if 0 ) ( F det .
1.5.2.10 Additive Decomposition of Tensors
Given two arbitrary tensors S , 0 T , and a scalar c, we can represent S by means of the
following additive decomposition of tensors:
T S U U T S c c - - where (1.224)
Note that, depending on the value of c, we have an infinite number of possibilities for
representing S . But, if 0 ) ( ) (
T T
T U U T Tr Tr , the additive decomposition is unique.
From the relationship in (1.224), we can evaluate the value of c as follows:
) ( ) ( ) ( ) (
0
T T T T T T T
T T T U T T T S T U T T T S - -

Tr Tr Tr Tr c c c
_

(1.225)
) (
) (
T
T
T T
T S


Tr
Tr
c (1.226)
For example, let us suppose that 1 T . In this case c is evaluated as follows:
3
) (
) (
) (
) (
) (
) (
) ( S
1
S
1 1
1 S
T T
T S Tr
Tr
Tr
Tr
Tr
Tr
Tr

T
T
c (1.227)
We can then define U as:
dev
S 1
S
S T S U = - -
3
) ( Tr
c (1.228)
Thus:
dev sph dev
S S S 1
S
S - -
3
) ( Tr
(1.229)
NOTES ON CONTINUUM MECHANICS

54
NOTE: 1
S
S
3
) ( Tr

sph
is the spherical part of the tensor S , and 1
S
S S
3
) ( Tr
-
dev
is
known as a deviatoric tensor.
Now suppose that T is given by ) (
2
1
T
S S T - then c can be evaluated as follows:
[ ]
[ ]
1
) ( ) (
4
1
) (
2
1
) (
) (

- -
-

T T T
T T
T
T
S S S S
S S S
T T
T S
Tr
Tr
Tr
Tr
c
(1.230)
We can then define U as ) (
2
1
) (
2
1
T T
S S S S S T S U - - - - c . Then we obtain S
represented by the additive decomposition as follows:
skew sym T T
S S S S S S S - - - - ) (
2
1
) (
2
1
(1.231)
which is the same as the equation obtained in (1.150) in which we split the tensor into
symmetric and antisymmetric parts.
Problem 1.26: Find a fourth-order tensor P so that
dev
A A : P , where A is a second-
order tensor.
Solution: Taking into account the additive decomposition into spherical and deviatoric parts,
we obtain:
1
A
A A A 1
A
A A A
3
) (
3
) ( Tr Tr
- - -
dev dev dev sph

Referring to the definition of fourth-order unit tensors seen in (1.172), and (1.174), where
the relations 1 A A ) ( Tr : I and A A : I hold, we can now state:
A 1 1 A A A 1
A
A A : : : : |
.
|

\
|
- |
.
|

\
|
- - -
3
1
3
1
3
1
3
) (
I I I I I
Tr
dev

Therefore, we can conclude that:
1 1 -
3
1
I P

The tensor P is known as a fourth-order projection tensor, Holzapfel(2000).
1.5.3 Transformation Law of the Tensor Components
The tensor components depend on the coordinate system, so, if the coordinate system is
changed due to a rotation so do the tensor components. The tensor components between
these coordinate systems are interrelated to each other by the component transformation
law, which is defined below, (see Figure 1.19).
Consider a Cartesian coordinate system ( )
3 2 1
, , x x x formed by the orthogonal basis
( )
3 2 1

e e e , (see Figure 1.20). In this system, an arbitrary vector v


,
is represented by its
components as follows:
3 3 2 2 1 1

e e e e v v v v v - -
i i
,
(1.232)

1 TENSORS

55




















Figure 1.19: Transformation law of the tensor components.














Figure 1.20: Rotation of the Cartesian system.

2
x

1
x

3
x

1
x

2
x

3
x

1
c

1
,

1


2

e

3

e

1

e

3

e

1

e

2

e
COMPONENTS
Representation of a tensor in
a coordinate system
COORDINATE SYSTEM
I

COORDINATE SYSTEM
II
COMPONENT
TRANSFORMATION
LAW
TENSORS
Mathematical interpretation of physical concepts
(Independent of the coordinate system)
NOTES ON CONTINUUM MECHANICS

56
The components,
i
v , are represented in matrix form as:
(
(
(


3
2
1
) (
v
v
v
v v
i i
v
,

(1.233)
Now consider a new coordinate system ( )
3 2 1
, , x x x represented by the orthogonal basis
( )
3 2 1

e e e , (see Figure 1.20). In this new system, the vector v


,
is represented by
j j
e

v . As
mentioned before, a tensor is independent of the adopted system, so:
j j k k
e e v

v v
,

(1.234)
To obtain the components of a tensor in a given system one need only make the dot
product between the tensor and the system basis:
i i
i j j ki k
i j j i k k
e e e e
e e
e e e e
- -

)

(

(

3 3 2 2 1 1
v v v v
v v
v v
c (1.235)
Or in matrix form:
(
(
(

- -
- -
- -

(
(
(

3 3 3 2 2 1 1
2 3 3 2 2 1 1
1 3 3 2 2 1 1
3
2
1

)

(

)

(

)

(
e e e e
e e e e
e e e e
v v v
v v v
v v v
v
v
v
(1.236)
After restructuring, the previous equation looks like:
j i i j ij
a e e e e
e e e e e e
e e e e e e
e e e e e e




3
2
1
3 3 3 2 3 1
2 3 2 2 2 1
1 3 1 2 1 1
3
2
1




.
(
(
(

(
(
(

(
(
(

v
v
v
v
v
v
(1.237)
Or in indicial notation:
j ij i
a v v
(1.238)
where we have introduced the transformation matrix
ij
a = A as:
(
(
(

(
(
(







=
3 3 2 3 1 3
3 2 2 2 1 2
3 1 2 1 1 1
3 3 3 2 3 1
2 3 2 2 2 1
1 3 1 2 1 1






e e e e e e
e e e e e e
e e e e e e
e e e e e e
e e e e e e
e e e e e e
ij
a A
(
(
(

=
33 32 31
23 22 21
13 12 11
a a a
a a a
a a a
a
ij
A
(1.239)
The matrix ( A) is not symmetric, i.e.
T
A A . With reference to the scalar product
( ) ( )
j i j i j i j i
x x x x , cos , cos

e e e e , (see equation (1.4)), the relationship in (1.237) is
expressed by means of the direction cosines as:
1 TENSORS

57

( ) ( ) ( )
( ) ( ) ( )
( ) ( ) ( )

v
v
(
(
(

(
(
(

(
(
(

3
2
1
3 3 2 3 1 3
3 2 2 2 1 2
3 1 2 1 1 1
3
2
1

, cos , cos , cos
, cos , cos , cos
, cos , cos , cos
v
v
v
v
v
v
_
A
x x x x x x
x x x x x x
x x x x x x

v v A

(1.240)
The direction cosines of a vector are those of the angles between the vector and the three
coordinate axes. According to Figure 1.20 we can verify that ( )
1 1 1
, cos cos x x c ,
( )
2 1 1
, cos cos x x , and ( )
3 1 1
, cos cos x x .
In the equation (1.235) we have projected the vector onto
i
e

. Now, we can project the


vector onto
i
e

:
v v



T
ji j i
ji j ki k
i j j i k k
a
a
A
v v
v v
v v
c
e e e e

(1.241)
Therefore, it is also true that:
j ji i
a e e

(1.242)
The inverse relationship of equation (1.240) is obtained as follows:
v v v v
- - - 1 1 1
A A A A (1.243)
and by comparing the equations (1.243) with (1.241) we can conclude that the matrix A is
an orthogonal matrix, i.e.:
ij kj ki
T T
a a c
-

notation
Indicial
1
1 A A A A
(1.244)
Second-order tensor
Consider a coordinate system formed by the orthogonal basis
i
e

then, how the basis


changes from the
i
e

system to a new one represented by the orthogonal basis


i
e

. This is
illustrated in transformation law as
i ik k
a e e

, which allow us to represent a second-order
tensor T as follows:
j i ij
j i jl ik kl
j jl i ik kl
l k kl
a a
a a
e e
e e
e e
e e T








T
T
T
T
(1.245)
Then, the transformation law of the components between systems for a second-order
tensor is given by:
T
jl kl ik jl ik kl ij
a a a a A T A T
form Matrix
T T T (1.246)
Third-order tensor
A third-order tensor ( S ) can be shown in two systems represented by orthogonal bases
i
e


and
i
e

as follows:
NOTES ON CONTINUUM MECHANICS

58
k j i ijk
k j i kn jm il lmn
k kn j jm i il lmn
n m l lmn
a a a
a a a
e e e
e e e
e e e
e e e S








S
S
S
S
(1.247)
In conclusion the components of the third-order tensor in the new basis (
i
e

) are:
kn jm il lmn ijk
a a a S S
(1.248)
The following table summarizes the transformation law of the components according to
the tensor rank:
rank
from ( ) ( )
3 2 1 3 2 1
, , , , x x x x x x
to
from ( ) ( )
3 2 1 3 2 1
, , , , x x x x x x
to

0 (scalar) / / / /
1 (vector)
j ij i
S a S
j ji i
S a S
2
kl jl ik ij
S a a S
kl lj ki ij
S a a S
3
lmn kn jm il ijk
S a a a S
lmn nk mj li ijk
S a a a S
4
mnpq lq kp jn im ijkl
S a a a a S
mnpq ql pk nj mi ijkl
S a a a a S
(1.249)

Problem 1.27: Obtain the components of T , given by the transformation:
T
A T A T
where the components of T and A are shown, respectively, as
ij
T and
ij
a . Afterwards,
given that
ij
a are the components of the transformation matrix, represent graphically the
components of the tensors T and T on both systems.
Solution: The expression
T
A T A T in symbolic notation is given by:
)

(
)

(
)

( )

( )

( )

(
k r kq pq rp
k r ql sp kl pq rs
k l kl q p pq s r rs b a ab
a a
a a
a a
e e
e e
e e e e e e e e



T
T
T T
c c
To obtain the components of T one only need make the double scalar product with the
basis )

(
j i
e e , the result of which is:
jq pq ip ij
kj ri kq pq rp bj ai ab
j i k r kq pq rp j i b a ab
a a
a a
a a
T T
T T
T T



c c c c
)

( )

( )

( )

( e e e e e e e e : :

The above eqaution is shown in matrix notation as:
T inverse T - -
A T A T A T A T
1

Since A is an orthogonal matrix, it holds that
1 -
A A
T
. Thus, A T A T
T
. The
graphical representation of the tensor components in both systems can be seen in Figure
1.21.




1 TENSORS

59




















Figure 1.21: Transformation law of the second-order tensor components.
Problem 1.28: Let T be a symmetric second-order tensor and
T
I ,
T
I I ,
T
I I I be scalars,
where:
{ ) ( ; ) (
2
1
; ) (
2 2
T T T
T T T T
det Tr T Tr - I I I I I I I
ii

Show that
T
I ,
T
I I ,
T
I I I are invariant with a change of basis.
Solution:
a) Taking into account the transformation law for the second-order tensor components
given in (1.249), i.e.
kl jl ik ij
a a T T or in matrix form
T
A T A T . Then,
ii
T is:
T
I a a
kk kl kl kl il ik ii
T T T T c
Hence we have proved that
T
I is independent of the adopted system.
b) To prove that
T
I I is an invariant, one only need show that ) (
2
T Tr is one also, since
2
T
I
is already an invariant.
) ( ) (
) )( ( ) ( ) (
2
2
T T T T T
T T T T T
Tr Tr
T T
T T
T T T T Tr Tr

:
:
pl pl
pq kl
lq
jq jl
kp
ip ik
pq jq ip kl jl ik ij ij
a a a a
a a a a
_ _
c c

c)
) ( ) ( ) ( ) ( ) ( ) ( ) (
1 1
T A T A A T A T det det det det det det det

_ _
T T
T
Consider now four sets of coordinate systems, represented by ( )
3 2 1
, , x x x , ( )
3 2 1
, , x x x ,
( )
3 2 1
, , x x x and ( )
3 2 1
, , x x x , (see Figure 1.22), and consider also the following
transformation matrices:
A: Transformation matrix from ( )
3 2 1
, , x x x to ( )
3 2 1
, , x x x ;

1
x
2
x

3
x

11
T

21
T

31
T

12
T

32
T

22
T

33
T

13
T

23
T

3
x

2
x

1
x

11
T

21
T

31
T

12
T

32
T

22
T

33
T

13
T

23
T

T
A T A T
A T A T
T


3
x

1
x

2
x
NOTES ON CONTINUUM MECHANICS

60
B : Transformation matrix from ( )
3 2 1
, , x x x to ( )
3 2 1
, , x x x ;
C : Transformation matrix from ( )
3 2 1
, , x x x to ( )
3 2 1
, , x x x .











Figure 1.22: Transformations matrices between several systems.
If we consider a v column matrix made up of components of v
,
in the coordinate system
( )
3 2 1
, , x x x , the components of this vector in the system ( )
3 2 1
, , x x x are given by:
v v A (1.250)
and the inverse transformation of relation (1.250) is:
v v
T
A
(1.251)
Now, starting with the system ( )
3 2 1
, , x x x , the components of the vector in the system
( )
3 2 1
, , x x x are given by:
v v B (1.252)
and the inverse transformation is:
v v
T
B
(1.253)
By substituting the equation (1.250) into (1.252) we obtain:
v v BA (1.254)
The resulting matrix BA is also an orthogonal matrix, and shows the transformation
matrix from ( )
3 2 1
, , x x x to ( )
3 2 1
, , x x x , (see Figure 1.22). The inverse form of (1.254) is
evaluated by substituting (1.253) into (1.251), the result of which is:
v v
T T
B A
(1.255)
This equation could have been obtained by using equation (1.254), i.e.:
( ) ( ) ( ) ( ) v v v v v v
- - - - - T T
B A B A BA BA BA BA
1 1 1 1 1

(1.256)
Then, it is easy to find the components of the vector in the coordinate system ( )
3 2 1
, , x x x ,
(see Figure 1.22):
v v v v
T T T
C B A CBA
form inverse
(1.257)

X
X X
B

T
B B
-1

A

T
A A
-1


T T T
) (BA B A
BA
X
T T T T
C B A CBA ) (
CBA

T
C C
-1

C
1 TENSORS

61
1.5.3.1 Component Transformation Law in Two Dimensions (2D)
Now, consider two sets of coordinate systems, shown in Figure 1.23.







Figure 1.23: Transformation of a coordinate system in 2D.
The transformation matrix from ( y x - ) to ( y x - ) is given by direction cosines, (see
Figure 1.23), as:
(
(
(

(
(
(



1 0 0
0 ) cos( ) cos(
0 ) cos( ) cos(
1 0 0
0
0
22 21
12 11
y y x y
y x x x
a a
a a
c c
c c
A
(1.258)
By using trigonometric identities we can deduce that:
) cos( ) cos( ) cos( c c c c c
y y x x y y x x
, ) sin(
2
cos ) cos( c c c |
.
|

\
|
-

r
y x
,
) sin(
2
cos ) cos( c c c - |
.
|

\
|
-

r
x y

(1.259)
Thus, the transformation matrix in 2D is dependent on a single parameter, c, i.e.:
(

) cos( ) sin(
) sin( ) cos(
c c
c c
A (1.260)
Another way to prove (1.260) is by considering the vector position of the point P in both
systems, (see Figure 1.24).
Moreover, in view of Figure 1.24, said coordinates are interrelated as shown below:

|
.
|

\
|
-
r
- -
-
, ,
, c
2
cos ) cos(
) cos( ) cos(
P P P
P P P
y x y
y x x

- |
.
|

\
|
-
r
-
|
.
|

\
|
-
r
-
) cos(
2
cos
2
cos ) cos(
c c
c c
P P P
P P P
y x y
y x x

- -
-

) cos( ) sin(
) sin( ) cos(
c c
c c
P P P
P P P
y x y
y x x

(1.261)
Or in matrix form:
(

-
P
P
P
P tion transforma
Inverse
P
P
P
P
y
x
y
x
y
x
y
x

) cos( ) sin(
) sin( ) cos(

) cos( ) sin(
) sin( ) cos(
1

c c
c c
c c
c c
(1.262)
Since
T
A A
-1
, it is true that:

y y
c

y x
c
c c
x x

x
y
x
y

x y
c

c c
c c
c c
-
r

-
r

2
2
x y
y x
y y

NOTES ON CONTINUUM MECHANICS

62
(

P
P
P
P
y
x
y
x
) cos( ) sin(
) sin( ) cos(
c c
c c
(1.263)













Figure 1.24: Transformation of a coordinate system in 2D.





Problem 1.29: Find the transformation matrix between the systems: z y x , , and z y x , , .
These systems are represented in Figure 1.25.















Figure 1.25: Rotation.

y y
x
y
z z
x
z z
y
x
x
,
c
c


c
P
x
P
y
,
x
y
x
y

P
y
r
,

P

P
x

)
c
o
s
(
o
P
x
)
c
o
s
(
o
P
x

c

)
s
i
n
(
o
P
y
)
s
i
n
(
o
P
y


)
s
i
n
(
o
P
x
)
s
i
n
(
o
P
x


)
c
o
s
(
o
P
y
)
c
o
s
(
o
P
y


P
y

P
y


1 TENSORS

63
Solution: The coordinate system z y x , , can be obtained by different combinations of
rotations as follows:

+ Rotation along the z -axis











+ Rotation along the y -axis















+ Rotation along the z -axis

















from z y x , , to z y x , ,

(
(
(

-
1 0 0
0 cos sin
0 sin cos


C
with 360 0 s s
y y
x
y
z z
x
z z
y
x
x
,
o
o


from z y x , , to z y x , ,

(
(
(

-
1 0 0
0 cos sin
0 sin cos
c c
c c
A
with 360 0 s s c
x
y
z z
x
y
c
c
from z y x , , to z y x , ,

(
(
(

, ,
, ,
cos 0 sin
0 1 0
sin 0 cos
B
with 180 0 s s ,
y y
x
y
z z
x
z
x
,
c
c
,
x
z z
z
x
NOTES ON CONTINUUM MECHANICS

64
The transformation matrix from ( z y x , , ) to ( z y x , , ), (see Figure 1.22), is given by:
CBA D
After multiplying the matrices, we obtain:
(
(
(

- - - -
- - -

, , c , c
, c , c c , c
, c , c c , c
cos sin sin sin cos
sin sin ) cos cos sin cos sin ( ) cos sin sin cos cos (
cos sin ) sin cos cos cos (sin ) sin sin cos cos (cos
D
The angles , c , , are known as Euler angles and were introduced by Leonhard Euler to
describe the orientation of a rigid body motion.

Problem 1.30: Let T be a second-order tensor whose components in the Cartesian system
( )
3 2 1
, , x x x are given by:
( )
(
(
(

-
-

1 0 0
0 3 1
0 1 3
T
ij ij
T T
Given that the transformation matrix between two systems, ( )
3 2 1
, , x x x - ( )
3 2 1
, , x x x , is:
(
(
(
(
(
(

0
2
2
2
2
0
2
2
2
2
1 0 0
A
Obtain the tensor components
ij
T in the new coordinate system ( )
3 2 1
, , x x x .
Solution: As defined in equation (1.249), the transformation law for second-order tensor
components is:
kl jl ik ij
a a T T
To enable the previous calculation to be carried out in matrix form we use:
[ ] [ ] [ ]
T
j l l k k i ij
a a

T T
Thus
T
A T A T
(
(
(
(
(
(
(

-
(
(
(

-
-
(
(
(
(
(
(

-

0 0 1
2
2
2
2
0
2
2
2
2
0
1 0 0
0 3 1
0 1 3
0
2
2
2
2
0
2
2
2
2
1 0 0
T
On carrying out the operation of the previous matrices we now have:
(
(
(


4 0 0
0 2 0
0 0 1
T
NOTE: As we can verify in the above example, the components of the tensor T , in the
new basis, have one particular feature, i.e. the off-diagonal terms are equal to zero. The
question now is: Given an arbitrary tensor T , is there a transformation which results in the
1 TENSORS

65
off-diagonal terms being zero? This type of problem is called the eigenvalue and eigenvector
problem.
1.5.4 Eigenvalue and Eigenvector Problem
As we have seen, the scalar product between a second-order tensor T and a vector (or unit
vector n

) leads to a vector. In other words, projecting a second-order tensor onto a


certain direction results in a vector that does not necessarily have the same direction as n

,
(see Figure 1.26(a)).
The aim of the eigenvalue and eigenvector problem is to find a direction n

, in such a way
that the resulting vector, n T t
n

(

,
, coincides with it, (see Figure 1.26 (b)).














Figure 1.26: Projecting a tensor onto a direction.
Let T be a second-order tensor. A vector n

is said to be eigenvector of T if there is a scalar


/ , called the eigenvalue, so that:
n n T

/
(1.264)
The equation (1.264) can be rearranged in indicial notation as:
( ) ( ) 0 n 1 T
,
/ - / -
/ -
/




notation
Tensorial
i j ij ij
i i j ij
i j ij
0 n T
0 n n T
n n T
c

(1.265)
The previous set of homogeneous equations only have nontrivial solution, i.e. 0 n
,

, if and
only if:
0 ; 0 ) ( / - / -
ij ij
c T det 1 T
(1.266)
The determinant (1.266) is called the characteristic determinant of the tensor T , explicitly given
by:
n T t
n

(
,

n



1
x

2
x

3
x
n n T t
n

)

(
/
,

n

- principal direction of T
/ - eigenvalue of T associated
with the direction n

.
b) Principal direction. a) Projection of T onto an
arbitrary plane.
NOTES ON CONTINUUM MECHANICS

66
0
33 32 31
23 22 21
13 12 11

/ -
/ -
/ -
T T T
T T T
T T T

(1.267)
Developing this determinant, we obtain the characteristic polynomial, which is shown by a
cubic equation in / :
0
2 3
- / - / - /
T T T
I I I I I I (1.268)
where
T
I ,
T
I I ,
T
I I I are the principal invariants of T , and are defined in components terms
as:
ii
I T Tr ) (T
T

[ ]
( ) ( ) [ ] {
( ) [ ] {
{
{ [ ] ) cof( Tr M T T T T
T T T T
Tr T T T T
T T Tr T Tr T Tr
Tr Tr
T
e e
e e e e e e e e
T T
T
-
-
-
-
-

ii ji ij kk ii
il jk kl ij kk ii
l i jk kl ij kl kl ij ij
l k kl j i ij l k kl j i ij
I I
2
1
2
1

2
1

)

( )

(
2
1
) ( ) (
2
1
2 2
c c
c c c
3 2 1
) (
k j i ijk ij
I I I T T T T det T
T

(1.269)
where
ii
M is the matrix trace defined in equation (1.210),
33 22 11
M M M M - -
ii
. More
explicitly the invariants are given by:
( ) ( ) ( )
22 31 32 21 13 23 31 33 21 12 23 32 33 22 11
21 12 22 11 31 13 33 11 32 23 33 22
22 21
12 11
33 31
13 11
33 32
23 22
33 22 11
T T T T T T T T T T T T T T T
T T T T T T T T T T T T
T T
T T
T T
T T
T T
T T
T T T
- - - - -
- - - - -
- -
- -
T
T
T
I I I
I I
I

(1.270)
If T is a symmetric tensor, the principal invariants are summarized as follows:
22
2
13 11
2
23 33
2
12 23 12 13 23 13 12 33 22 11
2
23
2
13
2
12 33 22 33 11 22 11
33 22 11
T T T T T T T T T T T T T T T
T T T T T T T T T
T T T
- - - - -
- - - - -
- -
T
T
T
I I I
I I
I

(1.271)
The eigenvalues,
3 2 1
, , / / / , are found by solving the characteristic polynomial (1.268).
Once the eigenvalues are evaluated, the eigenvectors are found by applying equation
(1.265), i.e.
i j ij ij
0 n T / -
) 1 (
1

) ( c ,
i j ij ij
0 n T / -
) 2 (
2

) ( c ,
i j ij ij
0 n T / -
) 3 (
3

) ( c . These
eigenvectors constitute a new space denoted as the principal space.
If T is a symmetric tensor, the principal space is defined by an orthonormal basis and all
eigenvalues are real numbers. If the three eigenvalues are different,
3 2 1
/ / / , the three
principal directions are unique. If two of them are equal, e.g.
3 2 1
/ / / , we can state that
the principal direction,
) 3 (

n , associated with the eigenvalue


3
/ , is unique, and, any
direction defined in the plane normal to
) 3 (

n is a principal direction, and othorgonality is



jk
c
1 TENSORS

67
the only constraint to determining
) 1 (

n and
) 2 (

n . If
3 2 1
/ / / , any direction is principal.
A tensor that has three equal eigenvalues is called a Spherical Tensor, (see Appendix A-The
Tensor ellipsoid).
The T -components in the principal space are only made up of normal components, i.e.:
(
(
(

(
(
(

/
/
/

3
2
1
2
2
1
0 0
0 0
0 0
0 0
0 0
0 0
T
T
T
T
ij

(1.272)
Therefore, the principal invariants can also be evaluated by:
3 2 1
T T T - -
T
I ,
3 1 3 2 2 1
T T T T T T - -
T
I I ,
3 2 1
T T T
T
I I I (1.273)
whose values must match the values obtained in (1.270), since they are invariant with a
change of basis.
If T is a spherical tensor, i.e. T T T T
3 2 1
, it holds that
T T
I I I 3
2
,
3
T
T
I I I .
Let W be an antisymmetric tensor. The principal invariants of W are given by:
[ ]
0
0
0
0
0
0
0
2
) (
) ( ) (
2
1
0 ) (
2
12 12 13 13 23 23
12
12
13
13
23
23
2
2 2

- -
-
-
-
-
-

-
-

W
W
W
W
W W
W
I I I
I I
I
.
W W W W W W
W
W
W
W
W
W
Tr
Tr Tr
Tr
(1.274)
where
2
12
2
13
2
23
2
2
W W W - - w w w
, , ,
. as defined in (1.132). Then, the characteristic
equation for an antisymmetric tensor is reduced to:
( ) 0 0 0
2 2 2 3 2 3
- / / / - / - / - / - / . .
W W W
I I I I I I (1.275)
In this case, one eigenvalue is real and equal to zero and the others are imaginary roots:
i 1 0 0
) 2 , 1 (
2 2 2 2
. . . . - / - / - / (1.276)
1.5.4.1 The Orthogonality of the Eigenvectors
Consider a symmetric second-order tensor T . By the definition of eigenvalues, given in
(1.264), if
1
/ ,
2
/ ,
3
/ are the eigenvalues of T , then it follows that:
) 3 (
3
) 3 ( ) 2 (
2
) 2 ( ) 1 (
1
) 1 (

;

n n T n n T n n T / / / (1.277)
Applying the dot product between
) 2 (

n and
) 1 (
1
) 1 (

n n T / , and the dot product between
) 1 (

n and
) 2 (
2
) 2 (

n n T / we obtain:
) 2 ( ) 1 (
2
) 2 ( ) 1 (
) 1 ( ) 2 (
1
) 1 ( ) 2 (


n n n T n
n n n T n


/
/

(1.278)
Since T is symmetric, it holds that
) 2 ( ) 1 ( ) 1 ( ) 2 (

n T n n T n , so:
) 1 ( ) 2 (
2
) 2 ( ) 1 (
2
) 1 ( ) 2 (
1

n n n n n n / / / (1.279)
NOTES ON CONTINUUM MECHANICS

68
( ) 0

) 2 ( ) 1 (
2 1
/ - / n n (1.280)
To satisfy the equation (1.280), with 0
2 1
/ / , the following must be true:
0

) 2 ( ) 1 (
n n (1.281)
Similarly, it is possible to show that 0

) 3 ( ) 1 (
n n and 0

) 3 ( ) 2 (
n n and then we can
conclude that the eigenvectors are mutually orthogonal, and constitute an orthogonal basis,
(see Figure 1.27), where the transformation matrix between systems is:
(
(
(

(
(
(

) 3 (
3
) 3 (
2
) 3 (
1
) 2 (
3
) 2 (
2
) 2 (
1
) 1 (
3
) 1 (
2
) 1 (
1
) 3 (
) 2 (
) 1 (


n n n
n n n
n n n
n
n
n
A
(1.282)
















Figure 1.27: Diagonalization.
Problem 1.31: Show that the following relations are invariants:
4
3
4
2
4
1
3
3
3
2
3
1
2
3
2
2
2
1
; ; C C C C C C C C C - - - - - -
where
1
C ,
2
C ,
3
C are the eigenvalues of the second-order tensor C .
Solution: Any combination of invariants is also an invariant, so, on this basis, we can try to
express the above expressions in terms of their principal invariants.
( ) ( )
C C
C
C
I I I C C C C C C C C C C C C C C C I
I I
2 2
2 2
3
2
2
2
1 3 2 3 1 2 1
2
3
2
2
2
1
2
3 2 1
2
- - - - - - - - - -
_

So, we have proved that
2
3
2
2
2
1
C C C - - is an invariant. Similarly, we can obtain:
2
C C C
2
C C C
C C C C
I I I I I I I I I I C C C
I I I I I I I C C C
2 4 4
3 3
4 4
3
4
2
4
1
3 3
3
3
2
3
1
- - - - -
- - - -


1
x

2
x

3
x

1
x

2
x
diagonalization

3
x

1
x

2
x

3
x

22
T

12
T
12
T

11
T

23
T

23
T

13
T

13
T

33
T

2
T

1
T

3
T

) 1 (

n

) 2 (

n
) 3 (

n
Principal Space

T
A T A T
A T A T
T

1 TENSORS

69

Problem 1.32: Let Q be a proper orthogonal tensor, and E be an arbitrary second-order
tensor. Show that the eigenvalues of E do not change with the following orthogonal
transformation:
T
Q E Q E
*

Solution: We can prove this as follows:
( )
( )
( )
( ) [ ]
( ) ( ) ( )
( )
( )
( )
( )
( ) [ ]
( ) ( ) ( )
( )
kp kp
jp kp kp ik
jp kp kp ik
kp jp ik jp kp ik
ij jp kp ik
ij ij
T
T
T T
T
c
c
c
c
c
c
/ -
/ -
/ -
/ -
/ -
/ -
/ -
/ -
/ -
/ -
/ -
/ -



E det
Q det E det Q det
Q E Q det
Q Q Q E Q det
Q E Q det
E det
det
det det det
det
det
det
det

0

0
*
1 1
*
1 E
Q 1 E Q
Q 1 E Q
Q 1 Q Q E Q
1 Q E Q
1 E
_ _

Thus, we have proved that E and
*
E have the same eigenvalues.
1.5.4.2 Solution of the Cubic Equation
Let T be a symmetric second-order tensor. The roots of the characteristic equation
( 0
2 3
- / - / - /
T T T
I I I I I I ) are all real numbers, and are expressed as:
3 3
4
3
cos 2
3 3
2
3
cos 2
3 3
cos 2
3
2
1
T
T
T
I
S
I
S
I
S
-
(

|
.
|

\
| r
- /
-
(

|
.
|

\
| r
- /
-
(

|
.
|

\
|
/
c
c
c
(1.283)
where
|
.
|

\
|
- - -
-

T
Q R
T
I
I I I
I I I
Q
R
S
I I I
R
2
arccos ;
27
;
27
2
3
;
3
;
3
3
3 3 2
c
T
T
T T T T

where c is in radians.
(1.284)
By restructuring the solution (1.283) in matrix form, we obtain:
_
_
part Deviatoric
part Spherical
3
2
1
3
4
3
cos 0 0
0
3
2
3
cos 0
0 0
3
cos
2
1 0 0
0 1 0
0 0 1
3
0 0
0 0
0 0
(
(
(
(
(
(
(

|
.
|

\
| r
-
|
.
|

\
| r
-
|
.
|

\
|
-
(
(
(

(
(
(

/
/
/
c
c
c
S
I
T

(1.285)
where we clearly distinguish the spherical and the deviatoric part of the tensor in the
principal space. Note that, if T is a spherical tensor the following relationship holds
T T
I I I 3
2
, then 0 S .
NOTES ON CONTINUUM MECHANICS

70

Problem 1.33: Find the principal values and directions of the second-order tensor T ,
where the Cartesian components of T are:
( )
(
(
(

-
-

1 0 0
0 3 1
0 1 3
T
ij ij
T T

Solution: We need to find nontrivial solutions for ( )
i j ij ij
0

/ - n T c , which are constrained


by 1


j j
n n (unit vector). As we have seen, the nontrivial solution requires that:
0 / -
ij ij
c T
Explicitly, the above equation is:
0
1 0 0
0 3 1
0 1 3
33 32 31
23 22 21
13 12 11

/ -
/ - -
- / -

/ -
/ -
/ -
T T T
T T T
T T T

Developing the above determinant, we can obtain the cubic equation:
[ ]
0 8 14 7
0 1 ) 3 ( ) 1 (
2 3
2
- / - / - /
- / - / -

We could have obtained the characteristic equation directly in terms of invariants:
7 ) (
33 22 11
- - T T T T T Tr
ii ij
I
T

( ) 14
2
1
22 21
12 11
33 31
13 11
33 32
23 22
- - -
T T
T T
T T
T T
T T
T T
T T T T
ij ij jj ii
I I
T

8
3 2 1

k j i ijk ij
I I I T T T T
T

Then, using the equation in (1.268), the characteristic equation is:
0 8 14 7 0
2 3 2 3
- / - / - / - / - / - /
T T T
I I I I I I
On solving the cubic equation we obtain three real roots, namely:
4 ; 2 ; 1
3 2 1
/ / /
We can also verify that:
8
14 1 4 4 2 2 1
7 4 2 1
3 2 1
1 3 3 2 2 1
3 2 1
/ / /
- - / / - / / - / /
- - / - / - /
T
T
T
I I I
I I
I

Thus, we can see that the invariants are the same as those evaluated previously.

Principal directions:
Each eigenvalue,
i
/ , is associated with a corresponding eigenvector,
) (

i
n . We can use the
equation in (1.265), i.e.
i j ij ij
0 n T / -

) ( c , to obtain the principal directions.


1
1
/
(
(
(

(
(
(

(
(
(

-
- -
- -

(
(
(

(
(
(

/ -
/ - -
- / -
0
0
0
1 1 0 0
0 1 3 1
0 1 1 3
1 0 0
0 3 1
0 1 3
3
2
1
3
2
1
1
1
1
n
n
n
n
n
n

These become the following system of equations:



1 TENSORS

71
1
0 0
0
0 2
0 2
2
3
2
2
2
1
3
2 1
2 1
2 1
- -


)
`

- -
-
n n n n n
n
n n
n n
n n
i i

Then we can conclude that: [ ] 1 0 0

1
) 1 (
1
/
i
n .
NOTE: This solution could have been directly determined by the specific features of the
T matrix. As the terms 0
32 31 23 13
T T T T imply that 1
33
T is already a principal
value, then, consequently, the original direction is a principal direction.
2
2
/
(
(
(

(
(
(

(
(
(

-
- -
- -

(
(
(

(
(
(

/ -
/ - -
- / -
0
0
0
2 1 0 0
0 2 3 1
0 1 2 3
1 0 0
0 3 1
0 1 3
3
2
1
3
2
1
2
2
2
n
n
n
n
n
n

-
- -
-
0
0
0
3
2 1
2 1 2 1
n
n n
n n n n

The first two equations are linearly dependent, after which we need an additional equation:
2
1
1 2 1
1
2
1
2
3
2
2
2
1
- - n n n n n n n
i i

Thus:
(
(

/ 0
2
1
2
1

2
) 2 (
2 i
n
4
3
/
(
(
(

(
(
(

(
(
(

-
- -
- -

(
(
(

(
(
(

/ -
/ - -
- / -
0
0
0
4 1 0 0
0 4 3 1
0 1 4 3
1 0 0
0 3 1
0 1 3
3
2
1
3
2
1
3
3
3
n
n
n
n
n
n

2
1
1 2 1
0 3
0
0
2
2
2
2
3
2
2
2
1
3
2 1
2 1
2 1
- -

-
-
)
`

- -
- -
n n n n n n n
n
n n
n n
n n
i i

Then:
(
(

/ 0
2
1
2
1

4
) 3 (
3
+
i
n
Afterwards, we summarize the eigenvalues and eigenvectors of T :
[ ]
(
(

/
(
(

/
/
0
2
1
2
1

4
0
2
1
2
1

2
1 0 0

1
) 3 (
3
) 2 (
2
) 1 (
1
+
i
i
i
n
n
n

NOTES ON CONTINUUM MECHANICS

72
NOTE: The tensor components of this problem are the same as those used in Problem
1.30. Additionally, we can verify that the eigenvectors make up the transformation matrix,
A, between the original system, ( )
3 2 1
, , x x x , and the principal space, ( )
3 2 1
, , x x x , (see
Problem 1.30).
1.5.5 Spectral Representation of Tensors
Based on the solution of the equation in (1.268), if T is a symmetric second-order tensor
there are three real eigenvalues:
1
T ,
2
T ,
3
T each of which is associated with an
eigenvector, i.e.:
[ ]
[ ]
[ ]




) 3 (
3
) 3 (
2
) 3 (
1
) 3 (
3
) 2 (
3
) 2 (
2
) 2 (
1
) 2 (
2
) 1 (
3
) 1 (
2
) 1 (
1
) 1 (
1
n n n n T
n n n n T
n n n n T



i
i
i
(1.286)
The principal space is formed by the orthogonal basis
) 3 ( ) 2 ( ) 1 (

n n n , and the tensor


components are represented by their eigenvalues as:
(
(
(


3
2
1
0 0
0 0
0 0
T
T
T
T T
ij

(1.287)
With reference to the fact that eigenvectors form a transformation matrix, A, so that:
T
A T A T (1.288)
Since
T
A A
-1
, the inverse form is:
A T A T
T

(1.289)
where
(
(
(

(
(
(

) 3 (
3
) 3 (
2
) 3 (
1
) 2 (
3
) 2 (
2
) 2 (
1
) 1 (
3
) 1 (
2
) 1 (
1
) 3 (
) 2 (
) 1 (


n n n
n n n
n n n
n
n
n
A
(1.290)
Explicitly, the relation in (1.289) is given by:
A A A A A A
A A
(
(
(

(
(
(

(
(
(

(
(
(

(
(
(

(
(
(

(
(
(

(
(
(

- -
1 0 0
0 0 0
0 0 0
0 0 0
0 1 0
0 0 0
0 0 0
0 0 0
0 0 1
0 0
0 0
0 0




0 0
0 0
0 0




3 2 1
3
2
1
) 3 (
3
) 3 (
2
) 3 (
1
) 2 (
3
) 2 (
2
) 2 (
1
) 1 (
3
) 1 (
2
) 1 (
1
3
2
1
) 3 (
3
) 2 (
3
) 1 (
3
) 3 (
2
) 2 (
2
) 1 (
2
) 3 (
1
) 2 (
1
) 1 (
1
33 23 13
23 22 12
13 12 11
T T T
T
T T T
T
T
T
n n n
n n n
n n n
T
T
T
n n n
n n n
n n n
T T T
T T T
T T T

(1.291)
Whereas:
1 TENSORS

73
) 3 ( ) 3 (
) 3 (
3
) 3 (
3
) 3 (
2
) 3 (
3
) 3 (
1
) 3 (
3
) 3 (
3
) 3 (
2
) 3 (
2
) 3 (
2
) 3 (
1
) 3 (
2
) 3 (
3
) 3 (
1
) 3 (
2
) 3 (
1
) 3 (
1
) 3 (
1
) 2 ( ) 2 (
) 2 (
3
) 2 (
3
) 2 (
2
) 2 (
3
) 2 (
1
) 2 (
3
) 2 (
3
) 2 (
2
) 2 (
2
) 2 (
2
) 2 (
1
) 2 (
2
) 2 (
3
) 2 (
1
) 2 (
2
) 2 (
1
) 2 (
1
) 2 (
1
) 1 ( ) 1 (
) 1 (
3
) 1 (
3
) 1 (
2
) 1 (
3
) 1 (
1
) 1 (
3
) 1 (
3
) 1 (
2
) 1 (
2
) 1 (
2
) 1 (
1
) 1 (
2
) 1 (
3
) 1 (
1
) 1 (
2
) 1 (
1
) 1 (
1
) 1 (
1




1 0 0
0 0 0
0 0 0




0 0 0
0 1 0
0 0 0




0 0 0
0 0 0
0 0 1
j i
T
j i
T
j i
T
n n
n n n n n n
n n n n n n
n n n n n n
n n
n n n n n n
n n n n n n
n n n n n n
n n
n n n n n n
n n n n n n
n n n n n n

(
(
(

(
(
(

(
(
(

(
(
(

(
(
(

(
(
(

A A
A A
A A

(1.292)
Then, it is possible to represent the components of a second-order tensor in function of
their eigenvalues and eigenvectors (spectral representation) as:
) 3 ( ) 3 (
3
) 2 ( ) 2 (
2
) 1 ( ) 1 (
1

j i j i j i ij
n n T n n T n n T T - - (1.293)
As we can see, the tensor is represented as a linear combination of dyads and the above
representation in tensorial notation becomes:
) 3 ( ) 3 (
3
) 2 ( ) 2 (
2
) 1 ( ) 1 (
1

n n n n n n T - - T T T (1.294)
or:


3
1
) ( ) (

a
a a
a
n n T T
Spectral representation of a
second-order tensor
(1.295)
which is the spectral representation of the tensor. Note that, in the above equation we have to
resort to the summation symbol, because the dummy index appears thrice in the
expression.
NOTE: The spectral representation in (1.295) could easily have been obtained from the
definition of the second-order unit tensor, given in (1.168), i.e.
i i
n n 1

, which can also
be represented by means of the summation symbol as


3
1
) ( ) (

a
a a
n n 1 . Then, it follows
that:



|
|
.
|

\
|

3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (

a
a a
a
a
a a
a
a a
n n n n T n n T 1 T T T (1.296)
where we have used the definition of eigenvalue and eigenvector
) ( ) (

a
a
a
n n T T .
We now consider the orthogonal tensor R . The orthogonal transformation applied to the
unit vector N

leads to the unit vector n

, i.e.

N R n . Therefore, it is also possible to


represent the orthogonal tensor R as follows:



|
|
.
|

\
|

3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (


a
a a
a
a a
a
a a
N n N N R N N R 1 R R (1.297)
The spectral representation is very useful for making algebraic operations with tensors. For
example, tensor power in the principal space can be expressed as:
NOTES ON CONTINUUM MECHANICS

74
( )
(
(
(

n
n
n
ij
n
3
2
1
0 0
0 0
0 0
T
T
T
T
(1.298)
So, the spectral representation of
n
T is given by:


3
1
) ( ) (

a
a a n
a
n
n n T T (1.299)
Now, if we need the square root of the tensor, T , this can easily be obtained from the
spectral representation as:


3
1
) ( ) (

a
a a
a
n n T T (1.300)
Next, we can show that a positive definite tensor has positive eigenvalues. For this
purpose, we can consider a semi-positive definite tensor, T , by which the condition
0

x T x holds for all 0 x
,

. Replacing the tensor by its spectral representation, we


obtain:
0

0

0

3
1
) ( ) (
3
1
) ( ) (

|
|
.
|

\
|

x n n x
x n n x
x T x
a
a a
a
a
a a
a
T
T
(1.301)
Note that the result of the operation )

(
) (a
n x is a scalar, thus:
[ ]
0 )

( )

( )

(
0 )

( 0

2 ) 3 (
3
2 ) 2 (
2
2 ) 1 (
1
3
1
2 ) (
3
1
) ( ) (
0
- -





>
n x n x n x
n x x n n x
T T T
T T
a
a
a
a
a a
a
_

(1.302)
The above expression must hold for all 0 x
,

. If we take
) 1 (

n x , the above equation is
reduced to 0 )

(
1
2 ) 1 ( ) 1 (
1
T T n n . The same is true for
2
T and
3
T . Thus, we have
demonstrated that if a tensor is semi-positive definite, its eigenvalues are greater than or
equal to zero, i.e. 0
1
T , 0
2
T , 0
3
T . Therefore we can conclude that a tensor is positive
definite, i.e. 0

> x T x , if and only if its eigenvalues are positive and nonzero, i.e. 0
1
> T ,
0
2
> T , 0
3
> T . Consequently, the positive definite tensor trace is greater than zero. If the
positive definite tensor trace is zero, this implies that the tensor is the zero tensor.
The spectral representation of the fourth-order unit tensor, I , can be obtained starting
from the definition in (1.169), i.e.:



3
1
3
1

a b
b a b a j i j i k j i j ik
e e e e e e e e e e e e

c c I (1.303)
As I is an isotropic tensor, (see 1.5.8 Isotropic and Anisotropic tensors), then the
representation in (1.303) is also valid in any orthonormal basis,
) (

a
n , so:



3
1
3
1
) ( ) ( ) ( ) (

a b
b a b a
n n n n I (1.304)
1 TENSORS

75
Similarly, we obtain the spectral representation for I and I as:
i j j i k j i jk i
e e e e e e e e



c c I
(1.305)


3
1 ,
) ( ) ( ) ( ) (

b a
a b b a
n n n n I
(1.306)
and
k k i i k j i k ij
e e e e e e e e



c c I
(1.307)


3
1 ,
) ( ) ( ) ( ) (

b a
b b a a
n n n n I
(1.308)
Problem 1.34: Let be an antisymmetric second-order tensor and V be a positive
definite symmetric tensor whose spectral representation is given by:

/
3
1
) ( ) (

a
a a
a
n n V
Show that the antisymmetric tensor can be represented by:


3
1 ,
) ( ) (

b a
b a
b a
ab
n n
Demonstrate also that:

/ - / -
3
1 ,
) ( ) (

) (
b a
b a
b a
a b ab
n n V V
Solution:
It is true that
( )
( )


r
r
|
|
.
|

\
|

3
1 ,
) ( ) ( ) (
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (


b a
a a b
b
a
a a
a
a a
a
a a
w n n n
n n n n n n 1 w
,


where we have applied an antisymmetric tensor property n n

r w
,
, where w
,
is the
axial vector associated with . Expanding the above equation, we obtain:
( ) ( ) ( )
( ) ( ) ( )
( ) ( ) ( )
( ) ( ) ( )
) 3 ( ) 3 ( ) 3 (
3
) 3 ( ) 3 ( ) 2 (
2
) 3 ( ) 3 ( ) 1 (
1
) 2 ( ) 2 ( ) 3 (
3
) 2 ( ) 2 ( ) 2 (
2
) 2 ( ) 2 ( ) 1 (
1
) 1 ( ) 1 ( ) 3 (
3
) 1 ( ) 1 ( ) 2 (
2
) 1 ( ) 1 ( ) 1 (
1
) 3 ( ) 3 ( ) ( ) 2 ( ) 2 ( ) ( ) 1 ( ) 1 ( ) (




n n n n n n n n n
n n n n n n n n n
n n n n n n n n n
n n n n n n n n n
r - r - r -
- r - r - r -
- r - r - r
r - r - r
w w w
w w w
w w w
w w w
b
b
b
b
b
b


On simplifying the above expression we obtain:
( ) ( )
( ) ( )
( ) ( )
) 3 ( ) 1 (
2
) 3 ( ) 2 (
1
) 2 ( ) 1 (
3
) 2 ( ) 3 (
1
) 1 ( ) 2 (
3
) 1 ( ) 3 (
2



n n n n
n n n n
n n n n
- -
- - -
- - -
w w
w w
w w

Taking into account that
32 23 1
- w ,
31 13 2
- w ,
21 12 3
- w , the
above equation becomes:
NOTES ON CONTINUUM MECHANICS

76
) 3 ( ) 1 (
13
) 3 ( ) 2 (
23
) 2 ( ) 1 (
12
) 2 ( ) 3 (
32
) 1 ( ) 2 (
21
) 1 ( ) 3 (
31



n n n n
n n n n
n n n n
- -
- - -
- -




which is the same as:


3
1 ,
) ( ) (

b a
b a
b a
ab
n n
The terms V and V can be expressed as follows:

/ /
|
|
.
|

\
|
/
|
|
|
.
|

\
|


3
1 ,
) ( ) (
3
1 ,
) ( ) ( ) ( ) (
3
1
) ( ) (
3
1 ,
) ( ) (


b a
b a
b a
ab b
b a
b a
b b b a
ab b
b
b b
b
b a
b a
b a
ab
n n n n n n
n n n n V



and


/
|
|
|
.
|

\
|

|
|
.
|

\
|
/
3
1 ,
) ( ) (
3
1 ,
) ( ) (
3
1
) ( ) (

b a
b a
b a
ab a
b a
b a
b a
ab
a
a a
a
n n n n n n V
Then,

/ - /
|
|
|
.
|

\
|
/ -
|
|
|
.
|

\
|
/ -
3
1 ,
) ( ) (
3
1 ,
) ( ) (
3
1 ,
) ( ) (

) (

b a
b a
b a
a b ab
b a
b a
b a
ab a
b a
b a
b a
ab b
n n
n n n n V V



Similarly, it is possible to show that:

/ - / -
3
1 ,
) ( ) ( 2 2 2 2

) (
b a
b a
b a
a b ab
n n V V

1.5.6 Cayley-Hamilton Theorem
The Cayley-Hamilton theorem states that any tensor, T , satisfies its own characteristic
equation, i.e. if the eigenvalues of T satisfy the equation 0
2 3
- / - / - /
T T T
I I I I I I , so
does the tensor T :
0 1 T T T
T T T
- - - I I I I I I
2 3
(1.309)
One of the applications of the Cayley-Hamilton theorem is to express the power of tensor,
n
T , as a combination of
1 - n
T ,
2 - n
T ,
3 - n
T . For example,
4
T is obtained as:
T T T T 0 T 1 T T T T T T
T T T T T T
I I I I I I I I I I I I - - - - -
2 3 4 2 3
(1.310)
Using the Cayley-Hamilton theorem, it is possible to express the third invariant as a
function of traces. According to the Cayley-Hamilton theorem, the expression
1 TENSORS

77
0 1 T T T
T T T
- - - I I I I I I
2 3
remains valid. Additionally, by applying the double scalar
product with the second-order unit tensor, 1, we obtain:
1 0 1 1 1 T 1 T 1 T
T T T
: : : : : - - - I I I I I I
2 3
(1.311)
Taking into consideration ) (
3 3
T 1 T Tr : , ) (
2 2
T 1 T Tr : , ) (T 1 T Tr : , 3 ) ( 1 1 1 Tr : ,
0 ) ( 0 1 0 Tr : in the equation (1.311) we obtain:
[ ] ) ( ) ( ) (
3
1
0 ) ( ) ( ) ( ) (
2 3
3
2 3
T T T
1 T T T
T T T
T T T
Tr Tr Tr
Tr Tr Tr Tr
I I I I I I
I I I I I I
- -
- - -

_

(1.312)
Replacing the values of the invariants,
T
I ,
T
I I , given by equation (1.269), we obtain:
[ ]
)
`

- -
3
2 3
) (
2
1
) ( ) (
2
3
) (
3
1
T T T T
T
Tr Tr Tr Tr I I I
(1.313)
or in indicial notation
)
`

- -
kk jj ii kk ji ij ki jk ij
I I I T T T T T T T T T
2
1
2
3
3
1
T
(1.314)

Problem 1.35: Based on the Cayley-Hamilton theorem, find the inverse of a tensor T in
terms of tensor power.
Solution: The Cayley-Hamilton theorem states that:
0 1 T T T
T T T
- - - I I I I I I
2 3

Carrying out the dot product between the previous equation and the tensor
1 -
T , we
obtain:
( ) 1 T T T
0 T 1 T T
T 0 T 1 T T T T T T
T T
T
T T T
T T T
I I I
I I I
I I I I I I
I I I I I I
- -
- - -
- - -
-
-
- - - - -

2 1
1 2
1 1 1 1 2 1 3
1

The Cayley-Hamilton theorem also applies to square matrices of order n . Let
n n
A be a
square n by n matrix. The characteristic determinant is given by:
0 - /

A
n n
1
(1.315)
where
n n
1 is the identity n by n matrix. Developing the determinant (1.315) we ontain:
0 ) 1 (
2
2
1
1
- - / - / - /
- -
n
n n n n
I I I (1.316)
where
n
I I I , , ,
2 1
are the invariants of A. In the particular case when 3 n , the
invariants are the same obtained for a second-order tensor, i.e.:
A
I I
1
,
A
I I I
2
,
A
I I I I
3
.
Applying the Cayley-Hamilton theorem it is true that:
0 1 - - - - -
- -
n
n n n n
I I I ) 1 (
2
2
1
1
A A A (1.317)
By means of the relationship (1.317), we can obtain the inverse of the matrix
n n
A by
multiplying all the terms by the inverse,
1 -
A , i.e.:
NOTES ON CONTINUUM MECHANICS

78
0 1
0 1
- - - - - -
- - - - -
-
-
- - - -
- - - - - -
1
1
1 3
2
2
1
1
1 1 2
2
1 1
1
1
) 1 ( ) 1 (
) 1 (
A A A A
A A A A A A A
n
n
n
n n n n
n
n n n n
I I I I
I I I


(1.318)
then
( ) 1
1
1 3
2
2
1
1
1
1
) 1 (
) 1 (
-
- - - -
-
-
- - - -
-

n
n n n n
n
n
I I I
I
A A A A (1.319)
n
I is the determinant of
n n
A . Then, the inverse exists if 0 ) ( A det
n
I .
Problem 1.36: Check the Cayley-Hamilton theorem by using a second-order tensor whose
Cartesian components are given by:
(
(
(

1 0 0
0 2 0
0 0 5
T
Solution:
The Cayley-Hamilton theorem states that:
0 1 - - -
T T T
I I I I I I T T T
2 3

where 8 1 2 5 - -
T
I , 17 5 2 10 - -
T
I I , 10
T
I I I , and
(
(
(

(
(
(

1 0 0
0 8 0
0 0 125
1 0 0
0 2 0
0 0 5
3
3
3
T ;
(
(
(

(
(
(

1 0 0
0 4 0
0 0 25
1 0 0
0 2 0
0 0 5
2
2
2
T
By applying the Cayley-Hamilton theorem, we can verify that it is true:
(
(
(

(
(
(

-
(
(
(

-
(
(
(

-
(
(
(

0 0 0
0 0 0
0 0 0
1 0 0
0 1 0
0 0 1
10
1 0 0
0 2 0
0 0 5
17
1 0 0
0 4 0
0 0 25
8
1 0 0
0 8 0
0 0 125

1.5.7 Norms of Tensors
The magnitude (module) of a tensor, also known as the Frobenius norm, is given below:
i i
v v v v v
, , ,
(vector) (1.320)
ij ij
T T T T T : (second-order tensor)
(1.321)
ijk ijk
A A A A A : (third-order tensor)
(1.322)
ijkl ijkl
C C C C C :: (fourth-order tensor)
(1.323)
Interpreting the Frobenius norm of T is done by considering the principal space of T
where
3 2 1
, , T T T are the eigenvalues of T . In this space, it follows that:
T T
T T T I I I
ij ij
2
2 2
3
2
2
2
1
- - - T T T T T : (1.324)
As we can verify T is an invariant, and T represents a measurement of distance as
shown in Figure 1.28.
1 TENSORS

79








Figure 1.28: Norm of a second-order tensor.
1.5.8 Isotropic and Anisotropic Tensor
A tensor is called isotropic when its components are the same in any coordinate system,
otherwise the tensor is said to be anisotropic.
Let T and T represent the tensor components T in the systems
i
e

and
i
e

,
respectively, so, the tensor is isotropic if T T on any arbitrary basis.
Isotropic first-order tensor
Let v
,
be a vector that is represented by its components,
3 2 1
, , v v v , in the coordinate
system
3 2 1
, , x x x . The representation of these components in a new coordinate system,
3 2 1
, , x x x , are given by
3 2 1
, , v v v , so the transformation law for these components is:
j ij i j j i i
a v v v v e e v

,

(1.325)
By definition, v
,
is an isotropic tensor if it holds that
i i
v v , and this is only possible if
j i
e e

, i.e. there is no change of system, or if the tensor is the zero vector, i.e.
i i i
0 v v . Then, the unique isotropic first-order tensor is the zero vector 0
,
.
Isotropic second-order tensor
An example of a second-order isotropic tensor is the unit tensor, 1, whose components
are represented by
kl
c (Kronecker delta). In the demonstration, we use the transformation
law for a second-order tensor components, obtained in (1.248), thus:
ij
T
jk ik kl jl ik ij
a a a a c c c

_
1 AA

(1.326)
An immediate observation of the isotropy of unit tensor 1 is that any spherical tensor
( 1 c ) is also an isotropic tensor. So, if a second-order tensor is isotropic it is spherical and
vice versa.
Isotropic third-order tensor
An example of a third-order isotropic tensor is the Levi-Civita pseudo-tensor, defined in
(1.182), which is not a real tensor in the strict meaning of the word. With reference to
the transformation law for the third-order tensor components, (see equation (1.248)), we
can conclude that:

2
x

1
x

3
x

1
T

2
T

3
T
T

T T
T T T I I I 2
2
- :
NOTES ON CONTINUUM MECHANICS

80


1
ijk ijk lmn kn jm il ijk
a a a A (see Problem 1.21)
(1.327)
Isotropic fourth-order tensor
With reference to the transformation law for fourth-order tensor components, (see
equation (1.249)), it is possible to demonstrate that the following tensors are isotropic:
jk il ijkl jl ik ijkl kl ij ijkl
c c c c c c I I I ; ; (1.328)
Therefore, any fourth-order isotropic tensor can be represented by a linear combination of
the three tensors given in (1.328), e.g.:
I I I D
2 1 0
a a a - -
1 1 1 1 1 1 - -
2 1 0
a a a D
jk il jl ik kl ij ijkl
a a a c c c c c c
2 1 0
- - D
(1.329)
Problem 1.37: Let C be a fourth-order tensor, whose components are given by:
( )
jk il jl ik kl ij ijkl
c c c c j c Ac - - C
where A, j are constant real numbers. Show that C is an isotropic tensor.
Solution:
Applying the transformation law for fourth-order tensor components:
mnpq lq kp jn im ijkl
a a a a C C
and by replacing the relation ( )
np mq nq mp pq mn mnpq
c c c c j c Ac - - C in the above equation,
we obtain:
( ) [ ]
( )
( )
( )
ijkl
jk il jl ik kl ij
lq kn jn iq lq kp jq ip lq kq jn in
np mq lq kp jn im nq mp lq kp jn im pq mn lq kp jn im
np mq nq mp pq mn lq kp jn im ijkl
a a a a a a a a a a a a
a a a a a a a a a a a a
a a a a
C
C

- -
- -
- -
- -
c c c c j c Ac
j A
c c c c j c c A
c c c c j c Ac

which is proof that C is an isotropic tensor.

1.5.9 Coaxial Tensors
Two arbitrary second-order tensors, T and S , are coaxial tensors if they have the same
eigenvectors. It is easy to show that if two tensors are coaxial, this means the dot product
between them is commutative, and vice versa, i.e.:
coaxial are if , T S T S S T =
(1.330)
If T and S are coaxial as well as symmetric tensors, the spectral representations of these
tensors are given by:



3
1
) ( ) (
3
1
) ( ) (

;

a
a a
a
a
a a
a
n n S n n T S T (1.331)
An immediate result of (1.330) is that the tensor S and its inverse
1 -
S are coaxial tensors:
1 TENSORS

81


-

- -


3
1
) ( ) ( 1
3
1
) ( ) (
1 1

1
;

a
a a
a a
a a
a
n n S n n S
1 S S S S
S
S
(1.332)
where
a
S ,
a
S
1
, are the eigenvalues of S and
1 -
S , respectively.
If S and T are coaxial symmetric tensors, the resulting tensor ( T S ) becomes another
symmetric tensor. To prove this we start from the definition of coaxial tensors:
0 S T 0 S T S T 0 T S S T T S S T - -
skew T
) ( 2 ) ( (1.333)
Then, if the antisymmetric part of a tensor is a zero tensor, it follows that this tensor is
symmetric:
sym skew
) ( ) ( ) ( S T S T 0 S T = (1.334)
1.5.10 Polar Decomposition
Let F be an arbitrary nonsingular second-order tensor, i.e. (
1
0 ) (
-
F F det ).
Additionally, as previously seen, it satisfies the condition 0 n n N
n
N N
, , ,
/

(
)

( )

(
f f F ,
since 0 ) ( F det . After that, given an orthonormal basis
) (

a
N , we can obtain:

-
/


3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) ( 1



a
a a
a
a
a a
a
a a
a
a a
N n
N N N N 1
N N 1
F
F F F F
F F

(1.335)
NOTE: The representation of F , given in (1.335), is not the spectral representation of F
in the strict sense of the word, i.e.,
a
/ are not eigenvalues of F , and neither
) (

a
n nor
) (

a
N
are eigenvectors of F .










Figure 1.29: Projecting F onto
) (

a
N .

) 2 (

n

) 1 (

n

) 3 (

N

) 1 (

N

) 2 (

N

) 3 (

n

) 1 ( )

) 1 (
N
N
F f
,


) 2 ( )

) 2 (
N
N
F f
,


) 3 ( )

) 3 (
N
N
F f
,


) 3 (
3
) 3 ( )

( )

( ) 3 (
) 2 (
2
) 2 ( )

( )

( ) 2 (
) 1 (
1
) 1 ( )

( )

( ) 1 (

) 3 ( ) 3 (
) 2 ( ) 2 (
) 1 ( ) 1 (
n n N
n n N
n n N
N N
N N
N N
/
/
/

f f F
f f F
f f F
, ,
, ,
, ,

NOTES ON CONTINUUM MECHANICS

82

Note that for the arbitrary orthonormal basis
) (

a
N , the new basis
) (

a
n will not necessarily
be orthonormal. We seek to find a basis
) (

a
N so that the new basis
) (

a
n is orthonormal,
(see Figure 1.29), i.e. 0
)

( )

(
) 2 ( ) 1 (

N N
f f
, ,
, 0
)

( )

(
) 3 ( ) 2 (

N N
f f
, ,
, 0
)

( )

(
) 1 ( ) 3 (

N N
f f
, ,
. Then we
look for a space in accordance with the following orthogonal transformation
) ( ) (

a a
N R n , which ensures
) (

a
n orthonormality since an orthogonal transformation
changes neither angles between vectors nor their magnitudes.
Now, consider that there is a transformation from
) (

a
N to
) (

a
n , which is given by the
following orthogonal transformation
) ( ) (

a a
N R n , then we can state that:
F F
F



/ / /


T
a
a a
a
a
a a
a
a
a a
a
R U U R
U R N N R N N R N n
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (

(1.336)
where we have defined the tensor

/
3
1
) ( ) (

a
a a
a
N N U . Note that U is a symmetric
tensor, i.e.
T
U U . This condition is easily verified by the fact that
) ( ) (

a a
N N is also
symmetric. Now considering that R n n R N N R n
) ( ) ( ) ( ) ( ) (

a a T a a a
, we obtain:
T
a
a a
a
a
a a
a
a
a a
a
R V R V
R V R n n R n n N n



/ / /


F F
F
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (

(1.337)
where we have defined the symmetric second-order tensor

/
3
1
) ( ) (

a
a a
a
n n V . By
comparing the spectral representation of U with V , we can conclude that they have the
same eigenvalues but different eigenvectors, and they are related by
) ( ) (

a a
N R n .
With reference to the above considerations, we can define the polar decomposition:
R V U R F Polar Decomposition (1.338)
Carrying out the dot product between
T
F and U R F , we obtain:
C F F F F F F
T T T T T
C
T
U U U U U R U R ) (
2
_

(1.339)
Moreover, by carrying out the dot product between R V F and
T
F , we obtain:
b F F F F F F
T T T T T
b
T
V V V V R V R V ) (
2
_

(1.340)
Since 0 ) ( F det , the tensors C and b are positive definite symmetric tensors, (see
Problem 1.25), which implies that the eigenvalues of C and b are all real and positive.
However, up to now, 0 ) ( F det is the only restriction imposed on the tensor F .
Therefore, we have the following possibilities:
If 0 ) ( > F det
In this scenario, we have 0 ) ( ) ( ) ( ) ( ) ( > R V U R det det det det det F , which results in the
following cases:
1 TENSORS

83

-
-
tensors definite Positive ,
tensor ortogonal Proper
V U
R
or

-
-
tensors definite Negative ,
tensor orthogonal Improper
V U
R

If 0 ) ( < F det
In this situation, we have 0 ) ( ) ( ) ( ) ( ) ( < R V U R det det det det det F , which give us the
following cases:

-
-
tensors definite Negative ,
tensor orthogonal Proper
V U
R
or

-
-
tensors definite Positive ,
tensor orthogonal Improper
V U
R

NOTE: In Chapter 2 we will work with some special tensors where F is a nonsingular
tensor, 0 ) ( F det , and 0 ) ( > F det . U and V are positive definite tensors and R is a
rotation tensor, i.e. a proper orthogonal tensor.
1.5.11 Partial Derivative with Tensors
The first derivative of a tensor with respect to itself is defined as:
I
o
o
=
o
o
)

( )

( ,
l k j i jl ik l k j i
kl
ij
e e e e e e e e A
A
A
A
c c
A
A
(1.341)
The derivative of a tensor trace with respect to a tensor:
[ ]
[ ] 1 e e e e e e A
A
A
A

o
o
=
o
o
)

( )

( )

( ) (
) (
, j i ij j i kj ki j i
ij
kk
c c c
A
A
Tr
Tr

(1.342)
The derivative of the tensor trace squared with respect to the tensor is given by:
[ ] [ ]
1 A
A
A
A
A
A
) ( 2
) (
) ( 2
) (
2
Tr
Tr
Tr
Tr

o
o

o
o

(1.343)
And, the derivative of the trace of the tensor squared with respect to tensor is given by:
[ ]
[ ] [ ]
T
j i ji
j i ji ji j i sj ri sr rj si rs
j i
ij
rs
sr
ij
sr
rs j i
ij
rs sr
A e e
e e e e
e e e e
A
A
2 )

( 2
)

( )

(
)

(
) ( ) (
)

(
) ( ) (
2

- -

(
(

o
o
-
o
o

o
o

o
o
A
A A A A
A
A
A
A
A
A
A
A A Tr
c c c c
(1.344)
We leave the reader with the following demonstration:
[ ]
T
) ( 3
) (
2
3
A
A
A

o
o Tr

(1.345)
Then, if we are considering a symmetric second-order tensor, C, it is true that
[ ]
1
C
C

o
o ) ( Tr
,
[ ]
1 C
C
C
) ( 2
) (
2
Tr
Tr

o
o
,
[ ]
C C
C
C
2 2
) (
2

o
o
T
Tr
,
[ ]
2 2
3
3 ) ( 3
) (
C C
C
C

o
o
T
Tr
.
Moreover, we can say that the derivative of the Frobenius norm of C is given by:
NOTES ON CONTINUUM MECHANICS

84
( )
[ ] [ ]
[ ] C C
C C
C
C
C
C C
C
C C
C
C
C
2 ) (
2
1
) ( ) (
2
1
) ( ) (
2
1
2
2
2
1
2
2
,
-
-

o
|
.
|

\
|
o

o
|
.
|

\
|
o

o
o

o
o

Tr
Tr Tr
Tr Tr
T
:
(1.346)
or:
C
C
C
C
C
C

o
o
) (
2
Tr

(1.347)
Another interesting derivative is presented below:
( )
j kj j
sym
kj j jk kj j jk j kj
ik i j kj jk ij i j ij ik
k
j
ij i j ij
k
i
k
j ij i
n C n C n C C n C n C
C n n C C n n C
n
n
C n n C
n
n
n
n C n
2 2 ) ( - -
- -
o
o
-
o
o

o
o
c c

(1.348)
where we have assumed that C is symmetric, i.e.,
jk kj
C C .
Let C be a symmetric second-order tensor. The partial derivative of
1 -
C with respect to
the tensor C is obtained by using the following relationship:
( )
O
o
o

o
o
-
C
C C
C
1
1

(1.349)
where O is the fourth-order zero tensor and the above equation in indicial notation
becomes:
( ) ( ) ( )
( ) ( ) ( ) ( )
( ) ( )
1 1
1
1 1 1
1
1
1
1
1 1
- -
-
- - -
-
-
-
-
- -
o
o
-
o
o

o
o
-
o
o

o
o
-
o
o

o
o
-
o
o

o
o
jr
kl
qj
iq qr
kl
iq
jr
kl
qj
iq jr qj
kl
iq
kl
qj
iq qj
kl
iq
ikjl
kl
qj
iq qj
kl
iq
kl
qj iq
C
C
C
C
C
C
C
C
C
C C C
C
C
C
C
C C
C
C
C
C
C C
C
C
C
C C
c
O
(1.350)
whereas ( )
jq qj qj
C C C -
2
1
, so we can conclude that:
( ) ( )
( )
[ ] [ ]
( )
[ ]
1 1 1 1
1
1 1 1 1 1 1
1
1 1
1
2
1
2
1
2
1
2
1
- - - -
-
- - - - - -
-
- -
-
- -
o
o
- - - -
o
o
(

o
- o
-
o
o
kr il lr ik
kl
ir
jr ql jk iq jr jl qk iq jr ql jk jl qk iq
kl
ir
jr
kl
jq qj
iq qr
kl
iq
C C C C
C
C
C C C C C C
C
C
C
C
C C
C
C
C
c c c c c c c c
c

(1.351)
Or in tensorial notation:
[ ]
1 1 1 1
1
2
1
- - - -
-
- -
o
o
C C C C
C
C

(1.352)
1 TENSORS

85
NOTE: Note that, if we had not replaced the symmetric part of
qj
C in (1.351), we would
have found that
( ) ( )
1 1 1 1 1 1
1
- - - - - -
-
- -
o
o
-
o
o
lr ik jr jl qk iq jr
kl
qj
iq qr
kl
iq
C C C C C
C
C
C
C
C
c c c , which is a non-
symmetric tensor.
1.5.11.1 Partial Derivative of Invariants
Let T be a second-order tensor. The partial derivative of
T
I with respect to T , (see
equation (1.342)), is:
[ ] [ ]
[ ] 1 T
T
T
T
T
T

o
o

o
o
,
) (
) (
Tr
Tr I
(1.353)
The partial derivative of
T
I I with respect to T , (see equation (1.342)), is:
[ ]
[ ] [ ]
[ ] [ ]
[ ]
T
T
I I
T 1 T
T 1 T
T
T
T
T
T T
T T
T
-
-
(
(

o
o
-
o
o

)
`

-
o
o

o
o
) (
2 ) ( 2
2
1
) ( ) (
2
1
) ( ) (
2
1
2 2
2 2
Tr
Tr
Tr Tr
Tr Tr

(1.354)
Next, we apply the Cayley-Hamilton theorem so as to represent T as:
2 1
2 1
2 2 2 2 2 3
- -
- -
- - - -
- -
- - -
- - -
T T 1 T
0 T T 1 T
0 T 1 T T T T T T
T T T
T T T
T T T
I I I I I I
I I I I I I
I I I I I I : : : :
(1.355)
By substituting (1.355) into the equation in (1.354), we obtain:
[ ]
( ) ( )
T T
T
I I I I I I I I I I I
I I
2 1 2 1
) ( ) (
- - - -
- - - - -
o
o
T T T T 1 1 T T 1 T
T
T T T T T
T
Tr Tr (1.356)
To find the partial derivative of the third invariant, we can start with the definition given in
(1.313), so:
[ ]
[ ]
[ ] [ ]
[ ]
[ ]
[ ] [ ]
1 T T
1 T T T T T
1 T 1 T T T T
1 T
T
T
T T
T
T
T
T T T T
T T
T T
T
I I I
I I I
T T
T T
T T
T
- -
- - -
- - -
-
o
o
-
o
o
-
)
`

- -
o
o

o
o
) (
) ( ) (
2
1
) ( ) (
) (
2
1
) (
2
1
) ( ) (
) (
6
3 ) (
) (
2
1
) (
) (
2
1
) ( 3
3
1
) (
6
1
) ( ) (
2
1
) (
3
1
2
2
2
2
2
2 2
2
2
2
2
3
2 3
Tr Tr Tr
Tr Tr Tr
Tr
Tr
Tr Tr
Tr
Tr Tr Tr Tr

(1.357)
Once again using the Cayley-Hamilton theorem we obtain:
1 T T T
0 T 1 T T
0 T 1 T T T T T T
T T T
T T T
T T T
I I I I I I
I I I I I I
I I I I I I
- -
- - -
- - -
-
-
- - - -

2 1
1 2
1 1 1 2 1 3
(1.358)
and the transpose:
NOTES ON CONTINUUM MECHANICS

86
( ) ( ) ( ) 1 T T 1 T T T
T T T T T
I I I I I I I I I
T
T T T
- - - -
- 2 2 1

(1.359)
By comparing (1.357) with (1.359) we find another way to express the derivative of
T
I I I
with respect to T , i.e.:
[ ]
( )
T
T
I I I I I I
I I I
- -

o
o
T T
T
T T
T 1
(1.360)
1.5.11.2 Time Derivative of Tensors
Let us assume a second-order tensor depends on the time, t , i.e. ) t T( T . Then, we define
the first time derivative and the second time derivative of the tensor T , respectively, as:
T T T T
` ` `

2
2
;
Dt
D
Dt
D
(1.361)
The time derivative of a tensor determinant is defined as:
( ) [ ] ( )
ij
ij
Dt
D
Dt
D
T cof
T
det T
(1.362)
where ( )
ij
T cof is the cofactor of
ij
T and defined as ( ) [ ] ( ) ( )
ij
T
ij
1

-
T T det T cof .

Problem 1.38: Consider that [ ] ( )2
1
2
1
) (
b
b I I I J det , where b is a symmetric second-order
tensor, i.e.
T
b b . Obtain the partial derivatives of J and ) (J ln with respect to b .

Solution:
( )
( ) ( )
( )
1 1
2
1
2
1
2
1
2
1

2
1

2
1
2
1
2
1
- -
-
- -

o
o

o
(

o
o
b b
b
b
b b
b
b b
b
b
b
J I I I
I I I I I I
I I I
I I I
I I I
J
T

( ) [ ]
1
2
1

2
1
2
1
-

o
o

o
(

|
.
|

\
|
o

o
o
b
b b b
b
b
b
I I I
I I I
I I I
J
ln
ln

1.5.12 Spherical and Deviatoric Tensors
Any tensor can be decomposed into a spherical and a deviatoric part, so, for a given
second-order tensor T , this decomposition is represented by:
dev
m
dev dev dev sph
I
T 1 T 1 T 1
T
T T T
T
- - - - T
Tr
3 3
) (
(1.363)
The deviatoric part of the tensor T is defined as:


1 TENSORS

87
1 T 1
T
T T
m
dev
T
Tr
- -
3
) (
(1.364)
For the following operations, we consider that T is a symmetric tensor,
T
T T , then
under this condition the deviatoric tensor components,
dev
ij
T , become:
(
(
(

- -
- -
- -

(
(
(

-
-
-

(
(
(

) 2 (
) 2 (
) 2 (
22 11 33 3
1
23 13
23 33 11 22
3
1
12
13 12 33 22 11 3
1
33 23 13
23 22 12
13 12 11
33 23 13
23 22 12
13 12 11
T T T T T
T T T T T
T T T T T
T T T T
T T T T
T T T T
T T T
T T T
T T T
T
m
m
m
dev dev dev
dev dev dev
dev dev dev
dev
ij

(1.365)
Graphical representations of the Cartesian components of the spherical and deviatoric
parts are shown in Figure 1.30.
In the following subsections we obtain the deviatoric tensor invariants in terms of the
principal invariants of T .
1.5.12.1 First Invariant of the Deviatoric Tensor
0 ) (
3
) (
) (
3
) (
) (
3
-
(

_
ii
dev
dev
I
c
1
T
T 1
T
T T
T
Tr
Tr
Tr
Tr
Tr Tr
(1.366)
Thus, we can conclude that the trace of any deviatoric tensor is equal to zero.
1.5.12.2 Second Invariant of the Deviatoric Tensor
For simplicity we can use the principal space to obtain the second and third invariant of the
deviatoric tensor. In the principal space the components of T are given by:
(
(
(

3
2
1
0 0
0 0
0 0
T
T
T
T
ij

(1.367)
The principal invariants of T :
3 2 1
T T T - -
T
I ,
1 3 3 2 2 1
T T T T T T - -
T
I I ,
3 2 1
T T T
T
I I I .
The deviatoric components, 1 T T
m
dev
T - , in the principal space are:
(
(
(

-
-
-

m
m
m
dev
ij
T T
T T
T T
T
3
2
1
0 0
0 0
0 0

(1.368)
So, the second invariant of deviatoric tensor
dev
T is evaluated as follows:
( )
2
2
2
3 2 1 3 2 3 1 2 1
3 2 3 1 2 1
3
3
1
3
) (
3
2
3 ) ( 2 ) (
) )( ( ) )( ( ) )( (
T T
T
T
T
T
T
I I I
I
I
I
I I
I I
m m
m m m m m m dev
-
- -
- - - - - -
- - - - - - - -
T T T T T T T T T T T
T T T T T T T T T T T T
(1.369)
NOTES ON CONTINUUM MECHANICS

88
We could also have obtained the above result, by directly starting from the definition of the
second invariant of a tensor given in (1.269), i.e.:
[ ] { [ ] {
[ ] {
[ ] {
[ ]
(

- -
(

- - -
- - -
- - -
- -
- -

3
) (
2
1
3
9 3
2 ) (
2
1
) ( ) ( 2 ) (
2
1
) 2 (
2
1
) (
2
1
) (
2
1
) ( ) (
2
1
2
2
2
2
2 2
2 2
2
2 2 2
T
T
T
T
T
T
1 T T
1 1 T T
1 T
T T
T T
I
I
I
I
I I I
m m
m m
m
dev dev
dev dev
Tr
Tr
Tr T Tr T Tr
T T Tr
T Tr
Tr Tr
(1.370)
Observing that
T
T
T I I I 2 ) (
2 2
3
2
2
2
1
2
- - - T T T Tr , (see Problem 1.31), the equation
(1.370) becomes:
( )
2
2 2
2
3
3
1
3
2
2
2
1
3
2
2
1
T T
T T
T
T T
T
I I I
I
I I
I
I I I I I
dev
-
(

-
(

- - -
(1.371)




















Figure 1.30: Spherical and deviatoric part.

_


11
T

12
T
13
T

33
T

23
T

13
T

22
T

23
T

12
T

1
x

2
x

m
T

m
T

m
T

1
x

2
x

3
x

dev
11
T

12
T
13
T

dev
33
T

23
T

13
T

dev
22
T

23
T

12
T

1
x

2
x

3
x
+

3
x
1 TENSORS

89
Another equation for
dev
I I
T
is presented in terms of deviatoric tensor components. To
calculate this, we can apply the equation (1.370):
[ ] [ ]
dev
ji
dev
ij
dev dev dev dev dev
dev
I I T T Tr Tr
2
1
2
1
) (
2
1
) (
2
1
2
- - - - T T T T T
T
(1.372)
Expanding the previous equation we obtain:
[ ]
2
23
2
13
2
12
2
33
2
22
2
11
) ( 2 ) ( 2 ) ( 2 ) ( ) ( ) (
2
1
dev dev dev dev dev dev
dev
I I T T T T T T - - - - - -
T
(1.373)
Additionally, in the space of the principal directions we obtain:
[ ]
2
3
2
2
2
1
) ( ) ( ) (
2
1
2
1
dev dev dev dev
ji
dev
ij dev
I I T T T T T - - - -
T
(1.374)
Another way to express the second invariant is shown below:
[ ]
2
13
2
23
2
12 22 11 33 11 33 22
22 12
12 11
33 13
13 11
33 23
23 22
) ( ) ( ) ( 2 2 2
2
1
dev dev dev dev dev dev dev dev dev
dev dev
dev dev
dev dev
dev dev
dev dev
dev dev
dev
I I
T T T T T T T T T
T T
T T
T T
T T
T T
T T
- - - - - - -
- -
T
(1.375)
or
( ) ( ) ( ) ( )
( ) ( ) ( ) ( ) ( )
2
13
2
23
2
12
2
33
2
22
2
11
2
22 22 11
2
11
2
33 33 11
2
11
2
33 33 22
2
22
) ( ) ( ) (
2
2 2
2
1
dev dev dev
dev dev dev dev dev dev dev
dev dev dev dev dev dev dev dev
dev
I I
T T T
T T T T T T T
T T T T T T T T
- - -
- - -
(

(
- -

- - - - - - -
T

(1.376)
Note that, from equation (1.373), we can state that:
2
23
2
13
2
12
2
33
2
22
2
11
) ( 2 ) ( 2 ) ( 2 2 ) ( ) ( ) (
dev dev dev
dev
dev dev dev
I I T T T T T T - - - - - -
T

(1.377)
Substituting (1.377) into (1.376), we find:
[ ]
2
13
2
23
2
12
2
22 11
2
33 11
2
33 22
) ( ) ( ) ( ) ( ) ( ) (
6
1
dev dev dev dev dev dev dev dev dev
dev
I I T T T T T T T T T - - - - - - - - -
T


(1.378)
Moreover, if we consider the principal space we obtain:
[ ]
2
2 1
2
3 1
2
3 2
) ( ) ( ) (
6
1
dev dev dev dev dev dev
dev
I I T T T T T T - - - - - -
T
(1.379)
1.5.12.3 Third Invariant of Deviatoric Tensor
The third invariant of the deviatoric tensor is given by:
( )
T T T T
T T T
T
T
T
T
T
T
T
T
I I I I I I I
I I I I
I I I
I
I
I
I I
I
I I I
I I I
m m m
m m m dev
27 9 2
27
1
27
2
3
27 9 3
) ( ) (
) )( )( (
3
3
3 2
3
3 2 1
2
3 2 3 1 2 1 3 2 1
3 2 1
- -
- -
- - -
- - - - - - -
- - -
T T T T T T T T T T T T T T T
T T T T T T
(1.380)
NOTES ON CONTINUUM MECHANICS

90
Another way of expressing the third invariant is:
dev
ki
dev
jk
dev
ij
dev dev dev
dev
I I I T T T T T T
3
1
3 2 1

T
(1.381)
Problem 1.39: Let be a symmetric second-order tensor, and
dev
s = be a deviatoric
tensor. Prove that s

s
s
o
o
: . Also show that and
dev
are coaxial tensors.
Solution: First, we make use of the definition of a deviatoric tensor:
1 s s 1 s

3 3
I I
sph dev sph
- - - - .
Afterwards we calculate:
[ ] [ ]
1

o
o
-
o
o

o
(

- o

o
o I
I
3
1 3

which in indicial notation is:
[ ]
ij kl jl ik ij
kl kl
ij
kl
ij
I
c c c c c
3
1
3
1
-
o o
o
-
o o
o o

o o
o

s

Therefore

kl
ii kl kl ij kl ij jl ik ij ij kl jl ik ij
kl
ij
ij
s
s s s s s
s
s

- - |
.
|

\
|
-
o o
o
0
3
1
3
1
3
1
c c c c c c c c c

s

s
s
o
o
:
To show that two tensors are coaxial, we must prove that
dev dev
:
1
1 1
1




|
.
|

\
|
-
- -
- - -
dev
sph sph dev
I
I I
I
3
3 3
3
) (

Therefore, we have shown that and
dev
are coaxial tensors. In other words, they have
the same principal directions.

1 TENSORS

91
1.6 The Tensor-Valued Tensor Function
Tensor-valued tensor function can be of the types: scalar, vector, or higher-order tensors.
As examples of scalar-valued tensor functions we can list:
( )
S T S T
T T
:

) , (
) (
1 1
1 1 det
(1.382)
where T and S are second-order tensors. Additionally, as an example of a second-order-
valued tensor function we have:
( ) T 1 T , c - H H
(1.383)
where c and , are scalars.
1.6.1 The Tensor Series
The function ) (x f can be approximated by the Taylor series as
n
n
n
n
a x
x
a f
n
x f ) (
) (
!
1
) (
0
-
o
o

, where ! n denotes the factorial of n , and ) (a f is the value of


the function at the application point a x . We can extrapolate that definition for use on
tensors. For example, let us suppose we have a scalar-valued tensor function in terms of
a second-order tensor, E , then we can approximate ) (E as:

- -
o o
o
- - -
o
o
- =
- - -
o o
o
- -
o
o
- =
) (
) (
) (
2
1
) (
) (
) )( (
) (
! 2
1
) (
) (
! 1
1
) (
! 0
1
) (
0
0
2
0 0
0
0
0
0
0
2
0
0
0
E E
E E
E
E E E E
E
E
E E
E E
: : :



kl kl ij ij
kl ij
ij ij
ij
E E E E
E E
E E
E
(1.384)
A second-order-valued tensor function, ) (E S , can be approximated as:

- -
o o
o
- - -
o
o
- =
- -
o o
o
- - -
o
o
- =
) (
) (
) (
2
1
) (
) (
) (
) (
) (
! 2
1
) (
) (
! 1
1
) (
! 0
1
) (
0
0
2
0 0
0
0
0
0
2
0 0
0
0
E E
E E
E
E E E E
E
E
E E
E E
E
E E E E
E
E
E E
: : :
: : :
S S
S
S S
S S
(1.385)
Other tensor algebraic expressions can be represented by series, e.g.:

- - -
- - - -
- - - -
5 3
3 2
3 2
! 5
1
! 3
1
) sin(
3
1
2
1
) (
! 3
1
! 2
1
S S S S
S S S S 1
S S S 1
S
ln
exp
(1.386)
With reference to the spectral representation of a symmetric second-order tensor, S , it is
also true that:
NOTES ON CONTINUUM MECHANICS

92




/ - |
.
|

\
|
- / - / - / -

|
|
.
|

\
|
-
/
-
/
- / -
/
3
1
) ( ) (
3
1
) ( ) ( 3 2
3
1
) ( ) (
3
1
) ( ) (
3 2

) 1 (

3
1
2
1
) (

! 3 ! 2
1
a
a a
a
a
a a
a a a
a
a a
a
a
a a a a
a
n n n n S 1
n n n n
S
ln ln
exp exp


(1.387)
where
a
/ and
) (

a
n are the eigenvalues and eigenvectors, respectively, of the tensor S .
1.6.2 The Tensor-Valued Isotropic Tensor Function
A second-order-valued tensor function, ) (T H H , is isotropic if after an orthogonal
transformation the following condition is satisfied:
( ) ( ) ( )
( )
_
*
*
T
Q T Q Q T Q T
H
H H H
T T

(1.388)
We can show that ) (T H has the same principal directions of T , i.e. ) (T H and T are
coaxial tensors. To demonstrate this we can regard the components of T in the principal
space as:
( )
(
(
(

/
/
/

3
2
1
0 0
0 0
0 0
ij
T
(1.389)
Then the tensor function is given in terms of the principal values of T : ( )
3 2 1
, , / / / H H
and, the transformation of T is given by:
T
Q T Q T
*
(1.390)
Likewise, for the tensor function H :
( ) ( )
T
Q T Q T H H
*
(1.391)
If we take as the orthogonal tensor components:
( )
(
(
(

-
-
1 0 0
0 1 0
0 0 1
ij
Q
(1.392)
After having done the calculation for the matrices (1.391), we obtain:
(
(
(

H
H
H
H
H
(
(
(

H H H
H H H
H H H

(
(
(

H H - H -
H - H H -
H - H - H
H
33
22
11
*
33 23 13
23 22 12
13 12 11
33 23 13
23 22 12
13 12 11
*
0 0
0 0
0 0

(1.393)
To satisfy that H H
*
(isotropy), we conclude that 0
23 13 12
H H H . Therefore, ) (T H
and T have the same principal directions.
Once again we observe, a tensor function ) (T H . This tensor function is isotropic if and
only if it can be represented by the following linear transformation, Truesdell & Noll
(1965):
1 TENSORS

93
2
2 1 0
) ( T T 1 T d - d - d H H (1.394)
where
0
d ,
1
d ,
2
d are functions of the tensor T invariants or functions of the T
eigenvalues.
A brief demonstration follows. We can now consider the spectral representations of T and
H , respectively:
) 3 ( ) 3 (
3
) 2 ( ) 2 (
2
) 1 ( ) 1 (
1
3
1
) ( ) (

n n n n n n n n T / - / - / /

a
a a
a
(1.395)
) 3 ( ) 3 (
3
) 2 ( ) 2 (
2
) 1 ( ) 1 (
1
3
1
) ( ) (

n n n n n n n n c - c - c c H

a
a a
a
(1.396)
Note that T and H have the same principal directions
) (

i
n . Then, we can put the
following set of equations together:

/ - / - /
/ - / - /
- -
) 3 ( ) 3 ( 2
3
) 2 ( ) 2 ( 2
3
) 1 ( ) 1 ( 2
1
2
) 3 ( ) 3 (
3
) 2 ( ) 2 (
2
) 1 ( ) 1 (
1
) 3 ( ) 3 ( ) 2 ( ) 2 ( ) 1 ( ) 1 (



n n n n n n T
n n n n n n T
n n n n n n 1

(1.397)
Solving the set above, we obtain
) ( ) ( ) (

a a a
M n n = as a function of the tensor T , and we
obtain:
( )
( )
( )
) )( ( ) )( ( ) )( (
) )( ( ) )( ( ) )( (
) )( ( ) )( ( ) )( (
2 3 1 3
2
2 3 1 3
2 1
2 3 1 3
2 1 ) 3 (
3 2 1 2
2
3 2 1 2
3 1
3 2 1 2
3 1 ) 2 (
2 1 3 1
2
2 1 3 1
3 2
2 1 3 1
3 2 ) 1 (
/ - / / - /
-
/ - / / - /
/ - /
-
/ - / / - /
/ /

/ - / / - /
-
/ - / / - /
/ - /
-
/ - / / - /
/ /

/ - / / - /
-
/ - / / - /
/ - /
-
/ - / / - /
/ /

T
T 1 M
T
T 1 M
T
T 1 M
(1.398)
It is evident that, if we substitute the values of
) ( ) ( ) (

a a a
M n n = in equation (1.395) we
obtain: T T . Now, if we substitute the values of
) ( ) ( ) (

a a a
M n n = in equation (1.396),
we obtain:
2
2 1 0
) ( T T 1 T d - d - d H H (1.399)
where the coefficients
0
d ,
1
d , and
2
d are functions of the eigenvalues of T ,
) (
3 2 1
/ / / , and given by:
( ) ( ) ( )
) )( ( ) )( ( ) )( (
) )( ( ) )( ( ) )( (
) )( ( ) )( ( ) )( (
2 3 1 3
3
3 2 1 2
2
2 1 3 1
1
2
2 3 1 3
2 1 3
3 2 1 2
3 1 2
2 1 3 1
3 2 1
1
2 3 1 3
2 1 3
3 2 1 2
3 1 2
2 1 3 1
3 2 1
0
/ - / / - /
c
-
/ - / / - /
c
-
/ - / / - /
c
d
/ - / / - /
/ - / c
-
/ - / / - /
/ - / c
-
/ - / / - /
/ - / c
- d
/ - / / - /
/ / c
-
/ - / / - /
/ / c
-
/ - / / - /
/ / c
d

(1.400)
We can now show that if a tensor function ) (T H is given in (1.399), this tensor function is
isotropic:
NOTES ON CONTINUUM MECHANICS

94
( ) ( ) ( )
) (
*
2
*
2
*
1 0
2
2 1 0
2
2 1 0
*
T
T T 1 Q T Q Q T Q Q 1 Q
Q T T 1 Q Q T Q T
H
d - d - d d - d - d
d - d - d H H


T T T
T T

(1.401)
1.6.3 The Derivative of the Tensor-Valued Tensor Function
Firstly, we refer to a scalar-valued tensor function:
) (A H H
(1.402)
The partial derivative of ) (A H with respect to A is defined as:
)

( ,
j i
ij
e e
A
A

o
H o
H =
o
H o
A

(1.403)
where the comma denotes a partial derivative.
Then the second derivative of ) (A H becomes a fourth-order tensor:
)

( )

( ,
2 2
l k j i ijkl l k j i
kl ij
e e e e e e e e
A A
AA

o o
H o
H
o o
H o
D
A A

(1.404)
Let C and b be positive definite symmetric second-order tensors defined as:
T T
F F b F F C ; (1.405)
where F is an arbitrary second-order tensor with the restriction 0 ) ( > F det imposed on it.
We must also bear in mind that there is a scalar-valued isotropic tensor function,
( )
C C C
I I I I I I , , 1 1 , expressed in terms of the principal invariants of C , where
b C
I I ,
b C
I I I I ,
b C
I I I I I I . Next, we can find the partial derivative of 1 with respect to C , and
with respect to b . We must also verify that the following relation holds:
b F F
b C
, , 1 1
T
(1.406)
By applying the chain rule for derivative we obtain:
( )
C C C C
C
C
C
C
C
C
C C C
C
o
o
o
o
-
o
o
o
o
-
o
o
o
o

o
o

I I I
I I I
I I
I I
I
I
I I I I I I 1 1 1 1
1
, ,
,
(1.407)
Considering the partial derivatives of the principal invariants, we can state that:
1
o
o
C
C
I

2 1 - -
- - -
o
o
C C C C
C
C C C C
C
I I I I I I I
I I
T
1 1
1
C C C C
C
C C C C
C
I I I I I I I I I
I I I
T
- -
o
o
- - 2 1

(1.408)
Now, by substituting the following values 1
o
o
C
C
I
, C
C
C
C
-
o
o
1 I
I I
and
1 -

o
o
C
C
C
C
I I I
I I I

into the equation in (1.407), we obtain:
( )
1
,
-
o
o
- -
o
o
-
o
o
C C
C
C
C
C C
C
I I I
I I I
I
I I I
1 1 1
1 1 1
(1.409)
1 TENSORS

95
1
,
-
|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o
C C
C
C C
C
C C
C
I I I
I I I I I
I
I I I
1 1 1 1
1 1
(1.410)
Another way to express the relation (1.410) is by substituting 1
o
o
C
C
I
, C
C
C
C
-
o
o
1 I
I I

and 1
C C
C
C C
C
I I I
I I I
- -
o
o
2
into the equation in (1.407), thus:
2
, C C
C
C
C C
C
C
C
C C
C
|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o
-
o
o

I I I
I
I I I I I
I I
I I I
I
I I I
1 1 1 1 1 1
1 1
(1.411)
If we now consider the values 1
o
o
C
C
I
,
2 1 - -
-
o
o
C C
C
C C
C
I I I I I
I I
,
1 -

o
o
C
C
C
C
I I I
I I I
, in the
equation in (1.407), we obtain:
2 1
,
- -
|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o
-
|
|
.
|

\
|
o
o
C C
C
C
C
C
C
C C
C
I I I
I I
I I I
I I I
I I
I I I
1 1 1 1
1 1
(1.412)
If we now observe both
b C
I I ,
b C
I I I I ,
b C
I I I I I I , and the equation in (1.410), we can
draw the conclusion that:
1
,
-
o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o
b b
b
b b
b
b b
b
I I I
I I I I I
I
I I I
1 1 1 1
1 1
(1.413)
Using the equation in (1.410), the equation
T
F F
C

,
1 becomes:
T T T T
I I I
I I I I I
I
I I I
F C F F C F F F F F
C
C C
C
C C
C

-
o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o

1
,
1 1 1 1
1 1
(1.414)
Then, if we observe that:
b F F F F
T T
1
2
b b b F F F F F C F
F F C



T T T
T

(1.415)
b b F F b F F F C F
F b F C


- - - -
- - -


1 1 1 1
1 1 1
T T
(1.416)
The equation (1.414) can be rewritten as:
b b b
b b b b F F
C
C C
C
C C
C
C C
C
C C
C


(
(

o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o

o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o

-
-
1
1 2
,
I I I
I I I I I
I
I I I
I I I
I I I I I
I
I I I
T
1 1 1 1
1 1 1 1
1
1

(1.417)
In light of the equation in (1.413) and (1.417), we can draw the conclusion that:
b b
b
b b
b
b b
C
b b
b b b F F
, ,
1
,
1 1
1 1 1 1
1



(
(

o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o

-
I I I
I I I I I
I
I I I
T
1

(1.418)
NOTES ON CONTINUUM MECHANICS

96
which indicates that
b
, 1 and b are coaxial tensors.
Once again, we can observe C given by the equation in (1.405). Next, we can evaluate the
derivative of the scalar-valued tensor function, ) (C 1 1 , with respect to the tensor F :
( )
( )
kl
ij
ij
kl
F
C
C o
o
o
o

o
o
o
o

o
o

1
1
1 1
1
F F
F
C
C F
C
, ,
notation indicial
:
(1.419)
The derivative of tensor C with respect to F is evaluated as follows:
( )
( ) ( )
ki jl kj il
qi jl qk qj il qk
kl
qj
qi qj
kl
qi
kl
qj qi
kl
ij
F F
F F
F
F
F F
F
F
F
F F
F
C
c c
c c c c
-
-
o
o
-
o
o

o
o

o
o

(1.420)
Then, by substituting (1.420) into (1.419), we obtain:
( ) ( )
il
ki
lj
kj
ki jl kj il
ij
kl
C
F
C
F
F F
C
o
o
-
o
o

-
o
o

1 1
c c
1
1
F
,

(1.421)
Due to the symmetry of C , i.e.
jl lj
C C , we can draw the conclusion that:
( )
o
o

o
o

kj
jl
kj
lj
kl
F
C
F
C
1 1
1 2 2 ,
F
C C F
F F , 2 , 2 , 1 1 1
T

(1.422)
Now, suppose that C is given by the equation
2
U U U U U
T
C , where U is a
symmetric second-order tensor. To find
U ,
) (C 1 we can use the same equation as in
(1.422), i.e.:
C C
, 2 , 2 , 1 1 1 U U
U

(1.423)
Therefore, we can draw the conclusion that
C
, 1 and U are coaxial tensors.
Let A be a symmetric second-order tensor, and ) (A 1 1 be a scalar-valued tensor
function. The following relationships hold:
T
T
T
b b b b b b
b b
b b b
b b b
b
b
b
-







and for
for
for
A
A
A
A A
A A
A
A
, ,
, 2 , 2 ,
, 2 ,
, 2 ,
1 1
1 1 1
1 1
1 1

(1.424)
1.7 The Voigt Notation
When dealing with symmetric tensors, it may be advantageous to just work with the
independent components. For example, a symmetric second-order tensor has 6
1 TENSORS

97
independent components, so, it is possible to represent these components by a column
matrix as follows:
{
(
(
(
(
(
(
(
(


(
(
(
(
(
(

13
23
12
33
22
11
33 23 13
23 22 12
13 12 11
T
T
T
T
T
T
T T T
T T T
T T T
T T
Voigt
ij

(1.425)
This representation is called the Voigt Notation. It is also possible to represent a second-
order tensor as:
{
(
(
(
(
(
(
(
(


(
(
(
(
(
(

13
23
12
33
22
11
33 23 13
23 22 12
13 12 11
2
2
2
E
E
E
E
E
E
E E E
E E E
E E E
E E
Voigt
ij

(1.426)
As we have seen before, a fourth-order tensor, C , that presents minor symmetry, i.e.
jilk ijlk jikl ijkl
C C C C , has 36 6 6 independent components. Note that, due to the
symmetry of ) (ij we have 6 independent components, and due to the symmetry of ) (kl
we have 6 independent components. In Voigt Notation we can represent these
components in a 6 -by- 6 matrix as:
[ ]
(
(
(
(
(
(
(
(

1313 1323 1312 1333 1322 1311


2313 2323 2312 2333 2322 2311
1213 1223 1212 1233 1222 1211
3313 3323 3312 3333 3322 3311
2213 2223 2212 2233 2222 2211
1113 1123 1112 1133 1122 1111
C C C C C C
C C C C C C
C C C C C C
C C C C C C
C C C C C C
C C C C C C
C
(1.427)
In addition to minor symmetry the tensor also has major symmetry, i.e.
klij ijkl
C C , and the
number of independent components have reduced to 21. One can easily memorize the
order of the components in the matrix [ ] C if we consider the order of the second-order
tensor in Voigt Notation, i.e.:
[ ] ) 13 ( ) 23 ( ) 12 ( ) 33 ( ) 22 ( ) 11 (
) 13 (
) 23 (
) 12 (
) 33 (
) 22 (
) 11 (
(
(
(
(
(
(
(
(

(1.428)
1.7.1 The Unit Tensors in Voigt Notation
The second-order unit tensor is represented in the Voigt notation as:
NOTES ON CONTINUUM MECHANICS

98
{
(
(
(
(
(
(
(
(


(
(
(

=
0
0
0
1
1
1

1 0 0
0 1 0
0 0 1
Voigt
ij
1 c (1.429)
In the subsection 1.5.2.5.1 Unit Tensors we have defined three fourth-order unit tensors,
namely,
j ik ijk
c c I ,
jk i ijk
c c

I and
k ij ijk
c c I , among which only
k ij ijk
c c I is a
symmetric tensor. The representation of
k ij ijk
c c I in Voigt notation can be evaluated by
observing how a symmetric fourth-order tensor is represented in (1.427), thus:
[ ]
(
(
(
(
(
(
(
(


0 0 0 0 0 0
0 0 0 0 0 0
0 0 0 0 0 0
0 0 0 1 1 1
0 0 0 1 1 1
0 0 0 1 1 1
I
Voigt
k ij ijk
c c I (1.430)
where 1
11 11 1111
c c I , 1
22 11 1122
c c I , and so on.
The components of a fourth-order unit tensor,
sym
I , are represented by
( )
jk i j ik ijk
c c c c

-
2
1
I , which in Voigt notation becomes:
[ ]
(
(
(
(
(
(
(
(

(
(
(
(
(
(
(
(


2
1
2
1
2
1
1313 1323 1312 1333 1322 1311
2313 2323 2312 2333 2322 2311
1213 1223 1212 1233 1222 1211
3313 3323 3312 3333 3322 3311
2213 2223 2212 2233 2222 2211
1113 1123 1112 1133 1122 1111
0 0 0 0 0
0 0 0 0 0
0 0 0 0 0
0 0 0 1 0 0
0 0 0 0 1 0
0 0 0 0 0 1
I I I I I I
I I I I I I
I I I I I I
I I I I I I
I I I I I I
I I I I I I
I I
Voigt
ijk
(1.431)
and the inverse of the equation in (1.431) becomes:
[ ]
(
(
(
(
(
(
(
(

-
2 0 0 0 0 0
0 2 0 0 0 0
0 0 2 0 0 0
0 0 0 1 0 0
0 0 0 0 1 0
0 0 0 0 0 1
1
I (1.432)

1.7.2 The Scalar Product in Voigt Notation
The dot product between a symmetric second-order tensor, T , and a vector n
,
, is given by
n T b
,
,
where the components of b
,
can be evaluated as follows:
1 TENSORS

99

- -
- -
- -

(
(
(

(
(
(

(
(
(

3 33 2 23 1 13 3
3 23 2 22 1 12 2
3 13 2 12 1 11 1
3
2
1
33 23 13
23 22 12
13 12 11
3
2
1

n T n T n T b
n T n T n T b
n T n T n T b
n
n
n
T T T
T T T
T T T
b
b
b

(1.433)
By observing how a second-order tensor is presented in Voigt notation, as in (1.425), the
scalar product (1.433) can be represented in the Voigt notation as:
[ ]

(
(
(
(
(
(
(
(

(
(
(

(
(
(

13
23
12
33
22
11
1 2 3
3 1 2
3 2 1
3
2
1

0 0 0
0 0 0
0 0 0
T
T
T
T
T
T
n n n
n n n
n n n
b
b
b
_
T
N
{ [ ] { T N
T
b
(1.434)
1.7.3 The Component Transformation Law in Voigt Notation
The component transformation law for a second-order tensor is defined as:
jl ik kl ij
a a T T

(1.435)
or in matrix form:
T
a a a
a a a
a a a
a a a
a a a
a a a
(
(
(

(
(
(

(
(
(

(
(
(




33 32 31
23 22 21
13 12 11
33 23 13
23 22 12
13 12 11
33 32 31
23 22 21
13 12 11
33 23 13
23 22 12
13 12 11

T T T
T T T
T T T
T T T
T T T
T T T
(1.436)
By multiplying the matrices and by rearranging the result in Voigt notation we obtain:
{ [ ]{ T M T
(1.437)
where:
{ {
(
(
(
(
(
(
(
(

(
(
(
(
(
(
(
(


13
23
12
33
22
11
13
23
12
33
22
11
;
T
T
T
T
T
T
T
T
T
T
T
T
T T
(1.438)
and [ ] M is the transformation matrix for the second-order tensor components in Voigt
Notation. The matrix [ ] M is given by:
[ ]
( ) ( ) ( )
( ) ( ) ( )
( ) ( ) ( )(
(
(
(
(
(
(
(

- - -
- - -
- - -

13 31 11 33 13 32 12 33 11 32 12 31 13 33 12 32 11 31
23 31 21 33 23 32 22 33 21 32 22 31 23 33 22 32 21 31
23 11 21 13 23 12 22 13 21 12 22 11 23 13 12 22 11 21
33 31 33 32 32 31
2
33
2
32
2
31
23 21 23 22 22 21
2
23
2
22
2
21
13 11 13 12 12 11
2
13
2
12
2
11
2 2 2
2 2 2
2 2 2
a a a a a a a a a a a a a a a a a a
a a a a a a a a a a a a a a a a a a
a a a a a a a a a a a a a a a a a a
a a a a a a a a a
a a a a a a a a a
a a a a a a a a a
M

(1.439)
If the representation of tensor components is shown in (1.426), equation (1.436) in Voigt
Notation becomes:
NOTES ON CONTINUUM MECHANICS

100
{ [ ]{ E N E (1.440)
where
[ ]
( ) ( ) ( )
( ) ( ) ( )
( ) ( ) ( )(
(
(
(
(
(
(
(

- - -
- - -
- - -

13 31 11 33 13 32 12 33 11 32 12 31 13 33 12 32 11 31
23 31 21 33 23 32 22 33 21 32 22 31 23 33 22 32 21 31
23 11 21 13 23 12 22 13 21 12 22 11 23 13 12 22 11 21
33 31 33 32 32 31
2
33
2
32
2
31
23 21 23 22 22 21
2
23
2
22
2
21
13 11 13 12 12 11
2
13
2
12
2
11
2 2 2
2 2 2
2 2 2
a a a a a a a a a a a a a a a a a a
a a a a a a a a a a a a a a a a a a
a a a a a a a a a a a a a a a a a a
a a a a a a a a a
a a a a a a a a a
a a a a a a a a a
N

(1.441)
The matrices (1.439) and (1.441) are not orthogonal matrices, i.e. [ ] [ ]
T
M M
-1
and
[ ] [ ]
T
N N
-1
. However, it is possible to show that [ ] [ ]
T
N M
-1
.
1.7.4 Spectral Representation in Voigt Notation
Regarding the spectral representation of a symmetric tensor T :
A T A T

form Matricial
3
1
) ( ) (

T
a
a a
a
n n T T (1.442)
where A is the transformation matrix between the original set and the principal space,
made up of the eigenvectors
) (

a
n . The above equation can be rewritten in terms of
components as follows:
A A A A A A
(
(
(

-
(
(
(

-
(
(
(

(
(
(

3
2
1
33 23 13
23 22 12
13 12 11
0 0
0 0 0
0 0 0
0 0 0
0 0
0 0 0
0 0 0
0 0 0
0 0

T
T
T
T T T
T T T
T T T
T T T

(1.443)
or
(
(
(

-
(
(
(

-
(
(
(

(
(
(

2
33 33 32 33 31
33 32
2
32 32 31
33 31 32 31
2
31
3
2
23 23 22 23 21
23 22
2
22 22 21
23 21 22 21
2
21
2
2
13 13 12 13 11
13 12
2
12 12 11
13 11 12 11
2
11
1
33 23 13
23 22 12
13 12 11
a a a a a
a a a a a
a a a a a
a a a a a
a a a a a
a a a a a
a a a a a
a a a a a
a a a a a
T
T T
T T T
T T T
T T T
(1.444)
By regarding how second-order tensors are presented in Voigt Notation as in (1.438), the
spectral representation of a second-order tensor in Voigt notation becomes:
{
(
(
(
(
(
(
(
(

-
(
(
(
(
(
(
(
(

-
(
(
(
(
(
(
(
(

(
(
(
(
(
(
(
(

33 31
33 32
32 31
2
33
2
32
2
31
3
23 21
23 22
22 21
2
23
2
22
2
21
2
13 11
13 12
12 11
2
13
2
12
2
11
1
13
23
12
33
22
11
a a
a a
a a
a
a
a
a a
a a
a a
a
a
a
a a
a a
a a
a
a
a
T T T
T
T
T
T
T
T
T
(1.445)
1 TENSORS

101
1.7.5 Deviatoric Tensor Components in Voigt Notation
Observing the components of the deviatoric tensor:
(
(
(

- -
- -
- -

) 2 (
) 2 (
) 2 (
22 11 33 3
1
23 13
23 33 11 22
3
1
12
13 12 33 22 11 3
1
T T T T T
T T T T T
T T T T T
T
dev
ij

(1.446)
dev
ij
T in Voigt notation is given by:
(
(
(
(
(
(
(
(

(
(
(
(
(
(
(
(

- -
- -
- -

(
(
(
(
(
(
(
(

13
23
12
33
22
11
13
23
12
33
22
11
3 0 0 0 0 0
0 3 0 0 0 0
0 0 3 0 0 0
0 0 0 2 1 1
0 0 0 1 2 1
0 0 0 1 1 2
3
1
T
T
T
T
T
T
T
T
T
T
T
T
dev
dev
dev
dev
dev
dev

(1.447)
Problem 1.40: Let ) , ( t x
,
T be a symmetric second-order tensor, which is expressed in
terms of the position ) ( x
,
and time ) (t . Also, bear in mind that the tensor components,
along direction
3
x , are equal to zero, i.e. 0
33 23 13
T T T .
NOTE: In the next section we will define ) , ( t x
,
T as a field tensor, i.e. the value of T
depends on position and time. As we will see later, if the tensor is independent of any one
direction at all points ) ( x
,
, e.g. if ) , ( t x
,
T is independent of the
3
x -direction, (see Figure
1.31), the problem becomes a two-dimensional problem (plane state) so that the problem is
greatly simplified.














Figure 1.31: A two-dimensional problem (2D).

a) Obtain
11
T ,
22
T ,
12
T in the new reference system (
2 1
x x - ) defined in Figure 1.32.
b) Obtain the value of 0 so that 0 corresponds to the principal direction of T , and also
find an equation for the principal values of T .
c) Evaluate the values of
ij
T , ) 2 , 1 , ( j i , when 1
11
T , 2
22
T , 4
12
- T and 45 0 .
Also, obtain the principal values and principal directions.
d) Draw a graph that shows the relationship between 0 and components
11
T ,
22
T and
12
T , and in which the angle varies from 0 to 360 .
Hint: Use the Voigt Notation, and express the results in terms of 0 2 .

1
x

2
x

3
x

22
T

11
T
12
T

12
T

1
x

2
x

22
T

11
T
12
T

12
T

11
T

22
T
2D
NOTES ON CONTINUUM MECHANICS

102









Figure 1.32: A two-dimensional problem (2D).
Solution:
a) Here we can apply the transformation law obtained in (1.437), which after removing
rows and columns associated with the
3
x -direction becomes:
(
(
(

(
(
(

(
(
(

12
22
11
21 12 22 11 12 22 11 21
22 21
2
22
2
21
12 11
2
12
2
11
12
22
11
2
2
T
T
T
T
T
T
a a a a a a a a
a a a a
a a a a
(1.448)

















Figure 1.33: Transformation law for (2D) tensor components.

The transformation matrix,
ij
a , in the plane, can be evaluated in terms of a single
parameter, 0 :
(
(
(

-
(
(
(

1 0 0
0 cos sin
0 sin cos
33 32 31
23 22 21
13 12 11
0 0
0 0
a a a
a a a
a a a
a
ij

(1.449)
By substituting the matrix components
ij
a given in (1.449) into (1.448) we obtain:
(
(
(

(
(
(

-
-
(
(
(

-
12
22
11
2 2
2 2
2 2
12
22
11

sin cos cos cos sin
cos sin 2 cos sin
cos 2 sin cos
sin
sin
T
T
T
T
T
T
0 0 0 0 0 0
0 0 0 0
0 0 0 0

(1.450)
Making use of the following trigonometric identities, 0 0 0 2 cos 2 sin sin ,
0 0 0 2 cos sin cos
2 2
- ,
2
2 cos 1
sin
2
0
0
-
,
2
2 cos 1
cos
2
0
0
-
, (1.450) becomes:
P

22
T

11
T

1
x

2
x

12
T

11
T

12
T

22
T
P
0

T
A T A T
A T A T
T


1
x
1
x

2
x

22
T

11
T
12
T

12
T

11
T

22
T
0

1
x

2
x

2
x
1
x
1 TENSORS

103
(
(
(

(
(
(
(
(
(
(

|
.
|

\
|
|
.
|

\
|
-
- |
.
|

\
| -
|
.
|

\
| -
|
.
|

\
| -
|
.
|

\
| -

(
(
(

12
22
11
12
22
11

2 cos
2
2
2
2
2
2
2 cos 1
2
2 cos 1
2
2
2 cos 1
2
2 cos 1
sin sin
sin
sin
T
T
T
T
T
T
0
0 0
0
0 0
0
0 0

Explicitly, the above components are given by:

- |
.
|

\
|
- |
.
|

\
|
-
- |
.
|

\
| -
- |
.
|

\
| -

- |
.
|

\
| -
- |
.
|

\
| -

0
0 0
0
0 0
0
0 0
2 cos
2
2
2
2
2
2
2 cos 1
2
2 cos 1
2
2
2 cos 1
2
2 cos 1
12 22 11 12
12 22 11 22
12 22 11 11
sin sin
sin
sin
T T T T
T T T T
T T T T

Reordering the previous equation, we obtain:

- |
.
|

\
| -
-
|
.
|

\
| -
- |
.
|

\
| -

- |
.
|

\
| -
- |
.
|

\
| -

-
0 0
0 0
0 0
2 cos 2
2
2 2 cos
2 2
2 2 cos
2 2
12
22 11
12
12
22 11 22 11
22
12
22 11 22 11
11
sin
sin
sin
T
T T
T
T
T T T T
T
T
T T T T
T
(1.451)
b) Recalling that the principal directions are characterized by the lack of any tangential
components, i.e. 0
ij
T if j i , in order to find the principal directions in the plane, we let
0
12
T , hence:
22 11
12
22 11
12
12
22 11
12
22 11
12
2
) 2 (
2
2 cos
2
2 cos 2
2
0 2 cos 2
2
sin
sin sin
T T
T
tg
T T
T
T
T T
T
T T
T
-

|
.
|

\
| -
- |
.
|

\
| -
-

0
0
0
0 0 0 0

Then, the angle corresponding to the principal direction is:
|
|
.
|

\
|
-

22 11
12
2
2
1
T T
T
arctg 0
(1.452)
To find the principal values (eigenvalues) we must solve the following characteristic
equation:
( ) ( ) 0 0
2
12 22 11 22 11
2
22 12
12 11
- - - -
-
-
T T T T T T T
T T T
T T T

And by evaluating the quadratic equation we obtain:
( ) [ ] ( ) [ ] ( )
( ) [ ] ( )
4
4
2
) 1 ( 2
) 1 ( 4
2
12 22 11
2
22 11 22 11
2
12 22 11
2
22 11 22 11
) 2 , 1 (
T T T T T T T
T T T T T T T
T
- - -

- - - - - - -


By rearranging the above equation we obtain the principal values for the two-dimensional
case as:
NOTES ON CONTINUUM MECHANICS

104
2
12
2
22 11 22 11
) 2 , 1 (
2 2
T
T T T T
T - |
.
|

\
| -


(1.453)
c) We directly apply equation (1.451) to evaluate the values of the components
ij
T ,
) 2 , 1 , ( j i , where 1
11
T , 2
22
T , 4
12
- T and 45 0 , i.e.:

|
.
|

\
| -
-
|
.
|

\
| -
- |
.
|

\
| -

- - |
.
|

\
| -
- |
.
|

\
| -

-
-
5 . 0 90 cos 90
2
2 1
5 . 5 90 90 cos
2
2 1
2
2 1
5 . 2 90 4 90 cos
2
2 1
2
2 1
4 sin
sin 4
sin
12
22
11
T
T
T

And the angle corresponding to the principal direction is:
( ) 4375 . 41
2 1
) 4 ( 2 2
2
1
22 11
12
0
-
-

|
|
.
|

\
|
-

T T
T
arctg 0
The principal values of ) , ( t x
,
T can be evaluated as follows:

- |
.
|

\
| -

5311 . 2
5311 . 5

2 2
2
1 2
12
2
22 11 22 11
) 2 , 1 (
T
T
T
T T T T
T
d) By referring to equation in (1.451) and by varying 0 from 0 to 360 , we can obtain
different values of
11
T ,
22
T ,
12
T , which are illustrated in the following graph:




























-6
-4
-2
0
2
4
6
8
0 50 100 150 200 250 300 350
0
5311 . 5
1
T

22
T


12
T


11
T


22
T

12
T

11
T
437 . 41 0

2
T

0311 . 4
max

S
T

C
o
m
p
o
n
e
n
t
s


1
x
437 . 131 0

2
o

437 . 86 0
45

1
x

1
x
5311 . 2
2
- T

1
T

1 TENSORS

105
1.8 Tensor Fields
A tensor field indicates how the tensor, ) , ( t x
,
T , varies in space ( x
,
) and time ( t ). In this
section, we regard the tensor field as a differentiable function of position and time. For
more information about it, we need to define some operators, e.g. gradient, divergence, curl,
which we can use as indicators of how these fields vary in space.
A tensor field which is independent of time is called a stationary or steady-state tensor
field, i.e. ) ( x
,
T T . However, if the field is only dependent on t then it is said to be
homogeneous or uniform. That is, ) (t T has the same value at every x
,
position.
Tensor fields can be classified according to their order as: scalar, vector, second-order
tensor fields, etc. As an example of a scalar field we can quote temperature ) , ( t T x
,
and in
Figure 1.34(a) we can see temperature distribution over time
1
t t . Then, as an example of
a vector field we can quote velocity ) , ( t x v
, ,
and Figure 1.34(b) shows velocity distribution,
in which each point is associated with a vector v
,
over time
1
t t .












Figure 1.34: Examples of tensor fields.
Scalar Field

) , ( t x
,
c c (1.454)
Vector Field
Tensorial notation
) , ( t x
, , ,
v v
Indicial notation ) , ( t
i i
x
,
v v
(1.455)
Second-Order Tensor Field
Tensorial notation ) , ( t x
,
T T
Indicial notation ) , ( t
ij ij
x
,
T T
(1.456)

2
x

1
x

3
x

5
T

6
T
) , (
1
) 4 (
4
t T x
,


3
T

7
T

2
T

1
T

8
T
a) Scalar field

2
x

1
x

3
x
b) Vector field

1
t t
1
t t ) , (
1
t x v
, ,

NOTES ON CONTINUUM MECHANICS

106
1.8.1 Scalar Fields
The next analysis is carry out with reference to a stationary scalar field, i.e. ) ( x
,
c c , with
continuous values of
1
/ x o oc ,
2
/ x o oc and
3
/ x o oc . Then, observe that the value of the
scalar function at point ) ( x
,
is ) ( x
,
c , and if we observe a second point located at ) ( x x
, ,
d - ,
the total derivative (differential) of the function c is defined as:
c c c
c c c
d x x x dx x dx x dx x
d d
= - - - -
= - -
) , , ( ) , , (
) ( ) (
3 2 1 3 3 2 2 1 1
x x x
, , ,

(1.457)
For any continuous function ) , , (
3 2 1
x x x c , c d is linearly related to
1
dx ,
2
dx ,
3
dx . This
linear relationship can be evaluated by the chain rule of differentiation as:
i i
dx d
dx
x
dx
x
dx
x
d
,
3
3
2
2
1
1
c c
c c c
c

o
o
-
o
o
-
o
o


(1.458)
The differentiation of the components of a tensor, with respect to coordinates
i
x , is
expressed by the differential operator:
i
i
x
, - =
o
- o

(1.459)
1.8.2 Gradient
The gradient of a scalar field
The gradient c
x
,
V or c grad is defined as:
x
x x
,
, ,
d d c c c V V (1.460)
where the operator
x
,
V is known as the Nabla symbol. Expressing the equation (1.460) in
the Cartesian basis we obtain:

o
o
-
o
o
-
o
o
3
3
2
2
1
1
dx
x
dx
x
dx
x
c c c

[ ] [ ]
3 3 2 2 1 1 3
3
2
2
1
1

) (

) (

) (

) (

) (

) ( e e e e e e dx dx dx
x x x
- - - - c c c
x x x
, , ,
V V V
(1.461)
Evaluating the above scalar product we find:
3
3
2
2
1
1
3
3
2
2
1
1
) ( ) ( ) ( dx dx dx dx
x
dx
x
dx
x
x x x
c c c
c c c
x x x
, , ,
V V V - -
o
o
-
o
o
-
o
o

(1.462)
Therefore, we can draw the conclusion that the c
x
,
V components in the Cartesian basis
are:
3
3
2
2
1
1
) ( ; ) ( ; ) (
x x x o
o
=
o
o
=
o
o
=
c
c
c
c
c
c
x x x
, , ,
V V V
(1.463)
Hence, the gradient in terms of components is defined as such:

3
3
2
2
1
1

e e e
x x x o
o
-
o
o
-
o
o

c c c
c
x
,
V
(1.464)
The Nabla symbol
x
,
V is defined as:
1 TENSORS

107
i i i
i
x
e e

,

o =
o
o

x
,
V
Nabla symbol (1.465)
The geometric meaning of c
x
,
V
The direction of c
x
,
V is normal to the equiscalar surface, i.e. it is perpendicular to
the isosurface const c . The direction of c
x
,
V points to the direction where c is
increasing the most, (see Figure 1.35).
The magnitude of c
x
,
V is the rate of change of c , i.e. the gradient of c .
The normal vector to this surface is obtained as follows:
c
c
x
x
,
,
V
V
n


(1.466)
The surface const c , called the surface level, or isosurface or equiscalar surface, is the
surface formed by points which all have the same value of c , so, if we move along the
level surface the values of the function do not change.
The gradient of a vector field ) ( x
, ,
v :
v v
, ,
,
x
V = ) ( grad (1.467)
Using the definition of
x
,
V , given in (1.465), the gradient of the vector field becomes:
j i j i j j i i j
j
i i
x
e e e e e
e
v

)

(
, ,

o
o
v v
v ,
,
x
V
(1.468)










Figure 1.35: Gradient of c .
Therefore, we can define the gradient of a tensor field )) , ( ( t x
,
- in the Cartesian basis as:
j
j
x
e

) (
) (
o
- o
-
x
,
V

Gradient of a tensor field in the
Cartesian basis
(1.469)
As noted, the gradient of a vector field becomes a second-order tensor field, whose
components are:

1
c const c

2
c const c

3
c const c
c
x
,
V
n



3 2 1
c c c > >
NOTES ON CONTINUUM MECHANICS

108
(
(
(
(
(
(
(

o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o

o
o
=
3
3
2
3
1
3
3
2
2
2
1
2
3
1
2
1
1
1
,
x x x
x x x
x x x
x
j
i
j i
v v v
v v v
v v v
v
v (1.470)
The gradient of a second-order tensor field ) ( x
,
T :
k j i k ij k
k
j i ij
T
x
T
e e e e
e e
T

)

(
,

o
o

x
,
V (1.471)
and its components are represented by:
( )
k ij ijk
T
,
= T
x
,
V
(1.472)
Problem 1.41: Find the gradient of the function
2 1
) cos( ) , (
1 2 1
x x
x x x f exp - at the point
) 1 , 0 (
2 1
x x .
Solution: By definition, the gradient of a scalar function is given by:
2
2
1
1

e e
x
f
x
f
f
o
o
-
o
o

x
,
V
where:
2 1
2 1
1
) sin(
x x
x x
x
f
exp - -
o
o
;
2 1
1
2
x x
x
x
f
exp
o
o

[ ] [ ]
2 1 1 2 1 2 1

) sin( ) , (
2 1 2 1
e e
x x x x
x x x x x f exp exp - - -
x
,
V [ ] [ ]
1 2 1

2

1 ) 1 , 0 ( e e e - f
x
,
V
Problem 1.42: Let ) ( x
, ,
u be a stationary vector field. a) Obtain the components of the
differential u
,
d . b) Now, consider that ) ( x
, ,
u represents a displacement field, and is
independent of
3
x . With these conditions, graphically illustrate the displacement field in
the differential area element
2 1
dx dx .
Solution: According to the differential and gradient definitions, it holds that:











Thus, the components are defined as:
(
(
(

(
(
(
(
(
(
(

o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o

(
(
(

o
o

3
2
1
3
3
2
3
1
3
3
2
2
2
1
2
3
1
2
1
1
1
3
2
1

dx
dx
dx
x x x
x x x
x x x
d
d
d
dx
x
d
j
j
i
i
u u u
u u u
u u u
u
u
u
u
u
) ( x
, ,
u
x
,
d

1
x

2
x

3
x
) ( x x
, , ,
d - u
x
,

x x
, ,
d -

x
x x x
x
,
, ,
, , , , , ,
,
d d
d d

- - =
u u
u u u
V
) ( ) (

1 TENSORS

109
or:

o
o
-
o
o
-
o
o

o
o
-
o
o
-
o
o

o
o
-
o
o
-
o
o

3
3
3
2
2
3
1
1
3
3
3
3
2
2
2
2
1
1
2
2
3
3
1
2
2
1
1
1
1
1
dx
x
dx
x
dx
x
d
dx
x
dx
x
dx
x
d
dx
x
dx
x
dx
x
d
u u u
u
u u u
u
u u u
u

with

- - - -
- - - -
- - - -
) , , ( ) , , (
) , , ( ) , , (
) , , ( ) , , (
3 2 1 3 3 3 2 2 1 1 3 3
3 2 1 2 3 3 2 2 1 1 2 2
3 2 1 1 3 3 2 2 1 1 1 1
x x x dx x dx x dx x d
x x x dx x dx x dx x d
x x x dx x dx x dx x d
u u u
u u u
u u u

As the field is independent of
3
x , the displacement field in the differential area element is
defined as:

o
o
-
o
o
- - -
o
o
-
o
o
- - -
2
2
2
1
1
2
2 1 2 2 2 1 1 2 2
2
2
1
1
1
1
2 1 1 2 2 1 1 1 1
) , ( ) , (
) , ( ) , (
dx
x
dx
x
x x dx x dx x d
dx
x
dx
x
x x dx x dx x d
u u
u u u
u u
u u u



or:

o
o
-
o
o
- - -
o
o
-
o
o
- - -
2
2
2
1
1
2
2 1 2 2 2 1 1 2
2
2
1
1
1
1
2 1 1 2 2 1 1 1
) , ( ) , (
) , ( ) , (
dx
x
dx
x
x x dx x dx x
dx
x
dx
x
x x dx x dx x
u u
u u
u u
u u

Note that the above equation is equivalent to the Taylor series expansion taking into
account only up to linear terms. The representation of the displacement field in the
differential area element is shown in Figure 1.36.




















NOTES ON CONTINUUM MECHANICS

110















































Figure 1.36: Displacement field in the differential area element.



2
2
1
dx
x o
ou


1
1
2
dx
x o
ou


2
dx

2
u
A
O

1
dx
B
A

2
dx
B

1
u

1
1
1
1
dx
x o
o
-
u
u
O

1
dx

1 1
, u x

2 2
, u x
B
A
A

2
2
2
2
dx
x o
o
-
u
u
B
O
+
) (
1
u
) , (
2 1
x x ) , (
2 1 1
x dx x -
) , (
2 2 1
dx x x -
) , (
2 2 1 1
dx x dx x - -

1
x

2
x
u
,
d
) (
2
u

2
2
2
1
1
2
2
dx
x
dx
x o
o
-
o
o
-
u u
u

2
2
1
1
1
1
1
dx
x
dx
x o
o
-
o
o
-
u u
u

2
2
2
2
dx
x o
o
-
u
u

2
2
1
1
dx
x o
o
-
u
u

1
1
1
1
dx
x o
o
-
u
u

1
1
2
2
dx
x o
o
-
u
u

1
dx

2
dx

_




=
1 TENSORS

111
1.8.3 Divergence
The divergence of a vector field, ) ( x
, ,
v , is denoted as follows:
v v
, ,
,
=
x
V ) ( div (1.473)
which by definition is:
) ( ) ( v 1 v v v
, , , ,
, , ,
x x x
V V V Tr div = : (1.474)
Then:
[ ] [ ]
3
3
2
2
1
1
, , ,

x x x
k k jl ik kl j i l k kl j i j i
o
o
-
o
o
-
o
o


v v v
v v v c c c c e e e e 1 v v : :
, ,
, ,
x x
V V
(1.475)
or
[ ] [ ]
[ ]
[ ]
[ ]
k
k
i i
k i k i
k i lj kl j i
l k kl j i j i
x
e
e
e e
e e
e e e e 1 v v




,
,
,

o
o


v
v
v
v
c c
c : :
, ,
, ,
x x
V V
(1.476)
Which we can use to insert the following operator into the Cartesian basis:
k
k
x
e

) (
) (
o
- o
-
x
,
V
Divergence of ) (- in Cartesian
basis
(1.477)
We can also verify that, when divergence is applied to a tensor field its rank decreases by
one order.
Divergence of a second-order tensor field ) ( x
,
T
The divergence of a second-order tensor field T is denoted by 1 T T :
x x
, ,
V V , which
becomes a vector:
i k ik
i jk
k
ij
k
k
j i ij
x
T
x
T
e
e
e
e e
T T

)

(
,
T
div

o
o

o
o
=
c
x
,
V

(1.478)
NOTE: In this text book, when dealing with gradient or divergence of a tensor field, e.g. v
,
,
x
V
(the gradient of the vector field), T
x
,
V (the gradient of a second-order tensor field), T
x
,
V
(divergence of a second-order tensor field), this does not indicate that we are making a
tensor operation between a vector and a tensor, i.e. ) ( ) ( v v
,
,
,
, ,

x x
V V , ) ( ) ( T T
x x
, ,
,
V V
and ) ( ) ( T T
x x
, ,
,
V V and so on. In this textbook,
x
,
V is an operator which must be
applied to the entire tensor field, so, the tensor must be inside the operator, (see equations
(1.477) and (1.469)). Nevertheless, it is possible to relate v
,
,
x
V , T
x
,
V or T
x
,
V to tensor
operations between tensors, and it is easy to show that:
NOTES ON CONTINUUM MECHANICS

112

) ( ) ( ) ( ) (
) ( ) (
) ( ) (
T
T T T
T T
v v



x x x
x x
x x
, , ,
, ,
, ,
, ,
,
,
, ,
V V V
V V
V V


(1.479)
Once the Nabla symbol is defined we introduce the Laplacian operator
2
V as:
kk k k
i i
ij
j i
i j
j i
x x x
x x x x x x
, , ,

2
3
2
2
2
2
2
1
2
2
2
2
o o o
o
o
-
o
o
-
o
o
=
o o
o

o
o
o
o

|
|
.
|

\
|
|
|
.
|

\
|
o
o
o
o

x
x x x
,
, , ,
V
V V V c e e

(1.480)
Then, the vector Laplacian of a vector field, ) ( x
, ,
v , is given by:
[ ] [ ]
kk i
i i
components
,
2 2
) ( ) ( v v v v v
, , , ,
, , , , , ,
x x x x x x
V V V V V V (1.481)
Problem 1.43: Let a
,
and b
,
be vectors. Show that the following identity
b a b a
,
,
,
,
, , ,
- -
x x x
V V V ) ( holds.
Solution:
Observing that
j j
e a

a
,
,
k k
e b

b
,
,
i
i
x o
o
e

x
,
V , we can express ) ( b a
,
,
,
-
x
V as:
b a e e e e e
e e
,
,
, ,
-
o
o
-
o
o

o
o
-
o
o

o
- o
x x
V V
i
i
i
i
i k
i
k
i j
i
j
i
i
k k j j
x x x x x
b a b
a b a

)

(

Working directly with indicial notation we obtain:
b a b a
,
,
,
,
, , ,
- - - -
x x x
V V V
i i i i i i i
, , ), ( ) ( b a b a
Problem 1.44: Find the components of b a
,
,
,
) (
x
V .
Solution: Bearing in mind that
j j
e a

a
,
,
k k
e b

b
,
,
i
i
x o
o
e

x
,
V ( 3 , 2 , 1 i ), the following is
true:
j
k
j
k j
i
j
ik k k k i j
i
j
k k i
i
j j
x x x x
e e e e e e e
e
b a

)

(

)

(
) (
o
o

o
o

|
|
.
|

\
|

o
o

|
|
.
|

\
|

o
o

a
b
a
b b
a
b
a
c
,
,
,
x
V
Expanding the dummy index k , we obtain:
3
3
2
2
1
1
x x x x
j j j
k
j
k
o
o
-
o
o
-
o
o

o
o a
b
a
b
a
b
a
b
Thus,
3
3
3
2
3
2
1
3
1
3
2
3
2
2
2
1
2
1
3
1
3
2
1
2
1
1
1
3
2
1
x x x
j
x x x
j
x x x
j
o
o
-
o
o
-
o
o

o
o
-
o
o
-
o
o

o
o
-
o
o
-
o
o

a
b
a
b
a
b
a
b
a
b
a
b
a
b
a
b
a
b

Problem 1.45: Prove that the following relationship is valid:
1 TENSORS

113
T
T T T
x x x
, , ,
, ,
,
V V V -
|
|
.
|

\
|
q q
q
2
1 1

where ) , ( t x
,
,
q is an arbitrary vector field, and ) , ( t T x
,
is a scalar field.
Solution:
(scalar)
1 1
1 1
2
,
2
,
,
T
T T
T
T T T T x T
i i i i
i
i i
i
x x
x
, ,
,
, ,
,
V V
V

-
- |
.
|

\
|
= |
.
|

\
|
o
o

|
|
.
|

\
|
q q
q
q q
q q

1.8.4 The Curl
The curl of a vector field
The curl (or rotor) of a vector field, ) ( x
, ,
v is denoted by v v v
,
,
, ,
,
r = =
x
V ) ( ) ( rot curl , and is
defined in the Cartesian basis as:
) (

) ( - r
o
o
- r
j
j
x
e
x
,
,
V

The curl (rotor) of a tensor field in
the Cartesian basis
(1.482)
Note that the curl is already a tensor operator between two vectors. Using the definition of
the vector product we obtain the curl of a vector field as:
i j k ijk i ijk
j
k
k j
j
k
k k j
j
x x x
e e e e e e v v

)

) (
,
v
v v
v rot
o
o
r
o
o
r
o
o
r
,
,
,
,
x
V
(1.483)
where
ijk
is the permutation symbol defined in (1.55). Moreover, we have applied the
definition
i ijk k j
e e e

r and we can also note that:
3
2
1
1
2
2
1
3
3
1
1
3
2
2
3
,
3 2 1
3 2 1
3 2 1


) (
e e e
e
e e e
v v
|
|
.
|

\
|
o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o

o
o
o
o
o
o
r
x x x x x x
x x x
i j k ijk
v v v v v v
v
v v v
rot
,
,
,
,
x
V
(1.484)
We can verify that the antisymmetric part of a vector field gradient, which is illustrated by
W v =
skew
) (
,
,
x
V , has as components:
[ ]
(
(
(

-
-
-

(
(
(

- -
-
(
(
(

(
(
(
(
(
(
(
(

|
|
.
|

\
|
o
o
-
o
o
|
|
.
|

\
|
o
o
-
o
o
|
|
.
|

\
|
o
o
-
o
o
|
|
.
|

\
|
o
o
-
o
o
|
|
.
|

\
|
o
o
-
o
o
|
|
.
|

\
|
o
o
-
o
o
=
0
0
0
0
0
0
0
0
0
0
2
1
2
1
2
1
0
2
1
2
1
2
1
0
) (
1 2
1 3
2 3
23 13
23 12
13 12
32 31
23 21
13 12
3
2
2
3
3
1
1
3
2
3
3
2
2
1
1
2
1
3
3
1
1
2
2
1
,
w w
w w
w w
x x x x
x x x x
x x x x
skew
j i ij
skew
W W
W W
W W
W W
W W
W W
v v v v
v v v v
v v v v
v v
,
,
x
V
(1.485)
NOTES ON CONTINUUM MECHANICS

114
where
1
w ,
2
w ,
3
w are the components of the axial vector w
,
associated with W, (see
subsection: 1.5.2.2.2. Antisymmetric Tensor).
With reference to the definition of the curl in (1.484) and the relationship in (1.485), we can
conclude that:
( )
w
x
,
,
,
,
,
2

2

2

2

) (
3 3 2 2 1 1
3 21 2 13 1 32
3
2
1
1
2
2
1
3
3
1
1
3
2
2
3

- -
- -
|
|
.
|

\
|
o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o
r =
e e e
e e e
e e e v v
w w w
x x x x x x
W W W
v v v v v v
rot V
(1.486)
And, if we use the identity in (1.141), we obtain:
( ) v v v v W
, ,
,
, , ,
,
r r r
x
w V
2
1
(1.487)
It could be interesting to note that the equation in (1.486) can be obtained by means of
Problem 1.18, in which we showed that ) (
2
1
x a
, ,
r is the axial vector associated with the
antisymmetric tensor
skew
) ( a x
, ,
. Therefore, the axial vector associated with the
antisymmetric tensor [ ]
skew
skew
) ( ) ( ) ( V V
,
, ,
,
v v W
x
is the vector ( ) v
,
,
,
r
x
V
2
1
.
As we can see, the curl describes the rotational tendency of the vector field.
Summary
Divergence
- = -
x
,
V ) ( div
Gradient
- = -
x
,
V ) ( grad
Curl
- r = -
x
,
,
V ) ( rot
Scalar vector
Vector Scalar Second-order tensor Vector
Second-order tensor Vector Third-order tensor Second-order tensor

We can now present some equations:
) ( ) ( ) ( ) ( a a a a
, ,
,
,
,
,
, , ,
r / - r / / r /
x x x
V V V rot
The result of the algebraic operation ) ( a
,
,
,
/ r
x
V is a vector, whose components are
given by:
[ ]
i i
k j ijk i
k j ijk j k ijk
j k k j ijk
j k ijk i
) ( ) (
) ( ) (
) (
) ( ) (
, ,
, ,
,
a a
a
a
, ,
,
,
,
, ,
, ,
,
r / r /
/ r /
/ /
/ - /
/ / r
x x
x x
x
V V
V V
V

a
a a
a a
a

(1.488)
We can use the above equation to check that the relationship
) ( ) ( ) ( ) ( a a a a
, ,
,
,
,
,
, , ,
r / - r / / r /
x x x
V V V rot holds.
-
1 TENSORS

115
a b b a b a a b b a
,
, ,
,
,
, ,
, ,
,
,
, , , , ,
- - - r r ) ( ) ( ) ( ) ( ) (
x x x x x
V V V V V
(1.489)
The components of the vector product ) ( b a
,
,
r are given by
j i kij k
b a r ) ( b a
,
,
, thus:
[ ] ) ( ) ) (
, , ,
(
p j i j p i lpk kij p j i kij lpk l
b a b a b a - r r b a
,
,
,
,
x
V
(1.490)
Regarding that
ijk kij
and
jl ip jp il lpk ijk
c c c c - , the above equation becomes:
[ ]
p l p p p l l p p p p l
p j i jl ip p j i jp il j p i jl ip j p i jp il
p j i j p i jl ip jp il p j i j p i lpk kij l
, , , ,
, , , ,
, , , ,
) )( ( ) ( ) (
b a b a b a b a
b a b a b a b a
b a b a b a b a
- - -
- - -
- - - r r
c c c c c c c c
c c c c b a
,
,
,
,
x
V

(1.491)
We can also verify that [ ]
p p l l
b a
,
) ( b a
,
,
,
x
V , [ ]
l p p l
b a
,
) ( b a
,
,
,
x
V , [ ]
p p l l ,
) ( b a a b
,
,
,
x
V ,
[ ]
p l p l ,
) ( b a a b
,
,
,
x
V .
a a a
, , ,
, ,
, , , , ,
2
) ( ) (
x x x x x
V V V V V - r r (1.492)
The components of ) ( a
,
,
,
r
x
V are given by
_
,
,
,
i
j k ijk i
c
a
,
) ( r a
x
V , thus:
[ ]
jl k ijk qli l j k ijk qli l i qli q , , , ,
) ) ( ( a a c r r a
,
, ,
, ,
x x
V V
(1.493)
Once again considering that
lj qk lk qj jki qli ijk qli
c c c c - , the above equation becomes:
[ ]
ll q kq k
jl k lj qk jl k lk qj jl k lj qk lk qj jl k ijk qli q
, ,
, , , ,
) ( ) (
a a
a a a a
-
- - r r c c c c c c c c a
,
, ,
, ,
x x
V V

(1.494)
where [ ]
kq k q ,
) ( a a
,
, ,
x x
V V and [ ]
ll q q ,
2
a a
,
,
x
V .
) ( ) ( ) (
2
c c c
x x x x x
, , , , ,
V V V V V - (1.495)
) ( ) ( ) ( ) (
2
, , , , ,
c c c c c c
x x x x x
, , , , ,
V V V V V - -
i i ii i i

(1.496)
where c and are scalar fields. Other interesting equations derived from the above are:
) ( ) ( ) (
) ( ) ( ) (
2
2
c c c
c c c
x x x x x
x x x x x
, , , , ,
, , , , ,
V V V V V
V V V V V


-
-
(1.497)
After subtracting the above two identities we obtain:
c c c c
c c c c
2 2
2 2
) (
) ( ) (
x x x x x
x x x x x x
, , , , ,
, , , , , ,
V V V V V
V V V V V V
- -
- -


(1.498)
1.8.5 The Conservative Field
A vector field, ) , ( t x
,
,
b , is said to be conservative if there exists a differentiable scalar field,
) , ( t x
,
c , so that:
c
x
,
,
V b (1.499)
NOTES ON CONTINUUM MECHANICS

116
If the function c satisfies the relation (1.499), then c is a potential function of ) , ( t x
,
,
b .
A necessary but insufficient condition for ) , ( t x
,
,
b to be conservative is that 0 b
, ,
,
,
r
x
V . In
other words, given a conservative field, the curl b
,
,
,
r
x
V equals zero. However, if the curl
of a vector field equals zero, this does not necessarily mean that the field is conservative.
Problem 1.46: Let c be a scalar field, and u
,
be a vector field. a) Show that
( ) 0 r v
x x
,
,
, ,
V V and ( ) 0
, ,
, ,
r c
x x
V V .
b) Show that ( ) [ ] [ ] ) ( ) ( ) ( ) )( ( v v v v v v v v
x x x x x x x x
,
,
, , ,
,
,
,
, , ,
, ,
, , , , , , , ,
r - r - r r r r V V V V V V V V ;
c) Referring v
x
, ,
,
r V , show that
, ,
,
,
,
, , , , ,
2 2 2
) ( ) (
x x x x x
v v V V V V V r r .

Solution:
Regarding that:
i j k ijk
v e

,
r v
x
,
,
,
V
( ) ( ) ( ) ( )
ji k ijk j k
i
ijk il j k
l
ijk l i j k ijk
l
v v
x
v
x
v
x
, , , ,



o
o

o
o

o
o
r c e e v
x x
,
,
, ,
V V
The second derivative of v
,
is symmetrical with ij , i.e.
ij k ji k
v v
, ,
, while
ijk
is
antisymmetric with ij , i.e.,
jik ijk
- , thus:
0
, 3 3 , 2 2 , 1 1 ,
- -
ji ij ji ij ji ij ji k ijk
v v v v
We can observe that
ji ij
v
, 1 1
equals the double scalar product by using a symmetric and an
antisymmetric tensor, so 0
, 1 1

ji ij
v .
Likewise, we can show that:
( ) 0 e e
,
,
, ,
r
i i i kj ijk

, 0 c c
x x
V V
b) Denoting by v
x
,
,
,
,
r V we obtain:
( ) [ ] ) ( v v v
x x x
, ,
,
, ,
, ,
, , ,
r r r r r V V V
Observing the equation in (1.489), it holds that:

, , , , , , , , , ,
,
, , , , ,
- - - r r ) ( ) ( ) ( ) ( ) ( v v v v v
x x x x x
V V V V V
Note that 0 ) ( r v
x x x
,
,
,
, , ,
V V V . Then, we can draw the conclusion that:
[ ] ) ( ) ( ) ( ) )( (
) ( ) ( ) ( ) (
v v v v v v
v v v v
x x x x x x
x x x x
,
,
, , ,
,
,
,
,
, , , , , , , ,
,
, , , , , ,
, , , ,
r - r - r
- - r r


V V V V V V
V V V V

c) Observing the equation in (1.492) we obtain:

,
,
,
,
, ,
, ,
, , ,
, , , , ,
r -
r r -

x x x
x x x x x
v
v v v
V V V
V V V V V
) (
) ( ) (
2

Applying the curl to the above equation we obtain:
[ ] ) ( ) ( ) (
2

0
,
, ,
_
,
,
,
,
, , , , , , ,
,
r r - r r

x x x x x x x
v v V V V V V V V
Referring once again to the equation in (1.492) to express the term ) (
,
, ,
, ,
r r
x x
V V :
[ ]
) (
) ( ) ( ) ( ) (
2
2 2 2
0
v
v v
x x
x x x x x x x x x x x
,
,
,
_
,
,
, , ,
, ,
,
,
, ,
, , , , , , , , , , ,
r
- r - - - r r - r


V V
V V V V V V V V V V V

1 TENSORS

117
1.9 Theorems Involving Integrals
1.9.1 Integration by Parts
Integration by parts states that:
} }
-
b
) ( ) ( ) ( ) ( ) ( ) (
a
b
a
b
a
dx x u x v x v x u dx x v x u
(1.500)
where
dx
dv
x v ) ( , and the functions ) (x u , ) (x v are differentiable in b x a s s .
1.9.2 The Divergence Theorem
Given a domain B with a volume V , and bounded by the surface S , (see Figure 1.37), the
divergence theorem, also called the Gauss theorem, applied to the vector field states that:
} } }
} } }


S
i i
S
i i
V
i i
S S V
dS dS dV
d dS dV


,
v n v v
S
x
,
, , ,
,
v n v v V
(1.501)
where n

is the outward unit normal to surface S .













Figure 1.37.
Let T be a second-order tensor field defined in the domain B. The divergence theorem
applied to this field is defined as:
} } }

S S V
d dS dV S
x
,
,

T n T T V
} } }

S
j ij
S
j ij
V
j ij
dS dS dV

,
T n T T (1.502)
By using the divergence theorem we can also demonstrate that:
n


B
S
dS
x
,


1
x

2
x

3
x
S
,
d
NOTES ON CONTINUUM MECHANICS

118
[ ]
} }
} }
} } }

-

S
j k
V
j k
S
j k
V
j i ik i j ik
S
j i ik
V
j i ik
V
j k
dS x dV x
dS x dV x x
dS x dV x dV x

), ( ), (
,
, ,
n
n
n
c c
c c
(1.503)
in which we have assumed that
ikj j ik
0
,
c . Additionally, by observing that
kj j k
x c
,
, we
can obtain:
}
} } }


S
S
j k kj
S
j k
V
kj
dS V
dS x V dS x dV

n 1 x
,
n n c c
(1.504)
Given a second-order tensor defined in the domain B, the following is valid:
[ ]
[ ]
} }
} }
} } }
o o - o
o o - o
o o o
S
k jk i
V
k jk i jk ik
S
k jk i
V
k jk i jk k i
S
k jk i
V
k jk i
V
k jk i
dS x dV x
dS x dV x x
dS x dV x dV x

), ( ), (
,
, ,
n
n
n
c

(1.505)
Hence proving that:
} } }
} } }
-
o - o o

V
T
S V
V
ji
S
k jk i
V
k jk i
dV dS dV
dV dS x dV x
)

,
n x x
x
, ,
,
V
n
(1.506)
or
( )
} } }
-
V
T
S V
dV d dV S x x
x
,
, ,
,
V (1.507)
Likewise, one can prove that:
} } }
-
V S V
dV dS dV )

( ) ( n x x
x
, ,
,
V (1.508)
Problem 1.47: Let ! be a domain bounded by 1 as shown in Figure 1.38. Further
consider that m is a second-order tensor field and . is a scalar field. Show that the
following relationship holds:
[ ] [ ] [ ]
} } }
-
! 1 !
! . 1 . ! . d d d
x x x x x
, , , , ,
V V V V V ) (

) ( ) ( m n m m:
1 TENSORS

119
[ ] [ ]
} } }
-
! 1 !
! . 1 . ! . d d d
i j ij j ij i ij ij
,

) , ( ,
,
m n m m











Figure 1.38

Solution: We could directly apply the definition of integration by parts to demonstrate the
above relationship. But, here we will start with the definition of the divergence theorem.
That is, given a tensor field v
,
, it is true that:
} } } }

1 ! 1 !
1 ! 1 ! d d d d
j j j j
indicial



,
n v v
x
n v v
, ,
,
V
Observing that the tensor v
,
is the result of the algebraic operation m v .
x
,
,
V and the
equivalent in indicial notation to
ij i j
m v , . , and by substituting it in the above equation
we obtain:
[ ]
[ ]
[ ] [ ]
} } }
} }
} } } }
-
-

! 1 !
1 !
1 ! 1 !
! . 1 . ! .
1 . ! . .
1 . ! . 1 !
d d d
d d
d d d d
j ij i j ij i ij ij
j ij i j ij i ij ij
j ij i
j
ij i j j j j
,

, ,

, , ,

, ,

,
,
,
,
m n m m
n m m m
n m m n v v

The above equation in tensorial notation becomes:
[ ] [ ] [ ]
} } }
-
! 1 !
! . 1 . ! . d d d ) (

) ( ) ( m n m m
x x x x x
, , , , ,
V V V V V :
NOTE: Consider now the domain defined by the volume V , which is bounded by the
surface S with the outward unit normal to the surface n

. If N
,
is a vector field and T is a
scalar field, it is also true that:
} } }
} } }
-
-
V S V
V
i j i
S
j i i
V
ij i
dV T dS T dV T
dV T N dS T N dV T N

) ( ) (
,

, ,
,
N N N
x x x x x
, , ,
, , , , ,
V V V V V n
n

where we have directly applied the definition of integration by parts.
!
n


1 1
x

2
x
NOTES ON CONTINUUM MECHANICS

120
1.9.3 Independence of Path
A curve which connects two points A and B is denoted by the path from A to B , (see
Figure 1.39). We can then establish the condition by which a line integral is independent of
path, (see Figure 1.39).








Figure 1.39: Path independence.
Let ) ( x
,
,
b be a continuous vector fields, then the integral
}

C
r
,
,
d b is independent of the path if
and only if b
,
is a conservative field. This means that there is a scalar field c so that c
x
,
,
V b .
Regarding the above, we can draw the conclusion that:
} }
} }


|
|
.
|

\
|
o
o
-
o
o
-
o
o
- -

B
A
B
A
B
A
x
B
A
r r
r r
, ,
, ,
,
,
d
x x x
d
d d
3
3
2
2
1
1
3 3 2 2 1 1

)

( e e e e e e
b
c c c
c
b b b
V

(1.509)
Thus
3
3
2
2
1
1
; ;
x x x o
o

o
o

o
o

c c c
b b b
(1.510)

As the field is conservative, the curl of b
,
is the zero vector:
i
x x x
0

3 2 1
3 2 1
3 2 1

o
o
o
o
o
o
r
b b b
e e e
0 b
, ,
,
,
x
V
(1.511)
We can therefore conclude that:

o
o

o
o
o
o

o
o
o
o

o
o

o
o
-
o
o

o
o
-
o
o

o
o
-
o
o
2
1
1
2
1
3
3
1
3
2
2
3
2
1
1
2
1
3
3
1
3
2
2
3
0
0
0
x x
x x
x x
x x
x x
x x
b b
b b
b b
b b
b b
b b

(1.512)
Therefore, if the above condition is not satisfied, the field is not conservative.

1
x

2
x
3
x
A
B
1
C
r
,
d
b
,


2
C
If
} }

2 1
C C
r r
,
,
,
,
d d b b
b
,
- Conservative field
1 TENSORS

121
1.9.4 The Kelvin-Stokes Theorem
Let S be a regular surface, (see Figure 1.40), and ) , ( t x
,
,
F be a vector field. According to the
Kelvin-Stokes Theorem:
} } }
r r
! ! 1
dS d d n F F F

) ( ) (
, , , , , ,
, ,
,
x x
S V V

(1.513)
If p

denotes the unit vector tangent to the boundary 1 , the Stokes theorem becomes:
} } }
r r
! ! 1
1 dS d d n F F p F

) ( ) (

, , , , , ,
, ,
x x
S V V
(1.514)
With reference to the vector representation in the Cartesian basis:
3 3 2 2 1 1

e e e F F F F - -
,
,
3 3 2 2 1 1

e e e dS dS dS d - - S
,
,
3 3 2 2 1 1

e e e dx dx dx d - -
,
, the components of the curl of F
,

are given by:
( )
3
2
1
1
2
2
1
3
3
1
1
3
2
2
3
3 2 1
3 2 1
3 2 1


e e e
e e e
F
|
|
.
|

\
|
o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o

o
o
o
o
o
o
r
x x x x x x x x x
i
F F F F F F
F F F
, ,
,
x
V
(1.515)
Next, the Stokes theorem expressed in terms of components becomes:
3
2
1
1
2
2
1
3
3
1
1
3
2
2
3
3 3 2 2 1 1
dS
x x
dS
x x
dS
x x
dx dx dx
} } |
|
.
|

\
|
o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o
- -
! 1
F F F F F F
F F F
(1.516)









Figure 1.40: Stokes theorem.
In the special case when the surface S coincides with the plane !, (see Figure 1.41), the
equation (1.516) remains valid. Then, if the domain ! coincides with the plane
2 1
x x - ,
the equation (1.513) becomes:
} }
r
! 1
dS d
3

) ( e F F
, , ,
,
,
x
V
(1.517)
which is known as the Stokes theorem in the plane or Greens theorem, which is expressed in
terms of components as:
3
2
1
1
2
2 2 1 1
dS
x x
dx dx
} } |
|
.
|

\
|
o
o
-
o
o
-
! 1
F F
F F
(1.518)

2
x

3
x

1
x
n


S
!
1
p


NOTES ON CONTINUUM MECHANICS

122












Figure 1.41.











Figure 1.42: Greens theorem.
1.9.5 Greens Identities
Let F
,
be a vector field, and by applying the divergence theorem we obtain:
} }

S V
dS dV

n F F
, ,
,
x
V (1.519)
With reference to the equations (1.496) and (1.498), i.e.:
) ( ) ( ) (
2
c c c
x x x x x
, , , , ,
V V V V V - (1.520)
c c c c
2 2
) (
x x x x x
, , , , ,
V V V V V - - (1.521)
and also regarding that c
x
,
,
V F , and by substituting (1.520) into (1.519) we obtain:
!
1

2
x

1
x

3
x
n



2
x

1
x

3
x
!

3

e dS d S
,

1

3

e
1 TENSORS

123
} } }
} }
-
-


V S V
S V
dV dS dV
dS dV

) ( ) (

) ( ) (
2
2
c c c
c c c
x x x x
x x x x
, , , ,
, , , ,
V V V V
V V V V
n
n
(1.522)
which is known as Greens first identity.
Now, if we substituting (1.521) into (1.519) we obtain:
} }
- -
S V
dS dV

) (
2 2
n c c c c
x x x x
, , , ,
V V V V (1.523)
which is known as Greens second identity.
Problem 1.48: Let b
,
be a vector field, which is defined as v b
,
,
,
,
r
x
V . Show that:
} }
/ /
V
i i
S
i i
dV S d ,

b n b
where ) ( x
,
/ / .

Solution: The Cartesian components of v b
,
,
,
,
r
x
V are
j k ijk i ,
v b and by substituting
them in the above surface integral we obtain:
} }
/ /
S
i j k ijk
S
i i
dS dS

,
n v n b
Applying the divergence theorem we obtain:
} }
}
} } }
/ / - /
/ - /
/ / /
V V
i i ji k ijk j k ijk i
V
ji k ijk j k i ijk
V
i j k ijk
S
i j k ijk
S
i i
dV dV
dV
dV dS dS
i
, ) , (
) , (
), (

0
, ,
, ,
, ,
b v v
v v
v n v n b
b
_ _















A. Graphical Representation of a Second-Order Tensor











A.1 Projecting a Second-Order Tensor onto a
Particular Direction
A.1.1 Normal and Tangential Components
Projecting a second-order tensor ( T ) onto the n

-direction results in n T t
n

(

,
whose
vector
)

(n
t
,
can be split into:
S N
T T t
n
, , ,
-
)

(

(A.1)
where
N
T
,
is the normal vector, and
S
T
,
is the tangential vector, (see Figure A.1).
If we then bear in mind that n

and s

are unit vectors according to the directions


N
T
,
and
S
T
,
, respectively, the vector
)

(n
t
,
can also be expressed as:
s n t
n

)

(
S N
T T -
,

(A.2)
where
N
T and
S
T are the magnitudes of
N
T
,
and
S
T
,
, respectively.
The vector
N
T
,
can also be evaluated as follows:
[ ]n n T n
n t n n n t
n n T
n n

(

)

( )

(
_
, ,
,
N
N N N
T
T T


=

[ ]
i
N
j kj k
i k k i k k
i N i N
n n T n
n t n n n t
n T T
T

( )

(
_

n n

(A.3)
A
Graphical Representation
of a Second-Order Tensor

Appendix
, Notes on Continuum Mechanics, Lecture Notes on Numerical
n Engineering and Sciences 4, DOI 10.1007/978-94-007-5986-2_3, Methods i
International Center for Numerical Methods in Engineering (CIMNE), 2013
E.W.V. Chaves 125
NOTES ON CONTINUUM MECHANICS

126
Thus:
j kj k N
n T n T

)

(
n T n n t
n
,
(A.4)











Figure A.1: Normal and tangential vectors.
As we saw in Chapter 1, T is a positive definite tensor if 0

> n T n
N
T for all 0 n
,

. It
is also true that n T n n T n


sym
N
T , thus, if the symmetric part of a tensor is a
positive definite tensor then the tensor will be also.
The vector
S
T
,
can also be expressed in terms of
S
T and s

as:
[ ]s n T s
s t s s s t
s T
n n

( )

(
_
, ,
,
S
S S
T
T


[ ]

(
i j k jk
i j j
i S i S
S
s s n T
s s t
s T T
T
_
,

n

(A.5)
Then, another way to evaluate the tangential vector can be by means of the equation in
(A.1):
[ ]n n n T n T T t T
n

)

(

(
- - :
N S
, , ,

(A.6)
Note, the magnitude of
S
T
,
can also be obtained by means of the Pythagorean Theorem,
i.e.:
2 )

( )

( 2
2 2 2
)

(
N i i S S N
T t t T - -
n n n
T T t
, , ,
(A.7)
where
k j ik ij i i
n n T T t t

)

( )

n n
.
The question now is: on what plane is the maximum normal and tangential component?
The answer to this problem is related to the maximum and minimum values of a function,
which will be discussed in the next subsection.

1

e

2

e

)

(n
t
,

n



N
T
,


S
T
,


1
x

2
x

3
x
s


P

3

e
APPENDIX A: GRAPHICAL REPRESENTATION OF A SECOND-ORDER TENSOR

127
Problem A.1: Show that )

(
)

(
n n 1 t T
n
-
, ,
S
, where
)

(n
t
,
is vector resulting from
projecting the second-order tensor T onto the n

-direction, and
S
T
,
is the tangential
vector.
Solution 1: If we consider that
S N
T T t
n
, , ,
-
)

(
and (A.3) we can state that:
[ ] )

(

)

( )

( )

( )

( )

(
n n 1 t n n t t n n t t T
n n n n n
- - -
, , , , , ,
S

Solution 2: We can also solve the problem just using the components of the equation (A.6),
[ ]n n n t T
n

)

(
)

(
- :
, ,
S
, i.e.:
[ ] ( )
k i ik k k k i ik k k k i i i kl l k i i S
n n t t n n t t n n t n T n n t T

)

(
)

( ) ( )

( )

( )

( )

(
- - - - c c
n n n n n n

A.1.2 The Maximum and Minimum Normal Components
As we have seen previously the normal component is given by n T n


N
T with the
constraint 1

n n (the unit vector). The maximum and minimum values of
N
T with this
constraint can be evaluated by means of the Lagrange multiplier method which consists in
constructing a function such as:
) 1

(

) 1

( ) ,

( - - - - n n n T n n n n
N
T L (A.8)
where is known as the Lagrange multiplier. Then, the derivative of the function )

(n L
with respect to n

and , yields the following set of equations:


1

0 1

) ,

) (

) ,

(
-
o
o
- -
o
o


n n n n
n
0 n 1 T 0 n n T
n
n
L
L
, ,
sym sym

(A.9)
The first set can only be solved if and only if 0 ) ( - 1 T
sym
det , which is the eigenvalue
problem of the symmetric part of T . That is, the maximum and minimum of
N
T
correspond to the eigenvalues of
sym
T . Now, if we consider that
sym
1
T ,
sym
2
T ,
sym
3
T , are the
eigenvalues of
sym
T , we can then restructure these such that:
sym sym sym
III II I
T T T > > (A.10)
Then, the maximum value of
N
T is defined by
sym
I
T , and the minimum value by
sym
III
T .

NOTE: As expected, the extreme values of
N
T relate to the symmetric part
sym
T because
the antisymmetric part plays no role in the normal component, i.e.
n T n n T n


sym
N
T .
OBS.: When the nomenclature
III II I
, , T T T are used to represent the eigenvalues,
III II I
T T T > > is implicit.
NOTES ON CONTINUUM MECHANICS

128
A.1.3 The Maximum and Minimum Tangential Component
In this section we will consider a symmetric second-order tensor, i.e.
T
T T . For the sake
of convenience, we will work here in the principal space of T where the tensor
components are just represented by the normal components, (see Figure A.2).










Figure A.2: Principal space.
Then, the normal component
N
T for the n

-direction can be obtained by means of the


equation in (A.4) as follows:
2
3 3
2
2 2
2
1 1
)

(

n T n T n T n n T n t T - -
i j ij i i N
n
(A.11)
Note that, on the particular plane [ ]
1
0 , 0 , 1

T T n
N i
.
The tangential component can then be evaluated as follows:
2 2 )

( )

( 2
2
)

( 2

N k j ik ij N i i N S
T n n T T T t t T T - - -
n n n
t
,
(A.12)
Next, by combining the equations (A.11) and (A.12) we find:
( )
2
2
3 3
2
2 2
2
1 1
2
3
2
3
2
2
2
2
2
1
2
1
2

n T n T n T n T n T n T T - - - - -
S

(A.13)
If we now ask what values of
i
n

maximize the function


2
S
T , this problem is equivalent to
find extreme values of the function:
( ) 1

)

(
2
- -
i i S
F n n T n (A.14)
where is the Lagrange multiplier, with the constraint 1


i i
n n . Then:
0
)

(
; 0
)

o
o

o
o n n F F
j
j
n

(A.15)
and by solving the above set of equations we obtain:
( ) {
( ) {
( ) { 0

2

0

2

0

2

2
3 3
2
2 2
2
1 1 3
2
3 3
2
3 3
2
2 2
2
1 1 2
2
2 2
2
3 3
2
2 2
2
1 1 1
2
1 1
- - - -
- - - -
- - - -
n T n T n T T T n
n T n T n T T T n
n T n T n T T T n
(A.16)
with the constraint 1


i i
n n . Then, the analytical solution of the previous set of equations
results in the following possible solutions:

1
x

2
x

3
x

1
T

3
T

2
T

1
x

2
x

3
x
n

(n
t
,


S
T
,


N
T
,

arbitrary plane
APPENDIX A: GRAPHICAL REPRESENTATION OF A SECOND-ORDER TENSOR

129

solutions
1

n
2

n
3

n
S
T
(1) 1

) 1 (
1
n 0

) 1 (
2
n 0

) 1 (
3
n 0
S
T

(2) 0

) 2 (
1
n 1

) 2 (
2
n 0

) 2 (
3
n 0
S
T

(3) 0

) 3 (
1
n 0

) 3 (
1
n 1

) 3 (
1
n 0
S
T

(4) 0

2
1

2
1


2
3 2
T T
T
-

S

(5)
2
1


0

2
1


2
3 1
T T
T
-

S

(6)
2
1


2
1


0

2
2 1
T T
T
-

S

(A.17)
where the values of
S
T were obtained by substituting the values of
i
n

into (A.13).
The first three sets of solutions give us the minimum values of
S
T , which correspond
precisely to the principal directions.
For solutions (4), (5) and (6) the planes are outlined as shown in Figure A.3, Figure A.4 and
Figure A.5, respectively.
Then, by ordering the eigenvalues (the principal values
1
T ,
2
T ,
3
T ) such that
III II I
T T T > > we can find the absolute maximum tangential component:
2
III I
max
T T
T
-

S
(A.18)
which corresponds to solution (5) of (A.17).











Figure A.3: The maximum relative for
S
T , solution (4).




3
T

2
T

1
T

3
T

2
T

1
T
[ ]
2
1
2
1
0


i
n
[ ]
2
1
2
1
0

- -
i
n
[ ]
2
1
2
1
0

-
i
n
[ ]
2
1
2
1
0

-
i
n
NOTES ON CONTINUUM MECHANICS

130










Figure A.4: The maximum relative for
S
T , solution (5).










Figure A.5: The maximum relative for
S
T , solution (6).
A.2 Graphical Representation of an Arbitrary
Second-Order Tensor
If the second-order tensor Cartesian components are known, it is possible to evaluate the
normal and tangential components (
N
T ,
S
T ) on any plane defined by the normal n

, with
the constraint 1

2
3
2
2
2
1
- - n n n n n . Our goal in this section is to draw a graph in which
the abscissa is represented by the normal components (
N
T ) and the ordinate represents the
tangential component (
S
T ). This procedure can be carried out either numerically or
analytically.
Firstly, we will adopt a numerical procedure, i.e. we will randomly evaluate different values
for the normal n

, and then we will obtain the corresponding values of (


N
T ,
S
T ), whose
values will be plotted on the graph
S N
T T . In this way we will obtain all feasible values of
the vector resulting from projecting the tensor onto a direction. Similarly, we will also
construct a graph which corresponds to the symmetric part, i.e.
) ( ) (
) (
sym
S N
sym
N
T T T .
[ ]
2
1
2
1
0


i
n

3
T
[ ]
2
1
2
1
0

-
i
n
3
T

2
T

1
T

2
T

1
T
[ ]
2
1
2
1
0

- -
i
n
[ ]
2
1
2
1
0

-
i
n
[ ] 0

2
1
2
1
- -
i
n

3
T

2
T

1
T

3
T

2
T

1
T
[ ] 0

2
1
2
1

i
n
[ ] 0

2
1
2
1
-
i
n
[ ] 0

2
1
2
1
-
i
n
APPENDIX A: GRAPHICAL REPRESENTATION OF A SECOND-ORDER TENSOR

131
The first example, (see Figure A.6), is a non-symmetric tensor and it is noteworthy that it is
positive definite since 0

> n T n
N
T for all 0 n
,

. We will also verify that it has three


real eigenvalues, which correspond to 0
S
T . Note, the maximum and minimum values
for the
N
T coincide with the eigenvalues of the symmetric part of T .
Then, for the tangential vector, we can carry out the following decomposition:
[ ] [ ]s n T s n T s s n T s T

-
skew sym
S
S
_
,
T

(A.19)
When n

is one of the principal directions of the symmetric part, we have:


[ ] [ ] [ ]s n T s s n T s n s s n T s n T s T

- / -
skew skew skew sym
S
,

(A.20)
where we have considered that 0

n s , since the unit vectors s

and n

are orthogonal to
each other. So, the tangential component can be obtained as:
)

( )

( )

(

n n n
t s s t t s n T s
skew skew skew
skew
S
, , ,
T (A.21)
We must emphasize that this procedure is only valid when n

is a principal direction of
sym
T , but not on an arbitrary plane.
For example, for the eigenvalue 55 . 10
sym
I
T , which is associated with the eigenvector
[ ] 692913086 . 0 ; 561517458 . 0 ; 45229371 . 0

) 1 (
- - -
j
n , we have:
424378 . 2
5753286 . 1
8381199 . 1
1313956 . 0

692913086 . 0
561517458 . 0
45229371 . 0

0 2 1
2 0 1
1 1 0

) 1 (
)


(
(
(

-

(
(
(

-
-
-
(
(
(

- -
-

_ _
,
j
anti
ij
skew S
n
T
T
n
t
(A.22)
for the eigenvalue 61 . 3
sym
I I
T , we find:
55947 . 0
50644304 . 0
036633496 . 0
234905 . 0

42439659 . 0
18949182 . 0
88542667 . 0

0 2 1
2 0 1
1 1 0

)

(

(
(
(

-
-

(
(
(

(
(
(

- -
-

n
t
skew S
,
T
(A.23)
and, for the eigenvalue 84 . 0
sym
I I I
T , we obtain:
41 . 2
5039465 . 1
2727817 . 1
3883641 . 1

582888489 . 0
80547563 . 0
107004733 . 0

0 2 1
2 0 1
1 1 0

)

(

(
(
(

-
-
-

(
(
(

-
(
(
(

- -
-

n
t
skew S
,
T
(A.24)
Note that the maximum and minimum normal components of T corresponds to the
eigenvalues of
sym
T , i.e., 55 . 10
max

sym
I N
T T and 84 . 0
min

sym
I I I N
T T . With respect to the
symmetric part, the maximum tangential component is equal to the radius of the circle
formed by 55 . 10
sym
I
T and 84 . 0
sym
I I I
T , (see Figure A.6), which is
86 . 4
2
84 . 0 55 . 10
max

-

sym
S
T .

NOTES ON CONTINUUM MECHANICS

132
























Figure A.6: Graphical representation of a positive definite tensor.
The second example is a symmetric tensor, which has two equal eigenvalues. We can now
verify that the possible values for (
N
T ,
S
T ) are limited to the circumference radius
5 . 2
2
III I

T T
R and centered at the point ( 5 . 1
2
III I

T T
T
N
, 0
S
T ), (see Figure A.7).
Intuitively, this leads us to believe that the graphical representation of a spherical tensor
reduces to a single point.









N
T

S
T

(
(
(

6 6 3
2 4 1
1 3 5
ij
T
52 . 1
I I I
T
59 . 3
I I
T 89 . 9
I
T


(
(
(

6 4 2
4 4 2
2 2 5
sym
ij
T

N
T

sym
S
T
55 . 10
sym
I
T
61 . 3
sym
I I
T
84 . 0
sym
I I I
T
424 . 2 ; 55 . 10
max

S N
T T
41 . 2 ; 84 . 0
min

S N
T T
86 . 4
max

sym
S
T
424 . 2 ; 55 . 10
max
-
S N
T T
41 . 2 ; 84 . 0
min
-
S N
T T
APPENDIX A: GRAPHICAL REPRESENTATION OF A SECOND-ORDER TENSOR

133














Figure A.7: Graphical representation of a symmetric tensor with two equal eigenvalues.

The third example is a non-symmetric tensor, which has one real eigenvalue that is equal to
964 . 0
1
- T , (see Figure A.8).
In the principal direction of the symmetric part of T , we have the following values for the
tangential component:
For the eigenvalue 894 . 9
sym
I
T , we have:
0 . 6
482120 0
991187 . 4
299127 . 3

484682632 . 0
514420622 . 0
707427855 . 0

0 3 5 . 1
3 0 5
5 . 1 5 0

)

(

(
(
(

-
-

(
(
(

-
-
-
(
(
(

-
- -
.
skew S
n
t
,
T
(A.25)
and the eigenvalue 02 . 2 -
sym
I I I
T , we have:
1676 . 4
2624170 . 2
3453965 . 2
5979967 . 2

97 6859401168 . 0
5 7253813847 . 0
0575152387 . 0

0 3 5 . 1
3 0 5
5 . 1 5 0

)

(

(
(
(

-
-
-

(
(
(

-
(
(
(

-
- -
n
t
skew S
,
T
(A.26)










N
T

S
T

(
(
(

-
-
1 0 0
0 1 0
0 0 4
ij
T
1 -
I I I I I
T T
4
max

N I
T T
5 . 2
max

S
T
NOTES ON CONTINUUM MECHANICS

134





























Figure A.8: Graphical representation of a non-symmetric tensor with only one real
eigenvalue.
A.2.1 Graphical Representation of a Symmetric Second-Order
Tensor (Mohrs Circle)
In this subsection we will focus our attention on the graphical representation of a
symmetric second-order tensor, which is denoted by the Mohrs Circle.
Once again, we will work in the principal space, and we will assume that the eigenvalues of
T are ordered such that
III II I
T T T > > . We will start from the equation in (A.7), i.e.:



N
T

S
T

(
(
(

-
2 7 1
1 2 2
4 8 6
ij
T

N
T

sym
S
T

(
(
(

2 4 5 . 2
4 2 3
5 . 2 3 6
sym
ij
T

max S
T

max
sym
S
T
02 . 2 -
sym
I I I
T
126 . 2
sym
I I
T
894 . 9
sym
I
T
964 . 0 -
I
T
0 . 6 ; 894 . 9
max

S N
T T
17 . 4 ; 02 . 2
min
-
S N
T T
0 . 6 ; 894 . 9
max
-
S N
T T
17 . 4 ; 02 . 2
min
- -
S N
T T
APPENDIX A: GRAPHICAL REPRESENTATION OF A SECOND-ORDER TENSOR

135
2
)

( )

( )

( 2 2 n n n
t t t
, , ,
-
N S
T T (A.27)












Figure A.9: An arbitrary plane in the principal space.
The components of the vector
)

(n
t
,
, in an arbitrary plane, were obtained in the Eq. (A.1).
Then the components of n T t
n

(

,
in the principal space, (see Figure A.9(b)), are given
by:
3 III
)

(
3 2 II
)

(
2 1 I
)

(
1

n T t n T t n T t
n n n

(A.28)
Next, the dot product
)

( )

( n n
t t
, ,
is evaluated as follows:
2
3
2
III
2
2
2
II
2
1
2
I
)

(
3
)

(
3
)

(
2
)

(
2
)

(
1
)

(
1
)

( )

( )

( )

( )

( )

( )

( )

(


n T n T n T
t t t t t t
t t t t t t
- -
- -

n n n n n n
n n n n n n n n
e e t t
i i ij j i j j i i
c
, ,

(A.29)
and by combining the equations (A.27) and (A.29), we obtain:
2
3
2
III
2
2
2
II
2
1
2
I
2 2

n T n T n T T T - - -
N S
(A.30)
Then, the normal component
N
T , in the principal space, can be expressed as:
2
3 III
2
2 II
2
1 I
)

(

n T n T n T n n T T - -
i j ij N
n t
n
,
(A.31)
where we have used the equation in (A.4).
Let us consider the constraint 1

i i
n n
2
3
2
2
2
1

1

n n n - - which if we substitute into (A.31)


give us the value of
2
2

n :
) (


)

1 (
I II
I
2
3 I
2
3 III 2
2
2
3 III
2
2 II
2
3
2
2 I
T T
T n T n T T
n n T n T n n T T
-
- - -
- - - -
N
N
(A.32)
Then, substituting (
2
3
2
2
2
1

1

n n n - - ) into the equation (A.30) yields:


2
3
2
III
2
2
2
II
2
3
2
2
2
I
2
3
2
III
2
2
2
II
2
1
2
I
2 2

)

1 (

n T n T n n T
n T n T n T T T
- - - -
- - -
N S

(A.33)

I
T

III
T

II
T

1
x

2
x

3
x

)

(n
t
,


I
T

III
T

1
x

2
x

3
x
S
A
B
C
O
n


II
T

1

e

2

e

3

e
a) b)
NOTES ON CONTINUUM MECHANICS

136
and substituting
2
2

n , obtained in (A.32), into the above equation results in:


[ ]
III II I II
2
3 II III I III
2 2

) )( ( T T T T T T n T T T T T T - - - - - -
N N N S
(A.34)
Next, if we evaluate the term
2
3

n we obtain:
) )( (
) )( (

II III I III
2
II I 2
3
T T T T
T T T T T
n
- -
- - -

S N N
(A.35)
and we can find
2
1

n and
2
2

n in a similar fashion. Finally, we can summarize the results as


follows:
0
) )( (
) )( (

III I II I
2
III II 2
1

- -
- - -

T T T T
T T T T T
n
S N N
(a)
0
) )( (
) )( (

I II III II
2
I III 2
2

- -
- - -

T T T T
T T T T T
n
S N N
(b)
0
) )( (
) )( (

II III I III
2
II I 2
3

- -
- - -

T T T T
T T T T T
n
S N N
(c)
(A.36)
Then, if we consider that
III II I
T T T > > we can verify that the equations (A.36) (a) and (c)
have positive denominators; so, their nominators must also be positive. However, the
equation in (A.36) (b) has negative denominators, so its nominator must be negative too.
i.e.:
[ ]
[ ]
[ ]
[ ]
[ ]
[ ]

- - -
s - - -
- - -

> - -
- - -

< - -
s - - -

> - -
- - -

0 ) )( (
0 ) )( (
0 ) )( (
0
0 ) )( (
0 ) )( (

0
0 ) )( (
0 ) )( (

0
0 ) )( (
0 ) )( (

2
II I
2
I III
2
III II
II III I III
2
II I 2
3
I II III II
2
I III 2
2
III I II I
2
III II 2
1
S N N
S N N
S N N
S N N
S N N
S N N
T T T T T
T T T T T
T T T T T
T T T T
T T T T T
n
T T T T
T T T T T
n
T T T T
T T T T T
n

(A.37)
Then after some algebraic manipulations, the previous inequalities in (A.37) become:
[ ] [ ]
[ ] [ ]
[ ] [ ]
2
II I
2
1
2
II I
2
1
2
2
III I
2
1
2
III I
2
1
2
2
III II
2
1
2
III II
2
1
2
) ( ) (
) ( ) (
) ( ) (
T T T T T T
T T T T T T
T T T T T T
- - - -
- s - - -
- - - -
N S
N S
N S
(A.38)
The above shows equations of circles. The first circle with the center ) (
III II
2
1
T T - and
radius ) (
III II
2
1
T T - , indicates that the feasible points for the pair (
S N
T T ; ) are outside the
circle
1
C , including the circumference, (see Figure A.10). The second circle with the
center ) (
III I
2
1
T T - and radius ) (
III I
2
1
T T - , indicates that the feasible values for the pair
(
S N
T T ; ) are inside the circle
2
C , including the circumference. The third inequality
indicates that the feasible values for the pair (
S N
T T ; ) are outside the circle
3
C , which
is defined by, the radius ) (
II I
2
1
T T - and center ) (
II I
2
1
T T - . Then, taking into account the
three equations simultaneously, the feasible region is formed by the pair (
S N
T T ; ), defined
in the gray area which includes the circumferences
1
C ,
2
C and
3
C , (see Figure A.10).

APPENDIX A: GRAPHICAL REPRESENTATION OF A SECOND-ORDER TENSOR

137












Figure A.10: Mohrs circle feasible region.
In the Mohrs circle, (see Figure A.11), we can identify the maximum values of
max S
T ,
which were also obtained in the set of solutions (A.17).





















Figure A.11: Mohrs circle.

S
T
feasible region

N
T

2
C

1
C

I
T
II
T
III
T

3
C

III
T

II
T

I
T

III
T

II
T

I
T

III
T

II
T

I
T
Point N
Point N
Point M
Point M
Point Q
Point Q
Q

N
T
max I N
T T
III
T
II
T

S
T

max S
T
M
N
Q
N
M

2
III I
max
T T
T
-

S

NOTES ON CONTINUUM MECHANICS

138
A.3 The Tensor Ellipsoid
Let us consider a symmetric second-order tensor T which is represented in the principal
space by its eigenvalues (
1
T ,
2
T ,
3
T ) and where the following is fulfilled:


3
)

(
3
3
2
)

(
2
2
1
)

(
1
1
3 3
)

(
3
2 2
)

(
2
1 1
)

(
1
)

T
t
n
T
t
n
T
t
n
n T t
n T t
n T t
n
n
n
n
n
n
n
t n T
components
,

(A.39)
Our goal now is to define the surface, in the principal space, that describes the vector
)

(n
t
,

for all possible values of n

. Then, if we consider the constraint of n

, i.e. 1

2
3
2
2
2
1
- - n n n ,
and substitute the
i
n equation given by (A.39) we obtain:
1
2
3
2
)

(
3
2
2
2
)

(
2
2
1
2
)

(
1
- -
T
t
T
t
T
t
n n n

(A.40)
which represents an ellipsoid in the principal space of T , (see Figure A.12). This surface
describes the feasible values for
1
t ,
2
t and
3
t . When two eigenvalues are equal, the surface
becomes an ellipsoid of revolution, whereas when the three eigenvalues are equal, the
surface is a sphere, so, tensors that exhibit this property are called Spherical Tensors, and any
direction for these is a principal direction.









Figure A.12: The tensor ellipsoid.

)

(
2 2
,
n
t x

)

(
1 1
,
n
t x

1
T
2
T

3
T

)

(
3 3
,
n
t x
n

(n
t
,

APPENDIX A: GRAPHICAL REPRESENTATION OF A SECOND-ORDER TENSOR

139
A.4 Graphical Representation of the Spherical
and Deviatoric Parts
A.4.1 The Octahedral Vector
Firstly, we define an octahedral plane, also called the deviatoric plane, which is that in which the
normal is at an equal angle with each principal axis. In Figure A.13, the plane formed by
the points ABC is an octahedral plane whose normal is defined by n

and can easily be


evaluated by using the definition of the unit vector 1

2
3
2
2
2
1
- - n n n n n
i i
. Then, by
considering that
3 2 1

n n n we obtain 1

3
2
1
n , after which the plane with the normal
] [

3
1
3
1
3
1

i
n , in the principal space, is said to be an octahedral plane. So, the vector
that comes out of projecting a second-order tensor onto the octahedral plane is denoted by
the octahedral vector
)

(n
t
,
which can be decomposed into normal and tangential components,
(see Figure A.13), so defining the normal octahedral vector,
oct
N
T
,
, and the tangential octahedral
vector,
oct
S
T
,
.












Figure A.13: The octahedral plane and octahedral vector (principal space).
The octahedral vector can be expressed in terms of components in the form:
3
3
2
2
1
1 )

e e e n t
n
- -
T T T
,

(A.41)
Then, the magnitude of
oct
N
T
,
can be evaluated by projecting
oct
N
T
,
onto n

:
( )
m ii
oct
N
I
T T T T T
T T T
T
- -
|
|
.
|

\
|
-

|
|
.
|

\
|
- -
3 3
1
3
1
3

3 2 1
3 2 1
3
3
2
2
1
1 )

(
T
n
e e e
e e e n t
,
(A.42)

(
(
(

1
1
1
3
1

i
n

1
T

2
T

3
T
A
B
C
c
c
c
O
OC OB OA

)

(n
t
,

n


oct
S
T
,


oct
N
T
,


3
T

2
T

1
T

oct
N
T
,
- normal octahedral vector

oct
S
T
,
- tangential octahedral vector
NOTES ON CONTINUUM MECHANICS

140
where
oct
N
T is called the octahedral normal component. Then, the magnitude of
oct
S
T
,
, called the
octahedral tangential component, is obtained as follows:
[ ] [ ] ( )
T T
n n
t t I I I
oct
N
oct
S
6 2
9
1
9
1
3
1
2 2
3 2 1
2
3
2
2
2
1
2
)

( )

(
2
- - - - - - - T T T T T T T T
, ,
(A.43)
Note, the above equation can also be expressed as:
( ) ( ) ( )
( ) ( ) ( ) ( )
dev
I I
oct
S
T
3
2
6
3
1
3
1
2
13
2
23
2
12
2
11 33
2
33 22
2
22 11
2
1 3
2
3 2
2
2 1
-

- - - - - - - -
- - - - -
T T T T T T T T T
T T T T T T T

(A.44)
or in terms of the principal values of the deviatoric tensor
dev
T :
( ) ( ) ( )
3
2
3
2
2
2
1
dev dev dev
oct
S
T T T
T
- -

(A.45)
Then, in summary, we have:
m
oct
N
I
T T
3
T
Octahedral
normal component
(A.46)
( ) ( ) ( )
3 3
2
6 2
3
1
2
3
2
2
2
1 2
dev dev dev
oct
S
dev
I I I I I
T T T
T
- -

-
-
T
T T
Octahedral
tangential
component
(A.47)
Notice that the octahedral normal and tangential components are the same for the 8
octahedral planes, (see Figure A.14).
Then, let us consider the principal space once again, here represented by the orthonormal
basis (
3 2 1

e e e ), (see Figure A.15). Now, we can plot the coordinates ) , , (


3 2 1
T T T , which
are denoted by the P in Figure A.15.
Note, any plane normal to the straight line OA is an octahedral (deviatoric) plane and the
specific plane passing through the origin is denoted by H . Finally, the straight line OA is
called the spherical axis (or hydrostatic axis).
Let us now consider a deviatoric plane passing through the point P , which is denoted by
H , (see Figure A.15). Then, we will define three vectors, namely,

OP,

OA and

AP.
Next, the vector

OP can be expressed in terms of the principal values of T as:


3 3 2 2 1 1

e e e - -

T T T OP
(A.48)
and according to Figure A.15, the magnitude of

OA is:
( )
oct
N m
m
OP p OA
T T
T
T T T
T T T
3 3
3
3
3 3


3 2 1 3 2 1
3 3 2 2 1 1

- -

|
|
.
|

\
|
-

- -

e e e
e e e n
(A.49)
APPENDIX A: GRAPHICAL REPRESENTATION OF A SECOND-ORDER TENSOR

141
oct
N m
p T T 3 3

(A.50)












Figure A.14: Vectors on the octahedral planes.



Figure A.15: Principal space.
Therefore, the vector

OA is defined as:
3 2 1
3 2 1

3

e e e
e e e
n - -
|
|
.
|

\
|
-

m m m m
OA OA T T T T (A.51)
Then, the coordinates of point A are (
m m m
T T T , , ) and once the vectors

OP and

OA
have been defined, the vector

AP can be evaluated by adding the following vectors, (see


Figure A.15):

1
T

2
T

3
T

oct
S
T
,


oct
N
T
,

oct
N
T
,


oct
N
T
,


oct
N
T
,


oct
S
T
,

oct
S
T
,


oct
S
T
,


(
(
(

1
1
1
3
1

i
n

2
T

3
T

1
T
O
Spherical axis
( )
m m m
A T T T , ,
) , , (
3 2 1
T T T P

3 2 1
T T T
H
n


p
q
H
Deviatoric plane
(Octahedral plane)
o
o
o

3

e

1

e

2

e
H -plane
(deviatoric plane)
NOTES ON CONTINUUM MECHANICS

142


- OA OP AP
(A.52)
Then, taking into account (A.48) and (A.51), the above equation becomes:
( )
3 3 2 2 1 1
3 3 2 2 1 1
3 2 1 3 3 2 2 1 1

) (

) (

) (

e e e
e e e
e e e e e e
- -
- - - - -
- - - - -

dev dev dev


m m m
m m m
AP
T T T
T T T T T T
T T T T T T
(A.53)
and using the definition
ij m ij
dev
ij
c T T T - , the above can also be expressed as:
3 3 2 2 1 1

e e e - -

dev dev dev


AP T T T
(A.54)
Notice that, the components of

AP represent the principal values of the deviatoric part


dev
ij
T . The magnitude of

AP is then given by:


( ) ( ) ( )
2
3
2
2
2
1
dev dev dev
AP q T T T - -


(A.55)
Taking into account the expression of
dev
I I
T
, given by
2
3
2
2
2
1
) ( ) ( ) ( 2
dev dev dev
dev
I I T T T - - -
T
,
the above equation becomes:
oct
S
dev
I I q T 3 2 -
T

(A.56)
Note, we could also have obtained the

AP module by using Pythagoras theorem:


( )
2
3 2 1
2
3
2
2
2
1
2 2 2
3
1
T T T T T T - - - - - -

OA OP AP
(A.57)
( ) q AP - - - - -

1 3 3 2 2 1
2
3
2
2
2
1
3
2
T T T T T T T T T (A.58)
where q indicates how faraway is the tensor state is from the spherical state, (see Figure
A.15).
Note, in some cases working in the H -plane could be advantageous and for this reason we
define some parameters on it.
We will next analyze the projection of the principal space onto the H -plane, (see Figure
A.16).
Then, to find the components of the unit vector
3 3 2 2 1 1 1

e e e e - - a a a described in
Figure A.16, we will consider the principal space as shown in Figure A.17, where
1
3
2
sin cos a c , ,
3 2
a a holds. If we also consider that the hydrostatic axis is
orthogonal to the deviatoric plane, we obtain:
( ) ( ) ( )
6
2
3
2
0
3
1
0

3
1

1 3 2
3 2 1 3 2 1 3 3 2 2 1 1 1
- - - -
- - - - - -
a a a
a a a a a a e e e e e e n e
(A.59)
In addition to this we have:
APPENDIX A: GRAPHICAL REPRESENTATION OF A SECOND-ORDER TENSOR

143
6
1
3 2
2
2
1
1 3 2
- - - a a a (A.60)












Figure A.16: Projection of the principal space onto the H -plane.









Figure A.17: Principal space.
Thus:
)

2 (
6
1

3 2 1 1
e e e e - -
(A.61)
The, the projection of OP onto
1

e is evaluated as follows:
0 - -
- - - -
cos ) 2 (
6
1
)

2 (
6
1
)

(

3 2 1
3 2 1 3 3 2 2 1 1 1
q
OP
dev dev dev
dev dev dev
T T T
T T T e e e e e e e
(A.62)
Then, if we consider that
dev dev dev dev dev dev
3 2 1 3 2 1
0 T T T T T T - - - - , the above equation
becomes:

1
T

3
T
2
T
O
0
P
0 cos q
Q
H , H

1

e
120
q

1
T

3 2
, T T
c

3
1
cos c
n

e

3 2

e e
n



2
T

3
T

1
T
3
2

1

e

1

e

2

e

3

e
c ,
NOTES ON CONTINUUM MECHANICS

144
dev dev dev dev
OP q
1 1 1 1 1

2
3
6
3
) 2 (
6
1

cos T T T T - 0 e (A.63)
and by considering that
dev
I I q
T
2 - , we obtain:
0 -
-
0
0 - 0
cos
3
2
2
3
cos

2
3
cos 2
2
3
cos
1
1
1 1
dev
dev
dev
I I
I I
I I q
dev
dev
dev dev
T
T
T
T
T
T T

(A.64)
Likewise, we can find
dev
2
T and
dev
3
T . Then, we can express the principal values,
dev
ij ij m ij
T T T - c , as follows:
(
(
(

- 0
- 0
0
- -
(
(
(

(
(
(

-
(
(
(

(
(
(

r
r
) cos( 0 0
0 ) cos( 0
0 0 cos
3
2
0 0
0 0
0 0
0 0
0 0
0 0
0 0
0 0
0 0
0 0
0 0
0 0
3
2
3
2
3
2
1
3
2
1
dev
I I
m
m
m
dev
dev
dev
m
m
m
T
T
T
T
T
T
T
T
T
T
T
T
T

(A.65)
with 3 / 0 r s 0 s . Note, the tensor state can also be expressed in terms of ( 0 , , q p ):
_ _
part
Deviatoric
part
Spherical
q
p
p
p
(
(
(

- 0
- 0
0
-
(
(
(

(
(
(

r
r
) cos( 0 0
0 ) cos( 0
0 0 cos

3
2
0 0
0 0
0 0
3
1
0 0
0 0
0 0
3
2
3
2
3
2
1
T
T
T

(A.66)
Then, substituting the 0 cos , given by the equation in (A.64), into the trigonometric
relationship 0 - 0 0 cos 3 cos 4 3 cos
3
, yields:
( )
(

-
-

|
|
.
|

\
|
-
-
|
|
.
|

\
|
-
0
dev
dev
dev dev
I I
I I
I I I I
dev dev
dev dev
T
T
T T
1
3
1
3
1
3
1
2
3 3
2
3
3
2
3
4 3 cos T T
T T
(A.67)
and if we take into account that ) (
3 1 3 2 2 1
dev dev dev dev dev dev
dev
I I T T T T T T - -
T
, the above
equation becomes:
( ) ( ) ( )
(

(
- - -

-
0
|
|
.
|

\
|
-
-
|
|
.
|

\
|
-
0
-
_ _
dev
dev dev
I I I
dev dev dev dev dev dev dev
dev dev
I I
I I I I
T
T
T T
3
3 2 1 3 2
2
1
3
1
3
1
3
1
1
2
3 3
3 cos
2
3
3
2
3
4 3 cos
J
T T T T T T T
T T
T

(A.68)
3
2
3 3
3 cos
dev
dev
I I
I I I
T
T
-
0
(A.69)
Note, the terms
dev
I I
T
,
dev
I I I
T
are invariants, hence 0 3 cos is an invariant too.
2 Continuum Kinematics











2.1 Introduction
A material body (continuum) in motion, starting from an initial state
0
) 0 ( t t = , will have
different configurations over time, (see Figure 2.1). In this Chapter we will study the
description of motion also called Kinematics, thus establishing the equations of motion that allow
us to characterize how the continuum evolves and how continuum properties, e.g.
displacement, velocity, acceleration, mass density, temperature, etc., change over time. To
do this, we will consider the initial configuration, also known as the reference or undeformed
configuration, characterized by material body
0
B at time 0 t , and we will also consider the
generic configuration
t
B at time t called the current configuration, also known as the actual or
deformed configuration.






Figure 2.1: Motion of a material body.
We will start by describing the motion of a single particle in the continuum. Then, we will
study how the relative distances between particles change during this motion and
afterwards we will define some deformation and strain tensors, but before we can carry out
these objectives we will define some continuum properties.
2
Continuum Kinematics

0
B

t
B
Reference
Configuration -
0
t
Configuration at
1
t

1
B
Current Configuration - t
,
145 , Notes on Continuum Mechanics, Lecture Notes on Numerical
Methods in Engineering and Sciences 4, DOI 10.1007/978-94-007-5986-2_4,
International Center for Numerical Methods in Engineering (CIMNE), 2013
E.W.V. Chaves
NOTES ON CONTINUUM MECHANICS

146
2.2 The Continuous Medium
Any continuous medium (continuum) is assigned a positive scalar quantity called mass. It is
assumed that the mass is continuously distributed in the continuum, without there being
any discontinuities. A continuum is said to be homogeneous if its properties are the same
throughout the continuum.
Now, let us consider a sphere centered at point P in the initial configuration, (see Figure
2.2). The volume and mass of this sphere are denoted by
0
V A and m A , respectively. Then,
we can define the mass density of the particle in the initial configuration as:
0 0
0
0
0
) (
dV
dm
V
m
im
V

A
A

,
X j
(

3
m
kg
(2.1)
Starting from this concept we can define a particle as a dimensionless element that has
physical properties such as mass density, velocity, temperature, etc.








Figure 2.2: Mass density in the initial configuration.
We can define a continuum medium as a set of particles arranged in an area without
discontinuities, in which there is a one-to-one correspondence (i.e. bijection) between
possible configurations. Then we can define some terminologies that are used throughout
this chapter, e.g.:
Particle (material point): a small volume element that has certain physical properties, e.g. mass
density ( j ), velocity ( v
,
), temperature ( T ), etc.
Points: A place in space, position;
Particle trajectory (or path line): The locus of the points occupied by a single particle during
motion, (see Figure 2.3).







Figure 2.3: Particle trajectory.

_
,
) (
0
0
0
X j

AV

2
X

3
X

1
X

0
) 0 ( t t =
Reference
configuration

0
B

3
t - Current position
P

0
t

1
t

2
t
Trajectory of particle P
P
P
P
j , v
,
, T
at
2
t
2 CONTINUUM KINEMATICS

147
2.2.1 Kinds of Motion
Motion of a continuous medium, also denoted by deformation, is characterized by the
following types:
Rigid Body Motion: Characterized by maintaining the original shape of the body after motion,
i.e. it is characterized by preserving the distance between particles. The rigid body motion
can be classified by: translation and/or rotation.
Motion with Deformation: Characterized by changes of distance between particles.
In general, motion is characterized by deformation and rigid body motion simultaneously.
2.2.1.1 Rigid Body Motion
As we have seen previously, in rigid body motion the distances between particles remain
unchanged. Then we can establish an equation that governs this motion. To do this let us
consider a Cartesian system
3 2 1
X X OX which is attached to the body, so, the position
vector of any particle with respect to this system remains unchanged during motion. We
can also adopt a second Cartesian system
3 2 1
x x ox , which is represented by the
orthonormal basis (
1

e ,
2

e ,
3

e ), (see Figure 2.4).















Figure 2.4: Rigid body motion.
If X
,
and x
,
are the position vectors of material point P with respect to the systems
i
e


and
i
I

, respectively, (see Figure 2.4), the following relationship is satisfied:


X x
,
,
,
- c
(2.2)
where ) (t c
,
is a time-dependent vector that describes the translation motion of the system
i
I

. The above equation (2.2) in symbolic notation can be represented as:


j j k k p p
X x I e e


- c (2.3)
o
3

e

1
x

2
x

3
x
) (t c
,


1

e

2

I

3

I

3
X

1
X

2
X

1

I
O
P
x
,

X
,


2

e
time - t

0
t
X x
,
,
,
- ) ( ) ( t t Q c
NOTES ON CONTINUUM MECHANICS

148
The components of (2.3) in the system
3 2 1
x x ox are obtained by means of the dot product
with respect to the basis
i
e

, i.e.:
ji j i i
ji j ik k pi p
i j j i k k i p p
a X x
a X x
X x
-
-
-
c
c
c
c c
e I e e e e


(2.4)
in where
ji i j
a e I

is the transformation matrix from the system


i
I

to the system
i
e

,
and where it holds that
ij kj ik
a a c , i.e.
ji
a is an orthogonal matrix. Note also that the
equation (2.4) holds for any adopted system. If we consider that
ji ij
a Q , where
ij
Q are
orthogonal tensor components, we can sum up the equation in (2.4) as:
X x
,
,
,
- Q c Rigid body motion equations (2.5)
which describes rigid body motion.
NOTE: The transformation law of components and orthogonal transformation are closely
interrelated although they have completely different meanings.
Problem 2.1: A continuum is defined by a square with sides b , subjected to rigid body
motion which is defined by rotating the continuum counterclockwise by an angle of 30 to
the origin. Find the equations of motion. Also obtain the new position of particle D.
Hint: Consider the systems x
,
and X
,
to be superimposed.













Solution: We apply the rigid body motion equations X X x
, ,
,
,
- Q Q c , to 0 c
,
,
. The
components of Q are the same as the components of the transformation matrix from the
x
,
-system to the x
,
-system, i.e.:
(
(
(

0 0
0 - 0

1 0 0
0 cos sin
0 sin cos
ij
Q
So, the continuum particles are governed by the equations of motion:

(
(
(

3
2
1
3
2
1
1 0 0
0 30 cos 30 sin
0 30 sin 30 cos
X
X
X
x
x
x

A particle which initially was at point D ( 0
1
X , b X
2
, 0
3
X ) moves into the following
position:
C
D
C
30
B
D
B

1 1
, x X

2 2
, x X
b
b
A A

1
x

2
x
30
2 CONTINUUM KINEMATICS

149

(
(
(

0
30 cos
30 sin
0
0
1 0 0
0 30 cos 30 sin
0 30 sin 30 cos
3
2
1
b
b
b
x
x
x
D
D
D


In the above example we have adopted the system, X
,
, fixed in space and time. When we
are establishing the equations of motion for a continuum subjected to deformation, we can
adopt a system fixed in space and time which is called the material system. There is also a
second system defined by x
,
and is called the spatial system.
In general, throughout this chapter we will adopt 0 c
,
,
, (see Figure 2.5), or to put it
another way, the spatial and the material axes will be superimposed as shown in Figure 2.5.











Figure 2.5: Spatial and material axes superimposed.
2.2.2 Types of Configurations
We define two types of configurations adopted in this chapter, (see Figure 2.6), namely:
The Reference configuration or the initial configuration: the configuration at the instant of
time
0
) 0 ( t t = , is considered to be the undeformed configuration in which the
particle P is identified by the position vector
P
X
,
.
The Current configuration or deformed configuration: the configuration at the instant of time
t .
As we have seen before, a continuum is defined as a set of particles arranged in an area
without discontinuities, in which there is a one-to-one correspondence (i.e. bijection)
between possible configurations. So, if motion is characterized by the bijective function, ,,
this ensures the existence of the inverse function
1 -
, , (see Figure 2.6).




3 3
, x X

1 1
, x X

2 2
, x X

3 3

e I

2 2

e I

1 1

e I
O
NOTES ON CONTINUUM MECHANICS

150












Figure 2.6: Initial and current configurations.
2.2.2.1 Mass Density
As with the definition of mass density in the reference configuration given in (2.1), we
define mass density in the current configuration, (see Figure 2.7), as:
dV
dm
V
m
im
V

A
A

A 0
j
(

3
m
kg

(2.6)
Mass density is a scalar field, which is a function of position and time, i.e. ) , ( t x
,
j j .












Figure 2.7: Mass density.

1 -
,

1 1
, X x

2 2
, X x

3 3
, X x

3 3

e I

2 2

e I

1 1

e I
O

0
B

t
B
P
P
Reference
configuration -
0
t
Current configuration - t
Trajectory of particle - P
) (
0
X
,
j
) , ( t x
,
j
,

_
,
) (
0
0
0
X j

AV

2
x

3
x

1
x

0
) 0 ( t t =
t

_
,
) , (
0
t
V
x j

A
Initial configuration
Current configuration
,
2 CONTINUUM KINEMATICS

151
2.3 Description of Motion
2.3.1 Material and Spatial Coordinates
Let us consider the material body
0
B in the initial configuration, (see Figure 2.8). At an
arbitrary time ) (t , the material body occupies a new position in space,
t
B . We now focus
our attention on a particle of the continuum, which is denoted by the particle P .
NOTE: With regard to nomenclature, the particles are identified by labels. These labels are
the positions they occupied in the reference configuration. For instance, the particle which
occupied the point ) , , (
3 2 1
X X X P in the reference configuration will be denoted by
particle P , (see Figure 2.8).













Figure 2.8: Initial and current configurations.
The position of a particle is characterized by the position vector. The position vector of
particle P in the reference configuration,
0
0 t t = , is given by:
3 3 2 2 1 1

e e e X X X - - X
,
(2.7)
which thus defines the material coordinate:
(
(
(

3
2
1
X
X
X
X
i

(2.8)
In the current configuration (deformed configuration) particle P occupies the position P ,
and the position vector is given by:
3 3 2 2 1 1

e e e x x x - - x
,
(2.9)
which defines the spatial coordinate:

1 1
, X x

2 2
, X x

3 3
, X x

0
B

t
B

3 3

e I

2 2

e I

1 1

e I
O
P
P
X
,

x
,

Initial configuration -
0
t
Current configuration - t
u
,

Trajectory of particle P
NOTES ON CONTINUUM MECHANICS

152
(
(
(

3
2
1
x
x
x
x
i

(2.10)
2.3.2 The Displacement Vector
By definition, the displacement vector ( u
,
) of a particle is the difference between the
position vector in the current configuration ( x
,
) and the position vector in the reference
configuration ( X
,
), (see Figure 2.8), i.e.:
X x
,
, ,
- u

i i i
X x - u
[ ] m (2.11)
2.3.3 The Velocity Vector
The velocity of a particle is defined by the rate of change of the position vector, i.e.:
( )

u
u u u
0
`
,
,
,
,
,
,
`
,
,
,
,
= -
-
=

dt
d
dt
d
dt
d
dt
d
dt
d X X
x
x
V
(

s
m
(2.12)
2.3.4 The Acceleration Vector
The acceleration of a particle is the rate of change of velocity, i.e.:
u
` `
,
` `
,
,
`
,
,
,
= = x
x
V
V
A
2
2
dt
d
dt
d

(

2
s
m

(2.13)
2.3.5 Lagrangian and Eulerian Descriptions
Continuum properties, e.g.: mass density, temperature, velocity, acceleration, etc., are
intrinsic in particles (material points), and such properties may change over time. As
mentioned before, continuum motion is characterized by the bijective function (,), and
the inverse function (
1 -
, ). This ensures that we can correlate continuum properties
between the current and reference configurations. In other words, the study of motion can
be carried out either in the current or reference configuration.
2.3.5.1 Lagrangian Description of Motion
The particle in motion can be described in terms of material coordinates ( X
,
) and time as:
( ) t , X x x
,
, ,
( ) ( ) t x t X X X x x
i i i
, , , ,
3 2 1
X
,
(2.14)
The equations of motion (2.14) are called the Lagrangian or Material Description of the motion.
The above parametric equation gives us the current position x
,
, at time t , of a particle that
occupied position X
,
in the reference configuration, at time
0
t . The equation in (2.14),
applied to particle P , provides us with the unique path line (trajectory) of the particle, (see
Figure 2.9).
2 CONTINUUM KINEMATICS

153
2.3.5.2 Eulerian Description of Motion
Particle motion can also be described in terms of spatial coordinates ( x
,
) and time as:
( ) t , x X X
,
, ,

( ) ( ) t X t x x x X X
i i i
, , , ,
3 2 1
x
,

(2.15)
The above equation give us the original position X
,
, at time
0
t , of a particle which at the
present time ( t ) has the coordinates ) , , (
3 2 1
x x x , (see Figure 2.9).
The necessary and sufficient condition for there to be an inverse is:
0
3
3
2
3
1
3
3
2
2
2
1
2
3
1
2
1
1
1

o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o

o
o

X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
J
j
i

(

3
3
m
m
(2.16)
where J is called the Jacobian determinant, also knows as the volume ration.

In view of Figure 2.9 we can observe that:
( )
P P
t X X x
, ,
,
0 , (2.17)
and

( )
( )
( )
_
,
,
,
,
,
,
,
,
,
P particle of Trajectory
P P
P P
P P
t
t
t

x X x
x X x
x X x
2
1
0
,
,
,

( )
( )
( )
_
,
,
,
,
,
,
,
,
,
time different at
P, point at Particles
Q Q
S S
P P
t
t
t
X x X
X x X
X x X

2
1
0
,
,
,

(2.18)
2.3.5.3 Lagrangian and Eulerian Variables
Any physical quantity ( Z) assigned to the continuum ( B) can be expressed as a
Lagrangian ( ) , ( t X
,
Z ) or Eulerian ( ) , ( t x
,
z ) description, and are related in the following
ways:
) , ( ) ), , ( ( ) , ( ; ) , ( ) ), , ( ( ) , ( t t t t t t t t X X x x x x X X
, ,
, , , ,
, ,
Z z z z Z Z (2.19)
NOTE: Some authors try to differentiate a Lagrangian from Eulerian variable by using
upper and lower-case letters, respectively. As a general rule we have not adopted this
convention in this publication. In this textbook when we are dealing with a Lagrangian
variable we will indicate explicitly by its arguments, i.e. ) , ( t X v V
,
,
,
= . And if we are dealing
with an Eulerian variable it will be indicated as follows ) , ( t x v
, ,
.

OBS.: The Axiom of Impenetrability: Two particles can not occupy the same place at
the same time. As discussed later, this condition is ensured when the Jacobian
determinant is positive, i.e. 0 > J .
NOTES ON CONTINUUM MECHANICS

154

















Figure 2.9: Lagrangian and Eulerian description.
Problem 2.2: Consider the following equations of motion in the Lagrangian description:

(
(
(

-
-
3
2
1
2
3
2
1
form Matrix
3 3
2 3 2
1
2
2 1
1 0 0
1 0
0 1
) , (
) , (
) , (
X
X
X
t
t
x
x
x
X t x
X t X t x
X t X t x
X
X
X
,
,
,

(2.20)
Is the motion above possible? If so, find the displacement, velocity and acceleration fields
in Lagrangian and Eulerian descriptions. Consider a particle P that at time 0 t was at the
point defined by the triple equation 3 , 1 , 2
3 2 1
X X X . Find the velocity of P at time
s t 1 and s t 2 .

Solution: Motion is possible if 0 J , thus
0 1
1 0 0
1 0
0 1
2
3
3
2
3
1
3
3
2
2
2
1
2
3
1
2
1
1
1

o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o

o
o
t
t
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
J
j
i

The displacement vector field is given by the definition in (2.11), X x
,
, ,
- u . Using the
equations of motion (2.20) we obtain:
( ) [ ]
( ) [ ]
( ) [ ]

- -
- - -
- - -
0 ) , ( ,
) , ( ,
) , ( ,
3 3 3 3 3
3 2 2 3 2 2 2
2
2 1 1
2
2 1 1 1
X X X t x t
t X X X t X X t x t
t X X X t X X t x t
X X
X X
X X
, ,
, ,
, ,
u
u
u
(2.21)
Q

Q
X
,


S
X
,


P P
x X
,
,

S
P

0
t
Reference configuration
at X x
,
,

0
t t

1
t
P S =
P

2
t
S
P

P
x
,


P
x
,


S S
t X x X
,
,
,

) , (
1

Trajectory of particle P
Particle P
Particle P
Particle Q
Q

S
x
,


Q
x
,


S S P
t X x X
,
,
,

=
) , (
1


P P
t

x X x
,
,
,
) , (
1


Q P
t X x X
,
,
,
) , (
2


P P
t

x X x
,
,
,
) , (
2

point Q P =
2 CONTINUUM KINEMATICS

155
which are the components of the displacement vector in the Lagrangian description. Here,
velocity and acceleration can be evaluated as follows:
( )
( )
( )
( )
( )
( )

=
=
=

=
=
=
0 ) , (
0 ) , (
2 ) , (
;
0
,
) , (
,
) , (
2
,
) , (
3
3 3
2
2 2
2
1
1 1
2
3
3 3
3 3
2
2 2
2
2
2
1
1 1
dt
dV
t a A
dt
dV
t a A
X
dt
dV
t a A
t X
dt
d
dt
t d
t v V
X t X
dt
d
dt
t d
t v V
t X t X
dt
d
dt
t d
t v V
X
X
X
X
X
X
X
X
X
,
,
,
,
,
,
,
,
,
u
u
u

(2.22)
The inverse form of (2.20) provides us the equations of motion in the Eulerian description:

-
- -

(
(
(

-
-

3 3
3 2 2
3
3
2
2
1 1
3
2
1
3 2
3
2
1
) , (
) , (
) , (
1 0 0
1 0
1
x t X
tx x t X
x t x t x t X
x
x
x
t
t t
X
X
X
x
x
x
,
,
,
(2.23)
Then, the displacement, velocity and acceleration fields in Eulerian description can be
evaluated by substituting equation (2.23) into the equations (2.21) and (2.22), i.e.:


( )
( )
( )



-
0 ) , ( ), , (
) , ( ) , ( ), , (
) , ( ) ( ) , ( ), , (
3 3
2 3 3 2
1
2
3 2
2
2 1
t t t
t t x t t X t t
t t tx x t t X t t
x x X
x x x X
x x x X
, ,
,
, , ,
,
, , ,
,
u u
u u
u u

(2.24)

( )
( )
( )



-
0 ) , ( ), , (
) , ( ) , ( ), , (
) , ( ) ( 2 ) , ( 2 ), , (
3 3
2 3 3 2
1 3 2 2 1
t v t t V
t v x t X t t V
t v t tx x t t X t t V
x x X
x x x X
x x x X
, ,
,
, , ,
,
, , ,
,
(2.25)

( )
( )
( )



-
0 ) , ( ), , (
0 ) , ( ), , (
) , ( ) ( 2 ) , ( 2 ), , (
3 3
2 2
1 3 2 2 1
t a t t A
t a t t A
t a tx x t X t t A
x x X
x x X
x x x X
, ,
,
, ,
,
, , ,
,
(2.26)
Taking into account the Lagrangian description of velocity given in (2.22), the velocity of
particle P ( 3 , 1 , 2
3 2 1
X X X ) at time s t 1 is given by:
0 ) , ( ; / 3 ) , ( ; / 2 2 ) , (
3 3 2 2 1
t v s m X t v s m t X t v X X X
, , ,

We can also observe that at time s t 1 the particle P occupies the position:
3 ; 4 ; 3
3 3 2 3 2 1
2
2 1
- - X x X t X x X t X x
So, the velocity of the particle P can also be evaluated by (2.25) as:


- -
0 ) , (
/ 3 ) , (
/ 2 1 ) 3 1 4 ( 2 ) ( 2 ) , (
3
3 2
3 2 1
t v
s m x t v
s m t tx x t v
x
x
x
,
,
,

Note that, the velocities obtained via the Lagrangian or Eulerian description are the same,
since velocity is an intrinsic property of the particle.
We can also provide the velocity of the particle P at time s t 2 :
( )
( )
( )

=
=
=
0 ,
/ 3 ,
/ 4 1 2 2 2 ,
3 3
3 2 2
2 1 1
t v V
s m X t v V
s m t X t v V
X
X
X
,
,
,

At time s t 2 the new position of P is:
NOTES ON CONTINUUM MECHANICS

156


-
-
3 ) , (
7 ) , (
6 ) , (
3 3
2 3 2
1
2
2 1
X t x
X t X t x
X t X t x
X
X
X
,
,
,

As we can verify the Lagrangian description of motion ) , ( t X x
,
,
describes the trajectory of
P .



















2.4 The Material Time Derivative
The rate of change of a physical quantity, e.g.: velocity, temperature, mass density, etc., is
called the material time derivative, and is denoted by
Dt
D
. For example, let us consider an
observer who is travelling with particle P and is recording how temperature changes over
time, (see Figure 2.10); this rate of change of temperature is denoted by the material time
derivative of temperature. By this example, it seemed reasonable to conclude that the
material time derivative of a property depends on whether the property is a Lagrangian or
Eulerian variable.
If the property is in Lagrangian description we have:
) , , , (
3 2 1
t X X X 0 0
(2.27)
In this case, the material time derivative is expressed in the form:
dt
t d
Dt
t D
t
) , ( ) , (
) , (
X X
X
, ,
,
`
0

0
= 0

(2.28)
When the property is described in terms of material coordinates, this implication is that the
property represented is connected to the same particle during motion.


P
[ ] 3 ; 1 ; 2
P
i
X

0
t
Trajectory of particle P
s t 1
P
[ ] 3 ; 4 ; 3
P
i
x
[ ] 0 ; 3 ; 2 ) 1 , ( s t V
P P
i
X
,

[ ] 0 ; 3 ; 2 ) 1 , ( s t v
P
i
x
,

s t 2
P
[ ] 0 ; 3 ; 4 ) 2 , ( s t v
P
i
x
,

[ ] 3 ; 7 ; 6
P
i
x
[ ] 0 ; 3 ; 4 ) 2 , ( s t V
P P
i
X
,

2 CONTINUUM KINEMATICS

157









Figure 2.10: The rate of change of temperature Lagrangian description.
If the property is in Eulerian description we have:
) , , , (
3 2 1
t x x x 0 0
(2.29)
In this description the observer is not traveling with the particle, but fixed at one point
) , , (
3 2 1
x x x watching the particles passing. According to Figure 2.11, equation (2.29), at
time
1
t , provides us with the property of particle Q , which takes the point ) , , (
3 2 1
x x x .
Later,
2
t , in equation (2.29) give us the property of another particle, e.g. R , and at time
3
t ,
the equation (2.29) give us the value of the property of particle P , (see Figure 2.11).










Figure 2.11: The rate of change of temperature Eulerian description.
It must be emphasized that the material time derivative is related to the derivative with
respect to time of an intrinsic property of the particle, i.e. it is related to the same particle.
But an observer fixed at a spatial point ) , , (
3 2 1
x x x only has information on the local rate of
change. In order to be fully informed, we need to know how the property of this particle
changes along its path line- this additional information is known as the convective rate of
change, which is related to mass transport. Afterwards, to obtain the material time derivative
of an Eulerian property ) , ( t x
,
0 0 we must take into consideration:
The local rate of change;
The convective rate of change
So,
P
P

0
0
P

0
t
1
t
2
t

X X x
, ,
,
) 0 , (

P
particle trajectory

) , , (
3 2 1
X X X

P
particle trajectory

) , , (
3 2 1
X X X

particle trajectory

) , (
1
t X x
,
,


) , (
2
t X x
,
,


1
0

2
0

Q

Q

Q

R
P
Trajectory of particle
Q


1
t
2
t
R
P
trajectory of particle R
R
P
trajectory of particle P

3
t

) ( x
,

) ( x
,

) ( x
,


t
t
P
o
0 o ) , (
3
x
,
t
t
R
o
0 o ) , (
2
x
,

t
t
Q
o
0 o ) , (
1
x
,

NOTES ON CONTINUUM MECHANICS

158
) , (
) , ( ) , (
) , ( ) , ( ) , ( ) , (
) , (
change of rate
Convective
change of rate
local
t v
x
t
t
t
t
t x
x
t
t
t
Dt
t D
t
k
k
k
k
X
x x
X x x x
x
,
, ,
_
,
,
_
, ,
,
`
o
0 o
-
o
0 o

o
o
o
0 o
-
o
0 o

0
= 0
(2.30)
where ) , ( ) , ( t t X x X v
,
`
,
,
,
= is the particle velocity, which can also be expressed in Eulerian
description by substituting the equations of motion, i.e. ) , ( ) ), , ( ( t t t x v x X v
, , ,
,
,
.
Then, we can define the material time derivative operator for an Eulerian property, ) , ( t x
,
- ,
as:
) , ( ) , (
) , ( ) , (
t t
t
t
Dt
t D
x v x
x x
x
, , ,
, ,
,
-
o
o


V
Material time derivative
for an Eulerian variable
(2.31)
or in indicial notation as:
k
k
v
x
t
t
t
Dt
t D
o
o
-
o
o

) , ( ) , ( ) , ( x x x
, , ,


(2.32)
2.4.1 Velocity and Acceleration in Eulerian Description
We have defined the velocity of particle P as:
( )
dt
t d
dt
d
dt
d
t
Dt
D
t
P
P P
) , (
) , ( ) , (
X X x
x X x X V
x x
,
,
,
, ,
`
,
,
,
, ,
u u

-
|
.
|

\
|

(2.33)
which is the Lagrangian velocity. To obtain the Eulerian velocity, ) , ( t x v
, ,
, we have to
substituting the inverse equation of motion, i.e. ) , ( ) ), , ( ( ) , ( t t t t
P P P
x v x X V X V
, , ,
, , , ,
.
The Lagrangian acceleration was obtained as follows:
) , ( ) , (
2
2
t
Dt
D
t
P P P
X x x V X A
,
,
` `
,
`
, , ,
= (2.34)
The Eulerian acceleration can be evaluated either by substituting the inverse equation of
motion into the equation in (2.34) or via the definition of the material time derivative for
an Eulerian property, i.e.:
_
,
, , , ,
, ,
on accelerati
convective
k
k
i i k
k
i i
i
P
i
t v
x
t v
t
t v
t
x
x
t v
t
t v
t v
Dt
D
t a ) , (
) , ( ) , ( ) , ( ) , (
) , ( ) , ( x
x x x x
x x
o
o
-
o
o

o
o
o
o
-
o
o

) , (
) , ( ) , (
) , ( t
t
t
Dt
t D
t x v v
x v x v
x a
x
, , ,
, , , ,
, ,
,
-
o
o
= V
(2.35)
The Eulerian acceleration in matrix form can be evaluated as follows:
2 CONTINUUM KINEMATICS

159

(
(
(
(
(
(
(

o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
-

o
o
o
o
o
o

3
2
1
3
3
2
3
1
3
3
2
2
2
1
2
3
1
2
1
1
1
3
2
1
3
2
1
) , (
v
v
v
x
v
x
v
x
v
x
v
x
v
x
v
x
v
x
v
x
v
t
v
t
v
t
v
a
a
a
t a
i
x
,
(2.36)
Returning to Problem 2.2, the Eulerian velocity field was obtained as:
0 ) , ( ; ) , ( ; ) ( 2 ) , (
3 3 2 3 2 1
- t v x t v t tx x t v x x x
, , ,
(2.37)
Explicitly, the Eulerian acceleration can also be evaluated by using the definition in (2.35):
(

o
o
-
o
o
-
o
o
-
o
o

o
o
-
o
o

) , (
) , (
) , (
) , (
) , (
) , ( ) , (
) , (
) , ( ) , (
) , (
3
3
2
2
1
1
t v
x
t v
t v
x
t v
t v
x
t v
t
t v
t v
x
t v
t
t v
t a
i i i i
k
k
i i P
i
x
x
x
x
x
x x
x
x x
x
,
,
,
,
,
, ,
,
, ,
,
(2.38)
Thus, the components
i
a are given by:
[ ] [ ]
0
) , (
) , (
) , (
) , (
) , (
) , ( ) , (
) , (
0
) , (
) , (
) , (
) , (
) , (
) , ( ) , (
) , (
) ( 2
0 2 0 4 2
) , (
) , (
) , (
) , (
) , (
) , ( ) , (
) , (
3
3
3
2
2
3
1
1
3 3
3
3
3
2
2
2
2
1
1
2 2
2
3 2
3 3 2
3
3
1
2
2
1
1
1
1 1
1

o
o
-
o
o
-
o
o
-
o
o

o
o
-
o
o
-
o
o
-
o
o

-
- - - -
(

o
o
-
o
o
-
o
o
-
o
o

t v
x
t v
t v
x
t v
t v
x
t v
t
t v
t a
t v
x
t v
t v
x
t v
t v
x
t v
t
t v
t a
t x x
t x t x x
t v
x
t v
t v
x
t v
t v
x
t v
t
t v
t a
P
P
P
x
x
x
x
x
x x
x
x
x
x
x
x
x x
x
x
x
x
x
x
x x
x
,
,
,
,
,
, ,
,
,
,
,
,
,
, ,
,
,
,
,
,
,
, ,
,
(2.39)
whose components are the same as those obtained in (2.26).
2.4.2 Stationary Fields
A field ) , ( t x
,
c is said to be stationary if the local rate of change does not vary over time:
) (
) , (
x
x ,
,
c c
c

o
o
0
t
t
Steady state (stationary) field (2.40)
For example, let us consider a stationary (steady state) velocity field as shown in Figure
2.12. Then, as we can verify, the field representation for any time, e.g.
1
t and
2
t , does not
change. However, that does not mean that the velocities of the particles do not change
over time. In light of Figure 2.12, we can now focus our attention on the fixed spatial point
*
x
,
. At time
1
t the particle Q is passing through point
*
x
,
with velocity
*
v
,
. Let us also
consider another particle P , which is passing through another point with velocity
*
1
) ( v v
, ,
t
P
. At time
2
t the particle P is now passing through the point
*
x
,
. It follows that
if we are dealing with a steady state velocity field, then the velocity of particle P at
*
x
,

must be
*
v
,
, i.e.
*
2
) ( v v
, ,
t
P
. We can easily contrast this with the material time derivative of
velocity, which is always associated with the same particle, i.e.:
NOTES ON CONTINUUM MECHANICS

160
) ( ) ( ) (
) , (
) , (
) , (
) (Stationay
x a x v v x v v
x v
x a
x v
x x
, , , , , , , ,
_
, ,
, ,
, ,
, ,
,
-
o
o
=

V V
0
t
t
t
Dt
t D

(2.41)
The rate of change of velocity (acceleration) will be zero if the velocity field is stationary
|
.
|

\
|

o
o
0
,
, ,
t
t) , ( x v
and homogeneous ( 0 v
x
,
,
V ).
We can also verify that, although spatial velocity is independent of time, that does not
mean material velocity is also, since:
) , ( )) , ( ( ) ( t t X v X x v x v
,
,
,
, , , ,
(2.42)
























Figure 2.12: Steady velocity field.
NOTE: As we can verify, there are two ways of analyzing the continuum. Either we can
follow the particles and see how their properties change over time or we can focus our
attention on a fixed spatial region and check how the continuum properties change over
time. While the first option is most often used in solid mechanics, the second one is
widespread in the field of fluid mechanics.

*
x
,


Q
t v v x v
, , , ,

*
1
*
) , (

1
t
Particle - Q
Particle- P
) ( x v
, ,


*
v v
, ,

P

) ( x v
, ,


P
t v v x v
, , , ,

*
2
*
) , (

2
t
Particle - P

*
x
,

2 CONTINUUM KINEMATICS

161
2.4.3 Streamlines
Given a spatial velocity field at time t , we can define a streamline to the curve in which the
tangent at each point has the same direction as the velocity. In general, the streamline and
the trajectory do not coincide, but in steady state motion they do.

Problem 2.3: The acceleration vector field is described by:
v v
v v
x a
x
, ,
, ,
, ,
,
-
o
o
V
t Dt
D
t) , (
Show that acceleration can also be written as:
v v
v
v v
v
v v
v v
x x x x
, ,
,
, ,
,
,
,
,
, ,
, , , ,
r -
|
|
.
|

\
|
-
o
o
r -
|
|
.
|

\
|
-
o
o
r r -
|
|
.
|

\
|
-
o
o

2 2
) (
2
2 2 2
rot rot
v
t
v
t
v
t Dt
D
V V V V
Solution:
To prove the above relationship one need only demonstrate that:
) (
2
2
v v v v
x x x
,
,
, , ,
, , ,
r r -
|
|
.
|

\
|
V V V
v

Expressing the terms on the right of the equation in symbolic notation we obtain:
( ) ( ) ( )
(

r
o
o
r -
(

o
o
r r -
|
|
.
|

\
|
s s r
r
i i j j
i
i
v
x
v v v
x
v
e e e e

2
1
) (
2
2
v v
x x
,
,
,
, ,
V V
Using the definition of the permutation symbol (see Chapter 1) we can express the vector
product as:
( ) ( )
k
r
s
i itk rst
i
j
j i
t
r
s
rst i i j j
i
i
x
v
v
x
v
v
x
v
v v v
x
v
e e
e e e

2
1

2
1
) (
2
2
o
o
-
(

o
o

o
o
r -
(

o
o
r r -
|
|
.
|

\
|

v v
x x
,
,
,
, ,
V V

where we have used the equation
k itk t i
e e e

r . In Chapter 1 we also proved that
sk ri si rk kit rst itk rst
c c c c - , then:
( )
k
i
k
i
k
s
s i
i
j
j
k
r
s
i sk ri
r
s
i si rk i
i
j
j
k
r
s
i sk ri si rk i
i
j
j
x
v
v
x
v
v
x
v
v
x
v
v
x
v
v
x
v
v
x
v
v
x
v
v
v
e e
e e
e e



) (
2
2
|
|
.
|

\
|
o
o
-
o
o
-
o
o

|
|
.
|

\
|
o
o
-
o
o
-
o
o

o
o
- -
o
o
r r -
|
|
.
|

\
|
c c c c
c c c c v v
x x
,
,
,
, ,
V V

v v
v
v v
x
x x
, ,
,
,
,
,
,
, ,

o
o

o
o

o
o
-
o
o
-
o
o

o
o
-
o
o
-
o
o

o
o
-
o
o
-
o
o
r r -
|
|
.
|

\
|
V
V V
i
i
i
i
k k
k
i
k
i i
i
s
s i
i
s
s
k
i
k
i i ik
k
s
s i
i
j
s sj
k
i
k
i k
k
s
s i
i
j
j
v
x
v
x
v
x
v
v
x
v
v
x
v
v
x
v
v
x
v
v
x
v
v
x
v
v
x
v
v
x
v
v
v
) (




) (
2
2
e
e e e
e e e
e e e
c c


NOTES ON CONTINUUM MECHANICS

162
Problem 2.4: Consider the equations of motion ) , ( t X x
,
,
and the temperature field ) , ( t T x
,

given by:
2
2
2
1
3 3
2 2
1 1
) ( ; ) 1 (
) 1 (
x x T
X x
t X x
t X x
-

-
-
x
,

Find the rate of change of temperature for the particle P at time s t 1 given that particle
P was at point ) 0 , 1 , 3 (
3 2 1
X X X at time 0 t .

Solution 1:
In this first solution we first obtain the material time derivative of the Lagrangian
temperature, so, we have to obtain the temperature in Lagrangian description ) , ( t T X
,

(Lagrangian temperature):
2 2
2
2 2
1
2
2
2
1
) 1 ( ) 1 ( ) , (
) (
t X t X t T
x x T
motion of equation the
tuting By substi
- - -
-
!
!
X
x
,
,

The material time derivative of the Lagrangian temperature is given by:
) 1 ( 2 ) 1 ( 2
) , (
) , (
2
2
2
1
t X t X
dt
t dT
Dt
DT
t T - - - =
X
X
,
,
`

By substituting s t 1 , ) 0 , 1 , 3 (
3 2 1
X X X , into the above equation we obtain:
40 ) 1 1 ( ) 1 ( 2 ) 1 1 ( ) 3 ( 2 ) 1 ( 2 ) 1 ( 2 ) , (
2 2 2
2
2
1
- - - - - - t X t X t T X
,
`

Solution 2:
In this second solution we directly use the definition of material time derivative of the
Eulerian variable, (see Eq. (2.30)). Then ) , (
) ( ) (
) , ( t v
x
T
t
T
Dt
DT
t T
k
k
x
x x
x
,
, ,
,
`
o
o
-
o
o
.
From the equations of motion we obtain:

-
-
0 ) , (
) , (
) , (
) 1 (
) 1 (
3
2 2
1 1
3 3
2 2
1 1
t v
X t v
X t v
X x
t X x
t X x
velocity
X
X
X
,
,
,

The equations of motion in Eulerian description are given by:

-
-
3 3
2
2
1
1

3 3
2 2
1 1
) 1 (
) 1 (
) 1 (
) 1 (
x X
t
x
X
t
x
X
X x
t X x
t X x
motion the of inverse

So, it is possible to obtain the Eulerian velocity as follows:

-

0 ) , (
) , (
) 1 (
) , ( ) ), , ( (
) , (
) 1 (
) , ( ) ), , ( (
3 3
2
2
2 2
1
1
1 1
t v V
t v
t
x
t X t t V
t v
t
x
t X t t V
x
x x x X
x x x X
,
, , ,
,
, , ,
,

2 CONTINUUM KINEMATICS

163
Afterwards, the material time derivative of the Eulerian temperature, ) , ( t T x
,
, is given by:
(

o
o
-
o
o
-
o
o
-
o
o
=

3
3
2
2
1
1
field) y (Stationar 0
) (
) , (
) , (
v
x
T
v
x
T
v
x
T
t
T
t T
Dt
t DT
_
,
,
`
,
x
x
x

) (
1
2
1
2
1
2
) , ( 0
1
2
1
2 ) , (
2
2
2
1
2
2
2
1 2
2
1
1
x x
t t
x
t
x
t T
t
x
x
t
x
x t T -
-

-
-
-
-
-
-
-
x x
,
`
,
`

The position of particle P at time s t 1 is evaluated as follows:


- -
- -
0
2 ) 1 1 ( 1 ) 1 (
6 ) 1 1 ( 3 ) 1 (
3 3
2 2
1 1
X x
t X x
t X x

Then, by substituting the spatial coordinates in the expression of the material time
derivative of temperature we obtain:
40 ) 2 6 (
1 1
2
) (
1
2
) 1 , 0 , 2 , 6 ( ) , (
2 2 2
2
2
1 3 2 1
-
-
-
-
x x
t
t x x x T t T
`
,
`
x
Alternatively, the expression ) , ( t T x
,
`
could also have been obtained as:
[ ] [ ]
) , ( ) (
) 1 (
2
) 1 (
) 1 (
2 ) 1 (
) 1 (
2 ) 1 ( ) , ( 2 ) 1 ( ) , ( 2 ) ), , ( (
) 1 ( 2 ) 1 ( 2 ) , (
2
2
2
1
2
2
2
1
2
2
2
1
2
2
2
1
t T x x
t
t
t
x
t
t
x
t t X t t X t t T
t X t X t T
x
x x x X
X
,
`
, , ,
,
`
,
`
-
-

-
(

-
- -
(

-
- - -
- - -

2.5 The Deformation Gradient
2.5.1 Introduction
In the previous section we studied the description of a particle in motion without looking
at how the relative motion between particles changed. In this section we analyze how
distances between particles change during motion after which we define some deformation
and strain tensors. To do this, let us consider two neighboring particles in the reference
configuration, which are denoted by P and Q , (see Figure 2.13).
2.5.2 Stretch and Unit Extension
Let X
,
d be a vector joining two points P and Q in the reference configuration, defining a
line element. The unit vector associated with the X
,
d -direction is represented by M

. After
motion, particles P and Q occupy new positions P and Q , respectively. In this new
configuration (current configuration), the vector joining the points P and Q is
represented by x
,
d , which is associated with the unit vector m , (see Figure 2.13), and the
magnitudes of X
,
d and x
,
d are denoted, respectively, by:
ds d Q P dS d PQ

x X
,
,
;
(2.43)

NOTES ON CONTINUUM MECHANICS

164














Figure 2.13: Continuum deformation.
Next we can define some parameters related to the magnitudes of line elements:
The Stretch (or Stretch ratio),
m
/ , associated with the m -direction is given by:
0

> / /
m m
X
x
where
dS
ds
d
d
,
,
Stretch (2.44)
The possible values of
m
/ are in the range:


< / <
ds ds
m
0
0 . And due to the axiom of
impenetrability 0 ds , the stretch has to be nonzero, 0

/
m
, if this is not so, the
implication is that two particles are occupying the same place at the same time which has
no physical meaning. Then, we can draw the conclusion that:
1

/
m
- There is no elongation;
1 0

< / <
m
- There is a shortening of the line element PQ;
1

> /
m
- There is an increase in distance between particles (elongation).
The Unit Extension,
m
, is defined as:
dS
dS ds
d
d d
-

X
X x
m
,
,
,

Unit extension (2.45)


The possible values of the unit extension are within the range of < < -
m
1 . The stretch
is related to the unit extension by:
dS dS ds
dS
ds
dS
dS ds
m m m m
) 1 ( 1 1 / - = - / -
-
(2.46)

1 1
, x X

3 3
, x X

Q
X
,

X X
, ,
=
P


Q
x
,


2 2
, x X

0
B


t
B

3 3

e I

2 2

e I

1 1

e I
O
P
Reference
configuration - 0 t
Current
configuration -
1
t t
Q
P
Q
X d
,

x d
,


P
x x
, ,
=
X X F x
, ,
,
d t d ) , (

ds d Q P
dS d PQ

x
X
,
,

X
,
d x
,
d
M

m
2 CONTINUUM KINEMATICS

165
2.5.3 The Material and Spatial Deformation Gradient
Our goal now is to find the relationship between the line elements X
,
d and x
,
d . By
considering the material description of motion ) , ( t X x x
,
, ,
, and by applying vector
addition, (see Figure 2.13), we obtain:
) , ( ) , ( ; t t d d
P P Q Q P Q
X x X x x X X X
,
,
,
, ,
, , ,
- - (2.47)
If we observe that ) , ( ) , ( ) , ( t d t d t
P P Q Q
X X x X X x X x
, ,
,
, ,
,
,
,
- - , the vector field x
,
d in the
current configuration becomes:
[ ]
i i i
t X X X x t dX X dX X dX X x d
t t d d
e

) , , , ( ) , , , (
) , ( ) , (
3 2 1 3 3 2 2 1 1
- - - -
- -
x
X x X X x x
,
,
,
, ,
, ,
(2.48)
Then by applying Taylor series to represent the function in (2.48) we obtain:
) (

) (

2 2
3
3
2
2
1
1
X x X x
,
,
,
,
d O dX
X
x
d d O dX
X
x
dX
X
x
dX
X
x
d
i j
j
i
i
i i i
-
o
o
-
(

o
o
-
o
o
-
o
o
e e
(2.49)
Since points P and Q are close together, higher order terms can be discarded, which leads
to:
i k ik i k
k
i
dX F dX
X
x
d e e

o
o
x
,

(2.50)
or expressed in compact form:
X F x
,
,
d d
(2.51)
where F is a two-point tensor and is known as the material deformation gradient or simply the
deformation gradient. The relation in (2.51) is a linear transformation so F relates X
,
d
(undeformed configuration) to x
,
d (deformed configuration), (see Figure 2.13). The
equation in (2.51) could have been obtained by directly starting from the gradient
definition, (see Chapter 1-Tensors). That is, if ) , ( t x
,
c c is a scalar field, the total
derivative ( c d ) is given by the equation: x
x
x
x x
,
,
,
, ,
d
t
d t d
o
o

) , (
) , (
c
c c V . Then, if we have
the vector field ) , ( t X x x
,
, ,
, the total derivative (differential) is:
X F X X x X
X
X x
x
X
, , ,
,
,
,
,
,
,
,
d d t d
t
d
o
o
) , (
) , (
V (2.52)
The components of x
,
d in Cartesian system can be evaluated by means of the scalar
product (dot product):
k jk
ij
j i k ik j j
dX F dX F d d
_
, ,
c
e e e

) ( x x
(2.53)
The tensor F can also be expressed as:
j i J i j i
j
i
x
X
x
X X X
e e e e e e e

, 3
3
2
2
1
1

o
o

o
o
-
o
o
-
o
o

x x x
F
, , ,

(2.54)
whose components can be expressed in matrix form as:
NOTES ON CONTINUUM MECHANICS

166
(
(
(
(
(
(
(

o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o

o
o

3
3
2
3
1
3
3
2
2
2
1
2
3
1
2
1
1
1
,
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
x
X
x
F
J i
j
i
ij
F (2.55)
NOTE: Sometimes, for purposes of clarification, we use subscript in capital letters to
differentiate material coordinates, i.e.
j
i
j i
j
i
J i
x X o
- o
= -
o
- o
= -
, ,
.
NOTE: In this publication, we denote the material gradient by
i
i
X
e

) (
) ( ) (
o
- o
- = -
X
,
V Grad and the spatial gradient by
i
i
x
e

) (
) ( ) (
o
- o
- = -
x
,
V grad .
We can also find the inverse transformation of the equation in (2.51) i.e.:
x F X
,
,
d d
-

1
(2.56)
where
1 -
F is the spatial deformation gradient, which is defined as:
) , (
1
t x X F
x
,
,
,
V
-

j I
j
i
ij
X
x
X
F
,
1

o
o

-

(2.57)
Explicitly, the components of
1 -
F are given by:
(
(
(
(
(
(
(

o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o

o
o

- -
3
3
2
3
1
3
3
2
2
2
1
2
3
1
2
1
1
1
1 1
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
F
j
i
ij
F (2.58)
We can now show, (see Chapter 1), that the following relationships are valid:
( )
[ ]
K n J m IJK lmn nk mj ijk lmn
li
il
x x
J
F F
J J
, ,
1
2
1
2
1 ) (

-
F
F
cof
(2.59)
K n J m I l IJK lmn nk mj li IJK lmn
x x x F F F J
, , ,
6
1
6
1
) ( F det (2.60)
The derivative of the equation (2.60) with respect to F becomes:
(
(

o
o
-
o
o
-
o
o

Q p
K n
J m I l K n
Q p
J m
I l K n J m
Q p
I l
IJK lmn
Q p pq
x
x
x x x
x
x
x x x
x
x
dx
dJ
dF
dJ
,
,
, , ,
,
,
, , ,
,
,
,
6
1
(2.61)
which becomes:
2 CONTINUUM KINEMATICS

167
[ ]
[ ]
[ ]
nk mj qjk pmn K n J m QJK pmn
J m I l QIJ plm I l K n QKI pnl K n J m QJK pmn
J m I l IJQ lmp K n I l IQK lpn K n J m QJK pmn
J m I l KQ np K n I l JQ mp K n J m IQ lp IJK lmn
Q p
F F x x
x x x x x x
x x x x x x
x x x x x x
dx
dJ




2
1

2
1
6
1
6
1
6
1
, ,
, , , , , ,
, , , , , ,
, , , , , ,
,

- -
- -
- - c c c c c c
(2.62)
Referring to the definitions for cofactors and inverting tensors, the following relationship
holds:
[ ]
p Q qp pq
nk mj qjk pmn
Q p pq
JX F
F F
dx
dJ
dF
dJ
,
1
,
) (
2
1


-
F F cof

(2.63)
We could have obtained the above equation by means of the definition for third invariant
derivatives in respect to tensors, (see Chapter 1), i.e.:
[ ] ( ) [ ]
T T
J I I I
I I I
- -

o
o

o
o
F F
F
F
F
F
F

det

(2.64)
In comparison with (2.63) it is also true that:
[ ]
1 1 1
,
1
2
1
) (
- - - -

nk mj qjk pmn pq P q
F F x J F cof (2.65)
The derivative of the equation (2.59) with respect to X
,
becomes:
( )
( ) 0
,
,
=
o
o
- -

T
p q qp
Q
p Q
J JF
X
JX
F
X
V ; 0 ) (
,
1
,
(2.66)
or
( )
( ) 0
,
,
=
o
o
- -
-

T
p q qp
q
p q
J F J
x
x J
F
x
1
,
1
,
1
; 0 ) ( V
(2.67)
The following provides proof of the above. From (2.63) we obtain
( ) [ ]
( )
p kq n j m k n jq m qjk pmn
q nk mj nk q mj qjk pmn
q
nk mj qjk pmn
q
p q
x x x x
F F F F F F JX
0 -
-
(

, , , ,
, ,
,
,
2
1
2
1
2
1
,


(2.68)
Note that, in kq the tensor
jqk jkq qjk
- is antisymmetric while the tensor
kq n
x
,
is
symmetric. Therefore
jn kq n qjk
x 0
,
. We can obtain the same result for
km jq m qjk
x 0
,
.
Likewise, it is possible to demonstrate that ( )
p
q
p q
x J 0
,
,
1

-
.
Using the above definitions, it can be shown that, if ) , ( t x
, ,
u and ) , ( t x
,
represent a vector
and a second-order tensor field, respectively, they satisfy the following relationships:
[ ]
[ ] ) , ( ) , (
) , ( ) , (
1
1
t J J t
t J J t
x F X
x F X
x
X
x
X
,
,
, ,
,
,
, ,
, ,

u u


-
-

V V
V V

(2.69)
NOTES ON CONTINUUM MECHANICS

168
To prove this we can use indicial notation:
[ ] ( ) ( ) ( )
( ) ) , (
) , ( ) , ( ) , (
) , (
,
,
1
,
1
,
1
,
1 1
t
X
t
X
x
x
t
x
t
X
x
F F J J
F J F J J F J J t J J
j
j
j
i
i
j
i
j
j
i
i j ij
i
j ij
i j ij j
j
i
ij
i
j ij
indicial
X
X x x
x F
X
x
,
,
,
, ,
_
, ,
,
,
u
u

o
o

o
o
o
o

o
o
o
o

(

(
-


-
- - - -

V
V
u u u
u u
u u u
0
(2.70)
Likewise:
[ ] ( ) ( ) ( )
( )
[ ] ) , ( ) , (
) , ( ) , ( ) , (

) , (
1
, ,
1
,
1
,
1
,
1 1
t t J J
X
t
X
x
x
t
x
t
X
x
F F J J
F J F J J F J J t J J
j
kj
j
i
i
kj
i
kj
k
i
i kj ik i kj ik
i kj ik kj
k
i ik
i
kj ik
indicial
X x F
X x x
x F
X
x
x
,
,
,
, ,
_
,
, ,
,




o
o o

o
o
o
o o

o
o o
o
o
o o
(

(
o - o

o
-
-
- - - -

V V
V
0

(2.71)

Problem 2.5: Let ) , ( t X
,
c be a scalar field in Lagrangian (material) description. Find the
relationship between the material gradient of ) , ( t X
,
c , i.e. ) , ( t X
X
,
,
c V , and the spatial
gradient of ) , ( t x
,
c , i.e. ) , ( t x
x
,
,
c V .
Solution: Remember that a Lagrangian variable ) , ( t X
,
c can be expressed in the Eulerian
(current) configuration by means of the equations of motion, i.e.:
) , ( ) ), , ( ( ) , ( t t t t x x X X
, ,
, ,
c c c .
Then, from the scalar gradient definition we obtain:
F x F
x
x
X
x
x
x X
X
X
X
x
X

o
o

o
o
o
o

o
o
) , (
) , ( ) ), , ( ( ) , (
) , ( t
t t t t
t
,
,
,
,
,
,
,
,
,
,
,
, ,
c
c c c
c V V
In addition we have the inverse form:
1 1
) , (
) , ( ) ), , ( ( ) , (
) , (
- -

o
o

o
o
o
o

o
o
F X F
X
X
x
X
X
X x
x
x
x
X
x
t
t t t t
t
,
,
,
,
,
,
,
,
,
,
,
, ,
c
c c c
c V V
2.5.4 Displacement Gradient Tensors (Material and Spatial)
Let u
,
be a displacement field, (see equation (2.11)). The displacement components in
Lagrangian (material) and Eulerian (spatial) descriptions are, respectively:
i i i
X t x t - ) , ( ) , ( X X
, ,
u
and ) , ( ) , ( t X x t
i i i
x x
, ,
- u
(2.72)
Taking the partial derivative of ) , ( t
i
X
,
u with respect to the material coordinates X
,
, we
obtain:
Indicial notation Tensorial notation
ij ij J i
j
i
j
i
j
i
F t
X
X
X
t x
X
t
c -
o
o
-
o
o

o
o
) , (
) , ( ) , (
,
X
X X
,
, ,
u
u

1 u - = F X X
X
) , ( ) , ( t t
, ,
,
,
J V
(2.73)
where J is known as the material displacement gradient tensor.
2 CONTINUUM KINEMATICS

169
Taking now the partial derivative of equation (2.72) with respect to spatial coordinates x
,
,
we obtain:
Indicial notation Tensorial notation
1
,
) , (
) , ( ) , (
-
-
o
o
-
o
o

o
o
ij ij j i
j
i
j
i
j
i
F t
x
t X
x
x
x
t
c x
x x
,
, ,
u
u

1
) , ( ) , (
-
- = F x x
x
1 u t t
, , ,
,
j V
(2.74)
where j is known as the spatial displacement gradient tensor.
The components of J and j can be represented, respectively, as:
(
(
(
(
(
(
(

-
o
o
o
o
o
o
o
o
-
o
o
o
o
o
o
o
o
-
o
o

(
(
(
(
(
(
(

o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o

o
o

1
1
1
) , (
3
3
2
3
1
3
3
2
2
2
1
2
3
1
2
1
1
1
3
3
2
3
1
3
3
2
2
2
1
2
3
1
2
1
1
1
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X X X
X X X
X X X
X
t
j
i
ij
u u u
u u u
u u u
u X
,
J (2.75)
(
(
(
(
(
(
(

o
o
-
o
o
-
o
o
-
o
o
-
o
o
-
o
o
-
o
o
-
o
o
-
o
o
-

(
(
(
(
(
(
(

o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o

o
o

3
3
2
3
1
3
3
2
2
2
1
2
3
1
2
1
1
1
3
3
2
3
1
3
3
2
2
2
1
2
3
1
2
1
1
1
1
1
1
) , (
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x x x
x x x
x x x
x
t
j
i
ij
u u u
u u u
u u u
u x
,
j (2.76)
By referring to (2.74) we obtain:
( ) 1 1 1 - - -
- -
F F F F x F x
1 1
) , ( ) , ( t t
, ,
j j (2.77)
and by comparing the above equation with that in (2.73) we can draw the conclusion that
) , ( t X
,
J and ) , ( t x
,
j are interrelated by:
F x X F x X
x
X
) , ( ) , ( ; ) , ( ) , ( t t t t
, ,
,
, ,
,
, ,
u u V V j J (2.78)
It is interesting to compare the above with the outcome of Problem 2.5.
Problem 2.6: Consider a continuum in which the displacement field is described by the
following equations:

-
0
2
3
2
2 2
2 1
2
1 1
u
u
u
X
X X X

By definition, a material curve is always formed by the same particles. Let OP and OT be
material lines in the reference configuration, where ) 0 , 0 , 0 (
3 2 1
X X X O ,
) 0 , 1 , 1 (
3 2 1
X X X P and ) 0 , 0 , 1 (
3 2 1
X X X T . Find the material curves in the
current configuration. Also find the deformation gradient.
Solution:
a) The equations of motion can be obtained by means of the displacement field, (see Eq.
(2.72)), i.e.:
i i i
X x - u
NOTES ON CONTINUUM MECHANICS

170

-
- -

-
-
-
3 3
2
2 2 2
2 1
2
1 1 1
3 3 3
2 2 2
1 1 1
2

X x
X X x
X X X X x
X x
X x
X x
3 2 1
, , of values the
ng substituti
u u u
u
u
u

Then, to obtain the material curve, one need only substitute the material coordinates with
the particles belonging to the line OP in the equations of motion, (see Figure 2.14). Notice
that the material curve OP in the current configuration is no longer a straight line, but the
line OT is still a straight line in the current configuration (see Figure 2.15).
The components of the deformation gradient, (see Eq. (2.55)), can be obtained as follows:
(
(
(

-
- -

(
(
(
(
(
(
(

o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o

1 0 0
0 2 1 0
0 ) 4 1 (
2
1 2 1
3
3
2
3
1
3
3
2
2
2
1
2
3
1
2
1
1
1
X
X X X
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
F
jk






















Figure 2.14: Deformation of the material curve OP.













0
0.5
1
1.5
2
2.5
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
x1
x
2
Current Conf.
Reference Conf.

P
P
Q Q
material curve
O
2 CONTINUUM KINEMATICS

171





















Figure 2.15: Deformation of the material curve OT .

2.5.5 Material Time Derivative of the Deformation Gradient.
Material Time Derivative of the Jacobian Determinant
2.5.5.1 Material Time Derivative of F . The Spatial Velocity Gradient
The material time derivative of F is given by
J I J i
j
i
x
i
j j
i
ij ij
v x
X
t v
t
t x
X X
t x
t
F F
Dt
D
i
, ,
) , ( ) , ( ) , (

o
o

o
o
o
o

o
o
o
o
=
`
,
_
, ,
`
`
X X X

(2.79)
Expressing velocity in Eulerian coordinates, i.e. ) ), , ( ( t t v
i
x X
,
,
, and by using the chain rule of
the derivatives we obtain:
kj ik
j
k
ik
j
k
k i
J k k i J I
j
k
k
i
j
i
ij
F
X
x
X
x
v
x v v
X
t x
x
t v
X
t t v
F
l l
o
o

o
o


o
o
o
o

o
o

,
, , ,
) , ( ) , ( ) ), , ( ( X x x X
,
, ,
,
`
(2.80)
The above equation is represented in tensor notation as:
F F l
`
(2.81)

Current Conf .
0
0.02
0.04
0.06
0.08
0.1
0 0.5 1 1.5 2 2.5 3 3.5
x1
x
2
Current Conf.
Ref erence Conf .
0
0.02
0.04
0.06
0.08
0.1
0 0.5 1 1.5 2 2.5 3 3.5
x
2
Reference Conf.

T
T
O
O
NOTES ON CONTINUUM MECHANICS

172
where l is the spatial velocity gradient, and is defined as:
1
) , ( ) , (
-
F F x v x
x
`
, , ,
,
t t V l Spatial velocity gradient
(

s m
m

(2.82)
Problem 2.7: Let x
,
d be a differential line element in the current configuration. Find the
material time derivative of x
,
d .
Solution:
x v x
X F X F X F X F x
x
x
, , ,
_
,
_
, , ,
,
,
,
,
d d
d d
Dt
D
d
Dt
D
d
Dt
D
d
Dt
D
d


=
-
V l
l
0
) ( ) ( ) (

And, whose components are represented by:
k
k
i
k k i
i
dx
x
t v
dx v d
Dt
D
o
o
|
.
|

\
|
) , (
,
x
x
,
,

2.5.5.2 Rate-of-Deformation and Spin Tensors
The spatial velocity gradient l can be decomposed into a symmetric and an antisymmetric
part, i.e.:
W D- - - - - ) (
2
1
) (
2
1
T T skew sym
l l l l l l l
(

s m
m

(2.83)
Whereby, we can define the following tensors:
) , ( t
sym
x
,
D = l - the rate-of-deformation tensor;
) , ( t
skew
x
,
W = l - the spin, rate-of-rotation tensor or vorticity tensor.
The components of D and W, respectively, are:
|
|
.
|

\
|
o
o
-
o
o

|
|
.
|

\
|
o
o
-
o
o

i
j
j
i
ij
i
j
j
i
ij
x
v
x
v
x
v
x
v
2
1
;
2
1
W D (2.84)
The spin tensor has three independent components and can be represented as:
(
(
(

-
-
-

(
(
(

- -
-
(
(
(

0
0
0
0
0
0
0
0
0
1 2
1 3
2 3
23 13
23 12
13 12
32 31
23 21
13 12
w w
w w
w w
ij
W W
W W
W W
W W
W W
W W
W
(2.85)
where
i
w are the axial vector components associated with the antisymmetric tensor W.
We can also define the vorticity vector field as w
, ,
2 . Moreover, as we saw in the chapter on
tensors: given an antisymmetric tensor, the following holds:
ij kij k kij k ij
w or w W W
2
1
- - (2.86)
In Chapter 1 we proved that v v w
x
,
,
, ,
,
r V ) ( 2 rot where w
,
is the axial vector associated
with the antisymmetric tensor
skew
) ( v
x
,
,
V . Therefore, the vorticity vector can be expressed
as:
2 CONTINUUM KINEMATICS

173
v v w
x
,
,
, , ,
,
r V ) ( 2 rot Vorticity vector (2.87)
Also in Chapter 1 we showed the following relationship was satisfied:
( ) v v v w v
x
, ,
,
, , ,
,
r r r V
2
1
W (2.88)
When 0 D motion is characterized by rigid body motion, i.e. the distance between
particles does not change. Furthermore, the condition x w
x , ,
,
d
Dt
d D
r
) (
is satisfied which is
proved by:
_
, , , ,
_
,
,
property tensor
ric Antisymmet
) 2.7 Problem See (
) (
) (
x w x x x
x
d d d d
Dt
d D
r - W W D l
(2.89)
To prove that 0 D characterized by rigid body motion our starting point is the definition
of rigid body motion in which the distance between particles does not change, hence the
magnitude of x
,
d does not change over time. Taking the material time derivative of
2
x
,
d
we obtain:
x x
x x x x x x x x
x x x x
x x
x x x x x x x
, ,
, , , , , , , ,
, , , ,
, ,
, , , , , , ,
d d
d d d d d d d d
d d d d
d
Dt
D
d
d
Dt
D
d d d
Dt
D
d d
Dt
D
d
Dt
D


- -
-

-
D
W D W D
W D
2
) ( 2 2 2 2
) ( 2 2
2.7) Problem (see ) ( 2
) ( ) ( ) (
2
:
l
(2.90)
where we have used the property 0 ) ( 0 x x
, ,
d d
sym skew
: : W B A . So, according to
(2.90), the magnitude of x
,
d does not change over time if 0 D .
If the spin tensor is a zero tensor, 0 W , the velocity field is said to be irrotational, thus
0
,
,
,
,
r v
x
V . In Problem 2.3 the following relationship was validated
( ) v v v v
x x x
, ,
,
, ,
, , ,
r r -
|
|
.
|

\
|
V V V
2
1
2
2
v
which can contrast with:
( )
[ ] ( ) ( ) [ ]
[ ] ( ) [ ] ( ) v v v v v v
v v v v
v v v v v v v v v
x x
x
, ,
,
, , , ,
, , , ,
, , , , , , , , ,
, ,
,
r r - -
- - - - - -
- - - - -



V V
V
T T
T T
2
2
1
2 2
2
1
2
2
1
2
) (
W
W W W D
W W W D W D W D
l
l l l l
(2.91)
The term ( ) v v
x
, ,
,

T
V 2 can be written in indicial notation as
j i j
v v
,
2 , which is equivalent to
i j j i j j j i j i j j i i i
v v v v v v v v v
, , ,
2
2
2 ), ( ), ( ), ( ), ( - v v v
, , ,
, Thus:
( ) v v v v
x x x
, ,
,
, ,
, , ,
r r -
|
|
.
|

\
|
V V V
2
2
v
(2.92)
NOTES ON CONTINUUM MECHANICS

174
2.5.5.3 The Material Time Derivative of
1 -
F
The material time derivative of the spatial deformation gradient (
1 -
F ) is obtained from the
material time derivative of 1
-
F F
1
, i.e.:
0
0
1
-
-

- -
-
-
-

1 1
1
1
1
) (
F F F F
F
F F
F
F F
` `
Dt
D
Dt
D
Dt
D
Dt
D

(2.93)
Therefore:
_
` `
_
` `
l
1 1 1 1 1 1 - - - - - -
- - F F F F F F F F F F
1

(2.94)
which leads to:
Tensorial notation Indicial notation
l
- -
-
1 1
F F
`
kj ik ij
F F l
1 1 - -
-
`

(2.95)
NOTE: In this publication we adopt the following notation to represent the material time
derivative of the inverse of a tensor: ( )
1 1
1
- -
-
= = F F
F
`
.
Dt
D
.
2.5.5.4 The Material Time Derivative of the Jacobian Determinant
The material time derivative of the Jacobian determinant can be evaluated by the definition
of the second-order tensor determinant:
( )
R Q P R Q P R Q P PQR
R Q P PQR
R Q P
PQR
j
i
x x x x x x x x x J
Dt
J D
x x x
X
x
X
x
X
x
X
x
J
, 3 , 2 , 1 , 3 , 2 , 1 , 3 , 2 , 1
, 3 , 2 , 1
3 2 1
) (
` ` `
`
- - =

o
o
o
o
o
o

o
o

F
(2.96)
According to equations (2.79) and (2.80) the following relationships are valid:
R s s
R
s
s
R Q s s
Q
s
s
Q
P s s
P
s
s p p
P
x v
X
x
x
v
x x v
X
x
x
v
x
x v
X
t x
x
t v
X
t v
X
t x
x
, , 3
3
, 3 , , 2
2
, 2
, , 1
1 1 1
, 1
;
) , ( ) , ( ) , ( ) , (

o
o
o
o

o
o
o
o

o
o
o
o

o
o

o
o
=
` `
,
,
, ,
`
`
X x X X
(2.97)
By substituting
P
x
, 1
` ,
Q
x
, 2
` ,
R
x
, 3
` , given by (2.97), into equation (2.96), we obtain:
( )
R s s Q P PQR R Q s s P PQR R Q P s s PQR
R Q P R Q P R Q P PQR
x v x x x x v x x x x v
x x x x x x x x x J
, , 3 , 2 , 1 , 3 , , 2 , 1 , 3 , 2 , , 1
, 3 , 2 , 1 , 3 , 2 , 1 , 3 , 2 , 1

- -
- - ` ` `
`
(2.98)
The first term on the right hand side of the equation in (2.98) can be expressed as:
J v
x x x v x x x v x x x v x x x v
R Q P PQR R Q P PQR
J
R Q P PQR R Q P s s PQR
1 , 1
0
, 3 , 2 , 3 3 , 1
0
, 3 , 2 , 2 2 , 1 , 3 , 2 , 1 1 , 1 , 3 , 2 , , 1

- -

_ _ _

(2.99)
2 CONTINUUM KINEMATICS

175
in which the following was validated: 0
, 3 , 2 , 3 , 3 , 2 , 2

R Q P PQR R Q P PQR
x x x x x x , since these
relationships represent a matrix determinant that has two equal rows (linearly dependent).
Similarly, we obtain: J v x x v x
R Q s s P PQR 2 , 2 , 3 , , 2 , 1
and J v x v x x
R s s Q P PQR 3 , 3 , , 3 , 2 , 1
after which
the equation in (2.98) can be rewritten as:
J v J v J v J v J
k k, 3 , 3 2 , 2 1 , 1
- -
`
(2.100)
which is the same as:
) (
) (
) (

D Tr
Tr
Tr
J
J
J
J
J
Dt
D

l
v
v
v F F
x
x
x
,
,
,
`
,
,
,
V
V
V
Material time derivative of the
Jacobian determinant
(2.101)
where we have used the equation in which the trace of an antisymmetric tensor is zero,
) ( ) ( ) ( ) ( ) ( D W D W D Tr Tr Tr Tr Tr - - l .
The material time derivative of the Jacobian determinant could also have been obtained as,
(see Chapter 1):
kk ik ik kj ik ji
T
I I I I I I F F I I I I I I
Dt
D
D
I I I D
Dt
I I I D
l l l
F F F F
F F
F F
F
F

- -
c
1
) ( ) ( l : : (2.102)
Problem 2.8: Starting from the definition ( ) [ ] ( )
ij
ij
F
Dt
F D
Dt
D
cof det F , show that the
equation in (2.101), v
x
,
`
,
V J J , is valid.
Solution: Considering that
j
i
ij
X
x
F
o
o
, the material time derivative of ) (F F det = is given
by:
[ ] ( ) ( ) ( ) ( )
ij i
j
ij
i
j
ij
j
i
F v
X
D
F
Dt
t x
X
D
F
X
t x
Dt
D
Dt
D
cof cof cof det
o
|
.
|

\
| o
o

|
|
.
|

\
|
o
o

) , ( ) , (
) (
X X
F
and considering that ) ), , ( ( t t v
i
X x
,
, we can state that:
( ) [ ] ( )
ij
j
k
k
i
F
X
x
x
v
Dt
D
cof det
o
o
o
o
F
By referring to the definition of the cofactor: ( ) [ ] ( ) ( )
ij ij
T
ij
F F F det cof
1 -
, we can also state
the following is valid:
( ) [ ] ( ) ( ) ( ) ( ) ( )
( )
i i ij
i
i
ij ki
k
i
ij ji kj
k
i
ij
T
ij
j
k
k
i
Jv F
x
v
F
x
v
F F F
x
v
F F
X
x
x
v
Dt
D
,
1

o
o

o
o

o
o

o
o
o
o

- -
det
det det det det c F

NOTES ON CONTINUUM MECHANICS

176

2.6 Finite Strain Tensors
Before outlining the different ways we can define the strain tensors, it must be stressed that
displacement is a measurable quantity, whereas strain is based on concepts that have been
introduced for convenience. The strain definition used in this section is the dimensionless
quantity
2
2 2
) (
) ( ) (
dS
dS ds -
(material description) or
2
2 2
) (
) ( ) (
ds
dS ds -
(spatial configuration).
Let us consider, once again, two particles P and Q , connected by the vector X
,
d in the
reference configuration. After motion, the particles occupying the points P and Q are
moved to the points P and Q , respectively, and the new vector joining these material
points is defined by x
,
d , (see Figure 2.16). The magnitudes of these vectors squared are:
X X X
, , ,
d d dS d
2
2
) ( (2.103)
and
x x x
, , ,
d d ds d
2
2
) ( (2.104)















Figure 2.16: Deformation of the continuum.



1 1
, x X

2 2
, x X

3 3
, x X

0
B

B

3 3

e I

2 2

e I

1 1

e I
O
P
X
,

Reference
configuration - ) ( 0
0
t t
Current
configuration -
1
t t
Q
P
Q
X d
,

x d
,


ds d Q P
dS d PQ

x
X
,
,

x
,

X X F x
, ,
,
d t d ) , (
X d
,

x d
,

2 CONTINUUM KINEMATICS

177
2.6.1 The Material Finite Strain Tensor
The relationship
2 2
) ( ) ( dS ds - can be expressed in the material description as:
( )
( )
( ) X E X
X C X
X F F X
X X X F F X
X X X F X F
X X x x
, ,
, ,
, ,
, , , ,
, , ,
, ,
, ,
d d
d d
d d
d d d d
d d d d
d d d d dS ds
T
T





-
-
-
-
- -
2
) ( ) (
2 2
1
1
j k kj
j k kj kj
j k kj ij ik
j k kj j ij k ik
k k i i
dX dX E
dX dX C
dX dX F F
dX dX dX F dX F
X d dX x d x d dS ds
) 2 (
) (
) (
) ( ) (
2 2

-
-
-
- -
c
c
c

(2.105)
where C is known as the right Cauchy-Green deformation tensor, also known as the Green
deformation tensor, and is defined as:
F F X C
T
t) , (
,
The right Cauchy-Green deformation tensor (2.106)
C is a symmetric tensor, i.e. ( ) C F F F F C
T
T
T T
, and is also a positive definite
tensor, since 0 F (see Problem 1.25 in Chapter 1). The inverse of C , which is also in
the reference configuration, is given by:
T
t
- - -
F F X C
1 1
) , (
,
(2.107)
We can now introduce the left Cauchy-Green deformation tensor ) (b , also known as the Finger
deformation tensor, which we can find in the spatial configuration, and is defined as:
T
t F F x b ) , (
,
The left Cauchy-Green deformation tensor (2.108)
b is a positive definite symmetric tensor.
NOTE: The word right is always associated with material configuration, meanwhile left is
related to spatial configuration.
The inverse of b , which is also in the current configuration, is given by:
1 1
) , (
- - -
F F x b
T
t
,
(2.109)
We can also define the Piola deformation tensor ( B ) as:
C F F B C F F X B
- - - - T inverse the T
t
1 1 1
) , (
,
(2.110)
In the subsection Polar Decomposition more details will be provided about the configurations
in which these tensors appear.
We can now present some relationships and properties of C and b :
The tensors C and b are related by:
2
1 1 1 1
1 1 1 1
;
;
b b b F F F F F C F
F C F b F C F b
F b F C F b F C






- - - -
- - - -
T T T
(2.111)
The determinant of b is:
) ( ) ( ) ( ) ( ) ( ) ( ) (
2
C F F F F F F b det det det det det det det J
T T

(2.112)
Then the Jacobian determinant can also be expressed as
C
C I I I J ) ( det .
NOTES ON CONTINUUM MECHANICS

178
The invariants of C and b :
[ ]
( )
(

- -
-

3 3 3 2
2 2
2
1
2
3
3
1
6
1
) (
) (
2
1
) (
C C C C C
C
C
C b
C C b
C b
Tr Tr Tr Tr det
Tr
Tr
J C C C I I I I I I
I I I I I
C I I
kr jq ip pqr ijk
ii


(2.113)
The relation
C b
I I is proven by applying the trace property, (see Chapter 1),
) ( ) ( ) ( ) ( b F F F F C Tr Tr Tr Tr
T T
. Furthermore, the relationship
) ( ) (
n n
b C Tr Tr is also valid.
C b
I I I I I I can be proved by using the determinant property, i.e.:
[ ]
b C
F F F F F F F C I I I I I I
T T T

2
) ( ) ( ) ( ) ( ) ( ) ( det det det det det det
The tensors C and b are positive definite symmetric tensors which was proven in
Problem 1.25 in Chapter 1.

Returning to equation (2.105), we now introduce the Green-Lagrange strain tensor denoted by
E , also called the Lagrangian finite strain tensor or the Green-St_Venant strain tensor, and
defined as:
Tensorial notation Indicial notation
The Green-Lagrange
strain tensor
) (
2
1
) (
2
1
) , (
1
1
-
-
C
F F X E
T
t
,

) (
2
1
) (
2
1
ij ij
ij kj ki ij
C
F F E
c
c
-
-

(2.114)
The Green-Lagrange strain tensor ( E ) is a symmetric tensor, i.e.:
E F F F F E - - ) (
2
1
) (
2
1
1 1
T T T T
(2.115)
The Green-Lagrange strain tensor ( E ) can also be expressed in function of the material
displacement gradient tensor, ) , ( t X
X
,
,
,
u V = J . To do this, we start from the equation in (2.73),
i.e.:
1 - J F ij J i ij
j
i
ij ij ij
X
F c c c - -
o
o
-
,
u
u
J

(2.116)
Afterwards, the right Cauchy-Green deformation tensor ( C ) can be expressed in terms of
J as:
1
1 1
1 1
- - -
- -
- -

J J J J
J J
J J
T T
T
T
T
) ( ) (
) ( ) (
F F C

ij J i I j J k I k
kj ki J k ki kj I k J k I k
kj J k ki I k
kj ki ij
F F C
c
c c c c
c c
- - -
- - -
- -

, , , ,
, , , ,
, ,
) )( (
u u u u
u u u u
u u

(2.117)
or
J J J J - - -
T T
_
E
C
2
1
J i I j J k I k
E
ij ij
ij
C
, , , ,
2
u u u u - - -
_
c

(2.118)
2 CONTINUUM KINEMATICS

179
then:
( ) J J J J - -
T T
2
1
E
|
|
.
|

\
|
o
o
o
o
-
o
o
-
o
o

j
k
i
k
i
j
j
i
ij
X X X X
E
u u
u
u
2
1

(2.119)
Explicitly, the components of E are given by:
(
(

|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o

2
1
3
2
1
2
2
1
1
1
1
11
2
1
X X X X
E
u u u u
,
(
(

|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o

2
2
3
2
2
2
2
2
1
2
2
22
2
1
X X X X
E
u u u u

(
(

|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o

2
3
3
2
3
2
2
3
1
3
3
33
2
1
X X X X
E
u u u u

21
2
3
1
3
2
2
1
2
2
1
1
1
1
2
2
1
12
2
1
E
X X X X X X X X
E
(

o
o
o
o
-
o
o
o
o
-
o
o
o
o
-
o
o
-
o
o

u u u u u u u u

31
3
3
1
3
3
2
1
2
3
1
1
1
1
3
3
1
13
2
1
E
X X X X X X X X
E
(

o
o
o
o
-
o
o
o
o
-
o
o
o
o
-
o
o
-
o
o

u u u u u u u u

32
3
3
2
3
3
2
2
2
3
1
2
1
2
3
3
2
23
2
1
E
X X X X X X X X
E
(

o
o
o
o
-
o
o
o
o
-
o
o
o
o
-
o
o
-
o
o

u u u u u u u u


Problem 2.9: Let us consider the equations of motion:

-
-
-
2
3 3 3
2
2 2 2
2 1 1 1
4
X X x
X X x
X X X x

Find the Green-Lagrange strain tensor ( E ).
Solution:
Referring to the E equation given in (2.114):
) (
2
1
; ) (
2
1
ij kj ki ij
T
F F E c - - 1 F F E (2.120)
where the components of F are derived as:
(
(
(

-
-
-

(
(
(
(
(
(
(

o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o

o
o

3
2
1 2
3
3
2
3
1
3
3
2
2
2
1
2
3
1
2
1
1
1
2 1 0 0
0 2 1 0
0 4 ) 4 1 (
X
X
X X
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
F
j
k
kj

And,
NOTES ON CONTINUUM MECHANICS

180
(
(
(

-
- - -
- -

(
(
(

-
-
-
(
(
(

-
-
-

2
3
2
2
2
1 1 2
1 2
2
2
3
2
1 2
3
2 1
2
) 2 1 ( 0 0
0 ) 2 1 ( ) 4 ( 4 ) 4 1 (
0 4 ) 4 1 ( ) 4 1 (
2 1 0 0
0 2 1 0
0 4 ) 4 1 (

2 1 0 0
0 2 1 4
0 0 ) 4 1 (
X
X X X X
X X X
X
X
X X
X
X X
X
F F
kj ki

Then substituting the above into the equation in (2.120) we obtain:
(
(
(

- -
- - - -
- - -

1 ) 2 1 ( 0 0
0 1 ) 2 1 ( ) 4 ( 4 ) 4 1 (
0 4 ) 4 1 ( 1 ) 4 1 (
2
1
2
3
2
2
2
1 1 2
1 2
2
2
X
X X X X
X X X
E
ij

Problem 2.10: Obtain the principal invariants of E in terms of the principal invariants of
C and b .
Solution:
The principal invariants of E are given by:
[ ] ) ( ; ) (
2
1
; ) (
2 2
E E E
E E E E
det Tr Tr - I I I I I I I
Referring to the fact ) (
2
1
1 - C E , the principal invariants can also be expressed as:
The First Invariant:
[ ] ( ) 3
2
1
) ( ) (
2
1
) (
2
1
) (
2
1
) ( - - -
(

-
C E
C C C E I I 1 1 1 Tr Tr Tr Tr Tr
The Second Invariant:
[ ] ) (
2
1
2 2
E
E E
Tr - I I I
where
( ) ( ) 9 6
4
1
3
2
1
2
2
2
- -
(

-
C C C E
I I I I
[ ] ( ) ( ) ( ) ( ) [ ]
( ) [ ] 3 2
4
1
2
4
1
2
4
1
) (
4
1
) (
2
1
) (
2
2 2 2
2
2
- -
- - - - -
(

-
C
C
C C C C C C E
I Tr
Tr Tr Tr Tr Tr Tr Tr 1 1 1 1

The term ( )
2
C Tr can be obtained as follows:
( )
2
3
2
2
2
1
2
2
3
2
2
2
1
2 2
0 0
0 0
0 0
C C C
C
C
C
C
ij
- -
(
(
(

C C C C Tr
It is also true that:
( ) ( )
C C
C
C
I I I C C C
C C C C C C C C C C C C I
I I
2
2
2 2
3
2
2
2
1
3 2 3 1 2 1
2
3
2
2
2
1
2
3 2 1
2
- - -
- - - - - - -
_

Therefore we have:
( ) 3 2 2
4
1
) (
2 2
- - -
C C C
E I I I I Tr
Whereupon, the second invariant can also be expressed as:
2 CONTINUUM KINEMATICS

181
( ) ( ) ( ) 3 2
4
1
3 2 2
4
1
9 6
4
1
2
1
2 2
- - -
(

- - - - - -
C C C C C C C E
I I I I I I I I I I I
The Third Invariant:
[ ] ) (
2
1
) (
2
1
) (
3
1 1 - |
.
|

\
|

- C C E
E
det det det I I I
The term [ ] ) ( 1 - C det can also be expressed as:
( )( )( )
1 1
1 1 1
1 0 0
0 1 0
0 0 1
) (
3 2 1 3 2 3 1 2 1 3 2 1
3 2 1
3
2
1
- - - - - - - - - -
- - -
-
-
-
-
C C C
C
I I I I I I C C C C C C C C C C C C
C C C
C
C
C
1 det

Then:
( ) 1
8
1
- - -
C C C E
I I I I I I I I I
In short we have:

( )
( )
( ) 1
8
1
3 2
4
1
3
2
1
- - -
- - -
-
C C C E
C C E
C E
I I I I I I I I I
I I I I I
I I
;
1 2 4 8
3 4 4
3 2
- - -
- -
-
E E E C
E E C
E C
I I I I I I I I I
I I I I I
I I

2.6.2 The Spatial Finite Strain Tensor (The Almansi Strain
Tensor)
In the previous subsection,
2 2
) ( ) ( dS ds - was expressed in the material description.
Alternatively, it can be expressed in spatial description, i.e.:
( )
( )
( ) x e x
x c x
x F F x
x F F x x x
x F x F x x
X X x x
, ,
, ,
, ,
, , , ,
, , , ,
, ,
, ,
d d
d d
d d
d d d d
d d d d
d d d d dS ds
T
T





-
-
-
-
- -
- -
- -
- -
2
) ( ) (
1
1
1 1
2 2
1
1
j i ij
j i ij ij
j i j k i k ij
j j k i i k j i ij
k k i i
dx dx e
dx dx c
dx dx X X
dx X dx X dx dx
X d dX x d x d dS ds
2
) (
) (
) ( ) (
, ,
, ,
2 2

-
-
-
- -
c
c
c

(2.121)
where we have introduced c known as the Cauchy deformation tensor, and defined as:
1
) , (
- -
F F x c
T
t
,
The Cauchy deformation tensor (2.122)
Additionally, it holds that b c
-1
, where ) , ( t x b
,
is the left Cauchy-Green deformation tensor,
defined in (2.108).
We can also define the Almansi strain tensor or Eulerian finite strain tensor, e , as:
) (
2
1
) (
2
1
) (
2
1
) , (
1 1 - - -
- - - b F F c x e 1 1 1
T
t
,
The Almansi strain tensor (2.123)
The components of e are given by:
NOTES ON CONTINUUM MECHANICS

182
) (
2
1
) (
2
1
1 1 - -
- -
kj ki ij ij ij ij
F F c e c c (2.124)
It is also true that:
1 1
1 1


- -
- -


T T
T T
F F F F B C C B
F F F F b c c b
1
1
(2.125)
The Almansi strain tensor ( e ) can also be expressed in terms of the spatial displacement
gradient tensor,
1
) , (
-
- = F x
x
1 u j t
, ,
,
V :
( ) [ ] [ ]
( ) [ ] ( ) j j j j j j j j
j j j j


- - - - - -
- - - - - - -
- -
T T T T
T T T
2
1
2
1
) ( ) (
2
1
) ( ) (
2
1
2
1
1
1 1
1 1 1 1 1 1 1 F F e

(2.126)
After which we have:
Tensorial notation Indicial notation
( ) j j j j - -
T T
2
1
e
|
|
.
|

\
|
o
o
o
o
-
o
o
-
o
o

j
k
i
k
i
j
j
i
ij
x x x x
e
u u
u
u
2
1

(2.127)
Both tensors E and e are symmetric second-order tensors and the relationship between
them can be obtained starting from the definition in (2.123), ) ( 2
1 - -
- F F e
T
1 :
E
F F
F F F F F F
F F F F F e F
2
) ( 2
1
1

-
-
-



- -
- -
1
1
1
T
T T T
T T T

(2.128)
Thus, we can draw the conclusion that:
F e F E
T
(2.129)
and:
1 - -
F E F e
T

(2.130)


Figure 2.17 summarizes some equations by the use of deformation and strain tensors.





OBS.: In rigid body motion the relation
2 2
) ( ) ( dS ds - is zero, so the strain
tensors ( E , e ) must be zero tensors at any time during motion.
2 CONTINUUM KINEMATICS

183












Figure 2.17: Deformation and strain tensors (Kinematic tensors).
Problem 2.11: Show that the Green-Lagrange strain tensor ( E ) and the right Cauchy-
Green deformation tensor ( C ) are coaxial tensors.
Solution:
Two tensors are coaxial if they have the same principal directions. Coaxiality can also be
demonstrated if the relation C E E C holds.
Starting with the definition E C 2 - 1 , we can conclude that:
( ) ( ) C E E E E E E E E E C - - - 2 2 2 1 1 1
Thus, we can prove that E and C are coaxial tensors.
2.6.3 The Material Time Derivative of Strain Tensors
2.6.3.1 The Material Time Derivative of the Right Cauchy-Green
Deformation Tensor
The material time derivative of the right Cauchy-Green deformation tensor, C
`
, is obtained
as follows:
( )
F F
F F F F F F
F F F F F F C C


- -
- =

D
T
sym
T T T T T
T T T
Dt
D
Dt
D
2
) ( ) (
2
_
` ` `
l
l l l l
(2.131)
2.6.3.2 The Material Time Derivative of the Green-Lagrange Strain
Tensor
The material time derivative of the Green-Lagrange strain tensor, E
`
, is obtained by means
of the equation in (2.114), the result of which is:
C F F F F C F F E E
` ` ` `
2
1
) (
2
1
) (
2
1
) (
2
1
) ( -
(

-
(

- =
T T T
Dt
D
Dt
D
Dt
D
1 1 (2.132)
By comparing the equation in (2.132) with (2.131), we can conclude that:
X
,

x
,

F
Reference
configuration
Current
configuration

0
B
B
Current Conf. Reference Conf.

( ) 1 -

- - -

C X E
C F F X B
F F X C
2
1
) , (
) , (
) , (
1 1
t
t
t
T
T
,
,
,

( ) c x e
b F F x c
c F F x b
-


- - -
-

1
2
1
) , (
) , (
) , (
1 1
1
t
t
t
T
T
,
,
,


1
1 1 1
1
- -
- - -
-


F E F e
F C F b
F C F b
T


F e F E
F b F C
F b F C


- - -
-
T
1 1 1
1

NOTES ON CONTINUUM MECHANICS

184
F F C E D
T
` `
2
1

(2.133)
and after some algebraic work we can obtain the inverse relationship:
1 1
2
1
- - - -
F C F F E F
` `
T T
D
(2.134)
The equation in (2.133) could have been obtained by means of the equation in (2.90), i.e.:
[ ] [ ] [ ]
[ ]

[ ] X E X x x
X E X X E X X E X x x
X E X
,
`
,
, ,
`
, , ,
`
, ,
`
,
, ,
, ,
, ,
d d d d
Dt
D
d d d d d d d d
Dt
D
d d
Dt
D
ds
Dt
D
dS ds
Dt
D




- -
-

2
2 2 2
2 ) ( ) ( ) (
2 2 2
0 0

(2.135)
Then we have:
[ ] X F F X x x x x X E X
, ,
, , , ,
,
`
,
d d d d d d
Dt
D
d d
T
D D 2 2 2 (2.136)
Therefore, we can conclude that F F E D
T
`
.
2.6.3.3 The Material Time Derivative of
1 -
C
The material time derivative of the inverse of the right Cauchy-Green deformation tensor
can be obtained by considering that if 1
-
C C
1
, it follows that:
1 1 1
1 1 1 1 1
) ( ) (
- - -
- - - - -


-
- -
C C C C
C C C C C C C C C C
` `
` ` ` `
0 1
Dt
D
Dt
D

(2.137)
Also by referring to
T - - -
F F C
1 1
and F F C D
T
2
`
, (see Eq. (2.131)), we obtain:
T - - -
- F F C D
1 1
2
`

(2.138)
Note that the tensors C
`
and
1 -
C
`
are still symmetric tensors.
2.6.3.4 Material Time Derivative of the Left Cauchy-Green
Deformation Tensor
The material time derivative of the left Cauchy-Green deformation tensor ( b ), (see
equation (2.108)), is given by:
T T T T T T T
Dt
D
Dt
D
l l l l - - - = b b F F F F F F F F F F b b
` ` `
) ( ) ( (2.139)
The material time derivative of ) (
1 -
b is obtained as follows:
l l
- - - - - - - - - -
- - -
1 1 1 1 1 1 1
) ( ) ( b b b F F F F F F b
T T T T
Dt
D
Dt
D
` ` `
(2.140)
So, we can see the tensors b
`
and
1 -
b
`
are still symmetric tensors.
The material time derivative of the Piola deformation tensor B is given by:
2 CONTINUUM KINEMATICS

185
T T
Dt
D
Dt
D
- - - -
- = F F F F B B D 2 ) ( ) (
1 1
`
(2.141)
2.6.3.5 The Material Time Derivative of the Almansi Strain Tensor
The material time derivative of the Almansi strain tensor, e` , can be obtained by means of
equation (2.123), the outcome of which is:
) (
2
1
) (
2
1
) (
2
1
) (
2
1
) (
1 1
1
c F F F F
c F F e e
`
` `
`
- -
(

-
(

- =
- - - -
- -


T T
T
Dt
D
Dt
D
Dt
D
1 1
(2.142)
We can also obtain the relationship between ( e` ) and ( D ). In order to do so we consider
the material time derivative of the equation in (2.129), F e F E
T
:
F e F F e F F e F E
`
`
` `
- -
T T T
(2.143)
Referring to the fact that F F E D
T
`
and F F l
`
, we obtain:
( ) ( )
1
1
1 1 1
- -
- -
- - - - - -




- -
- -
- -
- -
F F e e e F F
F F e e e F F
F F e F F F F e F F F F e F F
F e F F e F F e F F F
l l `
`
`
`
`
`
`
`
`
`
T
T
T T
T T T T T T
T T T T
D
D
D
D

(2.144)
Thus,
e e e - -
T
l l ` D
(2.145)
Problem 2.12: Obtain the material time derivative of the Jacobian determinant ( J
`
) in
terms of ( E
`
), ( C
`
), ( F
`
).
Solution:
This was obtained in (2.101) when ) (D Tr J J
`
, where D is the rate-of-deformation tensor
which is related to E
`
by means of the relationship
1 - -
F E F
`
T
D , (see equation
(2.134)), then:
( ) ( ) 1 D :
1 1
) (
- - - -
F E F F E F
` ` `
T T
J J J J Tr Tr
In indicial notation we have:
C C E C E F F
` ` ` ` ` `
: : :
1 1 1 1 1 1 1
2
) (
- - - - - - - -

J
J J E F F J F E F J J
T
kp pi ki ij pj kp ki
c
The J
`
can still be expressed in terms of F
`
. To this end let us consider the following
equation ( )
sp sk sp sk kp
F F F F E
` ` `
-
2
1
, (see Eq. (2.132)). Then, J
`
can also be expressed by:
( ) ( )
( ) ( )
T T
ts st st ts sp ps sk ks sp pi si sk ki si
sp sk pi ki sp sk pi ki sp sk sp sk pi ki kp pi ki
J J
F F J F JF F F F F
J
F F F F
J
F F F F F F F F
J
F F F F F F J E F F J J
- -
- - - - - -
- - - - - - - -

- -
- -
F F F F : :
` `
` ` ` ` ` `
` ` ` ` ` `
1 1 1 1 1 1
1 1 1 1 1 1 1 1
2 2
2 2
1

c c
In short, there are various different ways to express the material time derivative of the
Jacobian determinant:
NOTES ON CONTINUUM MECHANICS

186
) ( ) (
2
) (
2
) (
1 1 1
1 1
- - -
- - -


F F C C E C
F F C C E C
` ` `
` ` ` `
Tr Tr Tr
Tr
J
J
J
J
J
J J J
T
: : : D

where we have used the trace property: ) ( ) ( B A B A B A
T T
Tr Tr : in which A and B
are arbitrary second-order tensors.
2.6.4 Interpreting Deformation/Strain Tensors
Now we can consider two vectors in the reference configuration defined by
M X

) 1 ( ) 1 (
dS d
,
and N X

) 2 ( ) 2 (
dS d
,
, where O is the angle formed between them, (see
Figure 2.18). After motion, these vectors are transformed into m x
) 1 ( ) 1 (
ds d
,
and
n x
) 2 ( ) 2 (
ds d
,
, respectively.















Figure 2.18: Change of angle.
The vectors
) 1 (
x
,
d and
) 2 (
x
,
d are given, respectively, by:
) 1 ( ) 1 (
X F x
,
,
d d
P

) 1 ( ) 1 (
) (
k
P
jk j
X d F d
,
,
x

(2.146)
and
) 2 ( ) 2 (
X F x
,
,
d d
P

) 2 ( ) 2 (
) (
k
P
jk j
X d F d
,
,
x

(2.147)
where
P
jk
F show us that the deformation gradient is evaluated at the material point P .
Afterwards, the scalar product (
) 2 ( ) 1 (
x x
, ,
d d ) is expressed in the following manner:

) 2 ( ) 2 (
) 1 ( ) 1 (
) 2 ( ) 2 (
) 1 ( ) 1 (
ds d R P
ds d Q P
dS d PR
dS d PQ



x
x
X
X
,
,
,
,


1 1
, x X

2 2
, x X

3 3
, x X

0
B


t
B

3 3

e I

2 2

e I

1 1

e I
O
P
X
,

Reference
configuration - 0 t
Current
configuration - t
Q
P
Q ) 2 (
X
,
d

) 2 (
x
,
d
x
,


) 1 (
X
,
d
R

) 1 (
x
,
d
R
O
0
N


n
m
X X F x
, ,
,
d t d ) , (
2 CONTINUUM KINEMATICS

187
N E M
N C M
X C X
X F F X
X F X F x x
C

) 2 (


) 2 ( ) 1 (
) 2 ( ) 1 (
) 2 ( ) 1 (
) 2 ( ) 1 (
) 2 ( ) 1 ( ) 2 ( ) 1 (





-

1 dS dS
dS dS
d d
d d
d d d d
T
, ,
,
_
,
, ,
, ,
j ij ij i
j ij i
j ij i
j
ij
kj ki i
j kj i ki k k
N E M dS dS
N C M dS dS
dX C dX
dX
C
F F dX
dX F dX F dx dx

) 2 (


) )( (
) 2 ( ) 1 (
) 2 ( ) 1 (
) 2 ( ) 1 (
) 2 ( ) 1 (
) 2 ( ) 1 ( ) 2 ( ) 1 (
-

c
_

(2.148)
Additionally, the scalar product (
) 2 ( ) 1 (
X X
, ,
d d ) can be expressed as:
n e m
n c m
x c x
x F F x
x F x F X X
c
) 2 (


) 2 ( ) 1 (
) 2 ( ) 1 (
) 2 ( ) 1 (
) 2 (
1
) 1 (
) 2 ( 1 ) 1 ( 1 ) 2 ( ) 1 (





-

- -
- -
1 ds ds
ds ds
d d
d d
d d d d
T
, ,
,
_
,
, ,
, ,

j ij ij i
j ij i
j ij i
j
ij
kj ki i
j kj i ki k k
n e m ds ds
n c m ds ds
dx c dx
dx
c
F F dx
dx F dx F dX dX
) 2 (

) )( (
) 2 ( ) 1 (
) 2 ( ) 1 (
) 2 ( ) 1 (
) 2 (
1 1
) 1 (
) 2 ( 1 ) 1 ( 1 ) 2 ( ) 1 (
-

- -
- -
c
_

(2.149)
We use the equation in (2.148) to evaluate the magnitude of
) 1 (
x
,
d and
) 2 (
x
,
d in terms of the
deformation tensors. To do this, in equation (2.148) we enforce that
) 1 ( ) 2 (
x x
, ,
d d , which
leads to:
M C M x
M E M
M C M x x x

) 2 (


) 1 ( ) 1 (
) 1 ( ) 1 (
) 1 ( ) 1 (
2
) 1 ( ) 1 ( ) 1 (



-

dS d
dS dS
dS dS d d d
,
, , ,
1

(2.150)
Similarly, we can obtain the magnitude of
) 2 (
x
,
d as:
N C N x
N E N
N C N x x x

) 2 (


) 2 ( ) 2 (
) 2 ( ) 2 (
) 2 ( ) 2 (
2
) 2 ( ) 2 ( ) 2 (



-

dS d
dS dS
dS dS d d d
,
, , ,
1

(2.151)
Now by using the definition in (2.149) we can express the magnitude of
) 1 (
X
,
d and
) 2 (
X
,
d
as:
m c m X
m e m
m c m X X X

) 2 (

) 1 ( ) 1 (
) 1 ( ) 1 (
) 1 ( ) 1 (
2
) 1 ( ) 1 ( ) 1 (



-

ds d
ds ds
ds ds d d d
,
, , ,
1

(2.152)
and
n c n X
n e n
n c n X X X

) 2 (

) 2 ( ) 2 (
) 2 ( ) 2 (
) 2 ( ) 2 (
2
) 2 ( ) 2 ( ) 2 (



-

ds d
ds ds
ds ds d d d
,
, , ,
1

(2.153)
2.6.4.1 The Relationship between the Strain and Stretch Tensors
Next we can establish the relationship between the stretch, unit extension and strain
tensors. To do so we can start by defining the stretch (see equation (2.44)). Then the
stretch along direction M

, (see Figure 2.18), can be obtained by means of Lagrangian


variables as:

NOTES ON CONTINUUM MECHANICS

188
M E M M E M
M C M
M C M
X
x
M

2 1

) 2 (



) 1 (
) 1 (
) 1 (
) 1 (




- -
/
1
dS
dS
d
d
,
,

(2.154)
where we have used the term
) 1 (
x
,
d given in (2.150). If we now use the Eulerian variable,
the stretch along direction m , (see Figure 2.18), is defined as:
m e m m e m
m c m m c m X
x
m
2 1
1
) 2 (
1

1

) 1 (
) 1 (
) 1 (
) 1 (

/
1
ds
ds
d
d
,
,
(2.155)
Later on, we will show that
M
m
/ / . Once the stretch has been defined in terms of strain
tensors, and bearing in mind that the unit extension and stretch are related by the definition
in (2.46), i.e. 1 - / , we can express the unit extension along direction M

in terms of
Lagrangian variables as:

1

2 1 1

) 2 (

1

1

- - - -
- - /


M E M M E M
M C M
M M
1


(2.156)
We can also evaluate the unit extension in terms of Eulerian variables as:

1
2 1
1
1
) 2 (
1
1

1
1

-
-
-
-

- - /


m e m m e m
m c m
m m
1


(2.157)
In short, we can state:

- - - /
- - /


1

2 1 1

) 2 (

2 1

M E M
M C M M E M M E M
M M
M

1
Stretch and unit
extension according to
the M

-direction, in
terms of C and E
(2.158)
and

- -
-
- /

-
/



-
1

1
1
2 1
1
1

1

1
2 1
1

1

m c m m e m
m b m
m c m m e m
m m
m

Stretch and unit


extension according
to the m -direction,
in terms of c and e
(2.159)
Notice that, for any given motion, if there is no stretch ( 1

/
m
) in a particular direction
( m ), it holds that 1 m c m or 0 m e m .
2.6.4.2 Change of Angle
The angle between the vectors
) 1 (
x
,
d and
) 2 (
x
,
d , (see Figure 2.18), can be obtained by means
of the definition of the scalar product 0 cos
) 2 ( ) 1 ( ) 2 ( ) 1 (
x x x x
, , , ,
d d d d , the outcome of
which is:
2 CONTINUUM KINEMATICS

189
( )
N M
N C M
N C N M C M
N C M
N C N M C M
N C M
x x
x x

) 2 ( ) 1 (
) 2 ( ) 1 (
) 2 ( ) 1 (
) 2 ( ) 1 (






cos
/ /

0





dS dS
dS dS
d d
d d
, ,
, ,
(2.160)
where we have used the equations in (2.148), (2.150) and (2.151). We can summarize the
different ways of expressing 0 cos as:
( )
( ) ( ) N E N M E M
N E M N M
N E N M E M
N E M
N C M
N C N M C M
N C M
N M

2 1

2 1

2




cos








- -
-

- -
-

/ /
0
1 1
1
(2.161)
Likewise, we can evaluate the angle in the reference configuration as:
( )
( ) ( ) n e n m e m
n e m
n c m
n c n m c m
n c m
n c n m c m
n c m
X X
X X
n m
2 2
2






cos

) 2 ( ) 1 (
) 2 ( ) 1 (
) 2 ( ) 1 (
) 2 ( ) 1 (







- -
-
/ /

O
1 1
1
ds ds
ds ds
d d
d d
, ,
, ,
(2.162)
where we have used the equations in (2.149), (2.152) and (2.153). Then, we can summarize
O cos as:
( )
( ) ( ) n e n m e m
n e m n m
n e n m e m
n e m
n c m
n c n m c m
n c m
n m
2 1 2 1
2
2 2
2
) (


cos








- -
-

- -
-

/ / O
1 1
1
(2.163)
Taking into account that O cos

N M and 0 cos n m , the equations in (2.161) and
(2.163) become:
( )
N M N M N M
N E M N E M N M N E M



2 cos


2

cos
/ /
- O

/ /
-

/ /
-
0
1
(2.164)
and
( ) [ ] [ ] [ ]
n m n m n m
n e m n e m n m n e m

2 cos 2 2 cos / / - 0 / / - / / - O 1 (2.165)

2.6.4.3 The Physical Interpretation of the Deformation/Strain Tensor
Components. The Right Stretch Tensor
2.6.4.3.1 The Normal Components
Let us consider the Cartesian components of the right Cauchy-Green deformation tensor
at the material point P (particle), (see Figure 2.18).

NOTES ON CONTINUUM MECHANICS

190











Figure 2.19: Cartesian components of C .
Now let us state that the unit vector M

, shown in Figure 2.18, has the same direction as


the
1
X -axis, i.e.
1

e M . So, the product M C M



becomes:
[ ]
11
33 23 13
23 22 12
13 12 11
0
0
1
0 0 1

C
C C C
C C C
C C C
M C M
j ij i

(
(
(

(
(
(

M C M
(2.166)
Referring to the definition of stretch given in (2.44), we can conclude that:
0 ; 2 1

1 1
11 11
) 1 (
) 1 (
> / - /
X X
E C
d
d
M C M
X
x
,
,
(2.167)
As we can see,
11
C is the stretch measurement along the
1
X -axis. Similarly,
22
C and
33
C
show the stretch along
2
X and
3
X , respectively, i.e.:
33 33
22 22
11 11
2 1
2 1
2 1
3
2
1
E C
E C
E C
X
X
X
- /
- /
- /

( )
( )
( ) 1
2
1
1
2
1
1
2
1
2
33
2
22
2
11
3
2
1
- /
- /
- /
X
X
X
E
E
E
(2.168)
Therefore, the conclusion is that the diagonal terms E and C are related to the stretches.
Notice that, C is a symmetric positive definite tensor, and if we are working in the C
principal space, it follows that:
( )
( )
( )
(
(
(
(
(
(

- /
- /
- /

(
(
(

/
/
/

1
2
1
0 0
0 1
2
1
0
0 0 1
2
1
0 0
0 0
0 0
2
3
2
2
2
1
2
3
2
2
2
1
ij ij
E C (2.169)
where 0
1
> / , 0
2
> / , 0
3
> / are the principal stretches, which by definition are positive real
numbers, (see equation (2.44)). Then, the spectral representations of C and E are
expressed as follows:

22
C

2
X

3
X

11
C

12
C

13
C

12
C

23
C

33
C

13
C

23
C

1
X

3

e

2

e

1

e

(
(
(


33 23 13
23 22 12
13 12 11
C C C
C C C
C C C
C C
ji ij

2 CONTINUUM KINEMATICS

191

/
3
1
) ( ) ( 2

a
a a
a
N N C
The spectral representation of
the right Cauchy-Green
deformation tensor
(2.170)

- /
3
1
) ( ) ( 2

) 1 (
2
1
a
a a
a
N N E
The spectral representation of
the Green- Lagrange strain
tensor
(2.171)
In addition, by means of the spectral representation of C we can define a new tensor such
that C
2
U , where U denotes the right stretch tensor, and where the only possible solution
for U is C - U . Since the stretches are by definition positive real numbers, it follows
that the tensor U is definite positive. We can then define the right stretch tensor as:

/
3
1
) ( ) (

) , (
a
a a
a
t N N U X
,
The spectral representation of
the right stretch tensor
(2.172)
2.6.4.3.2 The Tangential Components
Now let us state that
1

e M and
2

e N . So, the product N C M



becomes:
[ ]
12
33 23 13
23 22 12
13 12 11
0
1
0
0 0 1

C
C C C
C C C
C C C
N C M
j ij i

(
(
(

(
(
(

N C M
(2.173)
With the following we can verify that the above term is related to 0 cos , (see equation
(2.161)):
22 11
12
22 11
12

2 1 2 1
2 1



cos
E E
E
C C
C
- -
-

/ /
0



N M
N C M
N C N M C M
N C M

(2.174)
So,
12
C measures the angle change between two differential line elements in the reference
configuration. Therefore, the off-diagonal terms E and C contain information about the
angle change.
2.7 Particular Cases of Motion
2.7.1 Homogeneous Deformation
If we consider an example in which the motion of all the particles is characterized by the
same deformation gradient, it follows that F is independent of the position vector X
,
, and
it is therefore only dependent on time, ) (t F F . This type of motion is an example of
homogeneous deformation. By integrating the equation X F x
,
,
d t d ) ( , we obtain:
) ( ) ( t t c
,
,
,
- X F x (2.175)
where the constant of integration c
,
shows translational motion, which is only dependent
on time.
Motion characterized by homogeneous deformation has the following characteristics:
NOTES ON CONTINUUM MECHANICS

192
A material surface defined by a plane in the reference configuration will remain a
plane in the current (deformed) configuration. Therefore, any material line in the
reference configuration will remain a line in the deformed configuration;
A material surface defined by a sphere in the reference configuration, will appear as
an ellipsoid in the current configuration. Therefore, any material curve defined by a
circle in the reference configuration will become an ellipse in the deformed
configuration.
2.7.2 Rigid Body Motion
We can state that a body undergoes rigid body motion when the distance between particles
are constant during motion. Under these conditions we can conclude that rigid body
motion is a specific case of homogenous deformation. Let us consider a vector in the
reference configuration A
,
. After motion, this vector is represented by a
,
(deformed
configuration). Then according to equation (2.175), it follows that A F a
,
,
. Moreover, as
the distances between particles do not change it holds that: A a
,
,
. In this situation, we
can conclude that ) (t F is an orthogonal tensor, i.e. ) ( ) ( ) ( ) (
1
t t t t
T
Q
-
F F F , (see
orthogonal tensor, Chapter 1). Hence, here the left and right Cauchy-Green deformation
tensors become:
1 Q Q
1 Q Q




T T
T T
F F b
F F C
The right and left Cauchy-Green
deformation tensor related to rigid body
motion
(2.176)
In addition, we find that for rigid body motion the following is satisfied:
( )
( )
0 D
0 1
0 1

-
-
-1
2
1
2
1
b e
C E
The strain tensors for rigid body motion (2.177)
In rigid body motion the stretches are unitary, since the distance between particles does not
change, and it is possible to check the previous result by means of the spectral
representation of C and E , (see equations (2.170) and (2.171)):
1 N N N N /


3
1
) ( ) (
3
1
) ( ) ( 2

a
a a
a
a a
a
C (2.178)
0 N N N N - /


3
1
) ( ) (
3
1
) ( ) ( 2

0

) 1 (
2
1
a
a a
a
a
a a
a
E (2.179)
NOTE: To ensure that the continuum is subjected to rigid body motion, the equation
0 E
`
or 0 D must be valid for all material points throughout the continuum.
Problem 2.13: Let us consider the following equations of motion:
3 3 2 1 2 2 1 1
;
2
1
;
2
1
X x X X x X X x - - (2.180)
a) Obtain the displacement field ( u
,
) in the Lagrangian and Eulerian descriptions;
b) Determine the material curve in the current configuration for a material circle defined in
the reference configuration as:
2 CONTINUUM KINEMATICS

193
0 2
3
2
2
2
1
- X X X
c) Obtain the components of the right Cauchy-Green deformation tensor and the Green-
Lagrange strain tensor;
d) Obtain the principal stretches.
Solution:
The deformation gradient is given by:
75 . 0 ;
2 0 0
0 2 1
0 1 2
2
1

(
(
(

o
o
F J
X
x
F
j
i
ij

And by comparing this with the equations of motion in (2.180) we have:
j ij i
X F x
X
X
X
x
x
x

(
(
(

(
(
(

(
(
(

;
2 0 0
0 2 1
0 1 2
2
1
3
2
1
3
2
1

So, we can verify that the proposed example is a case of homogeneous deformation in
which 0 c
,
,
. The inverse form of the above equation is given by:

- -
-

(
(
(

(
(
(

-
-

(
(
(

3 3
2 1 2
2 1 1
3
2
1
3
2
1
3
4
3
2
3
2
3
4
3 0 0
0 4 2
0 2 4
3
1
x X
x x X
x x X
x
x
x
X
X
X
(2.181)
The displacement field is defined by X x
,
, ,
- u , after which the components of the
Lagrangian displacement become:

-
- - -
- - -
-
0 ) , (
2
1
2
1
) , (
2
1
2
1
) , (
3 3 3
1 2 2 1 2 2 2
2 1 2 1 1 1 1
X x t
X X X X X x t
X X X X X x t
X x
i i i
X
X
X
,
,
,
u
u
u
u
(2.182)
The components of the Eulerian displacement can be obtained by substituting the Eulerian
description of motion (2.181) into (2.182), the result of which is:

- -

- -
0 ) , ( ) ), , ( (
) , (
3
4
3
2
2
1
) , (
2
1
) ), , ( (
) , (
3
4
3
2
2
1
) , (
2
1
) ), , ( (
2 3
2 2 1 1 2
1 2 1 2 1
t t t
t x x t X t t
t x x t X t t
x x X
x x x X
x x x X
, ,
,
, , ,
,
, , ,
,
u u
u u
u u
(2.183)
The particles belonging to the circle 2
2
2
2
1
- X X in the reference configuration will form
a new curve in the current configuration which is defined by:
18 20 32 20 2
3
4
3
2
3
2
3
4
2
2
2 2 1
2
1
2
2 1
2
2 1
2
2
2
1
- -
(

- - -
(

- - x x x x x x x x X X
which is an ellipse equation (Figure 2.20 shows the material curve in different
configurations).
NOTES ON CONTINUUM MECHANICS

194
The components of C and E can be obtained by using the definitions F F C
T
and
( ) 1 - C E
2
1
:
(
(
(

(
(
(

(
(
(


1 0 0
0 25 . 1 1
0 1 25 . 1
2 0 0
0 2 1
0 1 2

2 0 0
0 2 1
0 1 2
4
1
ij kj ki ij
C F F C
( )
(
(
(

|
|
|
.
|

\
|
(
(
(

-
(
(
(

o -
0 0 0
0 125 . 0 5 . 0
0 5 . 0 125 . 0
1 0 0
0 1 0
0 0 1
1 0 0
0 25 . 1 1
0 1 25 . 1
2
1
2
1
ij ij ij ij
E C E
In the principal space of C its components are given by:
(
(
(

/
/
/

(
(
(

/
/
/

3
2
1
2
3
2
2
2
1
0 0
0 0
0 0
0 0
0 0
0 0
ij ij
C C
where
i
/ show the principal stretches. Therefore, to calculate these we need to obtain the
C eigenvalues:

- -
-
-
25 . 0
25 . 2
0 5625 . 0 5 . 2 0
25 . 1 1
1 25 . 1
2
1 2
C
C
C C
C
C

(
(
(

(
(
(

/
/
/

(
(
(

(
(
(

/
/
/

1 0 0
0 5 . 0 0
0 0 5 . 1
0 0
0 0
0 0
1 0 0
0 25 . 0 0
0 0 25 . 2
0 0
0 0
0 0
3
2
1
2
3
2
2
2
1
ij
C

-2.0
-1.5
-1.0
-0.5
0.0
0.5
1.0
1.5
2.0
-2 -1 0 1 2
x1
x
2
Reference Conf.
Current Conf.

Figure 2.20: Material curve.

material curve
2 CONTINUUM KINEMATICS

195
Problem 2.14: Let us consider the following velocity field:

- -
-
- -
2 1 3
3 1 2
3 2 1
5 1
5 3
1 3
x x v
x x v
x x v

Show that this motion corresponds to rigid body motion.
Solution: First we obtain the components of the spatial velocity gradient ( ) l :
skew
ij
j
i
ij
x
v
x
v
x
v
x
v
x
v
x
v
x
v
x
v
x
v
x
t v
l l
(
(
(

-
-
-

(
(
(
(
(
(
(

o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o

o
o

0 5 1
5 0 3
1 3 0
) , (
3
3
2
3
1
3
3
2
2
2
1
2
3
1
2
1
1
1
x
,

Taking into account that l can be decomposed into a symmetric ( D =
sym
l ) and an
antisymmetric ( W =
skew
l ) part, i.e. W D- l , we can thus conclude that 0 D , which is
a characteristic of rigid body motion.
2.8 Polar Decomposition of F
As mentioned in Chapter 1, a non-singular second-order tensor can be decomposed
multiplicatively by means of the polar decomposition theorem. By applying polar decomposition
to the deformation gradient F which is a non-singular tensor 0 ) ( F det and 0 ) ( > F det ,
we obtain:


ion decomposit polar
Left
ion decomposit polar
Right


R V U R F
(2.184)
where R is a proper orthogonal tensor (rotation tensor), which must meet:
_
ity orthogonal
T T T 1 -
R R 1 R R R R and
_
proper
1 ) ( R det . U and V are symmetric positive definite
tensors, and are known as:
U - The right stretch tensor, the Lagrangian stretch tensor, or the material stretch tensor.
V - The left stretch tensor, the Eulerian stretch tensor, or the spatial stretch tensor.
In the right polar decomposition, we first carry out a transformation just with strain, and
then we make a transformation characterized by a rotation, (see Figure 2.21), whereas, in
the left polar decomposition, we first carry out an orthogonal transformation (rotation) and
then transformation only with strain is applied.
With the right polar decomposition it holds that F
T
R U , and also by applying the
scalar product between
T
F and the equation in (2.184) we obtain:
2
U U U U R
T T T
C F F F
C
_

(2.185)
NOTES ON CONTINUUM MECHANICS

196
In addition, based on the spectral representation of C , i.e.

/
3
1
) ( ) ( 2

a
a a
a
N N C , (see
equation (2.170)), we can conclude that

/
3
1
) ( ) (

a
a a
a
N N U , where
a
/ are the principal
stretches, and are positive numbers by definition, and so the tensor U is a positive definite
tensor.
Since the determinant of F is positive, 0 ) ( > F det , and the determinant of a positive
definite tensor is also positive 0 ) ( > U det , we can conclude that R is a proper orthogonal
tensor, i.e. a rotation tensor:
0 ) ( ) ( ) ( ) ( ) (
0 1
>
>
U U R U R det det det det det
_ _
F
(2.186)


















Figure 2.21: Polar decomposition of F.
Taking into account the polar decomposition of F , it is possible to express the right
Cauchy-Green deformation tensor, C , as:
2
) ( ) (
U U U
U R R U
U R U R

T T
T
T
F F C

R R
R V V R
R V R V


b
F F C
T
T
T
T
) ( ) (

(2.187)
and the left Cauchy-Green deformation tensor, b , as:

) 2 (

n

0
B
X
,
d

) 2 (

N

) 3 (

N

) 1 (

N
R V F
R
R
V
U
X
,
d
x
,
d
U R F
Reference
configuration
Current
configuration
X F x
,
,
d d
0
B
B
B
x
,
d

) 2 (

N

) 3 (

N

) 1 (

N

) 2 (

n

) 3 (

n

) 1 (

n

) 3 (

n

) 1 (

n
2 CONTINUUM KINEMATICS

197
2
1
) ( ) (
V V V
V R R V
R V R V

-
T T
T
T
F F c b

T
T
T
T
R R
R U U R
U R U R


C
F F b
) ( ) (

(2.188)
where c is the Cauchy deformation tensor. And the tensors C and b are interrelated to
each other by:
T T
R U R V R V R U ;

(2.189)
2.8.1 Spectral Representation of Kinematic Tensors
As we have seen before, the eigenvalues of U represent the principal stretches,
i
/ . Each
principal stretch (
i
/ ) is associated with a principal direction (
) (

i
N ), i.e.:
[ ]
[ ]
[ ]


for

for

for
) 3 (
3
) 3 (
2
) 3 (
1
) 3 (
3
) 2 (
3
) 2 (
2
) 2 (
1
) 2 (
2
) 1 (
3
) 1 (
2
) 1 (
1
) 1 (
1
N N N
N N N
N N N
/
/
/
N
N
N
(2.190)
Then, the spectral representation of U is given by:

/
3
1
) ( ) (

a
a a
a
N N U (2.191)




/ / /
|
|
.
|

\
|
/
|
|
.
|

\
|
/

3
1
) ( ) ( 2
3
1
) ( ) ( ) ( ) (
3
1
) ( ) (
3
1
) ( ) ( 2


1
a
a a
a
a
a a a a
a a
a
a a
a
a
a a
a
N N N N N N
N N N N U
_
C
(2.192)
Thus,

/
3
1
) ( ) ( 2 2

a
a a
a
N N U C
The spectral representation of
the right Cauchy-Green
deformation tensor
(2.193)
If we can verify that, U and C are coaxial tensors, then it holds that
3
U U U C C .
Based on the principle that C and b have the same principal invariants, (see equation
(2.113)), then C and b have the same eigenvalues. We can prove this after having defined
the eigenvalues and eigenvectors of U:


N N U
a
/
(2.194)
By substituting the equation R V R U
T
into that in (2.194), we obtain:

N N R V R N U
n

a
T
/
(2.195)
Now, by applying the scalar product between R and (2.195) we obtain:

n n V
N R n V R R
n

a
a
T
/
/


(2.196)
NOTES ON CONTINUUM MECHANICS

198
Thus, we find that U and V have the same eigenvalues
i
/ , but different eigenvectors so
the spectral representation of V and b are given by:

/
3
1
) ( ) (

a
a a
a
n n V
The spectral representation of the left
stretch tensor
(2.197)

/
3
1
) ( ) ( 2 2

a
a a
a
n n V b
The spectral representation of the left
Cauchy-Green deformation tensor
(2.198)
Next, we show F in terms of the eigenvalues of U,
a
/ , and eigenvectors N

, n

. If we
consider that F
T
R U and R n n R N

T
, it holds that:
( ) ( ) n R N R
N N R
N N U



a
T T
a
T
a
/
/
/



F
F
(2.199)
Thus, we can conclude that:
n N

a
/ F (2.200)
Taking into account that R n N

, the deformation gradient can also be expressed as:





/
|
|
.
|

\
|
/
3
1 ) (
) ( ) (
3
1
) ( ) (


a a
a a
a
a
a a
a _
N
R n n R n n R V F
(2.201)
whereas its inverse
1 1 1
) (
- - -
V R R V
T
F is:


- -

|
|
.
|

\
|


3
1
) (
) (
) (
3
1
) ( ) ( 1 1

1

1

a
a
a
a T
a a
a a
a
T T
n n R n n R V R
N
_
F
(2.202)
Thus, we can conclude that:


-


/
/
3
1
) ( ) ( 1
3
1
) ( ) (

1
;

a
a a
a a
a a
a
n N N n F F
Spectral representation of
the deformation gradient
(2.203)
By means of the spectral representation of the deformation gradient, we can see that F
is neither in the reference nor in the current configuration. It is as if it were straddling both
of them.
By making use of the left polar decomposition, F F
-

1
V R R V , the spectral
representation of the orthogonal tensor of the polar decomposition can be obtained as
follows:


/
/

|
|
.
|

\
|
/
|
|
.
|

\
|

3
1
) ( ) ( ) ( ) (
3
1
) ( ) (
3
1
) ( ) (


1
a
a a a a
a
a
a
a a
a
a
a a
a
N n n n
N n n n R

(2.204)
Thus:
2 CONTINUUM KINEMATICS

199
) ( ) (
3
1
) ( ) (

i i
a
a a
N n N n R

The spectral representation of the


orthogonal tensor
(2.205)
Additionally, the spectral representation of E and e are given by:
( ) ( )

- / - -
3
1
) ( ) ( 2 2

) 1 (
2
1
2
1
2
1
a
a a
a
N N 1 U 1 C E

The spectral representation of
the Green-Lagrange strain
tensor
( ) ( )

- - -
/ - - -
3
1
) ( ) ( 2 2 1

) 1 (
2
1
2
1
2
1
a
a a
a
n n V 1 1 b e
Spectral representation of the
Almansi strain tensor
(2.206)
Next, we can establish a connection between the configurations
0
B ,
0
B , B and B , (see
Figure 2.22). To start with we can observe:
X X X F x
,
, ,
,
d d d d V R V
x X X F x
, , ,
,
d d d d R U R
(2.207)
then:
( ) ( ) x X X x x X
F F
,
_
, ,
_
, ,
,
d d d d d d
T T
T

-
-

R V V R R V
1

(2.208)
Using the equation shown in (2.111) we obtain:
1
1
1
-
-
-


V V
V R R V
C
C
F C F b
T

T
T
R R
R U U R


-
-
C
C
F C F b
1
1
(2.209)
Notice also that C C C
- -

2 1 2 1
U U U U U U . So we can conclude that:
R R
R V R
R V V V R
R V V R




-
-
-
b
b
F b F C
T
T
T
T
2
2 1
1
1
U U
U R R U



-
-
-

b
b
F b F C
1
1
1
T

(2.210)
All the equations obtained above can be appreciated in Figure 2.22. We leave the reader to
make the necessary algebraic operations with the inverse tensors.









NOTES ON CONTINUUM MECHANICS

200


























Figure 2.22: Polar decomposition of F.
Problem 2.15: Let us consider the Cartesian components of the deformation gradient:
(
(
(

4 2 2
3 6 2
3 3 5
ij
F
obtain the tensors U (right stretch tensor), V (left stretch tensor), and R (rotation tensor).
Solution:
Before obtaining the tensors U, V , R , we analyze the deformation gradient F .
The motion is possible if the determinant of F is greater than zero, 0 60 ) ( > F det . The
eigenvalues and eigenvectors of F are given by:
10
11
F associated with eigenvector [ ] 4264014327 . 0 ; 6396021491 . 0 ; 6396021491 . 0
) 1 (

i
m
3
22
F associated with [ ] 3713906764 . 0 ; 7427813527 . 0 ; 5570860145 . 0
) 2 (
- -
i
m
R
R
V
U
X
,
d
x
,
d
Reference
configuration
Current
configuration
X X
,
,
d d R

0
B
B
B
x
,
d

0
B

T T
F V R

U U
U U
U

- - -
-


1 1 1
1
2
b C
b C
F F C
T


T - -
F R V
1


T
T
R R
R R

C b
C C


2 1
2 1
1 *
- -
-
-



V
V V V
V V
b
b b
C b


R R
R R


- -

1 1
b b
b b
T
T


R R
R R


- -

1 1
b C
b C
T
T


T
T
T
R R
R R
R R



- -
- -

1 1
1 1
C b
b b
b b


2 1 1 1
2 1
- - - -
-




U U U
U U U
C C
C C

X x
,
,
d d V
x x
,
,
d d R
X x
, ,
d d U

2 1
2
- -

V
V
b
b

X
,
d
b C
C b C
2 CONTINUUM KINEMATICS

201
2
33
F associated with [ ] 8164965809 . 0 ; 4082482905 . 0 ; 4082482905 . 0
) 3 (
- -
i
m
It is easy to check that the basis formed by these eigenvectors does not form an orthogonal
basis, i.e. 0
) 2 ( ) 1 (

i i
m m , 0
) 3 ( ) 1 (

i i
m m , 0
) 3 ( ) 2 (

i i
m m . We can also verify that if D is the matrix
containing the eigenvectors of F :
(
(
(

- -
- -
(
(
(

8164965809 . 0 ; 4082482905 . 0 ; 4082482905 . 0


3713906764 . 0 ; 7427813527 . 0 ; 5570860145 . 0
4264014327 . 0 ; 6396021491 . 0 ; 6396021491 . 0

) 3 (
) 2 (
) 1 (
i
i
i
m
m
m
D
we find that 1 905 . 0 ) ( D det , and
T
D D
-1
. However, it holds that:
(
(
(

(
(
(

(
(
(

(
(
(

- -
2 0 0
0 3 0
0 0 10
4 3 3
2 6 3
2 2 5
) (
4 3 3
2 6 3
2 2 5
2 0 0
0 3 0
0 0 10
1 1
D D D D and
ij
T
F
The right Cauchy-Green deformation tensor components, F F C
T
, are given by:
(
(
(


34 35 29
35 49 31
29 31 33
kj ki ij
F F C
Then the eigenvalues and eigenvectors of C are given by:

r eigenvecto
C 274739 . 9
11
[ ] 1894472683 . 0 ; 7023576528 . 0 ; 6861511933 . 0

) 1 (
-
i
N

r eigenvecto
C 770098 . 3
22
[ ] 8132215099 . 0 ; 2793856273 . 0 ; 5105143234 . 0

) 2 (
-
i
N

r eigenvecto
C 955163 . 102
33
[ ] 550264423 . 0 ; 65470405 . 0 ; 518239 . 0

) 3 (
- - -
i
N
These eigenvectors constitute an orthogonal basis, so, it holds that
T
C C
A A
-1
, and
1 ) ( -
C
A det (improper orthogonal tensor):
(
(
(

- - -
-
-

(
(
(

550264423 . 0 65470405 . 0 518239 . 0


8132215099 . 0 2793856273 . 0 5105143234 . 0
1894472683 . 0 7023576528 . 0 6861511933 . 0

) 3 (
) 2 (
) 1 (
i
i
i
N
N
N
C
A
Furthermore, it holds that:
(
(
(

(
(
(

(
(
(

(
(
(

33
22
11
33
22
11
0 0
0 0
0 0
34 35 29
35 49 31
29 31 33
;
34 35 29
35 49 31
29 31 33
0 0
0 0
0 0
C
C
C
C
C
C
C
T
ij
T
C C C C
A A A A
In the C principal space we obtain the components of the right stretch tensor, U, as:
(
(
(

(
(
(

(
(
(

/
/
/

1466824 . 10 0 0
0 9416741 . 1 0
0 0 0454455 . 3
0 0
0 0
0 0
0 0
0 0
0 0
33
22
11
3
2
1
C
C
C
ij
U U
and its inverse:
(
(
(
(
(
(

(
(
(
(
(
(
(

/
/
/

- -

1466824 . 10
1
0 0
0
9416741 . 1
1
0
0 0
0454455 . 3
1
1
0 0
0
1
0
0 0
1
3
2
1
1 1
ij
U U
We can evaluate the components of the tensor U in the original space by means of the
transformation law:
NOTES ON CONTINUUM MECHANICS

202
ij
T
U
(
(
(


46569091 . 4 80907159 . 2 48328843 . 2
80907159 . 2 00314487 . 6 25196988 . 2
48328843 . 2 25196988 . 2 66496626 . 4
C C
A U A
and
1 1
38221833 . 0 12519889 . 0 14302659 . 0
12519889 . 0 24442627 . 0 25196988 . 2
14302659 . 0 05134777 . 0 31528844 . 0
- -

(
(
(

- -
-
- -

ij
T
U
C C
A U A
Then, the rotation tensor of the polar decomposition is given by the equation
1 -
U R F ,
which is a proper orthogonal tensor, i.e. 1 ) ( R det .
(
(
(

- -
-
-
9924224 . 0 11463858 . 0 04422505 . 0
10940847 . 0 98826538 . 0 10658955 . 0
05592536 . 0 10094326 . 0 9933191 . 0
1
kj ik ij
F U R
The left Cauchy-Green deformation tensor components,
T
F F b , are given by:
(
(
(


24 28 28
28 49 37
28 37 43
jk ik ij
F F b
Next, the eigenvalues and eigenvectors of b are given by:

r eigenvecto
b 274739 . 9
11
[ ] 238183919 . 0 7465251613 . 0 6212637156 . 0

) 1 (
-
i
n

r eigenvecto
b 770098 . 3
22
[ ] 8616587383 . 0 1327190337 . 0 4898263742 . 0

) 2 (
-
i
n

r eigenvecto
b 95516 . 102
33
[ ] 448121233 . 0 6519860747 . 0 611638389 . 0

) 3 (
- - -
i
n
Note that, the tensors b and C have the same eigenvalues but different eigenvectors. If
the eigenvectors of b constitute an orthogonal basis then it holds that
T
b b
A A
-1
, and
1 ) ( -
b
A det :
(
(
(

- - -
-
-

(
(
(

448121233 . 0 6519860747 . 0 611638389 . 0


8616587383 . 0 1327190337 . 0 4898263742 . 0
238183919 . 0 7465251613 . 0 6212637156 . 0

) 3 (
) 2 (
) 1 (
i
i
i
n
n
n
b
A
and, it also holds that:
(
(
(

(
(
(

(
(
(

(
(
(

33
22
11
33
22
11
0 0
0 0
0 0
24 28 28
28 49 37
28 37 43
;
24 28 28
28 49 37
28 37 43
0 0
0 0
0 0
b
b
b
b
b
b
b
T
ij
T
b b b b
A A A A
Since C and b have the same eigenvalues, it follows that
ij ij
V U , i.e. they have the same
components in their respectively principal space. Additionally, it holds that
1 1 - -

ij ij
V U .
The components of the tensor V in the original space can be evaluated by:
ij
T T
V
(
(
(


3.6519622 2.20098553 2.41222612
2.20098553 6.04463857 2.76007379
2.41222612 2.76007379 5.3720129
b b b b
A U A A V A
and
1 1 1
0.42079849 0.08848799 0.14176921
0.08848799 0.23396031 0.07950684
0.14176921 0.07950684 0.28717424
- - -

(
(
(

- -
- -
- -

ij
T T
V
b b b b
A U A A V A
The polar decomposition rotation tensor obtained previously has to be the same as the one
obtained by F
-

1
V R .
2 CONTINUUM KINEMATICS

203
We could also have obtained the tensors U, V , R , by means of their spectral
representation. That is, if we know the principal stretches,
i
/ , and the eigenvectors of C
(
) (

i
N ), and the eigenvectors of b (
) (

i
n ), it is easy to show that:
) 3 ( ) 3 (
3
) 2 ( ) 2 (
2
) 1 ( ) 1 (
1
3
1
) ( ) (

j i j i j i
ij
a
a a
a ij
N N N N N N U / - / - /
|
|
.
|

\
|
/

N N
) 3 ( ) 3 (
3
) 2 ( ) 2 (
2
) 1 ( ) 1 (
1
3
1
) ( ) (

j i j i j i
ij
a
a a
a ij
n n n n n n V / - / - /
|
|
.
|

\
|
/

n n
) 3 ( ) 3 ( ) 2 ( ) 2 ( ) 1 ( ) 1 (
3
1
) ( ) (

j i j i j i
ij
a
a a
ij
N n N n N n R - -
|
|
.
|

\
|

N n
R V U R
R n n N N R
R n n N N R
N n




|
|
.
|

\
|
/
|
|
.
|

\
|
/
/ /
/ - / - /
|
|
.
|

\
|
/

3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (
) 3 ( ) 3 (
3
) 2 ( ) 2 (
2
) 1 ( ) 1 (
1
3
1
) ( ) (



a
a a
a
a
a a
a
a
a a
a
a
a a
a
j i j i j i
ij
a
a a
a ij
F
F
N n N n N n

As we can verify, the representations of the tensors R and F are not the spectral
representations in the strict sense of the word, i.e.,
i
/ are not eigenvalues of F , and
neither
) (

i
n nor
) (

i
N are eigenvectors of F .
2.8.2 Evolution of the Polar Decomposition
Using the right polar decomposition ( U R F ) as seen in (2.184), the material time
derivative of the deformation gradient ( F ) can also be evaluated by:
U R U R
` ` `
- F (2.211)
By considering equation (2.81), i.e. U R l l F F
`
, and by incorporating it into the
above equation we can obtain an equation for the spatial velocity gradient ( l ):
T T
R U U R R U U R
U U R U U R R
U R U R U R
1



- -
- -
-
-
-
1 1
1 1
`
_
`
` `
` `
l
l
l

(2.212)
Thus,
T T
R U U R R R
-
-
1
` `
l (2.213)
Notice that, in rigid body motion, 0 U 1 U
`
, the spatial velocity gradient becomes
T
R R
`
l . This is a prompt for us to introduce an antisymmetric second-order tensor, the
rate of the material rotation tensor (also called the angular-velocity tensor), and defined as:
T T
- R R
`
The rate of the material rotation tensor (2.214)
Additionally, the axial vector associated with is called the angular-velocity vector and is
denoted by c
,
.
NOTES ON CONTINUUM MECHANICS

204
We want to show that
T
- . To do so, we start from the orthogonality condition
1 R R
T
, it then follows that:

0
0 R R R R
1 R R
-
-



T
T T
T
Dt
D
Dt
D

` `
) ( ) (

T
-
(2.215)
It is also true that:

-
-
-
-
-





T
T T T
T T
T T
T T T
Dt
D
Dt
D
Dt
D
R R
R R R R R R
R 1 R R R
R R R R
R R R R R R
` `
` ` ` `
` ` ` `
` ` ` ` `
` ` `
) ( ) ( ) (
-
T
R R
` ` `

(2.216)
Taking into account (2.213), the rate-of-deformation tensor, D , can also be expressed as:
( ) ( )
( )
(

- - -
(

- - - -


- -
- -
T T T
T
T T T
T
T T T T T
R U U R R R R U U R R R
R U U R R R R U U R R R D
` ` ` `
` ` ` `
1
1 1
2
1
2
1
) (
2
1
l l

(2.217)
Notice that ( )
T
T T
R R R R -
` `
(antisymmetric tensor) and
T
U U (symmetric tensor).
Therefore, the above relationship becomes:
[ ]
T
R U U U U R D
- -
-
` `
1 1
2
1
(2.218)
Following the same reasoning, W can be expressed in terms of R and U as:
[ ]
T T T
R U U U U R R R W
- -
- - -
` ` `
1 1
2
1
) (
2
1
l l (2.219)
We can now attempt to graphically visualize the tensors obtained above. To do so, let us
consider Figure 2.23, in which the time domain is discretized by means of time increments
t A . And, at each time step we represent the right polar decomposition.










2 CONTINUUM KINEMATICS

205
























Figure 2.23: Evolution of the right polar decomposition.

As we can verify in Figure 2.23 we have represented the rate of change of U by means of
the tensor
U
l , which is in the intermediate configuration
) (t
B :
1 -
U U U U
U U
` `
l l
(2.220)
Moreover, we can see this in the current configuration (
U
l ) by means of an orthogonal
transformation, i.e.
T
R R
U U
l l , (see Figure 2.23). In general,
U
l is not a symmetric
tensor.
Likewise, we have represented the material time derivative of R by means of
R
l (current
configuration), and it follows that:

- T
R R R R R R
R R R
` ` `
l l l
1
(2.221)
Note that
R
l is the rate of the material rotation tensor (antisymmetric tensor). It is also
true that:
F

) 1 (
F

) 2 (
F
F F l
`

R
U

) 1 (
R

) 2 (
R

R
R R
R

l
`

) 1 (
U

) 2 (
U U U
U
l
`


1 -
U U
U
`
l

T
R R
R

`
l

T
R R
U U
l l
W D- l
B

) 1 (
B

) 2 (
B

0
B

) 1 (
B

) 2 (
B
B
.
.
.
.
.
.
t A
t A
t A
t A
t A
t A
0 At
0 At
0 At
NOTES ON CONTINUUM MECHANICS

206
T T
R U U R R R
U
U R

-
- -
_
` `
l
l l l
1

(2.222)
which is the same as that obtained in (2.213). The symmetric part of l can also be
expressed as:

( ) ( )
( ) ( )
[ ]
T
T
T T
T
T T sym sym sym sym
R U U U U R
R U U R R U U R
R R R R D
U U U R U
0



- -
- -
-
(

-
(

- - =

` `
` `
1 1
1 1
2
1
2
1
2
1
l l l l l l

(2.223)
which is the same as that obtained in (2.218). The antisymmetric part of l can also be
expressed as:
( ) ( )
( ) ( )
[ ]
T T
T
T
T T
T
T
T T skew skew skew skew
R R R U U U U R
R R R U U R R U U R
R R R R R R W
U U R U R U



- -
-
(

-
-
(

- - - =
- -
- -
` ` `
` ` `
`
1 1
1 1
2
1
2
1
2
1
l l l l l l l
(2.224)
which matches the equation in (2.219).
Now, if we refer to the left polar decomposition, R V F , the material time derivative of
F is:
R V R V
` ` `
- F
(2.225)
Using R V l l F F
`
, the above equation can be rewritten as:
R V R V R V
` `
- l
(2.226)
In addition, by applying the dot product with
1 -
V R
T
we obtain:
1 1 - -
- V R R V V V
T
` `
l
(2.227)
Based on (2.227), it is also true that:
( ) ( )
( ) ( )
( ) ( ) V V V V V V V V
V R R V V R R V V V V V
V R R V V V V R R V V V D



- - - -
- - - -
- - - -
- - -
- - -
- - - -

1 1 1 1
1 1 1 1
1 1 1 1
2
1
2
1
2
1
2
1
2
1
2
1
` `
` ` ` `
` ` ` `
T T
T T T
l l
(2.228)
and
( ) ( )
( )
( ) ( ) V V V V V V V V
V R R V V R R V V V V V
V R R V V V V R R V V V W



- - - -
- - - -
- - - -
- - -
- - -
- - - -

1 1 1 1
1 1 1 1
1 1 1 1
2
1
2
1
2
1
2
1
2
1
` `
` ` ` `
` ` ` `
T T
T T T
l l
(2.229)
2 CONTINUUM KINEMATICS

207
The tensors obtained above can be appreciated in Figure 2.24, where we represent the rate
of change of R by means of
T
R R
R
`
l , which is in the intermediate configuration
) (
0
t
B .
























Figure 2.24: Evolution of the left polar decomposition.
Additionally, the representation of
R
l in the current configuration is given by:
1 1 1 - - -
V V V R R V V V
R R

T
`
l l
(2.230)
The rate of change of V is represented by means of
V
l , which is in the current
configuration, and is given by:
1 -
V V
V
`
l (2.231)
Then, l can be represented as:
1 1 1 1 - - - -
- - - V R R V V V V V V V
R R V
T
` ` `
l l l l
(2.232)
F

) 1 (
F

) 2 (
F
F F l
`

R
V

) 1 (
R

) 2 (
R R
R R
R

l
`


) 1 (
V

) 2 (
V
V V
V
l
`


T
R R
R

`
l

1 -
V V
V
`
l

1 -
V V
R R
l l
W D- l

0
B

) 1 (
0
B

) 2 (
0
B

0
B

) 1 (
B

) 2 (
B
B
.
.
.
.
.
.
t A
t A
t A
t A
t A
t A
0 At
0 At
0 At
NOTES ON CONTINUUM MECHANICS

208
which is the same equation as that obtained in (2.227). The symmetric part of l can be
obtained as:
( ) ( ) V V V V V V V V D
R V

- - - -
- - - - =
T sym sym sym

1 1 1 1
2
1
2
1
` `
l l l (2.233)
And the antisymmetric part of l is given by:
( ) ( ) V V V V V V V V W
R V

- - - -
- - - - =
T skew skew skew

1 1 1 1
2
1
2
1
` `
l l l (2.234)
which matches the equation in (2.229).
2.8.2.1 The Alternative Way to Express the Rate of Kinematic Tensors
Let us consider the motion and evolution of the polar decomposition as shown in Figure
2.25. We can introduce a new configuration
0
B , which does not change over time. So, we
denote by
0
R the orthogonal transformation between the fixed basis
0

N and the principal


direction of the right stretch tensor N

, (see Figure 2.25). In fact, the basis


0

N is a system
that is fixed in
0
B . Here we have separated these configurations, i.e.
0
B and
0
B , so as to
have a better understanding of the process involved.
Remember that:
; ) (

x x X F x
, ,
,
,
d d
Dt
D
d d
change of rate
l
F F l
`
(2.235)
Then, by comparison we have:
;

0
0

0 0
N N N N N
R
R l
`
change of rate
0 0 0
R R
`
(2.236)
;


0
n n n N n
R
R l
` change of rate
R R
`

(2.237)
where we have made a name change:
0
0
=
R
l , =
R
l . It is now possible to show that
0
R R R , (see Figure 2.25) and the following condition is satisfied:
( ) ( )
0 0 0
0 0 0 0
0 0 0
0 0
R R
R R R
R R R
R R R




-
-
-
-



R R R
R R R
R R
R R

` ` `

(2.238)
where we have considered the following relationships: R R
`
,
0 0 0
R R
`
,
R R
`
. According to (2.238) we can conclude that:
T
R R R R R - = -
0 0


(2.239)
Starting from (2.239) and referring to the fact that
T
R R
`
, we can obtain the following
equation for R
`
:
-
0
) ( R R R R R
T
`

0
- R R R
`
(2.240)

2 CONTINUUM KINEMATICS

209


























Figure 2.25: Evolution of the polar decomposition.
The material time derivative of N R n

can be evaluated as follows:


( )
n
N R
N R R
N R N R
N R N R n

0
0

-
-
-

`
`
`

( )
( )
n
n R R
n R R R
N R N R
N R N R n

0
0
0

-
-
-
-

T
T

`
`
`

(2.241)
And, that of
0 0

N N R is given by:
N



0 0
0
0
0
R
R R
R

l
`


0

N

0

N
N


U
R

0
R

0
R R R
R
V
F

0
B

0
B
B

0
B

F F l
`


R R
R
l
`


V V
V
l
`


R
R R
R

l
`


U U
U
l
`

with

-
-
T
T
T
R R
R R
R
R
`
`
`
`
`
l
l
l l
l
l
0 0 0
0
1
1
R R
U U
R R
U
F F


R
R R
R

l
`

B
U

T
0 0
R R U U
n


NOTES ON CONTINUUM MECHANICS

210

N N N N N N N
0

0 0 0 0 0 0 0 0 0 0
- =

R R R R
` `
,
` `
Dt
D

(2.242)

By referring to the spectral representation of the second-order unit tensor, i.e.


3
1
) ( ) (

a
a a
N N 1 , and the equation in (2.236), N N

0

`
, we can conclude that:
0
3
1
) ( ) (
3
1
) ( ) (
0
3
1
) ( ) (
0 0





a
a a
a
a a
a
a a
N N N N N N 1
`
(2.243)
Similarly,




3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (

a
a a
a
a a
a
a a
n n n n n n 1
`
(2.244)
The antisymmetric tensors
0
, can also be expressed by means of the summation
symbol, (see Problem 1.34), as


3
1 ,
) ( ) (
0 0

b a
b a
b a
ab
N N


3
1 ,
) ( ) (

b a
b a
b a
ab
n n
(2.245)
Then, the equation in (2.240) can also be expressed as:
( )

- -
|
|
|
.
|

\
|
-
|
|
|
.
|

\
|

3
1 ,
) ( ) (
0
3
1 ,
) ( ) (
0
3
1 ,
) ( ) (
3
1 ,
) ( ) (
0
3
1 ,
) ( ) (



b a
b a
b a
ab ab
b a
b a
b a
ab
b a
b a
b a
ab
b a
b a
b a
ab
b a
b a
b a
ab
N n N n N n
N N R R n n R


`
(2.246)
With the above, the term
T
R R
`
can also be shown as:
( ) ( )

-
|
|
|
.
|

\
|
-
3
1 ,
) ( ) (
0
3
1 ,
) ( ) (
0

b a
b a
b a
ab ab
T
b a
b a
b a
ab ab
T
n n R N n R R
`

(2.247)
Moreover, we can express U by starting from its spectral representation:
T
T
a
a a
a
a
T a a
a
a
a a
a
a
a a
a
0 0
0
3
1
) (
0
) (
0 0
3
1
0
) (
0
) (
0 0
3
1
) (
0 0
) (
0 0
3
1
) ( ) (




R R
R R
R R R R


|
|
.
|

\
|
/
/ / /


U
N N
N N N N N N U

(2.248)
where we have introduced the tensor U , which is in the configuration
0
B , (see Figure
2.25), and U is given by:


/ /
3
1
) (
0
) (
0

3
1
) (
0
) (
0


a
a a
a
change of rate
a
a a
a
N N U N N U
`
`
(2.249)
2 CONTINUUM KINEMATICS

211
Then, the material time derivative of
T
0 0
R R U U becomes:

/ - -
|
|
.
|

\
|
/ - -
|
|
.
|

\
|
/ - -
- -
- -
- -






3
1
) ( ) (
0 0
3
1
0
) (
0
) (
0 0 0 0
0
3
1
) (
0
) (
0 0 0 0
0 0 0 0
0 0 0 0 0 0 0 0
0 0 0 0 0 0






a
a a
a
a
T a a
a
T
a
a a
a
T
T T
T T T T
T T T
N N U U
N N U U
N N U U
U U U
U U U
U U U U
`
`
`
`
`
`
` ` `





R R
R R
R R
R R R R R R
R R R R R R

(2.250)
In Problem 1.34 we demonstrated that:

/ - / -
/ - / -


3
1 ,
) ( ) ( 2 2
0 0
2 2
0
3
1 ,
) ( ) (
0 0 0

) (

) (
b a
b a
b a
a b ab
b a
b a
b a
a b ab
N N U U
N N U U



(2.251)
and

/ - / -
3
1 ,
) ( ) ( 2 2 2 2

) (
b a
b a
b a
a b ab
n n V V
(2.252)
Then, the equation in (2.250) can be rewritten as:


/ - / - /
3
1 ,
) ( ) (
0
3
1
) ( ) (

) (

b a
b a
b a
a b ab
a
a a
a
N N N N U
` `

(2.253)
Moreover, the left stretch tensor
T
R U R V can also be expressed as:
T
T T
T
R R
R R


U
R U R
R U R V
0 0

(2.254)
and, the material time derivative of
T
R R U V becomes:

/ - -
-
|
|
.
|

\
|
/ - - -
- -
- -




3
1
) ( ) (
3
1
) (
0
) (
0




a
a a
a
T T
a
a a
a
T T
T T T T
T T T
n n V V
U N N V V U V
U U U
U U U V
`
`
`
`
`
` ` `



R R R R
R R R R R R
R R R R R R

(2.255)
or
NOTES ON CONTINUUM MECHANICS

212


/ - / - /
3
1 ,
) ( ) (
3
1
) ( ) (

) (

b a
b a
b a
a b ab
a
a a
a
n n n n V
` `

(2.256)
The right Cauchy-Green deformation tensor ( C ) is also expressed as:
( ) ( ) ( )
T T T T
0
2
0 0 0 0 0
2
0 0
2
R R R R R R R R U U U U U C
(2.257)
and its material time derivative is given by:

/ / - -
|
|
.
|

\
|
/ / - -
- -
- -
- -





3
1
) ( ) (
0 0
0
3
1
) (
0
) (
0 0 0 0
0 0 0 0
0 0
2
0 0 0 0
2
0 0
0
2
0 0 0 0
2
0


2


2
2
2
2
a
a a
a a
T
a
a a
a a
T
T T
T T T T
T T T
N N
N N
U U
U U U U
U U U U
`
`
`
`
`
` ` `




C C
C C
C C
C
R R
R R
R R R R R R
R R R R R R

(2.258)
or


/ - / - / /
3
1 ,
) ( ) ( 2 2
0
3
1
) ( ) (

) (

2
b a
b a
b a
a b ab
a
a a
a a
N N N N
` `
C
(2.259)
Similarly, it is possible to define the left Cauchy-Green deformation tensor as:
( ) ( ) ( )
T T T T
R R R R R R R R
2
2
2
U U U U V b
(2.260)
and, its material time derivative as:

/ / - -
3
1
) ( ) (


2
a
a a
a a
n n
` `
b b b (2.261)
By fixing - - b b
2 2
V V , and by referring to (2.252) we obtain:


/ - / - / /
3
1 ,
) ( ) ( 2 2
3
1
) ( ) (

) (

2
b a
b a
b a
a b ab
a
a a
a a
n n n n
` `
b
(2.262)
The material time derivative of the deformation gradient,

/
3
1
) ( ) (

a
a a
a
N n F , becomes:
( )




- - / - - /
|
|
.
|

\
|
/ -
|
|
.
|

\
|
/ - /
/ - / - /
|
.
|

\
|
/ - / - /
3
1
0
) ( ) (
3
1
0
) ( ) (
3
1
0
3
1
) ( ) (
3
1
) ( ) ( ) ( ) (
3
1
) (
0
) ( ) ( ) ( ) ( ) (
3
1
) ( ) ( ) ( ) ( ) ( ) (

a
a a
a
a
T a a
a
a
T
a
a a
a
a
a a
a
a a
a
a
a a
a
a a
a
a a
a
a
a a
a
a a
a
a a
a



F F F F
F
N n N n
N n N n N n
N n N n N n
N n N n N n
` `
`
`
`
`
` `
(2.263)
2 CONTINUUM KINEMATICS

213
Using the same reasoning we made to solve Problem 1.34, we can state that:

/
|
|
.
|

\
|
/
|
|
|
.
|

\
|

3
1 ,
) ( ) (
3
1
) ( ) (
3
1 ,
) ( ) (


b a
b a
b a
b ab
b
b b
b
b a
b a
b a
ab
N n N n n n F
(2.264)


/
|
|
|
.
|

\
|

|
|
.
|

\
|
/
3
1 ,
) ( ) (
0
3
1 ,
) ( ) (
0
3
1
) ( ) (
0

b a
b a
b a
b ab
b a
b a
b a
ab
a
a a
a
N n N N N n F
(2.265)
So, the equation in (2.263) can also be written as:
( )

/ - / - /
3
1
3
1 ,
) ( ) (
0
) ( ) (

a
b a
b a
b a
ab a ab b
a a
a
N n N n
` `
F
(2.266)
Then, the spatial velocity gradient
1 -
F F
`
l can also be shown as:
( )
|
|
.
|

\
|

/
|
|
|
.
|

\
|
/ - / -
-
|
|
.
|

\
|

/
|
|
.
|

\
|
/

3
1
) ( ) (
3
1 ,
) ( ) (
0
3
1
) ( ) (
3
1
) ( ) (

b
b b
b
b a
b a
b a
ab a ab b
a
a a
a a
a a
a
n N N n
n N N n

`
l
(2.267)
which becomes:



|
|
.
|

\
|
/
/
- -
/
/

3
1 ,
) ( ) (
0
3
1
) ( ) (

b a
b a
b a
ab
b
a
ab
a
a a
a
a
n n n n
`
l

(2.268)
By referring to
1 1
2
1
- - - -
F C F F E F
` `
T T
D , (see equation (2.134)), and using the
expression of C
`
given in (2.259), we can also verify that:
|
|
.
|

\
|

/
|
|
|
.
|

\
|
/ - /
|
|
.
|

\
|

/
-
|
|
.
|

\
|

/
|
|
.
|

\
|
/ /
|
|
.
|

\
|





3
1
) ( ) (
3
1 ,
) ( ) ( 2 2
0
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (

1

) (

1
2
1

1


2

1
2
1
b
b b
b
b a
b a
b a
a b ab
a
a a
a
a
a a
a a
a a
a a
a
a a
a
n N N N N n
n N N N N n D

`
(2.269)
which becomes:



/ /
/ - /
-
/
/

3
1 ,
) ( ) (
2 2
0
3
1
) ( ) (

2
) (



b a
b a
b a
b a
a b
ab
a
a a
a
a
n n n n D
`

(2.270)
By referring to D W - l , and the equation in (2.268) and (2.270) we obtain:


|
|
.
|

\
|
/ /
/ - /
-
3
1 ,
) ( ) (
2 2
0


2
b a
b a
b a
b a
a b
ab ab
n n W

(2.271)

NOTES ON CONTINUUM MECHANICS

214
Problem 2.16: A rigid body motion is characterized by the following equation:
X x
,
,
,
- ) ( ) ( t t Q c (2.272)
Find the velocity and the acceleration fields as a function of c
,
, where c
,
is the axial vector
associated with the antisymmetric tensor (
T
Q Q
`
).

Solution:
The material time derivative of X x
,
,
,
- ) ( ) ( t t Q c is given by
X x x v
,
`
`
,
`
, , ,
- = Q c
Dt
D

Let us consider that Q Q Q Q
` `
T
. The above equation can also be expressed as:
) ( c c
Q c
,
,
`
,
,
,
`
,
,
- -
-


x v
X v


If is an antisymmetric tensor, it holds that a a
,
,
,
r c , where c
,
(angular velocity vector)
is the axial vector associated with the antisymmetric tensor , (see equation (2.88)). Then,
the associated velocity can be expressed as:
) (
) (
c c
c c
,
, ,
`
,
,
,
`
,
,
- r -
- -
x
x v
c

(2.273)
Note that ) (t Q is only dependent on time, hence the axial vector (angular velocity)
associated with is also time-dependent, i.e. ) (t c c
, ,
.
Then, its acceleration is given by:
X x v a
,
` `
` `
,
` `
,
`
, ,
- Q c
By referring to Q Q Q
` ` ` `
- , the above equation can also be expressed as:
) ( ) (
) (
c c c
Q Q c
Q Q c
Q Q c
,
,
,
,
`
` `
,
, ,
`
` `
,
,
`
,
`
` `
,
,
` `
` `
,
,
- - - -
- -
- -
- -




x x
X X
X X
X a





Then by using the property in (2.88) again we can state that:
[ ] ) ( ) ( c c c
,
, , ,
,
,
`
,
` `
,
,
- r r - - r - x x a c c c

(2.274)
where c
`
, ,
= shows the angular acceleration.
For a rigid body motion where 0 c
,
,
, the velocity becomes x v
, , ,
r c whose components
are
q p ipq i
x v c , and the rate-of-deformation tensor D becomes:
( ) ( ) ( )
ij p ipj p ipj p jpi p ipj qi p jpq qj p ipq
i
q
p jpq
j
q
p ipq
i
q p jpq
j
q p ipq
i
j
j
i
ij
x
x
x
x
x
x
x
x
x
v
x
v
0
D
c - c c - c c - c
|
|
.
|

\
|
o
o
c -
o
o
c
|
|
.
|

\
|
o
c o
-
o
c o

|
|
.
|

\
|
o
o
-
o
o




2
1
2
1
2
1
2
1
) ( ) (
2
1
2
1
c c

So, once again we have proved that 0 D for a rigid body motion.




2 CONTINUUM KINEMATICS

215
2.9 Area and Volume Elements Deformation
2.9.1 Area Element Deformation
Let us consider two line elements
) 1 (
X
,
d and
) 2 (
X
,
d in the reference configuration that
define the area element A
,
d , (see Figure 2.26). After motion, theses vectors are transformed
into
) 1 (
x
,
d and
) 2 (
x
,
d , thus defining the new area element a
,
d , (see Figure 2.26).














Figure 2.26: Area element deformation.
The area element A
,
d can be found using the definition of the vector product (the cross
product) used in Chapter 1, i.e.:
N N A X X X A

) (
) 2 ( ) 1 (
dA d d d PR PQ d r r

, , , , ,

(2.275)
where dA d = A
,
is the magnitude of A
,
d , and N

is the unit vector which is normal to the


area element, i.e. codirectional with A
,
d . In indicial notation, the cross product is
represented via the permutation symbol as:
) 2 ( ) 1 (
k j ijk i
dX dX dA (2.276)
The deformed area element (current configuration) is given by:
n n a x x a
) 2 ( ) 1 (
da d d d R P Q P d r r

, , , ,

(2.277)
where da d = a
,
is the magnitude of a
,
d , and n is the unit vector associated with the a
,
d -
direction. In indicial notation, the deformed area element can be expressed as:
) 2 ( ) 1 (
k j ijk i
dx dx a d (2.278)
From the equation in (2.278), we obtain:
A
,
d
a
,
d
A F a
,
,
d J d
T

-


1 1
, x X

2 2
, x X

3 3
, x X

3 3

,

e I

2 2

,

e I

1 1

,

e I
O
Reference
configuration - 0 t
Current configuration -
1
t t
a
,
d
P
Q ) 2 (
X
,
d

) 1 (
X
,
d
R
A
,
d
N



A
,
d


) 2 (
x
,
d
P
Q

) 1 (
x
,
d
R
n
a
,
d
NOTES ON CONTINUUM MECHANICS

216
A F
A F F
X X F
X F X F
x x x a
,
,
, ,
, ,
, , , ,
d J
d
d d
d d
d d t d
T
T


-
-

r
r
r

) ( ) (
) , (
) 2 ( ) 1 (
) 2 ( ) 1 (
) 2 ( ) 1 (
cof

( )
) 2 ( ) 1 ( 1
) 2 ( ) 1 ( 1
) 2 ( ) 1 (
) 2 ( ) 1 (
) 2 ( ) 1 (
q p ti kq jp rt rjk
q p kq jp ti rt rjk
q p kq jp ri rjk
q kq p jp ijk
k j ijk i
dX dX F F F F
dX dX F F F F
dX dX F F
dX F dX F
dx dx a d
-
-

c

(2.279)
In the above demonstration, by using tensor notation, we applied the tensor cofactor
definition, i.e. given a tensor T and two vector a
,
, b
,
, it holds that
b T a T b a T
,
,
,
,
r r ) ( ) ( cof . Then, [ ]
T
) (
1
1
T T T cof
-
-
, (see Chapter 1). We also proved
in the same chapter that
tpq kq jp rt rjk
F F F F , with which the equation in (2.279), in
indicial notation, can be expressed as:
t ti q p tpq ti q p ti kq jp rt rjk i
dA F J dX dX F dX dX F F F F a d
1 ) 2 ( ) 1 ( 1 ) 2 ( ) 1 ( 1

- - -
F
(2.280)
or in tensor notation as:
A F F A a
, ,
,
d J d J d
T

- -

1
Nansons formula (2.281)
which is known as Nansons formula, and can also be written in terms of N

and n as:
1

- -
F N N F n a dA J dA J da d
T
,
(2.282)
whose inverse relationship is given by:
a F F a N A
, ,
,
d
J
d
J
dA d
T

1 1

(2.283)
The magnitude of a
,
d is evaluated as follows:
1

-
F N a dA J d
,

(2.284)
Using equations (2.281) and (2.283), the magnitudes of da and dA are interrelated by:
N B N
A F F A
A F A F
a a
C B



2 2
1
1 2
2
2


-
- -
- -

dA J
d d J
d d J
d d da
T
T T
,
_
,
, ,
, ,

n b n
a F F a
F a F a
A A
b

1

1

1
2
2
2
2
2


da
J
d d
J
d d
J
d d dA
T
,
_
,
, ,
, ,

(2.285)
Then:
N B N

dA J da
n b n
1
da
J
dA (2.286)
Thus, it is also valid that:
) (
1

2 2
2
n b n
N B N

|
.
|

\
|
J J
dA
da
(2.287)
Taking into account the equation in (2.282) the expression of n is obtained as:
2 CONTINUUM KINEMATICS

217
N B N
N F
N B N
F N A F F A
n

1 1




- - - -

dA J
dA J
dA J
dA J
da
d J
da
d J
T T
, ,
(2.288)
or
N B N
N F
N B N
F N
n

- -

T
(2.289)
The material time derivative of n can be evaluated as follows:
( ) ( )
( )
( ) ( )
( ) N F F N
N B N
F N
N B N
F N
N B N
N B N
F N
N B N
F N
n


2
1

) (
1
2
3
1
2
1
1
2
3
1
2
1
1










- -
- -
- -
- -
-
T
Dt
D
`
`
`
D
l
(2.290)
or
( )
( )
( ) [ ] n n n
n n n n
n n n n n n
n n n n
n




) (
0




-
- -
|
|
.
|

\
|
- - -
- -

T
Dt
D
l l
l l
l l
l
1
W
D
_ (2.291)
2.9.1.1 The Material Time Derivative of the Area Element
Let us consider the undeformed and deformed area elements:
A F n a x a N A X A
,
, , ,
, , ,
d J d t d d t d
T
) , ( ;

) , (
-

(2.292)
The material time derivative of the deformed are element, a
,
d , is given by:
[ ]
[ ] [ ]
_
,
_
,
,
_
, , ,
,
,
, ,
,
,
a a
A F A F v
A F A F A F
A F a
x
d d
d J d J
d
Dt
D
J d
Dt
D
J d
Dt
DJ
d J
Dt
D
d
Dt
D
T T T
T T T
T
1
) (

) (
- - -
- - -
-



-
- -

l V
0

(2.293)
Then:
[ ] a
a a
a a v a
x
,
, ,
, , , ,
,
d
d d
d d d
Dt
D
T
T
T


-
-
-
l
l
l
) (
) (
) ( ) (
1 D
D
Tr
Tr
V

(2.294)
NOTES ON CONTINUUM MECHANICS

218
2.9.2 The Volume Element Deformation
Let us consider a parallelepiped formed by the line elements
) 1 (
X
,
d ,
) 2 (
X
,
d ,
) 3 (
X
,
d , (in the
reference configuration) whose volume is denoted by
0
dV . After motion, the vectors
) 1 (
X
,
d ,
) 2 (
X
,
d ,
) 3 (
X
,
d are transformed into the line elements
) 1 (
x
,
d ,
) 2 (
x
,
d ,
) 3 (
x
,
d , respectively,
and describe a new parallelepiped volume denoted by dV , (see Figure 2.27). Next, we can
establish the relationship between ) (
0
X
,
dV and ) , ( t dV x
,
.











Figure 2.27: Volume element deformation.
The parallelepiped volume, in the reference configuration, can be obtained by means of the
scalar triple product by including the three vectors
) 3 ( ) 2 ( ) 1 (
, , X X X
, , ,
d d d , i.e.:
( )
) 3 ( ) 2 ( ) 1 ( ) 3 ( ) 2 ( ) 1 (
0 k j i ijk
dX dX dX d d d dV r X X X
, , ,
(2.295)
Similarly, the deformed volume element can be evaluated as:
( )
) 3 ( ) 2 ( ) 1 ( ) 3 ( ) 2 ( ) 1 (
k j i ijk
dx dx dx d d d dV r x x x
, , ,
(2.296)
The element volume dV can also be expressed as:
( )
_
_
, , ,
0
) 3 ( ) 2 ( ) 1 (
) 3 ( ) 2 ( ) 1 (
) 3 ( ) 2 ( ) 1 (
) 3 ( ) 2 ( ) 1 (
) 3 ( ) 2 ( ) 1 (
dV
D C B BCD
D C B kD jC iB ijk
D kD C jC B iB ijk
k j i ijk
dX dX dX
dX dX dX F F F
dX F dX F dX F
dx dx dx
d d d dV
BCD

F
x x x
F

r

( )
( )
[ ]
[ ]
[ ]
0
) 3 ( ) 2 ( ) 1 (
) 3 ( 1 ) 2 ( ) 1 (
) 3 ( ) 2 ( ) 1 (
) 3 ( ) 2 ( ) 1 (
) 3 ( ) 2 ( ) 1 (
) 3 ( ) 2 ( ) 1 (
) (
) (
) (
) ( ) (
JdV
d d d J
d d d J
d d d J
d d d
d d d
d d d dV
T

r
r
r
r
r
r






-
-
X X X
X F F X X
X F X X F
X F X X F
X F X F X F
x x x
, , ,
, , ,
, , ,
, , ,
, , ,
, , ,
cof

(2.297)
Thus proving that the relationship between dV and
0
dV is given by:
0 0
dV J dV dV F Deformation of the volume element (2.298)


0
dV J dV

) 1 (
x d
S
P
Reference configuration - 0 t
Q
) 2 (
X d

) 1 (
X d
R

1 1
, x X

2 2
, x X

3 3
, x X

3 3

,

e I

2 2

,

e I

1 1

,

e I
O
A
,
d
N



0
V d
Current configuration - t
P
Q
R a
,
d
n
V d

) 2 (
x d
S

0
V d
V d

) 3 (
X d

) 3 (
x d
2 CONTINUUM KINEMATICS

219

Let ) (
0
X
,
j and ) , ( t x
,
j be mass densities in the reference and current configurations,
respectively. The differential mass in the reference configuration (
0
dm ) and in the current
configuration ( dm) are related by:
0 0 0
dV dm j ; dV dm j
(2.299)
Based on the principle of conservation of mass, we find that:
J
dV
dV
dV dV
dm dm

F
0
0
0 0
0

j
j
j j
(2.300)
) , ( ) (
0
t J x X
,
,
j j
(2.301)
Problem 2.17: Obtain an equation for mass density in terms of the third invariant of the
right Cauchy-Green deformation tensor, i.e. ( )
C
I I I
0 0
j j .
Solution:
From the equation in (2.301) we obtain:
J t) , ( ) (
0
x X
,
,
j j
and considering that the third invariant
2
) ( ) ( J I I I
T
F F C
C
det det , we obtain
C
I I I J , then:
C
I I I
0
j j (2.302)


2.9.2.1 The Material Time Derivative of the Volume Element
The material time derivative of dV is given by:
) ( ) ( ) ( ) (
0
0 0 0
_

- dV
Dt
D
J J
Dt
D
dV dV J
Dt
D
dV
Dt
D

(2.303)
Then we have:
dV
dV
dV J
dV
Dt
DJ
dV
Dt
D
) (


) (
0
0
D Tr


v
v
x
x
,
,
,
,
V
V

(2.304)
OBS.: If 1 J F , the volume is preserved during motion. If 1 > J , the volume
expands. If 1 0 < < J , the volume shrinks. If 0 s J , there is particle penetration,
thus violating the Axiom of impenetrability, i.e. without any physical meaning in
classical Mechanics.
NOTES ON CONTINUUM MECHANICS

220
2.9.2.2 Dilatation
Dilatation is the relative variation of the volume element, i.e.:
0
0
) 0 , (
) 0 , ( ) , (
) , (
dV
dV dV
dV
dV t dV
t D
V
-

X
X x
X ,
,
,
,
(2.305)
If we considering that
0 0
dV J dV dV F , the above equation becomes:
1 ) , (
0
0 0
0
0
-
-

-
J
dV
dV dV J
dV
dV dV
t D
V
X
,

(2.306)
2.9.2.3 Isochoric Motion. Incompressibility
If during motion the volume element remains unchanged this implies that the Jacobian
determinant field is unitary, since:
0 ) ( 1
0
0
dV
Dt
D
dV dV
dV
dV
J F
(2.307)
Then, if during motion the volume of every particle remains unchanged the motion is said
to be isochoric, i.e.:
) , ( ) (
0
t x X
,
,
j j (Isochoric motion) (2.308)
The continuous medium characterized by isochoric motion is said to be incompressible. An
incompressible medium can also be characterized by:


0 0 ) ( ) (
0
0
0
0 0

v v
x x
, ,
, ,
V V dV J dV
Dt
DJ
JdV
Dt
D
dV
Dt
D

(2.309)
where we have taken into account that 0
0
dV , 0 J . By using the equation in (2.102), it
is possible to express the incompressibility of the form:
0 ) ( ) ( ) (
,
D Tr Tr Tr l v v
x x
, ,
, ,
V V
k k
v (2.310)
NOTE: It is interesting to point out that incompressibility is an approximation. That is, all
continuous media are compressible, but depending on the material, such as liquids in
general, the compressibility can be insignificant.
2.10 Material and Control Domains
2.10.1 The Material Domain
The Material Curve
The material curve is a moving line that is always made up of the same particles.
The Material Surface
We can define the material surface, (see Figure 2.28), as a moving surface which is always
made up of the same particles. The material surfaces in the reference and current
configuration are represented, respectively, by:
2 CONTINUUM KINEMATICS

221
c t t C d d )) , ( ( ) , ( ; ) ( x X x X
,
,
,
,
c
(2.311)
By applying the chain rule of differentiation to the equation in (2.311) we obtain:
1
) , ( ) , ( ) , (
-
o
o

o
o
o
d o

o
o
ki
k i
k
k i
F
X
t
x
X
X
t
x
t x x x
, , ,
c c

) ( ) , ( X F x
X
x
,
,
, ,
d
-
V V
T
t c
(2.312)
The normal to the surface C d ) ( X
,
is given by:
) (
) (

X
X
N
X
X
,
,
,
,
d
d

V
V

(2.313)
and the derivative of C with respect to N

is given by:
) (

X
N
X
,
,
d V
d
dC

(2.314)
Then, the unit vector n can be expressed by:
) , (
) , (

t
t
x
x
n
x
x
,
,
,
,
c
c
V
V

(2.315)












Figure 2.28: Material surface.
The Material Volume
The material volume is a moving volume that is always made up of the same particles, (see
Figure 2.29).

2.10.2 The Control Domain
A spatial domain that is fixed in space is denoted by the control domain. Then, a control
surface is a fixed spatial surface and a control volume is a fixed spatial volume, (see Figure 2.29).



1
x

2
x

3
x
N


n
Reference configuration
Current configuration

0
S
S
NOTES ON CONTINUUM MECHANICS

222























Figure 2.29: Material volume and control volume.
2.11 Transport Equations
Now, we can establish how an intrinsic property of a material curve, material surface or material
volume, changes over time. A material curve is always made up of the same particles, so if
we consider that ) , ( t x
,
c is a property of each particle belonging to this material curve as
well as being a continuous and differentiable field, we obtain:
( ) ( ) [ ]
} } } }
- - |
.
|

\
|
-
C C C C
x x x x x x
, , , , , ,
` ` `
d d d d
Dt
D
d d
Dt
D
l l c c c c c c c 1 ) (
(2.316)
Likewise, for a material surface we have:

P
x

P
x
Control volume
Material volume

P
X

*
X

0 t


1
t


2
t


*
x

*
x
Material volume

0
v
) , (
1
*
t v x
,

) , (
2
*
t v x
,

Material volume
Control surface
Control volume
Control surface
Control volume
Control surface
2 CONTINUUM KINEMATICS

223
( ) ( )
( ) [ ]
}
} } }

- -
- - |
.
|

\
|
-
S
T
S
T
S S
d
d d d d
Dt
D
d d
Dt
D
a
a a a a a a
,
, , , , , ,
`
` `
l
l
c c c
c c c c c
1 D 1
D
) (
) ( ) (


Tr
Tr

(2.317)
Similarly, for a material volume we have:
}
}
} }
(

-
(

-
(

V
V
V V
dV t t
Dt
D
dV t t
Dt
D
dV
dV
Dt
D
t t
Dt
D
dV dV t
Dt
D
) , ( ) , (
) , ( ) , (
) ( ) , ( ) , ( ) , (
v x x
v x x
x x x
x
x
, , ,
, , ,
, , ,
,
,
V
V
c c
c c
c c c
(2.318)
These equations are known as transport equations:
a) Material curve
( )
} }
-
C C
x x
, ,
`
d d
Dt
D
l c c c 1

b) Material surface
( )
} }
- -
S
T
S
d d
Dt
D
a a
, ,
`
l c c c c 1 D 1 ) ( Tr

Transport
Equations
c) Material volume
} } (

-
V V
dV t t
Dt
D
dV t
Dt
D
) , ( ) , ( ) , ( v x x x
x
, , , ,
,
V c c c

(2.319)
Next, we can apply the material time derivative to the equation in (2.319)(c) to obtain:
[ ]
} }
} }
}
} }

-
(

o
o

-
o
o
-
(

o
o

-
o
o
-
o
o

-
V V
V V
V
V V
dV dV t
t
dV t
t
dV t
t
dV t t
t
t
t
dV t t
Dt
D
dV t
Dt
D
) ( ) , (
) , (
) , (
) , (
) , ( ) , (
) , (
) , (
) , ( ) , ( ) , (
v x
v x v
x
x
x
v x x v
x
x
x
v x x x
x
x
x
x
, ,
, , ,
,
,
,
, , , ,
,
,
,
, , , ,
,
,
,
,
c c
c
c
c
c
c
c
c c c
V
V
V
V

(2.320)
Now we apply the divergence theorem to the second integral on the right side of the
equation and we obtain:


,
_
,
,
S surface the
acrossing of flux
of flux
local
) (
) , (
) , (
c
c
c
c
c
} } }
-
o
o

S V V
dS dV
t
t
dV t
Dt
D
n v
x
x
(2.321)
As we can verify the volume integral on the right of the equation is a control volume and
the surface integral is a control surface. The term ) ( v
,
c , (as discussed in Chapter 5),
represents the property flux ( c ) that crosses the control surface. We can also see that only
the normal component of the flux (
n
q
,
) crosses the surface, whereas the tangential
NOTES ON CONTINUUM MECHANICS

224
component remains on the surface. Moreover, when the property is mass density, the
equation in (2.320) is known as the mass continuity equation, which is also discussed in
Chapter 5.









Figure 2.30: Surface and volume control.
2.12 Circulation and Vorticity
Let us consider a circuit 1 (closed curve) where the Eulerian velocity line integral around
the circuit 1 is given by:
}

1
1 x x v
, , ,
d t) , ( ) ( C
Circulation around 1 (2.322)
The line integral (2.322) is denoted by the circulation around 1 , (Chadwick (1976)) and by
using the Stokes Theorem we can obtain:
}
} }

r
!
! 1
!
! 1
d
d d t
n
n v x x v
x

) ( ) , ( ) (

,
,
,
, , ,
,
V C

(2.323)
where
,
is the vorticity vector defined in (2.87) for which the physical interpretation in
(2.323) follows. Let us consider the region !, which consists of molecules subject to
circulatory motion, (see Figure 2.31). The equation in (2.323) ensures that the vorticity
contribution of all molecules in the region ! is equal to the total circulation of the circuit
1 (the boundary of !). We can also verify that if the circulation around any closed curve
is zero, it holds that 0
,
,
,
,
r v
x
V . If this is the case, then the flow is said to be irrotational.
The rate of change of circulation around 1 can be obtained directly from (2.319)(b) by the
following change of variables ) , ( t x
, ,
c ,
,
,
d d a , so, we obtain:
( ) [ ]
} }
- -
! !
1
,
, ,
`
,
,
, ,
d d t
Dt
D
Dt
D
T
l ) ( ) , ( ) ( D Tr x C
(2.324)
The motion is said to be circulation preserving if and only if 0 D
,
, ,
`
,
- -
T
l ) ( Tr , where
the following is satisfied:
n
) ( v
,
c

t
t
o
o ) , ( x
,
c

x
,

V
S
[ ]n n v ) (
,
,
c
n
q
control surface
control volume
2 CONTINUUM KINEMATICS

225
( ) 0
0
0
0
0 D
,
,
,
,
`
,
`
`
,
,
,
`
`
,
`
,
,
,
`
`
,
`
,
,
, ,
`
, , ,
`
,

- -
- -
- -
- - - -





-
- - -
- - - -
-
1
1 1 1
1 1 1 1
1
) ( ) ( ) (
) ( ) (
F
F F F
F F F F F
F F
J
Dt
D
J J J
J
J
J
J J
J
J
T
l l l Tr Tr

(2.325)
where we have used ) (l Tr J J
`
,
1 -
- F F
`
l . From the equation in (2.325) we can
conclude that circulation which preserves the particle vorticity can be expressed as:
0
1
0
1

, , , ,

-
F F
J
J Cauchys vorticity formula (2.326)
which is known as Cauchys vorticity formula, and links the vorticity between the reference
and current configurations with the circulation preserving motion, (Chadwick (1976)).















Figure 2.31: Circulation and vorticity.
2.13 Motion Decomposition: Volumetric and
Isochoric Motions
Sometimes, when establishing certain constitutive equations, it may be convenient (at the
time of the numerical implementation) to separate the motion into a volumetric and an
isochoric part. Then, we can introduce the multiplicative decomposition of the
deformation gradient as:

,


1
x

3
x
2
x
!
1
n

,
d
) , ( t x v
, ,

n
NOTES ON CONTINUUM MECHANICS

226
vol vol iso
F F F F F
~

(2.327)
where
iso
F F =
~
is the transformation characterized by a volume-preserving (isochoric)
transformation, and
vol
F characterizes a volume-changing (dilatational) transformation,
(see Figure 2.32), where:
1
3
1
3
1
;
~
J J
vol

-
F F F
(2.328)
According to Simo&Hughes (1998) this decomposition was originally introduced by Flory
in 1961.














Figure 2.32: Multiplicative decomposition of the deformation gradient into a volumetric
and an isochoric part.

If we consider the right Cauchy-Green deformation tensor, F F C
T
, the isochoric part
of C is represented by C
~
, and the volumetric part by
vol
C . There are given respectively
by:
C
F F
F F
F F C
C
3
2
3
1
3
1
3
1
3
1
~ ~ ~
-
- -
- -

|
|
.
|

\
|
|
|
.
|

\
|

J
J J
J J
T
T
T
_

1
1 1
1 1
1
3
2
3
1
3
1
3
1
3
1
J
J J
J J
T
T
vol
T
vol vol

|
|
.
|

\
|
|
|
.
|

\
|

_
F F C

(2.329)
Similarly, the isochoric part of the left Cauchy-Green deformation tensor
T
F F b is
expressed as:
b b b b b b
~ ~
) (
~
3
2
3
2

-
vol
J J 1
(2.330)
B
X
,

x
,

X
,


vol
F F
~


vol
F F F
~

reference
configuration
current
configuration

0
B
B
pure dilatation
F F
~
3
1
J
2 CONTINUUM KINEMATICS

227
where ) (
3
2
1 J
vol
b is the volumetric part of b .
We can now obtain the material time derivative of C C
3
2
~
-
J :
C C C
` `
`
3
2
3
5
3
2 ~
- -
- - J J J
(2.331)
Taking into account that ) (D Tr J J
`
and F F C D
T
2
`
, the equation in (2.331)
becomes:

F F
F F F F F F
F F C C C C
F F


|
.
|

\
|
- |
.
|

\
|
- -
- - - -
- -
-

- - -
dev T
T T T
T
J J
J J J J J J
T
D
1 D D D 1 D
D D
~
2
) (
3
1
2 ) (
3
1
2
2 ) (
3
2
3
2 ~
3
2
3
2
3
2
3
5
3
2
3
5
Tr Tr
Tr
` `
`

(2.332)
where it holds that
dev dev sph
D 1 D D D D - - ) (
3
1
Tr . In addition, it holds that:
[ ] [ ]
0
) ( 2 2 2 2
2 2
~
3
2
3
2
3
2
1 1
3
2
1 1
3
2
1
3
2
1


(
(

(
(

- - -
- -
-
- -
-
- -
-
-

dev dev
qq pq
dev
kp kq jq pj
dev
kp iq ki
jq iq pj
dev
kp ki
T dev T
J J J F F F F J
F F F F J J
D
D
Tr D D D
D
c c
F F F F C C : :
`

(2.333)
2.13.1 The Principal Invariants
The principal invariants of F
~
and
vol
F are given, respectively, by:
( ) ( )
[ ] ( ) [ ]


|
|
.
|

\
|

|
|
.
|

\
|

-
(
(

|
|
.
|

\
|
- -

|
|
.
|

\
|

-
- -
- - - -
- - -
1 ) ( )
~
(
2
1
2
1
)
~
(
2
1
~
1
3
3
1
3
1
~
3
2
2 2
3
2
2
3
2
2
3
2
2 2
~ ~
3
1
3
1
3
1
~
J J J J I I I
I I J I J J I J I I I
I J J J I
F F F
F F F
F F F
F
F F F
F F
F
F
det det det
Tr Tr Tr
Tr Tr Tr
(2.334)
and
( )

|
|
.
|

\
|

(
(

-
(

|
|
.
|

\
|

J J I I I
J J J I I I
J J I
vol
vol
vol
vol
vol vol
vol
1
1
3
1
3
2
3
2
3
2
2
2
3
1
3
1
) (
3 3 9
2
1
) (
2
1
3
det det
Tr
Tr Tr
F
F
F
F
F F
F

(2.335)
NOTES ON CONTINUUM MECHANICS

228
If we consider that the Jacobian determinant can be expressed as
C
I I I J , the principal
invariants of C
~
are:
( )
3
3
2
3
2
~
~
C
C
C
C
C C
I I I
I
I J J I
|
|
.
|

\
|

- -
Tr Tr
(2.336)
[ ]
3
2
3
4
2
3
4
2
3
4
2 2
~ ~
2
1
)
~
(
2
1
C
C
C C
C C
C C
I I I
I I
I I J J I J I I I
(
(

|
|
.
|

\
|
- -
- - -
Tr Tr
(2.337)
1 ) ( )
~
(
2 2
3
3
2
3
2
~
|
|
.
|

\
|

|
|
.
|

\
|

-
- -
J J J J I I I C C C
C
det det det (2.338)
Likewise, we obtain:
1 ; ; ~
3
2
~
3
~
b
b
b
b
b
b
b
I I I
I I I
I I
I I
I I I
I
I
(2.339)
Taking into account that
C b
I I ,
C b
I I I I ,
C b
I I I I I I , we can also conclude that:
C b C b C b
~ ~ ~ ~ ~ ~ ; ; I I I I I I I I I I I I
(2.340)
2.14 The Small Deformation Regime
2.14.1 Introduction
Given a quadratic ( c bx ax y
C
- -
2
) and a linear ( c bx y
L
- ) function, we can ask the
following question: In which situation are these two functions approximately the same? To
answer this, let us consider the following numerical values for the constants 2 a , 1 b ,
0 c . We can verify that for very small values of x , these functions are very close to each
other,
L C
y y x = <<1 , (see Figure 2.33). That is, if x is very small compared with unity,
the quadratic or higher terms can be discarded.







Figure 2.33: Linear and quadratic functions.
We can extend this reasoning to tensors. For example, let us consider the material strain
tensor ) (
2
1
J J J J - -
T T
E . If the material displacement gradient components
ij
J are
x
x y
L

x x y
C
-
2
2
y
1 x

L C
y y x = <<1
2 CONTINUUM KINEMATICS

229
much smaller than the unity, i.e. 1 <<
ij
J , the quadratic terms ) ( J J
T
are even smaller and
can be discarded.
There are many cases in engineering in which the displacement gradient components are
small compared with the unity:
1 <<
o
o

j
i
ij
X
u
J
(2.341)
The approximation in (2.341) takes place when the continuum (structure) is made up of
very rigid material and the loads (forces) to which theses structures are subjected produce
small displacements. In such cases, motion can be approximated by the infinitesimal strain
theory, also known as the small deformation theory, or the small displacement theory.
2.14.2 Infinitesimal Strain and Spin Tensors
As previously discussed, if the displacement gradient is already small, higher-order terms
are even smaller. In this case, the Green-Lagrange strain tensor can be approached by
means of the linear Green-Lagrange strain tensor (
L
ij
E ), which is defined as:
|
|
.
|

\
|
o
o
-
o
o
=
|
|
.
|

\
|
o
o
o
o
-
o
o
-
o
o

i
j
j
i L
ij
n deformatio small
regime
j
k
i
k
i
j
j
i
ij
X X
E
X X X X
E
u
u u u
u
u
2
1
2
1

(2.342)
or in tensorial notation:
( ) [ ] ( )
[ ]
sym
X
sym
T T
X X X X
L
t t ) , ( ) , (
2
1
) (
2
1
2
1
X X
E
,
,
,
,
, , , ,
,
, , , ,
u u
u u u u
V V
V V V V
=
- - - J J
(2.343)
If we now consider the Almansi strain tensor in terms of the gradient of the displacements,
(see Eq. (2.128)), we can define the linear Almansi strain tensor,
L
ij
e , as:
|
|
.
|

\
|
o
o
-
o
o
=
|
|
.
|

\
|
o
o
o
o
-
o
o
-
o
o

i
j
j
i L
ij
n deformatio small
regime
j
k
i
k
i
j
j
i
ij
x x
e
x x x x
e
u
u u u
u
u
2
1
2
1

(2.344)
or in tensorial notation:
( ) [ ] ( )
[ ]
sym
x
sym
T T
x x x x
L
t t ) , ( ) , (
2
1
) (
2
1
2
1
x x
e
, , , ,
, , , ,
,
, , , ,
u u
u u u u
V V
V V V V
=
- - - j j

(2.345)
NOTE: To verify that the material time derivative of the linear Almansi strain tensor
L
e` is
equal to the rate-of-deformation tensor:

ij j
i
i
j
j
i
i
j i
j
j
i
i
j
j
i L
ij
v
x
v
x
t x t x x t x t
x x t
e
D
u
u
u
u
u
u

|
|
.
|

\
|
o
o
-
o
o

|
|
.
|

\
|
o
o
o
o
-
o
o
o
o

|
|
.
|

\
|
o
o
o
o
-
o
o
o
o

|
|
.
|

\
|
o
o
-
o
o
o
o

2
1
2
1
2
1
2
1
`

(2.346)
NOTES ON CONTINUUM MECHANICS

230
If both the displacement gradient and the displacement are small, it means there is very
little difference between the spatial and material configurations, so the linear strain tensors
can be considered equal, i.e.:
( ) u
, ,
,
, ,
,
sym T L L
t t t V - = = = J J
2
1
) , ( ) , ( ) , ( x X x e x X E (2.347)
which defines the infinitesimal strain tensor:
j i ij
sym
t e e u

) , ( r
, ,
V x
The infinitesimal strain tensor (2.348)
Explicitly, the components of are given by:
(
(
(
(
(
(
(
(

o
o
|
|
.
|

\
|
o
o
-
o
o
|
|
.
|

\
|
o
o
-
o
o
|
|
.
|

\
|
o
o
-
o
o
o
o
|
|
.
|

\
|
o
o
-
o
o
|
|
.
|

\
|
o
o
-
o
o
|
|
.
|

\
|
o
o
-
o
o
o
o

(
(
(

r r r
r r r
r r r
r
|
|
.
|

\
|
o
o
-
o
o
r
3
3
2
3
3
2
1
3
3
1
2
3
3
2
2
2
1
2
2
1
1
3
3
1
1
2
2
1
1
1
33 23 13
23 22 12
13 12 11
2
1
2
1
2
1
2
1
2
1
2
1
2
1
x x x x x
x x x x x
x x x x x
x x
ij
i
j
j
i
ij
u u u u u
u u u u u
u u u u u
u
u

(2.349)
The displacement gradient u
,
V can be split into a symmetric and antisymmetric part as:
[ ] [ ] u u u u u u u - - - - -
, , , , , , ,
skew sym T T
V V V V V V V ) (
2
1
) (
2
1
(2.350)
where the symmetric part is the infinitesimal strain tensor, and the antisymmetric part is
known as the infinitesimal spin tensor, which is also called the infinitesimal rotation tensor. For
rigid body motion the condition that strain tensor is zero, i.e. 0 must be satisfied and
for motion characterized only by strain 0 must be true.
Notice that, the tensor does not accurately measure strain, since it is affected by rigid
body motion. To illustrate this, let us consider that a material body is subjected to a
rotation as indicated in Figure 2.34. In this situation, the equations of motion are given by:

0 - 0
0 - 0

(
(
(

0 0
0 - 0

3 3
2 1 2
2 1 1
3
2
1
3
2
1
cos sin
sin cos
1 0 0
0 cos sin
0 sin cos
X x
X X x
X X x
X
X
X
x
x
x
(2.351)








Figure 2.34: A material body subjected to a rotation.

1
X

1 1
, x X

*
2
x

*
1
x
0

2
X

2 2
, x X
2 CONTINUUM KINEMATICS

231
The displacement field,
i i i
X x - u , can be obtained as follows:

-
- 0 - 0 -
0 - - 0 -
0
) 1 (cos sin
sin ) 1 (cos
3 3 3
2 1 2 2 2
2 1 1 1 1
X x
X X X x
X X X x
u
u
u
(2.352)
If we consider the equation in (2.349) we can obtain the components of the infinitesimal
strain tensor as:
(
(
(

- 0
- 0

|
|
.
|

\
|
o
o
-
o
o
r
0 0 0
0 ) 1 (cos 0
0 0 ) 1 (cos
2
1
i
j
j
i
ij
X X
u
u

(2.353)
As we can see
11
r and
22
r are not equal to zero, but with small rotations it is true that
1 cos = 0 , thus the terms
11
r and
22
r are insignificant. As discussed in the chapter on linear
elasticity, the small deformation approximation is widely used for various engineering
problems and such problems are subjected to small displacements and small rotations. Note
that, for the rigid body motion described in Figure 2.34, the Green-Lagrange strain tensor
components are equal to zero. For instance, the component
11
E works out as:
( ) [ ] 0 sin 1 cos
2
1
) 1 (cos
2
1
2 2
2
1
3
2
1
2
2
1
1
1
1
11
0 - - 0 - - 0
(
(

|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o

X X X X
E
u u u u
(2.354)
2.14.3 Stretch and Unit Extension
Let us consider the relationship between the stretch and the unit extension in terms of the
Green-Lagrange strain tensor E , (see equation (2.158)). Then, if we are dealing with a
small strain regime ( = E ) the stretch and the unity extension become:

- - - /
- /


1

2 1 1

2 1

M M
M M
m m
m


(2.355)
Remainder: Taking into account the binomial series we have:
( )
( )

- - - -
-
-
- - -
- -
2
2
1
2 2 1
8
1
2
1
1 1
! 2
) 1 (
x x x
x a
n n
x na a x a
n n n n

(2.356)
In the event that x is very small, we can discard higher order terms, i.e.:
( ) x x
2
1
1 1 2
1
- = - (2.357)
Taking into account the Remainder above, the stretch and the unit extension, in a small
strain regime, are represented by:

r = - - - /
- = - /


) (


1

2 1 1

1

2 1
m
m m
m
M M M M
M M M M
N


(2.358)
which verifies that in a small strain regime the unit extension is equal to the normal
engineering strain.
NOTES ON CONTINUUM MECHANICS

232
2.14.4 Change of Angle
By applying the equation obtained in (2.161) for a small strain regime, = E , the angle in
the reference configuration becomes:
( )
( ) ( )
N N M M
N M
N N M M
N M
N N M M
N M

2 1

2 1

2

2 1

2 1


2

cos






- -
-
- -

- -
-
0



1 1
1
(2.359)
If we consider that 1

2 << M M , 1

2 << N N , we can conclude that:
N M N M
N M
N M
N M N M

2 cos

2



2

cos

- O - - 0
(2.360)
where O is the angle between M

and N

, and is defined in the initial configuration. The


equation in (2.360) could have been obtained directly by applying the equation in (2.164)
and by considering that:
N M
N E M
N M

2 cos


2 cos
cos



- O =
/ /
- O
0
(2.361)
Additionally, if we consider that O - 0 0 A is the angle variation, then the equation in
(2.360) can be rewritten as:
( ) N M

2 cos cos - O 0 A - O (2.362)
Moreover, the term on the left of the equation can be rewritten by the following
trigonometric relationship:
( ) O 0 A - O 0 A O - 0 A O 0 A - O
0 A = =
sin cos sin sin cos cos cos
1
_ _

(2.363)
in which we have considered that the angle 0 A is very small. If we substitute the previous
result into (2.362) we obtain:
N M
N M

2 sin

2 cos sin cos


- O 0 A
- O O 0 A - O


(2.364)
Then the angle change for the small strain regime is given by:
O
-
0 A

sin

2 N M

The angle change for a small strain regime (2.365)
2.14.5 The Physical Interpretation of the Infinitesimal Strain
Tensor
First, let us consider the components of the infinitesimal strain tensor :
(
(
(

r r r
r r r
r r r

(
(
(

r r r
r r r
r r r
r
zz yz xz
yz yy xy
xz xy xx
ij
33 23 13
23 22 12
13 12 11
(2.366)
2 CONTINUUM KINEMATICS

233
The stretch
x
/ and the unit extension
x
according to the x x =
1
-direction can be
evaluated by considering the equations (2.358) and
1

e M , thus:
xx x
xx x
r = - /
r - - = /
1
1

1
x
M M


(2.367)
Hence, we can conclude that the diagonal terms are related to the unit extension as follows:
z y x
= r = r = r
zz yy xx
; ;
(2.368)
Now, let us consider that
1

e M and
2

e N , (see Figure 2.35). In this particular case we


obtain:
[ ]
xy
zz yz xz
yz yy xy
xz xy xx
r
(
(
(

(
(
(

r r r
r r r
r r r

0
1
0
0 0 1

N M (2.369)









Figure 2.35: Angle change.
Then by using the equation in (2.365) and by considering (2.369) the angle variation
becomes:
xy xy
0 A
-
r
O
-
0 A

2
1
sin

2 N M

(2.370)
Afterwards we can interpret
xy
r as the angular distortion between two of the line elements,
whereas if we consider all three directions we obtain:
yz yz xz xz xy xy
0 A
-
r 0 A
-
r 0 A
-
r
2
1
;
2
1
;
2
1
(2.371)
2.14.5.1 Engineering Strain
Traditionally in engineering we have the following notation for the axes
1
x x = ,
2
x y = ,
3
x z = , for the displacement components
1
u = u ,
2
u = v ,
3
u = w .
Now, let us consider a segment AB whose end is displaced by u A as shown in Figure 2.36.
In one dimension, the strain is defined as:
( )
dx
du
x
u u u
im
x

A
- A -
r
A 0
(2.372)
x x X =
1 1
,
y x X =
2 2
,
P
Q
z x X =
3 3
,
R

) 1 (
X
,
d

) 2 (
X
,
d
M



(
(
(

(
(
(

0
0
1

0
0
1

i
i
N
M

NOTES ON CONTINUUM MECHANICS

234






Figure 2.36: Strain in one dimension.
Now let us consider a differential element dxdy and the displacement field, ) , ( y x u and
) , ( y x v , as shown in Figure 2.37. The normal strains according to the x -direction and y -
direction, respectively, are given by:
y
v
dy
v dy
y
v
v
x
u
dx
u dx
x
u
u
y x
o
o

-
|
|
.
|

\
|
o
o
-
r
o
o

- |
.
|

\
|
o
o
-
r ;
(2.373)













Figure 2.37: Normal strain.
Similarly, if we take into account all three dimensions, the normal strain components are
given by:
z
z y x w
y
z y x v
x
z y x u
z y x
o
o
r
o
o
r
o
o
r
) , , (
;
) , , (
;
) , , (

(2.374)
To find the tangential strain (or the shear strain), let us consider that the differential
element is only distorted by the angle shown in Figure 2.38. For small angles it holds that
0 = 0 tan , then:
y
u
x
v
o
o
0 = 0
o
o
0 = 0
2 2 1 1
tan ; tan
(2.375)
So, the shear strain
xy
is defined as:
B
u
B A
A

i
t

1 - i
t
x A
u u A -
x
B
u
dx
x
u
u
o
o
-
O
O dx
u x,
v y,
B
A
A dy
v
dy
y
v
v
o
o
-
2 CONTINUUM KINEMATICS

235
y
u
x
v
xy
o
o
-
o
o
0 - 0
2 1

(2.376)
Similarly, if we consider the other dimensions, we can obtain:
z
u
x
w
y
w
z
v
y
u
x
v
xz yz xy
o
o
-
o
o

o
o
-
o
o

o
o
-
o
o
; ;
(2.377)
Note that
xy xy
- O - 0 0 A , and if we compare that with the equation in (2.371) we can
conclude that
xy xy
r 2 . Similarly, we can say that
yz yz
r 2 and
xz xz
r 2 , thus,
_
Notation g Engineerin
2
1
2
1
2
1
2
1
2
1
2
1
33 23 13
23 22 12
13 12 11
(
(
(

r
r
r

(
(
(

r r r
r r r
r r r

(
(
(

r r r
r r r
r r r
r
z yz xz
yz y xy
xz xy x
zz yz xz
yz yy xy
xz xy xx
ij

(2.378)
Then, the strain components in terms of displacement in engineering notation are given by:
(
(
(
(
(
(
(

o
o
|
|
.
|

\
|
o
o
-
o
o
|
.
|

\
|
o
o
-
o
o
|
|
.
|

\
|
o
o
-
o
o
o
o
|
|
.
|

\
|
o
o
-
o
o
|
.
|

\
|
o
o
-
o
o
|
|
.
|

\
|
o
o
-
o
o
o
o

(
(
(

r
r
r
r
z
w
y
w
z
v
x
w
z
u
y
w
z
v
y
v
x
v
y
u
x
w
z
u
x
v
y
u
x
u
z yz xz
yz y xy
xz xy x
ij
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1

(2.379)













Figure 2.38: Shear strain (small deformation regime).
2.14.6 The Volume Ratio (Dilatation)
Let us consider a cube defined by the line elements
1
dX ,
2
dX ,
3
dX , (see Figure 2.39).
The volume variation is given by:
B
dx
x
v
o
o


2
0
A

1
0
O
dx
u x,
v y,
B
A
dy
dy
y
u
o
o

O
0

2 1
0 - 0
xy

NOTES ON CONTINUUM MECHANICS

236
[ ]
3 2 1
3 2 1
3 2 1 3
3
2
2
1
1
3 2 1 3
3
2
2
1
1
0
1 ) 1 ( ) 1 ( ) 1 (
) 1 ( ) 1 ( ) 1 (
dX dX dX
dX dX dX dX dX dX
dX dX dX dX dX dX dV dV V
x x x
x x x
x x x
- - - -
- - - -
- / / / - A


(2.380)
In a small strain regime, 1 << r
ij
, the higher order terms can be discarded without there
being any significant change in the outcome. It is also true that with small strains unit
extensions are in keeping with the normal components of the strains
11
1
r
x
,
22
2
r
x
,
33
3
r
x
, (see equation in (2.368)), thus:
0 33 22 11 0
3 2 1
] [ ] [ dV dV V
x x x
r - r - r - - A
(2.381)
In this case, the dilatation (volume ratio) becomes:
3 2 1 33 22 11
0
) ( ) , ( r - r - r r - r - r
A
r =

Tr I
dV
V
t D
V
L
V
x
,

(2.382)
Additionally, if the continuum is incompressible, it is valid that:
0
33 22 11
0
r - r - r
A

I
dV
V

(2.383)











Figure 2.39: Dilatation.
2.14.7 The Plane Strain
When the strain tensor field is independent of any one direction we say that the continuum
represents a plane strain state. In general, the independent direction is adopted by the
3
x -
one. So, in this situation, the infinitesimal strain tensor components are given by:
) 2 , 1 , (
0 0 0
0
0
22 12
12 11 Strain Plane
22 12
12 11

r r
r r
r
(
(
(

r r
r r
r j i
ij ij
(2.384)
The displacement field for a plane strain state is only a function of
1
x and
2
x , i.e.:
) constant ( ; ) , ( ; ) , (
3 2 1 2 2 2 1 1 1
C x x x x u u u u u
(2.385)

3
dX

2
dX

1
dX

2
X

3
X

1
X
dV

0
dV

1 1
1
dx dX
x
/

3 3
3
dx dX
x
/

2 2
2
dx dX
x
/

i
x
/ - stretch

i
x
- unit extension

i
i
i
i
i
dX dX dx
x x
) 1 ( - /

2 CONTINUUM KINEMATICS

237
Problem 2.18: Consider a material body in a small deformation regime, which is subjected
to the following displacement field:
3
3 3
3
1 2 2
3
2 1 1
10 ; 10 ) 10 ( ; 10 ) 7 2 (
- - -
- - - - x x x x x u u u
a) Find the infinitesimal spin and strain tensor;
b) Find the principal invariants of the infinitesimal strain tensor, as well as the
correspondent characteristic equation;
c) Draw the Mohrs circle for strain, and obtain the maximum shear strain;
d) Find the dilatation and the deviatoric infinitesimal strain tensor.
Solution
a) For the displacement gradient we obtain:
( )
(

(
(
(

- -
-

(
(
(
(
(
(
(

o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o

o
o

-
m
m
x x x
x x x
x x x
x
j
i
ij
3
3
3
2
3
1
3
3
2
2
2
1
2
3
1
2
1
1
1
10
1 0 0
0 10 1
0 7 2
u u u
u u u
u u u
u
u
,
V

In the International System of Units the displacement gradient is dimensionless, i.e.
[ ]
m
m

o
o

x
,
,
,
u
u V .
As for the infinitesimal spin tensor we obtain:
( )
3
10
0 0 0
0 0 4
0 4 0
2
1
-

(
(
(

-
|
|
.
|

\
|
o
o
-
o
o
c
i
j
j
i
ij
skew
ij
x x
u
u
u
,
V
Then for the infinitesimal strain tensor we have:
( )
3
10
1 0 0
0 10 3
0 3 2
2
1
-

(
(
(

-
-

|
|
.
|

\
|
o
o
-
o
o
r
i
j
j
i
ij
sym
ij
x x
u
u
u
,
V
b) The principal invariants are defined as ) (

Tr I , [ ] { ) ( ) (
2
1
2 2

Tr Tr - I I ,
) (

det I I I , (see Chapter 1). Then, it follows that:


[ ] {
9
6 6 2 2
3 3
10 11 ) (
10 1 10
1 0 0
0 10 3
0 3 2
1 0 0
0 10 3
0 3 2
1 0 0
0 10 3
0 3 2
) ( ) (
2
1
10 11 10 ) 1 10 2 ( ) (
-
- -
- -

-
|
|
|
.
|

\
|
-
-
- -
-
- -
-
-
- - - -

det
Tr Tr
Tr
I I I
I I
I

Then, the characteristic determinant is:
0 10
1 0 0
0 10 3
0 3 2
3

(
(
(

r -
r - -
r - -
-

whilst the characteristic equation is:
0 10 11 10 11 10 11 0
9 6 2 3 3 2 3
- r - r - r - r - r - r
- - -

I I I I I I
c) To draw the Mohrs circle for strain, (see Appendix A), we need to evaluate the
eigenvalues of . But, if we take a look at the components of we can verify that 1 r is
NOTES ON CONTINUUM MECHANICS

238
already an eigenvalue associated with the direction [ ] 1 0 0


i
n . So, to obtain the
remaining eigenvalues one only need solve the following system:

- r
- r
- r - r
(

r - -
r - -
-
-
- - -
3
2
3
1 6 3 2 3
10 0 . 11
10 0 . 1
0 10 11 10 12 0 10
10 3
3 2

Then by restructuring the eigenvalues such that
I I I I I I
r > r > r , we obtain:
3 3 3
10 0 . 11 ; 10 0 . 1 ; 10 0 . 1
- - -
- r - r r
I I I I I I

Then the maximum shear (tangential) strain is evaluated as follows:
3
max
10 6
2
-

r - r
r
I I I I
S

Finally, the Mohrs circle for strain can be depicted as:















d) The volumetric strain (dilatation) -
V
r is:
3
10 12 ) (
-
- r

Tr I
V

The additive decomposition of into a spherical and a deviatoric part is denoted by
dev sph
- , where the spherical part is given by:
3
10
4 0 0
0 4 0
0 0 4
3
) (
-

(
(
(

-
-
-
r
ij
sph
ij
c
Tr

And, the deviatoric part is given by:
3 3
10
4 0 0
0 6 3
0 3 2
10
4 0 0
0 4 0
0 0 4
0 0 0
0 10 3
0 3 2
- -

(
(
(

-
|
|
|
.
|

\
|
(
(
(

-
-
-
-
(
(
(

-
-
r - r r
sph
ij ij
dev
ij


1 r
I
) 10 (
3 -
r
N

) 10 (
3 -
r
S

6
max
2
1
max
r
S

11 - r
I I I

1 - r
I I

2 CONTINUUM KINEMATICS

239
2.15 Other Ways to Define Strain
2.15.1 Motivation
In this section we will define other measures of strain that may be useful in addressing
the problem incrementally. We shall see that the strain tensors defined previously cannot
be obtained by adding incremental strains caused by successive motions.
Let us consider a continuum which is subjected to successive configurations, (see Figure
2.40).













Figure 2.40: Motion defined by successive configurations.
According to Figure 2.40 the following conditions are satisfied:
x F x X F x X F x
,
,
, , ,
,
d d d d d d
) 2 ( ) 1 (
; ; (2.386)
By substituting the x
,
d given by the second expression into the third one, we obtain:
( ) X F F X F F x F x
, , ,
,
d d d d
) 1 ( ) 2 ( ) 1 ( ) 2 ( ) 2 (
(2.387)
And by comparing (2.387) with X F x
,
,
d d , we can conclude that:
) 1 ( ) 2 (
F F F (2.388)
The Green-Lagrange strain tensor E , (reference configuration
0
B ) is defined as
( ) 1 - F F E
T
2
1
. We defined new tensors in a similar fashion because of the
transformations
) 1 (
F and
) 2 (
F , i.e.:
Initial configuration
0
B :
Intermediate configuration B :
|
.
|

\
|
- 1
) 1 ( ) 1 ( ) 1 (
2
1
) , ( F F X E
T
t
,
|
.
|

\
|
- 1
) 2 ( ) 2 ( ) 2 (
2
1
) , ( F F x E
T
t
,

(2.389)
The Green-Lagrange strain tensor ( E ) can be written in terms of
) 1 (
F and
) 2 (
F as:
B

1 1
, x X

2 2
, x X

3 3
, x X

3

e

2

e

1

e
O
Initial configuration -
0
t
Intermediate configuration
1
t
x
,
d

) 1 (
F
B
Current configuration
2
t

) 2 (
F

) 1 ( ) 2 (
F F F
x
,
d
0
B

X
,
d
NOTES ON CONTINUUM MECHANICS

240
( ) ( ) ( )
|
.
|

\
|
-
(

- -


1
1 1
) 1 ( ) 2 ( ) 2 ( ) 1 (
) 1 ( ) 2 ( ) 1 ( ) 2 (
2
1
2
1
2
1
F F F F
F F F F F F E
T T
T
T

(2.390)
Taking into account the
) 2 (
E equation given in (2.389), we obtain 1 -
) 2 ( ) 2 ( ) 2 (
2E F F
T
,
and by substituting this into (2.390) we obtain:
( )
) 1 ( ) 2 ( ) 1 ( ) 1 (
) 1 (
) 1 ( ) 1 ( ) 1 ( ) 2 ( ) 1 (
) 1 ( ) 2 ( ) 1 ( ) 1 ( ) 2 ( ) 2 ( ) 1 (
2
1
2
2
1
2
1
F E F E
F F F E F
F E F F F F F E
E



-
(

- -
(

- - |
.
|

\
|
-

T
T T
T T T
_
1
1 1 1

(2.391)
Thus, we can verify that
) 2 ( ) 1 (
E E E - , i.e. the Green-Lagrange strain tensor is not
additive for increments of motions. We can apply the same reasoning to the Almansi strain
tensor, ( )
1
2
1
- -
- F F e
T
1 , (current configuration). To demonstrate this, we define the
following tensors in the intermediate and current configurations because of the
transformations
) 1 (
F and
) 2 (
F :
intermediate configuration B :
current configuration B:
|
.
|

\
|
-
- -

1
) 1 ( ) 1 ( ) 1 (
2
1
) , ( F F x e
T
t 1
,
|
.
|

\
|
-
- -

1
) 2 ( ) 2 ( ) 2 (
2
1
) , ( F F x e
T
t 1
,

(2.392)
The Almansi strain tensor can be written in terms of
) 1 (
F and
) 2 (
F as:
( ) ( ) ( )
( )
|
.
|

\
|
-
|
|
.
|

\
|
|
.
|

\
|
-
(

- -
- - - -
-
-
- -
- -



1
) 2 (
1
) 1 ( ) 1 ( ) 2 (
1
) 1 ( ) 2 (
1
) 2 ( ) 1 (
1
) 1 ( ) 2 ( ) 1 ( ) 2 ( 1
2
1
2
1
2
1
2
1
F F F F
F F F F
F F F F F F e
T T
T T
T
T
1
1
1 1

(2.393)
Taking into account that ( )
) 1 (
1
) 1 ( ) 1 (
2e F F -
- -
1
T
, we obtain:
( )
) 2 (
1
) 2 ( ) 1 ( ) 2 (
1
) 2 ( ) 2 (
1
) 2 ( ) 1 ( ) 2 (
1
) 2 ( ) 1 ( ) 2 (
1
) 2 ( ) 2 (
1
) 2 ( ) 1 ( ) 2 (
1
) 2 (
1
) 1 ( ) 1 ( ) 2 (
2
1
2
2
1
2
2
1
2
1
e F e F
F F F e F
F e F F F
F e F F F F F e
-
|
.
|

\
|
- -
|
.
|

\
|
- -
|
.
|

\
|
- - |
.
|

\
|
-
- -
- - - -
- - - -
- - - - - -




T
T T
T T
T T T
1
1
1 1 1

(2.394)
Hence, we can see that
) 2 ( ) 1 (
e e e - and Figure 2.41 shows the strain tensors defined by
successive configurations.



2 CONTINUUM KINEMATICS

241













Figure 2.41: Motion defined by successive configurations.
2.15.2 The Logarithmic Strain Tensor
Let us consider a bar subjected to a stretching, in which we define the differential strain
along the axis of the bar as:
( ) /
|
|
.
|

\
|
r r
}
ln ln
0 0 0
0
1
L
L
dL
L L
dL
d
f
L
L
axial axial
f
ting by integra

(2.395)
where
axial
r is known as the logarithmic strain or true strain. Note that, if there are successive
increments of displacement, i.e.
) 1 (
0 f
L L , and
f f
L L
) 1 (
, it follows that the logarithmic
strain is additive, i.e.:
) 2 ( ) 1 (
) 1 (
0
) 1 (
) 1 ( 0
) 1 (
0
0
0
0
1 1 1
axial axial
f
f f
f
f
f f
Total
axial
L
L
L
L
dL
L
dL
L
dL
L
L
L
L
L
L
L
r - r
|
|
.
|

\
|
-
|
|
.
|

\
|
- r
} } }
ln ln (2.396)
Then, from the logarithmic strain definition in (2.395), we can define the logarithmic strain
tensor as:
( ) ( )

/
3
1
) ( ) ( ) ( ) (

;
a
a a
a
Ln Ln
N N U U U ln ln
The logarithmic strain
tensor
(2.397)
Likewise, we can define this tensor in the current configuration, which is also known as the
Hencky strain tensor:
( ) ( )

/
3
1
) ( ) ( ) ( ) (

;
a
a a
a
Ln Ln
n n V V V ln ln The Hencky strain tensor (2.398)
The tensors
) (Ln
V and
) (Ln
U have the same eigenvalues. Also, the following condition is
satisfied:

1 1
, x X

2 2
, x X

3 3
, x X

0
B

B

3

e

2

e

1

e
O
x
,
d

) 1 (
F
B

) 2 (
F

) 1 ( ) 2 (
F F F
x
,
d
X
,
d

) 1 ( ) 2 ( ) 1 ( ) 1 (
F E F E E -
T

) , (
) 1 (
t X E
,


) 2 (
1
) 2 ( ) 1 ( ) 2 (
e F e F e -
- -

T

) , (
) 2 (
t x e
,


) , (
) 1 (
t x e
,


) , (
) 2 (
t x E
,

NOTES ON CONTINUUM MECHANICS

242
( ) ( ) ( ) ( ) ( ) ( )
3 2 1 3 2 1
) ( ) (
/ / / / - / - / ln ln ln ln Tr Tr
Ln Ln
V U (2.399)
2.15.3 The Biot Strain Tensor
We define the Biot strain tensor ( H), in the reference configuration, as:
( )

- / -
3
1
) ( ) (

1 ;
a
a a
a
N N H 1 U H
The Biot strain tensor
(Reference configuration)
(2.400)
and in the current configuration as:
( )

-
/ - -
3
1
) ( ) ( 1

1 ;
a
a a
a
n n h V 1 h
The Biot strain tensor
(Current configuration)
(2.401)
2.15.4 Unifying the Strain Tensors
Using the above concepts, it is possible to define several tensors according to the basis of
U or V . Then, we can define the strain tensors in the material configuration as:
( )
( ) ( )

|
|
.
|

\
|
- -

0
2
1
0
1 1
) (
2
m for
m for
m m
m
m
m
C
C
E
ln ln U
1 1 U
U (2.402)
where m is a positive integer number. The tensors defined above may be represented by
the eigenvalues of U (principal stretches), then:
( )

/
- /
/

0

) (
0

1
1
) (
3
1
) ( ) (
3
1
) ( ) (
m for
m for
m
a
a a
a
a
a a m
a
i
m
N N
N N
ln
E (2.403)
Notice that, depending on the value of m we recover the following tensors
2 m
( )
( ) 1
1 U U
-
-
C
E E
2
1
2
1
) (
2 ) 2 (
Green-Lagrange strain tensor
1 m ( ) 1 U H U - ) (
) 1 (
E Biot strain tensor (2.404)
0 m ( ) U U U ln
) ( ) 0 (
) (
Ln
E Logarithmic strain tensor
Next, we can define the strain tensors in the spatial configuration as:
( )
( )

0
0
1
) (
m for
m for
m
m
m
V
1 V
U
ln
e (2.405)
where m is a negative integer number. The above equations can be expressed in the
spectral representation as:
2 CONTINUUM KINEMATICS

243
( )

/
- /
/

0

) (
0

1
1
) (
3
1
) ( ) (
3
1
) ( ) (
m for
m for
m
a
a a
a
a
a a m
a
i
m
n n
n n
ln
e (2.406)
Notice that, depending on the value of m we recover the following tensors:
- 2 m
( )
( )
1
2 ) 2 (
2
1
2
1
) (
-
- -
-
-
b
e e
1
V 1 U
The Almansi strain tensor
- 1 m ( )
1 ) 1 (
) (
- -
- V 1 h V e The Biot strain tensor (2.407)
0 m ( ) V V V ln
) ( ) 0 (
) (
Ln
e The Hencky strain tensor
2.15.5 One Dimensional Measurements of Strain (1D)
2.15.5.1 Cauchys strain or Engineering strain or the Linear strain
1
0
0
0
- /
-

A
r
L
L L
L
L
C

(2.408)
where / shows stretch.
2.15.5.2 The Logarithmic or True strain
) (
0
0
/
|
|
.
|

\
|
r
}
ln ln
L
L d
L
L
H


(2.409)
or:
- r - r r - r
2
) (
2
1
) 1 (
C C C H
ln (2.410)
2.15.5.3 The Green-Lagrange strain
Generally speaking we obtain the following relationship:
( )
2
2 2
2 2
) (
) ( ) (
2
1
1 2 ) ( ) (
dS
dS ds
D d d dS ds
G
-
r - X E X
, ,

(2.411)
In the one dimensional (uniaxial case), we have:
( ) ( )( ) ( ) ( )
2 2 2
2
0
2
0
2
2
1
1
2
1
1
2
1
1 1
2
1
1
2
1
2
C C C G
L
L L
r - r r - / - / - / - / - /
-
r
(2.412)
2.15.5.4 The Almansi strain
Generally speaking we obtained the following relationship:
( )
2
2 2
2 2
) (
) ( ) (
2
1
1 2 ) ( ) (
ds
dS ds
D d d dS ds
A
-
r - x e x
, ,

(2.413)
In the one dimensional (uniaxial case), we have:
NOTES ON CONTINUUM MECHANICS

244
( )
2
2
2
0
2
1
2
1
2
-
/ -
-
r
L
L L
A

(2.414)
Additionally, the following relationships are satisfied:
( )
2 2
2
1
1
1
2
1
/
r
r -
|
.
|

\
|
r - r r
G
C
C C A

A
G
r
r
/
2

(2.415)
2.15.5.5 The Swaiger strain
1 0
1
-
/ -
-

A
r
L
L L
L
L
S

(2.416)
Additionally, the following relationship is satisfied:
C
C
S
r -
r
r
1

(2.417)
2.15.5.6 The Kuhn strain
( )
1 2
2
0
3
0
3
3
1
3
-
/ - /
-
r
L L
L L
K

(2.418)
In a small strain regime (infinitesimal strain) it holds that:
K S A G C H
r = r = r = r = r = r = r
(2.419)
We can draw a graph where the abscissa is represented by stretch ( / ) and the ordinate
represents the strains, (see Figure 2.42). We can also verify that for stretch values close to
unity the relations in (2.419) are met.

-5
-3
-1
1
3
5
7
0 1 2 3 4
Engineering
Logarithmic
Green-Lagrange
Almansi
Swaiger
Kuhn

Figure 2.42: Curve stretch vs. strains.
/
r
Range of the infinitesimal strain regime
3 Stress










3.1 Introduction
When an external force is acting on a body, the atoms or molecules that make up the
continuum are affected and undergo a position change to achieve balance. Resistance to
this movement depends on the characteristics of the atoms or molecules that make up the
continuum. Internal resistance to movement is called internal force, and can be interpreted as
the average of the interatomic forces of a handful of atoms, thereby characterizing internal
force as a macroscopic variable. The internal force at each material point (particle) of the
continuum is represented by the traction vector field (force per unit area), which is the starting
point to establish the stress state at a material point.
3.2 Forces
When forces are applied directly to a body they are known as surface forces (e.g. contact forces
between two bodies), whereas when the body is immersed in, for instance, a gravitational
or electromagnetic field we have an indirect force. As regards forces, this chapter will only
deal with surface and gravitational forces.
3.2.1 Surface Forces (Traction)
An example of a surface force is illustrated in Figure 3.1 in which the water pressure on the
dam is substituted by the surface force ) (
*
x
,
,
t also called the traction vector. The total force
acting on the dam wall can be obtained by means of the surface integral over the surface

S :
3
Stress
245 , Notes on Continuum Mechanics, Lecture Notes on Numerical
Methods in Engineering and Sciences 4, DOI 10.1007/978-94-007-5986-2_5,
International Center for Numerical Methods in Engineering (CIMNE), 2013
E.W.V. Chaves
NOTES ON CONTINUUM MECHANICS

246
} }


t
S S
S d d ) (
*
x f f
,
, , ,

(3.1)
where f
,
d is the differential force acting on the differential area

S d (surface element),
where it holds that

t S d d ) (
*
x f
,
, ,
. The unit of the surface forces in the International
System of Units (SI) is Pa m N
2
/ (a Pascal is equivalent to one Newton per square
metre).








Figure 3.1: Surface force (traction).
3.2.2 Gravitational Force (Body Force)
When the continuous medium is submerged in a gravitational field, the continuum is also
subjected to a force which can be represented either by force per unit mass (body force),
b
,
, or by force per unit volume (force density), p
,
. These two forces are related to each
other by means of the equation:
p b
,
,
j i i
p b j
(3.2)
where ) , ( t x
,
j is the mass density (mass per unit volume). The units of j , b
,
, p
,
, in the
International System of Units (SI), are [ ]
3
m
kg
j , [ ]
2
s
m
kg
N
b
,
, [ ]
3
m
N
p
,
. Then, the total
force acting on the body defined by the domain B can be evaluated by means of the
integral:
dV dm d
V
b b
, , , ,
j
} } }

B B
F F (3.3)
where we must have take into account that dV dm j .
Problem 3.1: Ignoring the curvature of the earths surface, the gravitational field can be
assumed to be uniform as shown in Figure 3.2, where g is the acceleration caused by
gravity (the gravity of the Earth). Find the resultant force acting on the body B.
Solution:
All bodies immersed in a force field are subjected to the body force b
,
, and in the special
case presented in Figure 3.2 this is given by:
=
) (
*
x
,
,
t

S
3 STRESS

247
(

(
(
(

2
0
0
) , (
s
m
g
t
i
x
,
b
Hence, the total force acting on the body can be evaluated as follows:
(
(
(
(

(
-


}
}
V
i
V
i
dV g
V d t

0
0
) , (
j
j x
,
b F
We can also verify the F unit: [ ]
[ ]

) (
2
3
2 3
Newton N
s
m kg
dV
s
m
m
kg
V
m

(

}
F .













Figure 3.2: Gravitational field.
3.3 Stress Tensors
Let
0
B be a material body in the reference configuration, which is defined by the volume
0
V and bounded by the surface
0
S . After motion, the material body
t
B occupies a new
configuration characterized by volume
t
V which is delimited by the surface
t
S , (see Figure
3.3).








Figure 3.3: Reference and current configuration.

2
x
B

1
x

3
x
g

0
B

0
P
0 t
Reference
configuration

1
x

2
x

3
x

0
S

0
V

t
B
) , ( t P x
,

t
Current
configuration

t
S

t
V
NOTES ON CONTINUUM MECHANICS

248
If the continuum is subjected to external forces (surface, gravitational or otherwise), the
body is also subjected to internal forces. In this section, we will define a continuous and
differentiable tensor field so as to represent the internal force, thereby characterizing the
stress state in the continuous medium.
3.3.1 The Cauchy Stress Tensor
3.3.1.1 The Traction Vector
Let us consider a continuum in the current configuration (deformed) which has been
divided into two parts by a plane H passing through the point ) , ( t P x
,
, (see Figure 3.4).
This plane is defined by said point and by the normal n (unit vector). Let us also consider
a deformed area a A centered at the point ) , ( t P x
,
, so, the outcome of the internal force
acting on this area element is denoted by

Af . Then, we can define the traction vector (also


called the stress vector) at the point ) , ( t P x
,
and which is associated with the normal n , as:
|
|
|
.
|

\
|
A
A


A
a
im t
a
f
n x
n
0
) (
) , , (
,
,
t
(

Pa
m
N
2
(3.4)
Note that, the traction vector
)

(n
t
,
can vary from point to point and said variation defines the
traction vector field. Additionally, at a point ) , ( t P x
,
the traction vector is only dependent on the
normal n . This traction vector represents the force per unit deformed area and its limit (3.4) exists
because the medium was assumed to be continuous.











Figure 3.4: Traction vector.
3.3.1.2 Cauchys Fundamental Postulate
Cauchys Fundamental Postulate: the traction
) (n
t
,
is a function of the position ) ( x
,
and
normal n .
For instance, let us consider a plane
1
H defined by the normal
) 1 (
n that passes through the
point ) , ( t P x
,
, thereby, defining the traction vector
) (
) 1 (
n
t
,
at the point, which is associated
with the normal
) 1 (
n , (see Figure 3.5(a)). Now if we consider a second plane
2
H defined by
0 Aa

t
B

) (n
t
,

n
n
P
H

1
x

2
x

3
x
x
,

Current configuration - t

) (n
t
,

3 STRESS

249
the normal
) 2 (
n , which also passes through the point ) , ( t P x
,
, there will be another traction
vector
) (
) 2 (
n
t
,
which is associated with this new plane, (see Figure 3.5(a)). Then an
immediate result of Cauchys Fundamental Postulate is the principle of action and reaction, (see
Figure 3.5(b)):
) , ( ) , ( n x n x
,
,
,
,
t t - - (3.5)








Figure 3.5: Traction vector.
The stress state at a point ) , ( t P x
,
is completely described when the traction vector ) , ( n x
,
,
t
can be obtained for any arbitrary plane passing through this point ) , ( t P x
,
. Cauchy showed
that if we define the traction vector on three mutually perpendicular planes passing through
the point ) , ( t P x
,
we can fully describe the stress state at that point, (see Figure 3.6).













Figure 3.6: The stress state at the point P.
If we adopt three planes that are perpendicular to the unit vectors
1

e ,
2

e and
3

e , we can
obtain three traction vectors associated with each direction and represented by
)

(
1
e
t
,
,
)

(
2
e
t
,

and
)

(
3
e
t
,
respectively, (see Figure 3.6). Then by breaking each traction vector down
according to the directions
1
x ,
2
x and
3
x , (see Figure 3.7), we can obtain:

) 1 (
n

) (
) 1 (
n
t
,

) 2 (
n

) (
) 2 (
n
t
,

P
1
H
2
H
b) a)

) (
) 1 (
n
t
,

) 1 (
n
P
1
H

) 1 (
n -
) (
) 1 (
n
t
,
-

1
x

2
x

3
x

)

(
2
e
t
,


)

(
3
e
t
,


)

(
1
e
t
,


2

e
2

e

3

e

1

e

P

1
x

2
x

3
x
Current configuration
NOTES ON CONTINUUM MECHANICS

250

- -
- -
- -
3

3 2

2 1

1
)

(
3

3 2

2 1

1
)

(
3

3 2

2 1

1
)

(



3 3 3 3
2 2 2 2
1 1 1 1
e e e t
e e e t
e e e t
e e e e
e e e e
e e e e
t t t
t t t
t t t
,
,
,
(3.6)














Figure 3.7: Representation of the three tractions in the Cartesian basis.
In order to consider these three vectors simultaneously we can define a second-order
tensor as:
3 3

3 2 3

3 1 3

3
3 2

2 2 2

2 1 2

2
3 1

1 2 1

1 1 1

1
3
)

(
2
)

(
1
)

(




3 2 1
3 2 1
3 2 1
3 2 1
e e e e e e
e e e e e e
e e e e e e
e t e t e t
e e e
e e e
e e e
e e e
- - -
- - - -
- - -
- -
t t t
t t t
t t t
, , ,
(3.7)
Note that, in Chapter 1 we established that any second-order tensor can be represented by
a linear combination of dyads.
Then we can rearrange the components of into matrix form, and make a change in the
nomenclature so that we obtain:

33 32 31
23 22 21
13 12 11

1
3 2 1
3 2 1
3 2 1
(
(
(

o o o
o o o
o o o

(
(
(

e e e
e e e
e e e
t t t
t t t
t t t

(3.8)
thus, defining
ij
o as the components of the Cauchy stress tensor also called the true stress tensor,
:
)

(
j i ij
e e o The Cauchy stress tensor [ ] Pa (3.9)
The representation of the Cauchy stress tensor components in the Cartesian system is
shown in Figure 3.8(a).

1
x

2
x

3
x

1

1

1
e
e
t

2

2

1
e
e
t
3

3

1
e
e
t

1

1

2
e
e
t

3

3

2
e
e
t

2

2

2
e
e
t

3

3

3
e
e
t

1

1

3
e
e
t
2

2

3
e
e
t

)

(
1
e
t
,


)

(
2
e
t
,


)

(
3
e
t
,

3 STRESS

251












Figure 3.8: Stress state at a point P.
In the literature, we can find other nomenclature for the Cauchy stress tensor components,
distinguishing the scientific and engineering notations, (see Figure 3.9):
_ _
Notation
g Engineerin
z zy zx
yz y yx
xz xy x
Notation
Scientific
zz zy zx
yz yy yx
xz xy xx
ij
(
(
(

o t t
t o t
t t o

(
(
(

o o o
o o o
o o o

(
(
(

o o o
o o o
o o o
o
33 32 31
23 22 21
13 12 11

(3.10)
NOTE: It is important to note that many authors (mostly engineers) reverse the
convention of the indices, so to avoid misunderstandings, when we refer to engineering
notation the symmetry of the Cauchy stress tensor is already implicit.











Figure 3.9: Stress state at a point P Engineering notation.
OBS.: The Cauchy stress tensor is a symmetric tensor
sym
= . Therefore it
holds that
T
. Proof of this is provided in Chapter 5.

x
o

xy
t
xz
t

z
o

yz
t

xz
t

y
o

yz
t

xy
t
x
y
z

11
o

21
o

31
o

33
o

23
o

13
o

22
o

32
o

12
o

1
x

2
x

3
x

11
o

21
o
31
o

33
o

23
o

13
o

22
o

32
o

12
o

1
x

2
x

3
x
b) rear faces a) front faces
NOTES ON CONTINUUM MECHANICS

252
Taking into account the symmetry of the Cauchy stress tensor, the representation of its
components in Voigt notation is given by:
{
(
(
(
(
(
(
(
(

t
t
t
o
o
o

(
(
(
(
(
(
(
(

o
o
o
o
o
o

(
(
(
(
(
(
(
(

o
o
o
o
o
o

(
(
(
(
(
(

o o o
o o o
o o o
o
xz
yz
xy
z
y
x
xz
yz
xy
zz
yy
xx
Voigt
ij
13
23
12
33
22
11
33 23 13
23 22 12
13 12 11

(3.11)
3.3.2 The Relationship between the Traction and the
Cauchy Stress Tensor
Our goal now is that given the nine components of the Cauchy stress tensor, how can we
find the traction vector associated with an arbitrary plane? It is very easy to answer this
question if we consider that the projection of the second-order tensor ( ) according to the
direction ( n ) is given by n
n

) (
t
,
, (see Chapter 1).
We will prove we can obtain the same result by starting from the forces equilibrium at the
material point. To do this, we define an arbitrary plane ABC with the normal n , (see
Figure 3.10), where the plane ABC passes through the point P .














Figure 3.10: The traction vector in an arbitrary plane.
Associated with this plane is the traction vector
) (n
t
,
. Taking into account the Cauchy stress
tensor components in the rear faces of the tetrahedron, (see Figure 3.8(b)), and by
considering that the point is in equilibrium, the balance of forces according to the
1
x -
direction is evaluated as follows:
0
13 3 12 2 11 1
) (
1
o - o - o - n S n S n S S
n
t (3.12)

T
A

d
n

) (n
t
,


11
o

21
o

31
o

33
o

23
o

13
o

22
o

32
o

12
o

1
x

2
x

3
x

2
n S

3
n S

1
n S
S
A
B
C
O
n n A A ) (
2
1
S d AC AB d = r
T T


S ABC

i i
S d e

) ( n A
T


BOC n S S d
1 1 1

) ( e n A
T


AOC n S S d
2 2 2

) ( e n A
T


AOB n S S d
3 3 3

) ( e n A
T


3 STRESS

253
where S is the triangle ABC area and the projection of the area S according to the planes
3 2
x x - ,
3 1
x x - and
2 1
x x - is given respectively by
1
n S ,
2
n S and
3
n S , (see Figure 3.10).
Then, if we simplify the equation (3.12) we obtain:
13 3 12 2 11 1
) (
1
o - o - o n n n
n
t (3.13)
Similarly, the balance of forces according to the directions
2
x and
3
x provide us, the
following relationships respectively:
32 3 22 2 12 1
) (
2
o - o - o n n n
n
t (3.14)
33 3 23 2 13 1
) (
3
o - o - o n n n
n
t (3.15)
Then by rearranging the equations (3.13), (3.14) and (3.15) into matrix form we obtain:
(
(
(

(
(
(

o o o
o o o
o o o

(
(
(

3
2
1
33 32 31
23 22 21
13 12 11
) (
3
) (
2
) (
1


n
n
n
n
n
n
t
t
t

(3.16)
The above equations can still be represented as:
Indicial notation Tensorial notation
j ij i
n
) (
o
n
t n
n

) (
t
,

(3.17)
Then, by referring to the symmetry of the Cauchy stress tensor ( ), the traction vector
components can be represented in Voigt notation as follows:
{ [ ] {
T
n n n
n n n
n n n
N T
(
(
(
(
(
(
(
(

o
o
o
o
o
o
(
(
(


0 0 0
0 0 0
0 0 0
13
23
12
33
22
11
1 2 3
3 1 2
3 2 1
) (
3
) (
2
) (
1
n
n
n
t
t
t

(3.18)

Problem 3.2: The Cauchy stress tensor components at a point P are given by:

Pa
ij

2 5 . 0 1
5 . 0 3 4
1 4 8
(
(
(

-
-
o

a) Calculate the traction vector (
) (n
t
,
) at P
which is associated with the plane ABC
defined in Figure 3.11.
b) With reference to paragraph a).
Obtain the normal (
N

,
) and tangential (
S

,
)
traction vectors at P (see Appendix A).

Figure 3.11: Plane ABC .
Solution:
First, we obtain the unit vector which is normal to the plane ABC . To do this we choose
two vectors on the plane:
) 0 , 2 , 0 ( B
) 5 , 0 , 0 ( C
) 0 , 0 , 3 ( A
1
x

2
x

3
x
O
n
NOTES ON CONTINUUM MECHANICS

254
3 2 1
3 2 1

3
e e e
e e e
- - -
- - -


OB OC BC
OB OA BA

Then, the normal vector associated with the plane ABC is obtained by means of the cross
product between

BA and

BC , i.e.:
3 2 1
3 2 1

15

10
0 2 3
5 2 0

e e e
e e e
- -
-
- r

BA BC n
,

Additionally, the unit vector codirectional with n
,
is given by:
3 2 1

19
6

19
15

19
10
e e e - -
n
n
n


Then by using the equation in (3.16), we can obtain the traction components as:
Pa
j ij i

6
15
10

2 5 . 0 1
5 . 0 3 4
1 4 8
19
1

3
2
1
) (
(
(
(

(
(
(

-
-

(
(
(

o
t
t
t
n t
n
Pa
5 . 29
8
26
19
1
3
2
1
(
(
(

(
(
(

t
t
t

b) The traction vector
) (n
t
,
associated with the normal n can be broken down into a
normal (
N

,
) and a tangential (
S

,
) vector as shown in Figure 3.12. Then,
S N
t
, ,
,
-
) (n
or s n
n

) (
S N
o - o t
,

where
N
o and
S
o are the magnitudes of
N

,
and
S

,
, respectively.
















Figure 3.12: Normal and tangential stress vector.

As we have seen in Appendix A, the normal component,
N
o , can be evaluated as follows:
) ( ) ( ) ( ) (
) ( ) (
j i ij j ij i i j ij i i N
n n n n n n n o o o o
n n
n n n n n n n t : t
,

Thus:
[ ] Pa n
N i i N
54 . 1
6
15
10
5 . 29 8 26
19
1

2
=
(
(
(

o o t
Then the tangential component,
S
o , can be obtained by means of the Pythagorean
Theorem, i.e.:

1

e

2

e

) (n
t
,

n

N

,


S

,


1
x

2
x

3
x
s
P

3

e
n
n

) (
t
,


S N
t
, ,
,
-
) (n

n o
N N

,


2 2
2
) (
S N
o - o
n
t
,

3 STRESS

255
2 ) ( ) ( 2 2 2
2
) (
N i i S S N
o - o o - o
n n n
t t t
,

where
[ ] 46 . 4
5 . 29
8
26
5 . 29 8 26
19
1
2
) ( ) (
=
(
(
(

n n
i i
t t
Thus,
Pa
N i i S
0884 . 2 3716 . 2 46 . 4
2 ) ( ) (
= - o - o
n n
t t

Problem 3.3: The stress state at a point in the continuum is represented by the
components of the Cauchy stress tensor as:
Pa
ij
(
(
(

o
2 0 0
0 2 1
0 1 2

a) Obtain the components of in a new system
3 2 1
, , x x x , where the transformation
matrix is given by:















b) Obtain the principal invariants of ;
c) Obtain the eigenvalues and eigenvectors of . Also verify if the eigenvectors form a
basis transformation between the original and the principal space;
d) Illustrate the Cauchy stress tensor graphically, i.e. with the Mohrs circle in stress, (see
Appendix A);
e) Obtain the spherical (
sph
) and the deviatoric (
dev
) part of . Also, find the principal
invariants of
dev
;
f) Obtain the octahedral normal (
oct
N
o ) and tangential (
oct
S
o ) components of , (see
Appendix A).
Solution:
a) As we have seen in Chapter 1, the transformation law for the components of a second-
order tensor is given by:
T
kl jl ik ij
a a A A
form Matrix
o o
Thus,

2
x

1
x

3
x

1
x

2
x

3
x

1
c

1
,

1


2

e

3

e

1

e

3

e

1

e
2

e

(
(
(

-

3 0 4
0 5 0
4 0 3

5
1
A
ij
a
where

.
1 13
1 12
1 11
cos
cos
cos

,
c

a
a
a

NOTES ON CONTINUUM MECHANICS

256
(
(
(

(
(
(

-
(
(
(

(
(
(

-
o
2 8 . 0 0
8 . 0 2 6 . 0
0 6 . 0 2
3 0 4
0 5 0
4 0 3
2 0 0
0 2 2
0 1 1

3 0 4
0 5 0
4 0 3

5
1
2
ij

These new components
ij
o can be appreciated in Figure 3.13.

































Figure 3.13: Basis transformation.

b) The principal invariants of the Cauchy stress tensor can be calculated as follows:

33 22 11
) ( o - o - o o
ii
I

Tr

[ ] ( )
2
23
2
13
2
12 22 33 33 11 22 11
2 2
2
1
) ( ) (
2
1
o - o - o - o o - o o - o o
o o - o o -
ij ij jj ii
I I

Tr Tr


( )
2
12 33
2
13 22
2
23 11 13 23 12 33 22 11
3 2 1
2
2 3
6
1
) (
o o - o o - o o - o o o - o o o
o o o - o o o - o o o o o o
ki jk ij jk jk ii kk jj ii k j i ijk
I I I

det

By substituting the values of
ij
o for those in the proposed problem we obtain:

P

1
x
1
x

3
x

2
x
3
x

2
x

T
A A
A A
T


11
o

12
o
13
o

33
o

23
o

13
o

22
o

23
o

12
o

1
x

2
x

3
x

11
o

12
o
13
o

33
o

23
o

13
o

22
o

12
o

1
x

2
x

3
x

23
o
3 STRESS

257
6 ; 11
2 1
1 2
2 0
0 2
2 0
0 2
; 6 - -

I I I I I I
c) The principal stresses (
i
o ) and principal directions (
) (

i
n ) are obtained by solving the
following set of equations:
(
(
(

(
(
(

(
(
(

o -
o -
o -
0
0
0

2 0 0
0 2 1
0 1 2
3
2
1
n
n
n

To obtain the nontrivial solutions of
) (

i
n we have to solve the characteristic determinant,
which is a cubic equation for the unknown magnitude o:
0 0
2 3
- o - o - o o - o

I I I I I I
ij ij
c
However, if we look at the format of the Cauchy stress tensor components, we can notice
that we already have one solution as in the
3
x -direction the tangential components are
equal to zero, then:
o
direction Principal
3
2 0
) 3 (
2
) 3 (
n n
1
, 1
) 3 (
3
n
To obtain the other two eigenvalues, one only need solve:
( )

o
o
- o -
o -
o -
3
1
0 1 2
2 1
1 2
2
1 2

Then we can express the Cauchy stress tensor components in the principal space as:
Pa
ij
(
(
(

o
2 0 0
0 3 0
0 0 1

Additionally, the principal direction associated with 1
1
o is calculated as follows:
) 1 (
2
) 1 (
1
) 1 (
2
) 1 (
1
) 1 (
2
) 1 (
1
) 1 (
3
) 1 (
2
) 1 (
1
0
0
0
0
0

1 2 0 0
0 1 2 1
0 1 1 2
n n
n n
n n
n
n
n
-

-
-

(
(
(

(
(
(

(
(
(

-
-
-

with 0
) 1 (
3
n and by using the condition 1
2
) 1 (
2
2
) 1 (
1
- n n we obtain:
2
1
) 1 (
2
) 1 (
1
- n n then
(

-
0
2
1
2
1

) 1 (
i
n
Since is a symmetric tensor, the principal space is formed by an orthogonal basis, so, it
is valid that:
) 2 ( ) 1 ( ) 3 ( ) 1 ( ) 3 ( ) 2 ( ) 3 ( ) 2 ( ) 1 (

;

;

n n n n n n n n n r r r
Thus, the second principal direction can be obtained by the cross product between
) 3 (

n
and
) 1 (

n , i.e.:
2 1
3 2 1
) 1 ( ) 3 ( ) 2 (

2
1

2
1
0
2
1
2
1
1 0 0


e e
e e e
n n n -
-
r
which can also be checked by the following analysis:
The Principal direction associated with 3
2
o :
) 2 (
2
) 2 (
1
) 2 (
2
) 2 (
1
) 2 (
2
) 2 (
1
) 2 (
3
) 2 (
2
) 2 (
1
0
0
0
0
0

3 2 0 0
0 3 2 1
0 1 3 2
n n
n n
n n
n
n
n

-
- -

(
(
(

(
(
(

(
(
(

-
-
-

NOTES ON CONTINUUM MECHANICS

258
With 0
) 3 (
3
n and using the condition 1
2
) 3 (
2
2
) 3 (
1
- n n we obtain:
2
1
) 2 (
2
) 2 (
1
n n then
(

0
2
1
2
1

) 2 (
i
n
As we have seen in Chapter 1, the eigenvectors of a symmetric tensor form the
transformation matrix D , from the original system to the principal space, i.e.
T
D D , thus:
(
(
(
(
(
(
(

-
(
(
(

(
(
(
(
(
(
(

(
(
(

o
o
o
1 0 0
0
2
1
2
1
0
2
1
2
1

2 0 0
0 2 1
0 1 2

1 0 0
0
2
1
2
1
0
2
1
2
1
2 0 0
0 3 0
0 0 1
3
2
1

d) The graphical representation of a second-order tensor can be obtained from the
description in Appendix A. To do this we have to restructure the eigenvalues of so that
I I I I I I
o > o > o , thus:
1 ; 2 ; 3 o o o
I I I I I I

Then the three circumferences are defined by:
5 . 0 ) (
2
1
) ( ; 5 . 2 ) (
2
1
) ( ; 3 Circle
0 . 1 ) (
2
1
) ( ; 0 . 2 ) (
2
1
) ( ; 2 Circle
5 . 0 ) (
2
1
) ( ; 5 . 1 ) (
2
1
) ( ; 1 Circle
3 3
2 2
1 1
o - o o - o
o - o o - o
o - o o - o
I I I I I I
I I I I I I I I
I I I I I I I I I I
R radius C center
R radius C center
R radius C center


Then, we can illustrate the Cauchy stress tensor at P by means of Mohrs circle in stress as
shown in Figure 3.14.
















Figure 3.14: Mohrs circle in stress at the point P .






N
o
1 o
III

1
max
o
S


S
o
2 o
II


1
R

3
R

2
R

1
C

3
C

max
3
N I
o o
0 . 1 ) (
2
1
max
o - o o
I I I I S

3 STRESS

259
e) As defined in Chapter 1, a second-order tensor can be broken down additively into a
spherical and a deviatoric part, i.e.:
Tensorial notation
Indicial notation
dev
m
dev sph
1

- o
-

dev
ij ij m
dev
ij ij kk
dev
ij
sph
ij ij
o - o
o - o
o - o o
c
c
3
1

(3.19)
A schematic representation of these components in the Cartesian basis can be appreciated
in Figure 3.15 and the value of the scalar
m
o is evaluated as follows:
2
3
6
3
) (
3
1
3
1
3 3
3 2 1 33 22 11
o
o - o - o

o - o - o
o

I
kk m
Tr
Then the spherical part becomes:
(
(
(

o o
2 0 0
0 2 0
0 0 2
2
ij ij m
sph
ij
c c
And, the deviatoric part can be evaluated as follows:
(
(
(

o - o - o o o
o o - o - o o
o o o - o - o

(
(
(

o
o
o
-
(
(
(

o o o
o o o
o o o
o
) 2 (
) 2 (
) 2 (
0 0
0 0
0 0
22 11 33
3
1
23 13
23 33 11 22
3
1
12
13 12 33 22 11
3
1
33 23 13
23 22 12
13 12 11
m
m
m
dev
ij

Thus,
(
(
(

(
(
(

-
-
-
o
0 0 0
0 0 1
0 1 0
2 2 0 0
0 2 2 1
0 1 2 2
dev
ij

Now let us remember from Chapter 1 that and
dev
are coaxial tensors, i.e., they have
the same principal directions, so we can use this information to operate in the principal
space of to obtain the eigenvalues of
sph dev
- . With that we obtain:
(
(
(

(
(
(

o
o
o
-
(
(
(

o
o
o
o
0 0 0
0 1 0
0 0 1
0 0
0 0
0 0
0 0
0 0
0 0
3
2
1
m
m
m
dev
ij

Then the invariants of
dev
are given by:
0 ; 1 ; 0 ) ( -
dev dev dev
I I I I I I
dev

Tr
Traditionally, in engineering, the invariants of the deviatoric stress tensor are represented
by:
( )
( )

I I I I I I I I I I
I I I I I
I
dev
dev
dev
27 9 2
27
1
3
3
1
0
3
3
2
2
1
- -
- -

J
J
J




NOTES ON CONTINUUM MECHANICS

260




























Figure 3.15: The spherical and deviatoric part of .

f) The octahedral normal and tangential components, (see Appendix A), can be expressed
as:
( )
m ii
oct
N
I
o o o - o - o o
3 3
1
3
1
3 2 1



( ) ( ) ( )
3 3
2
6 2
3
1
2
3
2
2
2
1
2
2
dev dev dev
oct
oct
S
I I I
o - o - o
- t = o J


Then, by substituting the values of the proposed problem we obtain:
3
2
3
2
; 6
2
t o o J
oct m
oct
N


3.3.3 Other Measures of Stress
3.3.3.1 The First Piola-Kirchhoff Stress Tensor
As we have seen before, the Cauchy stress tensor was derived in the current configuration
(deformed). In some cases we may wish to adopt the Lagrangian description for studying

_


11
o

12
o
13
o

33
o

23
o

13
o

22
o

23
o

12
o

1
x

2
x

3
x

m
o

m
o

m
o

1
x

2
x

3
x

dev
11
o

12
o
13
o

dev
33
o

23
o

13
o

dev
22
o

23
o

12
o

1
x

2
x

3
x
+

sph


dev


3 STRESS

261
motion, and then it will be necessary to correlate the Cauchy stress tensor with a
hypothetical stress tensor in the reference configuration, (see Figure 3.16).











Figure 3.16: Traction vector Current and reference configuration.
In the reference configuration, we adopt an area element A
,
d with the normal N

and
associated with that plane we can define a pseudo traction vector
)

(
0
N
t
,
. After motion, this area
element becomes a
,
d in the deformed configuration, associated with which we have the
traction vector
) (n
t
,
, (see Figure 3.16). Then by using the definition in (3.4) we can define
)

(
0
N
t
,
and
) (n
t
,
, respectively, as:
a
f
n x
A
N X
n N
,
,

,
,
,
,

, ,
d
d
a
f
im t
d
d
A
F
im t
a A
|
.
|

\
|
A
A
|
.
|

\
|
A
A

A A 0
) (
0
)

(
0
) , , ( ; )

, , ( t t
?

(3.20)
Based on the principle that:
f
, ,
d d ? i i
f d d ?
(3.21)
we can conclude that:
a A
n N
,
, , ,
d d
) ( )

(
0
t t
a A
n N
,
,
d d
i i
) ( )

(
0
t t
(3.22)
a A
a n A N
,
,
,
,
d d
d d




P
P


k ik k ik
k ik k ik
da dA
d n d N
o
o
P
P a A
,
,


(3.23)
where we have introduced a second-order tensor P so that the projection of P according
to the N

-direction results in the traction vector


)

(
0
N
t
,
.
Then by referring to the Nansons formula ( A F a
,
,
d J d
T

-
) obtained in Chapter 2, where
J is the Jacobian determinant and is equal to the determinant of the deformation gradient
F , i.e. F J , the equation in (3.23) can be rewritten as:
( ) A F A F a A
, ,
,
,
d J d J d d
T T

- -
P (3.24)
Thus, we can conclude that:
T T
J
J F F =
-
P P
1

jk ik ij jk ik ij
F
J
F J P P
1

1
o = o
-

(3.25)

3 3
, x X

)

(
0
N
t
,

N


0
P

1 1
, x X

2 2
, x X

A
,
d


) (n
t
,

n
P
A
,
d
Reference configuration - ) 0 (
0
t t Current configuration - t
a
,
d
a
,
d
t
B
0
B
F
NOTES ON CONTINUUM MECHANICS

262
where P is the first Piola-Kirchhoff stress tensor which is also called the nominal stress tensor. This
tensor represents the force in the current configuration per unit undeformed area, so, it is
both a two-point second-order tensor and a non-symmetric tensor, i.e.
T
P P .
In addition to the traction vectors (
)

(
0
N
t
,
,
) (n
t
,
) defined previously, we can find in the
literature other traction vectors that have a purely mathematical transformation, with no
physical meaning, (see Figure 3.17), namely:
t t t R t t t
, , , , , ,
; ;
*
0 0 0
1
0
J
T

O - A
F (3.26)
where R is the polar decomposition rotation tensor, (see Chapter 2). In addition, the
following relationships are valid:
t R t R t R t
t t R t t
t R t t t
, , , ,
, , , ,
, , , ,






A O
- O - - A
O A
T T T
dA
da
dA
da
dA
da
0 0 0
1
0
1
0
1
0
0 0 0
F
F F F
F
(3.27)
3.3.3.2 The Kirchhoff Stress Tensor
We define the Kirchhoff stress tensor , which is related to the traction vector
*
t
,
(current
configuration), as:
t t
*
J J J n n
, ,
(3.28)
As we can verify the Kirchhoff stress tensor is a symmetric second-order tensor and is
related to the Cauchy stress tensor and to the first Piola-Kirchhoff stress tensor by means
of the following relationships:

-

1
; ; F F P P
T
J (3.29)
3.3.3.3 The Second Piola-Kirchhoff Stress Tensor
We can also introduce the second Piola-Kirchhoff stress tensor, S , which is defined in the
reference configuration, as:
Tensorial notation Indicial notation
T
T
J
- -
- - -


F F
F F F

1
1 1

P S

1 1
1 1 1

- -
- - -
o
t
jl kl ik
jl kl ik kj ik ij
F F J
F F F P S

(3.30)
or
T T
F F F F ; S P S P (3.31)
The Cauchy stress tensor can be expressed in terms of the second Piola-Kirchhoff stress
tensor as:
T
J
F F
1
S (3.32)
Next, we can prove that S is a symmetric tensor, i.e. S S
T
:
( ) S S
- - - -

T
T
T T
J J F F F F
1 1

(3.33)

3 STRESS

263



































Figure 3.17: Traction vectors Reference and current configuration.

Reference configuration - ) 0 (
0
t t
Current configuration - t
da
) (n
t
,

n
a
,
d

t
B
dA
)

(
0
N
t
,

A
,
d


0
B
N

0
P t
,


P
-First Piola-Kirchhoff stress tensor
n t
,

- Cauchy stress tensor
da
) (n
t
,

n
a
,
d

t
B
dA
)

(
0
N
t
,

A
,
d


0
B
dA
A
0
t
,


1 -
F
N

0

A
S t
,

S -Second Piola-Kirchhoff stress tensor
n
*
t t
, ,
J
- Kirchhoff stress tensor
dA
)

(
0
N
t
,

A
,
d


0
B
dA
O
0
t
,


T
R
N

0

O
T t
,

T - Biot stress tensor
N


NOTES ON CONTINUUM MECHANICS

264
3.3.3.4 The Biot Stress Tensor
We can also introduce the Biot stress tensor, T , which in general is non-symmetric and is
related to the traction vector
O
0
t
,
:
N

0

O
T t
,
(3.34)
From (3.34) we can obtain:
N N
N

P R T
t R T
T
T
,

(3.35)
Thus,
P R T
T
(3.36)
and by considering that
T
J
-
F P we can obtain:
T T
J
-
F R T (3.37)
If we refer to the equations in (3.36) and (3.31), the tensor T can also be expressed as:
S U S U R R S R P R T
T T T
F (3.38)
where we have applied the right polar decomposition U R F . Note that the tensor T
will be symmetrical if U and S are coaxial tensors, hence 0 S U S U U S
skew
) ( ,
(see Chapter 2 in subsection 1.5.9 Coaxial Tensors).
3.3.3.5 The Mandel Stress Tensor
We can also introduce the Mandel stress tensor denoted by M, which in general is non-
symmetric and is defined as:
( ) ( )
T T T T - -
F F F F F F C
1
P P S M

(3.39)
where C is the right Cauchy-Green deformation tensor, which was defined in Chapter 2.
We can now summarize the relationships between stress tensors defined above as:
T T T T T
J J J J J
F F F F F F
-
M T R S P
1

1 1 1 1

The Cauchy stress
tensor
(3.40)
M T R S P
1

- - -

T T
J F F F F

The first Piola-Kirchhoff stress
tensor
(3.41)
T T T T T
J F F F F F F
-
M T R S P The Kirchhoff stress tensor (3.42)
M T U P S
1 1 1 1 1

- - - - - - -
C F F F F F
T T
J
The second Piola-
Kirchhoff stress tensor
(3.43)
M U S U R P R R T
1

- - -

T T T T T
J F F The Biot stress tensor (3.44)
T U S P M
- -
C F F F F F
T T T T T
J
The Mandel stress tensor (3.45)
Figure 3.18 shows the representation of tensors in different configurations.
NOTE: If the current configuration is very close to the reference configuration, the
deformation gradient is approximately equal to the identity tensor 1, i.e.:
3 STRESS

265
1 ) (
1
= = =
-
F F F det J and 1 (3.46)
In this condition all the stress tensors are equal, i.e. M T S P = = = = = .












Figure 3.18: Stress tensors.
3.3.4 Spectral Representation of the Stress Tensors
For the next stress tensor representation let us consider that the Cauchy stress tensor ( )
and the left stretch tensor ( V ) are coaxial, i.e. they present the same principal directions.
Then, the spectral representation of the Cauchy stress tensor is given by:

o
3
1
) ( ) (

a
a a
a
n n (3.47)
where
a
o are the eigenvalues of , and
) (

a
n are the eigenvectors of or V or b .
Remember from Chapter 2 that the following representations are valid:


-


/

/
/
3
1
) ( ) (
3
1
) ( ) ( 1
3
1
) ( ) (

1
;

1
;

a
a a
a
T
a
a a
a a
a a
a
N n n N N n F F F (3.48)
and also that the polar decomposition rotation tensor is represented by:


3
1
) ( ) (

a
a a
N n R (3.49)
where
a
/ are the principal stretches, and
) (

a
N are the eigenvectors of the right stretch
tensor ( U). If we refer to the relationships between stress tensors given in (3.40) - (3.45),
we can obtain the spectral representation of the stress tensor as follows:
The Kirchhoff stress tensor:


t o o
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (

a
a a
a
a
a a
a
a
a a
a
J J J n n n n n n (3.50)



T T
T T
T
J
J
J
-
-
- -


F
F F
F F
R T
M
S


1

X
,

x
,

F
Reference
configuration
Current
configuration

0
B
B
Current Conf.
Ref. Conf.

S M
T
S
C X
X
X
) , (
) , (
) , (
t
t
t
,
,
,

) , (
) , (
J t
t
x
x
,
,



T
J
-
F P
NOTES ON CONTINUUM MECHANICS

266
The first Piola-Kirchhoff stress tensor:




-

/
o

/
o

|
|
.
|

\
|

/
|
|
.
|

\
|
o


3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) ( ) ( ) (
3
1
) ( ) (
3
1
) ( ) (

1

a
a a
a
a
a a
a
a
a
a a a a
a
a
a
a a
a a
a a
a
T
J J
J J
N n N n N n n n
N n n n P
P
F

(3.51)
As we can verify the first Piola-Kirchhoff stress tensor is neither in the current nor in the
reference configuration, i.e. P is a two-point tensor, and
a
P are not the eigenvalues of P .
The second Piola-Kirchhoff stress tensor is shown as:


- -

/
o


/
o

|
|
.
|

\
|

/
|
|
.
|

\
|
o
|
|
.
|

\
|




3
1
) ( ) (
3
1
) ( ) (
2
3
1
) ( ) ( ) ( ) ( ) ( ) (
2
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (
1

1
a
a a
a
a
a a
a
a
a
a a a a a a
a
a
a
a a
a a
a a
a
a
a a
a
T
J
J
J
J
N N N N
N n n n n N
N n n n n N
S
S
F F

(3.52)
As we can verify, the second Piola-Kirchhoff stress tensor has the same principal directions
as the right stretch tensor ( U).
The Biot stress tensor can be shown as:






-
/
|
|
.
|

\
|

|
|
.
|

\
|
/
/
o

|
|
.
|

\
|

/
|
|
.
|

\
|
o
|
|
.
|

\
|



3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (


a
a a
a a
a
a a
a
a
a a
a
a
a a
a
a
a a
a
a
a
a a
a a
a a
a
a
a a
T T
J
J
J
N N N N
N N N N N N
N n n n n N
S U R T
S T
S
F
(3.53)
Then the Mandel stress tensor is shown as:




-
o
|
|
.
|

\
|

/
|
|
.
|

\
|
o
|
|
.
|

\
|
/



3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (


a
a a
a
a
a a
a
a
a a
a a
a a
a
a
a a
a
T T
J
J
J
N N N N
N n n n n N
M
M
F F

(3.54)
Then, the following is valid:
a
a
a
a
a
a
a
a
a
a
a
J
M T P S
2 2 2
1 1 1 1
/

/
t
/
o
/

(3.55)



3 STRESS

267
Problem 3.4: Prove that the following relationship are valid:
1 1
;
- - - - -
o - o - C F F F F
m
T dev T
m
T dev
J J J J S P
where P and S are the first and second Piola-Kirchhoff stress tensors, respectively, C is
the right Cauchy-Green deformation tensor, F is the deformation gradient, J is the
Jacobian determinant, and the scalar
m
o is the mean normal Cauchy tress. Also prove that
the following relationships are true:
m
Jo 3 C F : : S P
Solution:
First of all we prove that C F : : S P :
C
C F F
F
:
:
S
P

kj kj kj
T
kj
ij ik kj ij kj ik
ij ij
F F F F
F
) ( ) (
) ( ) (
S S
S S
P

Secondly, by referring to the definition
T
J
-
F P , (see equation (3.25)), and the
different components of by
dev sph
- , we obtain:
T
m
T dev
T
m
T dev
T
m
dev
J J
J J
J
- -
- -
-
o -
o -
o -



F F
F F
F

1
1 P

) (

Thirdly, by taking into account the definition
T
J
- -
F F
1
S , (see equation (3.30)), and
by breaking down into
dev sph
- , we obtain:
1 1
1 1
1 1
1
1
) (
) (
- - -
- - -
- - - -
- -
- -
o -
o -
o -
o -






C F F
F F F F
F F F F
F F
F F
m
T dev
T
m
T dev
T
m
T dev
T
m
dev
T
J J
J J
J J
J
J

1
1
S

1
) (
1 1
1 1
) (
1 1
1
) (
1 1 1
1
) (
1
1 1
) (
- - -
- - - -
- - - -
- -
- -
o - o
o - o
o - o
o - o
o
ij m jp
dev
kp ik
jk ik m jp
dev
kp ik
jp kp m ik jp
dev
kp ik
jp kp m
dev
kp ik
jp kp ik ij
C J F JF
F F J F JF
F JF F JF
F JF
F JF
c
c
S

Then by applying the double scalar product between S and C we can obtain:
C C C F F
C C F F C
: :
: :
1 1
1 1
) (
- - -
- - -
o -
o -


m
T dev
m
T dev
J J
J J

S

where the term C F F :
T dev
J
- -

1
becomes:

0


) )( ( ) ( ) (
) (
0 ) (
1 1 1
1 1

o o
o
o

- - - -
- - - -

_
dev
T
dev
dev
kk pk
dev
pk
dev
pk qk qp
qj qi jk
dev
pk ip ij
T
ij
T dev
T dev T dev
J
J J
J
F F F F J
J J


Tr
:
: :
c
c c
F F F F
C F F C F F
F F

Thus:
m m m m
J J J J o o o o
- -
3 ) ( ) (
1 1
1 S Tr Tr C C C C C : :
Now, by taking the double scalar product between P and F we obtain:
F F F F F : : :
T
m
T dev
J J
- -
o - P
NOTES ON CONTINUUM MECHANICS

268
Then by analyzing the term F F :
T dev
J
-
we can conclude that:
0

) ( ) (
0 ) (
1

o o

-
- -

_
dev
dev
ik
dev
ik ij jk
dev
ik
ij ij
T dev T dev
J
J F F J
J J

1

Tr
:
:
c
F F F F

Thus,
m m
T T
m
T
m
J J J J o o o o
- -
3 ) ( ) ( 1 P Tr Tr F F F F F : :






























4. Objectivity of Tensors









4.1 Introduction
Any physical quantity must be invariant for different observers. For example, let us
suppose that two observers are located at different positions, (see Figure 4.1), this means
they must both detect the same stress state acting on the body for there to be physical
meaning.















Figure 4.1: Superimposed rigid-body motion.
4
The Objectivity of Tensors
x x
,
,
,
- ) ( ) (
*
t t Q c


_



B
observer 1
Current configuration

B

*
B
observer
Current configuration
observer 2
269 , Notes on Continuum Mechanics, Lecture Notes on Numerical
Methods in Engineering and Sciences 4, DOI 10.1007/978-94-007-5986-2_6,
International Center for Numerical Methods in Engineering (CIMNE), 2013
E.W.V. Chaves
NOTES ON CONTINUUM MECHANICS

270
The equivalent to the two observers is that one single observer that records the stress state
in the current configuration must be able to compute the same stress state if the continuum
undergoes rigid body motion.
When we are dealing with nonlinear problems it is necessary to approach the constitutive
equations in rates. As we shall see, in general, the rate of change of the tensor, e.g. velocity,
acceleration, etc., is not objective, which can be inconvenient when formulating the
constitutive equation, which by definition must be objective. Therefore, to overcome this
drawback, we will define some rates that are objective.
4.2 The Objectivity of Tensors
Let us consider two possible motions defined by F and
*
F , where the latter, only differs
from the former, by a rigid body motion which in turn is characterized by a proper
orthogonal tensor Q, i.e. a rotation tensor 1 ) ( - Q det , (see Figure 4.2). Then the tensor
( -) is said to be objective, or frame-indifferent, when its counterpart
*
- can be obtained by
the corresponding orthogonal transformation. By virtue of the fact that the motion
characterized by F generates the stress state , and the motion
*
F generates the stress
state
T
Q Q
*
we have the principle of objectivity or material frame indifference.














Figure 4.2: Motions.
Scalars
A scalar is objective if:
c c
*
(4.1)
Then, all scalars are objective.

Reference configuration
F
Q
G
,

F F Q
*

X
,

g
,

A
x
,

c

*
g
,


*
A

*
x
,


*
c
Current configuration
Rotated current configuration
4 THE OBJECTIVITY OF TENSORS

271
Vectors
If g
,
is an Eulerian vector which is generated by the motion F , then we can state that g
,
is
objective if its counterpart
*
g
,
, which is generated by
*
F , is related to g
,
by means of the
equation:
g g
, ,
Q
*
(4.2)
NOTE: We will take this opportunity to mention that the orthogonal transformation law
for two-point tensors (pseudo-tensors) is the same as that for vectors. As examples of two-
point tensor we can quote: the deformation gradient, F F Q
*
; the first Piola-Kirchhoff
stress tensor, P Q P
*
; and the Polar Decomposition rotation tensor, R Q R
*
.
As an example of an objective vector we can quote the area element vector a
,
d . To prove
that, let us consider Figure 4.3, where the area element in the rotated current configuration
is defined as:
a
x x x x
x x X F X F
X F X F x x a
,
, , , ,
, ,
, ,
, ,
, , ,
d
d d d d
d d d d
d d d d d
T

r r
r r
r r
-
Q
Q Q Q
Q Q Q Q
) )( ( ) )( (
) ( ) (
) 2 ( ) 1 ( ) 2 ( ) 1 (
) 2 ( ) 1 ( ) 2 ( ) 1 (
) 2 ( * ) 1 ( *
) 2 (
*
) 1 (
* *
cof

(4.3)
Hence, we have demonstrated that a
,
d is objective.















Figure 4.3: Differential area element.
The velocity is the rate of change of displacement and is not objective and neither is the
acceleration. To prove this, let us consider a homogeneous motion represented by:
X F x
,
,
,
- c (4.4)
Its velocity is obtained as follows:
X F v x
,
`
`
,
,
`
,
- = c
(4.5)
F
Q

*
F

) 1 (
X
,
d
Reference configuration
Rotated current configuration

) 2 (
X
,
d
A
,
d

) 1 (
x
,
d

) 2 (
x
,
d
a
,
d

) 2 (
*
x
,
d

) 1 (
*
x
,
d
*
a
,
d
Current configuration
NOTES ON CONTINUUM MECHANICS

272
and by applying an orthogonal transformation we obtain:
X F v x
,
`
`
,
,
`
,
- Q c Q Q Q
(4.6)
We can also define that
X F X F x
X F X F x
X F X F x X F x
,
`
,
`
`
,
`
,
, ,
`
,
`
,
,
`
,
,
`
`
,
`
,
,
,
,



- -
- -
- - -
Q Q c
Q Q c
Q c c c
*
*
* * * *
) ( ) (
) (
Dt
D
Dt
D
Dt
D

(4.7)
If we compare the last line of the equation (4.7) with (4.6) we can conclude that the
velocity is not objective as the only way to achieve this is when 0 c
,
`
,
and 0 Q
`
.
Then the rate of change of (4.5) provides us the acceleration, i.e.:
X F a x
,
` `
` `
,
,
` `
,
- = c
(4.8)
Then by applying an orthogonal transformation we obtain:
X F a x
,
` `
` `
,
,
` `
,
- = Q c Q Q Q
(4.9)
Similarly, by applying the rate of change in (4.7) and we obtain:
X F X F X F X F x
,
` `
,
` `
,
` `
,
` `
` `
,
` `
,
- - - - Q Q Q Q c
*

(4.10)
If we compare the equations (4.9) with (4.10) we conclude that the acceleration is not
objective as this will only be so if and only if 0 c
,
` `
,
and 0 Q
`
.
Second-order tensors
If ) , ( t x
,
A is an Eulerian second-order tensor produced by the motion F , then A is
objective if
*
A is related to A by:
T
t t Q A Q A ) , ( ) , (
* *
x x
, ,
The Eulerian second-order tensor (4.11)
Another special case is the two-point tensor which is neither in the current nor in the
reference configuration. In this case the orthogonal transformation is characterized by:
A Q A
*
The two-point tensor (4.12)
The Lagrangian second-order tensor, ) , ( t X
,
A , is objective when the following condition is
satisfied:
) , ( ) , (
*
t t X X
, ,
A A The Lagrangian second-order tensor (4.13)
Note that, the reference configuration has not been rotated, (see Figure 4.4).
4.2.1 The Deformation Gradient
As we saw in Chapter 2, the deformation gradient (two-point tensor) relates line elements
between reference and current configurations, i.e. X F x
,
,
d d , (see Figure 4.4). With
components this relation becomes:
j
j
i
j ij i
dX
X
x
dX F dx
o
o

(4.14)
4 THE OBJECTIVITY OF TENSORS

273
Additionally, we can define the deformation gradient in the rotated current configuration
as:

o
o
o
o

o
o

kj ik
j
k
k
i
j
i
ij
F
X
x
x
x
X
x
F Q
* *
*
F F Q
*
(4.15)
Here we can notice that F is objective, since F is a two-point tensor.
Now, if we start from F F Q
*
we can prove that X F x
,
,
d d is also objective, i.e.:
x X F X F x
,
, ,
,
d d d d Q Q
* *
(4.16)
It is now simple to show, (see Figure 4.4), that the following relationship is valid:
F F F F Q Q Q
T T *
(4.17)













Figure 4.4: Deformation gradient, objectivity.
The inverse of (4.15) is given by:
( )
-
-

1
1
*
F F Q
T
Q
-
-

1
1
*
F F

(4.18)
If we start from the definition of the Jacobian determinant ) (F det J , then the following
is valid:
J J

) ( ) ( ) ( ) ( ) (
1
* *
F F F F det det det det det
_
Q Q
(4.19)
4.2.2 Kinematic Tensors
Taking into account the rotated current configuration, the right Cauchy-Green deformation
tensor ( F F X C
T
t) , (
,
) is defined as:
( ) ( ) C F F F F F F F F C
T T T T
T
Q Q Q Q
* * *
(4.20)

*
B
x
,

Q
F
Reference
configuration
Current
configuration

0
B
B
*
X X
, ,
d d =
x
,
d
X F x
,
,
d d
F F Q
*


*
x
,
d

*
x
,

x x
, ,
d d Q
*
( ) X F
X F x
,
, ,
d
d d

Q
* *


T
Q

*
X X
, ,
=
Rotated current configuration
NOTES ON CONTINUUM MECHANICS

274
Then it follows from the above equation that the Green-Lagrange strain tensor,
( ) 1 - C X E
2
1
) , ( t
,
, is given by the following transformation law:
( ) ( ) E C C E - - 1 1
2
1
2
1
* * *
(4.21)
Note that the tensors C and E are objective, since they are defined in the reference
configuration.
Given the left Cauchy-Green deformation tensor,
T
t F F x b ) , (
,
, we can conclude that:
( ) ( )
T T T T
T
Q Q Q Q Q Q b F F F F F F b
* * *
(4.22)
We can find the same result if we start from the following definition,
1 -
F C F b , (see
Chapter 2), i.e.:
( ) ( )
T
T T
T
T
T T
Q Q
Q Q
Q Q
Q Q






- -
-
-
-
-
b
F F
F F F F
F F F F
F C F
F F F F F F
F C F b
1 1
1
1
* *
* *
1
* * * *
1
* * * *

T
T
T
Q Q
Q Q
Q Q



-
- - -
-
-

1
1 1 1
1
1
*
) (
) (
b
b
b b

(4.23)
Consequently, the Almansi strain tensor ( )
1
2
1
) , (
-
- b x e 1 t
,
in the rotated current
configuration becomes:
T
Q Q e e
*
(4.24)
Starting from the polar decomposition defined in Chapter 2, (see Figure 4.5), it follows
that:
* * * * *
; R V U R R V U R F F (4.25)
where R is the polar decomposition (rotation tensor) proper orthogonal tensor, ) , ( t X
,
U is
the right stretch tensor, and ) , ( t x
,
V is the left stretch tensor. Then by taking into account
that
2
) , ( U F F X C
T
t
,
and the equation in (4.20) we deduce that:
U U U U
* 2
2
* *
C C
(4.26)
Then, taking as a starting point the equation F F Q
*
we can obtain:



*
* * U U
U R Q U R
R Q R
*
T
R R Q
*

(4.27)
Thus, R is an objective two-point tensor.
It is also true that:
T T
Q Q V Q Q V
-
- -

1
2
*
1
*
2
* *
b b b b
(4.28)
If we use the left polar decomposition we have:

* * * * *
R V Q R V F F
* *
R V R V Q (4.29)
4 THE OBJECTIVITY OF TENSORS

275
Moreover, if we bear in mind that
T
R R Q
*
, (see Eq. (4.27)), we can conclude that:
T
T T
Q V Q V
Q V V Q R R V R R V Q R V R V Q




*
* * * * *
(4.30)
Therefore, tensors F , R , C , U, E , b , V , e are objective.





















Figure 4.5: Objectivity of the kinematic tensors.
4.2.3 Stress Tensors
The Cauchy stress tensor, ) , ( t x
,
, is objective, since it is true that:
T
Q Q
*

(4.31)
We can prove the equation (4.31) based on the following equation
)

(
* * *
*

n
t n
,
defined
in the rotated current configuration
*
t
B , (see Figure 4.6):
)

(
)

( * )

( *
)

(
* * *


*
n
n

n
n
t n
t n Q Q t Q n Q t n
,
,
_
, ,

T
(4.32)

*
B
X
,

x
,

Q
F
Reference
configuration
Current
configuration

0
B
B
F F Q
*


*
x
,

1
* * *
-
F C F b

*
2
* * 2
E E
C C

U U


1
, ,
-
F C F b
e b V


T
T
Q Q
Q Q

e e
b b
*
*

U U
*

R Q R
*

B B =
*


T
Q

*
R Q R
T

Rotated current
configuration
NOTES ON CONTINUUM MECHANICS

276
The first Piola-Kirchhoff stress tensor, defined in Chapter 3, is given by the equation
T -
F F P ) ( det . Then, P can be defined in the rotated current configuration as:
T -

* * * *
) ( F F P det
(4.33)
Then if we consider that
T
Q Q
*
and F F Q
*
, the equation in (4.33) becomes:

-
-
-
-
-





_
P
Q
Q Q Q
Q Q Q Q
Q Q Q Q
P
T
T T
T T T
T T
T
F F
F F
F F
F F
F F
) (
) (
) ( ) ( ) (
) ( ) (
) (
1
* * * *
det
det
det det
det
det

P Q P
*
(4.34)
Note that the first Piola-Kirchhoff stress tensor is a two-point tensor whose
transformation is defined according to the transformation law of vectors, (see Figure 4.6),
hence, the first Piola-Kirchhoff stress tensor is objective.


















Figure 4.6: Objectivity of the stress tensors.
Then if we refer to the definition of the second Piola-Kirchhoff stress tensor, ) , ( t X
,
S , we
can conclude that:

- -
- -
- -



T
T T T
T
F F F
F F F
F F F

Q Q Q Q Q
S
1
1
* *
1
* * *
) (
) (
) (
det
det
det
S S
*
(4.35)
dA
)

(
0
N
t
,

N



A
,
d


0
B
da
)

(
*
*
n
t
,


*

n

*
t
B
Q
F
F F Q
*

J


*
*
*
T T
M M
S S



T
J
-
F P
P Q P
*


* * *
*

Q Q
J
T


a
,
d

t
B
da
)

(n
t
,

Reference configuration
Current configuration
Rotated current configuration
4 THE OBJECTIVITY OF TENSORS

277
The tensor S S
*
is defined in the reference configuration, so it is objective.
Now, if we consider the Kirchhoff stress tensor, ) , ( J t x
,
, it follows that in the rotated
current configuration we obtain:
T T T
J J J Q Q Q Q Q Q
* * *
(4.36)
We can notice from the above that is also objective.
Similarly, we can verify the objectivity of the Mandel ( P M
T
F ) and Biot ( S U T )
stress tensors:
( ) ( )
) , (
* * *
t
T
T
T
X
F
F F
,
M
P
P Q Q P M



) , (
* * *
t X
,
T
S U
S U T


(4.37)
4.3 Tensor Rates
Before introducing objective rates we need to evaluate some equations that will be useful in
generating these.
The material time derivative of F F Q
*
is given by:
F F F
` ` `
- Q Q
*
(4.38)
Additionally, the material time derivative of the inverse
T
Q
-
-

1
1
*
F F becomes:
T T
Q Q
- -
-
-
1 1
1
*
F F F
` ` `

(4.39)
Then the spatial velocity gradient
1 -
F F
`
l can be evaluated in the rotated current
configuration as follows:
( ) ( )
T T T
Q Q Q Q Q Q Q
- - -
-
- -
1 1 1
1
* * *
F F F F F F F F F
` ` ` ` `
l (4.40)
T T
Q Q Q Q - l l
`
*

(4.41)
Hence, the tensor l is not objective due to the additional term (
T
Q Q
`
) that appears in
the equation (4.41) by means of which we can obtain Q
`
as:
T T
T
T T
Q Q Q Q Q Q - - l l l l
* *
` `

(4.42)
Then if we consider that 1 Q Q
T
, we obtain:
( ) ( )
T T T
Dt
D
Dt
D
Q Q Q Q 0 1 Q Q
` `
- (4.43)
Afterwards the equation in (4.41) can still be rewritten as:
T T
Q Q Q Q - - l l
`
*

(4.44)
Then if we use (4.44) we can obtain another way to express the rate of Q:
NOTES ON CONTINUUM MECHANICS

278
* *
; l l l l - -
T T T
T
T
Q Q Q Q Q Q
` `

(4.45)
Let us now consider the symmetric part of
*
l , which by definition is
*
D , i.e.:
|
.
|

\
|
- =
T sym
* * * *
2
1
l l l D (4.46)
Then by substituting
*
l , given in the equation (4.41), into the above equation (4.46), we
obtain:
( ) ( ) [ ]
T T T T T
Q Q Q Q Q Q Q Q D - - - l l
` `
2
1
*
(4.47)
If we then bear in mind the equation in (4.43), the above becomes:
[ ] [ ]
T T T T T
Q Q Q Q Q Q D - - l l l l
2
1
2
1
*
(4.48)
T
Q D Q D
*

(4.49)
Thus, the rate-of-deformation tensor ( D ) is objective.
If we now return to the equation in (4.41) and consider that l has been broken down into
a symmetric ( D : rate-of-deformation tensor) and an antisymmetric ( W: spin tensor) part,
i.e. W D- l , we can state that:
T T T
T T
T T
Q W Q Q D Q Q Q
Q W D Q Q Q
Q Q Q Q W D



- -
- -
- -
`
`
`
) (
* * *
l l
(4.50)
and if we consider
T
Q D Q D
*
, we can obtain
*
W as:
T T
Q W Q Q Q W -
`
*

(4.51)
Thus,
*
W is not objective since
T
Q W Q W
*
. Therefore, from the equation in (4.51)
we can obtain:
W Q Q W Q W Q W Q Q Q - - -
* *
` `
T T
(4.52)
Moreover, if we also consider that ( )
T
T
T T
Q Q Q Q Q Q
` ` `
- , the above equation
becomes:
* *
W Q Q W Q W Q W Q Q Q Q Q - - -
T T T T T T
` ` `
(4.53)
4.3.1 Objective Rates
If we consider an arbitrary vector a
,
, then the orthogonal transformation is given by:
a Q a
, ,

*
(4.54)
whose rate becomes:
a Q a Q a
`
, ,
`
`
,
-
*

(4.55)
4 THE OBJECTIVITY OF TENSORS

279
The above proves that the rate of change of a
,
is not objective, since an additional term
( a Q
,
`
) appears in the above equation.
As we have seen before, a second-order tensor that is defined in the current configuration
is objective if it holds that:
T
Q A Q A
*
(4.56)
Then the material time derivative of
T
Q A Q A
*
can be evaluated as follows:
T T T
Q A Q Q A Q Q A Q A
` ` ` `
- -
*
(4.57)
Thus, we can conclude that A
`
is not objective, since
T
Q A Q A
` `
*
. We can then define
some objective rates.
4.3.1.1 The Convective Rate
Let us now consider the expression of Q
`
, given in (4.45), and by substituting it into the
equation in (4.55) we obtain:
( )
C C
a Q a
a a Q a a
a Q a Q a Q a
a Q a Q a Q a Q a Q a
, ,
,
`
, ,
`
,
`
, , ,
`
,
`
, , ,
`
, ,
`
`
,

- -
- -
- - - -
*
* * *
* *
* *
T
T
T
T
T
T
l l
l l
l l

(4.58)
The rate (
C
-
,
) indicates the convective rate with which we can introduce a new vector rate
C
a
,
,
which is objective and is defined as:
a a a
,
`
, ,
-
T
l
C
The convective rate (4.59)
4.3.1.2 The Oldroyd Rate
If we use the equation given in (4.42), i.e. l l - Q Q Q
*
`
and
T T
T
T T
Q Q Q - l l
*
`
,
and by substituting them into the equation in (4.57) we obtain:
( ) |
.
|

\
|
- - - -
T T
T
T T T
Q Q A Q Q A Q Q A Q Q A l l l l
* * *
` `

(4.60)
or
( )
( )
T
T T
T
T T
T
T T
T T T T
T
T T
Q A Q A
Q A A A Q A A A
Q A A A Q Q A Q Q A Q A
Q A Q Q A Q Q A Q Q A Q Q A Q A
A A



- - - -
- - - -
- - - - -

*
* * * * *
*
* *
* *
* * *
l l l l
l l l l
l l l l
` `
`
_ _
`
` `
(4.61)
which defines a new objective rate known as the Oldroyd rate:
T
l l - - A A A A
`

The Oldroyd rate (4.62)


NOTES ON CONTINUUM MECHANICS

280
Problem 4.1: Obtain the Oldroyd rate of the left Cauchy-Green deformation tensor ( b ).
Solution:
Based on the definition of the Oldroyd rate in (4.62), we can obtain the Oldroyd rate of b
as
T
l l - - b b b b
`

. Then if we refer to
T
F F b , the material time derivative of b
becomes ( ) ( )
T
T
T T T
l l l l - - - b b F F F F F F F F b
` ` `
. Thus, we can
conclude that 0 - - - - -
T T T
l l l l l l b b b b b b b b
`

.
4.3.1.3 The Cotter-Rivlin Rate
If instead of using the equation related to Q
`
and
T
Q
`
given in (4.42), we use the equations
given in (4.45), and by substituting them into (4.57) we obtain:
( )
( )
( )
T
T T
T
T T T T
T
T T T T T T
T
T T T T
T
T
Q A Q A
Q A A A Q A A A
Q A A A Q Q A Q Q A Q A
Q A Q Q A Q Q A Q Q A Q Q A Q A
Q Q A Q Q A Q Q A Q Q A
A A





A A

- - - -
- - - -
- - - -
- - - |
.
|

\
|
-
*
* * * * *
*
* *
* *
* * *
* * *
l l l l
l l l l
l l l l
l l l l
` `
`
_ _
`
` `
` `

(4.63)
Thus, we can obtain a new objective rate, the Cotter-Rivlin rate:
l l - -
A
A A A A
T
`
The Cotter-Rivlin rate (4.64)
Problem 4.2: Obtain the Cotter-Rivlin rate of the Almansi strain tensor ( e ) in terms of the
rate-of-deformation tensor ( D ).
Solution: Based on the definition of the Cotter-Rivlin rate in (4.64), the Cotter-Rivlin rate of
e is l l - -
A
e e e e
T
` . Remember that in Chapter 2 we saw that the rate of the Almansi
strain tensor ( e` ) is related to the rate-of-deformation tensor D by the equation:
l l l l - - - - e e e e e e
T T
D D ` `
And by substituting e` into the Cotter-Rivlin rate
A
e we obtain D
A
e .
4.3.1.4 The Jaumann-Zaremba Rate
If we now consider Q
`
given in (4.52) and
T
Q
`
given in (4.53), and substitute them into the
equation in (4.57) we obtain:
( )
* * * * *
W A A W Q W A A W A Q A - - - -
T
` `
(4.65)
Then by rearranging the above equation, we obtain:
( )
T T
Q A Q A Q W A A W A Q W A A W A - - - -

` ` * * * * * *

(4.66)
We can conclude by the above that the rate

A , called the Jaumann-Zaremba rate, is objective


and is given by:
4 THE OBJECTIVITY OF TENSORS

281
W A A W A A - -
`

The Jaumann-Zaremba rate (4.67)


Next, we can interrelate the rates

A ,

A and
A
A . To do this let us consider the following
equations:
T T
l l l l - - - - A A A A A A A A

` `

(4.68)
W A A W A A W A A W A A - - - -

` `

(4.69)
l l l l - - - -
A A
A A A A A A A A
T T
` `

(4.70)
By combining (4.68) and (4.69) we obtain:
( ) ( ) W A A W W D A A W D A A
W A A W A A A A
A A A W A A W A



- - - - - -
- - - -
- - - -
T
T
T

l l
l l

(4.71)
Then, we can connect the Jaumann-Zaremba rate to the Oldroyd rate by:
D A A D A A - -

Relationship between the Jaumann-


Zaremba rate and the Oldroyd rate
(4.72)
Afterwards, by combining (4.69) and (4.70) we obtain:
( ) ( ) W A A W W D A A W D A A
W A A W A A A A A A A W A A W A


- - - - - -
- - - - - - - -
A
A A
T
T T


l l l l
(4.73)
Then, we obtain the relationship between the Jaumann-Zaremba rate and the Cotter-Rivlin
rate:
D A A D A A - -
A

Relationship between the Jaumann-Zaremba


rate and the Cotter-Rivlin rate
(4.74)
Now by adding equations (4.72) and (4.74) we can reach the following conclusion:
A
- A A A

2

(4.75)
Problem 4.3: Let A be a symmetric second-order tensor. Prove that

A A A A : : 2 ) (
Dt
D
,
where

A is the Jaumann-Zaremba rate of A .


Solution:
A A A A A A
` `
: : : - ) (
Dt
D

If we now incorporate W A A W A A - -

`
into the above equation, we obtain:



` `
A A W A A A W A A A
W A A A W A A A A W A A A W A A
W A A W A A A W A A W A A A A A A A
: : : :
: : : : : :
: : : : :
2 ) ( 2 ) ( 2 2
) ( ) ( ) ( ) (
) ( ) ( ) (
- -
- - - - -
- - - - - -



Dt
D

NOTES ON CONTINUUM MECHANICS

282
where we have applied the commutative property of the double scalar product, i.e.
A B B A : : . Note that, due to the symmetry of A the following condition is satisfied:
) ( ) ( ) ( ) ( W A A A W A : :
ik ji jk kj ik ij
W A A A W A
4.3.1.5 The Green-Naghdi Rate (Polar Rate)
Let us now refer back to the Polar Decomposition of F , (see Chapter 2), in which we
obtained the following equation for the spin tensor:
( )
T T
R R R U U U U R W - -
- -
` ` `
1 1
2
1
(4.76)
If 0 U
`
the above equation is reduced to
T
R R W
`
. Therefore
T
R R W
`
l , and by
substituting this into the equation in (4.64) we obtain:
W A A W A
W A A W A A
A A
W


- -
- -


A
`
`

T
l

(4.77)
which is the same equation as that obtained in (4.67). Then, we can define the Green-Naghdi
rate, also known as the Polar rate or Green-McInnis rate, by:
( ) ( )
T T
R R A A R R A A - -
V
` ` `

Green-Naghdi rate or Polar rate or Green-
McInnis rate
(4.78)
4.3.2 The Objective Rate of Stress Tensors
The material time derivative of the first Piola-Kirchhoff stress tensor, (see Eq. (4.34)),
becomes:
P Q P Q P P Q P
` ` `
-
* *
(4.79)
Then by substituting Q Q Q -
T
T *
l l
`
, (see equation (4.42)), into the above equation
we obtain:

( ) P P Q P P
P Q P Q P Q P Q P Q Q P Q P Q P
P


- -
- - - |
.
|

\
|
- -
T
T
T
T
T
T
l l
l l l l
` `
` ` ` ` `
* * *
* * *
*
(4.80)
Note that the orthogonal transformation of P obeys the vector transformation law so the
rate ( ) P P -
T
l
`
is objective because it has the same structure as the convective rate
presented in (4.59).
If we now apply the material time derivative to
T
J
-
F P we obtain:
T T T
J J J
Dt
D
- - -
- - = F F F
`
`
` `
P P ) ( (4.81)
Additionally, if we take into account the material time derivative of the Jacobian
determinant, (see Chapter 2), we have:
F F C C E C
` ` ` `
: : :
T
J
J
J J J J J
- - -
-
1 1
2
) ( ) ( ) ( D W D Tr Tr Tr l (4.82)
4 THE OBJECTIVITY OF TENSORS

283
and l
- -
-
1 1
F F
`
. Then, the equation in (4.81) becomes:
T T T T
J J J
- - -
- - F F F l l P ) (
`
`
Tr (4.83)
or:
[ ]
T T
J
-
- - F l D P ) (
`
`
Tr

(4.84)
The Truesdell Stress Rate
Now let us consider the Kirchhoff stress tensor J and
* * * *
J J , so the
material time derivative of
*
becomes:
( )
T T T
J J
J J J J
Q Q Q Q Q Q D
D
`
`
`
` `
`
`
- - -
- -
*
* * * * *
) (
) (
Tr
Tr

(4.85)
where we have substituted
*

`
into the equation in (4.57). Then by substituting Q
`
and
T
Q
`

given in the equation (4.42), i.e. l l - Q Q Q
*
`
and
T T
T
T T
Q Q Q - l l
*
`
, we
obtain:

T T
T
T
T T T
T J
Q Q Q Q
Q Q Q Q Q Q D

Q Q


- -
- - -

l l
l l
*
* *
*
*
*
) (
_
`
_
`
Tr

(4.86)
( )
T T
T
J
Q D Q - - - - - l l l l
`
`
) (
* * * *
*
Tr


(4.87)
If we now bear in mind that
* * *
) ( D
`
` J J - Tr , the above equation becomes:
( )
T T
T
Q D Q D - - - - - - l l l l
` `
) ( ) (
* * * * * *
Tr Tr
(4.88)
Hence, we obtain a new objective stress rate called the Truesdell stress rate:
) (D Tr - - -
T
T
l l
`
The Truesdell stress rate (4.89)
Relationship between Objective Stress Rates
Let us consider the Oldroyd rate of the Cauchy stress tensor
T
l l - -
`

, and if
we use the Truesdell stress rate we can conclude that:
D ) ( Tr -
T

(4.90)
We can also relate the Oldroyd rate of the Kirchhoff stress tensor
T
l l - - `


with the Oldroyd rate of the Cauchy stress tensor
T
l l - -
`

as:
T
J


(4.91)
We can also prove the above equation is valid by starting from (4.87):
NOTES ON CONTINUUM MECHANICS

284
( )
( )
( )
T T
T
T T
T
T T
T
J
J J J
J
Q D Q
Q D Q
Q D Q



- - - - -
- - - - -
- - - - -
l l l l
l l l l
l l l l
`
`
`
`
`
`
) (
) (
) (
* * * * *
*
* *
*
*
* * * *
*
Tr
Tr
Tr


(4.92)
thus,
( )
T
- - - |
.
|

\
|
- - D Q Q J J
T
T
T

l l l l
`
_
` ) (
*
* * * * *
Tr
(4.93)
If we consider the equation of the second Piola-Kirchhoff stress tensor
T - -
F F
1
S ,
(see Chapter 3), the material time derivative becomes:
( )
T T
T T T T
T T T
- -
- - - - - -
- - - - - -



- -
- - -
- -
F F
F F F F F F
F F F F F F
l l
l l



`
`
`
` ` `
1
1 1 1
1 1 1
S

(4.94)
where we have considered that l
- -
-
1 1
F F
`
. If we refer to the Oldroyd rate of the
Kirchhoff stress tensor
T
l l - - `

the equation in (4.94) becomes:


T T
J
-
T
- - -
F F F F S
1 1

`

(4.95)
Now, taking the material time derivative of
* *
J we obtain:
( )
T T T T
J J
Dt
D
J J J J Q Q Q Q Q Q Q Q
`
`
` ` `
`
`
` - - - - -
* * * * *
) ( (4.96)
Also if we consider the relationship W Q Q W Q -
*
`
and ) (D Tr J J
`
in the above
equation we obtain:
( ) ( ) |
.
|

\
|
- - - - -
T
T T
J J W Q Q W Q Q Q Q W Q Q W D
* * * *
) (
`
` Tr
(4.97)
(
(
)
T T T
T
T
T T
T T
T
T
T T T
J J J J
J J
Q W Q Q Q
Q W Q D W Q Q Q Q W
Q W Q W Q Q
Q Q Q W Q Q Q W D




-
- - - - -
|
.
|
-
- - - -
`
`
`
`
* * * *
*
* * *
) (
) (
Tr
Tr


(4.98)
or
[ ]
T T
T
J
T
Q W W D Q W W
Q Q
- - - - -

`
_

` ) (
* * * * *
Tr


(4.99)
which give us the relationship between the Jaumann-Zaremba rate of the Kirchhoff stress
tensor and the Cauchy stress tensor:
[ ]
(

- - - -

`
D W W D ) ( ) ( Tr Tr J J
T


(4.100)
5 The Fundamental Equations of Continuum Mechanics












5.1 Introduction
The fundamental equations of continuum mechanics are based on the conservation
principles of certain physical quantities. We consider five of these to establish the basic
equations that govern the Initial Boundary Value Problem (IBVP), namely:
The principle of conservation of mass;
The principle of conservation of linear momentum;
The principle of conservation of angular momentum;
The principle of conservation of energy;
The principle of irreversibility.
In this chapter we will address the fundamental principle of mechanics in the reference and
current configurations. At the end of the Chapter we will show that these principles are
insufficient to establish the IBVP set of partial differential equations, so, it is necessary to
add certain equations to fully resolve this problem. Then, we will introduce some concepts
and theorems to develop the concepts in this chapter.
5.2 Density
Density, denoted by ) , ( t f x
,
, is a scalar function that measures the amount of a property
per unit volume around a material point ( x
,
) at a time t . One very important density
function is mass density, denoted by ) , ( t x
,
j , which measures the amount of mass per unit
5
The Fundamental Equations
of Continuum Mechanics
285 , Notes on Continuum Mechanics, Lecture Notes on Numerical
Methods in Engineering and Sciences 4, DOI 10.1007/978-94-007-5986-2_7,
International Center for Numerical Methods in Engineering (CIMNE), 2013
E.W.V. Chaves
NOTES ON CONTINUUM MECHANICS

286
volume. Another density function we can quote is energy density, which measures stored
energy per unit volume. The term specific will be used to denote the amount of the
property per unit mass.
5.2.1 Mass Density
Any continuous medium is caused by a positive scalar quantity called mass. It is assumed
that the mass is continuously distributed throughout the continuum.
We will next review the concept of mass density introduced in Chapter 2. Let us consider a
sphere of infinitesimal radius centered at point P in the reference configuration, (see
Figure 5.1). The material contained in this sphere is denoted by m A and the sphere volume
is represented by
0
V A . Then, the mass density
0
j , in the reference configuration, is
defined by the limit:
0 0
0
0
0
) ( ) 0 , (
dV
dm
V
m
im t
V

A
A

A

,
,
X x j j
(5.1)








Figure 5.1: Mass density.
Likewise, mass density in the current configuration is given by:
dV
dm
V
m
im t
V

A
A

A 0
) , (
,
x j j (5.2)
Then the functions ) , ( t x
,
j and ) (
0
X
,
j are continuous density functions and are
interrelated to each other, (see Chapter 2), by:
) , ( ) (
0
t J X x
,
,
j j
(5.3)
5.3 Flux
The properties conferred by density (e.g. mass, energy, entropy, etc.) are mobile and the rate
of change and direction of these quantities are assigned by the flux vector, usually denoted
by ) , ( t x
,
,
q . With this information, we can define the amount of property that passes
through a differential area element da per unit time, (see Figure 5.2), as:
da da da
n
cos

q o q n q
, ,
(5.4)

_
,
) (
0
0
0
X j

AV

2
x

3
x

1
x

0
) 0 ( t t =
t

_
,
) , (
0
t
V
x j

A
Reference configuration
Current configuration
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

287
where n

is the unit normal vector, and o is the angle formed between q


,
and
n
q

,
. Note
that, only the normal vector
n
q

,
crosses the surface, since the tangential vector
s
q

,
remains
on the surface da . As an example of flux, we can mention the mass flux vector which is
represented by v
,
,
j q . With regard to the SI unit we have: [ ]
s m
kg
2
q
,
where q
,
represents
the mass flux vector, and [ ]
s m
J
2
q
,
where q
,
refers to the energy flux vector.








Figure 5.2: Flux vector.
5.4 The Reynolds Transport Theorem
Let ) , ( t x
,
d be an Eulerian scalar field which describes a certain physical quantity per unit
volume. If ) , ( t x
,
d is continuous and differentiable, we can state that:
}
}
} }
(

d - d
(

d - d
(

d - d d

V
V
V V
dV t t
Dt
D
dV t t
Dt
D
dV
dV
Dt
D
t t
Dt
D
dV dV t
Dt
D
) , ( ) , (
) , ( ) , (
) ( ) , ( ) , ( ) , (
v x x
v x x
x x x
x
x
, , ,
, , ,
, , ,
,
,
V
V (5.5)
whose equivalent in indicial notation is:
} }
(

o
o
d - d d
V
k
k
V
dV
x
v
t t
Dt
D
dV t
Dt
D
) , ( ) , ( ) , ( x x x
, , ,
(5.6)
Then by using the material time derivative operator we can still state that:
( )
}
} }
(
(

d
o
o
-
o
d o

(
(

o
o
d -
o
d o
-
o
d o
d
V
p
p
V
k
k
p
p
V
dV v t
x t
t
dV
x
v
t
x
t
v
t
t
dV t
Dt
D
) , (
) , (
) , (
) , ( ) , (
) , (
x
x
x
x x
x
,
,
,
, ,
,

(5.7)
This last equation is known as the Reynolds transport theorem and can be represented by the
following equations:

1
x

2
x

3
x

t
B
q
,

n



n
q

,

da
o
NOTES ON CONTINUUM MECHANICS

288
( )
} }
}
} }

d -
o
d o

d -
o
d o

d - d d
t t
S V
V
V V
dS t dV
t
t
dV t
t
t
dV t t
Dt
D
dV t
Dt
D

) , (
) , (
) , (
) , (
) , ( ) , ( ) , (
n v x
x
v x
x
v x x x
x
x
, ,
,
, ,
,
, , , ,
,
,
V
V
The Reynolds transport
theorem
(5.8)
where
t
V is the control volume,
t
S is the control surface, and n

is the outward unit


normal to the boundary
t
S of
t
B . The first term on the right of equation is the local rate
of change of the property d in the domain
t
V , while the second term characterizes the
transport of v
,
d , that leaves the domain
t
V via the surface
t
S , (see Figure 5.3).









Figure 5.3: Control volume.

Problem 5.1: Prove that Reynolds transport theorem is valid in the following equation:
} }
d d
0
0
V V
JdV
Dt
D
dV
Dt
D

(5.9)
Solution:
} } } }
|
.
|

\
|
d -
d
|
.
|

\
|
d -
d
|
.
|

\
|
d -
d
d
V V V V
dV
Dt
D
dV J
Dt
D
J dV
Dt
DJ
Dt
D
J JdV
Dt
D
v v
x x
, ,
, ,
V V
0 0 0
0 0 0



5.4.1 Reynolds Transport Theorem for Volumes with
Discontinuities
Let us consider a material volume that is intersected by a discontinuous surface ) (t Z , a
singular surface which is moving over time at velocity c
,
, (see Figure 5.4). The surface ) (t Z
divides the material volume into two parts viz.
-
B and
-
B . We can also define the
boundaries
-
Z and
-
Z situated ahead of and to the rear of the singular surface Z as shown
in Figure 5.4.
n
) ( v
,
c

t
t
o
o ) , ( x
,
c

x
,


t
V

t
S
[ ]n n v ) (
,
,
c
n
q
control surface

control volume
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

289

















Figure 5.4: Material volume with discontinuity.
The mobile discontinuity ) (t Z is defined by the surface equation:
) ( 0 ) , ( t t f Z =
Z
x x
, ,

(5.10)
Then, the unit normal vector n

on the surface ) (t Z is given by the equation:


Z
Z

f
f
t
V
V
) , (

x
,
n
(5.11)
For material points belonging to the surface ) (t Z , the normal component of the velocity,
n
c , is defined as:
Z
Z
o
o
- c
f
t
f
n
V

(5.12)
Then by combining the equation in (5.11) with (5.12) we obtain:

o
o
- c
Z
Z
Z
Z
c
V
V
V
c c
, , ,
f
f
f
t
f
n
n n

0 0 -
o
o
Z
Z
Z

Dt
Df
f
t
f
c V
,

(5.13)
Let A be a second-order tensor field and let us consider that
-
A and
-
A are the values of
A in the boundaries
-
Z and
-
Z , respectively. Then, we can define the jump of A as:
[ ] [ ]
- -
- A A A (5.14)
Then by applying Gauss theorem (the divergence theorem) for the two domains
-
B and
-
B we can obtain, respectively:

_

B
) (t Z

-
B
-
B

-
S
n



-
S


-
Z

-
Z
n


c
,

) (t Z

-
n



-
n


NOTES ON CONTINUUM MECHANICS

290
} } }
- - -
Z
- -
- dS dS dV
S V


n A n A A V
; } } }
- - -
Z
- -
- dS dS dV
S V


n A n A A V
(5.15)
Then, by summing up these two above equations we obtain:
} } } }
- - - - - -
Z
- -
Z
- -
- -
- - dS dS dS dV
S S V V


n A n A n A A V
(5.16)
Additionally, if we bear in mind that n n n

- -
- -
and [ ] [ ]
- -
- A A A the above
equation becomes:
[ ] [ ]
} } }
Z - -
-
- - - -
dS dS dV
S S V V


n A n A A V
(5.17)
The Reynolds transport theorem can be modified for the case in which there is a singular
surface ) (t Z , which moves at the velocity c
,
, (see Figure 5.4). Then by applying the
equation in (5.8) to the two domains
-
B and
-
B , whose contours are
- -
Z - S and
- -
Z - S , respectively, we obtain:
} } } }
} } } }
- - - -
- - - -
Z
- -
Z
- -


d - d -
o
d o
d
d - d -
o
d o
d
dS t dS t dV
t
t
dV t
Dt
D
dS t dS t dV
t
t
dV t
Dt
D
S V V
S V V

) , (

) , (
) , (
) , (

) , (

) , (
) , (
) , (
n n
n n
c
c
, , , ,
,
,
, , , ,
,
,
x v x
x
x
x v x
x
x

(5.18)
Then by adding the two equations above, and once again considering that n n n

- -
- -

and [ ] [ ]
- -
d - d d , we can conclude that:
[ ] [ ] ( )
} } } }
Z - - -
d - d -
o
d o
d
- - - - - -
dS dS t dV
t
t
dV t
Dt
D
S S V V V V


) , (
) , (
) , ( n n c
, , ,
,
,
v x
x
x
(5.19)
Additionally, by using the definition in (5.17) we can state that:
( ) [ ] [ ]
} } }
Z - -
d - d d
- - - -
dS t dV t dS t
V V S S

) , ( ) , (

) , ( n n v x v x v x
x
, , , , , ,
,
V
(5.20)
Then by combining the above equation with that in (5.19) we obtain:
( ) [ ] [ ] [ ] [ ] ( )
} } }
Z - -
d - d - |
.
|

\
|
d -
o
d o
d
- - - -
dS t dV t
t
t
dV t
Dt
D
V V V V

) , ( ) , (
) , (
) , ( n c V
, , , , ,
,
,
,
v x v x
x
x
x

(5.21)
which results in the Reynolds transport theorem for domains with discontinuities:
( ) ( ) [ ] [ ]
( ) [ ] [ ]
} } }
} } }
Z Z - Z -
Z Z - Z -


- d - |
.
|

\
|
d -
d
d
- d - |
.
|

\
|
d -
o
d o
d
dS dV t
Dt
t D
dV t
Dt
D
or
dS dV t
t
t
dV t
Dt
D
V V
V V

) , (
) , (
) , (

) , (
) , (
) , (
n
n
c V
c V
, , , ,
,
,
, , , ,
,
,
,
,
v v x
x
x
v v x
x
x
x
x
(5.22)
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

291
5.5 Conservation Law
Conservation Law when applied to a particular physical quantity per unit volume, in a part
of the domain, states that no physical quantity (mass density, energy density, etc.) can be
created or destroyed, but merely moves from one place to another. The conservation law in
global form (weak) is established from Reynolds transport theorem:
( )
} } (

d -
o
d o
d

V
sink or source
V
dV t
t
t
dV t
Dt
D
) , (
) , (
) , (
0
v x
x
x
x
, ,
,
_
,
,
V
(5.23)
If the term on the left of the equation is nonzero this means that somewhere in the domain
there is a property source or sink, which can be represented locally by the variable Q .
Then, 0 > Q indicates that there is a source, and 0 < Q that there is a sink. For example, if
the property in question is mass density (mass per unit volume) in general 0 Q . However,
if there is a tumor (cells with abnormal growth) in a biological organism, we can establish a
law (at the macroscopic level) that indicates how the mass changes over time (source),
without regard to individual cells. Then, another example of a source we can cite is the
internal heat generated by a chemical reaction, such as in cement hydration. The effect of
the chemical reaction at the macroscopic level can be represented by a variable that
provides the amount of heat generated per unit volume and per unit time (the internal heat
source).
Note that the term ) ( v
,
d shows the flux of the property d. Then, if ) ( v
,
d represents the
energy flux we have the following unit [ ]
s m
J
2
dv
,
, and if we are dealing with mass
transport we have [ ]
s m
kg
2
dv
,
. As we have seen before, in general, the flux is represented
by q
,
, with which we can establish the local form (strong) of conservation law and which is
denoted by the following continuity equation:
( ) ) , (
) , (
t
t
t
Q x
x
x
,
,
,
,
q -
o
d o
V Continuity equation
[ ]
s
d
(5.24)
where [ ] d is the SI unit of the physical quantity per unit volume.
5.6 The Principle of Conservation of Mass. The
Mass Continuity Equation
The law of conservation of mass states that the total mass of a continuum does not change.
This implies that the total mass in the reference configuration is equal to the total mass in
the current configuration:
} }

V V
dV dV m
0
0
j j

[ ] kg (5.25)
NOTES ON CONTINUUM MECHANICS

292
As a result of conservation of mass, the material time derivative of the total mass is zero,
i.e. 0 m
Dt
D
, then:
[ ] [ ] [ ]
[ ] 0 ) , ( ) , (
0 ) , ( ) , ( ) , ( ) , (

-
-
}
} } }

V
V V V
dV t t
Dt
D
dV
Dt
D
t t
Dt
D
dV dV t
Dt
D
dV t
Dt
D
m
Dt
D
v x x
x x x x
x
, , ,
, , , ,
,
V j j
j j j j
(5.26)
or in indicial notation:
[ ] 0 ) , (
(

o
o
-
}
dV
x
v
t
Dt
D
V
k
k
j j x
,
(

s
kg
(5.27)
If the above equation is valid for the entire domain, then it must also be satisfied locally:
0
,
-
k k
v
Dt
D
j
j

(

3
sm
kg
(5.28)
which is the mass continuity equation in Eulerian description and is expressed in tensorial
notation as:
0 ) ( - v
x
,
,
V j
j
Dt
D
The mass continuity equation
(Eulerian description)
(5.29)
Then by applying the material time derivative operator, i.e.
k
k
v
x t Dt
D
o
- o
-
o
- o

-
, the mass
continuity equation (5.28) becomes:
0 ) ( ) (
, ,
-
o
o

o
o
-
o
o

o
o
-
o
o
-
o
o
-
k k k
k k
k
p
p k k
v
t
v
x t x
v
x
v
t
v
Dt
D
j
j
j
j
j
j j
j
j

(5.30)
Hence, we have another way to express the mass continuity equation:
0 ) ( -
o
o
v
x
,
,
j
j
V
t
The mass continuity equation
(Eulerian description)
(5.31)
We could have obtained the same equation in (5.31) by means of Reynolds transport
theorem, i.e. in the equation (5.8) we substitute ) , ( t x
,
d for ) , ( t x
,
j , which means:
( ) ( ) 0 ) , (
) , (
0 ) , (
) , (
) , ( -
o
o

(

-
o
o

} }
v x
x
v x
x
x
x x
, ,
,
, ,
,
,
, ,
t
t
t
dV t
t
t
dV t
Dt
D
V V
j
j
j
j
j V V (5.32)
We could also have obtained the mass continuity equation in (5.29) by means of the
principle of conservation of mass in a differential volume element
3 2 1
dx dx dx , (see Figure
5.5), in which the following is satisfied:



The rate of mass entering through face A is represented by the mass flux
3 2 1
1
) ( dx dx v
x
j ,
while the rate of mass that goes through face B is given by
3 2 1
1
1
1
) (
dx dx dx
x
v
v
(

o
o
-
j
j .
Mass
accumulation
Inward mass
flux
Outward mass
flux -
=
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

293
Likewise, we can obtain the rate of change of mass in other faces. Moreover, by applying
the conservation of mass for the differential volume element we obtain:
( )

o
o
- -
(

o
o
-

-
(

o
o
- - - -
o
o
2 1 3
3
3
3 3 1 2
2
2
2
3 2 1
1
1
1 2 1 3 3 1 2 3 2 1 3 2 1
) ( ) (
) (
dx dx dx
x
v
v dx dx dx
x
v
v
dx dx dx
x
v
v dx dx v dx dx v dx dx v
t
dx dx dx
j
j
j
j
j
j j j j
j

(5.33)
Then by simplifying the above equation we obtain
3
3
2
2
1
1
) ( ) ( ) (
x
v
x
v
x
v
t o
o
-
o
o
-
o
o
-
o
o j j j j
and
by using the chain rule of derivative we find that:
( ) 0
) (
3
3
2
2
1
1
3
3
2
2
1
1
-
|
|
.
|

\
|
o
o
-
o
o
-
o
o

|
|
.
|

\
|
o
o
-
o
o
-
o
o
-
o
o
-
o
o
-
o
o
-
o
o


v
x
v v
,
_
_
,
, ,
V
V V
j
j
j
j j
j
j j j j
j
Dt
D
x
v
v
x t
x
v
x
v
x
v
v
x
v
x
v
x t
i
i
i
i
Dt
D
Tr
(5.34)










Figure 5.5: Conservation of mass in a differential volume element.
5.6.1 The Mass Continuity Equation in Lagrangian
Description
The mass continuity equation in (5.29) can also be expressed in Lagrangian description
(material). To do this, we can start from the conservation of mass which establishes:
} } }

0 0
0 0 0
) , ( ) , ( ) (
) (
V V V
dV J t dV t dV
f
_
, ,
,
,
X
x x X j j j
(5.35)
Since the above equation is valid for any volume it means that it will be valid locally too,
i.e.:
j j ) (
0
J X
,
(5.36)
Note that
0
j is independent of time and so is j J which results in the Lagrangian
description of the mass continuity equation:

3
dx

1
x
1
dx
face A
face B

3 2 1
dx dx v j

1
x

2
x

3
x

2
dx

3 2 1
1
1
1
) (
dx dx dx
x
v
v
(

o
o
-
j
j

NOTES ON CONTINUUM MECHANICS

294
0 ) ( J
Dt
D
j
The mass continuity equation
(Lagrangian description)
(5.37)
Problem 5.2: Show that
} }

V
ij
V
ij
dV
Dt
t DP
dV t P
Dt
D
) , (
) , (
x
x
,
,

j j (5.38)
where ) , ( t P
ij
x
,

is a continuum property per unit mass, which can be a scalar, a vector or


higher order tensor.
Solution: It was proven in equation (5.6) that:
dV
x
v
t t
Dt
D
dV t
Dt
D
V
p
p
V
} }
(
(

o
o
d - d d ) , ( ) , ( ) , ( x x x
, , ,

Then by making
ij
P j d , and by considering it in the above equation we obtain:
dV
x
v
Dt
D
P P
Dt
D
dV
x
v
P
Dt
D
P P
Dt
D
dV
x
v
P P
Dt
D
dV P
Dt
D
V
equation continuity mass
k
k
ij ij
V
k
k
ij ij ij
V
p
p
ij ij
V
ij
}
} } }
(

(
|
|
.
|
o
o
-

\
|
-

o
o
- -
(
(

o
o
-


) (
0
_
. .
. . . . . .
j
j
j
j
j
j j j j

Thus, we can conclude that:
dV
Dt
DP
dV P
Dt
D
V
ij
V
ij

} }
(
(

.
.
j j
Problem 5.3: Prove that the following relationship is valid:
) ( ) ( v v v a
x
, , , ,
,
-
o
o
j j j V
t
(5.39)
Solution: Based on the Reynolds transport theorem:
} } }
d -
o
d o
d
S V V
dS dV
t
dV
Dt
D
)

( n v
,

and if we consider that v
,
j d we obtain:
} } }
-
o
o

S V V
dS dV
t
dV
Dt
D
)

(
) (
n v v
v
v
, ,
,
,
j
j
j
Then, the above equation in indicial notation becomes:
} } } } } }
-
o
o
-
o
o

S
k k i
V
i
V
i
i
S
k k i
V
i
V
i
dS v v dV
t
v
dV v
Dt
D
dS v v dV
t
v
dV v
Dt
D
a

) (
) (
)

(
) (
n n j
j
j j
j
j
_

Additionally, by applying the divergence theorem to the surface integral we obtain:
} } } } (

-
o
o
-
o
o

V
k k i
i
V
k k i
V
i
V
i
dV v v
t
v
dV v v dV
t
v
dV a
, ,
) (
) (
) (
) (
j
j
j
j
j
which in tensorial notation is:

-
o
o

} }

V V
dV
t
dV ) (
) (
v v
v
a
x
, ,
,
,
,
j
j
j V ) (
) (
v v
v
a
x
, ,
,
,
,
-
o
o
j
j
j V
t

5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

295
5.6.2 Incompressibility
Compressibility is the ability to change the volume of a continuous medium. It is common
knowledge that gases are more compressible than liquids, but for practical purposes, the
liquid can be considered to be incompressible.
An incompressible medium is characterized by an isochoric motion, i.e. 1 J , hence the mass
density field (for all particle) is independent of time. In this case the mass continuity
equation in (5.29) boils down to:
( ) 0 0 ) (
,
- - -
k k
v
Dt
D
Dt
D
j j
j
j
j
v v
x x
, ,
, ,
V V (5.40)
thus
0 v
x
,
,
V
Mass continuity equation for a
incompressible medium
(5.41)
Thus, an incompressible medium can be characterized by:
[ ] 1 ; ; 0 ; 0 ) (
0
= = J
Dt
D
J
Dt
D
j j j
j
`
`
F det (5.42)
or
0 ) ( ) ( 0
3
3
2
2
1
1
,

o
o
-
o
o
-
o
o
D Tr Tr l
x
v
x
v
x
v
v
k k

(5.43)
where l denotes the spatial velocity gradient, and D is the rate-of-deformation tensor
which is equal to the symmetrical part of l , (see Chapter 2).
5.6.3 The Mass Continuity Equation for Volume with
Discontinuities
Now, let us consider a domain where there is a singular surface ) (t Z as established in
subsection 5.5.1, (see Figure 5.4). Based on the conservation of mass we have:
0
}
V
dV
Dt
D
j (5.44)
and if we consider the Reynolds transport theorem with discontinuities, (see equation
(5.22)), in which j d , we obtain:
( ) [ ] [ ] 0

- - |
.
|

\
|
-
} } }
Z - -

Z Z
dS dV
Dt
D
dV
Dt
D
V V
n c V
, , ,
,
v v
x
j j
j
j (5.45)
where the mass density ) , ( t x
,
j , and the velocity ) , ( t x v
, ,
are continuous differentiable
functions in Z - V , and ( ) [ ] [ ] c
, ,
- v j is also a continuous differentiable function on Z. The
global balance law is valid for any arbitrary parts of the volume and for the discontinuous
surface, hence it holds that:
( ) [ ] [ ] Z -
Z - -

on
V in
Dt
D
0

0
n c
V
, ,
,
,
v
v
x
j
j
j
The mass continuity equation
with discontinuities
(Eulerian description)
(5.46)
NOTES ON CONTINUUM MECHANICS

296

Problem 5.4: Let us consider the following velocity field:
t
x
v
i
i
-

1
for 0 t
1) Find the mass density field;
2) Prove that this motion satisfies
3 2 1 0 3 2 1
X X X x x x j j .
Solution: 1) By applying the mass continuity equation we obtain:
k
k
k
k
x
v
dt
d
Dt x
v
Dt
D D
o
o
- =
o
o
- j
j j
j
j
0
and by using the given velocity field, we find that:
t t x
x
t x
v
ii
i
i
i
i
-

o
o
-

o
o
1
3
1 1
1 c

Thus,
t
dt d
t dt
d
-
-
-
-
1
3
1
3
j
j j j

Then by integrating the both sides of the above equation we obtain:

-
-
} }
t
dt d
1
3
j
j
C t - - - ) 1 ( 3ln lnj
The constant of integration C is obtained by means of the above equation if we refer to
the initial condition 0 t , in which
0
) 0 , ( j j t x
,
, thus
0 0
) 0 1 ( 3 j j ln ln ln - - - C C
|
|
.
|

\
|
-
-
|
|
.
|

\
|
-
- - -
3
0
0
3
0
) 1 ( ) 1 (
1
) 1 ( 3
t t
t
j
j j j ln ln ln ln ln ln
Thus, we can conclude that:
( )
3
0
1 t -

j
j
2) Then by using the velocity definition we obtain:
t
x
dt
dx
v
i i
i
-

1

t
dt
x
dx
i
i
-

1

Additionally, by integrating the both sides of the above equation we obtain:

} }
t
dt
x
dx
i
i
1
i i
K t x - - ) 1 ( ln ln
(5.47)
Then by applying the initial condition, i.e. at time 0 t
i i
X x , we obtain:
i i i i
X K K X ln ln ln - - ) 0 1 (
Additionally, by substituting the value of
i
K into the equation (5.47) we obtain:
[ ] ) 1 ( ) ( ) 1 ( t X x X t x
i i i i
- - - ln ln ln ln ln
Hence we can conclude that ) 1 ( t X x
i i
- , which gives us ) 1 (
1 1
t X x - , ) 1 (
2 2
t X x - ,
) 1 (
3 3
t X x - , and if we consider that
( )
3
0
1 t -

j
j , we obtain:
( )( )( ) - - -
0
3
3
2
2
1
1
1 1 1 j j
_ _ _
X
x
X
x
X
x
t t t
3 2 1 0 3 2 1
X X X x x x j j
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

297
5.7 The Principle of Conservation of Linear
Momentum. The Equations of Motion
5.7.1 Linear Momentum
Let us consider the body,
t
B , in motion which is subjected both to body forces (per unit
mass), ) , ( t x
,
,
b , and to surface forces, ) , (
*
t x
,
,
t , acting on the surface

S , (see Figure 5.6).


Let ) , ( t x v
, ,
be the Eulerian velocity field, then we can define the linear momentum of the
mass system
t
B as:
} }

V
dV dm
t
v v L
, ,
,
j
B
Linear momentum
(

s
m kg
(5.48)









Figure 5.6: Continuum in motion.
5.7.2 The Principle of Conservation of Linear Momentum
The principle of conservation of linear momentum, based on Newtons second law, states
that the rate of change of the linear momentum of an arbitrary part of a continuous
medium is equal to the resultant force (body and surface forces) acting on the part in
question, then:
} } }
-
V V S
dV
Dt
D
dV dS v
,
, ,

*
j j b t


} } }
-
V V
i i
S
i
dV v
Dt
D
dV dS
*
j j b t


(5.49)
The equation in (5.49) represents the global form of the principle of conservation of linear
momentum and by applying
j ij i
n t

*
o we obtain:
} } }
}
p
}
o
- o
V
V
i
V
i i
Theorem
Guass
V
j ij
S
j ij
dV v
dV
dV v
Dt
D
dV dS
_
_

`

'
,

j jb n


(5.50)
thus,

t
B

S
) ,

, (
*
t n t x
,
,

n



1
x

2
x

3
x
x
,

) , ( t x
,
,
b j
) , ( t x v
, ,
j
dV
O
NOTES ON CONTINUUM MECHANICS

298
i
V
i i j ij
dV v 0 b - - o
}
) (
,
` j j
(

N
s
m kg
2
(5.51)
If the above equation is valid for the entire volume, it is also valid locally, i.e.:
i i i j ij
v 0 b - - o ` j j
,

(


m
Pa
m
N
m s
kg
3 2 2

(5.52)
which are known as the equations of motion or Cauchys first equation of motion:
a v
x
,
`
,
,
,
j j j - b V
The equations of motion
(Eulerian description)
(5.53)
Sometimes it is useful to express the equations of motion in the reference configuration.
To do this we can rewrite the equation in (5.49) in the undeformed configuration, i.e.:
} } } }
} } } }
-
-

0 0 0 0
0 0 0 0
0 0 0 0 0 0 0 0
0 0 0 0 0 0
*
0


V V V S
V V V S
dV dV
Dt
D
dV dS
dV J dV J
Dt
D
dV dS
A V
A V
, ,
,
, ,
, ,
j j j
j j j
b N P
b t

(5.54)
where ) , ( t X v V
,
,
,
= is the Lagrangian velocity, ) , ( t X a A
,
,
,
= is the Lagrangian acceleration
field, P is the first Piola-Kirchhoff stress tensor,
T
J
-
F P , and ( ) t ,
0
X
,
,
b is the body
forces vector per unit mass in the undeformed configuration. Note that P N N P

, since
P is a non-symmetric tensor. Then by applying Gauss theorem (the divergence theorem)
to the surface integral we obtain:
( )
}
- -
0
0 0 0 0

V
dV 0 b P
, , ,
,
A
X
j j V
(5.55)
Then, the local form of the equations of motion in material description (Lagrangian) can be
expressed as:
( ) A F
A
X
X
, ,
, ,
,
,
0 0 0
0 0 0
j j
j j
-
-

b S
b P
V
V
The equations of motion
(Lagrangian description)
(5.56)
5.7.2.1 The Equilibrium Equations
In the exceptional cases when we have a static or quasi-static equilibrium, the acceleration
components are zero, thus we obtain the equilibrium equations as:
0 b
, ,
,
- j
x
V
The equilibrium equations
(Eulerian description)
(5.57)
Explicitly, the equations in (5.57),
i i j ij
0 b - o j
,
, can be expressed as:

-
o
o o
-
o
o o
-
o
o o
-
o
o o
-
o
o o
-
o
o o
-
o
o o
-
o
o o
-
o
o o
0
0
0
3
3
33
2
32
1
31
2
3
23
2
22
1
21
1
3
13
2
12
1
11
b
b
b
j
j
j
x x x
x x x
x x x
(5.58)
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

299
Then by using both the engineering and Voigt notation, the equilibrium equations can be
expressed as follows:
[ ]

(
(
(

(
(
(

-
(
(
(
(
(
(
(
(

t
t
t
o
o
o
(
(
(
(
(
(
(

o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
0
0
0

0 0 0
0 0 0
0 0 0

3
2
1
b
b
b
j
j
j
xz
yz
xy
z
y
x
T
x y z
z x y
z y x
_
L
[ ] { { { 0 - M L
T

(5.59)
Additionally, the equilibrium equations in the Lagrangian description are given by:
0 b S
0 b P
, ,
, ,
,
,
-
-

0 0
0 0
) ( j
j
F
X
X
V
V
The equilibrium equations
(Lagrangian description)
(5.60)
Problem 5.5: Find the equilibrium equations in engineering notation by means of the
differential volume element equilibrium ( dxdydz ). For this purpose consider that the
Cauchy stress tensor field in the differential volume element varies as indicated in Figure
5.7.





























Figure 5.7: The stress field in the differential volume element.


yz
t

xy
t

y
o

z
o

yz
t

xz
t

xz
t

x
o

xy
t

dz
z
xz
xz
o
t o
- t


dz
z
yz
yz
o
o o
- o


dz
z
z
z
o
o o
- o


dy
y
yz
yz
o
t o
- t


dy
y
xy
xy
o
t o
- t


dy
y
y
y
o
o o
- o

dx
x
xz
xz
o
t o
- t


dx
x
xy
xy
o
t o
- t


dx
x
x
x
o
o o
- o

z
y
x
dz
dx
dy
Rear face

x
b
y
b

z
b
Rear face
Rear face
NOTES ON CONTINUUM MECHANICS

300

Solution:
To obtain the equilibrium equations we apply the force equilibrium condition in the
volume element. First, we evaluate the equilibrium force according to the x -direction:
0
x
F
0 t - |
.
|

\
|
o
t o
- t - t -
|
|
.
|

\
|
o
t o
- t - o - |
.
|

\
|
o
o o
- o -
dxdy dxdy dz
z
dxdz
dxdz dy
y
dydz dydz dx
x
dxdydz
xz
xz
xz xy
xy
xy x
x
x x
b j

Then by simplifying the above equation we obtain:
0
o
t o
-
o
t o
-
o
o o
- dxdydz
z
dxdydz
y
dxdydz
x
dxdydz
xz
xy
x
x
b j
0
o
t o
-
o
t o
-
o
o o
-
z y x
xz
xy
x
x
b j
The equilibrium force according to the y -direction, 0
y
F , can be expressed as follows
0
22
t -
|
|
.
|

\
|
o
t o
- t - t -
|
|
.
|

\
|
o
t o
- t - o -
|
|
.
|

\
|
o
o o
- o -
dydz dydz dx
x
dxdy
dxdy dz
z
dxdz dxdz dy
y
dxdydz
xy
xy
xy yz
yz
yz y
y
y
b j

Then by simplifying the above equation we obtain:
0
o
t o
-
o
o o
-
o
t o
-
z
yz y xy
y
x y x
b j
Finally, the equilibrium according to the z -direction, 0
z
F , is given by:
0 t -
|
|
.
|

\
|
o
t o
- t - t -
|
.
|

\
|
o
t o
- t - o - |
.
|

\
|
o
o o
- o -
dxdz dxdz dy
y
dzdy
dzdy dx
x
dxdy dxdy dz
z
dxdydz
yz
yz
yz xz
xz
xz z
z
z z
b j

Additionally, by simplifying the above equation we obtain:
0
o
o o
-
o
t o
-
o
t o
-
z y x
z
yz
xz
z
b j
Then, the equilibrium equations in engineering notation become:

-
o
o o
-
o
t o
-
o
t o
-
o
t o
-
o
o o
-
o
t o
-
o
t o
-
o
t o
-
o
o o
0
0
0
z
z
yz
xz
y
z
yz y xy
x
xz
xy
x
z y x
x y x
z y x
b
b
b
j
j
j

Problem 5.6: Let be the Cauchy stress tensor field, which is represented by its
components in the Cartesian basis as:
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

301
0 ; 2
; ;
13 31 32 23 2 1 21 12
2
2
2
1 33
2
2 22
2
1 11
o o o o o o
- o o o
x x
x x x x

Considering that the body is in equilibrium, find the body forces acting on the continuum.
Solution: By applying the equilibrium equations, 0 b
, ,
,
- j
x
V , we obtain:

- -
- -

-
o
o o
-
o
o o
-
o
o o
-
o
o o
-
o
o o
-
o
o o
-
o
o o
-
o
o o
-
o
o o
- o
0
0 2 2
0 2 2
0
0
0
3
2 2 2
1 1 1
3
3
33
2
32
1
31
2
3
23
2
22
1
21
1
3
13
2
12
1
11
,
b
b
b
b
b
b
0 b
j
j
j
j
j
j
j x x
x x
x x x
x x x
x x x
i i j ij

Thus, to satisfy the equilibrium equations the following condition must be met:
2 2 2 2
1 1 1 1
4 4
4 4
x x
x x
- -
- -
b b
b b
j j
j j

0
3
b j
)

( 4
2 2 1 1
e e b x x - -
,
j
Problem 5.7: The equations of motion of a body are given, in Lagrangian description, by:

- -
-
-
) (
2 1 3 3
3 2 2
3 1 1
X X t X x
tX X x
tX X x
c
c
c

where c is a constant scalar. Find the mass density in the current configuration ) (j in
terms of the mass density of the reference configuration ) (
0
j , i.e. ) (
0
j j j .
Solution:
We can apply the equation j j J
0
, where J is the Jacobian determinant and is given by:
2
3
3
2
3
1
3
3
2
2
2
1
2
3
1
2
1
1
1
) ( 2 1
1
1 0
0 1
t
t t
t
t
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
X
x
F J
j
i
c
c c
c
c
-
- -

o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o
o

o
o

Thus, we obtain
2
0 0
) ( 2 1 t J c
j j
j
-

5.7.3 The Equations of Motion with Discontinuities
Let us consider again a domain with a singular surface ) (t Z such as that discussed in
subsection 5.5.1, (see Figure 5.4). Then, the principle of conservation of linear momentum
becomes:
} } }
Z - Z - Z -
-
V S V
dV dS dV
Dt
D
b n
,
,

j j v
(5.61)
Then by applying the divergence theorem with discontinuities, (see Eq. (5.17)), we obtain:
( ) [ ] [ ]
} } }
Z Z - Z -
- - dV dS dV
Dt
D
V V
n b

,
,
,
j j
x
v V
(5.62)
NOTES ON CONTINUUM MECHANICS

302
Additionally, by using Reynolds transport theorem, (see equation (5.22)), with v
,
j d , we
obtain:
( ) [ ] [ ]
} } }
Z - -
- - |
.
|

\
|
-
Z Z
dS dV
Dt
D
dV
Dt
D
V V

) (
n c V
, , , , ,
,
,
,
v v v v
v
v
x
j j
j
j (5.63)
Then by combining the above equation with the equation in (5.62) and by considering that
Dt
D
Dt
D
Dt
D ) ( ) ( ) ( v
v
v
,
,
,
j
j j
- we obtain:
( ) [ ] [ ] 0 n b
,
, , ,
,
,
, ,
, ,
- - - |
.
|

\
|
- - - |
.
|

\
|
-
} }
Z -

Z
dS dV
Dt
D
Dt
D
V

) ( ) (
c V V v v
v
v v
x x
j j j j
j

(5.64)
Bearing in mind the mass continuity equation, 0
) (
- v
x
,
,
V j
j
Dt
D
, the equation in (5.64)
becomes:
( ) [ ] [ ] 0 n b
,
, , ,
,
,
,
- - - |
.
|

\
|
- -
} }
Z -

Z
dS dV
Dt
D
V

) (
c V v v
v
x
j j j (5.65)
Then, the local form can be expressed as:
( ) [ ] [ ] Z - -
Z - -

on
V in
0 n
b
,
, , ,
,
,
,

c
V
v v
a
x
j
j j
The equations of motion with
discontinuities
(Eulerian description)
(5.66)
For a static or quasi-static problem the equations in (5.66) become:
[ ] [ ]
Z



Z - -

- -
on
V in
n n
0 n
0 b

,
, ,
,
j
x
V
The equations of motion with
discontinuities (static problem)
(Eulerian description)
(5.67)
5.8 The Principle of Conservation of Angular
Momentum. Symmetry of the Cauchy Stress
Tensor
5.8.1 Angular Momentum
Once again let us consider Figure 5.6, and we can define the angular momentum of a mass
system with respect to the origin by:
dV v x t H
dV
V
k j ijk i O
V
O
}
}

r
) ( ) (
) (
j
j

v x H
, ,
,
Angular momentum (5.68)
The SI unit of
O
H
,
is [ ]
s
m kg
O
2

H
,
, and J Nm
s
m kg
O

(

2
2

H
`
,
.
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

303
5.8.2 The Principle of Conservation of Angular Momentum
The principle of conservation of angular momentum states that the rate of change of
angular momentum with respect to a point is equal to the resultant moment (with respect
to this point) produced by all forces acting on the body under consideration.
Then by obtaining the resultant momentum with respect to the origin, (see Figure 5.6), and
by applying the principle of angular momentum, we obtain:
dV
Dt
D
dV dS
V V S
} } }
r r - r ) ( ) ( ) (
*
v x x x
, ,
,
,
,
,
j jb t


[ ] Nm (5.69)
NOTE: The equation in (5.69) is valid for those continuous media in which the forces
between particles are equal, opposite and collinear, and without any distributed moments.
The equation in (5.69) can be rewritten in indicial notation as:

dV v x v x
dV v x
Dt
D
dV v x
Dt
D
dV x dS x
V
k j ijk
i
k
j
j ijk
V
k j ijk
V
k j ijk
V
k j ijk
S
k j ijk
v
}
} } } }
-
-

) (
) ( ) ( ) ( ) (
0
*
`
_
`

j
j j jb t


(5.70)
Then by substituting
l kl k
n t

*
o into the first integral of (5.70), and by applying the Gauss
theorem, we obtain:

dV a x dV x dV x
V
k j ijk
V
k j ijk
V
l kl j ijk
} } }
- o ) ( ) ( ) (
,
j j b
(5.71)

dV a x dV x x x
V
k j ijk
V
k j ijk l kl j ijk kl
jl
l j ijk
} }
- o - o ) ( ) (
, ,
j j
c
b
(5.72)


(

(
- - o - o

i
V
motion of Equations
k
k k l kl j ijk kj ijk
dV a b x 0 ) (
0
,
_
j j
i kj ijk
V
dV 0 o
}

(5.73)

i kj ijk
0 o


kj jk
o o

(5.74)
Thus obtaining Cauchys second law of motion, also known as the Boltzmann postulate, the
symmetry of the Cauchy stress tensor is:
T
Cauchys second law of motion
(5.75)
Then bearing in mind the relationship
T
J F
-
P
1
, the Boltzmann postulate in the
reference configuration becomes:

|
.
|

\
|


T
T T
T
J J
F F P P

1 1
T T
P P F F

(5.76)
and considering that S P F , where S is the second Piola-Kirchhoff stress tensor, we
obtain:
NOTES ON CONTINUUM MECHANICS

304
T T T T T
F F F F F F F F S S S S ; ) ( (5.77)
Thus
T
S S

(5.78)
Problem 5.8: Find the linear and angular momentum for a solid subjected to rigid body
motion.
















Solution: According to Problem 2.16 in Chapter 2, we obtained the velocity for rigid body
motion as:
) ( c c
,
, ,
`
,
,
- r - x v
where
,
is the axial vector (angular velocity) associated with the antisymmetric tensor W
(the spin tensor).
Linear momentum:
( )
} } }
} } } } }
r - r -
r - r - - r -
V V V
V V V V V
dV dV dV
dV dV dV dV dV

) (
j j j
j j j j j
c c
c c c c
,
, , ,
`
,
,
, , ,
`
, ,
, ,
`
,
,
,


x
x x v L

By definition x x
,
,
m dV
V

}
j is the first moment of inertia, where m is the total mass, and
k
x
,
is the vector position of the center of mass G . The first moment of inertia is equal to
zero if the Cartesian system originates at the center of mass, so, 0
, ,
,

}
x x m dV
V
j .

[ ]
v
x L
,
,
,
,
`
,
,

) (
m
m

- r - c c
(Linear momentum for rigid body motion)
where ) ( c c
,
,
,
`
,
,
- r - x v is the velocity of the center of mass.
Angular momentum:
( ) [ ] dV dV
V V
O
) ( ) (
} }
- r - r r c c
,
, ,
`
,
, , ,
,
x x v x H j j
Thus

t
B

1
x

2
x

3
x
x
,

v
,

G
O

1
x

3
x

2
x
Rigid body
G - mass center

) 1 (
F
,


) 2 (
F
,

) (n
F
,

5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

305

) ( ) (
) ( ) (
c c
c c
,
, , , , ,
`
,
,
,
, , , , ,
`
,
,
,
r r
(
(

- r r - r
(
(

r r - r r - r
} } }
} } }


dV dV dV
dV dV dV
V V V
V V V
O
x x x x
x x x x H
j j j
j j j

(5.79)
Next, we discuss the second integral of the previous equation.
It was proven in Chapter 1 that given three vectors a
,
, b
,
, c
,
, the relationship
c b a b c a c b a
,
,
,
,
, , ,
,
,
) ( ) ( ) ( - r r holds, thus when c a
, ,
it holds that
a b a b a a a b a
,
,
,
,
, , ,
,
,
) ( ) ( ) ( - r r , so, [ ]dV dV
V V
) ( ) ( ) (
} }
- r r x x x x x x
, , , , , , , , ,
j j , with
which we obtain:
[ ] [ ] [ ]
[ ]
p ip O p
V
i p pi k k
p
V
i p pi k k
V
i p p pi p k k
V
i p p i k k
dV x x x x
dV x x x x dV x x x x dV x x x x
c c -
c - c - c c - c
}
} } }
I

c j
c j c j j

or in tensorial notation:
[ ]
, , , , , , , , ,

(
(

- r r
} }
O
V V
dV dV I 1 ) ( ) ( ) ( x x x x x x j j
where [ ]dV
V
O
) ( ) (
}
- x x x x
, , , ,
1 I j is the inertia tensor with respect to the origin O . As
we can observe,
O
I is a second-order pseudo-tensor, since it depends on the reference
system, and the components [ ] dV x x x x
V
j i ij k k ij O

}
- c j I can be expressed explicitly as:

[ ] [ ]
[ ] [ ] dV x x dV x x
dV x x dV x x x x x x x x
V
O
V
O
V V
O
;
) (
2
2
2
1 33
2
3
2
1 22
2
3
2
2 1 1 11 3 3 2 2 1 1 11
} }
} }
- -
- - - -
j j
j c j
I I
I

[ ] [ ]
12 2 1 2 1 12 3 3 2 2 1 1 12
) (
O
V V
O
dV x x dV x x x x x x x x I - - - - -
} }
j c j I
[ ] ;
13 3 1 13 O
V
O
dV x x I - -
}
j I [ ]
23 3 2 23

O
V
O
dV x x I - -
}
j I
where
11 O
I ,
22 O
I ,
33 O
I , are moments of inertia of the body relative to the reference point O ,
and
12 O
I ,
13 O
I ,
23 O
I , are the products of inertia of the body relative to the reference point
O .
Returning to the equation in (5.79) we can state that:
[ ]

,
,
,
`
,
,
,
,
,
,
`
,
,
,
, , , , ,
`
,
,
,
- r - r r r - - r
r r
(
(

- r r - r
(
(

} } }
O O
V V V
O
m m m
dV dV dV
I c c c I c
c c
) ( ) (
) ( ) (
x x x
x x x x H j j j

Then by adding and subtracting the term x x
,
,
,
r r m in the above equation we obtain:
( ) [ ]
[ ] [ ] {
G
O O
O O O
m
m
m m m m
m m m
H v x
v x
x x x x v x x x x x v x
x x x x x H
, , ,
,
, ,
,
, , , , , ,
, ,
, , , , , ,
,
,
,
,
,
,
,
`
,
,
,
,
,
`
,
, ,
- r
- r
- - - r - - - r
- r r - - r - r - r - r





) ( ) ( ) ( ) (
) ( ) (



I
I 1 I 1
I c c I c c

NOTES ON CONTINUUM MECHANICS

306
where [ ] 1 I I ) ( ) ( x x x x
, , , ,
- - m
O
is the inertia pseudo-tensor, which is related to the
reference system at the center of mass. By means of this equation we can calculate the
inertia tensor in any reference system if we know the inertia tensor at the center of mass:
[ ]
ij j i ij ij O
x x x x x m c ) (
2
3
2
2
2
1
- - - - I I . Explicitly, these components can be expressed as:
) ( ; ) (
) ( ; ) (
) ( ; ) (
3 1 13 13
2
2
2
1 33 33
3 2 23 23
2
3
2
1 22 22
2 1 12 12
2
3
2
2 11 11
x x m x x m
x x m x x m
x x m x x m
O O
O O
O O
- - -
- - -
- - -
I I I I
I I I I
I I I I

Note that, the above equations represent the parallel axis theorem (Steiners theorem) from
Classical Mechanics.
Problem 5.9: Obtain the principle of conservation of linear momentum and angular
momentum for a solid subjected to rigid body motion.
Solution: We can start from the definition of the principle of conservation of linear
momentum which states that:
L v
`
,
,
,

}

V
dV
Dt
D
j F
Then we use the equation of linear momentum obtained in Problem 5.8, v L
, ,
m , to
obtain:
a v L v
,
`
,
`
,
,
,
m m dV
Dt
D
V

}

j F
Then we have:
a
, ,
m

F

Now let us consider the principle of conservation of angular momentum which states:
O O
V
O
Dt
D
dV
Dt
D
H H v x M
`
, ,
, ,
,
= r
}

) ( j
By which we obtain:
O O
H M
`
, ,

or
G G
H M
`
, ,


where the equation of angular momentum
O
H
,
was obtained in Problem 5.8. The set of
equations a
, ,
m

F and
G G
H M
`
, ,

inform us that the following systems are


equivalent:













=
G - center of mass
G

) 1 (
F
,


) 2 (
F
,

) (n
F
,

G
a
,
m

G
H
`
,

5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

307
5.9 The Principle of Conservation of Energy. The
Energy Equation
The principle of conservation of energy states:
The rate of change of the kinetic energy plus the rate of change of the internal
energy is equal to the sum of the rate of change of the work done by the system plus
the rate of change of any other energy supplied to, or removed from, the system.
(5.80)
The energy supplied to, or removed from, the system per unit time can be any of three
kinds: thermal; chemical; or electromagnetic energy. In this publication we only consider
thermal energy as the energy added to the system. In such circumstances, the principle of
conservation of energy is known as the first law of thermodynamics. Mathematically, the
principle of conservation of energy, for continuum thermodynamics, is given by:
Dt
Q D
Dt
D
Dt
DU
Dt
D
- -
W K
(

W
s
J
(5.81)
where K is the kinetic energy, U is the internal energy, W is the work done by the
system, and Q is the energy added to the system.
Next, we will introduce the types of energy involved in the energy equation.
5.9.1 Kinetic Energy
The kinetic energy of the system represented in Figure 5.6 is given by:
dV v v dV t
V
i i
V
} }
) (
2
1
) (
2
1
) ( j j v v
, ,
K
Kinetic energy [ ] J (5.82)
The SI unit of the energy is the joule: [ ] J Nm dV
m
Nm
dV
s
m
s
m
m
kg
V V

} } 3 3
K .
Then, the rate of change of the kinetic energy becomes:
dV v v v v dV v v
Dt
D
dV v v
Dt
D
t
Dt
D
V
i i i i
V
i i
V
i i
} } }
-
(
(

= ) (
2
1
) (
2
1
) (
2
1
) ( ` `
`
j j j K K
(5.83)
Thus
dV v v t
Dt
D
V
i i
}
= ) ( `
`
j K K
(5.84)
5.9.2 External and Internal Mechanical Power
Let us consider the equations of motion
i i j ij
v` j j - o b
,
, and if we substituting those into
the rate of kinetic energy given in (5.84) we obtain:
dV v
V
i j ij i
}
- o ) (
,
b j K
`

(5.85)
Then the term
j ij i
v
,
o can be substituted by:
NOTES ON CONTINUUM MECHANICS

308

ij j i j ij i j ij i j ij i ij j i j ij i
ij
v v v v v v o - o o o - o o
l
, , , , , ,
) ( ) (
(5.86)
where l = v
x
,
,
V is the spatial velocity gradient, which can be broken down into a
symmetric and an antisymmetric part, i.e. W D- l , (see Chapter 2), where D is the rate-
of-deformation tensor and W is the spin tensor. The components of these tensors can be
expressed in terms of Eulerian velocity as:
_ _
ij
i
j
j
i
ij
i
j
j
i
j i ij
x
v
x
v
x
v
x
v
v
W D
|
|
.
|

\
|
o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o

2
1
2
1
,
l

(

s m
m

(5.87)
Returning to the equation in (5.85), and considering the relationships in (5.86) and (5.87),
the rate of change of the kinetic energy becomes:
( ) [ ]
[ ]
( ) dV dV v dV v
dV v v dV
v v v
ij
V V
ij
V
j
ij i i i
V
i i ij ij ij j ij i
V
i i ij j i
j
ij i
D b
b W D
b
} } }
} }
o - o -
- - o - o
- o - o

,
,
,
,
) ( ) (
j
j j K
`

(5.88)
where we have taken into account that the double scalar product of a symmetric and
antisymmetric tensor is equal to zero, i.e. 0 o
ij ij
W or 0 W : . Then by applying the
divergence theorem to the second integral of the right side of the equation in (5.88), we
find that:
( ) dS v dS v dV v
i
S
i j ij
S
i
V
j
ij i
*
,

t n
} } }
o o


(5.89)
By combining the above relationship with the equation in (5.88), we can still express the
rate of change of the kinetic energy as:
_
_
`
Power Mechanical Internal
t
ij
V
ij
Power Mechanical External
t
S
i i
V
i i
int
ext
dV dS v dV v

) (

) (
*

P
P
D t b
} } }
o - -

j K
) ( ) ( t t
Dt
D
int ext
P P - K (5.90)
or
) ( ) ( t t
Dt
D
ext int
P P - K (5.91)
where we have introduced the external mechanical power ) (t
ext
P , which is the rate of change
of the work done by the external forces
Dt
DW
, as:
dV v dS v t
dV dS t
i
V
i i
S
i ext
V S
ext
} }
} }
-
-
b t ) (
) (
*
*
j
j

b t
P
P v v
,
,
,
,

The external mechanical power
(

W
s
J
(5.92)
and the internal mechanical power, also known as the stress power, which is the rate of change of
the work done by the internal forces:
( ) ( ) dV dV dV dV
V V
T
V
ij
V
ij int
} } } }
o D D D Tr Tr D : P

The stress power (5.93)
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

309
NOTE: The SI unit of power is the watt, s J W / = , i.e. one joule ( J ) per second ( s ),
which is equal to [ ] W
s
J
dV
s m
J
dV
s m
m
m
N
dV
s m
m
Pa
V V V
int

} } }

3 2
P .
We can also define the stress power per unit volume, denoted by ) (t
int
w , as:
) ( ) ( D D Tr w : t
int
Stress power per unit volume (5.94)
Then by starting from the stress power in the current configuration we can also express the
stress power as a function of the other stress tensors, i.e.:

} } }

0 0
0 0

V V V
dV dV J dV D D D : : :


(5.95)
Bearing in mind that
jk ik ij
F P t (Kirchhoff stress tensor components),
T T
P P F F
(Cauchys second law of motion in the reference configuration), and
1 1 - -

lj pl pi ij
F E F
`
D , we
obtain:
} } } }
} } }
} } }
} }
} } }



-

- -
- -
- -
o
V V V V
V V
ik ik
V
pk pk
V
jk ij ik
V
pk ik pi
V
pk pi ik
V
ij ij jk ik
V
lk
jk lj pl pi ik
V
ij jk ik
V
lj pl pi jk ik
V
dV dV
J
dV dV
dV dV F dV E
dV F dV E F dV E F
dV F dV F F E F
dV F dV F E F F dV

1

2
1





0
0 0
0 0 0
0 0
1
0
1
0 0
1 1
0 0
1 1
0
0 0
0 0 0
0 0 0
0 0
0 0 0
F F C E
F
` ` ` `
` ` `
` `
_
`
`
: : : :
:
:
P P S S
P
D
j
j
P S
P P P
) W ( P P
D P P
l
l
(5.96)
which proves that the rate of change of the deformation gradient and the first Piola-
Kirchhoff stress tensor are conjugate quantities ( F
`
: P ). Other conjugate quantities are: the
second Piola-Kirchhoff stress tensor and the rate of change of the Green-Lagrange strain
tensor ( E
`
: S ); the Kirchhoff stress tensor and the rate-of-deformation tensor ( D : ).
Furthermore, we can show that U T
`
: is already a conjugate pair. To prove this, let us
consider the relationship T R P , where S U T is the Biot stress tensor, and R is the
orthogonal tensor from the polar decomposition, and
T
U U is the right stretch tensor.
Then if we refer to the right polar decomposition, i.e. U R U R U R
` ` `
- F F , we
obtain:
( ) ( )
( ) ( ) ( ) ( )
( ) ( )
( ) ( )
U T
U T R R U S U
U T R R U T
U R T R U R T R
U R U R T R P
`
` `
` `
` `
` ` `
:
: :
: :
: :
: :

-
-
-
-




T T
T T
F ( )( )
kj kj
kj kj ik ip kj qj pq
kj kj ik ip kj pj
kj ik pj ip kj ik pj ip
kj ik kj ik pj ip ij ij
F
U T
U T R R U S U
U T R R U T
U R T R U R T R
U R U R T R P
`
` `
` `
` ` `
` ` `

-
-
-
-
) )( (
) )( (

(5.97)
NOTES ON CONTINUUM MECHANICS

310
Note that ( ) ( ) 0 R R U S U
`
T
: , since the tensor U S U U S U U S U
T T T
) ( is
symmetrical and ( ) R R
`

T
is an antisymmetric tensor. Thus, the equation in (5.97)
becomes:
H T U T P
` ` `
: : : F
(5.98)
where 1 U H - is the Biot strain tensor, (see Chapter 2) and if we know that U
`
is
symmetrical, it is also possible to express the above relationship as:
U T U T T P
` ` `
: : :
sym skew sym
- ) ( F (5.99)
Then, if we take into account all the equations obtained before, we can summarize the
stress power per unit volume by:
F F
` `
: : : P P D
J
int
1
0

j
j
w
The stress power per unit current
volume
(5.100)
H T U T S P S D
` ` ` ` `
: : : : : : C F E
2
1
int
w
The stress power per unit reference
volume
(5.101)
5.9.3 The Balance of Mechanical Energy
If we compare the equation given in (5.91) with the energy equation (5.81), i.e.:

0
) ( ) (

- - = -
Dt
Q D
Dt
D
Dt
DU
Dt
D
t t
Dt
D
ext int
W K
K P P
(5.102)
we can observe that the equation in (5.91) is an exceptional case of the energy equation
where only mechanical energy is considered. In this case the principle of conservation of
energy is known as the balance of mechanical energy which is otherwise known as the theorem of
power extended:
_
,
,
,
,
_ _
power mechanical External
power Stress
energy Kinetic the of
change of Rate
} } } }
} } } }
- -
- o -

t b D
S V V V
S
i i
V
i i
V
ij ij
V
dS dV dV dV v
Dt
D
dS v dV v dV dV v
Dt
D
v v
* 2
* 2
2
1
2
1
j j
j j
:
t b D
Balance of
mechanical energy
(5.103)




OBS.: In rigid body motion 0 D is satisfied, so, the stress power (internal
mechanical power) is zero 0 ) ( t
int
P , then it holds that ) (t
ext
P K
`
.
If K
`
is discarded, which characterizes a static or quasi-static regime, it holds that
) ( ) ( t t
ext int
P P .
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

311

Problem 5.10: Find the kinetic energy related to rigid body motion in terms of the inertia
tensor, (see Problem 5.8 and Problem 5.9).
Solution: The rigid body motion velocity can be expressed as ) ( c c
,
, ,
`
,
,
- r - x v . Then, the
kinetic energy becomes:
[ ] [ ]
dV dV t
V V

) ( ) (
2
1
) (
2
1
) (
} }
)
`

- r - - r -

c c c c
,
, ,
`
, ,
, ,
`
,
, ,
x x
v v

j j K
Using the following vector sum x x x -
,
,
,
, where x
,
is the mass center vector position,
and x
,
is the particle vector position with respect to the system that has its origin in the
center of mass, the energy equation becomes:
[ ] [ ] {
( ) [ ] ( ) [ ] { dV
dV t
V
V
) ( ) ( ) ( ) (
2
1
) ) (( ) ) ((
2
1
) (
}
}
r - - r - r - - r -
- - r - - - r -

x x x x
x x x x
, ,
,
,
,
`
,
, ,
,
,
,
`
,
,
,
,
,
`
, ,
,
,
,
`
,


c c c c
c c c c
j
j K

Note that ) ( c c
,
,
,
`
,
,
- r - x v is the center of mass velocity, thus:
[ ] [ ] { dV t
V
) ( ) (
2
1
) (
}
r - r - x v x v
, ,
,
, ,
,
j K
or:
) ( ) (
2
1
) (
2
1
) (
2
1

2
1
) (
} } } }
r r - - r - r -
V V V V
dV dV dV dV t x x v x x v v v
, , , ,
,
, , , ,
, , ,
j j j j K

Then by simplifying the above equation we obtain:
) ( ) (
2
1
) (
2
1
) (
} } }
r r - r -
V V V
dV dV dV t x x x v v v
, , , , , ,
, , ,
j j j K
Next, we discuss separately the terms of the previous equation:
1)
2
2
2
1

2
1

2
1
v m dV dV
V V

} }
j j v v v
, , ,

2)

0 ) ( ) ( r
(
(

r r


} }
0
,
,
,
,
, ,
,
, ,
,
x v x v x v m dV dV
V V
j j
Note that, the system x
,
is located at the center of mass ( G ), hence the center of mass
vector position related to the system x
,
is zero.
3) ) ( ) (
}
r r
V
dV x x
, , , ,
j
[ ]
p jp j
p
V
p j k k jp j
V
p j p k k jp j
V
p q k kp jq q k kq jp j
V
q p k j kp jq kq jp
V
q p ipq k j ijk
V
dV x x x x
dV x x x x
dV x x x x
dV x x dV x x dV
c c
c
|
|
.
|

\
|
- c
c - c
c - c
c c - c c r r
}
}
}
} } }

I
) (
) (
) (
) ( ) ( ) (
c j
c j
c c c c j
c c c c j j j x x
, , , ,


or in tensorial notation as:
NOTES ON CONTINUUM MECHANICS

312
[ ] [ ]


, ,
, , , , , , , , , ,

(
(

- r r
} }
I
1 dV dV
V V
) ( ) ( ) ( ) ( x x x x x x j j

where I is the inertia pseudo-tensor related to the system located at the center of mass,
(see Problem 5.8).
Then if we bear in mind all the above considerations, the kinetic energy equation for rigid
body motion becomes:
) ( ) (
2
1
) ( 2
2
1

2
1
) (
0
} } }
r r - r -

V V V
dV dV dV t x x x v v v
, , , ,
_
, ,
, , ,
j j j K

, ,
- I
2
1

2
1
) (
2
v m t K

Additionally, if we take into account that:
[ ] [ ] [ ]
[ ] [ ] [ ]
[ ] [ ] [ ]
(
(
(

- -
- -
- -

(
(
(
(
(
(
(

- - -
- - -
- - -

} } }
} } }
} } }
33 23 13
23 22 12
13 12 11
2
2
2
1 3 2 3 1
3 2
2
3
2
1 2 1
3 1 2 1
2
3
2
2



I I I
I I I
I I I
dV x x dV x x dV x x
dV x x dV x x dV x x
dV x x dV x x dV x x
V V V
V V V
V V V
ij
j j j
j j j
j j j
I
we obtain an explicit equation for the kinetic energy as:
[ ]
[ ]
3 2 23 3 1 13 2 1 12
2
3 33
2
2 22
2
1 11
2
3
2
1
33 23 13
23 22 12
13 12 11
3 2 1
2
2
2 2 2
2
1

2
1

2
1

2
1
2
1

2
1
) (
c c - c c - c c - c - c - c -
(
(
(

c
c
c
(
(
(

- -
- -
- -
c c c -
c c -
I I I I I I
I I I
I I I
I I I
v m
v m
v m t
j kj k
I K


[ ]
3 2 23 3 1 13 2 1 12
2
3 33
2
2 22
2
1 11
2
2 2 2
2
1

2
1
) ( c c - c c - c c - c - c - c - I I I I I I v m t K


5.9.4 The Internal Energy
If we take a handful of atoms (the material point) and we evaluate the average of all forms
of energy present in it we obtain what is known as the internal energy. Continuum
thermodynamics usually presents the rate of change of the internal energy as:
} }

V V
dV u dV u
Dt
D
Dt
DU
` j j

(

s
J
(5.104)
where u is the specific internal energy, i.e. energy per unit mass, [ ]
kg
J
u . For example, for an
ideal gas the specific internal energy is given by

j
j
p
T c T c u
p v
- , where T is the
temperature,
v
c is the specific heat capacity at a constant volume, j is the specific entropy,
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

313
p
c is the specific heat capacity at a constant pressure, p is the thermodynamic pressure,
and j is the mass density. We can give another example with the mechanical problem,
which was discussed in the previous subsection, where the rate of change of the internal
energy is given by dV
Dt
DU
V
}
D : .
5.9.5 Thermal Power
We define thermal power as the rate of increase of total heat in the continuum, which is
denoted by
Dt
Q D
. The contribution of thermal power considered here is caused by:
The Cauchy heat flux (non convective, i.e. without mass transport);
The heat sources.
1) The Cauchy heat flux
Let us assume that there is a temperature gradient in the continuum, so there is scientific
evidence of energy transfer (heat) from the hottest to the colder region. Then, we can
represent this transferred energy per unit area per unit time by the thermal flux vector
) , ( t x
,
,
q , which is also known as the Cauchy heat flux or true heat flux. Now, let us consider the
domain B bounded by the surface S , (see Figure 5.8). The amount of energy which is
transferred through the surface dS per unit time, (see Figure 5.8), is represented by
dS t n q

) , ( x
,
,
, where n

is the outward unit normal to the area element dS . Meanwhile, the


tangential component remains on the surface. Thus, the rate of increase of total heat, due
to thermal flux, in the continuum is given by:
}
-
S
dS t

) , ( n q x
,
,

(

W
s
J
(5.105)
2) The heat sources
If in a continuum there is a nuclear or chemical reaction which results in the release of
heat, we can represent this by means of the heat sources, (see Figure 5.8).
We represent the rate of increase of total heat in the continuum cause by the heat source
as:
}
V
dV r j

(

W
s
J
(5.106)
where ) , ( t r x
,
is the radiant heat constant (also called the heat source) per unit mass per unit
time, a scalar function, and the SI unit is [ ]
kg s
J
t r

) , ( x
,
, and ) , ( t x
,
j is the mass density.
Then by considering the heat flux (incoming) and the heat source, we can define the thermal
power (the rate of thermal work) as:
} }
-
S V
dS dV r
Dt
Q D

n q
,
j
The thermal power
(

W
s
J
(5.107)


NOTES ON CONTINUUM MECHANICS

314









Figure 5.8: Heat flux and heat source.
5.9.6 The First Law of Thermodynamics. The Energy
Equation
Once we know what forms of energy are involved in a system we can provide the energy
equation by starting from that in (5.81):
Dt
Q D
Dt
D
Dt
DU
Dt
D
- -
W K
(5.108)
The mechanical power and the thermal power are not exact differentials (
Dt
D-
), but there is
experimental evidence showing that the sum of mechanical and thermal power is already an
exact differential, (Mase(1977)).
Considering only the mechanical and thermal energy, the principle of conservation energy
becomes what is known as the first law of thermodynamics, which postulates the
interchangeability of mechanical and thermal energy. Then, the equation in (5.108)
becomes:

} } } } } }
- - - -
o
q
S
i i
V V
i i
S
i
j ij
i
V V
i i
dS rdV dV v dS v dV u dV
v v
Dt
D
n q b t
n

*
j j j j `
(5.109)
Then by using divergence theorem to transform the surface integral into the volume
integral we obtain:
( )
( ) [ ]
( )
} }
} } }
} } } } } }
- - - - o - o
- - - o -
- - - o -
V
i i i i i i j i ij i j ij
V
V
i i i i
j
i ij
V V
i i
V
i i
V V
i i
V
j
i ij
V V
i i
dV v v r v v v dV u
dV r v v dV u dV v v
dV rdV dV v dV v dV u dV
v v
Dt
D



2
, , ,
,
,
,
,
` `
` `
`
j j j j
j j j j
j j j j
q b
q b
q b

(5.110)
Additionally, by rearranging the above equation we obtain:
( )
} }
(

(
- - o - - - o

V
i i j i ij
motion of quations the
i
i i j ij i
V
dV r v v v dV u
, ,
e
,
q b
0
j j j j
_
` `
(5.111)

t
B
dS
) , ( t x
,
,
q
n



1
x

2
x

3
x
n n n q q

)

( ) , (
n n
t q
,
,
,
x

[ ]
[ ]
s kg
J
r
s m
J

2

q
,

Current configuration
dV
) , ( t r x
,
j
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

315
Then if we bear in mind that
ij ij j i
v W D -
,
, the above equation becomes:
( ) [ ] ( )
} } } }
- - o - - - o
V
i i ij ij
V V
i i ij ij ij
V
dV r dV u dV r dV u
, ,
q D q W D j j j j ` ` (5.112)
The local form of the above equation is known as the energy equation:
i i ij ij
r u
,
q D - - o j j`
(


3 3
m
W
s m
J
(5.113)
which is expressed in tensorial notation as:
r u j j - - q D
,
`
,
x
V : The energy equation (current configuration) (5.114)
NOTE: For a purely mechanical problem in which there is no internal heat production
( 0 r ) nor heat flux 0 q
,
,
, the energy equation becomes:
D :
j
1
u`
(

kg s
J


(5.115)
where the SI unit can easily be verified [ ]
kg s
J
kg s
m N
s m
m
m
N
kg
m
u



1
2
3

(

D :
j
`
5.9.6.1 The Energy Equation in Lagrangian Description
The energy equation (5.114) can also be established in Lagrangian description (material
description). From the equation in (5.112), the integral related to the integral energy can be
written in the reference configuration as:
} } }

0 0
0 0 0
) , ( ) ( ) , ( ) , (
V V V
dV t u dV u J dV t u t X X x x
,
`
,
`
,
`
,
j j j
(5.116)
The integral associated with stress power can be established in the reference and current
configuration, (see equations (5.100) and (5.101)), as shown bellow:

} } } } } } }

V V V V V V V
dV dV dV dV dV dV J dV
2
1

0
0 0 0 0 0
0 0 0 0 0
F F C E
` ` ` `
: : : : : : :

P P S S D D D
j
j
(5.117)
Similarly for the integral related with the heat source, i.e.:
_
, ,
_
, ,
ion configurat reference
V V
ion configurat current
V
dV t r dV r J dV t r t

0 0 0

0 0
) , ( ) ( ) , ( ) , (
} } }
X X x x j j j
(5.118)
Finally, we can address the integral related to the heat flux. The amount of heat that passes
through the area element da in the current configuration must in theory be the same as,
that which passes through the area element dA in the reference configuration, (see Figure
5.9). Then the following relationship must be met:
a A
,
,
,
,
d d q q
0
(5.119)
where
0
q
,
is the heat flux in the reference configuration. Then if we use Nansons formula
A F a
,
,
d J d
T

-
, obtained in Chapter 2, the equation in (5.119) becomes:
NOTES ON CONTINUUM MECHANICS

316
T T T
J J d J d F F A F A
- - -

0
1
0 0
q q q q q q
, , , ,
,
,
,
,
(5.120)








Figure 5.9: Heat flux.
Thus, the integral
}

V
dV q
,
,
x
V can be written in the reference configuration as:
}
} } } }
|
.
|

\
|
o
o
- |
.
|

\
|
o
o

|
.
|

\
|
o
o

o
o

0
0 0
0 0
0
0 0 0 ,

1

1


1

V
ik
i
k ik
i
k
V
ik k
i
V
i
i
V
i i
V
dV F
J x
J F
J x
J
dV F
J x
J dV
x
J dV dV
q
q
q
q
q q
,
,
x
V
(5.121)
It was proven in Chapter 2 that ( ) 0
,
,

-
F
x
1
J V , thus, the above equation becomes:
}
} } } }

o
o

|
|
.
|

\
|
o
o
o
o
|
.
|

\
|
o
o

0
0 0 0
0 0
0
0
0
0 0


1

V
V
k
k
V
k
i
i
k
V
ik
i
k
V
dV
dV
X
dV
X
x
x
dV F
J x
J dV
q
q
,
,
,
,
X
x
V
V
q q q
(5.122)
Bearing in mind the equations in (5.116), (5.117), (5.118) and (5.122), the energy equation
in the reference configuration can be established as:
( )
} }
- -
0 0
0 0 0 0 0

V V
dV r dV u j j q S
,
`
`
,
X
E V :
(5.123)
Additionally, the local form of the above equation is:
) , ( ) , (
0 0 0
t r t u X E X
X
,
,
`
,
`
,
j j - - q S V :
The energy equation
(reference configuration)
(5.124)
5.9.7 The Energy Equation with Discontinuity
In this subsection we obtain the energy equation for a domain with a singular surface ) (t Z
as discussed in subsection 5.5.1, (see Figure 5.4). In this case the energy equation becomes:
Dt
Q D
Dt
D
Dt
DU
Dt
D
- -
W K

} } } } } }
- - - -
- Z - - Z - Z - Z -
- - -
(
(

-
S S V S S V V V
dS rdV dS dV udV dV
Dt
D

) (
2
1
n q n b
,
, ,
,
, ,
j j j j v v v v
(5.125)
Reference configuration
Current configuration
A
,
d
a
,
d

0
q
,

q
,


T
T
J
d J d
F
A F a

-
-

0
1
q q
, ,
,
,

F
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

317
For the terms on the left of the equation in (5.125) we can apply Reynolds transport
theorem, (see the equation in (5.22)), to u j j - d ) ( v v
, ,
, thus:
( )
( ) ( ) ( ) [ ] [ ]
} }
}
Z Z -
Z -

- - -
|
|
.
|

\
|
- -
-

(
(

-
dS u dV u
Dt
u D
dV u
Dt
D
V
V

) ( ) (
) (
) (
2
1
2
1
2
1 2
1
n c V
, , , , , , ,
, ,
, ,
,
v v v v v v
v v
v v
x
j j j j
j j
j j

(5.126)
Then by mathematically manipulating the terms of the volume integral we can see that:
( )
( )
( )
( ) v v v
v v
v v v v v
v v
x
x
, , ,
, ,
, , , , ,
, ,
,
,

- - - -
- - - -
-
V
V
u
Dt
Du
Dt
D
u
Dt
D
Dt
D
u
Dt
u D
j j j
j
j
j
j j
j j
) (
) (
) ( ) (
) (
2
1
2
1
2
1
2
1 2
1

(5.127)
Moreover, by reorganizing the above equation, we find that:
( )
( )
( )
( ) ( )
( )
( )
Dt
D
Dt
Du
Dt
D
u
u
Dt
Du
u
Dt
D
Dt
D
u
Dt
u D
) (
) (
) ( ) (
) (
) (
) (
2
1
0
2
1
2
1
2
1
2
1
2
1 2
1
v v
v v v
v v v v v
v v
v v v
v v
x
x
x
, ,
_
, , ,
, , , , ,
, ,
,
`
, ,
, ,
,
,
,

- - |
.
|

\
|
- -
- - - - -
- - -
-

j j j
j
j j
j
j j j
j j
V
V
V

(5.128)
Thus,
( )
( ) v v v v v
v v
x
`
, , ,
`
, ,
, ,
,

- - -
-
j j j j
j j
Dt
Du
u
Dt
u D
V ) (
) (
2
1 2
1

(5.129)
Then if we return to the equation in (5.126), and if we refer to (5.129) we can conclude
that:
( ) ( ) [ ] [ ]
} }
}
Z Z -
Z -

- - - |
.
|

\
|
-

(
(

-
dS u dV
Dt
Du
dV u
Dt
D
V
V

) (
) (
2
1
2
1
n c
, , , ,
`
, ,
, ,
v v v v v
v v
j j j
j j

(5.130)
For the surface integrals on the right side of the equation in (5.125) we can apply Gauss
theorem to a volume with discontinuity, (see equation (5.17)):
( ) ( ) [ ] [ ]
} } }
Z - -
- - - -
- - - -
dS dV dS
V V S S


n q q n q
,
,
,
,
,
,
,
v v v
x
V
(5.131)
Then if we know that ( ) ( ) v v v
x x x
, , ,
, , ,
V V V : - o - o o
j i ij i j ij
j
ij i
v v v
, ,
,
, and if
we have observed the spatial velocity gradient has been broken down into a symmetric and
an antisymmetric part, we obtain D W D : : : : - v
x
,
,
V , so, we can conclude that:
( ) ( ) [ ] [ ]
( ) [ ] [ ]
} }
} } }
Z -
Z - -


- - - -
- - - - -
- -
- - - -
dS dV
dS dV dS
V V
V V S S

n q q D
n q q q
,
,
,
,
,
,
,
, ,
,
,
, ,
, , ,
v v
v v v n v
x x
x x x
V V
V V V
:
:
(5.132)
NOTES ON CONTINUUM MECHANICS

318
Then by substituting the equations (5.130) and (5.132) into the energy expression in (5.125)
we obtain:
( ) ( ) [ ] [ ]
( ) [ ] [ ]
} } } }
} }
- - - - - -
- - - |
.
|

\
|
-


Z Z -
Z Z -
V V V
V
rdV dV dS dV
dS u dV
Dt
Du

) (
2
1
j j
j j j
v v v
v v v v v
x x
,
,
,
,
,
,
, , , ,
`
, ,
, ,
b n q q D
n
V V
c
:
(5.133)
or
( )
( ) ( ) [ ] [ ] 0

) (

2
1
- - - - -
- - - - - - |
.
|

\
|
-
}
}
Z
Z -


dS u
dV r
Dt
Du
V
n q
b q D
,
, , , , ,
,
,
,
,
`
, ,
, ,
v v v v
v v v v
x x
c
V V
j j
j j j :
(5.134)
( ) ( )
( ) ( ) [ ] [ ] 0

) (

2
1
- - - - -
- - - - - - - |
.
|

\
|
}
}
Z
Z -

dS u
dV r
Dt
Du
V
n q
b q D
0
,
, , , , ,
_
,
`
, ,
,
,
, ,
v v v v
v v
x x
c
V V
j j
j j j j :

(5.135)
with which we can conclude that:
( ) ( ) [ ] [ ] 0

) (
2
1
- - - - - - - -
} }
Z Z -
dS u rdV u
V
n q q D
,
, , , , ,
,
`
,
v v v v
x
c V j j j j : (5.136)
which thereby results in the energy equation for volumes with discontinuity:
( ) ( ) [ ] [ ] Z - - - -
- -


0

) (

2
1
on u
V in r u
n q
q D
,
, , , , ,
,
`
,
v v v v
x
c
V
j j
j j :
The energy equation with
discontinuity
(5.137)
5.10 The Principle of Irreversibility. Entropy
Inequality
5.10.1 The Second Law of Thermodynamics
Before applying the second law of thermodynamics, we define entropy which is a state
function. In thermodynamics, entropy is the physical quantity that measures the energy that
can not be used to produce work. In a broader sense, entropy is interpreted as the
measurement of system disorder. The entropy unit is K J / , joules per Kelvin and a process
characterized by constant entropy is called the isentropic process.
The second law of thermodynamics imposes restrictions on the possible direction of the
thermodynamics process. For example, the first law of thermodynamics does not establish
the direction of the heat flux.
The second law of thermodynamics states that the rate of change of the total entropy H is never
less than the sum of the entropy flow s
,
that enters through the surface of the continuum plus the entropy
created inside the continuum B .
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

319
The total entropy of the system ( H ) is given by:
} }

0
0 0
) , ( ) , ( ) (
V V
dV t dV t t H X x
,
,
j j j j

(

K
J
(5.138)
where ) , ( t x
,
j is the specific entropy (per unit mass), [ ]
kgK
J
j .
The entropy supplied to the system ( B ) is given by:
} }

0
0 0
) , ( ) , (
V V
dV t dV t X x
,
,
b b B j j

(

sK
J
(5.139)
where b is the source of local entropy per unit mass per unit time [ ]
K s kg
J

b .
Then the entropy flux that enters the system through the material surface is defined by:
}
-
s
s
,
,
S
dS

n

(

sK
J
(5.140)
Thus, we can set the entropy inequality as:
} } }
} } }

-
- I
s
s
s b
s b
,
,
, , ,
, , ,
`
S V V
S V V
dS dV t dV t
dS dV t dV t
Dt
D
t

) , ( ) , (

) , ( ) , ( ) (
n
n
x x
x x
j j j
j j j
(5.141)
Then by applying the divergence theorem to the surface integral, we obtain:
} } }
- I
V V V
dV dV dV t ) ( s b
,
,
`
x
V j j j
The second law of thermodynamics
(Entropy inequality)
(5.142)
NOTE: The global form of the entropy inequality in (5.142) implies that: if entropy occurs
then the process is irreversible, that is, we can not return to the original system without
adding work to the system. And, the equality of (5.142) represents a reversible process.
The local form of the equation in (5.142) is given by:
s b
, ,
,
` -
x
x V j j j ) , ( t (5.143)
and if we consider that:
) 1 ( ) 1 (
; b b s s - -
T
r
T
,
,
, q
(5.144)
where 0 ) , ( t T x
,
is the absolute temperature, [ ] K T , and by assuming that
) 1 (
s
,
and
) 1 (
b
are equal to zero, the entropy inequality in (5.143) becomes:
T
T T T
r
T T
r
x x x
, , ,
, ,
,
` V V V - -
|
|
.
|

\
|
- q q
q
2
1 1
j j j j (5.145)
Thus,
NOTES ON CONTINUUM MECHANICS

320
0
1 1 ) , (
) , (
0
) , (
) , (
2
- - -

|
|
.
|

\
|
- -

T
T T T
t r
t
T T
t r
t
x x
x
x
x
x
x
, ,
,
, ,
,
,
,
,
,
`
`
V V
V
q q
q
j j j
j j j
Entropy inequality
(current configuration)
(5.146)
We can also express the entropy inequality given in (5.143) in the reference configuration
as:
) , ( ) , ( ) , (
0 0
t t t X X X
X
, , , ,
,
` S b - V j j j (5.147)
where S
,
is the entropy flux vector in Lagrangian description. For thermal processes, the
entropy flux vector and entropy source can be established, respectively, as:
1 1
0
) , (
) , ( ; ) , ( b b S S - -
T
t r
t
T
t
X
X X
,
, ,
,
, ,
q

(5.148)
Then if we take into account the equation in (5.148) where
1
S
,
and
1
b are equal to zero,
the equation in (5.147) becomes:
T
T T T
t r
t
T T
t r
t
X X
X
X
X
X
X
, ,
,
, ,
,
,
, ,
,
`
`
V V
V

- -
|
|
.
|

\
|
-
0
2
0 0 0
0
0 0
1 1 ) , (
) , (
) , (
) , (
q q
q
j j j
j j j

Entropy inequality
(reference configuration)
(5.149)
Then if we refer to Eq. (5.120), where we obtained
T
J
-
F q q
, ,

0
, or in indicial notation
1
0

-

ik k i
F J q q
, ,
, it is true that:
_
, , , ,
_
, ,
,
Spatial
k
k
p
pk k pi
p
ik k
i
p
p
ik k
Material
i
i
x
T
J
x
T
J F
x
T
F J
X
x
x
T
F J
X
T
T
o
o

o
o

o
o

o
o
o
o

o
o

- -
q q q q q
1 1
0 0
c
X
V q
Then, we can prove the following relationship is valid:
) , ( ) , ( ) , ( ) , (
0
t T t J t T t x x X X
x
X
, ,
,
, ,
,
, ,
V V q q (5.150)
5.10.2 The Clausius-Duhem Inequality
If we combine the entropy inequality in (5.146) with the energy equation given in (5.114),
q D q D
,
`
,
`
, ,
- - - -
x x
V V r u r u j j j j : : , we obtain:
( )
( ) 0
1

1
0
1 1 1 1
2
2 2

- - -
- - - - - -


T
T
u
T
T
T
r
T
T
T T T
r
x
x x x x
,
, , , ,
,
`
, , , ,
`
` `
V
V V V V
q D
q q q q
: j j j
j j j j j j
(5.151)
In this scenario, the entropy inequality is called the Clausius-Duhem inequality, and is given by:
0
1

1 1
) , (
2
- - - T
T
u
T T
t
x
x
,
,
`
,
` V q D j j j :
The Clausius-Duhem inequality
(current configuration)
(5.152)
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

321
We can also express the Clausius-Duhem inequality in the reference configuration. From
the equation in (5.124) we obtained E X
X
`
`
,
,
,
: S q - - u t r
0 0 0
) , ( j j V and by substituting
this into the entropy inequality given in (5.149) we obtain:
( )
( ) T
T
u
T
t
T
T
t r
T
t
X
X X
E X
X X
,
, ,
,
`
`
,
, ,
, ,
`
`
V
V V


- -
- -
0
2
0 0
0
2
0 0 0
1 1
) , (
1
) , (
1
) , (
q S
q q
: j j j
j j j
(5.153)
or:
0
1 1 1
) , (
0
1 1 1
) , (
0
2
0 0
0
2
0 0
- - -
- - -

T
T
u
T T
t
or
T
T
u
T T
t
X
X
F X
E X
,
,
,
`
`
,
,
`
`
,
`
`
V
V
q P
q S
j j j
j j j
:
:
The Clausius-Duhem inequality
(reference configuration)
(5.154)
5.10.3 The Clausius-Planck Inequality
Note that the inequality 0 s T
x
,
,
V q is always valid, since the orientation of the heat flux
vector ( q
,
) is always opposite to the temperature gradient ( T
x
,
V ), (see Figure 5.10). Then,
we can formulate the heat conduction inequality:
0 - T
x
,
,
V q (current configuration)
Heat conduction inequality
0
0
- T
X
,
,
V q (reference configuration)
(5.155)
If we now incorporate the restrictions in (5.155) into the Clausius-Duhem inequality
(5.152) and in (5.154) we will have a less restrictive inequality known as the Clausius-Planck
inequality:
0 ) , (
1 1
) , ( - - t u
T T
t
int
x x
,
`
,
` j j j D : D (current configuration)
Clausius-Planck
inequality
0 ) , (
1 1
) , (
0 0
- - t u
T T
t
int
X F X
,
`
`
,
` j j j : P D
(reference
configuration)
(5.156)
where
int
D is the internal energy dissipation, which requires positiveness at any time,
0
int
D .
5.10.4 The Alternative Form to Express the Clausius-Duhem
Inequality
An alternative form of entropy inequality is that expressed in terms of the Helmholtz free
energy, , which is a thermodynamic potential per unit mass and is given in Eulerian description
by:
j T u - The Helmholtz free energy
(

kg
J

(5.157)
NOTES ON CONTINUUM MECHANICS

322
NOTE: A thermodynamic potential indicates the amount of energy available in the
system. In this chapter we only work with the potentials ) , ( j E u u and ) , ( T E . We can
also use another potential, e.g. Gibbs free energy ( ) , ( T S G ), or Enthalpy ( ) , ( j S H ), (see Chapter
10). The choice to adopt one or the other depends on the independent variables under
consideration, ( E -volume, j -entropy, S -pressure, T -temperature). For more details
about these potentials see the chapter on Thermoelasticity.












Figure 5.10: Temperature gradient and heat flux vector.
If we calculate the rate of change of the Helmholtz free energy, we obtain:
[ ] j j j
j jj j j j
j j j j
` `
`
` `
` `
`
` `
` `
`
`
`
- -
- -
- - - -
T u
T u T
T u T T T u
(5.158)
Then if we consider that 0 > T (absolute temperature) and the entropy inequality given in
(5.145), we obtain:
T
T
r T T
T T T
r
x x x x
, , , ,
, , , ,
` ` V V V V - - - - q q q q
1 1 1

2
j j j j j j (5.159)
Afterwards by combining the above inequality with the equation in (5.158) we obtain:
[ ] T
T
r T u
x x
, ,
, ,
`
`
`
V V - - - - q q
1
j j j j (5.160)
Then by also considering the energy equation in (5.114), i.e. r u j j - - q D
,
`
,
x
V : , we
obtain:
[ ]
[ ] T
T
T
T
T
r T r
x
x x x
,
, , ,
,
`
, ,
`
,
`
`
V
V V V


- -
- - - - - -
q D
q q q D
1
1

j j
j j j j
:
:

(5.161)
by which we obtain the Clausius-Duhem inequality (current configuration) in terms of the
Helmholtz free energy:
[ ] 0
1
- - - T
T
x
,
,
` `
V q D j j T :
Clausius-Duhem inequality
(current configuration)
(5.162)

3 2 1
T T T > >
T
x
,
V

n
q

,


3
T

2
T

1
T


n



n q q
n
n q n n q q
n n


, ,
, , ,
, ,
, ,
T T
T T
x x
x x
V V
V V
q
,

T T T
x x x
, , ,
, , , ,
V V V -
n s n
q q q q

) (
s


5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

323
The Clausius-Duhem inequality in the reference configuration, (see Eq. (5.154)),
[ ] 0
1
0
1
0 0 0 0 0
- - - - - - T
T
T u T
T
u T
X X
E E
, ,
,
`
`
,
`
`
` ` V V q S q S j j j j j : : (5.163)
can also be written in terms of the Helmholtz free energy . To do this let us consider the
Helmholtz free energy in Lagrangian description ) , ( ) , ( ) , ( t t T t u X X X
, , ,
j - . Additionally,
the rate of change is given by j j j j T T u T T u
`
`
`
`
`
` `
`
- - - - with that the Clausius-
Duhem inequality in the reference configuration becomes:
[ ]
[ ] 0
1
0
1
0 0
0 0
- - -
- - -

T
T
T
T
T
T
X
X
F
E
,
,
,
` `
,
` `
`
`
V
V
q P
q S
j j
j j
:
:
Clausius-Duhem inequality
(reference configuration)
(5.164)
The Helmholtz free energy per unit reference volume is denoted by j 1 ) (
0
X
,
, and it holds
that j 1
` `
0
. Proof of this can be shown by:
j

j
j

j 1
1
` `
_
0 0
0
0 0
) ( ) ( ) (
-

=
Dt
D
Dt
D
Dt
D
Dt
D
.
5.10.5 The Alternative Form of the Clausius-Planck
Inequality
The Clausius-Planck inequality can also be expressed in terms of Helmholtz free energy.
Then, if we consider the heat conduction inequality, 0 - T
x
,
,
V q , the equation in (5.162)
becomes:
[ ] 0 - - j j
` `
T
int
D : D
(5.165)
which in the reference configuration is given by:
[ ] 0
0
- - j j
` ` `
T
int
E : S D
(5.166)
5.10.6 Reversible Process
A thermodynamic process is said to be reversible if there is no dissipation of energy, i.e.
0 ) ( I t , (see equation (5.142)). A reversible process is characterized by:
The work done by the forces between two points being independent of the path;
The work done in a closed cycle being zero.
If we take into account that the dissipation of energy is equal to zero in a reversible process
we obtain:
0 - j
`
D :
int
D
D : j
`
(5.167)
Then the equation in (5.167) in the reference configuration becomes:
0
2
1
0
0
0
-
-
j
j
`
`
`
`
C
E
:
:
S
S
int
D

E C
` ` `
: : S S
2
1
0
j
(5.168)
NOTES ON CONTINUUM MECHANICS

324
5.10.7 Entropy Inequality for a Domain with Discontinuity
If we applying the entropy inequality in (5.141) for a volume with discontinuity we obtain:
} } }
- - - - - -
- - -
- I
S S V V V V
dS dV t dV t
Dt
D
t

) , ( ) , ( ) ( n s b
, , ,
x x j j j
(5.169)
For the surface integral on the right side of the inequality in (5.169) we can apply the
divergence theorem with discontinuity given in Eq. (5.17), the result of which is:
[ ] [ ]
} } }
Z - -
-
- - - -
dS dV dS
V V S S


n n s s s
, , ,
V
(5.170)
For the volume integral on the left side of the inequality in (5.169) we can apply the
Reynolds transport theorem given in (5.22) in which j j d , then:
( )
( ) ( ) [ ] [ ]
} } }
Z Z - Z -
- - |
.
|

\
|
-
(
(

dS dV
Dt
D
dV
Dt
D
V V

n c V
, , ,
,
v v
x
j j j j
j j
j j
(5.171)
Then by substituting the equations in (5.171) and (5.170) into (5.169), we obtain:
( )
( ) ( ) [ ] [ ]
[ ] [ ]
} }
} } }
Z Z -
- Z Z -


- -
- - - |
.
|

\
|
-
- -
dS dV
dV t dS dV
Dt
D
V
V V V

) , (

n
n
s s
b
, ,
, , , ,
,
x
x v v
V
c V j j j j j
j j

(5.172)
Additionally by regrouping the integrands we obtain:
( ) ( ) [ ] [ ] 0

- - - |
.
|

\
|
- - - -
} }
Z Z -
dS dV
Dt
D
Dt
D
V
n s b s
, , , , ,
, ,
c V V v v
x x
j j j j j
j
j
j
j
(5.173)
Note that the equation ( ) 0 |
.
|

\
|
- - v v
x x
, ,
, ,
V V j
j
j j j
j
j
Dt
D
Dt
D
is valid due to the mass
continuity equation. Then, the equation in (5.173) becomes:
( ) [ ] [ ] 0

- - - |
.
|

\
|
- -
} }
Z Z -
dS dV
Dt
D
V
n s b s
, , , ,
,
c V v
x
j j j
j
j (5.174)
The local form of the above equation is expressed as:
( ) [ ] [ ] Z - -
Z - -

on
V in
Dt
D
0

n s
b s
, , ,
,
,
c
V
v
x
j j
j
j
j
Entropy inequality with
discontinuity
(5.175)
Problem 5.11: 1) Consider a continuum motion in which the stress power is equal to zero.
Also, consider that the heat flux is given by T T
x
,
,
V - ) ( K q , which is known as Fouriers
law of thermal conduction, where ) (T K is a second-order tensor called the thermal conductivity
tensor (the thermal property of the material), and
T
T u
c
o
o

) (
, where c is the specific heat
capacity at a constant deformation (the thermal property of the material) and is expressed in
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

325
units of joule per kelvin, i.e. [ ]
K
J
c . Taking into account all previous considerations, find
the energy equation for this process. Then also provide the unit of ) (T K in the
International System of Units (SI).
2) Consider the stress power is equal to zero, and that there is a continuous medium with
no internal heat source. Also consider that there is a heterogeneous material where
) ( x
,
K K is an arbitrary second-order tensor (not necessarily symmetrical). a) Show that
the thermal conductivity tensor is semi-definite positive, b) Check in which scenario the
skew part of ) ( x
,
K does not affect the outcome of the heat conduction problem. c) Taking
into account that the material is isotropic, in what format is K ?
Solution: For this problem we know that the stress power is equal to zero, 0 D : . It then
follows that, the energy equation becomes:

r r
t
T
T
u
u j j j j - - - -
o
o
o
o

q q D
, ,
`
, ,
x x
V V
0
:
[ ] r T T
t
T
c r
t
T
c j j j j - - -
o
o
- -
o
o

x x x
, , ,
,
V V V ) ( K q
or
[ ]
t
T
c r T T
o
o
- j j
x x
, ,
V V ) ( K
The above equation is called the heat flux equation which is applied to the thermal
conduction problem.
Then if we take into account the following units: [ ]
2 2
m
W
s m
J
q
,
,
m
K T
T
(

o
o
=
x
x
,
,
V , we
can ensure that the units are consistent if the following is met:
[ ] [ ] [ ]
(



m
K
K m
W
K m s
J
m
W
s m
J
T

2 2
x
,
,
V K q

thus, we can draw the conclusion that [ ]
(


K m
W
K m s
J

K .
NOTE: As we will see later, when the stress power is equal to zero, we can decouple the
thermal and mechanical problem. That is, we can study these problems separately.
2) a) We start from the heat conductivity inequality:
0 ) (
0 ) ) ( (

- - -


T T
T T T
x x
x x x
x
x
, ,
, , ,
,
,
,
V V
V V V
K
K q
or
0
0 ) (
, ,
, , ,

- - -
j ij i
i j ij i i
T K T
T T K T q

Remember that the arbitrary tensor A is semi-definite positive if it holds that 0 x x
, ,
A
for all 0
,
,
x thereby demonstrating that ) ( x
,
K is a semi-definite positive tensor. Then, as a
result the eigenvalues of ) ( x
,
K are all real values greater than or equal to zero, i.e. 0
1
K ,
0
2
K , 0
3
K . Also remember that since ) ( x
,
K is not symmetric, the principal space of
) ( x
,
K does not define an orthonormal basis. Moreover, it is noteworthy that: the
antisymmetric part of ) ( x
,
K does not affect the heat conduction inequality since:
[ ]
0 ) (
0
0 ) (
-
-
-



T T T T
T T T T
T T T T
skew sym
skew sym
skew sym
x x x x
x x x x
x x x x
x
, , , ,
, , , ,
, , , ,
,
V V V V
V V V V
V V V V
: K K
K K
K K K

NOTES ON CONTINUUM MECHANICS

326
Notice that 0 ) ( T T
skew
x x
, ,
V V : K , since the double scalar product between an
antisymmetric tensor (
skew
K ) and a symmetric one ) ( T T
x x
, ,
V V is equal to zero, then:
0 ) ( 0 s T T T T
sym
x x x x
x
, , , ,
,
V V V V K K
That is, the above inequality is always true whether ) ( x
,
K is symmetric or not.
b) For the proposed problem the only remaining governing equation is the energy
equation: q q D
, ,
, ,
` - - - =
x x
V V r u
Dt
Du
j j j : , where u is the specific internal
energy, D : is the stress power, and r j is the internal heat source per unit volume. Then:
[ ]
) ( ) ( ) (
) ( ) ( ) ( ) (
) ( ) ( ) (
) ( ) ( ) ( ) (
, , , , , ,
T T
T T T
T T
T T T T T
sym T
skew sym T
skew sym T
T
ji ij j i ij i j ij i i
u
x x x x
x x x x x x
x x x x
x x x x
, , , ,
, , , , , ,
, , , ,
, , , ,
`
V V V V
V V V V V V
V V V V
V V V V
:
: :
:
:
K K
K K K
K K K
K K
-
- -
- -
- - - - -



K K K q j

where we have considered the symmetry of [ ]
ji ij ij
T T T
, ,
) (
x x
, ,
V V . If the material is
homogeneous the implication is that the K field does not depend on ) ( x
,
, so
j i ij
0 K
,
. In
this scenario the heat equation reduces to:
) ( T
sym
u
x x
, ,
` V V : K j
Therefore, when the material is homogeneous, the antisymmetric part of K does not affect
the outcome.
c) The feature of isotropic materials is that their properties (at one material point) do not
change if the coordinate system is changed. It follows then that K must be an isotropic
tensor. An isotropic second-order tensor has the format of a spherical tensor, (see Chapter
1), then the tensor K must be of the type: 1 K K , where K is a scalar.
5.11 Fundamental Equations of Continuum
Mechanics
Then, we sum up the fundamental equations of continuum mechanics in the current
configuration as:
Fundamental Equations of the Continuum Mechanics
(Current configuration)

Mass Continuity Equation
(Principle of conservation of mass)
0 ) ( - v
x
,
,
V j
j
Dt
D
(1 equation) (5.176)
Equation of Motion
(Principle of conservation of linear
momentum)
v
x
`
,
,
,
j j - b V (3 equations) (5.177)
Symmetry of the Cauchy Stress Tensor
(Principle of conservation of angular
momentum)
T
(6 unknowns) (5.178)
Energy Equation
(Principle of conservation of energy)
r u j j - - q D
,
`
,
x
V : (1 equation) (5.179)
Entropy Inequality
(Principle of irreversibility)
0
1

1 1
) , (
2
- - - T
T
u
T T
t
x
x
,
,
`
,
` V q D j j j : (5.180)
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

327
The entropy inequality is not one more problem equation. Rather, it is used to establish
restrictions on the problem variables. The symmetry of the Cauchy stress tensor reduces
the number of -unknown from 9 to 6.
The mass continuity equation, the equations of motion and the energy equation give us in
total 5 equations. The unknowns are: the three components of velocity v
,
, temperature T ,
mass density j , six components of the Cauchy stress tensor , the specific internal energy
u , three components of the heat flux vector q
,
, and the entropy j , with a total of 16
unknowns.
To achieve the well-posedness of the problem eleven equations must be added. We must
add equations that connect the stress, heat and energy with other fields. These equations
are called constitutive equations, which is the subject of the next chapter.
5.11.1 Particular Cases
5.11.1.1 Rigid Body Motion
When we are dealing with rigid body motion without the effect of temperature, the only
principles needed to establish the set of equations are: the principle of conservation of the
linear momentum and the principle of conservation of angular momentum. Then, the
governing equations are characterized by a
, ,
m

F and
G G
H M
`
, ,

, (see Problem 5.9).


The problem can then be solved by introducing the appropriate initial and boundary
conditions.
5.11.1.2 Flux Problems
For problems which only involve the transport of a physical quantity (mass, energy, or
otherwise) the only principle necessary to establish the governing equations is the
conservation law of the physical quantity, or in its strong form: the physical quantity
continuity equation, (see equation (5.24)):
( ) ) , (
) , (
) (
) , (
t
t
t
t
t
Q x
x
v
x
x x
,
,
,
,
,
, ,
q -
o
d o
= d -
o
d o
V V
[ ]
s
d
(5.181)
The case in Problem 5.11 was related to energy transport (not mass transport). Said energy
transport exists due to the agitation of atoms, in which the degree of agitation at the
macroscopic level is characterized by the temperature. If a particle starts to increase the
degree of agitation, then neighboring particles also start behaving in a similar fashion. In
this way the energy in solids is transported without any mass transport. This energy
transport at the macroscopic level is represented by means of the flux, v
,
,
d q .
When we are working in the field of continuum mechanics we do not go down to the
atomic level and measure the average velocity (vibration) of a handful of atoms to establish
the flux. What we do is: we go to the laboratory with the material with which we want to
establish the heat flux (energy flow), we vary the temperature and we verify
macroscopically that the flux can be characterized by the following phenomenological law
T k q V - (in a one-dimensional case), where k is a thermal property of the material. This
procedure was performed by Fourier, thereby establishing Fouriers law of heat conduction.
Fourier also verified in the laboratory that heat flux is opposite to the temperature gradient,
a fact already proven by the second law of thermodynamics. The law T k q V - is a
phenomenological law or constitutive equation of heat flux, and connects two thermal
NOTES ON CONTINUUM MECHANICS

328
variables. It is also interesting to observe that the fundamental equations of continuum
mechanics (5.176)-(5.180) do not have such a relationship.
In Problem 5.11, the physical quantity in question is given by cT j d . According to the
SI units we have: [ ] K T , [ ]
3
m
kg
j , [ ]
K kg
J
c

with which we can verify the following SI
units:
[ ] [ ]
3 3
m
J
K
K kg
J
m
kg
cT d j (unit of energy per unit volume - energy density)
[ ] [ ]
s m
J
s
m
m
J
2 3
d = v
,
,
q (unit of energy flux)
There are several engineering problems which are characterized by the continuity equation,
some of which are: heat conduction problems (energy flux); filtration problems in porous
media (mass transport); diffusion problems (e.g. transport of contaminant in an aqueous
medium); and the Saint-Venant torsion problem (stress flux).
5.12 Flux Problems
5.12.1 Heat Transfer
Heat flow is a form of energy transfer in a continuous medium which occurs in three ways,
namely via: conduction; convection; radiation.
5.12.1.1 Thermal Conduction
Thermal Conduction: Transfer of energy in the form of heat, which is caused by the collision
and vibration of molecules and atoms (no mass transport).
Temperature: The temperature ( 0 ) , ( > t T x
,
) is not a form of energy. Rather it is a
measurement of how hot a particle is. In experiments it has been proven that the hot
particles tend to give heat to cooler particles. The SI unit of absolute temperature is the
kelvin, [ ] K T . Absolute zero C K T 15 , 273 0 - = is a theoretical temperature when even
atoms and electrons cease to move.
When a continuous medium undergoes a non-uniform temperature variation, heat is
transferred from a higher to a lower temperature region. When this phenomenon occurs
without mass transport, this is known as a heat conduction problem. The phenomenological law
(constitutive equation of heat flux) that governs heat conduction behavior can be defined by
means of Fouriers law of heat conduction, which states that the heat flux is proportional to the
temperature gradient:
T
T
x
x
,
,
,
V -
o
o
- K K q
Fouriers law of heat
conduction
(

s m
J
2
(5.182)
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

329
where q
,
is the heat flux per unit area per unit time, and its SI unit [ ]
2 2
m
W
s m
J
q
,
; T
x
,
V is
the temperature gradient whose SI unit is [ ]
m
K
T
x
,
V , and K is the thermal conductivity
tensor, whose SI unit is [ ]
mK
W
K , (see Problem 5.11).
NOTE: Fouriers law of heat conduction is not universal as there are complex materials in
which heat flow is governed by more complex laws.
The negative sign in Fouriers law is there because the heat flux vector is always opposite to
the temperature gradient. The temperature gradient vector ( T
x
,
V ) points from the coldest
to the warmest region, while the heat flux vector ( q
,
) points from the warmest to the
coldest region (physical fact), (see Figure 5.10).










Figure 5.11: Heat conduction.
The thermal conductivity tensor contains the thermal properties of the material, which are
obtained in the laboratory, and depends on porosity, mass density, composition, etc.
Explicitly, the components of K are:

(
(
(

material isotropic
ij
33 32 31
23 22 21
13 12 11
) (
K K K
K K K
K K K
K
(
(
(

1 0 0
0 1 0
0 0 1
) ( K
ij
K
(5.183)
For isotropic materials, i.e. those that have the same property in any direction, the thermal
conductivity tensor is represented by a spherical tensor, (see Problem 5.11).
If we are dealing with homogenous material, K is not dependent on x
,
. For isotropic
materials, the components of the heat flux vector ( T
x
,
,
V - K q ) are obtained as follows:
(
(
(
(
(
(
(
(

o
o
o
o
o
o
-
(
(
(
(
(
(
(
(

o
o
o
o
o
o
(
(
(

3
2
1
3
2
1
3
2
1

1 0 0
0 1 0
0 0 1
) (
x
T
x
T
x
T
x
T
x
T
x
T
i
K K
q
q
q
q
,
(5.184)
conduction



conduction
NOTES ON CONTINUUM MECHANICS

330
and the normal component n q


,
n
q , (see Figure 5.2), is evaluated as:
3
3
2
2
1
1
3 3 2 2 1 1

n K n K n K n q n q n q n q q
x
T
x
T
x
T
i i n
o
o
-
o
o
-
o
o
- - -
(5.185)
5.12.1.2 Thermal Convection Transfer
Heat transfer by convection occurs in a fluid environment where there are moving particles
between regions with different temperatures, (see Figure 5.12). In other words it shows the
transfer of energy (heat) due to the movement of fluid particles. This phenomenon is
governed by Newtons Law of Cooling, which is:
( )
ext
T T - c q
Newtons law of cooling (5.186)
where q is the thermal energy; c is the heat transfer coefficient per unit area; T is the
temperature of the bodys surface, and
ext
T is the temperature of the surrounding
environment.







Figure 5.12: Thermal convection.
If we consider a room in which there is a radiator, the air particles in contact with the hot
surface of the radiator increases their temperature and their mass density decreases, so that
the hot ascending particles, displace the cooler particles moving downwards, (see Figure
5.12) and because of this movement, the heat will be transferred to the whole room.
5.12.1.3 Thermal Radiation
Thermal radiation is the process by which thermal energy is transferred between two
surfaces, obeying the laws for electromagnetic radiation (photon transport). To give an
example we can mention how heat is transferred from the Sun to the Earth. The
phenomenological law governing this phenomenon is the Stefan-Boltzmann law.
5.12.1.4 The Heat Flux Equation
Next, we can obtain the partial differential equation that governs the heat transfer problem,
by means of an energy balance, i.e.:




+
Heat
generated
internally
_
=
Change of
internal
energy
Heat that
leaves the
system
Heat that
enters into
the system

warm air
cold air
radiator
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

331
Let us consider a differential volume element, (see Figure 5.13), in which there are inflows
and outflows of heat. In addition let us consider energy generated internally represented by
r Q j (per unit volume per unit time), whose SI unit is [ ]
s m
J
Q
3
. The scalar function r
describes the heat generated which could be caused by a phenomenon such as a chemical,
or nuclear reaction, and whose SI unit is [ ]
s kg
J
r

. As the temperature of the body
increases, part of the thermal energy is stored in the body. For a differential volume
element (
3 2 1
dx dx dx ) this stored energy is governed by the expression:
3 2 1
dx dx dx
t
T
c
o
o
v
j (5.187)
where j is the mass density; and the material property
v
c is the specific heat capacity at a
constant volume whose SI unit is [ ]
K kg
J
c
v

.










Figure 5.13: Source and heat flux in a differential volume element.
In the following demonstration, let us consider the following change of nomenclature:
coordinates: x x =
1
, y x =
2
, z x =
3
; heat flux components;
x
q q =
1
,
y
q q =
2
,
z
q q =
3
.
Notice that we are employing the engineering notation.
Then by applying the energy conservation law throughout the differential element, we
obtain:
t
T
c
z y x
Q
dxdydz
t
T
c dxdy dz
z
dxdz dy
y
dydz dx
x
dz dy dx Q dy dx dz dx dz dy
z
y
x
v
z
z
y
y
x
x z y x
o
o

o
o
-
o
o
-
o
o
-
o
o

(
|
.
|

\
|
o
o
- -
|
|
.
|

\
|
o
o
- -
|
.
|

\
|
o
o
-

- - - -
v
j
j
q
q
q
q
q
q
q
q
q q q q

(5.188)
which results in the heat equation:
( )
t
T
c T Q
t
T
c Q
o
o
-
o
o
-
v x x v x
j j
, , ,
,
V V V K q
The heat flux equation (5.189)

z
q
z

x
q
dy
y
y
y
o
o
-
q
q
dx
x
x
x
o
o
-
q
q
dz
z
z
z
o
o
-
q
q

y
q
x
y
Q
dy
dx
dz
NOTES ON CONTINUUM MECHANICS

332
where we have considered Fouriers law of heat conduction
j
ij i
x
T
o
o
- K q . Notice that the
above equation was obtained in Problem 5.11 and it should be pointed out that we have
one equation in (5.189) and one unknown (temperature). The solution of equation (5.189)
is unique if we are given the appropriate boundary and initial conditions. The governing
equation in (5.189) together with the boundary and initial conditions are called the Initial
Boundary Value Problem (IBVP) of thermal conduction.
Then by considering an isotropic homogeneous material, the heat equation in (5.189)
becomes:
t
T
c
z
T
z y
T
y x
T
x
Q
o
o
|
.
|

\
|
o
o
o
o
-
|
|
.
|

\
|
o
o
o
o
- |
.
|

\
|
o
o
o
o
-
v
j K K K
(5.190)
t
T
z
T
y
T
x
T Q
o
o

o
o
-
o
o
-
o
o
-
r
1
2
2
2
2
2
2
K
(5.191)
where r is known as the thermal diffusivity:
v
c j
r
K

(

s
m
2
(5.192)
Particular Cases
A steady state temperature field, i.e. ) ( x
,
T T :
0
o
o
t
T
(5.193)
The equation in (5.191) becomes:
0 0
2
2
2
2
2
2
2
-
o
o
-
o
o
-
o
o
- T
Q
z
T
y
T
x
T Q
x
,
V
K K
The Poissons equation (5.194)
From a mathematical point of view, the above equation is known as the Poissons
equation.
A steady state problem, and without internal heat generation:
0 ; 0
o
o
Q
t
T
(5.195)
In this scenario the equation in (5.191) becomes Laplaces equation:
0 0
2
2
2
2
2
2
2

o
o
-
o
o
-
o
o
T
z
T
y
T
x
T
x
,
V Laplaces equation (5.196)
Transient problem, ) , ( t T T x
,
(time dependent), but in the absence of internal heat
generation, 0 Q , the equation in (5.191) becomes Fouriers equation:
t
T
T
t
T
z
T
y
T
x
T
o
o

o
o

o
o
-
o
o
-
o
o
r r
1 1
2
2
2
2
2
2
2
x
,
V Fouriers equation (5.197)
Initial and Boundary Conditions
1. Prescribed value of the temperature:
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

333
1
*
0 ) , , , ( S on t to T t z y x T > (5.198)
Mathematically this condition is known as Dirichlet boundary condition.
2. Flux boundary condition:
2
0 0

S on t to
z
T
y
T
x
T
Q
z y x
>
o
o
-
o
o
-
o
o
- n K n K n K
(5.199)
Mathematically this condition is known as a Neumann boundary condition.
A combination of the boundary conditions of Dirichlet and Neumann is known as the
Robin boundary condition, i.e.:
3
0 0 ) (

S on t to T T
z
T
y
T
x
T
Q
ext z y x
> - -
o
o
-
o
o
-
o
o
- c n K n K n K
(5.200)
where
z y x
n n n

are the components of the outward unit normal vector on the surface.
Initial conditions
0
) 0 , , , ( T t z y x T .











Figure 5.14: Heat flux problem and boundary conditions.











o
cos

cos

cos

z
y
x
n
n
n


ext
T

3
S

1
S

*
q
,

B
r j

*
T

2
x

1
x

3
x
O
t
dV
x
z
y
n



NOTES ON CONTINUUM MECHANICS

334
5.13 Fluid Flow in Porous Media (filtration)
Let us consider a reservoir as shown in Figure 5.15. To obtain the partial differential
equation that governs the fluid flow in porous media we will make the same approach as
that made to the heat flux problem, but in this case we will consider the two-dimensional
case and steady state regime.














Figure 5.15: Fluid flow throughout the porous medium.
The partial differential equation governing the fluid flow in porous media for a steady state
case can be obtained by means of the differential element equilibrium, (see Figure 5.15), i.e.:
0 0
o
o
-
o
o
-
(
(

|
|
.
|

\
|
o
o
- - -
(

|
.
|

\
|
o
o
- - dxdy
y
dxdy
x
dx dy
y
dy dx
x
y
x
y
y y
x
x x
q
q
q
q q
q
q q (5.201)
The phenomenological law of mass flux in porous media is governed by Darcys law,
c
x
,
,
V - K q , where c is the total potential (water level), and K is the permeability tensor
which depends on the material. In the two-dimensional case, the components of q
,
are
given by:
y x
y x
o
o
-
o
o
-
c c
K q K q ;
(5.202)
where we have considered an isotropic material (one which has the same permeability in all
directions). Then by substituting the components of the flux vector in Equation (5.201), we
obtain:
0
|
|
.
|

\
|
o
o
o
o
- |
.
|

\
|
o
o
o
o
y y x x
c c
K K
(5.203)
0 0
2
2
2
2
2

|
|
.
|

\
|
o
o
-
o
o
c
c c
V
y x
K Laplaces equation (5.204)
dx
x
x
x
o
o
-
q
q
dam (impermeable)

2
h

1
h
x
y
dx
dy
soil (permeable)
Flux
P

x
q

y
q
L
rock (impermeable)
dy
y
y
y
o
o
-
q
q
b
P
material point - P
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

335
The boundary conditions are:
There is no flow at - x and - x : 0 ; 0
o
o

o
o
- - x x
x x
c c

There is no flow at the border (soil-rock interface): 0
0

o
o
y
y
c

There is no flow at the soil-dam interface:
2 2
; 0 ) , (
b
x
b
L x
y
s s -
o
oc

Additionally, the total potential is prescribed at the water-soil interface:
2
2
1
2
) , ( ; ) , ( h L x h L x b
x
b
x

>
-
<
c c
(5.205)
5.14 The Convection-Diffusion Equation
Diffusion: An irreversible physical process is one where particles which are in a high
concentration region tend to move to a region of low concentration. In general
this process is governed by Ficks law of diffusion:
c
x
,
,
V - D J Ficks law of diffusion
(

s m
mol
2
(5.206)
where 0 D> is the diffusion tensor (or diffusivity tensor), and ) , ( t c x
,
is the solute
concentration whose SI unit is [ ]
3
m
mol
c . This concentration is defined as follows:
mass fluid
mass solute
c (5.207)
or we can express this concentration as:
f
s
f
c
j
j
j

volume
mass solute 1

(5.208)
where
f
j ,
s
j are the mass densities of the fluid and solute, respectively.
In general, when we have a process where there is mass transport (a fluid+solute) two
mechanisms take place, namely: convection and diffusion. In this case the matter (solute) is
defined by the concentration c and we must consider the matter to be diffused throughout
the aqueous medium. Here, we can assume that the amount of the matter is too little to
affect the fluid velocity field.
Let us consider the solute flux v
,
,
c q (a convective term) to which we add the diffusive
term to obtain the total flux:

_
,
,
,
term
Diffusive
term
Convective
c
c
x
v
o
o
- D q
(5.209)
Then, to obtain the partial differential equation for the convection-diffusion problem we
consider the one-dimensional case, (see Figure 5.16).
NOTES ON CONTINUUM MECHANICS

336




Figure 5.16: Mass transport (solute).
Here we can put the conservation law down to:




Then, mathematically, the above expression becomes:
t
c
x
Q dxdy
t
c
dy dx
x
Qdxdy dy
x x
x x
o
o

o
o
-
o
o
|
.
|

\
|
o
o
- - -
q q
q q
(5.210)
Additionally, by substituting the flux given in (5.209) into the above equation, we obtain:
( )
t
c
x
c
x x
v c
Q
t
c
x
x
c
v c
Q
x
x
o
o
|
.
|

\
|
o
o
o
o
-
o
o
-
o
o

o
|
.
|

\
|
o
o
- o
- D
D


(5.211)
Therefore we can summarize the convection-diffusion equation in three dimensions as:
t
c
c c Q
o
o
- - ) ( ) (
x x x
v
, , ,
,
V V V D
Convection-diffusion
equation
(5.212)
where
t
c
o
o
is the local variation of the concentration with respect to time, c
x
v
,
,
V is the
convection term caused by fluid motion, and ) ( c
x x
, ,
V V D is the diffusion.
Next, we assume that at a material point there are two types of material that are
represented by a physical quantity per unit volume in such a way that
s f
c c c - , and the
following holds
s f
v v v
, , ,
- , (see Figure 5.17).








Figure 5.17: Heterogeneous medium.

x

x
q dx
x
x
x
o
o
-
q
q
dx x -
+
solute
generated
internally
_
=
Change of
the solute
internally
solute that
leaves the
system
solute that
enters into
the system

s
c
P

f
c

s
v
,


f
v
,

v
,

V
material point - P
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

337
Then, from the continuity equation for this physical quantity we obtain:
( ) [ ] ) )( (
) (
s f s f
s f
c c
t
c c
Q
t
Q v v
x
v
x
, ,
,
,
,
- -
o
o
-
o
- o
d -
o
d o
V
(5.213)
thus
[ ]
[ ]
[ ]
[ ]
s s s f f s
s
f f
f
s s f s s f f f
s f
s s f s s f f f
s f
s f s f
s f
c c c
t
c
c
t
c
Q
c c c c
t
c
t
c
Q
c c c c
t
c c
Q
c c
t
c c
Q
v v v v
v v v v
v v v v
x
v v
x
x x x
x
, , , ,
, , , ,
, , , ,
,
, ,
,
, , ,
,
- - -
o
o
-
(

-
o
o

- - - -
o
o
-
o
o

- - -
o
o
-
o
- o

- -
o
o
-
o
- o

V V V
V
) ( ) (
) (
) )( (
) (

(5.214)
If we assume that there is no ( f )-material source, then 0 ) ( -
o
o

f f
f
c
t
c
v
x
,
,
V holds,
which is the continuity equation of the quantity
f
c with which the equation in (5.214)
becomes:
[ ]
s f s f s s f s
s
s f s s f s
s
s s s f f s
s
c c c c
t
c
Q
c c c
t
c
Q
c c c
t
c
Q
v v v v
v v v
v v v
x x x x
x x x
x x
, , , ,
, , ,
, , ,
, , , ,
, , ,
, ,



- - - -
o
o

- - -
o
o

- - -
o
o

V V V V
V V V
V V
) ( ) (
) ( ) ( ) (
) (
(5.215)
If the physical quantity
f
c does not change with x
,
, then the gradient of
f
c becomes
0
,
,

f
c
x
V . In addition if we consider the medium ( s ) to be incompressible we obtain
0
s
v
x
,
,
V . These simplifications indicate that the material ( s ) does not affect the velocity
field of the material ( f ). So, if the amount of the material ( s ) is significant, this approach
is no longer valid. Then, with these approximations we obtain:
) (
) ( ) ( ) (
D f s
s
s s f s
s
c
t
c
c c
t
c
Q q
,
, , ,
, , , ,
- -
o
o
- -
o
o

x x x x
v v v V V V V
(5.216)
Notice that the term
) (
) (
D s s
c q
,
,
= v represents the flux caused by the ( s )-material
concentration, the diffusive term. The term
) (
) (
C f s
c q
,
,
= v is related to mass transport, the
convective term. Then, if
) (D
q
,
is defined by Ficks law we refer back to the equation in
(5.212).






NOTES ON CONTINUUM MECHANICS

338
5.14.1 The Generalization of the Flux Problem
Flux problems can be found in many branches of physics or engineering. These problems
are only governed by the continuity equation (sometimes called the transport equation):
( )
t
c Q
o
o
-
c
j c
x x
, ,
V V D (5.217)
where ) , , , (
3 2 1
t x x x c is the scalar variable to be solved. Depending on the problem the
variables take the following meanings:
( ) 0 - c
x x
, ,
V V D Q
Equation
Scalar field
c
D
Q
Flux vector
q
,

Phenomenological
law
Heat flux
Temperature
T
Thermal
conductivity
tensor
Heat
generated
Q
Heat flux
vector
q
,

Fouriers law
T
x
,
,
V - K q
Fluid flow in the
porous media
piezometric head
(or hydraulic head)
h
permeability
tensor
water
source
Volume flux
vector
Darcys law
c
x
,
,
V - K q
Diffusion
ion concentration
c
Constitutive
matrix for
diffusion
coefficient
Ion source
Ion flux
vector
J
,

Ficks law
c
x
,
,
V - D J
Saint-Venant
torsion
Prandtl stress
function
0

G
1


0 2


Hookes law


5.15 Initial Boundary Value Problem (IBVP) and
Computational Mechanics
An Initial Boundary Value Problem (IBVP) is defined by the governing equations (a set of
partial differential equations-PDEs) and by boundary and initial conditions. Said conditions are
restrictions on the governing equations. The IBVP solution is the one that is given by the
solution of the equations and which also satisfies the boundary and initial conditions. The
IBVP solution will be unique if the problem is well posed, i.e. given a boundary and initial
conditions, there is only one solution to the problem. Then, the governing equations are
defined by the fundamental and constitutive equations. In subsequent chapters of this
publication we will fundamentally deal with the constitutive models (constitutive
equations), which are used to complete the IBVP.
5 THE FUNDAMENTAL EQUATIONS OF CONTINUUM MECHANICS

339
It must be stressed that until the fundamental equations were established, we did not
discuss the type of material that the continuum is made up of. When we begin to specify
this is the time when the concept of the constitutive equation appears. Then, we can
automatically believe that every class of material has its particular constitutive equations.
In order to represent the real behavior of the material, the constitutive equation has to be
calibrated with macroscopic parameters, i.e. these parameters, which are obtained in the
laboratory, represent the material behavior at the macroscopic level. Remember that the scale
of study of continuum mechanics is macroscopic, and then we need to obtain some
representative macroscopic parameters of the phenomena that occur at the microscopic
scale. We can consider this to be the Achilles heel of Continuum Mechanics, because in
some cases we are not able to obtain a macroscopic parameter that characterize
phenomena that are taking place at the microscopic level. In fact, the evolution of the
constitutive equation is directly linked to the precision of the instrumentation and new
techniques used in laboratory testing of such materials. So in summary we can state that the
constitutive equations used to characterize a material must be capable of simulating any
phenomena that arise in the material during the loading/unloading/loading process, or at
least the most significant.
As discussed earlier, the constitutive equations complete the set of equations that govern a
particular physical problem. That is, they complement the IBVP. The solution of the
problem can be analytical (the exact solution) or numerical (an approximate solution). In
most cases it is impossible to obtain the analytical solution, so we turn to computational
mechanics to obtain the numerical solution of the physical problem.
Computational Mechanics resolve specific problems by using numerical simulation tools
incorporated in the computer. In general we can state that computational mechanics is not
an independent block, i.e., for its complete implementation it is directly dependent on three
areas: Theoretical Analysis (IBVP establishment), Experimental Analysis (Lab), and Numerical
Analysis (a numerical methodology incorporated into the computer to obtain the numerical
solution for IBVP), (see Figure 5.18). From a very general point of view, we can also
appreciate in Figure 5.18 how the constitutive models are embedded within the field of
Computational Mechanics.
NOTES ON CONTINUUM MECHANICS

340





































Figure 5.18: Role of the constitutive model in Computational Mechanics.
Initial Boundary Value Problem

STRUCTURE
LABORATORY
Proposition for a
CONSTITUTIVE
MODEL

IBVP
NUMERICAL SOLUTION
Input data

YES
NO
COMPUTATIONAL MECHANICS
NO
New testing
proposal

Testing proposal
Does the proposed model
accurately simulate lab testing?
Option 2
N
u
m
e
r
i
c
a
l

S
i
m
u
l
a
t
i
o
n

Option 3:
New numerical
method
If possible
Option 4:
New IBVP proposal.
The Continuum Theory
is not suitable.
Is the simulation
realistic?
Option 1
6 Introduction to Constitutive Equations











6.1 Introduction
Mathematically, the purpose of constitutive equations is to establish connections between
kinematic, thermal and mechanical variables. As an example, Figure 6.1 shows the stress-
strain relationship for a mechanical problem which represents the constitutive equation for
stress. In this case, constitutive equations should be understood as being bijective
relationships between stress and strain. Physically speaking, constitutive equations
represent different ways of idealizing the response of a material.












Figure 6.1: Stress-strain relationship (constitutive equation for stress).
6
Introduction to
Constitutive Equations
force/moment
stress
strain
displacement
Constitutive law

o


r

341 , Notes on Continuum Mechanics, Lecture Notes on Numerical
Methods in Engineering and Sciences 4, DOI 10.1007/978-94-007-5986-2_8,
International Center for Numerical Methods in Engineering (CIMNE), 2013
E.W.V. Chaves
NOTES ON CONTINUUM MECHANICS

342

Next, we will summarize the fundamental equations of continuum mechanics obtained in
Chapter 5:
The Fundamental Equations of Continuum Mechanics
(Current configuration)

The Mass Continuity Equation
(The principle of conservation of mass)
0 ) ( - v
x
,
,
V j
j
Dt
D
(6.1)
The Equations of Motion
(The principle of conservation of linear
momentum)
v
x
`
,
,
,
j j - b V (6.2)
Cauchy Stress Tensor symmetry
(The principle of conservation of angular
momentum)
T
(6.3)
The Energy Equation
(The principle of conservation of energy)
r u j j - - q D
,
`
,
x
V : (6.4)
The Entropy Inequality
(The principle of irreversibility)
0
1

1 1
) , (
2
- - - T
T
u
T T
t
x
x
,
,
`
,
` V q D j j j : (6.5)

The Fundamental Equations of Continuum Mechanics
(Reference Configuration)

The Mass Continuity Equation 0 ) ( j J
Dt
D
(6.6)
The Equations of Motion
( ) V F
V
X
X
`
,
,
`
,
,
,
,
0 0 0
0 0 0
j j
j j
-
-

b S
b P
V
V
(6.7)
Second Piola-Kirchhoff Stress Tensor
symmetry
T
S S or
T T
P P F F
(6.8)
The Energy Equation
) , ( ) , (
0 0 0
t r t u X E X
X
,
,
`
,
`
,
j j - - q S V :
or ) , ( ) , (
0 0 0
t r t u X F X
X
,
,
`
,
`
,
j j - - q P V :
(6.9)
The Entropy Inequality
0
1 1 1
0
2
0 0
- - - T
T
u
T T
X
E
,
,
`
`
` V q S j j j :
or 0
1 1 1
0
2
0 0
- - - T
T
u
T T
X
F
,
,
`
`
` V q P j j j :
(6.10)
As we saw in Chapter 5, the mass continuity equation, the equations of motion and the
energy equation give us in total 5 equations. The unknowns are: three components of
velocity v
,
, temperature T , mass density j , six components of the Cauchy stress tensor
T
, the specific internal energy u , three components of the heat flux vector q
,
, and
the entropy j , making a total of 16 unknowns.
For the problem to be well-posed eleven equations must be added which are ones that
connect stress (the constitutive equation for stress), heat (the constitutive equation for heat
conduction), energy, and entropy with other fields, i.e.:
6 INTRODUCTION TO CONSTITUTIVE EQUATIONS

343
The Constitutive Equations
The constitutive equation for stress
The relationship between stress and state
variables
(6.11)
The constitutive equation for heat conduction
The relationship between heat flux and
state variables
(6.12)
The equations of state
These relate energy and/or entropy with
state variables
(6.13)
where the constitutive equations for stress provide six equations, the heat conduction law
provides three equations, and thermodynamic state laws provide two equations, which
results in a well-posed system, with 16 unknowns and 16 equations.
The equations that relate state functions to state variables are called the equations of state or
constitutive equations and state variables, the selection of which depends on the problem in
hand, are those that depend only on themselves. For instance, when we are dealing with
solids, in general, we use strain and temperature as independent variables. In this scenario,
in thermodynamics, internal energy u (or Helmholtz free energy ), entropy j , heat flux
q
,
, and the Cauchy stress tensor are all considered to be state functions, which can be
determined by the state variables. Then, how a material responds is fully defined by the
fields ( , , j and q
,
). Depending on the problem it may be more appropriate to use
other thermodynamic potentials among: ) , ( j E u -specific internal energy, ) , ( T E -Helmholtz free
energy, ) , ( T S H -enthalpy or ) , ( T S G -Gibbs-free energy. For further details concerning these
potentials see Chapter 10. On a final note, the effects caused by an electromagnetic or
chemical change will not be considered here.
NOTE: It must be emphasized that as constitutive equations describe material
constitutions of systems from a macroscopic point of view, based on experimental
evidence, then, constitutive equations are, by their nature, approximations.
6.2 The Constitutive Principles
Due to the complexity presented by constitutive equation formulation, it may be helpful to
lay down certain principles (restrictions) when defining said constitutive equations,
Chadwick(1976), Gurtin(1963), Truesdell&Noll(1965). These principles include:
The Principle of Determinism;
The Principle of Local Action;
The Principle of Equipresence;
The Principle of Objectivity;
The Principle of Dissipation.



NOTES ON CONTINUUM MECHANICS

344
6.2.1 The Principle of Determinism
This principle states that the fields ( , , j , q
,
) at a material point ( X
,
) depend on the
entire history of the motion, ) , ( t X x
,
,
, and the temperature history, ) , ( t T X
,
, i.e., up until the
present time t , but they never depend on the future values of ( x
,
, T ).
In certain types of thermodynamic processes, it may be unrealistic for the current field
values ( , , j , q
,
) to depend on values that are too distant from the current ones,
hence, we have the principle of limited memory. The history of fields that are too far removed
from the current ones do not affect them. So, we must only consider the latest values of
the fields, which implies we need to define what is meant by recent time.
6.2.2 The Principle of Local Action
This principle states that the current values of the fields ( , , j , q
,
) at a material point
depend on the state of these field in the vicinity of the point. Motion information is given
locally by the deformation gradient ) , ( t X F
,
, and temperature by its gradient T V .
Then, the materials that satisfy the determinism and local action principles are called simple
thermoelastic materials.
6.2.3 The Principle of Equipresence
This principle states that: there is no reason to exclude a priori a state variable (independent
variable) of the constitutive equations. For example, if we have a simple thermoelastic
material it makes no sense if stress is only a function of the deformation gradient while the
heat flux vector is only a function of temperature.
6.2.4 The Principle of Objectivity
This principle states that the constitutive equations must be the same for any observer.
Therefore, any constitutive equation must satisfy the principle of objectivity, (see Chapter
4). That is, if an observer detects a stress state in the body B which undergoes a rigid body
motion, then, this observer must detect the same stress state in the body
*
B , (see Figure
6.2).
6.2.5 The Principle of Dissipation
This principle states that: constitutive equations must satisfy the entropy inequality for any
admissible thermodynamic process.






6 INTRODUCTION TO CONSTITUTIVE EQUATIONS

345









Figure 6.2: Rigid body motion.
6.3 Characterization of Constitutive Equations
for Simple Thermoelastic Materials
For a simple thermoelastic material, the state variables (independent variables) are the
deformation gradient ) , ( t X F
,
, temperature T , and the temperature gradient T
X
,
V . We
assume that , j ,
0
q
,
and P (reference configuration) have been established by the
history and current values of F , T and T
X
,
V . These quantities are expressed by means of
the following set of functionals:
) , , , (

) (
) , , , (

) (
) , , , ( ) (
) , , , ( ) (
) ( ) ( ) (
) ( ) ( ) (
0 0
) ( ) ( ) (
) ( ) ( ) (

t t t
t t t
t t t
t t t

T T t
T T t
T T t
T T t
X
X
X
X
F X
F X
F X
F X
,
,
,
,
,
,
, ,
,
,
V
V
V
V
P P
q q
j j

(6.14)
where
) (t
- represents the history of -, until the present time t , t s t . We can also verify
that

and j are scalar-valued functionals, q

,
is a vector-valued functional and P

is a
second-order-valued functional. Then, taking into account the principle of dissipation, the
Clausius-Duhem inequality must be satisfied for any thermodynamic process.
For a homogeneous simple thermoelastic material, the functionals described in (6.14) are
independent of X
,
:
) , , (

) (
) , , (

) (
) , , ( ) (
) , , ( ) (
) ( ) ( ) (
) ( ) ( ) (
0 0
) ( ) ( ) (
) ( ) ( ) (

t t t
t t t
t t t
t t t

T T t
T T t
T T t
T T t
X
X
X
X
F
F
F
F
,
,
,
,
, ,
V
V
V
V
P P
q q
j j

(6.15)
NOTE: The functions with hat - (functional) are distinct from those that are on the left
of the equation, i.e., - provides the current value of ) (t - taking into account the entire
history of the arguments of - .
X x
,
,
,
- ) ( ) (
*
t t Q c

*


B

*
B
observer
NOTES ON CONTINUUM MECHANICS

346
According to the principle of objectivity, the constitutive equations must be invariant under
rigid body motion. Then the constitutive equations must satisfy the following:
) , , (

) (
) , , (

) (
) , , ( ) (
) , , ( ) (
*
) ( ) (
*
) ( *
*
) ( ) (
*
) (
0
*
0
*
) ( ) (
*
) (
*
) ( ) (
*
) (

t t t
t t t
t t t
t t t

T T t
T T t
T T t
T T t
X
X
X
X
F
F
F
F
,
,
,
,
, ,
V
V
V
V
P P
q q
j j


(6.16)
where
*
- represents the tensor in a new system which undergoes an orthogonal
transformation, (see Chapter 4- The objectivity of tensors).
Then, by using the chain rule of derivative of the Helmholtz free energy (6.14) we obtain:
.
: T
T
T
T
T T
X
X
F
F
F
,
,
` ` `
V
V
V

o
o
-
o
o
-
o
o

) , , (

(6.17)
Additionally, the alternative form of the Clausius-Duhem inequality (entropy inequality) in
the reference configuration was obtained in Chapter 5 as:
[ ] 0
1
0 0
- - - T
T
T
X
F
,
,
` ` `
V q P j j : (6.18)
where is the Helmholtz free energy per unit mass, q is the entropy per unit mass, and
P is the first Piola-Kirchhoff stress tensor. Note that, although the stress power F
`
: P is
defined in the reference configuration, neither P nor F are in the reference configuration.
Then, by combining the equation in (6.17) with the entropy inequality in (6.18) we obtain:
0
1
0
1
0 0 0 0
0 0
-
o
o
-
(

-
o
o
-
(

o
o
-
-
(

-
o
o
-
o
o
-
o
o
-


T
T
T
T
T
T
T
T
T T
T
T
T
X
X
F
F
F
F
F
,
,
,
` `
,
` ` ` `
V V
V
V V
V
q P
q P
.
.
:
: :

j j

j
j

j

(6.19)
The above inequality must be satisfied for any admissible thermodynamic process.
Let us now consider the process such that 0 F
`
, and a system with uniform temperature,
thus 0
,
,
T
X
V , 0
,
,

.
T
X
V . In this particular thermodynamic process the inequality in (6.19)
becomes:
0
0

(

-
o
o
- T
T
`
j

j (6.20)
Note that the inequality in (6.20) must also be met for any thermodynamic process. Then,
if in the current process the condition in (6.20) is met, we can then apply another process
such that T T
` `
- , in which the entropy inequality is violated. Thus, the only way in which
the inequality in (6.20) is satisfied is when:
T T o
o
- -
o
o


j j

0 (6.21)
Then if we take into account (6.21), the inequality in (6.19) becomes:
6 INTRODUCTION TO CONSTITUTIVE EQUATIONS

347
0
1
0 0 0
-
o
o
-
(

o
o
- T
T
T
T
X
F
F
,
,
`
V V
V
q P
.
:

j

j (6.22)
Now let us consider a process where 0 F
`
, with which the inequality in (6.22) becomes:
0
1
0 0
-
o
o
- T
T
T
T
X
,
,
V V
V
q
.

j (6.23)
Note that the term 0
1
0

-
T
T
X
,
,
V q is always true, since the heat flux vector (
0
q
,
) is always
opposite to the temperature gradient ( T
X
,
V ). If 0
o
o
T
X
,
V

we have an inconsistency, i.e.


we can use
.
T
X
,
V in such a way that the condition in (6.23) is violated, with which we can
conclude that should not depend on the temperature gradient, i.e. ) , ( T F . Then
the entropy inequality in (6.22) becomes:
0
1
0 0
-
(

o
o
- T
T
X
F
F
,
,
`
V q P :

j (6.24)
Now let us consider a process where 0
,
,
T
X
V (a uniform temperature field), then the
inequality in (6.24) becomes:
0
0

(

o
o
- F
F
`
:

j P (6.25)
Starting from this point, we could apply another process where F F
` `
- , thus:
0
0

(

o
o
- - F
F
`
:

j P (6.26)
Then, the only way that the two equations (6.25) and (6.26) can remain valid is when:
F F o
o

o
o
-

j

j
0 0
P 0 P (6.27)
Afterwards, we can conclude that the constitutive equations for a simple thermoelastic
material are given by:
) , , (
) , (
) , (
) , (
) , (
) , (
0 0
0
T T
T
T
T
T
T
T
X
F
F
F
F
F
F
F
,
, ,
V q q
P

o
o
-
o
o

j

Constitutive equations for a simple
thermoelastic material

(6.28)
As we state before, the constitutive equations must satisfy the principle of objectivity and
any scalar, for instance energy and entropy, satisfies this principle. However, we can take
advantage of this principle in order to express these scalars in terms of other appropriate
parameters. Then by applying the principle of objectivity to the energy we have:
) , (
) , (
* * *
T
T
F
F

Q

(6.29)
NOTES ON CONTINUUM MECHANICS

348
where we have taken into account that F F Q
*
and T T
*
(see Chapter 4). Then as the
principle of objectivity must be met for any orthogonal tensor Q, we use the transpose
rotation tensor (
T
R Q ) of the polar decomposition ( U R F ) as the orthogonal tensor,
(see Figure 6.3), with which we obtain:
) , (
) , (
) , (
T
T
T
T
U
U R R
Q


F
(6.30)
That is, to satisfy the principle of objectivity, the energy must be a function of the right
stretch tensor. Then, if we take into account the equations
2
U C and 1 - E C 2 , (see
Chapter 2), we can still express the energy in terms of C or E , i.e.:
) , ( ; ) , (

T T E C
(6.31)
Then, for the entropy we have:
j j

j j

j j
) , (
) , (
) , (
) , ( ) , (
) , ( ) , (
*
* * *

o
o
-

o
o
-
o
o
-
T
T
T
T
T
T
T
T
T T
E
E
C
C F
F F

(6.32)
Likewise, the heat flux can be expressed as:
0 0 0 0
* * *
0
*
0

) , , ( ) , , ( ) , , ( q q q q q q
,
,
,
, , ,
, , ,
T T or T T T T
X X X
E C F V V V
(6.33)
To avoid excessive symbolism, we omit the symbols at the top part of the tensor.
As we saw in Chapter 3, P is related to the second Piola-Kirchhoff stress tensor S
(reference configuration) by means of the equation P S
-

1
F , then, by taking into
account (6.28) we can conclude that:
[ ]
ij
ij ji pj
ip
jq
iq
kp
pj
ik kq
jq
ik
qj rk rp rq pj rk
pq
ik
kj
rq
rp rq
kj
rp
pq
ik
kj
rq rp
pq
ik
kj
pq
pq
ik
kj
ik kj ik ij
C
T
C
T
C
T
C
T
C
T
F
C
T
F F
C
T
F
F F
C
T
F
F
F
F F
F
F
C
T
F
F
F F
C
T
F
F
C
C
T
F
F
T
F F
o
o

(
(

o
o
-
o
o

(
(

o
o
-
o
o

(
(

o
o
-
o
o

-
o
o

(
(

o
o
-
o
o
o
o

o
o
o
o

o
o
o
o

o
o

- -
-
- -
- - -
) , (
2
) , ( ) , ( ) , ( ) , (
) , ( ) , (
) , (
) , (
) (
) , (
) , ( ) , (
0
0 0
1 1
0
1
0
1
0
1
0
0
1
0
1 1
C
C C C C
C C
C
C C
C F

j

j

c j

j
c c c c

j P S

(6.34)
Likewise, it is possible to show that the following is also true:
E
E
C
C
o
o

o
o

) , ( ) , (
2
0 0
T T
j

j S (6.35)
6 INTRODUCTION TO CONSTITUTIVE EQUATIONS

349
Thus, the constitutive equations can be expressed in the reference configuration as follows:
) , , (

) , (
) , (
) , (
) , (
0 0
0

T T
T
T
T
T
T
X
E
E
E
E
E
E
,
,
,
V q q
S

o
o
-
o
o

j

Constitutive equations for a simple
thermoelastic material
(Reference configuration)
(6.36)
It is also possible to express the constitutive equations in the current configuration
(deformed). To do this, let us consider the relationship between the first Piola-Kirchhoff
stress tensor ( P ) and the Cauchy stress tensor, i.e.
T
J
F P
1
with which the constitutive
equation for stress given in (6.28) can be rewritten as follows:
T
T T T
T
T T
J J
T
F
F
F
F
F
F
F
F
F
F
F
F


o
o

o
o

o
o

o
o

) , (
) , ( ) , ( 1 1
) , (
0
0
0
0

j
j
j
j

P
P

(6.37)
It is also true that
T T
J J F F
- -
=
0
1
0
q q q q
, , , ,
, (see Chapter 2). Hence we can express
the constitutive equations in the current configuration as:
) , , (
) , , (
) , (
) , (
) , (
) , (
0
1
0
1
T T J
T T J
T
T
T
T
T
T
T
X
X
F F
F F
F
F
F
F
F
F
,
,
,
, ,
V
V
q
q q

-
-

o
o
-
o
o

j

Constitutive equations for a simple
thermoelastic material
(Current configuration)
(6.38)
Then, due to the principle of objectivity, the Helmholtz free energy can be written as a
function of C , i.e. ) , ( T C . Additionally, the constitutive equation for stress
jk
ik
ij
F
F
T
o
o
o
) , (F
j can still be rewritten as:
jk
ik
pq
pq
jk
ik
ij
F
F
C
C
T
F
F
T
o
o
o
o

o
o
o
) , (
) ), ( (
C
C F

j
(6.39)





NOTES ON CONTINUUM MECHANICS

350
















Figure 6.3: Right polar decomposition of the deformation gradient.

Then, by applying the definition of the right Cauchy-Green deformation tensor,
rq rp pq
F F C , the following is still valid:
[ ]
[ ]
jp
qp
iq
jq
pq
ip jp
qp
iq
jq
pq
ip jp
pq
iq
jq ip jp iq
pq
jk qk ri rp jk rq pk ri
pq
jk
ik
rq
rp rq
ik
rp
pq
jk
ik
rq rp
pq
jk
ik
pq
pq
ij
F
C
T
F
F
C
T
F F
C
T
F
F
C
T
F F
C
T
F
F F F F
C
T
F F F F
C
T
F
F
F
F F
F
F
C
T
F
F
F F
C
T
F
F
C
C
T
o
o

o
o
-
o
o

o
o
-
o
o

-
o
o

-
o
o

o
o
-
o
o
o
o

o
o
o
o

o
o
o
o
o
) , (
2
) , ( ) , (
) , ( ) , (
) , (
) , (
) , (
) (
) , ( ) , (
C
C C
C C
C
C
C
C C

j
c c c c

j
(6.40)
where we have considered the symmetry of C , i.e.
qp pq
C C . In tensorial notation the
constitutive equation for stress becomes:
T
T
F
C
C
F
o
o

) , (
2

j (6.41)
X
,

x
,

X
,

U
U R F
reference
configuration
current
configuration

0
B
B
B
R
Q R
T

) , ( T F

1 -
U
) , ( T U

) , (
) , (

T
T
E
C



S
E C,


J
e b,

6 INTRODUCTION TO CONSTITUTIVE EQUATIONS

351
Then if we take into account that j j J
0
in the above equation, we have:
T
T T T
J
T
J
T
J
T
J
F F
F
C
C
F F
C
C
F F
C
C
F

o
o

o
o

o
o

1
) , (
2
1 ) , (
2
1 ) , (
2
0
0
1
j
j

(6.42)
where S is the second Piola-Kirchhoff stress tensor, and ) , ( T C 1 is the strain energy per
unit reference volume.
The constitutive equation for heat conduction can also be expressed as:
[ ]
[ ]
T
T
T
T T J
T T J
T T J
U R Q U R Q q Q
Q Q q Q
q q



-
-
-

) , , (
) , , (
) , , (
0
1
0
1
* * * * *
0
1 *
X
X
X
F F
F F
,
,
,
,
,
, ,
V
V
V

(6.43)
Then by using
T
R Q , and by considering the symmetry of
T
U U , the above equation
becomes:
[ ]
[ ]
F
X
X
X
X
X





-
-
-
-
-

) , , (
) , , (
) , , (
) , , (
) , , (
0
1
0
1
0
1
0
1
0
1 *
T T J
T T J
T T J
T T J
T T J
T
T
T T T
T
,
,
,
,
,
,
,
,
,
, ,
V
V
V
V
V
U q
U R U q
U U q R
U R R U R R q R
U R Q U R Q q Q q

(6.44)
After that, so as to satisfy the principle of objectivity, the constitutive equations can be
expressed as:
F C F
C
C
F
C
C
F
C
X X


- -

o
o
-
o
o

) , , ( ) , , (
) , (
) , (
) , (
2
) , (
0
1
0
1
T T J T T J
T
T
T
T
T
T
, ,
,
, ,
V V q U q q

j

Constitutive equations for a simple
thermoelastic material
(Current configuration)
(6.45)
6.4 Characterization of the Constitutive
Equations for a Thermoviscoelastic
Material
Let us consider a material, (see Romano et al. (2006)), which has the behavioral
characteristics:
The stress state depends on the local deformations ( F ) and temperature ( T );
Phenomenon of energy dissipation (due to the internal friction) appears when one
part of the system has a relative shearing motion with respect to other part of the
NOTES ON CONTINUUM MECHANICS

352
system, (see Romano et al. (2006)). In this way, the material response depends on
the spatial velocity gradient (
1
) , (
-
= F F x v
x
`
, ,
,
l t V ).
Now we can observe that the functionals will also depend on the history of F
`
:
) , , , (

) (
) , , , (

) (
) , , , ( ) (
) , , , ( ) (
) ( ) ( ) ( ) (
) ( ) ( ) ( ) (
0 0
) ( ) ( ) ( ) (
) ( ) ( ) ( ) (

t t t t
t t t t
t t t t
t t t t

T T t
T T t
T T t
T T t
X
X
X
X
F F
F F
F F
F F
,
,
,
,
`
`
, ,
`
`
V
V
V
V
P P
q q
j j

(6.46)
Then, to obtain the constitutive equations, we use alternative proof to that made on a
simple thermoelastic material, (see Romano et al. (2006)).
Once again we can apply the Clausius-Duhem inequality:
[ ] 0
1
0 0
- - - T
T
T
X
F
,
,
` ` `
V q P j j : (6.47)
In addition we can calculate the rate of change of the energy ) , , , ( T T
X
F F
,
`
V :
.
: : T
T
T
T
X
X
F
F
F
F
,
,
` `
`
` `
V
V

o
o
-
o
o
-
o
o
-
o
o



(6.48)
Then, by combining (6.48) with (6.47) we obtain:
0
1
0
1
0 0 0 0 0
0 0 0 0 0
-
o
o
-
(

-
o
o
-
o
o
-
)
`

o
o
-
-
o
o
-
(

-
o
o
-
o
o
-
(

o
o
-


T
T
T
T
T
T
or
T
T
T
T
T
T
T
X X
X X
F
F
F
F
F
F
F
F
, ,
, ,
,
` ` `
`
`
,
` ` `
`
`
V V
V
V V
V
q P
q P
.
.
:
: :

j j

j j

j
Tr

(6.49)
where we have verified that given two tensors A and B , the following holds
) (
T
B A B A Tr : . Then, we can restructure the above equation as follows:
{ { 0 - b
T
u a (6.50)
where
{ {
T
T
T T
T T
T
b
X
X
X
F
F
F u
F
a
,
,
,
,
`
` ` `
`
V
V
V
-
)
`

-
o
o
)
`

o
o
-
(

-
o
o
-
o
o
-

0 0
0 0 0
1
, , ; ; ;
q P

j j

j
Tr
.

(6.51)
Since { a and b are independent of { u , the inequality in (6.50) holds for any arbitrary
value of { u , if and only if { { 0 a and 0 b with which we can make the conclusion
that:
0
o
o
F
`

j
0
(The energy does not depend on F
`
)
T T o
o
-
(

-
o
o
j j

j 0
0
(The constitutive equation for entropy)
(6.52)
6 INTRODUCTION TO CONSTITUTIVE EQUATIONS

353
0
,
,

o
o
T
X
V

j
0
(The energy is not a function of the temperature gradient T
X
,
V )
Then, if we take into account the considerations in (6.52), we can rewrite the Clausius-
Duhem inequality as:
0
1
0 0
-
(

o
o
- T
T
X
F
F
,
,
`
V q P :

j (6.53)
We can now break down the tensor P into static and dynamic equilibrium parts, i.e.
) ( ) ( d e
P P P - with which the above inequality becomes:
0
1
0
1
) (
0
) (
0
) (
0 0
) ( ) (
- -
(

o
o
-
-
(

o
o
- -

T
T
T
T
d e
d e
X
X
F F
F
F
F
,
,
,
` `
,
`
V
V
q P P
q P P
: :
:

j

(6.54)
Note that the above inequality must satisfy:
F
F
F F
X
o
o

) , (
) , , , (
0
) ( ) (
T
T T
e e

j 0 0 P P
,
`
,
V (6.55)
Thus, we can rewrite the Clausius-Duhem inequality as:
0
1
0
) (
- T
T
d
X
F
,
,
`
V q P : (6.56)
which must be met for any thermodynamic process. Note that
) (d
P and
0
q
,
cause the
energy dissipation in the system. In this way, we can summarize all the constitutive
equations for thermoviscoelastic materials as follows:
0
1
) , (
) , (
) , (
0
) (
0
) (
-
o
o

o
o
-

T
T
T
T
T
T
d
e
X
F
F
F
F
F
,
,
`
V q P
P
:

j

Constitutive equations for a
thermoviscoelastic material and
thermodynamic constraints
(Reference configuration)
(6.57)
Note that
) (d
P is a function of ) , , , (
) (
T T
d
X
F F
,
`
V P , and if we observe that
( ) F F F - W D l
`
, we can state that
) (d
P is a function of ) , , , , (
) (
T T
d
X
F
,
V W D P .
Then, by applying the principle of objectivity we obtain:
Q Q W Q Q Q Q D Q Q P Q W D P
Q W D P Q W D P
Q W D P Q W D P



-


) , , , , ( ) , , , , (
) , , , , ( ) , , , , (
) , , , , ( ) , , , , (
) ( ) (
* * * ) ( ) (
) ( * * * ) (
T T T T
T T T T
T T T T
T T T d T d
d T d
T d d
X X
X X
X X
F F
F F
F F
, ,
, ,
, ,
`
V V
V V
V V

(6.58)
where we have considered that
T T
Q W Q Q Q W -
`
*
, and
T
Q D Q D
*
, (see
Chapter 4).
NOTES ON CONTINUUM MECHANICS

354
The equation in (6.58) must satisfy for any orthogonal tensor, including the particular case
when 1 Q . In this situation, the term
T T
Q W Q Q Q W -
`
*
becomes
0 W W W Q W - - -
`
*
, thus:
) , , , , ( ) , , , , (
) ( ) (
T T T T
d d
X X
F F
, ,
V V 0 D P W D P (6.59)
Therefore, we can proven that
) (d
P is not a function of W, i.e. ) , , , (
) ( ) (
T T
d d
X
F
,
V D P P .
We can also express the tensor
) (d
P in the reference configuration by means of the tensor
) , , , (
) (
T T
d
X
E E
,
`
V S , since the terms E
`
and D are interrelated by F F E C D
T
` `
2
1
.
As we saw in Chapter 5 the relationship T J T
x
X
, ,
, ,
V V q q
0
holds, which can be proved
if we start from the equation
T
J
-
F q q
, ,

0
or
1
0

-

ik k i
F J q q
, ,
, i.e.:
k
k
p
pk k pi
p
ik k
i
p
p
ik k
i
i
x
T
J
x
T
J F
x
T
F J
X
x
x
T
F J
X
T
T
o
o

o
o

o
o

o
o
o
o

o
o

- -
q q q q q
1 1
0 0
c
X
,
,
V q
In Chapter 5 we obtained the following equation for stress power:

} } } } } } }

V V V V V V V
dV dV dV dV dV dV J dV
2
1

0
0 0 0 0 0
0 0 0 0 0
F F C E
` ` ` `
: : : : : : :

P P S S D D D
j
j

(6.60)
Hence we can express the constitutive equations for thermoviscoelastic materials in the
current configurations as:
0
) , (
) , (
) , (
) , (
) (
) (
-
o
o
-
o
o

T
T
J
J
T
T
T
T
T
d
T e
x
F
F
F
F
F
F
,
,
V q D

j

Constitutive equations for
thermoviscoelastic materials
(Current configuration)
(6.61)
The stress
) (d
in the current configuration can be obtained by comparison with the
equation in (6.42),
T
J
F F S
1
, thus:
T d d
T T
J
F E E F
X
) , , , (
1
) ( ) (
,
`
V S (6.62)
Thus:
F E C
F E E F
C
C
F
C
C
F
C
X
X

o
o
-
o
o

) , , , (

) , , , (
1
) , (
) , (
) , (
2
) , (
0
) ( ) (
) (
T T
T T
J
T
T
T
T
T
T d d
T e
,
,
`
`
V
V
q q
S

j

Constitutive equations for
thermoviscoelastic materials
(Current configuration)
(6.63)
6 INTRODUCTION TO CONSTITUTIVE EQUATIONS

355
or
F E E
F E E F
E
E
F
E
E
F
E
X
X

o
o
-
o
o

) , , , (

) , , , (
1
) , (
) , (
) , (
) , (
0
) ( ) (
) (

T T
T T
J
T
T
T
T
T
T d d
T e
,
,
`
`
V
V
q q
S

j

Constitutive equations for
thermoviscoelastic materials
(Current configuration)
(6.64)
6.4.1 Constitutive Equations with Internal Variables
The constitutive equations in (6.14) written in terms of functionals by means of the
histories of F , T and T
X
,
V are very general. An effective alternative to using functionals
is to apply the method known as thermodynamics with internal variables. In this method it is
postulated that the current state of an inelastic continuum can be characterized by the
current values of F , T , T
X
,
V and by a set of internal variables (
i
) whose evolution
indirectly includes the deformation history. Hence, the constitutive equations can be
defined as:
) , , , (
) , , , (
) , , , (
) , , , (
0 0
i
i
i
i
T T
T T
T T
T T

X
X
X
X
F
F
F
F
,
,
,
,
, ,
V
V
V
V
P P
q q

j j


(6.65)
where
i
, n i , , 2 , 1 , is a set of internal variables. These variables can be scalars, vectors
or higher order tensors.
In a process in which energy dissipation takes place, the theory with internal variables
together with the Clausius-Duhem inequality provides the conditions (restrictions) to the
constitutive equations.
From now on we will assume that the Helmholtz free energy is independent of the
temperature gradient, so, the Helmholtz free energy (6.65) is expressed by:
) , , (
i
T F
(6.66)
where {
n i
, ,
1
are the internal variables which must be added in order to
characterize the material behavior and whose presence requires that new equations be
included in the model. These additional equations are only dependent on the
thermodynamic state at the point in question, so they are local by nature.
Then, the rate of change of the equation in (6.66) is given by:
i
i
T
T

`
` ` `

o
o
-
o
o
-
o
o


F
F
:
(6.67)
The operator is substituting with the number of contractions of the order. That is, if
is a scalar, has no contractions, if is a vector, (scalar product), if is a
second-order tensor, : (double scalar product) and so on.
NOTES ON CONTINUUM MECHANICS

356
Then, by combining the equation in (6.67) with the entropy inequality we obtain:
0
1
0 0 0 0
-
o
o
-
(

-
o
o
-
(

o
o
- T
T
T
T
i
i
X
F
F
,
,
` `
` V q P

j j

j :
(6.68)
From the previous sections we have shown the following holds:
T o
o
-
o
o

j ;
0
F
P (6.69)
Then, the entropy inequality becomes:
0
1
0
1
0
0 0
- -
-
o
o
-

T
T
T
T
i i
i
i
X
X
,
,
,
,
`
`
V A
V
q
q

j

(6.70)
where we have introduced the thermodynamic forces, denoted by (
i
A - ):
i
i
o
o
- -

j
0
A
(6.71)
Then, to fully characterize the constitutive equations, the complementary laws associated
with the dissipative mechanism must be introduced, i.e. the equations for the variables
0
1
q
,
T
and
i
` . One way to ensure that the equations related to
0
q
,
and
i
` satisfy the
condition in (6.70) is by using a dissipative pseudo-potential (scalar-valued tensor function),
such that:
) , ( T
i
X
,
V A 1 1
(6.72)
which is a convex potential for any value of the pair (
i
A , T
X
,
V ). Then, the variables are
determined by:
T T
i
X
,
,
`
V A o
o
-
o
o
-
1 1
0
1
; q
(6.73)
Problem 6.1: Find the governing equations for a continuum solid which has the following
features: Isothermal and adiabatic processes; an infinitesimal strain regime and a linear
elastic relationship between stress and strain.
b) Once the linear elastic, stress-strain relationship has been established, find the equation
in which ) ( is a tensor-valued isotropic tensor function.
Solution:
When we have isothermal and adiabatic processes temperature and entropy play no role.
In an infinitesimal strain regime, the following is satisfied:
Strain tensors: u
,
sym
V = = e E
Stress tensors: S P = =
0
; ; j j = = = = V V V
x
X
F
, ,
1 . If we take this approach, mass density is no
longer unknown.
Then, taking into account the fundamental equations in (6.6)-(6.9), the remaining equations
for the proposed problem are:
1) The equations of motion
v
`
,
,
j j - b V
2) The energy equation
6 INTRODUCTION TO CONSTITUTIVE EQUATIONS

357
q S
`
`
,
,
`
,
`
,
: : - - u t r t u j j j ) , ( ) , (
0 0 0
X E X
X
V
or in terms of the Helmholtz free energy: [ ] j
`
- T
Dt
D
Dt
Du
:

`
` `
:
e
1 j
where
e
1 is the energy density (also known as strain energy density). Then if we bear in mind
the entropy inequality, we can observe that the proposed problem is characterized by a
process without any energy dissipation (an elastic process), i.e. all stored energy caused by
will recover when 0 .
3) In this problem, the constitutive equations in (6.36) become:
) (
) ( ) (
) (

o
o

o
o
=

e
1
j


Energy ( ) and stress are only functions of strain. Then, if we calculate the rate of change
of the Helmholtz free energy, i.e.

`
`
:
o
o

) (
) (

, and by substituting it with the equation

`
` `
:
e
1 j , we obtain:

o
o

o
o

o
o ) ( ) ( ) (
e e
1 1
j
` ` `
`
: : :
Thus, we can conclude that the energy equation is a redundant one, i.e. if the stress is
known the energy can be evaluated and vice-versa. So, we can summarize the governing
equations for the problem proposed with:
The equations of motion:
u b
` `
,
`
,
,
j j j - v V (3 equations)
The constitutive equations for stress:


o
o

) (
) (
e
1
(6 equations)
Kinematic equations:
u
,
sym
V (6 equations)
(6.74)
The unknowns of the proposed problem are: (6), u
,
(3) and (6), making a total of 15
unknowns and 15 equations, so the problem is well-posed. Then, to achieve the unique
solution of the set of partial differential equations given in (6.74) one must introduce the
initial and boundary conditions, hence defining the Initial Boundary Value Problem for the
linear elasticity problem.
NOTE: Although the energy equation is a redundant one, at the time of establishing an
analytical or numerical method for solving the problem, we will always start from energy
principles, hence the importance of studying the energy equation in a system.
In subsection 1.6.1 The Tensor Series (Chapter 1), we saw that we can approach a tensor-
valued tensor function by means of the following series:

- -
o o
o
- - -
o
o
- =
- -
o o
o
- - -
o
o
- =
) (
) (
) (
2
1
) (
) (
) (
) (
) (
! 2
1
) (
) (
! 1
1
) (
! 0
1
) (
0
0
2
0 0
0
0
0
0
2
0 0
0
0





: : :
: : :

Then, by considering the application point 0
0
and 0
0 0
) ( , and also taking
into account that the relationship - is linear, higher order terms can be discarded, thus:
NOTES ON CONTINUUM MECHANICS

358



: : :
e
e
C
o o
o

o
o

) ( ) (
) (
0
2
0
1

where


o o
o

) (
2 e
e
1
C is a symmetric fourth-order tensor which is known as the elasticity
tensor, which contains the material mechanical properties.
Note that, the energy equation has to be quadratic with which we can guarantee that the
relationship - is linear, since


o
o

) (
) (
e
1
. We can also use series expansion to
represent the strain energy density as follows:



: : : :
: : :
: : :
e
e
e
e
e e
e e
C
2
1 ) (
2
1
) (
) (
) (
2
1
) (
) (
) (
) (
! 2
1
) (
) (
! 1
1
) (
! 0
1
) (
0
2
0
0
2
0 0 0 0
0
0
2
0 0
0
0

o o
o

- -
o o
o
- - - -
- -
o o
o
- - -
o
o
-
1
1
1
1 1
1 1


where we have also considered that 0 0
0 0 0
, 0
e
1 .
To better illustrate the problem established here, let us consider a particular case (a one-
dimensional case) where the stress and strain components are given by:
r o r o
(
(
(

r
r
(
(
(

o
o E
e
ij ij 11 1111 11
0 0 0
0 0 0
0 0
;
0 0 0
0 0 0
0 0
C
In this case, the stress-strain linear relationship becomes r o E (Hookes law) and the
strain energy density is given by r r or E
e
2
1
2
1
1 , (see Figure 6.4).












Figure 6.4: Stress-strain relationship (a one-dimensional case).

NOTE: Here it should be pointed out that in the case of elastic processes the constitutive
equation ) ( is only dependent on the current value of , i.e. it is independent of the
deformation history.
b) The tensor-valued tensor function ) ( is isotropic if the following is satisfied:
) ( ) ( ) ( ) (
kl ij kl ij kl
e
kl
e
r o r o r r 1 1
Then, taking into account that the relationship - is given by
kl
e
ijkl ij
r o C ) ( (indicial
notation), we can conclude that:
e
ijkl
e
ijkl kl
e
ijkl kl
e
ijkl kl ij kl ij
C C C C r r r o r o ) ( ) (
) (r
e
1
E
r
0
0
o
r
r
) (r o
r
o
Stored energy
0
0
r

e
1
Current state
1
r r E
e
2
1
1
or
2
1
e
1
0
0

e
1
6 INTRODUCTION TO CONSTITUTIVE EQUATIONS

359
That is, the fourth-order tensor
e
C is isotropic. An isotropic symmetric fourth-order
tensor has the form ) (
jk il jl ik kl ij
e
ijkl
c c c c j c Ac - - C or I 1 1 j A 2 -
e
C , (see Chapter
1), and here the parameters A and j are known as Lam constants. Figure 6.5 shows the
stress-strain relationship for an isotropic material. Note that, for an isotropic linear elastic
material in an infinitesimal strain regime the constitutive equation for stress becomes:
1 I 1 1 j A j A 2 ) ( ) 2 ( ) ( ) ( - - Tr : :
e
C
It should be emphasized here that due to the fact that the
e
C -components are independent
of the coordinate system, the tensors and share the same principal space
(eigenvectors), (see Figure 6.5).



































Figure 6.5: Stress-strain relationship (isotropic material).



11
o

11
r

1
x

1
x
P

11
r

22
r

12
r

11
o

22
o

12
o

1
x
P
P
P
P
P
P

22
r

12
r

22
o

12
o

11
r

22
r

11
o

22
o

kl
e
ijkl ij
r o C


kl
e
ijkl ij
r o C


kl
e
ijkl ij
r o C


pq jq ip ij
a a o o
Isotropic material

e
ijkl
e
ijkl
e
ijkl
C C C


) ( ) (
kl
e
kl
e
r r 1 1

Principal space
NOTES ON CONTINUUM MECHANICS

360
6.5 Some Experimental Evidence
6.5.1 Behavior of Solids
In 1660, Robert Hooke discovered that for many materials (solids) displacement was
proportional to the applied force, hence the notion of elasticity was established, but this
was not the case in the sense of the stress-strain relationship. It was the Swiss
mathematician Jacob Bernoulli who observed that the proper way to describe any change
in length was by providing a force per unit area (stress) as a function of the elongation per
unit length (strain), (see Figure 6.1).
Now, if we consider the one-dimensional stress-strain relationship seen in a loading
/unloading process we can observe the following types of behavior (see Figure 6.6):
- Linear elastic behavior;
- Non-linear elastic behavior;
- Inelastic behavior.





















Figure 6.6: Behavior of solids (a one-dimensional case).
In elastic processes there is no energy dissipation, i.e., all energy that there is during the
loading process is stored, and after removing the entire load, that is, all the energy is
o
r
c) Inelastic behavior
Dissipated energy
Elastic energy
(recoverable)
r
o
r
o
a) Linear elastic behavior b) Non-linear elastic behavior
Stored energy
(recoverable)

Recoverable
energy
loading path
unloading path
loading path
unloading path
loading path
unloading path
6 INTRODUCTION TO CONSTITUTIVE EQUATIONS

361
completely recovered. In the linear elastic process stress-strain curve (see Figure 6.6(a)), the
paths of the loading and unloading processes are the same.
In the case of small deformations and isothermal processes, materials that behave
according to Figure 6.6(a) can be characterized by means of Linear Elasticity (Chapter 7).
A non-linear elastic process (Figure 6.6(b)) differs from a linear elastic one by the non-
linearity of the stress-strain curve. In general, materials that behave in this way tend to
show large deformations. These types of materials can be characterized by means of
Hyperelasticity (nonlinear elasticity), which is the theme in Chapter 8. It should be stressed here
that in the elastic process (whether this is linear or non-linear), the constitutive equation is
only dependent on the current value of r , i.e. it is independent of the deformation history.
In contrast to the above, inelastic behavior is characterized by involving energy dissipation,
and this energy at an atomic level can be interpreted as being the energy released for
restructuring atoms (dislocations). The tensile testing shown in Figure 6.7 is a typical
example of inelastic behavior or elastoplastic behavior to be precise. Then, materials having
these characteristics are analyzed by means of plasticity models (Chapter 9). At a macroscopic
level, elastoplastic behavior is characterized by the fact that once the stress state exceeds a
certain threshold (the yield stress) the material acquires a permanent strain,
p
r -plastic
strain, i.e. when the material is stress free it no longer returns to its initial state.



















Figure 6.7: Elastoplastic behavior.
Another type of inelastic behavior is shown in Figure 6.8 and is characterized by Damage
Models (Chapter 11), which are fundamentally used to show how the elastic modulus has
degraded, (see Figure 6.8). In this type of behavior when the load is removed, the material
E

e
r
o
r

Y
o

p
r

Y
o - yield stress

p
r - plastic strain

e
r - elastic strain

Y
o

Y
o
I II
III
I - elastic zone
II - plastification zone
III - completely unloading
II I
III
E
NOTES ON CONTINUUM MECHANICS

362
has no permanent strain, but internally the material will have undergone internal
degradation (an irreversible process). In Figure 6.8 we consider a loading/unloading/
loading process, where the steps 1-2-3 represent the loading process; the path 4 indicates
the unloading process which if applied will proceed as indicated by the path 5. In general,
brittle materials such as concrete, ice, ceramics and ice have these characteristics.








Figure 6.8: Inelastic behavior (damage model).
NOTE: Notice that materials are not strictly characterized by just one of the above
classifications. In general, there are materials which need a combination of the models
described above to be accurately represented, i.e. some materials exhibit both permanent
deformation and elastic modulus degradation. So, to characterize this material a plastic-
damage model can be employed.
6.5.1.1 Temperature Effect
When materials are subjected to a temperature change their mechanical properties change,
i.e. they are temperature dependent. There are two possible scenarios to be aware of when
analyzing the effect of temperature. The first is when the effect of temperature does not
significantly affect the material mechanical properties. In this case we can decouple the
problem, i.e. we can treat the thermal and mechanical effects independently. The second is
when temperature has a significant effect on the mechanical properties. In this case the
thermal and mechanical variables must be considered simultaneously in the constitutive
equations.
6.5.1.2 Some Mechanical Properties of Solids
A simple method to determine some mechanical properties of solids is by means of testing
the materials. There are destructive tests, which consist in using a material specimen and
testing it. Other types of testing are the non-destructive tests which use special devices
such as the ultra-sonic type with which we can obtain the material properties of the
structure without causing any damage to it. Within the class of destructive tests we can cite:
tensile testing, compressive testing, the triaxial compression test, to mention a few.
Tensile Testing
Tensile testing consists of a specimen whose ends are subjected to a tensile force as shown
in Figure 6.9. Then, if we known the dimensions of the specimen cross section, b , h , and
the tensile force F , it is possible to obtain the nominal stress (
0
A
F
nom
o ). Furthermore,
1
E
r r
o

d
E
E E
d
<
2
3
4
5
6 INTRODUCTION TO CONSTITUTIVE EQUATIONS

363
by means of a device, known as an extensometer, we can obtain the elongation of the
specimen ( A ) during the test. As we known (which is measured before the beginning of
the test) we can evaluate the engineering strain ( r ), so, it is possible to define the stress-
strain curve as shown in Figure 6.10.











Figure 6.9: A specimen under tensile forces.
Now, as we have the stress-strain curve, (see Figure 6.10), we can evaluate Youngs modulus
( E ) (also known as the elastic or tensile modulus). Then, if we bear in mind Figure 6.10 we
can introduce the tangent modulus (
tan
E ) and the secant modulus (
s
E ), where at a given
stress state it holds that r o
s
E and r o ` `
tan
E (the slope of the stress-strain curve). Note
that in the elastic zone theses modules coincide, i.e. r r o
s
E E and r r o ` ` ` E E
tan
.
If, in addition to the extensometer along to the tensile force direction we have a second
extensometer to measure the contraction of the cross section we could have obtained other
mechanical properties of the material: the Poissons ration, v . Remember that for an
isotropic linear elastic material the elasticity tensor ( C ) is a function of two independent
variables, here these parameters are represented by E and v . Later, in Chapter 7, we will
link these variables to the Lam constants (A, j).
Then, bearing in mind the curve r - o we can emphasize some important points:
The proportionality limit This point is denoted by the stress
e
o . The region between
the stress-free state and the proportionality limit defines the linear elastic behavior area
where there is no dissipation energy.
The yield point (the elastic limit) This point is denoted by the stress
Y
o . In the region
between
e
o and
Y
o the behavior of the material is assumed to be elastic, i.e. there is no
dissipation of energy, but the relationship o- r is no longer linear. For some materials the
proportionality point and the yield point coincide.
The ultimate strength point This point is denoted by the stress
u
o . In the region
between
Y
o and
u
o the material shows inelastic behavior, i.e. the internal structure of the
material has suffered irreversible changes, that is, it is characterized by energy dissipation.
The rupture strength point This point is defined by
r
o , which denotes the material
rupture.
b
h

A
F
F
extensometer
Initial cross section area
bh A
0
(undeformed area)
specimen

A
r
= o
0
A
F
S

NOTES ON CONTINUUM MECHANICS

364
In the region between the points
u
o and
r
o the phenomenon of concentration of
deformation in a body (inelastic process) occurs, this concentration is known as strain
localization.













Figure 6.10: Stress-strain curve.
Figure 6.11 shows some typical stress-strain curves, for instance those of steel, iron and
concrete.
There are materials in which the yield stress (
Y
o ) is well defined, since the stress is
maintained, while the strain increases in value, (see Figure 6.11 for steel). Such behavior is
characteristic for some types of steel. For materials where the yield point is not well
defined, we use an offset method in order to define
Y
o . For example,
Y
o can be used as the
point of intersection between the r - o curve and the slope line equal to Youngs modulus
and displaced by % 2 . 0 , (see Problem 6.2).
When defining the material constitutive equations, we can make idealizations
(simplifications) of the stress-strain curve, taking the curve that best fits the real material
behavior, (see Figure 6.12). For example, in Figure 6.12(a) we focus on perfectly plastic
behavior, in Figure 6.12(b) we idealize the elastoplastic stress-strain behavior by linear
hardening, in Figure 6.12(c) we can appreciate an elastoplastic stress-strain behavior by bi-
linear hardening, and Figure 6.12(d) shows behavior characterized by softening which is
that typical detected in brittle materials.
Depending on how the solids behave, these have traditionally been classified into two
categories, namely: Fragile and Ductile Materials. The most important characteristics of
these types of materials are listed below:
Fragile Materials: small deformation occurs; there is no previous warning of failure
(abrupt rupture), e.g. as found in concrete, ceramics, glass, ice, rocks, etc.
Ductile Materials: large deformation occurs; there is a previous warning of failure, e.g.
as found in steel, aluminum, etc.
Depending on the manufacturing process and the amount of carbon involved in its
manufacture, some steel may behave as if it were a brittle material.


nom
o

tan
E

s
E
1
1

e
o

u
o
r

Y
o
E
1

r
o
IV
III II I
I - Linear elastic zone (
tan S
E E E )
II - Non-linear elastic zone
IV III - - Inelastic zone
IV - Strain localization zone

r o
r o
` `
tan
S
E
E

6 INTRODUCTION TO CONSTITUTIVE EQUATIONS

365









Figure 6.11: The stress-strain curve for a few materials.


















Figure 6.12: Some idealizations of the stress-strain curve.
Problem 6.2: In tensile testing we evaluated the following points:
6 . 3 22 5
3 24 4
2 20 3
33 . 1 3 . 13 2
667 . 0 67 . 6 1
) 10 ( ) ( Point
3 -
r o Pa

Calculate Youngs modulus ( E ) and define the stress-strain curve limit points.
r
o
concrete
iron
steel

Y
o

) 2 (
Y
o

) 1 (
Y
o
1 1

) 2 (
p
E
1

) 1 (
p
E
1
a) Perfectly elastoplastic
model
b) Linear hardening
elastoplastic model
c) Bi-linear hardening
elastoplastic model
o o o
r r r
E
1

p
E
1
o
r

a
E
1 d) Softening behavior

Y
o
E E

Y
o

Y
o
1
E
NOTES ON CONTINUUM MECHANICS

366
Solution: First, we verify that the first three points maintain the same proportionalities:
kPa Pa E 10 000 10
10 2
20
3 ) 3 (
) 3 (
) 2 (
) 2 (
) 1 (
) 1 (

r
o

r
o

r
o

-

The stress-strain curve can be appreciated in Figure 6.13, in which we define the following
points:
e
o - the proportionality point;
Y
o - the yield point;
u
o - the ultimate strength
point; and
r
o - the rupture strength point.
















Figure 6.13: Stress-strain curve.
Brazilian Test
A direct test for tensile stress in brittle materials was put forward by Brazilian engineer
Fernando Lobo Carneiro. In this test a compression cylinder was used as shown in Figure
6.14, which had an advantage when working with fragile materials (concrete, ceramics),
since the manufacturing process of the specimen like the type described in Figure 6.9 can
affect the material properties.
Unconfined Compression Test
Conversely, the compression test is the opposite of the tensile test as the specimen is in the
shape of a solid circular cylinder, (see Figure 6.14). Here, the mechanical properties present
include the elastic modulus for compression, the yield stress and the rupture strength point.
In some materials, such as steel, the properties present in both the tensile and compression
tests are identical, whereas other materials, such as concrete, exhibit different properties
depending on which test is carried out. Figure 6.14 shows how concrete typically behave.
We can clearly see that it is a low tensile strength material.
Although tensile yield stress in metals is equal to compression, it was observed
experimentally that when these materials were subjected to cyclic loads their yield stress
changed. For example, let us assume that a metal which originally has as tensile and
compressive yield stresses the following values
Y
o and
Y
o - , respectively, (see Figure
6.15). Once it exceeds the tensile yield stress
Y
o and then is subjected to an unloading
process, the compressive yield stress changes to
*
Y
o - . This phenomenon was first detected
by Bauschinger, hence it is known as the Bauschinger effect.

0; 0
0.667; 6.67
1.33; 13.3
2; 20
3; 24
3.6; 22
0
5
10
15
20
25
30
0 0.5 1 1.5 2 2.5 3 3.5 4


e
o

u
o

r
o
Y
o
% 2 . 0
) (Pa o
) 10 (
3 -
r
E
E
6 INTRODUCTION TO CONSTITUTIVE EQUATIONS

367
















Figure 6.14: The stress-strain curve for concrete.












Figure 6.15: The Bauschinger effect.
Therefore, when establishing a constitutive model to show how a material behaves, it is
important to consider the whole process, i.e. loading/unloading/loading process.
Another interesting phenomenon observed in metals occurs when these materials are
subjected to cyclic loads of tension-compression. Although they do not reach the yield
stress in these cycles, they can reach a state of complete rupture (depending on the number
of cycles). This phenomenon is known as fatigue. It has become a subject of great interest
since the time of the First and Second World War, when ship hulls suffered unexplained
ruptures.
r
o

Y
o

Y
o -

*
Y
o -

t
E

c
E
r
o
Tension
Compression

t
Y
o

c
Y
o
Concrete:
t
Y
c
Y
o = o 10
Unconfined compressive test
Brazilian test
NOTES ON CONTINUUM MECHANICS

368
Triaxial Compression Test
The triaxial compression test, (see Figure 6.16), is used to obtain properties of cohesive
saturated (or unsaturated) soils. A triaxial test is outlined below:
1. The specimen is a cylindrical sample.
2. The specimen is enclosed within an elastic membrane (rubber), and both ends
of the specimen are supported by rigid plates.
3. The specimen is placed in a pressure chamber and confined to pressure
denoted by
3
o .
4. A device is previously fixed to the specimen in order to measure how its length
varies and which is used to calculate strain.











Figure 6.16: Cross section of the triaxial compression test apparatus.
With a fixed hydrostatic pressure during the test, the normal force increases gradually until
the soil sample fails by shear. The same test is repeated by varying the hydrostatic pressure
value. Each trial stress state at the time of failure is represented by the Mohrs circle. The
envelope curve to the circles is used to define some parameters, namely: the angle of
internal friction ( c ) and cohesion ( c ), (see Figure 6.17).
NOTE: The sign convention in soil and rock mechanics is the reverse of the one adopted
in solid mechanics, i.e. in soil mechanics compression is considered to be positive since
tensile strength in soils is very low or non-existence.
Some soils are formed by sediments that are linked by electrostatic forces between fine
particles, e.g. clay-water. Negatively charged clays are cohesive with water that has a strong
electrical polarity and cohesion has the same unit of measurement as stress.
Another important parameter for cohesive materials is dilatancy ( ), which can be
obtained from a test as indicated in Figure 6.18.




pressure
flexible membrane
drainage
soil
F
rigid plate
liquid
6 INTRODUCTION TO CONSTITUTIVE EQUATIONS

369








Figure 6.17: Mohrs circle for three specimens Triaxial test.











Figure 6.18: Dilatancy.
Pore Pressure
Soils (formed by sediments) have gaps which can be filled with liquid or gas (usually air),
(see Figure 6.19). Pore pressure acts by reducing contact between grains, (see Figure 6.19).
When this material is subjected to load, stress appears due to the presence of water and
depending on the permeability of the material this stress will cease to exist or be a function
of time. In this case we can define the effective stress
eff
o with:
p
eff
P - o o (6.75)
where o is total stress, and
p
P is the pore pressure.
6.5.2 Behavior of Fluids
Gases as well as liquids are materials made up of molecules (agglomerations of two or more
atoms). Fundamentally, we can state that solids can resist shear stress and consequently
have the ability to store mechanical energy while liquids have low or no resistance to shear
stress and have no capacity to store energy. Additionally, resistance to shear stress in fluids
is directly linked to fluid properties, namely, viscosity and fluids are classified as non-
viscous (e.g. water) or viscous (e.g. oil). In the case of viscous fluids all dissipated energy is
caused by viscosity.
Q
P dy
dx
tg
y

) 3 (
I
o
) 2 (
I I I
o

) 2 (
I
o
) 1 (
I
o
) 1 (
I I I
o
) 3 (
I I I
o
t

N
o
c
c
specimen 1
specimen 2
specimen 3
NOTES ON CONTINUUM MECHANICS

370
Viscosity is highly temperature dependent, decreasing in value as temperature increases.





















Figure 6.19: Porous material.
6.5.2.1 Viscosity
Viscosity can be kinematic or dynamic. Kinematic viscosity ( u) is not dependent on the
fluid mass density whereas dynamic viscosity is highly dependent on it. Thus, we can define
dynamic viscosity
v
j as:

`
v
j
(

s Pa
m
s
s
kgm
s
m
N
2 2 2
(6.76)
where t is shear stress, and ` is the rate of change of the shear strain. The SI unit of
dynamic viscosity is (Pascal x second) s Pa or
s
m kg
.
The most accurate way to measure dynamic viscosity is by viscometers (also called
viscosimeter), which are devices that measure the time it takes for a fluid to pass through a
very precise capillary diameter. The kinematic and dynamic viscosities are related by:
j
j
v
u
(

s
m
2
(6.77)
water
gas (air)
solid
V

f
V

air
V

water
V

solid
V
material point
6 INTRODUCTION TO CONSTITUTIVE EQUATIONS

371
where u is kinematic viscosity,
v
j is dynamic viscosity, and j is mass density.
Depending on the type of the constitutive equation for stress used, fluids can be classified
as follows:
Newtonian Fluids (Chapter 12)
A Newtonian fluid is characterized by one that shows a linear relation between the
viscous stress tensor ( t ) and the rate-of-deformation tensor ( D ). Some examples
of Newtonian fluids are: water and oil.
Non-Newtonian Fluids (Stokesian Fluids)
A Non-Newtonian fluid is characterized by one that shows a non-linear relation
between the viscous stress and the rate-of-deformation tensors. Some examples of
Non-Newtonian fluids are: blood, sauces, honey, toothpaste, heavy oil.
6.5.3 Behavior of Viscoelastic Materials
To understand viscoelastic behavior, we can carry out a simple experiment. For example,
we can take a gum (used) and stretch it in such a way that most of the gum is concentrated
at one end. Then, we place it in a vertical position so that the only force acting on it is the
gravity, (see Figure 6.20 at time
0
t ). Without any force added to the system, we will observe
that over time the gum will start to deform, (see Figure 6.20 during time
3 1
t t ). After it
has been deforming for a while, we cut its end off, i.e. we remove the force, and we will see
that part of the deformation recovers instantly, and we will also verify that over time
another part of the deformation recovers slowly.
That is, these materials have the ability to store mechanical energy as elastic solids and can
also dissipate energy due to their viscosity. Hence, when we are working with how to
approach the constitutive equation for these materials we have to take into account these
phenomena simultaneously, (see Findley et al. (1976), Christensen (1982)).











Figure 6.20: Viscoelastic behavior.

In other words, viscoelastic materials are those in which the stress-strain relationship is
time dependent. The most relevant viscoelastic phenomena are listed below:


0
t
1
t
3
t


4
t
4 5
t t >>
Instantaneous elastic
recovery
Slow recovery
NOTES ON CONTINUUM MECHANICS

372
Creep When stress is constant, strain increases over time. For example we can mention a
building column, which, when force is first applied shows an initial strain, which increases
over time with no corresponding increase in stress, (see Figure 6.21).
Relaxation When strain is constant, stress decreases over time. As an example we can
cite a prestressed cable bridge whose cable is initially subjected to an initial strain causing
an initial stress and over time this stress decreases while the strain remains constant, (see
Figure 6.22).
On a final note, creep and relaxation are reciprocal phenomena.











Figure 6.21: Creep phenomenon.








Figure 6.22: Relaxation phenomenon.
6.5.4 Rheological Models
Now we can introduce some simple devices that will help us to interpret constitutive
models and which will also help us to formulate more complex constitutive models.
Let us consider a rod (a one-dimensional case) subjected to tension where the stress state at
a material point is represented by o. If we are working with a linear elastic material the
stress-strain relationship is given by r o E (Hookes law), (see Problem 6.1). If we then
compare this with the governing law of a spring given by ku F , where k is the spring
constant and u is the displacement, we can state that the linear elastic model, r o E , can
be represented by the spring device, (see Figure 6.23).

) ( ) (
0
t t
e
o < o
) (
0
t
e
o

0
r
c) Stress over time.

0
r

0
r

0
r
a) Unloaded cable.
b) Imposed deformation.

0
t

0
t t >>
F F
a) unloaded column b) instantaneous deformation c) deformation over time

0
t
0
t t >>
6 INTRODUCTION TO CONSTITUTIVE EQUATIONS

373





















Figure 6.23: Spring device.
Let us suppose now that when a material reaches a certain stress value
Y
o permanent
deformation appears, i.e. if we remove the load, the material does not recover its initial
state. At a macroscopic level we can represent this phenomenon by means of a variable
called plastic strain (
p
r ) and the mechanical device that represents this phenomenon is the
coulombic frictional device which has a mechanical parameter equal to
Y
o , (see Figure 6.24).
We have seen before that the viscosity phenomenon is characterized by one that maintains
constant stress when the material undergoes strain that evolves over time. One mechanical
device that represents this phenomenon is the dashpot device which has a mechanical
property equal to
v
j , i.e. viscosity, (see Figure 6.25).
A rheological model characterizes the behavior of the material by means of combining
simple mechanical devices. For example, let us consider such a phenomenon that is initially
linear elastic and after the material reaches the threshold
Y
o it shows perfect plastic
behavior. The rheological model that represents this phenomenon can be made up of a
spring device connected in series with a coulombic frictional device, (see Figure 6.26). Note
that nothing happens on the coulombic frictional device if the stress is less than
Y
o . In the
elastic range [ )
Y
o , 0 all the stress is absorbed by the spring device. Then, when the material
reaches the value
Y
o the frictional device begins to deform freely and this deformation can
not be recovered.
Energy
F
E
Spring device
o o
r o,
r
o
F

r o E

material point
E
Nanoscopic level
m
9
10
-
=
Mesoscopic level
m
6
10
-
=
Macroscopic level
m
3
10
-
=
NOTES ON CONTINUUM MECHANICS

374






Figure 6.24: Coulombic frictional device.









Figure 6.25: Viscous dashpot device.







Figure 6.26: The rheological model for perfect elastoplasticity.
The type of device employed and their arrangements (series and/or parallel) depends on
the type of material and also how these materials behave during the loading/unloading/loading
process. This chapter will not go into detail on the mathematical formulation of these
models, since each representative model will be established in the relevant chapter.
NOTE: As we have seen in this brief introduction to material behavior, we can start from
simple models to develop more complex ones by combining them. Therefore, from now
on we will start by studying simple constitutive equations and then go on to more complex
ones.





Y
o
Coulombic frictional device
o o

p
r
o

Y
o
Unloading/loading

) 1 (
p
r
loading

v
j
r =
r
`
Dt
D

o

v
j
1
Dashpot device
r
o
r o `
v
j

Y
o E
o o
o
r

Y
o
1
E
7 Linear Elasticity











7.1 Introduction
Approaching the problem via the linear elasticity theory is perfectly acceptable in many
practical cases in engineering. Linear elasticity is used when the displacement gradient is
sufficiently small when compared with the unity. In this scenario we can apply the
infinitesimal strain regime (small deformation) which was discussed in Chapter 2 (see
subsection 2.14). In this approach the material strain (Green-Lagrange) and the spatial
strain tensor (Almansi) collapse into:
[ ] u u u
, , , ,
sym T
t V V V - ) ( ) (
2
1
) , ( x
( )
i j j i
i
j
j
i
ij
x x
, ,
2
1
2
1
u u
u
u
-
|
|
.
|

\
|
o
o
-
o
o
r (7.1)
where ) , ( t x
,
is a symmetric second-order tensor known as the small strain or infinitesimal
strain tensor, whose components are explicitly given by:
(
(
(
(
(
(
(
(

o
o
|
|
.
|

\
|
o
o
-
o
o
|
|
.
|

\
|
o
o
-
o
o
|
|
.
|

\
|
o
o
-
o
o
o
o
|
|
.
|

\
|
o
o
-
o
o
|
|
.
|

\
|
o
o
-
o
o
|
|
.
|

\
|
o
o
-
o
o
o
o

(
(
(

r r r
r r r
r r r
r
3
3
2
3
3
2
1
3
3
1
2
3
3
2
2
2
1
2
2
1
1
3
3
1
1
2
2
1
1
1
33 23 13
23 22 12
13 12 11
2
1
2
1
2
1
2
1
2
1
2
1
x x x x x
x x x x x
x x x x x
ij
u u u u u
u u u u u
u u u u u

(7.2)
Then, taking into account the following nomenclature used in engineering notation:
displacement: u
1
u , v
2
u , w
3
u , and the strain field:
x
r r
11
,
y
r r
22
,
z
r r
33
,
xy
r
12
2 ,
yz
r
23
2 ,
xz
r
13
2 , the equations in (7.2) can be rewritten as follows:
7
Linear Elasticity
375 , Notes on Continuum Mechanics, Lecture Notes on Numerical
Methods in Engineering and Sciences 4, DOI 10.1007/978-94-007-5986-2_9,
International Center for Numerical Methods in Engineering (CIMNE), 2013
E.W.V. Chaves
NOTES ON CONTINUUM MECHANICS

376
(
(
(
(
(
(
(

o
o
|
|
.
|

\
|
o
o
-
o
o
|
.
|

\
|
o
o
-
o
o
|
|
.
|

\
|
o
o
-
o
o
o
o
|
|
.
|

\
|
o
o
-
o
o
|
.
|

\
|
o
o
-
o
o
|
|
.
|

\
|
o
o
-
o
o
o
o

(
(
(

r
r
r
r
z
w
y
w
z
v
x
w
z
u
y
w
z
v
y
v
x
v
y
u
x
w
z
u
x
v
y
u
x
u
z yz xz
yz y xy
xz xy x
ij
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1
2
1

(7.3)
7.2 Initial Boundary Value Problem of Linear
Elasticity
Let us consider a three-dimensional body B (deformed configuration) which has a volume
V and mass density j . Let S be the boundary of B and n

be the outward unit normal to


the surface S . Then, we shall consider that the body is moving under the action of body
forces ) ( x
,
,
b and under traction forces ) (
*
x
,
,
t (prescribed value). The boundary consists of
a part
u
S in which the displacements are prescribed and a part

S where the traction


vector is prescribed (surface force), such that S S S .
u
and C
u
S S , (see Figure
7.1).









Figure 7.1: Body in motion.
7.2.1 Governing Equations
In Problem 6.1 we established the governing equations for the linear elasticity problem,
i.e.:
The equations of motion:
2
2
) , (
) , ( ) , (
t
t
t t
o
o
-
x
x x
, ,
,
,
, u
b j j V
2
2
t x
i
i
j
ij
o
o
-
o
o o
u
b j j
(7.4)
which provide three equations.
The constitutive equations for stress:
: C
kl ijkl ij
r o C
(7.5)
B

S
u
S
) (
*
x
,
,
t
n


) ( x
,
,
b j

2
x

1
x

3
x
O
x
,

dV V
S
7 LINEAR ELASTICITY

377
which provide six equations.
The kinematic equations (or strain-displacement equations):
) , ( ) , ( t t
sym
x x
, , ,
u V
|
|
.
|

\
|
o
o
-
o
o
r
i
j
j
i
ij
x x
u
u
2
1
(7.6)
which provide six equations.
For the problem in hand there are a total of 15 equations whose unknowns are: the
displacement vector ) , ( t x
, ,
u (with 3 unknowns); the strain tensor ) , ( t x
,
(with 6 unknowns)
and the stress tensor (with 6 unknowns) which results in a total of 15 unknowns, so, we
have a fully established problem when given the appropriate initial and boundary
conditions.
7.2.2 Initial and Boundary Conditions
The displacement boundary condition on the part
u
S of the surface:
) , ( ) , (
*
t t x x
, , , ,
u u ) , ( ) , (
*
t t
i i
x x
, ,
u u (7.7)
The stress boundary condition on the part

S of the surface:
) ,

, (

) , (
*
t t n t n x x
,
,
,
) , (
*
t
j k jk
x
,
t n o (7.8)
Initial condition (at 0 t ):
) ( ) , (
) , (
) 0 , (
0 0
0
0
0
x v x
x
x
, , ,
`
,
, ,
,
,
,

o
o

t
t
t
t
t
u
u
u u

i i
i i
v
t
0 0
0
) (
) ( ) 0 , (


x
x x
,
`
, ,
u
u u

(7.9)
In the particular case when we have a static or quasi-static problem, the equation of motion
becomes the equilibrium equations ( 0 b
, ,
- j V ), and the initial conditions become
redundant.
7.3 Generalized Hookes Law
The stress-strain linear elastic relationship ( - ) in its broadest form is known as the
generalized Hookes law which is given by the following equations:
Tensorial notation Indicial notation
: C

kl ijkl ij
r o C
(7.10)
where C is known as the elasticity tensor (or elastic stiffness tensor), which is of the symmetric
fourth-order type and contains the elastic constants (the material properties). In Problem
6.1 we showed that C has both minor (
ijlk jikl ijkl
C C C ), and major (
klij ijkl
C C )
symmetry, so the tensor features 21 independent components. It is said that a material is
homogeneous when its elastic properties do not vary from point to point throughout the
continuum, i.e. C is independent of the position vector. Moreover, a material is said to be
isotropic at any point when the components of C do not change if the reference system
undergoes a base change.
NOTES ON CONTINUUM MECHANICS

378
7.3.1 The Generalized Hookes Law in Voigt Notation
By referring to the Cauchy stress tensor symmetry we can use Voigt notation (see Chapter
1) to store the tensor components as follows:
{
_
Notation
g Engineerin
xz
yz
xy
z
y
x
xz
yz
xy
zz
yy
xx
Voigt
ij
(
(
(
(
(
(
(
(

t
t
t
o
o
o

(
(
(
(
(
(
(
(

o
o
o
o
o
o

(
(
(
(
(
(
(
(

o
o
o
o
o
o

(
(
(
(
(
(

o o o
o o o
o o o
o
13
23
12
33
22
11
33 23 13
23 22 12
13 12 11

(7.11)
Each Cauchy stress tensor component (
11
o ,
22
o ,
33
o ,
12
o ,
23
o ,
13
o ) can be obtained by
using the constitutive equation in (7.10). Then by expanding said equation for the
component
11
o , we find:
_ _ _
33 1133
32 1132
31 1131
3 113
23 1123
22 1122
21 1121
2 112
13 1113
12 1112
11 1111
1 111 11 11 11









r
-
r
-
r
r -
r
-
r
-
r
r -
r
-
r
-
r
r o r o
C
C
C
C
C
C
C
C
C
C
C
C C
l l l l l l kl kl

(7.12)
Then as we have the minor symmetry (
ijlk ijkl lk kl
C C r r ) the above equation
becomes
13 1113 23 1123 12 1112 33 1133 22 1122 11 1111 11
2 2 2 r - r - r - r - r - r o C C C C C C .
Likewise, we can obtain the expressions for
22
o ,
33
o ,
12
o ,
23
o and
13
o , and if we then
reorder them in matrix form we can obtain the generalized Hookes law in Voigt notation,
i.e.:
{ [ ]{
13
23
12
33
22
11
1313 1323 1312 1333 1322 1311
2313 2323 2312 2333 2322 2311
1213 1223 1212 1233 1222 1211
3313 3323 3312 3333 3322 3311
2213 2223 2212 2233 2222 2211
1113 1123 1112 1133 1122 1111
13
23
12
33
22
11
2
2
2
C
(
(
(
(
(
(
(
(

r
r
r
r
r
r
(
(
(
(
(
(
(
(

(
(
(
(
(
(
(
(

o
o
o
o
o
o

C C C C C C
C C C C C C
C C C C C C
C C C C C C
C C C C C C
C C C C C C

(7.13)
where [ ] C is the matrix with mechanical elastic properties. Then by application of major
symmetry, i.e.
klij ijkl
C C , the elasticity tensor components can be expressed in Voigt
notation as follows:
[ ]
(
(
(
(
(
(
(
(

66 56 46 36 26 16
56 55 45 35 25 15
46 45 44 34 24 14
36 35 34 33 23 13
26 25 24 23 22 12
16 15 14 13 12 11
C C C C C C
C C C C C C
C C C C C C
C C C C C C
C C C C C C
C C C C C C
C
(7.14)
7 LINEAR ELASTICITY

379
The equation in (7.13) indicates that the strain tensor components in Voigt notation are
shown in the following format:
{
_
Notation
g Engineerin
xz
yz
xy
z
y
x
xz
yz
xy
zz
yy
xx
Voigt
ij
(
(
(
(
(
(
(
(

r
r
r

(
(
(
(
(
(
(
(

r
r
r
r
r
r

(
(
(
(
(
(
(
(

r
r
r
r
r
r

(
(
(
(
(
(

r r r
r r r
r r r
r
2
2
2
2
2
2
13
23
12
33
22
11
33 23 13
23 22 12
13 12 11

(7.15)
Note that as the double scalar product : has the unit of stored energy (see Problem
6.1), then, this energy must be the same as when it is obtained either by : or when
{ {
T
, i.e.:

( ) { {
_
Voigt
Tensorial
T

2
1
2 2 2
2
1
2
1
13 13 23 23 12 12 33 33 22 22 11 11
r o - r o - r o - r o - r o - r o :
(7.16)
7.3.2 The Component Transformation Law for the
Generalized Hookes Law
If we consider the Cartesian coordinate system
1
x ,
2
x ,
3
x , the stress-strain relationship is
set by means of the generalized Hookes law:
Indicial notation Voigt notation
kl ijkl ij
r o C { [ ]{ C
(7.17)
These components are affected by any change to the coordinate system. Then, given the
new coordinate system
1
x ,
2
x ,
3
x , the generalized Hookes law therein is given by:
Indicial notation Voigt notation
kl ijkl ij
r o C { [ ]{ C
(7.18)
where
ij
o ,
kl
r and
ijkl
C show the stress, the strain, and the elasticity tensor components
in the new system
1
x ,
2
x ,
3
x respectively which are explicitly given in Voigt notation, by:
{ {
(
(
(
(
(
(
(
(

t
t
t
o
o
o

(
(
(
(
(
(
(
(

o
o
o
o
o
o

(
(
(
(
(
(
(
(

o
o
o
o
o
o

(
(
(
(
(
(
(
(

r
r
r

(
(
(
(
(
(
(
(

r
r
r
r
r
r

(
(
(
(
(
(
(
(

r
r
r
r
r
r

xz
yz
xy
z
y
x
xz
yz
xy
zz
yy
xx
xz
yz
xy
z
y
x
xz
yz
xy
zz
yy
xx
13
23
12
33
22
11
13
23
12
33
22
11
;
2
2
2
2
2
2

(7.19)
and
NOTES ON CONTINUUM MECHANICS

380
[ ]
(
(
(
(
(
(
(
(








66 56 46 36 26 16
56 55 45 35 25 15
46 45 44 34 24 14
36 35 34 33 23 13
26 25 24 23 22 12
16 15 14 13 12 11
C C C C C C
C C C C C C
C C C C C C
C C C C C C
C C C C C C
C C C C C C
C
(7.20)
We will next establish the transformation law for these tensor components in Voigt
notation.
7.3.2.1 The Matrix Transformation for Stress and Strain Components
At a given point in the continuum, the stress and the strain tensor components related to
the system
1
x ,
2
x ,
3
x , are represented by
ij
o and
ij
r , respectively. Then, the components
of these second-order tensors (
ij
o ,
ij
r ), in the new system, can be obtained as follows:
Indicial notation Matrix notation
kl jl ik ij
a a o o
T
A A
(7.21)

Indicial notation Matrix notation
kl jl ik ij
a a r r
T
A A
(7.22)
where A is the transformation matrix which is denoted by:
(
(
(


33 32 31
23 22 21
13 12 11
a a a
a a a
a a a
a
ij
A
(7.23)
In Voigt notation the transformation laws in (7.21) and (7.22), (see Chapter 1), are given
respectively by:
{ [ ]{ { [ ] {
-

1
M M
inverse
(7.24)
{ [ ]{ { [ ] {
-

1
N N
inverse
(7.25)
where [ ] M is the transformation matrix of the stress tensor components in Voigt notation,
which is given explicitly by:
[ ]
( ) ( ) ( )
( ) ( ) ( )
( ) ( ) ( )(
(
(
(
(
(
(
(

- - -
- - -
- - -

13 31 11 33 13 32 12 33 11 32 12 31 13 33 12 32 11 31
23 31 21 33 23 32 22 33 21 32 22 31 23 33 22 32 21 31
23 11 21 13 23 12 22 13 21 12 22 11 23 13 12 22 11 21
33 31 33 32 32 31
2
33
2
32
2
31
23 21 23 22 22 21
2
23
2
22
2
21
13 11 13 12 12 11
2
13
2
12
2
11
2 2 2
2 2 2
2 2 2
a a a a a a a a a a a a a a a a a a
a a a a a a a a a a a a a a a a a a
a a a a a a a a a a a a a a a a a a
a a a a a a a a a
a a a a a a a a a
a a a a a a a a a
M

(7.26)
and [ ] N is the transformation matrix of the strain tensor components in Voigt notation:
7 LINEAR ELASTICITY

381
[ ]
( ) ( ) ( )
( ) ( ) ( )
( ) ( ) ( )(
(
(
(
(
(
(
(

- - -
- - -
- - -

13 31 11 33 13 32 12 33 11 32 12 31 13 33 12 32 11 31
23 31 21 33 23 32 22 33 21 32 22 31 23 33 22 32 21 31
23 11 21 13 23 12 22 13 21 12 22 11 23 13 12 22 11 21
33 31 33 32 32 31
2
33
2
32
2
31
23 21 23 22 22 21
2
23
2
22
2
21
13 11 13 12 12 11
2
13
2
12
2
11
2 2 2
2 2 2
2 2 2
a a a a a a a a a a a a a a a a a a
a a a a a a a a a a a a a a a a a a
a a a a a a a a a a a a a a a a a a
a a a a a a a a a
a a a a a a a a a
a a a a a a a a a
N

(7.27)
Additionally, it can be shown that [ ] M and [ ] N are not orthogonal matrices, i.e.
[ ] [ ]
T
M M
-1
and [ ] [ ]
T

1
N N
-
, and that:
[ ] [ ]
1 -
N M
T
(7.28)
7.3.2.2 The Transformation Matrix of the Elasticity Tensor
Components
As we saw in Chapter 1, the transformation law of the fourth-order tensor components is
given by
pqrs ls kr jq ip ijkl
a a a a C C . Our goal now is to express the transformation law of the
elasticity tensor components in Voigt notation. To do this we will use the equation in (7.17)
as a starting point in order to obtain:
{ [ ]{
[ ]{ [ ][ ][ ] {
{
{ [ ][ ][ ] {
{ [ ][ ] [ ] {
{ [ ]{

-
-





1
1
C
M C M
N C M
N C M M
C

T


(7.29)
where [ ] C is the elasticity matrix in the new system (
1
x ,
2
x ,
3
x ). Therefore, we can define
the transformation law of the elasticity tensor components in Voigt notation as:
[ ] [ ][ ] [ ]
T
M C M C
Transformation law of the elasticity
tensor components in Voigt notation
(7.30)
7.4 The Elasticity Tensor
7.4.1 Anisotropy and Isotropy
Materials in general are anisotropic, i.e., material properties have different values for
different directions at any given point. Certain kinds of these at the microscopic and
mesoscopic scale have anisotropic properties, such as: concrete (made by mixing different
materials), but at a macroscopic level these properties can be considered as an average of
these properties at the mesoscopic scale, so, it is possible to consider material as
macroscopically isotropic, i.e., material properties are independent of the coordinates
system adopted. There are materials such as wood or man-made materials, e.g. composite
materials, which are made up of fibers that are directionally oriented and embedded in a
matrix, hence these materials exhibit a clear anisotropy even at the macroscopic level.
NOTES ON CONTINUUM MECHANICS

382
Most materials have some kind of symmetry along one or more axes, i.e. these axes can be
reversed without changing the material properties. For example, Figure 7.2(b) shows one
plane of symmetry, the plane
2 1
x x - , which implies that if we are dealing with just one of
these we can change a coordinate system from say
1
x ,
2
x ,
3
x to
1
x ,
2
x ,
3
x without
altering the elastic properties of the material. Then we can see another example of
symmetry in Figure 7.2(c) in which two planes of symmetry are displayed, namely:
2 1
x x -
and
3 2
x x - . Remember that the transformation law from
1
x -
2
x -
3
x to
1
x -
2
x -
3
x system is
given by the equation:
(
(
(

(
(
(

(
(
(

3
2
1
33 32 31
23 22 21
13 12 11
3
2
1

x
x
x
a a a
a a a
a a a
x
x
x

(7.31)
Next, we will study the different types of symmetry that appear in materials which may
include: one plane of symmetry (monoclinic symmetry), two planes of symmetry
(orthotropic symmetry), tetragonal symmetry, transversely isotropic symmetry, cubic
symmetry, and finally symmetry in all orientation (isotropy).








Figure 7.2: Symmetry planes.
7.4.2 Types of Elasticity Tensor Symmetry
7.4.2.1 Triclinic Materials
Triclinic materials are the most generic of anisotropic materials, i.e. there are no symmetry
planes. Then, the elasticity tensor features 21 independent components to be determined in
the laboratory:
[ ]
(
(
(
(
(
(
(
(

66 56 46 36 26 16
56 55 45 35 25 15
46 45 44 34 24 14
36 35 34 33 23 13
26 25 24 23 22 12
16 15 14 13 12 11
C C C C C C
C C C C C C
C C C C C C
C C C C C C
C C C C C C
C C C C C C
C
Triclinic materials
21 independent components
(7.32)
NOTE: The main drawback when dealing with high material anisotropy is the extreme
complexity that appears at the time of obtaining the constants (material properties) in the
laboratory.

2
x

3
x

1
x

2
x ,
2
x

3
x

1
x ,
1
x
a) Original
coordinate
system
b) One plane of
symmetry

3
x

2
x ,
2
x ,
2
x

3
x

1
x
c) Two planes
of symmetry

1
x ,
1
x

3
x
7 LINEAR ELASTICITY

383
7.4.2.2 Monoclinic Symmetry (One Plane of Symmetry)
Let us now consider a material that has a single plane of symmetry (plane
2 1
x x - ) as
illustrated in Figure 7.2(b). Then, the transformation law between the systems defined in
Figure 7.2(a) and Figure 7.2(b) is given by:
(
(
(

(
(
(

(
(
(

3
2
1
3
2
1

1 0 0
0 1 0
0 0 1
x
x
x
x
x
x
_
A

(7.33)
with which we can obtain the transformation matrix ( [ ] M ), previously defined in (7.26),
as:
[ ]
(
(
(
(
(
(
(
(

-
-

1 0 0 0 0 0
0 1 0 0 0 0
0 0 1 0 0 0
0 0 0 1 0 0
0 0 0 0 1 0
0 0 0 0 0 1
M

(7.34)
Then, to obtain the elasticity matrix in this new system, we can carry out the following
matrix operation:
[ ] [ ][ ] [ ]
T
M C M C
(7.35)
the result of which is:
[ ]
(
(
(
(
(
(
(
(

- - - -
- - - -
- -
- -
- -
- -

66 56 46 36 26 16
56 55 45 35 25 15
46 45 44 34 24 14
36 35 34 33 23 13
26 25 24 23 22 12
16 15 14 13 12 11
C C C C C C
C C C C C C
C C C C C C
C C C C C C
C C C C C C
C C C C C C
C
(7.36)
Since in this specific transformation, the elasticity matrix must provide symmetry, i.e.
[ ] [ ] C C , we can draw the conclusion that the terms in which negative signs appear should
be zero so as to satisfy the symmetry condition. Then for materials that exhibit one plane
of symmetry, the elasticity matrix has 13 independent components, namely:
[ ]
(
(
(
(
(
(
(
(

66 56
56 55
44 34 24 14
34 33 23 13
24 23 22 12
14 13 12 11
0 0 0 0
0 0 0 0
0 0
0 0
0 0
0 0
C C
C C
C C C C
C C C C
C C C C
C C C C
C
Monoclinic Symmetry
13 independent constants
(7.37)
NOTES ON CONTINUUM MECHANICS

384
7.4.2.3 Orthotropic Symmetry (Two Planes of Symmetry)
We will start from monoclinic symmetry to define the elasticity matrix format for a material
with two planes of symmetry. Then, by means of the transformation law between the
systems defined in Figure 7.2(b) and Figure 7.2(c) we obtain:
[ ]
(
(
(
(
(
(
(
(

-
-

(
(
(

1 0 0 0 0 0
0 1 0 0 0 0
0 0 1 0 0 0
0 0 0 1 0 0
0 0 0 0 1 0
0 0 0 0 0 1

1 0 0
0 1 0
0 0 1
M A

(7.38)
where we have considered the equation in (7.26) in order to evaluate the matrix [ ] M .
Then, to obtain the elasticity tensor components in the system x , we can carry out the
following matrix operation [ ] [ ][ ][ ]
T
M C M C the result of which is:
[ ]
(
(
(
(
(
(
(
(

-
-
- - -
-
-
-

66 56
56 55
44 34 24 14
34 33 23 13
24 23 22 12
14 13 12 11
0 0 0 0
0 0 0 0
0 0
0 0
0 0
0 0
C C
C C
C C C C
C C C C
C C C C
C C C C
C
(7.39)
In this specific transformation the following must be satisfied [ ] [ ] [ ] C C C , with which
we can draw the conclusion that the elasticity matrix features 9 independent constants to
be determined:
[ ]
(
(
(
(
(
(
(
(

66
55
44
33 23 13
23 22 12
13 12 11
0 0 0 0 0
0 0 0 0 0
0 0 0 0 0
0 0 0
0 0 0
0 0 0
C
C
C
C C C
C C C
C C C
C
Orthotropic Symmetry
9 independent constants
(7.40)
NOTE: Materials such as bones present a high degree of anisotropy. Nevertheless, some
researchers consider two planes of symmetry (orthotropic symmetry) to simulate
numerically the bone behavior.

7.4.2.4 Tetragonal Symmetry
Materials with tetragonal symmetry have 5 planes of symmetry one of which is the
2 1
x x -
plane and the other 4 are indicated in Figure 7.3. Note that this type also includes
orthotropic symmetry. So, to determine the format in which the matrix [ ] C is presented we
start from the elasticity matrix given in (7.40). Next, we apply the symmetry condition
according to the transformation from
3 2 1
x x x - - to
3 3 2 1
x x x x - - , in which the
transformation matrix is given by:
7 LINEAR ELASTICITY

385
[ ]
(
(
(
(
(
(
(
(

-
-
-

(
(
(

-
r r
r r
2
1
2
1
2
1
2
1
0 0 0 0
0 0 0 0
0 0 0 0 5 . 0 5 . 0
0 0 0 1 0 0
0 0 1 0 5 . 0 5 . 0
0 0 1 0 5 . 0 5 . 0

1 0 0
0 ) cos( ) sin(
0 ) sin( ) cos(
4 4
4 4
M A

(7.41)











Figure 7.3: Tetragonal symmetry.
The elasticity matrix components in this new system can then be obtained by the
transformation law [ ] [ ][ ] [ ]
T
M C M C which becomes:
_
(


(
(
(
(
(
(
(
(
(
(
(
(
(
(
(

- -
- -
- - -
|
|
.
|

\
| - -
-
|
|
.
|

\
| - -
- -
-
|
|
.
|

\
| - -
-
|
|
.
|

\
| - -
- -
-
|
|
.
|

\
| - -
-
|
|
.
|

\
| - -
C
2 2
0 0 0 0
2 2
0 0 0 0
0 0
4
2
2 4 4
0 0
2 2 2
0 0
4 2 4
2
4
2
0 0
4 2 4
2
4
2
66 55 66 55
66 55 66 55
12 22 11 13 23 11 22 11 22
13 23
33
23 13 23 13
11 22 23 13
44
12 22 11
44
12 22 11
11 22 23 13
44
12 22 11
44
12 22 11
C C C C
C C C C
C C C C C C C C C
C C
C
C C C C
C C C C
C
C C C
C
C C C
C C C C
C
C C C
C
C C C
(7.42)
Note that the plane (
3 1
x x - ) is also a plane of symmetry, (see Figure 7.3), so, the
components of (7.42) must be equal to the components obtained from the following
coordinate transformation:

1
x

2
x
3
x

4 / r


4 / r

4 / r

4 / r

4 / r


1
x

2
x

4 / r


4 / r


4 / r

NOTES ON CONTINUUM MECHANICS

386
[ ]
(
(
(
(
(
(
(
(

-
-

(
(
(

-
1 0 0 0 0 0
0 1 0 0 0 0
0 0 1 0 0 0
0 0 0 1 0 0
0 0 0 0 1 0
0 0 0 0 0 1

1 0 0
0 1 0
0 0 1
M A

(7.43)
Then, by once again applying the transformation seen in (7.30), we can obtain the
following matrix:
_
(


(
(
(
(
(
(
(
(
(
(
(
(
(
(
(

- -
- -
- - -
|
|
.
|

\
| - -
-
|
|
.
|

\
| - -
- -
-
|
|
.
|

\
| - -
-
|
|
.
|

\
| - -
- -
-
|
|
.
|

\
| - -
-
|
|
.
|

\
| - -
C
2 2
0 0 0 0
2 2
0 0 0 0
0 0
4
2
2 4 4
0 0
2 2 2
0 0
4 2 4
2
4
2
0 0
4 2 4
2
4
2
66 55 55 66
55 66 66 55
12 22 11 23 13 11 22 11 22
23 13
33
23 13 23 13
22 11 23 13
44
12 22 11
44
12 22 11
22 11 23 13
44
12 22 11
44
12 22 11
C C C C
C C C C
C C C C C C C C C
C C
C
C C C C
C C C C
C
C C C
C
C C C
C C C C
C
C C C
C
C C C

(7.44)
and by comparing the two matrices given in (7.42) and in (7.44), we can conclude that
22 11
C C ,
66 55
C C ,
23 13
C C . Then, the elasticity matrix for tetragonal symmetry material
features 6 independent constants to be determined:
[ ]
(
(
(
(
(
(
(
(

55
55
44
33 13 13
13 11 12
13 12 11
0 0 0 0 0
0 0 0 0 0
0 0 0 0 0
0 0 0
0 0 0
0 0 0
C
C
C
C C C
C C C
C C C
C
Tetragonal symmetry
6 independent constants
(7.45)
7.4.2.5 Transversely Isotropic Symmetry (Hexagonal Symmetry)
Material with transversely isotropic symmetry already includes orthotropic symmetry, (see
Eq. (7.40)). In addition, any transformation on the plane
2 1
x x - is also a plane of
symmetry, (see Figure 7.4).
For these kinds of material
2 1
x x - and
3 2
x x - are planes of symmetry, i.e. they have
orthotropic symmetry, so, by starting from the elasticity matrix for orthotropic symmetry in
(7.40) and by some transformations on the plane
2 1
x x - we can obtain the constants.
Initially, let us consider a transformation on the plane
2 1
x x - characterized by the angle
90 c , (see Figure 7.4), with which we can obtain the following transformation matrices:
7 LINEAR ELASTICITY

387
[ ]
(
(
(
(
(
(
(
(

-
-

(
(
(

-
0 1 0 0 0 0
1 0 0 0 0 0
0 0 1 0 0 0
0 0 0 1 0 0
0 0 0 0 0 1
0 0 0 0 1 0
1 0 0
0 0 1
0 1 0
M A

(7.46)









Figure 7.4: Transversely isotropic symmetry.
Then, using the equation in (7.30) we can obtain the elasticity matrix in this new system:
[ ]
(
(
(
(
(
(
(
(


55
66
44
33 13 23
13 11 12
23 12 22
0 0 0 0 0
0 0 0 0 0
0 0 0 0 0
0 0 0
0 0 0
0 0 0
C
C
C
C C C
C C C
C C C
C
(7.47)
and by comparing (7.40) and (7.47) we can deduce that:
22 11
C C ,
13 23
C C ,
66 55
C C .
Now, if we consider a transformation characterized by the angle 45 c , the
transformation matrix [ ] M becomes:
[ ]
(
(
(
(
(
(
(
(

-
-
-

(
(
(
(
(
(

-
2
1
2
1
2
1
2
1
0 0 0 0
0 0 0 0
0 0 0 0 5 . 0 5 . 0
0 0 0 1 0 0
0 0 1 0 5 . 0 5 . 0
0 0 1 0 5 . 0 5 . 0

1 0 0
0
2
1
2
1
0
2
1
2
1
M A

(7.48)
Then, by using the equation in (7.30) we obtain [ ] C as:
[ ]
(
(
(
(
(
(
(
(

-
- - - -
- - - -

55
55
12 11
2
1
33 13 13
13 44 12 22
2
1
44 12 22
2
1
13 44 12 22
2
1
44 12 11
2
1
0 0 0 0 0
0 0 0 0 0
0 0 ) ( 0 0 0
0 0 0
0 0 0 ) ( ) (
0 0 0 ) ( ) (
C
C
C C
C C C
C C C C C C C
C C C C C C C
C (7.49)

3 3
, x x

1
x

2
x
2
x

1
x
c
Transformation matrix:

(
(
(

-
1 0 0
0 ) cos( ) sin(
0 ) sin( ) cos(
c c
c c
A
NOTES ON CONTINUUM MECHANICS

388
Afterwards, if we compare (7.49) with (7.47) we can draw the conclusion that
) (
12 11
2
1
44
C C C - , hence the matrix [ ] C features 5 independent constants. Note that, any
other transformation on the plane
2 1
x x - will not reduce the number of constants. Thus,
matrices with the elastic mechanical properties for transversely isotropic symmetry material
appear in the following format:
[ ]
(
(
(
(
(
(
(
(

55
55
12 11
2
1
33 13 13
13 11 12
13 12 11
0 0 0 0 0
0 0 0 0 0
0 0 ) ( 0 0 0
0 0 0
0 0 0
0 0 0
C
C
C C
C C C
C C C
C C C
C
Transversely Isotropic Symmetry
5 independent constants
(7.50)
7.4.2.6 Cubic Symmetry
Some metals are formed by crystals which can be classified as cubic symmetry materials.
These exhibit two planes of symmetry (orthotropic symmetry) and also have the same
properties if we make a rotation along the
3
x -axis at an angle 90 c , and along the axis
1
x with 90 , , as shown in Figure 7.5.









Figure 7.5: Cubic symmetry.
As a starting point we can use the elasticity matrix for orthotropic symmetry defined in
(7.40). The transformation matrix from the x -system to the x -system, (see Figure 7.5), is
defined as:
[ ]
(
(
(
(
(
(
(
(

-
-

(
(
(

-
0 1 0 0 0 0
1 0 0 0 0 0
0 0 1 0 0 0
0 0 0 1 0 0
0 0 0 0 0 1
0 0 0 0 1 0
1 0 0
0 0 1
0 1 0
M A

(7.51)
By substituting (7.51) in the equation in (7.30) we obtain:
Rotation along
1
x -axis
90 ,
Rotation along
3
x -axis
90 c

1
x

2
x
3
x
c

3
x

1
x

2
x

1
x
3
x

2
x
,
7 LINEAR ELASTICITY

389
[ ]
(
(
(
(
(
(
(
(

55
55
44
33 31 32
13 11 12
23 21 22
0 0 0 0 0
0 0 0 0 0
0 0 0 0 0
0 0 0
0 0 0
0 0 0
C
C
C
C C C
C C C
C C C
C
(7.52)
Then, by comparing the above matrix with (7.40) we can conclude that:
[ ]
(
(
(
(
(
(
(
(

55
55
44
33 13 13
13 11 12
13 12 11
0 0 0 0 0
0 0 0 0 0
0 0 0 0 0
0 0 0
0 0 0
0 0 0
C
C
C
C C C
C C C
C C C
C
(7.53)
Then using the equation in (7.53), we rotate the
1
x -axis at an angle 90 , , resulting in the
transformation matrix:
[ ]
(
(
(
(
(
(
(
(

-
-

(
(
(

0 0 1 0 0 0
0 1 0 0 0 0
1 0 0 0 0 0
0 0 0 0 1 0
0 0 0 1 0 0
0 0 0 0 0 1
0 1 0
1 0 0
0 0 1
M A

(7.54)
After that, if we substitute (7.54) into the transformation law (7.30), we can obtain the
elasticity matrix [ ] C :
[ ]
(
(
(
(
(
(
(
(


44
55
55
11 13 12
13 33 13
12 13 11
0 0 0 0 0
0 0 0 0 0
0 0 0 0 0
0 0 0
0 0 0
0 0 0
C
C
C
C C C
C C C
C C C
C
(7.55)
Additionally, by comparing (7.55) with (7.53) we can conclude that
11 33
C C ;
44 55
C C ;
12 13
C C , since [ ] [ ] C C . Then, the elasticity matrix is defined by three independent
constants:
[ ]
(
(
(
(
(
(
(
(

44
44
44
11 12 12
12 11 12
12 12 11
0 0 0 0 0
0 0 0 0 0
0 0 0 0 0
0 0 0
0 0 0
0 0 0
C
C
C
C C C
C C C
C C C
C
Cubic Symmetry
3 independent constants
(7.56)
NOTES ON CONTINUUM MECHANICS

390
7.4.2.7 Symmetry in All Directions (Isotropy)
Finally, if the material has the same property in all directions it is called isotropic material.
Note that if we compare (7.56) with (7.52) we can conclude that ) (
12 11
2
1
55 44
C C C C - to
fulfill symmetry in all direction. Then, the elasticity matrix features 2 elastic constants to be
determined:
[ ]
(
(
(
(
(
(
(
(

-
-
-

) ( 0 0 0 0 0
0 ) ( 0 0 0 0
0 0 ) ( 0 0 0
0 0 0
0 0 0
0 0 0
12 11
2
1
12 11
2
1
12 11
2
1
11 12 12
12 11 12
12 12 11
C C
C C
C C
C C C
C C C
C C C
e
C
Isotropic Symmetry
2 independent
constants
(7.57)
Then substituting some variables so that:
12
C A and ( )
12 11
2
1
C C - j , the elasticity matrix
can be represented as follows:
[ ]
(
(
(
(
(
(
(
(

-
-
-

j
j
j
j A A A
A j A A
A A j A
0 0 0 0 0
0 0 0 0 0
0 0 0 0 0
0 0 0 2
0 0 0 2
0 0 0 2
e
C
(7.58)
where the constants A, j, are known as the Lam constants. We then split the matrix [ ]
e
C
as follows:
[ ]
[ ]
_ _
I
(
(
(
(
(
(
(
(

-
(
(
(
(
(
(
(
(

(
(
(
(
(
(
(
(

2
1
2
1
2
1
0 0 0 1 1 1
0
0
0
1
1
1
0 0 0 0 0
0 0 0 0 0
0 0 0 0 0
0 0 0 1 0 0
0 0 0 0 1 0
0 0 0 0 0 1
2
0 0 0 0 0 0
0 0 0 0 0 0
0 0 0 0 0 0
0 0 0 1 1 1
0 0 0 1 1 1
0 0 0 1 1 1
j A
e
C
(7.59)
Note that I is the matrix in Voigt notation that represents the symmetric fourth-order
unit tensor components ( [ ]
jk il jl ik ijkl
c c c c -
2
1
I ), (see Chapter 1). Then the tensor with
the elastic properties for isotropic materials is represented in tensorial and indicial notations
as follows:
Tensorial notation Indicial notation
The elasticity tensor
for isotropic material
I 1 1 j A 2 -
e
C

[ ]
jk il jl ik kl ij
e
ijkl
c c c c j c Ac - - C
(7.60)
7 LINEAR ELASTICITY

391
NOTE: Remember that in Chapter 1 we saw that any isotropic fourth-order tensor can be
written in terms of the following tensors:
jk il jl ik kl ij
c c c c c c , , , i.e.:
jk il jl ik kl ij ijkl
a a a c c c c c c
2 1 0
- - C

(7.61)
Then, because of the
e
C symmetry we can prove that
2 1
a a .
Additionally, the inverse of the elasticity matrix in (7.58) is given by:
[ ]
(
(
(
(
(
(
(
(
(
(
(
(
(
(
(

-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-

-
j
j
j
j A
j A
j j A
A
j j A
A
j
j A
A
j j A
j A
j j A
A
j
j A
A
j j A
A
j j A
j A
j
1
0 0 0 0 0
0
1
0 0 0 0
0 0
1
0 0 0
0 0 0
) 2 3 (
1
) 2 3 ( 2
1
) 2 3 ( 2
1
0 0 0
) 2 3 ( 2
1
) 2 3 (
1
) 2 3 ( 2
1
0 0 0
) 2 3 ( 2
1
) 2 3 ( 2
1
) 2 3 (
1
1
e
C (7.62)
Then, we can split the matrix [ ]
1 -
e
C as follows:
[ ]
[ ]
_ _
1
0 0 0 1 1 1
0
0
0
1
1
1
1
2 0 0 0 0 0
0 2 0 0 0 0
0 0 2 0 0 0
0 0 0 1 0 0
0 0 0 0 1 0
0 0 0 0 0 1
2
1
0 0 0 0 0 0
0 0 0 0 0 0
0 0 0 0 0 0
0 0 0 1 1 1
0 0 0 1 1 1
0 0 0 1 1 1
) 2 3 ( 2
-
(
(
(
(
(
(
(
(

-
(
(
(
(
(
(
(
(

-
(
(
(
(
(
(
(
(

-
-

I
j j A j
A
e
C
(7.63)
If we can verify firstly, that the second matrix of (7.63) is the Voigt notation representation
of the inverse of the symmetric fourth-order unit tensor components, and secondly,
1 -

ijkl ijkl
I I holds, then the inverse of the isotropic elasticity tensor is given as follows:
Tensorial notation Indicial notation
I 1 1
j j A j
A
2
1
) 2 3 ( 2
1
-
-
-

-
e
C ( ) [ ]
jk il jl ik
kl ij
e
ijkl
c c c c
j j A j
c Ac
- -
-
-

-
4
1
) 2 3 ( 2
1
C
(7.64)
where
1 -
e
C is known as the elastic compliance tensor. Here, we will left the reader work out
whether I =
-
sym e e
I C C
1
: .
NOTES ON CONTINUUM MECHANICS

392
7.5 Isotropic Materials
7.5.1 Constitutive Equations
The generalized Hookes law (7.10) for isotropic linear elastic materials can be written using
the equation in (7.60) as follows:
( )

1 I 1 1 I 1 1

j A j A j A 2 ) ( 2 2
) (
- - -

Tr
Tr
sym
e
: : : : C
(7.65)
thus
Tensorial notation Indicial notation
1 j A 2 ) ( - Tr

ij ij kk ij
r - r o j c A 2
(7.66)
Then, the inverse of (7.66) is obtained as follows:
Tensorial notation Indicial notation
1
1
1
) (
2 2
1
) ( 2
2 ) (
Tr
Tr
Tr
j
A
j
A j
j A
-
-
-

ij kk ij ij
ij kk ij ij
ij ij kk ij
c
j
A
j
c A j
j c A
r - o r
r - o r
r - r o
2 2
1
2
2

(7.67)
We now need to evaluate the trace
kk
r and to do so we need to obtain the trace of
ij
o , i.e.:
Tensorial notation Indicial notation
) (
) 2 3 (
1
) (
) ( 2 3 ) ( ) (
2 ) (


1 1 1 1
Tr Tr
Tr Tr Tr
Tr
j A
j A
j A
-

-
- : : :

( )
kk kk
kk kk
kk kk ii ii kk ii
ij ij kk ij
o
-
r
r - o
r - r r - r o
r - r o
) 2 3 (
1
2 3
2 3 2
2
j A
j A
j A j c A
j c A

(7.68)
Then, by substituting the ) ( Tr value given in (7.68) into the equation in (7.67) we obtain:
Tensorial notation Indicial notation
1
j j A j
A
2
1
) (
) 2 3 ( 2
-
-
-
Tr
ij ij kk ij
o - o
-
-
r
j
c
j A j
A
2
1
) 2 3 ( 2

(7.69)
Furthermore, the above equation could easily have been obtained by means of the
following relationship:
1 I 1 1
j j A j
A
j j A j
A
2
1
) (
) 2 3 ( 2 2
1
) 2 3 ( 2
1
-
-
-

|
|
.
|

\
|
-
-
-

-
Tr : :
e
C
(7.70)
where we have applied the compliance elasticity tensor given in (7.64). Then, from the
definition of the Cauchy stress tensor eigenvalues and eigenvectors, i.e. n n


, we can
obtain:
7 LINEAR ELASTICITY

393
( )
( ) n

n n n
n n n n n n
n n n 1 n n 1 n n

2
) (

) (

) (

2

2

) (

2

) (

2 ) (

|
|
.
|

\
| -
-
- -
- -



j
A
A j
A j j A
j A j A
Tr
Tr
Tr Tr
Tr Tr

n n n


2
) (


|
|
.
|

\
| -

j
ATr

(7.71)
Thus, as expected, we can see that the tensors and have the same principal directions
(eigenvectors), and their eigenvalues (principal values) are connected by
j
A
2
) (

Tr -
.
If we denote by
1
) 1 (
r

,
2
) 2 (
r

,
3
) 3 (
r

and
1
) 1 (
o

,
2
) 2 (
o

,
3
) 3 (
o

, the
eigenvalues of and can also be evaluated as follows:
(
(
(

r
r
r
-
(
(
(

(
(
(

o
o
o
3
2
1
3
2
1
0 0
0 0
0 0
2
1 0 0
0 1 0
0 0 1
) (
0 0
0 0
0 0
j A Tr
(7.72)
(
(
(

o
o
o
-
(
(
(

-
-

(
(
(

-
(
(
(

o
o
o

(
(
(

r
r
r
3
2
1
3
2
1
3
2
1
0 0
0 0
0 0
2
1
1 0 0
0 1 0
0 0 1
) (
) 2 3 ( 2
1 0 0
0 1 0
0 0 1
2
) (
0 0
0 0
0 0
2
1
0 0
0 0
0 0
j j A j
A
j
A
j

Tr
Tr

(7.73)
7.5.2 Experimental Determination of Elastic Constants
7.5.2.1 Youngs Modulus and Poissons Ratio
For a homogeneous and isotropic body, the following assumptions (experimentally
observed), Sechler (1952), are valid:
The normal stresses (
z y x
o o o , , ) can not produce angular shear with respect to
the same coordinate system;
Shear stress only produces shear strain;
Pure shear stress produces deformation only in the plane where shear is applied.
Based on these assumptions, we can conclude that the strain functions are:
) ( ) ( ) (
) , , ( ) , , ( ) , , (
zx zx zx yz yz yz xy xy xy
z y x z z z y x y y z y x x x
t t t
o o o r r o o o r r o o o r r
(7.74)
We can also assume that the normal strain is a linear function of the normal stresses, i.e.:
z y x x
o - o - o r
3 2 1
c c c
(7.75)
Then, because of material isotropy, the effect of
y
o upon
x
r is the same as the effect of
z
o upon
x
r , (see Figure 7.6(a)), thus:
0
3 2
< c c (experimentally observed) (7.76)
NOTES ON CONTINUUM MECHANICS

394
with
E E
v
-
3 2 1
;
1
c c c (7.77)
where E is Youngs modulus (unit of stress) and v is Poissons ratio (dimensionless). The values
of E and v can then be obtained by laboratory experiments, (see Chapter 6). Afterwards,
all of these values will enable us to write the first three equations in (7.74) as, (see Figure 7.6):
( ) [ ] ( ) [ ] ( ) [ ]
y x z z z x y y z y x x
E E E
o - o v - o r o - o v - o r o - o v - o r
1
;
1
;
1
(7.78)
7.5.2.2 The Shear and Bulk Moduli
The relationship between the types of shear stress and shear strain are given by:
zx zx yz yz xy xy
G G G
t t t
1
;
1
;
1
(7.79)
where G is known as the shear modulus, which is related to the parameters E and v by the
equation:
) 1 ( 2 v -

E
G (7.80)
Then, to define the bulk modulus ( x ) we can start from the strain tensor trace:
( ) ( ) ( ) [ ]
[ ]
m
z y x
y x z z x y z y x z y x
E
E
E
I
o
v -

o - o - o
v -

o - o v - o - o - o v - o - o - o v - o r - r - r
) 2 1 ( 3
) 2 1 (

1

(7.81)
In infinitesimal strain theory, (see Chapter 2), the strain tensor trace is equal to the volume
ratio, i.e.:

I
dV
V
z y x v
r - r - r
A
r
0

(7.82)
Then, if we compare the equation in (7.81) with (7.82) we can draw the conclusion that:
v m m v
E
p
E
r
v -
- o o
v -
r
) 2 1 ( 3
) 2 1 ( 3

(7.83)
where
E
) 2 1 ( 3 v -
is the compressibility factor. We can then define the bulk modulus (or
compressibility modulus) ( x ), (see Figure 7.7) as the inverse of the compressibility factor:
) 2 1 ( 3 v -
x
E

(7.84)
If we then look at said compressibility factor, (see equation (7.83)), when working with
incompressible material the following is satisfied: 0
) 2 1 ( 3

v -
E
, which is equivalent to
5 . 0 v . Then, material in which Poissons ratio equals 5 . 0 v is considered to be
incompressible.
7 LINEAR ELASTICITY

395





































Figure 7.6: The normal strains at a material point.

x
o
x
o
z
x
y

E
x
vo -


E
x
vo -


E
x
o

vo
- r
vo
- r
o
r
E
E
E
x
z
x
y
x
x


y
o

y
o
z
x
y

E
y
o


E
y
vo -


E
y
vo -

vo
- r
o
r
vo
- r
E
E
E
y
z
y
y
y
x

+
+

z
o

z
o
z
x
y

o
r
vo
- r
vo
- r
E
E
E
z
z
z
y
z
x


E
z
vo -


E
z
o


E
z
vo -

=

( ) [ ]
( ) [ ]
( ) [ ]

o - o v - o r
o - o v - o r
o - o v - o r
y x z z
z x y y
z y x x
E
E
E

1

1

1

a)
c)
b)
NOTES ON CONTINUUM MECHANICS

396
















Figure 7.7: Some material mechanical properties.
We can now regroup the equations in (7.78) and (7.79) into matrix form as follows:
(
(
(
(
(
(
(
(

t
t
t
o
o
o
(
(
(
(
(
(
(
(
(
(
(
(
(

v - v -
v - v -
v - v -

(
(
(
(
(
(
(
(

r
r
r
zx
yz
xy
z
y
x
zx
yz
xy
z
y
x
G
G
G
E E E
E E E
E E E

1
0 0 0 0 0
0
1
0 0 0 0
0 0
1
0 0 0
0 0 0
1
0 0 0
1
0 0 0
1
(7.85)
where
) 1 ( 2 v -

E
G . Then, the inverse of (7.85) is given by:
(
(
(
(
(
(
(
(

r
r
r
(
(
(
(
(
(
(
(
(
(

v -
v -
v -
v - v v
v v - v
v v v -
v - v -

(
(
(
(
(
(
(
(

t
t
t
o
o
o
zx
yz
xy
z
y
x
zx
yz
xy
z
y
x
E

2
2 1
0 0 0 0 0
0
2
2 1
0 0 0 0
0 0
2
2 1
0 0 0
0 0 0 1
0 0 0 1
0 0 0 1
) 2 1 )( 1 (
(7.86)
Now, the above relationship can be rewritten in tensorial or indicial notations as:
p -
1 1 1

x
o
x
y

x
o

xy
t
x
y
p
x
y
p
p

xy
t

x
o

x
r
xy


xy
t

v
r
E
G j x

v
r -volumetric
strain
E -Youngs modulus G -Shear modulus x -Bulk modulus
7 LINEAR ELASTICITY

397
Tensorial notation Indicial notation
1

) 1 ( ) 2 1 )( 1 (
) (
v -
-
v - v -
v

E E Tr
ij ij kk ij
E E
r
v -
- r
v - v -
v
o
) 1 ( ) 2 1 )( 1 (

c
(7.87)
If we then compare the equations in (7.87) and (7.66) we can conclude that the Lam
constants ( A, j) are connected to the parameters ( v , E ) via the following relationships:
) 1 ( 2
;
) 2 1 )( 1 (

v -

v - v -
v

E
G
E
j A
(7.88)
Additionally, the inverse of (7.87) is:
Tensorial notation Indicial notation
1
E E
v -
-
v -

1
) ( Tr
ij ij kk ij
E E
o
v -
- o
v -
r
1
c
(7.89)
Then if we compare the equations in (7.89) with those in (7.69) we can draw the conclusion
that:
) ( 2
;
) 2 3 (
j A
A
j A
j A j
-
v
-
-
E
(7.90)
Table 7.1 provides us with the different equations among the following mechanical
properties: E - Youngs modulus; v - Poissons ratio; x - bulk modulus; G - shear modulus,
and the Lam constants (A, j).
Now, using the information in the table, we can show the elasticity tensor in terms of the
parameters ) ; ( v E , ) ; ( G j A , ) ; ( G x j , i.e.:
(

- - x
-
v -
-
v - v -
v

1 1 I 1 1
I 1 1
I 1 1
3
1
2
2
) 1 ( ) 2 1 )( 1 (

j
j A
e
e
e
E E
C
C
C
Elasticity tensor (7.91)
and respectively the inverse tensors:
(

- -
x
=
-
-
-
=
v -
-
v -
=
-
-
-
1 1 I 1 1
I 1 1
I 1 1
3
1
2
1
9
1
2
1
) 2 3 ( 2
) 1 (
1
1
1
j
j j A j
A
e e
e e
e e
E E
D C
D C
D C
Elastic compliance tensor (7.92)






NOTES ON CONTINUUM MECHANICS

398
Table 7.1: Equations among material mechanical properties.

= j G
E x A v
) ; ( E G f
G
E
E G
GE
3 9 -

( )
E G
G E G
-
-
3
2

G
G E
2
2 -

) ; ( x G f
G
G
G
- x
x
3
9

x
3
2G
- x
( ) G
G
- x
- x
3 2
2 3

) ; ( A G f G
( )
G
G G
-
-
A
A 2 3
G
3
2
- A A
( ) G - A
A
2

) ; ( v G f
G ( ) v - 1 2G
( )
( ) v -
v -
2 1 3
1 2G

v - 2 1
2G
v
) ; ( x E f
E
E
- x
x
9
3

E x
( )
E
E
- x
- x x
9
3 9

x
- x
6
3 E

) ; ( v E f
( ) v - 1 2
E

E
( ) v - 2 1 3
E

( )( ) v - v -
v
2 1 1
E
v
) ; ( A x f
( )
2
3 A - x

( )
A
A
- x
- x x
3
9

x A
A
A
- x 3

) ; ( v x f
( )
( ) v -
v - x
1 2
2 1 3
( ) v - x 2 1 3
x
v -
xv
1
3
v
7.5.3 Restrictions on Elastic Mechanical Properties
One significant tensor in elasticity is the elastic acoustic tensor ( ) (N
e
Q ) defined as:
N N N

)

(
e e
C Q
(7.93)
which is used to find restrictions on elastic mechanical properties. The components of
(7.93) in isotropic materials are given by:
( ) ( ) [ ]
l j jl
j l jl k k l j k jk il i k jl ik i k kl ij i
k jk il jl ik kl ij i k
e
ijkl i jl
e
N N
N N N N N N N N N N N N
N N N N

) (


j A jc
j c j A c c j c c j c Ac
c c c c j c Ac
- -
- - - -
- - C Q
(7.94)
which in tensorial notation becomes ( ) N N N

)

( - - j A j1
e
Q , and whose inverse form
is given by:
( )
( )
(

-
-
-
-
N N

2
1 1
j A
j A
j
1
e
Q
( )
( )
(

-
-
-
-
l j jl
e
jl
Q N N

2
1 1
j A
j A
c
j

(7.95)
or in terms of the variables v , E as:
( )
(


v -
-
v -

-
N N

1 2
1 ) 1 ( 2 1
1
E
e
Q (7.96)
Next, the isotropic elastic acoustic tensor determinant can be evaluated as follows:
( ) j A j 2
2
-
e
Q (7.97)
In two-dimensional cases (2D) the determinant of
e
Q becomes ( ) j A j 2 -
e
Q .
7 LINEAR ELASTICITY

399
Then to obtain the eigenvalues of
e
Q one need only solve the following characteristic
determinant:
( ) [ ] ( ) ( )
( ) ( ) [ ] ( )
( ) ( ) ( ) [ ]
0



3 3 3 2 3 1
3 2 2 2 2 1
3 1 2 1 1 1

- - - - -
- - - - -
- - - - -
j j A j A j A
j A j j A j A
j A j A j j A
N N N N N N
N N N N N N
N N N N N N
(7.98)
which gives rise to the characteristic equation. Then, if we use the constraint
1
2
3
2
2
2
1
- - N N N , we can obtain the following eigenvalues:
( )(
(
(

-

j A
j
j
2 0 0
0 0
0 0
) (
ij
e
Q
(7.99)
The system for small perturbation is unstable in the presence of zero or negative roots. In
other words,
e
Q must be a positive definite tensor in order to guarantee the stability of the
system. Now, the necessary and sufficient conditions for strong ellipticity occurs when 0 > j
and 0 2 > - j A , but if said strong ellipticity conditions are violated, the material is
subjected to instability shown by a homogeneous deformation band. These conditions can
also be expressed as follows:

- < v
<

- > v
>
>
v -

1
0
1
0
0
) 1 ( 2 E
E
E
j and

> v
< v
>
v -
v -
-
1
5 . 0
0
) 2 1 (
) 1 (
2 2 j j A (7.100)
So, the material is stable if:
] [ ] [ ] [ 1 ; 0 ; 1 5 . 0 ; 1 0 - - c v < . - c v > E and E (7.101)
Now, in order to have physical meaning the bulk modulus ( x ) (Truesdell&Noll 1965)
must be positive and the stability condition point-to-point is ensured by:
0
) 2 1 ( 3
; 0
3
2
>
v -
- x >
E
j A j
(7.102)
Then for isotropic linear elastic material, strain energy is positive when it holds that:
5 . 0 1 ; 0 < v < - > E (7.103)
For most materials, Poissons ratio is between the range 5 . 0 0 < v < and materials with
negative Poissons ratio are called auxetic materials.
7.6 Strain Energy Density
As we saw in Problem 6.1, the energy equation is a redundant one for the elasticity
problem. However, the starting points for devising a strategy which can be used to obtain
the solution of the Initial Boundary Value Problem, whether analytically or numerically, are
the energy principles, hence the importance of studying the strain energy density. The
energy equation, for isotropic linear elastic material, was already obtained in Problem 6.1.
Here, we will show strain energy from an engineering standpoint.
NOTES ON CONTINUUM MECHANICS

400
In order to physically interpret the strain energy we will consider a differential volume
element dxdydz in which we have the normal stress
x
o , (see Figure 7.8).











Figure 7.8: Stored energy normal stress component.
The stored energy caused by the normal stress
x
o is equal to the triangular area of the
graph defined in Figure 7.8(b), i.e.:
dxdydz dx dydz U
x x x
x
r o r |
.
|

\
| o -

2
1
) (
2
0
0

(7.104)
Likewise, we can obtain the strain energy caused by the normal stress types
y
o and
z
o .
Now, if we consider the shear stress
xy
t , (see Figure 7.9), the strain energy is given as
follows:
Angle Moment U
0

(7.105)
dxdydz dy dxdz U
xy xy xy
xy
t
|
|
.
|

\
| t -

2
1
2
0
0
(7.106)
Next, if we consider the 6 Cauchy stress tensor components, the strain energy stored in a
differential volume element is:
( )dxdydz U
yz yz xz xz xy xy z z y y x x
t - t - t - r o - r o - r o
2
1
0
(7.107)
We can now introduce the strain energy per unit volume,
e
1 , which is known as the strain energy
density and is given by:
( )
yz yz xz xz xy xy z z y y x x
e
t - t - t - r o - r o - r o
2
1
1
(

3
m
J
(7.108)
Then, the total strain energy (U ) in the entire continuum can be evaluated by integrating
the strain energy density over the volume, i.e.:
}

V
e
dV U 1 [ ] J Nm = (7.109)


dz
dx
dy
x
y
z

x
r

x
o
Stored energy

x
o
x
o
a) b)

x
r

x
o
7 LINEAR ELASTICITY

401









Figure 7.9: Strain energy shear stress.
Note also that
e
1 is the elastic potential (the Helmholtz free energy per unit volume):
Tensorial notation Indicial notation
:
2
1

e
1
ij ij
e
r o
2
1
1
(7.110)
Then, using the generalized Hookes law :
e
C , the elastic potential becomes:
Tensorial notation Indicial notation Voigt notation
: :
e e
C
2
1
1
ijkl kl ij
e
C r r
2
1
1 { [ ]{ C
T e
2
1
1
(7.111)
Next, if we consider the equations in (7.87) and substitute them into equation (7.110), we
obtain:
( ) ( ) ( )
(

- -
v -
- r - r - r
v -
- r - r - r
v - v -
v

2 2 2 2 2 2 2
) 1 ( 2
1
1
1
) 2 1 )( 1 ( 2
zx yz xy z y x z y x
e
E
1
( ) ( ) ( ) [ ]
2 2 2 2 2 2 2
2
2
zx yz xy z y x z y x
e
- - - r - r - r - r - r - r j j
A
1
(7.112)
Additionally, if we take the derivative of the above equation with respect to the strain
x
r ,
we can obtain:
( )
x x z y x
x
e
E
o
(

r
v -
- r - r - r
v - v -
v

r o
o
1
2
2
) 2 1 )( 1 ( 2
1

(7.113)
Likewise, we can obtain:
x
x
e
o
r o
o1
;
y
y
e
o
r o
o1
;
z
z
e
o
r o
o1
;
xy
xy
e
t
o
o1
;
yz
yz
e
t
o
o1
;
zx
zx
e
t
o
o1
(7.114)
thus, we can draw the conclusion that:
Tensorial notation Indicial notation


o
o
e
1
ij
ij
e


o
o1

(7.115)
Then, the second derivative of
e
1 give us the elasticity tensor, i.e.:

xy

dx
x
y
z
dz
dy

xy
t
x
y

xy
t
dy
dx
NOTES ON CONTINUUM MECHANICS

402
e e
e
C C |
.
|

\
|
o o
o

o o
o


: :
2
1
2 2
1
(7.116)
7.6.1 Decoupling Strain Energy Density
Next, we can split the strain energy density into deviatoric and spherical parts. To do so, let
us consider the strain energy density, (see Eq. (7.110)), for an isotropic linear elastic
material:
[ ]
[ ]
[ ] [ ] [ ]
) (
2
) (
) (
2
) (
) (
2
) (

2
) (

2
) (
2 ) (
2
1
2
1
2
2 2 2
2
) (

Tr
Tr
Tr
Tr
Tr
Tr
Tr Tr
Tr
Tr
j
A
j
A
j
A
j
A
j
A
j A 1
- - -
- - -

T
e
: : : : :
_
(7.117)
where we have taken into account the constitutive equation in (7.66). The strain tensor can
be split into spherical and deviatoric parts, i.e. 1


3
) ( Tr
-
dev
, which can be substituted
in the equation in (7.117), to obtain:
[ ] [ ]
[ ] [ ]

[ ] [ ] [ ]
[ ]
dev dev
dev dev dev dev
dev
dev
dev
dev dev dev
dev dev e

1 1


:
: :
: : : :
: :

2
) (

2
) (
3
2
3
) (

2
) (
9
) (
3
) (
3
) (
2
) (
3
) (
3
) (

2
) (

2
) (
2
2 2 2
2 2
2 2
3
0 ) ( 0 ) (
j
j
j
A j j
A
j
A
j
A
j
A
1
-
x

- |
.
|

\
|
- - -
|
|
.
|
- - -

\
|
-
|
.
|

\
|
- |
.
|

\
|
- - -
x

Tr
Tr Tr Tr
Tr Tr Tr Tr
Tr Tr Tr Tr
Tr Tr
_
_ _


(7.118)
where:
[ ]
_
_
energy
al distortion purely
dev dev
energy
volumetric purely
e
: ) (
2
2
j 1 -
x
Tr
(7.119)
Then, we can conclude that the strain energy density allows for additive decomposition
into purely volumetric and distortional parts.
Next, if we consider that instead of substituting the equation of stress in (7.117) we
substitute the strain equation given in (7.69), the strain energy density becomes:

1 1

: : : :
j j A j
A
j j A j
A
1
4
1
) (
) 2 3 ( 4 2
1
) (
) 2 3 ( 2 2
1
2
1
) (
-
-
-

|
|
.
|

\
|
-
-
-

Tr
Tr Tr
e

(7.120)
The double scalar product : is a scalar and an invariant. Now, using the stress principal
space we can obtain
2
3
2
2
2
1
o - o - o o o
ij ij
: . Then, given that
3 2 1
) ( o - o - o

I Tr , we can conclude that
7 LINEAR ELASTICITY

403

o o - o o - o o - o - o - o
o - o - o
) ( 2
) (
3 2 3 1 2 1
2
3
2
2
2
1
2
3 2 1
2
_

I I
I

I I I 2
2 2
3
2
2
2
1
- o - o - o (7.121)
Now, if we also consider the equation of the second invariant of the deviatoric stress
tensor
2
J -
dev
I I

which is given by
3 3
2 2

I
I I I I
I
I I I I
dev dev
- - (see Chapter 1),
the equation in (7.121) becomes

: -
|
|
.
|

\
|
- - - o - o - o
2
2 2
2
2 2
3
2
2
2
1
2
3 3
2 J J
I I
I , with
which the strain energy density (7.120) can also be expressed as:

|
|
.
|

\
|
- -
-
-

2
2
2
2
3 4
1
) 2 3 ( 4
J

I
I
e
j j A j
A
1
_ _
energy
al distortion purely
energy
volumetric purely
2
2
2
1
) 2 3 ( 6
1
J
j j A
1 -
-


I
e

(7.122)
Problem 7.1: Given an isotropic linear elastic material whose elastic properties are
GPa E 71 , GPa G 6 . 26 , find the strain tensor components and the strain energy density
at the point in which the stress state, in Cartesian basis, is represented by:
MPa
ij
(
(
(

-
-
o
15 10 5
10 0 4
5 4 20

Solution: Poissons ratio can be obtained by means of the equation:
) 1 ( 2 v -

E
G
335 . 0 1
2
- v
G
E

( ) [ ] ( ) [ ]
6 6
9
33 22 11 11
10 211 10 15 0 335 . 0 20
10 71
1

1
-
- -

o - o v - o r
E

( ) [ ] ( ) [ ]
( ) [ ] ( ) [ ]
6 6
9
22 11 33 33
6 6
9
33 11 22 22
10 117 10 0 20 335 . 0 15
10 71
1

1
10 165 10 15 20 335 . 0 0
10 71
1

1
-
-
- -

o - o v - o r
- - -

o - o v - o r
E
E

6 6
9
23 23
6 6
9
13 13
6 6
9
12 12
10 188 ) 10 10 (
10 71
335 . 0 1 1
10 94 ) 10 5 (
10 71
335 . 0 1 1
10 75 ) 10 4 (
10 71
335 . 0 1 1
-
-
-

-
o
v -
r

-
o
v -
r
-

-
o
v -
r
E
E
E

thus:
6
10
117 188 94
188 165 75
94 75 211
-

(
(
(

- -
-
r
ij

Then, the strain energy density for an elastic material is obtained by the equation:
ij ij
e indicial e e
o r
2
1
2
1
2
1
1 1 : : : C
Next, by considering the symmetry of the tensors and , the strain energy density can
be calculated as follows:
NOTES ON CONTINUUM MECHANICS

404
[ ]
[ ]
3
13 13 23 23 12 12 33 33 22 22 11 11
/ 5 . 5637 ) 5 )( 94 ( 2 ) 10 )( 188 ( 2 ) 4 )( 75 ( 2 ) 15 )( 117 ( ) 0 )( 165 ( ) 20 )( 211 (
2
1
2 2 2
2
1
m J
e
- - - - - - - -
o r - o r - o r - o r - o r - o r 1
We can also obtain the strain energy density by using the equation in (7.122), i.e.:
2
2 2
2
1
) 2 3 ( 6
1
2
1
) 2 3 ( 6
1
J
j j A j j A
1 - -
-
- -
-

I I I I
dev
e

and if we consider that
7
10 5 . 3

I ;
14
10 4933 . 2 -

I I ; Pa
10
10 3804 . 5 = A ; G j , we
can obtain
3
/ 03 . 5638 m J
e
= 1 . Note that any discrepancies in the numerical results of
e
1
are due to numerical approximations.
Problem 7.2: Find the strain energy density in terms of the principal invariants of .
Solution:
[ ]
[ ] [ ]
[ ] [ ]
) (
2
) (
) (
2
) (
) (
2
) (

2
) (

2
) (
2 ) (
2
1
2
1
2
2 2
2 2
) (

Tr
Tr
Tr
Tr
Tr
Tr Tr
Tr
Tr
Tr
j
A
j
A
j
A
j
A
j
A
j A 1
- -
- -
- -

T
e
:
: : : :
_

We can add and subtract the term [ ]
2
) ( Tr j without altering the above outcome:
[ ]
[ ] [ ] ( )[ ] [ ] { ) ( ) ( ) ( 2
2
1
) ( ) ( ) (
2
) (
2 2 2 2 2 2
2

Tr Tr Tr Tr Tr Tr
Tr
- - - - - - j j A j j j
A
1
e

Finally, if we consider that the principal invariants of the strain tensor are ) (

Tr I ,
[ ] { ) ( ) (
2
1
2 2

Tr Tr - I I , we can obtain: ( )

I I I
e
j j A 1 2 2
2
1
2
- -

7.7 The Constitutive Law for Orthotropic
Material
For orthotropic material the stress-strain relationship is given by the following equation:
(
(
(
(
(
(
(
(

t
t
t
o
o
o
(
(
(
(
(
(
(
(
(
(
(
(
(
(
(

v - v -
v - v -
v - v -

(
(
(
(
(
(
(
(

r
r
r
zx
yz
xy
z
y
x
zx
yz
xy
z
y
x
G
G
G
E E E
E E E
E E E

1
0 0 0 0 0
0
1
0 0 0 0
0 0
1
0 0 0
0 0 0
1
0 0 0
1
0 0 0
1
13
23
12
3 2
23
1
13
3
32
2 1
12
3
31
2
21
1
(7.123)
7 LINEAR ELASTICITY

405
in which there are 12 constants:
1
E ;
2
E ;
3
E ;
12
v ;
13
v ;
23
v ;
21
v ;
31
v ;
32
v ;
12
G ;
23
G ;
13
G ,
but only 9 independents, (see equation (7.40)), since
2
23
3
32
1
13
3
31
1
12
2
21
; ;
E E E E E E
v

v v

v v

v

(7.124)
Next, the reciprocal of (7.123) provides the generalized Hookes law for orthotropic
material:
(
(
(
(
(
(
(
(

r
r
r
(
(
(
(
(
(
(
(
(
(
(

- v v v v - v - v v - v -
v v - v - - v v v v - v -
v v - v - v v - v - - v v

(
(
(
(
(
(
(
(

t
t
t
o
o
o
zx
yz
xy
z
y
x
zx
yz
xy
z
y
x
G
G
G
E E E
E E E
E E E

0 0 0 0 0
0 0 0 0 0
0 0 0 0 0
0 0 0
) 1 ( ) ( ) (
0 0 0
) ( ) 1 ( ) (
0 0 0
) ( ) ( ) 1 (
13
23
12
12 21 3 31 12 32 2 21 32 31 1
31 12 32 2 31 13 2 31 23 21 1
21 32 31 1 31 23 21 1 23 32 1



(7.125)
where: 1 2
32 13 21 12 21 13 31 23 32
- v v v - v v - v v - v v .
Note that when
3 2 1
E E E ;
23 31 21 23 13 12
v v v v v v ;
13 23 12
G G G are
satisfied, we revert to the isotropic case and thereby obtain the equations in (7.85) and
(7.86).
7.8 Transversely Isotropic Materials
We can represent the elasticity matrix for a transversely isotropic material as follows:
[ ]
0 0 0 0 0
0 0 0 0 0
0 0 0 0 0
0 0 0 2
0 0 0 2
0 0 0 2
0
0
0 0 0 0
0
0
(
(
(
(
(
(
(
(

-
-
-

j
j
j
j A A A
A j A A
A A j A
C
(7.126)
Now, by decoupling the elasticity matrix into an isotropic and anisotropic part we obtain:
[ ]
(
(
(
(
(
(
(
(

A
A
A - A A A
A
A
-
(
(
(
(
(
(
(
(

-
-
-

j
j
j A A A
A
A
j
j
j
j A A A
A j A A
A A j A
0 0 0 0 0
0 0 0 0 0
0 0 0 0 0 0
0 0 0 2
0 0 0 0 0
0 0 0 0 0

0 0 0 0 0
0 0 0 0 0
0 0 0 0 0
0 0 0 2
0 0 0 2
0 0 0 2
C
(7.127)
where we have taken into account that
0
A A A - - A and
0
j j j - - A . Moreover, the
strain energy density,
e
1 , can also be split into an isotropic and anisotropic part:
NOTES ON CONTINUUM MECHANICS

406
e
Ani
e
iso
e
1 1 1 - (7.128)
The isotropic part of the strain energy density,
e
iso
1 , is the same as that seen previously for
isotropic materials. The stress-strain relationship for the anisotropic part is considered as
follows:
(
(
(
(
(
(
(
(

r
r
r
r
r
r
(
(
(
(
(
(
(
(

A
A
A - A A A
A
A

(
(
(
(
(
(
(
(

o
o
o
o
o
o
13
23
12
33
22
11
13
23
12
33
22
11
2
2
2
0 0 0 0 0
0 0 0 0 0
0 0 0 0 0 0
0 0 0 2
0 0 0 0 0
0 0 0 0 0
j
j
j A A A
A
A

(7.129)
Then, the anisotropic part of the strain energy density is given by:
( )
13 13 23 23 12 12 33 33 22 22 11 11
2 2 2
2
1
2
1
r o - r o - r o - r o - r o - r o r o
ij ij
e
Ani
1 (7.130)
and by substituting the stresses given in (7.129) we obtain:
( ) [ ] {

13 13 23 23
33 33 22 11 22 33 11 33
4 4 0
2
2
1
r o A - r o A - -
r r A - A - r A - r A - r r A - r r A
j j
j A A A A A 1
e
Ani
(7.131)
Then, if we simplify the above equation we obtain:
[ ] ( )
2
23
2
13
2
33 33
2
2
) ( r - r A - r |
.
|

\
| A
- A - r A j
A
j A 1 Tr
e
Ani
(7.132)
7.9 The Saint-Venants and Superposition
Principles
If we consider two equivalent systems of forces as indicated in Figure 7.10, the Saint-
Venants principle states that when a point is far enough (damped area) from the point of
disturbance, the two force systems are equivalent, i.e. they have the same outcome.










Figure 7.10: Saint-Venants principle.

A
F
o
Zone with same
response for both
systems
Undisturbed zone Disturbed zone
o
A
F

F F
FORCE
SYSTEM (I)
FORCE
SYSTEM (II)
Zone with different
responses
Disturbed zone
Zone with different
responses
7 LINEAR ELASTICITY

407
The superposition principle states that the balance of a system in which several actions take
place is equal to the sum of all independent actions, (see Figure 7.11). This principle is valid
because the governing equations have been linearized. As example, we can decouple the
thermo-mechanical process, i.e. we can treat the different parts independently.









Figure 7.11: Superposition principle.
Problem 7.3: Let us consider a bar to which at one end we apply a force equal to N 6000
as shown in Figure 7.12. Find
z y x
r r r , , , and the length change of the bar. Let us consider
that the bar is made up of a material whose properties are: Youngs modulus: Pa E
7
10 ;
Poissons ratio: 3 . 0 v .















Figure 7.12

Solution: Using the normal strain expressions we can obtain:
( ) [ ]
( ) [ ] 0006 . 0
10
6000

1
00018 . 0
10
) 6000 )( 3 . 0 (

1
7
7

o
o - o v - o r
- - o
v
- o - o v - o r
E E
E E
y
z x y y
y z y x x

( ) [ ] 00018 . 0
1
- o
v
- o - o v - o r
y y x z z
E E

The total change in cross-sectional dimensions is m w u
4
10 8 . 1 1 00018 . 0
-
- - , and
the total change in length is m v
2
10 0 . 6 100 0006 . 0
-
.
m 1
N F 6000
u x,
v y,
w z,
m 1
m 100

1 1
6000

o
y

=

= +

u
,

) (
) (
) (

u
u
,
,

)) ( (
) (
)) ( (
T
T
T

u
,

= +
NOTES ON CONTINUUM MECHANICS

408
7.10 Initial Stress/Strain
Some physical phenomena can be directly added to the constitutive equations because of the
superposition principle. The effect of these phenomena can be represented by means of stress or
strain.
7.10.1 Thermal Deformation
When temperature changes, there is an increase in internal energy, so, the atoms/molecules
vibrate more intensely. This vibration causes the ligaments among the molecules to stretch,
thereby causing the body volume to increase, (see Figure 7.13).









Figure 7.13: Body under a temperature change.
NOTE: In general, material properties change with temperature, i.e. material properties are
temperature dependent. In this chapter, they are considered to be constant with regard to
temperature, since the temperature variation range under consideration is not large enough.

In decoupled thermomechanical problems, it is possible to apply the superposition
principle, i.e. we can obtain the strain field because of the mechanical problem ) (u
,
ij
r
(considering the isothermal process) and we can add the strain field due to the thermal
effect ) ( T
ij
A r , i.e.:
) ( ) ( T
ij ij ij
A r - r r u
,

(7.133)
where ) (u
,
ij
r shows the mechanical strain in terms of the displacement field ( u
,
) and
) ( T
ij
A r is the thermal strain in terms of the temperature variation ( T A ).
To obtain the temperature variation it is necessary to find the temperature distribution
within the body by means of the heat flux equation, (see Chapter 5). For isotropic
materials, the thermal strain caused by temperature variation is just represented by its
normal components:
ij ij
T T T c c ) ( ) (
0
- r

(7.134)
where
0
T is the initial temperature; T is the final temperature, and c is the coefficient of
thermal expansion. For further details about thermomechanical problems, see Chapter 10.
warming

Initial volume Final volume
7 LINEAR ELASTICITY

409
Next, we will show the coefficients of thermal expansion for some materials, namely:
C
steel

1
10 12
6 -
c ,
C
aluminum

1
10 23
6 -
c ,
C
copper

1
10 17
6 -
c .
Then, by using the equations in (7.133) and (7.89) we can obtain:
ij ij ij kk ij ij ij kk ij
T T
E E
T T c c c c c
j
c
j A j
A
) (
1
) (
2
1
) 2 3 (
0 0
- - o
v -
- o
v -
- - o - o
-
-
r
(7.135)
The Hookes law for isotropic materials including the thermal effect is given by the
reciprocal of the equation in (7.135), the result of which is:
ij ij ij kk ij
ij ij ij kk ij
T T
E E E
T T
c c c
c j A c j c A
) (
) 2 1 ( ) 1 ( ) 2 1 )( 1 (
) )( 2 3 ( 2
0
0
-
v -
- r
v -
- r
v - v -
v
o
- - - r - r o

(7.136)
Problem 7.4: Let us consider a length rod equal to m L 5 . 7 , whose diameter is equal to
m 1 . 0 , which is made up of a material whose properties are: Pa E
11
10 0 . 2 and
C
1
10 20
6 -
c . Initially the rod has a temperature equal to C 15 which later rises to
C 50 .
1) Considering that the rod can expand freely, calculate the total elongation of the rod, L A ;
2) Now assume that the rod can not expand freely because concrete blocks have been
placed at its ends, (see Figure 7.14(b)). Find the stress in the rod.
Hint: Consider the problem in one dimension.










Figure 7.14: Rod under thermal effect.

Solution: 1) To obtain the elongation, we pre-calculate the thermal strain according to the
rod axis direction
ij ij
T c c ) (A r . Since this is a one-dimensional case, we need only
consider the normal strain component according to the x -direction,
x
r r
11
, then:
4 6
11
10 7 ) 15 50 ( 10 20
- -
- r r
x

Then, the total elongation,
) 2 ( ) 1 (
L L L A - A A , is obtained by solving the integral:
m L dx L
x
L
x
3 4
0
10 25 . 5 5 . 7 10 7
- -
r r A
}

Note that as the rod can expand freely, it is stress-free.
2) If the ends can not move, there will be a homogeneous stress field equal to:
Pa E T E
x ij x
8 4 11
10 4 . 1 10 7 10 0 . 2 " " ) ( - - r - A - o
-
c c
Note that in the case 2) there is no strain, since 0 AL . Moreover, it is the same as when
the initial length is equal to L L A - in which we apply compression stress in order to
obtain a final length equal to L.

) 2 ( ) 1 (
L L L A - A A
a)
b)
L

) 1 (
L A
x
T A
T A

) 2 (
L A
L
NOTES ON CONTINUUM MECHANICS

410
7.11 The Navier-Lam Equations
By means of the constitutive law, (see the equations in (7.5) and (7.65)), we can calculate
the Cauchy stress tensor divergence as follows:
j ij ij j kk j ij ij ij kk ij , , ,
2 2 r - r o r - r o j c A j c A
(7.137)
Furthermore, by using the kinematic equations in (7.6) we can obtain the term
j ij ,
r and
therefore
j kk,
r , i.e.:
( ) ( )
kj k j kk ij j jj i j ij i j j i ij , , , , , , ,
2
2
1
u u u u u r - r - r (7.138)
Then, by combining the equations in (7.138) and (7.137) we can obtain:
( )
jj i ji j ij j jj i ij kj k j ij , , , , , ,
) ( u u u u u j j A j c A - - - - o
(7.139)
Finally, by substituting the equation (7.139) into the equations of motion given in (7.4),
i i j ij
u b
` `
j j - o
,
, we can obtain:
( )
2
2
2
, ,
) , (
) (
) (
t
t
i i jj i ji j
o
o
- - -
- - -

x
, ,
,
, ,
` `
u
b u u j j j j A
j j j j A
V V V
u b u u
The Navier-Lam equations (7.140)
which are known as the Navier-Lam equations. With that we have reduced the number of
equations as well as the number of unknowns. Note that the only remaining unknowns are
the displacement components. Finally, as for addressing specific problems, this equation
can be used to obtain an analytical solution of the linear elasticity problem.
7.12 Two-Dimensional Elasticity
Occasionally, three-dimensional structures have certain geometrical and load features that
enable us to treat them as two-dimensional problems (2D) which simplifies the problem
immensely in two aspects: when solving the problem and when interpreting the results.
Fundamentally, there are two kinds of simplifications:
1) Simplification on a conceptual level
Within this class of simplification there are two types of approach:
The state of plane stress;
The state of plane strain.
It should be stressed that such simplification are mere approximations of the real
problem. Nevertheless, in many cases they turn out to be quite satisfactory, i.e. the
error made when using them are insignificant.
2) Simplification on a mathematical level
We use these simplifications in structures that have radial symmetry. Such structures
are known as:
Solids of revolution (or Axisymmetric solids).
7 LINEAR ELASTICITY

411
The results obtained by using this simplification are exactly the same as considering
the problem from a three-dimensional point of view.
7.12.1 The State of Plane Stress
In this type of approach, one of the dimensions of the structural elements is very small
when compared to the other two, (see Figure 7.15), and the load is perpendicular to the
direction of smallest dimension. As a result of this the stress tensor components found in
this direction are equal to zero. The deep beam is an example where we can apply this
approach, (see Figure 7.16).











Figure 7.15: Plate










Figure 7.16: Deep beam.
The state of plane stress field, ) , (
2 1
x x , is characterized by the absence of stress in
one direction which we will show as z x =
3
. Then the stress tensor components can be
characterized by:
(
(
(

o t
t o

(
(
(

o o
o o
o
0 0 0
0
0
0 0 0
0
0
22 12
12 11
y xy
xy x
ij

(7.141)
y x ,
2

x x ,
1
z x ,
3

y x ,
2

q
2D
h
L
x
y
t
NOTES ON CONTINUUM MECHANICS

412
Then, if we consider the above, the equation in (7.85) becomes:
(
(
(
(
(
(
(
(
(
(
(
(
(
(
(

t
t
t
o
o
o
(
(
(
(
(
(
(
(
(
(
(
(
(

v - v -
v - v -
v - v -

(
(
(
(
(
(
(
(
(
(
(
(
(
(
(

r
r
r
zx
yz
xy
z
y
x
zx
yz
xy
z
y
x
G
G
G
E E E
E E E
E E E

1
0 0 0 0 0
0
1
0 0 0 0
0 0
1
0 0 0
0 0 0
1
0 0 0
1
0 0 0
1

(7.142)
Then, if we remove the columns and rows associated with the zero stresses, the stress-
strain relationship for the plane stress case is given by:
(
(
(

t
o
o
(
(
(

v -
v -
v -

(
(
(

r
r

(
(
(

t
o
o
(
(
(
(
(
(

v -
v -

(
(
(

r
r
v -

xy
y
x
xy
y
x
E
G
xy
y
x
xy
y
x
E
G
E E
E E

) 1 ( 2 0 0
0 1
0 1
1

1
0 0
0
1
0
1
) 1 ( 2
(7.143)
The reciprocal of the above equation will result in Hookes law for the state of plane stress:
(
(
(

r
r
(
(
(
(

v -
v
v
v -

(
(
(

t
o
o
xy
y
x
xy
y
x
E

2
1
0 0
0 1
0 1
1
2

( ) [ ] ( ) [ ]
y x y x z z
E E
o - o v - o - o v - o r
1

1

(7.144)
Note that the normal strain
z
r is not equal to zero, since
z
r is not just dependant on the
normal stress
z
o . Then, the strain tensor components are represented as follows:
(
(
(

r
r
r
r
z
y xy
xy x
ij
0 0
0
0
2
1
2
1

(7.145)
7.12.1.1 The Initial Strain
We can incorporate the initial strain (
0
) into the constitutive equation, such as those that
appear from thermal phenomena:
0 0
1
; : : :
e e e
C C C - -
-

(7.146)
If we are considering thermal effects, the equations for the state of plane stress become:



0
0
0
7 LINEAR ELASTICITY

413
Strain:
(
(
(

A -
(
(
(

t
o
o
(
(
(
(
(
(

v -
v -

(
(
(

r
r
0
1
1

1
0 0
0
1
0
1
T
G
E E
E E
xy
y
x
xy
y
x
c
( ) T
E
y x z
A - o - o
v -
r c
(7.147)
Stress:
(
(
(

v -
A
-
(
(
(

r
r
(
(
(
(

v -
v
v
v -

(
(
(

t
o
o
0
1
1
1
2
1
0 0
0 1
0 1
1
2
T E E
xy
y
x
xy
y
x
c
(7.148)
7.12.2 The State of Plane Strain
Let us now consider a structural element with prismatic features, in which the dimension
that corresponds to the direction of the prismatic axis is much larger than the other
dimensions. Additionally, the loads applied are normal to the prismatic axis, (see Figure
7.17). Under these conditions the strain components:
13
r ,
23
r and
33
r are zero. This state
is called the plane strain, examples of which include: retaining walls, cylinders under pressure
(see Figure 7.17), dams (see Figure 7.18), tunnels (see Figure 7.19) and spread footing
foundations.
It must be stressed that, in order to consider a state of plane strain the variables involved
(load, section, material) must be constant along the prismatic axis. Otherwise, significant
errors can occur.










Figure 7.17: Cylinder under pressure.
2D
x
y
z
Cross section
per unit length
x
y
p
p - pressure
prismatic axis
NOTES ON CONTINUUM MECHANICS

414








Figure 7.18: Dam.











Figure 7.19: Tunnel.
If we start from the generalized Hookes law (7.86) and by deleting the columns and rows
associated with the zero strain, i.e.:
(
(
(
(
(
(
(
(
(
(

r
r
r
(
(
(
(
(
(
(
(
(
(

v -
v -
v -
v - v v
v v - v
v v v -
v - v -

(
(
(
(
(
(
(
(
(
(

t
t
t
o
o
o
zx
yz
xy
z
y
x
zx
yz
xy
z
y
x
E

2
2 1
0 0 0 0 0
0
2
2 1
0 0 0 0
0 0
2
2 1
0 0 0
0 0 0 1
0 0 0 1
0 0 0 1
) 2 1 )( 1 (

(7.149)
we obtain:
0
0
0
Dam cross section
1
2D

2D
Tunnel cross section
7 LINEAR ELASTICITY

415
(
(
(

r
r
(
(
(
(

v -
v - v
v v -
v - v -

(
(
(

t
o
o
xy
y
x
xy
y
x
E

2
2 1
0 0
0 1
0 1
) 2 1 )( 1 (
(7.150)
Then, the stress according to the direction z is given by:
( )
y x z
E
r - r
v - v -
v
o
) 2 1 )( 1 (

(7.151)
Additionally, the reciprocal of (7.150) is:
(
(
(

t
o
o
(
(
(

v - v -
v - v -
v -

(
(
(

r
r
xy
y
x
xy
y
x
E

2 0 0
0 1
0 1
1
(7.152)
Afterwards, we can write the constitutive law for the state of stress and strain in a single
equation as:
(
(
(

r
r
(
(
(
(

v -
v
v
v -

(
(
(

t
o
o
xy
y
x
xy
y
x
E

2
1
0 0
0 1
0 1
1
2
(7.153)
where the values of v , E assume the following values:
State of Plane Stress State of Plane Strain
v v ; E E
v -
v
v
v -

1
;
1
2
E
E
(7.154)
7.12.2.1 Thermal Strain
If we take into account the thermal effect, the stress tensor components can be obtained by
means of the following equation:
(
(
(

v -
A
-
(
(
(

r
r
(
(
(
(

v -
v - v
v v -
v - v -

(
(
(

t
o
o
0
1
1
2 1
2
2 1
0 0
0 1
0 1
) 2 1 )( 1 (
T E E
xy
y
x
xy
y
x
c
(7.155)
Note that the above equation is the same as that given in (7.136) when we are dealing with
two-dimensional cases, i.e. when 2 , 1 , j i . Our goal now is to obtain the strain field and to
do so, we will restructure the above equation as:
(
(
(

r
r
(
(
(
(

v -
v - v
v v -
v - v -

(
(
(

v -
A
-
(
(
(

t
o
o
xy
y
x
xy
y
x
E T E

2
2 1
0 0
0 1
0 1
) 2 1 )( 1 (
0
1
1
2 1
c
(7.156)
Then if we multiply the above equation by the matrix given in (7.152) we can obtain:
NOTES ON CONTINUUM MECHANICS

416
(
(
(

r
r

(
(
(

(
(
(

v - v -
v - v -
v -
v -
A
-
(
(
(

t
o
o
(
(
(

v - v -
v - v -
v -
xy
y
x
xy
y
x
E
T E
E

0
1
1
2 0 0
0 1
0 1
1
2 1
2 0 0
0 1
0 1
1 c
(7.157)
with which the strain field becomes:
(
(
(

A v - -
(
(
(

t
o
o
(
(
(

v - v -
v - v -
v -

(
(
(

r
r
0
1
1
) 1 (
2 0 0
0 1
0 1
1
T
E
xy
y
x
xy
y
x
c (7.158)

Problem 7.5: A strain gauge (or strain gage) is a device used to obtain the strain in only one
direction. Consider a strain rosette that contains three strain gauges where there are 45
internal angles, (see Figure 7.20). At one point we have calculated the following strain
values:
3 3 3
10 05 . 0 ; 10 22 . 0 ; 10 33 . 0
- - -
- r r r
y x x

Find the maximum shear stress at the point in question.
Then consider an isotropic linear elastic material with the following mechanical properties:
Pa E 29000 (Youngs modulus); 3 . 0 v (Poissons ratio). Find:
a) the eigenvalues (principal strains) and eigenvectors (principal directions) of the strain
tensor;
b) the eigenvalues (principal stresses) and eigenvectors (principal directions) of the stress
tensor.
Hint: Consider the state of plane strain.












Figure 7.20: Strain rosette.
Solution:
Firstly, we have to obtain the strain tensor components in the system z y x , , and to do so
we will use the coordinate transformation law in order to obtain the component
12
2r
xy
.
Remember that in two-dimensional cases, the normal component in a new system is given
by (see Problem 1.40 in Chapter 1):
) 2 sin( ) 2 cos(
2 2
12
22 11 22 11
11
0 r - 0
r - r
-
r - r
r
The above equation was obtained by means of the transformation law, (see Chapter 1),
which in engineering notation becomes:
) 2 sin(
2
) 2 cos(
2 2
0

- 0
r - r
-
r - r
r
xy y x y x
x

Then,
xy
can be obtained as follows:
y
x
x
45
45
strain gauge
7 LINEAR ELASTICITY

417
3
10 16 . 0 ) 2 cos(
2
) (
2
) (
) 2 sin(
2
-

|
|
.
|

\
|
0
r - r
-
r - r
- r
0

y x y x
x xy

thus
3
10
0 0 0
0 05 . 0 08 . 0
0 08 . 0 33 . 0
-

(
(
(

- r
ij

Then, the stress components can be evaluated as follows:
[ ]
[ ]
[ ] Pa
E
Pa
E
Pa
E
Pa
E
y x z xy xy
x y y
y x x
684 . 4
) 2 1 )( 1 (
; 7846 . 1
) 1 ( 2
5692 . 3 ) 2 1 (
) 2 1 )( 1 (
0462 . 12 ) 2 1 (
) 2 1 )( 1 (
r - r
v - v -
v
o
v -
t
vr - r v -
v - v -
o
vr - r v -
v - v -
o

Additionally, the maximum shear stress is given by:
Pa
xy
y x
5988 . 4
2
2
2
max
t -
|
|
.
|

\
| o - o
t
a) The characteristic equation for the strain tensor (2D) is:
) 10 ( 0 10 29 . 2 28 . 0
3 2 2 - -
- r - r
Then, by solving the above equation we can find the eigenvalues (principal strains) given
by:
3
2
3
1
10 06615528 . 0 ; 10 346155 . 0
- -
- r r
Then, the eigenvectors of the strain tensor are:
(
(
(

-
r
r
r
1 0 0
0 9802 . 0 1979 . 0
0 1979 . 0 9802 . 0
3
2
1

b) Given the stress tensor components, we have:
Pa
ij
(
(
(

o
684 . 4 0 0
0 5692 . 3 7846 . 1
0 7846 . 1 0462 . 12

We now obtain the characteristic determinant and in turn the eigenvalues (principal
stresses) 40654 . 12
1
o , 208843 . 3
2
o . Additionally, the eigenvectors of the stress tensor
are:
(
(
(

-
o
o
o
1 0 0
0 9802 . 0 1979 . 0
0 1979 . 0 9802 . 0
3
2
1

As expected, the eigenvectors of stress and strain are the same; since we are working with
isotropic linear elastic material.

7.12.3 Axisymmetric Solids
In solids of revolution we use the cylindrical coordinate system to express the strain field:
r
u
r
o
o
r ;
z
w
z
o
o
r ;
r
w
z
u
rz
o
o
-
o
o
(7.159)
NOTES ON CONTINUUM MECHANICS

418
where
r
r is the radial strain,
z
r is the axial strain, and
rz
is the shear strain.
We can then introduce the strain in the circumferential direction
0
r as:
P P
P P
r
u
r
r u r

r
r - - r
r
0
2
2 ) ( 2

(7.160)
Then, if we regroup the strain tensor components we can obtain:
(
(
(
(
(
(
(
(

o
o
-
o
o
o
o
o
o

(
(
(
(

r
r
r
0
r
w
z
u
z
w
r
u
r
u
rz
z
r

(7.161)
Next, the generalized Hookes law for a solid of revolution is given by:
(
(
(
(

r
r
r
(
(
(
(
(

v -
v - v v
v v - v
v v v -
v - v -

(
(
(
(

t
o
o
o
0 0
rz
z
r
rz
z
r
E

2
2 1
0 0 0
0 1
0 1
0 1
) 2 1 )( 1 (
(7.162)









Figure 7.21: Stress components Axisymmetric solid.
7.13 The Unidimensional Approach
7.13.1 Beam Structural Elements
Structural elements in which one dimension is much larger than the other two are subject
to a particular stress/strain field and if we use this particular feature the problem can be
greatly simplified. That is, a problem which is three dimensional by nature can be treated as
if it were one-dimensional. A few of the structures that exemplify this problem type are:
beams, trusses, arches, frames.
As two dimensions are smaller than the third one, if we also consider the linear elastic
material and the small deformation regime, a planar cross section of the beam after

0
o

z
o

rz
t

r
o
7 LINEAR ELASTICITY

419
deformation remains planar. Consequentially, the strain and stress fields at the beam cross
section are defined by planes, (see Figure 7.22).
It must be pointed out that in the deep beam case, (see Figure 7.16), the approach adopted
in this subsection is invalidated, since the beam cross section does not remain planar after
deformation (bending).










Figure 7.22: Beam.
If we now make a cut in a cross section according to the orientation of the plane H , in
general, the stress state at a point in this cross section is given as shown in Figure 7.23. The
intensity (or even the nonexistence) of stress depends on the load type (external force) and
on the beam cross section.












Figure 7.23: Stress at a point in the beam cross section.
As we know how the stress is distributed in the cross section, we can define some resultant
internal forces caused by stress components, by integrating the cross-sectional area, namely:
N - internal normal force; M - bending moment; Q - shear force;
T
M - torsional moment.
Next, we will outline how to obtain these internal resultant forces.
Strain diagram Stress diagram
) (z
x
o ) (z
x
r
x
y
z
a) beam b) beam cross section
neutral axis
y
z
H
x
y
z

x
o

xz
t

xy
t
A- cross section area
H
NOTES ON CONTINUUM MECHANICS

420
7.13.1.1 The Internal Normal Force and the Bending Moments
As we mentioned above, the stress distribution in the cross section is defined by a planar
surface, (see Figure 7.24). The normal stress
x
o can be broken down as shown in Figure
7.24. Then, if we consider the normal stress
x
o by itself, it is possible to obtain the internal
normal force ( N ) and the bending moments (
y
M ,
z
M ).

























Figure 7.24: The internal normal force and the bending moments.
Then, by considering Figure 7.25, the bending moment
y
M is defined as follows:
y
S
A
S
A
S
A
x y
I
c
dA z
c
zdA
c
z
zdA M
o

o
o
} } }
2
(7.163)

_

x
y z
) , ( z y
x
o
x
y
z
N

}
o
A
x
dA N
) 1 (

x
y
z

z
M

}
o
A
x z
dA y M
) 3 (

x
y
z

}
o
A
x y
dA z M
) 2 (


y
M
+ +

) 1 (
x
o
) 2 (
x
o
) 3 (
x
o

_

0
) 2 (
o
}
A
x
dA

_

0
) 3 (
o
}
A
x
dA
7 LINEAR ELASTICITY

421
where
y
I is the moment of inertia of the cross section about the y -axis. Then, if we
observe that
z c
x S
o

o
, we can obtain:
z
I
M
z
y
y
x
o ) (
(7.164)
Similarly, we can obtain:
y
I
M
y
z
z
x
o ) (
(7.165)






Figure 7.25: Normal stress distribution in the beam cross section.
7.13.1.2 The Shear Forces and the Torsional Moment
Due to shear stresses, the shear forces
y
Q and
z
Q appear, (see Figure 7.26) as well as the
torsional moment (
T
M ), (see Figure 7.27):
( )
}
t - t
A
yz xz T
dA z y M
(7.166)












Figure 7.26: The shear stresses Shear forces.
Warping is a phenomenon that comes about because shear stress is increasing at one point
whilst decreasing at another, (see Figure 7.28(a)). In the circular cross-section there is no
warping phenomenon, since the shear stress is uniform for given radius ( r ), (see Figure
7.28(b)).

x
y
z
x
y
z

z
Q

}
t
A
xz
dA

}
t
A
xy
dA

y
Q
) , ( z y
xz
t ) , ( z y
xy
t

x
o
neutral axis

S
o
h
c
z
NOTES ON CONTINUUM MECHANICS

422






Figure 7.27: The shear stress (torsional moment).









Figure 7.28: Distribution of the shear stress.
7.13.1.3 The Strain Energy
The strain energy associated with the normal stress,
) 1 ( ) 1 (
x x
Er o (see Figure 7.24), can be
expressed in terms of internal normal force as follows:
} } } } }

o
r o
L
A
L
V
x
V
x x
dx
EA
N
dx dA
EA
N
dV
E
dV U
0
2
0
2
2
2
) 1 (
) 1 ( ) 1 (
2
1
2
1
2
1
2
1
(7.167)
Likewise, we can obtain the strain energy associated with the normal stress
) 2 ( ) 2 (
x x
Er o as:
} } } } } }
r o
L
y
y
A
L
y
y
A
y
y
y
y
L
V
x x
dx
EI
M
dx dA z
EI
M
dx zdA
EI
M
z
I
M
dV U
0
2
2
0
2
2
0
) 2 ( ) 2 (
2
1
2
1
2
1
2
1

(7.168)
In a similar fashion, if we consider the component
) 3 (
x
o , we can obtain:
}

L
z
z
dx
EI
M
U
0
2
2
1

(7.169)
Then, if we follow the same procedure for the other stress components, we can obtain the
strain energy of a bar in function of the internal forces as:
}
|
|
.
|

\
|
-

- - -
L
T
T z
y
z
z
y
y
dx
EJ
M
GA
Q
GA
Q
EI
M
EI
M
EA
N
U
0
2 2
2
2
2
2
2
1

(7.170)
where is the correction factor for the cross-section, and
T
J is the effective polar
moment.

y
z

xy
t

xz
t
b) Circular cross-section a) Rectangular cross-section
r
) , ( z y t

max
t
) (r t
8 Hyperelasticity












8.1 Introduction
Some materials such as elastomers, polymers, rubber and biological matter (arteries,
muscles, skin, etc.) may be subject to large deformations without there being any internal
energy dissipation (which is typical en elastic process). These materials are classified as
being hyperelastic and purely hyperelastic materials have no memory of motion history, i.e.
they are only dependent on the current values of the state variables.
Physically speaking, elastic materials (linear elasticity, hyperelasticity) return to their initial
state once their load disappears, (see Figure 8.1). In other words, the work done during the
loading process is recovered during the unloading process, i.e. there is no internal energy
dissipation (a reversible process).
Our goal in this chapter is to establish the constitutive equations for materials that behave
according to the hyperelasticity theory, also known as Green or nonlinear elasticity. Moreover, we
will limit our analysis to purely mechanical theories, so we have eliminated thermodynamic
variables such as temperature and entropy.
Among the researchers who have used the hyperelastic constitutive equations to model
rubberlike materials we can mention: Alexander (1968), Treloar (1975), Ogden (1984),
Morman (1986) and Holzapfel (2000).





8
Hyperelasticity
423 , Notes on Continuum Mechanics, Lecture Notes on Numerical
Methods in Engineering and Sciences 4, DOI 10.1007/978-94-007-5986-2_10,
International Center for Numerical Methods in Engineering (CIMNE), 2013
E.W.V. Chaves
NOTES ON CONTINUUM MECHANICS

424








Figure 8.1: The stress-strain curve for elastic materials.
8.2 Constitutive Equations
A hyperelastic material supposes the existence of a function which is denoted by the
Helmholtz free energy per unit reference volume ( 1 ). The energy 1 is also known as
strain energy density or the strain energy function, or elastic potential. In hyperelastic materials, the
strain energy function 1 is only dependent on the deformation gradient ) (F , i.e.,
) (F 1 1 .
In pure deformation processes, which do not involve changes caused by entropy or
temperature, internal energy dissipation is equal to zero ( 0
int
D ), which thereby describes
reversible processes. In this way, the Clausius-Planck inequality, (see Chapter 5), for
reversible processes, turns into the following equations:
D D : : - 1 1
` `
0
int
D (Current configuration)

C C
F F
` `
` `
` `
` `
: :
: :
S S
P P
2
1
0
2
1
0
-
-
1 1
1 1
int
D
(Reference configuration)
(8.1)
where is the Cauchy stress tensor, D is the rate-of-deformation tensor, P is the first
Piola-Kirchhoff stress tensor, S is the second Piola-Kirchhoff stress tensor, and C is the
right Cauchy-Green deformation tensor. Then, taking into account the conjugate relations
obtained in Chapter 5 we have:

} } } } } } }

V V V V V V V
dV dV dV dV dV dV J dV
2
1

0
0 0 0 0 0
0 0 0 0 0
F F C E
` ` ` `
: : : : : : :

P P S S D D D
j
j

(8.2)
Then, in summary we can state that the rate of change of the strain energy density can also
be expressed as follows:
_
` ` ` `
Power Stress
D S S P : : : : C E F
2
1
1
(8.3)
where E is the Green-Lagrange strain tensor, and is the Kirchhoff stress tensor with
which we can state that a material is considered to be hyperelastic if and only if the rate of
change of the strain energy is equal to the stress power.
I
o
r
loading
unloading
II
I - linear elastic zone
II - non-linear elastic zone
8 HYPERELASTICITY

425
Then, by evaluating the rate of change of the strain energy ) (F 1 we obtain:
F
F
F F
F
F
F
` `
: :
o
o

o
o
o
o

) ( ) (
) (
1 1
1
t
(8.4)
Next, by substituting the above equation into the internal energy dissipation given by the
equation in (8.1) we obtain:
F
F
F
F F
F
F
F
` ` ` `
: : : :
o
o

o
o
-
) (
0
) ( 1 1
P P (8.5)
Thus we can draw the conclusion that:
F
F
o
o

) ( 1
P (8.6)
NOTE: The tensor P is a function of X
,
and F , i.e. ) , ), ( ( t X X F
, ,
P P , and as P is
directly dependent of X
,
we can study non-homogeneous materials. Here, for the sake of
simplicity, we will omit the material coordinate X
,
.
As discussed in Chapter 6, strain energy has to be objective, i.e. ) ( ) ( ) (
*
F F F Q 1 1 1 ,
where Q is an orthogonal tensor, (see Figure 8.2). Then, by applying the polar
decomposition ( U R F ) and by adopting
T
R Q , we can obtain:
) ( ) ( ) (
) (
) ( ) (

E C
F F
1 1 1 1 1
1
1 1

U
U R Q
Q

(8.7)
where U is the right stretch tensor and R is the polar decomposition rotation tensor, (see
Chapter 2). Since the tensors C , U, and E are directly linked by
2
U C , and 1 - C E 2 ,
the strain energy density can also be expressed in terms of the Green-Lagrange strain
tensor ( E ). Then, similarly to (8.6), it is possible to express the constitutive equation for
stress in the material description. Next, if we take the rate of change of the energy ) (C 1
we can obtain:
C
C
C C
C
C
C
` `
: :
o
o

o
o
o
o

) ( ) (
) (
1 1
1
t
(8.8)
By substituting (8.8) into the internal energy dissipation (8.1) we obtain:
0
) (
2
1
0
) (
2
1
0 ) (
2
1
|
.
|

\
|
o
o
-

o
o
-
-
C
C
C
C
C
C
C
C C
`
` `
` `
:
: :
:
1
1
1
S
S
S
int
D

(8.9)
Note that, the condition in (8.9) must hold for any thermodynamic process. Now, in a
mechanical process with 0 C
`
, the only scenario in which the condition (8.9) remains
valid is when:
C
C
C
C
o
o

o
o
-
) (
2 0
) (
2
1 1 1
S S (8.10)
Likewise, we can show that:
NOTES ON CONTINUUM MECHANICS

426
E
E
C
C
o
o

o
o

) ( ) (
2
1 1
S (8.11)

















Figure 8.2: Objectivity of strain energy density.

Then, if we take into account the relationships between the stress tensors, (see Chapter 3),
we can still express the constitutive equation for stress as follows:
The Kirchhoff stress tensor ( ):
T T T
F
C
C
F F
E
E
F F F
o
o

o
o

) (
2
) ( 1 1
S (8.12)
The Cauchy stress tensor ( ):
T
T
J
F
C
C
F
F F


o
o


) (
2

1
S

T
T
J
J
-
-

o
o

F
F
F
F
) ( 1
P
(8.13)
The first Piola-Kirchhoff stress tensor ( P ):
C
C
F F F F F F
o
o

- -
) (
2
1
S S P
T T T
(8.14)
The Mandel stress tensor ( M):
F
F
F F
C
C
C C
o
o

o
o



) (
) (
2
1
1
T T
P
S M

(8.15)
B
X
,
x
,

X
,

F , ) (F 1
reference
configuration
current
configuration

0
B
B
F F Q
*

) ( F Q 1
U
) (U 1
R
Q
U R F

T
R Q
E C,
) (
) (

E
C
1
1

) ( ) ( ) ( U U R Q Q 1 1 1 F
8 HYPERELASTICITY

427
Hence, we can sum up the different ways of expressing the stress constitutive equations for
hyperelastic materials as:
F
F
F
C
C
C
C
C
E
E
F
C
C
F
C
C
F
F
F
o
o

o
o

o
o

o
o

o
o

o
o

o
o



) ( ) (
2 ;
) (
2
) (
) (
2 ;
) (
2
) (
1 1 1 1
1 1 1
T
T
J
M S
P
Stress constitutive
equations for
hyperelastic materials
(8.16)
The strain energy function ( 1 ) has to satisfy the following:
The normalization condition: 0 ) ( 1 1 , i.e. the strain energy function vanishes
when the material has been completely unloaded, i.e. when 1 F ;
0 ) ( F 1 . The strain energy must increase monotonically with deformation.
In a reversible process (without internal energy dissipation) the following must be satisfied:
The work done is independent of the path.
If we consider a particle which is deformed according to the path I at a given time
interval, said deformation is defined by the deformation gradient [ ]
2 1
, : t t F . Then, the
internal work done associated with this path is:
} } }

2
1
2
1
2
1

2
1

t
t
t
t
t
t
dt dt dt C E F
` ` `
: : : S S P
(8.17)
Now, for an elastic material (reversible) the internal work done is independent of the path,
so, the following must be met:
} }
I I
F F d d : : P P
(8.18)
for any deformation path I, I .
For any closed cycle of deformation the work done is equal to zero:
0 0
} }
E F d or d : : S P
(8.19)
8.2.1 Elastic Tangent Stiffness Tensors
8.2.1.1 The Material Elastic Tangent Stiffness Tensor
The rate of change of the constitutive equation in (8.16), ) (E S , can be expressed as:
E E
E E
E
` ` `
: :
tan
C
o o
o

) (
2
1
S kl
tan
ijkl kl
kl ij
ij
E E
E E
` `
`
C
o o
o


) (
2
E 1
S
(8.20)
where S
`
, E
`
are objective rates, and
tan
C is a fourth-order tensor known as the material
elastic tangent stiffness tensor also called the material tangent elasticity tensor. Remember that the
tensors E and C are related to each other by the equation ( ) C E C E
` `
- 2 2 1 , thus:
E
C C
C
C
C C
C
`
_
` ` `
: :
tan
C
o o
o

o o
o


) (
4

) (
2
2 2
1 1
S S
(8.21)
Then, taking into account the equations in (8.11), (8.20) and (8.21) we can conclude that:
NOTES ON CONTINUUM MECHANICS

428
C C C
C
E E E
E
o
o

o o
o

o
o

o o
o

S
S
2

) (
4

) (
2
2
1
1
tan
C

The material elastic tangent stiffness tensor
(Reference configuration)
(8.22)
Note that the tensors S
`
and E
`
are symmetrical, i.e.
ji ij
S S
` `
,
ji ij
E E
` `
, so, the fourth-
order tensor,
tan
C , must feature at least minor symmetry, i.e.:
tan
jilk
tan
ijlk
tan
jikl
tan
ijkl
C C C C (8.23)
Then, taking into account the equation in (8.22) we can conclude that the tensor
tan
C also
has major symmetry:
tan
klij
ij kl kl ij kl ij
tan
ijkl
E E E E E E
C C
|
|
.
|

\
|
o
o
o
o

|
|
.
|

\
|
o
o
o
o

o o
o


2
1 1 1
(8.24)
Therefore, we can conclude that the tensor
tan
C is symmetric. In the general case
tan
C is
anisotropic and has 21 independent components. For further details regarding symmetry
types see Chapter 7.
8.2.1.2 The Spatial Elastic Tangent Stiffness Tensor
The rate of change of the second Piola-Kirchhoff stress tensor can be obtained by means
of the equation
T - -
F F
1
S , i.e.:
( )
T T
T T T T
T T T
- -
- - - - - -
- - - - - -



- -
- - -
- -
F F
F F F F F F
F F F F F F
l l
l l



`
`
`
` ` `
1
1 1 1
1 1 1
S

(8.25)
where we have considered the equation l
- -
-
1 1
F F
`
which was obtained in Chapter 2.
Remember in Chapter 4 the Oldroyd rate of the Kirchhoff stress tensor was given by:
T
l l - - `


(8.26)
Remember also that ` is not objective, but

is. Then, by substituting (8.26) into (8.25) we


obtain:
T - -
F F

1
S
`

(8.27)
Then, given the relationship between the rate of change of the Green-Lagrange strain
tensor and the rate-of-deformation tensor we have:
F F E D
T
`

(8.28)
Additionally, by substituting the equations (8.27) and (8.28) into the constitutive equation
(8.20) we obtain:
F F F F
E

- -
D
S
T tan T
tan
:
:
C
C

1
` `

lq kl kp
tan
mnpq nt
st
ms
pq
tan
mnpq mn
F F F F
E
D
S
C
C
t

- - 1 1

` `

(8.29)
Then, taking into account the symmetry of the Oldroyd rate of the Kirchhoff stress tensor
(

) and D , we can obtain:


8 HYPERELASTICITY

429
( ) D
D
: : :
:

T T tan
T T tan T T
F F F F
F F F F F F F F


- -
C
C

1
kl
ijkl
lq kp
tan
mnpq im jn
ij
kl lq kp
tan
mnpq im jn jt
st
is
kl lq kp
tan
mnpq im jn nt jn
st
ms im
F F F F
F F F F
F F F F F F F F
D
D
D
_
L
C
C
C
t
t
t
- -

c c
1 1
(8.30)
D : L


(8.31)
where we have introduced the spatial elastic tangent stiffness tensor, also called the spatial
tangent elasticity tensor, in the current configuration, which is given by:
) ( ) (
T T tan
F F F F : : C L
lq kp
tan
mnpq jn im ijkl
F F F F C L (8.32)
Then, from the above equation we can obtain the inverse relationship:
tan
abcd
dq cp
tan
mnpq bn am lq dl kp ck
tan
mnpq jn bj im ai dl ck ijkl bj ai
F F F F F F F F F F F F
C
C C L


- - - - - - - -
c c c c
1 1 1 1 1 1 1 1
(8.33)
Thus
1 1 1 1 - - - -

dl ck ijkl bj ai
tan
abcd
F F F F L C ( ) ( )
T T tan - - - -
F F F F : : L C
1 1
(8.34)
In Chapter 4 we obtained the relationship between the Jaumann-Zaremba rate (

) and the
Oldroyd rate of the Kirchhoff stress tensor (

), i.e.:
D D D D - - - -



(8.35)
Next, by combining the above rate

with the constitutive equation in (8.31) we can


obtain:
D D D
D D D


- -
- -


:
:
L
L


pj ip pj ip kl ijkl ij
D D D t - t - t L


(8.36)
Notice that the tensor D is symmetric, so the double dot product between the symmetric
fourth-order unit tensor,
sym
I , and a symmetric second-order tensor turns out to be the
same tensor, so,
kl pk il pl ik kl
sym
ipkl ip
D D D ) (
2
1
c c c c - I
kl jk pl jl pk kl
sym
pjkl pj
D D D ) (
2
1
c c c c - I (8.37)
Then, by substituting (8.37) into (8.36) we obtain:
[ ]
kl ijkl ijkl kl jk il jl ik il kj ik lj ijkl
kl jk il jl ik il kj ik lj ijkl
kl jk pl ip jl pk ip pk il pj pl ik pj ijkl
kl jk pl jl pk ip kl pj pk il pl ik kl ijkl ij
D D
D
D
D D D
2 ) (
4
1
2
) (
2
1
) (
2
1
) (
2
1
) (
2
1
) (
2
1
) (
2
1
H L L
L
L
L
-
(

t - t - t - t -
(

t - t - t - t -
(

t - t - t - t -
- t - t - - t
c c c c
c c c c
c c c c c c c c
c c c c c c c c


(8.38)
NOTES ON CONTINUUM MECHANICS

430
Therefore, the rate of change of the constitutive equation in terms of Jaumann-Zaremba
rate of the Kirchhoff stress tensor becomes:
D : L


(8.39)
where L

is a fourth-order tensor and is defined by:


H L L 2

-
(8.40)
with
) (
4
1
jk il jl ik il kj ik lj ijkl
c c c c t - t - t - t H (8.41)
We can now summarize the relationships between the rate of change of the stress and the
rate-of-deformation tensor as:
[ ]

- t

t

kl ijkl ijkl ij kl ijkl
ij D D 2

;
H L L

D D : : L L

where

) (
4
1
2

jk il jl ik il kj ik lj ijkl
ijkl ijkl ijkl
lq kp
tan
mnpq im jn ijkl
F F F F
c c c c t - t - t - t
-

H
H L L
C L
The spatial elastic tangent stiffness tensor
(Current configuration)
(8.42)
8.2.1.3 The Instantaneous Elastic Tangent Stiffness Tensor
The relationship between the Cauchy stress tensor and the second Piola-Kirchhoff stress
tensors is given by
T
J
- -
F F
1
S whose rate of change becomes:
T T T T
J J J J
- - - - - - - -
- - - F F F F F F F F
`
`
` `
`

1 1 1 1
S
(8.43)
where ) (D Tr J J
`
and l
- -
-
1 1
F F
`
, (see Chapter 2), which if substituted into the
above equation yields the following result:
( )
T T
J
- - -
- - - F F ) (
1
D S Tr l l
`
`
(8.44)
Now, remember in Chapter 4 that the Truesdell stress rate, which is objective, is given by
) (D Tr - - -
T
T
l l
`
so, we can state that:
T
J
-
T
-
F F
1
S
`

(8.45)
Then, by substituting the equation (8.45) into the constitutive equation in (8.20) and if we
know that F F E D
T
`
, we can obtain:
( ) D
D
S
: : :
:
:
T T tan
T tan T
tan
J
J
F F F F
F F F F
E

T
-
T
-

C
C
C
1


1
` `

(8.46)
or
8 HYPERELASTICITY

431
D : A
T

(8.47)
where A is the instantaneous elastic tangent stiffness tensor, also called the instantaneous elastic
moduli, (see Asaro&Lubarda(2006)), which is defined by:
L C A
J J
T T tan
1 1
F F F F : :
ijkl lq kp
tan
mnpq jn im ijkl
J
F F F F
J
L C A
1 1
(8.48)
where L is the fourth-order tensor given in (8.32). So, in summary we have:
kl ijkl
ij
D A o
T T
; D : A
with

ijkl lq kp
tan
mnpq jn im ijkl
J
F F F F
J
L C A
1 1
The instantaneous elastic tangent stiffness tensor
(8.49)
Then, by taking into account the relationship between the Truesdell stress rate (
T
) and the
Oldroyd rate of the Kirchhoff stress tensor

, i.e.
T
J

, the equation in (8.47) becomes:


D D D : : : L L A
T
J J
1 1
(8.50)
which is the same as that obtained in (8.31).
8.2.1.4 The Elastic Tangent Stiffness Pseudo-Tensor
The constitutive equation for stress that relates the first Piola-Kirchhoff stress tensor to the
deformation gradient is:
F
F
o
o

) ( 1
P
ij
ij
F o
o

) (F 1
P
(8.51)
Remember that F and P are two-point tensors (pseudo-tensors), i.e. they are not defined
in any configuration. Then, the rate of change of the above constitutive equation is given
by:
F
F
F F
F
` `
` `
:
:
K
o o
o

P
P
) (
2
1

kl ijkl ij
kl
kl ij
ij
F
F
F F
` `
` `
K
|
|
.
|

\
|
o
o
o
o

P
P
) (F 1

(8.52)
We can now introduce the elastic tangent stiffness pseudo-tensor also called the elastic
pseudomoduli, (see Lubarda&Benson (2001)), as follows:
F F
F
o o
o

) (
2
1
K klij
ij kl kl ij
ijkl
F F F F
K K
o o
o

o o
o

) ( ) (
2 2
F F 1 1

(8.53)
The elastic tangent stiffness pseudo-tensor is not a real moduli, because it is partially
associated with the material spin tensor, (see Asaro&Lubarda (2006)).
Next, we can relate the tensors K and
tan
C . To do so, we need to evaluate the rate of
change of
pj ip ij
F S P , (see Eq. (8.14)):
pj ip ip pj ij
F F S S P
` ` `
- (8.54)
NOTES ON CONTINUUM MECHANICS

432
Then, by substituting (8.21) and (8.52) into the above equation we obtain:
pj ip ip kl pjkl kl ijkl
F F E F S
` ` `
- C K
(8.55)
and if we know that ( )
ql qk ql qk kl
F F F F E
` ` `
-
2
1
the above equation becomes:
( )
( )
pj ip ql ip qk
tan
pjkl qk ip ql
tan
pjkl
pj ip ip ql qk ql qk
tan
pjkl kl ijkl
F F F F F F F
F F F F F F F
S
S
` ` `
` ` ` `
- -
- -
C C
C K
2
1
2
1

(8.56)
Note that the dummy indices k and l from the expression
qk ip ql
tan
pjkl
F F F
`
C can be
exchanged without altering the result of the expression, and the dummy indices k and q
from
ql ip qk
tan
pjkl
F F F
`
C can also be exchanged, so:
( )
( )
ip pj kl ip kq
tan
pjql kl ip kq
tan
pjlq
ip pj kl ip kq
tan
pjql ql ip qk
tan
pjlk kl ijkl
F F F F F F F
F F F F F F F F
` ` `
` ` ` `
S
S
- -
- -
C C
C C K
2
1
2
1

(8.57)
Then, if we make use of the minor symmetry
tan
pjql
tan
pjlq
C C , we can still state that:
( )
pj ip kl ip kq
tan
pjlq kl ijkl
F F F F F S
` ` `
- C K (8.58)
Next, if we see that
kl ik lj il lj ip pj
F F F
` ` `
c S S S holds, we can conclude that:
( ) -
kl ik lj ip kq
tan
pjlq kl ijkl
F F F F
` `
c S C K
ik lj ip kq
tan
pjlq ijkl
F F c S - C K (8.59)
Thus:
kl ijkl ij
F
` ` ` `
K P ; F : K P

ik lj ip kq
tan
pjlq ijkl
kl ij
ijkl
F F
F F
c
1
S -
o o
o

C K
K
) (
2
F
The elastic tangent stiffness pseudo-tensor
(8.60)
8.3 Isotropic Hyperelastic Materials
If the scalar-valued tensor function ) (C 1 is isotropic it must satisfy the following:
) ( ) (
T
Q Q C C 1 1 (8.61)
for any orthogonal tensor Q. Then if we use the polar decomposition rotation tensor, i.e.
R Q , and if we know that F F C
T
, we can obtain:
) (
) ( ) ( ) ( ) ( ) (
2
b
F F C C
1
1 1 1 1 1

V V V R R R R
T T T T
(8.62)
8 HYPERELASTICITY

433
where
T
V V is the left stretch tensor which is related to the left Cauchy-Green
deformation tensor (
T
F F b ) by means of
2
V b , where
T
R V R V F F is
satisfied, (see Chapter 2). Thus, in isotropic materials, the energy function 1 can be
expressed in terms of the left Cauchy-Green deformation tensor as follows:
) , ( ) , ( x b X C
,
,
1 1
Energy function for isotropic
hyperelastic materials
(8.63)






























Figure 8.3: The elastic tangent stiffness tensors.
ELASTIC TANGENT STIFFNESS TENSORS
RATE-TYPE CONSTITUTIVE EQUATIONS

kl ijkl
ij D L t


kl ijkl ij
D

L t


kl ijkl
ij
D A o
T

X
,

x
,

Reference
configuration
Current
configuration

0
B
B
Material elastic tangent stiffness tensor

C C C
C
E E E
E
o
o

o o
o

o
o

o o
o

S
S
2

) (
4

) (
2
2
1
1
tan
C

F
E
`
`
:
tan
C S
Spatial elastic tangent stiffness tensor

) (
4
1
2

jk il jl ik il kj ik lj ijkl
ijkl ijkl ijkl
lq kp
tan
mnpq im jn ijkl
with
F F F F
c c c c t - t - t - t
-

H
H L L
C L

Instantaneous elastic tangent stiffness tensor

ijkl lq kp
tan
mnpq jn im ijkl
J
F F F F
J
L C A
1 1

kl ijkl ij
F
` `
K P
Elastic Tangent Stiffness Pseudo-Tensor
ik lj ip kq
tan
pjlq ijkl
kl ij
ijkl
F F
F F
c
1
S -
o o
o

C K
K
) (
2
F

NOTES ON CONTINUUM MECHANICS

434
8.3.1 The Constitutive Equation in terms of Invariants
8.3.1.1 The Constitutive Equation in terms of C and b
We showed in Chapter 1 that the scalar-valued isotropic tensor function, ) (C 1 , can be
written in terms of the principal invariants of C (
C C C
I I I I I I , , ), or in terms of the
invariants of b , i.e.:
) , , ( ) , , ( ) (
b b b C C C
C I I I I I I I I I I I I 1 1 1 1 (8.64)
Then, it was verified in Chapter 2 that
b C
I I ,
b C
I I I I and
b C
I I I I I I , which is obtained
from:
) (C
C
Tr I ,
[ ] { ) ( ) (
2
1
2 2
C C
C
Tr Tr - I I ,
[ ]
)
`

- -

3 2 3
2
) (
2
1
) ( ) (
2
3
) (
3
1
) (
C C C C
C
C
Tr Tr Tr Tr
det J I I I

Also in Chapter 1, we showed that for the scalar-valued tensor function,
) , , ( ) (
C C C
C I I I I I I 1 1 , the following equations are valid:
2 1
2
1
) , , (
- -
-
|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o
-
|
|
.
|

\
|
o
o

|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o
-
o
o

|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o

o
o
o
o
-
o
o
o
o
-
o
o
o
o

o
o
C C
C C
C C
C C C C
C
C
C
C
C
C C
C
C
C C
C
C
C
C C
C
C C
C
C C
C
C
C
C
C
C
C C C
I I I
I I
I I I
I I I
I I
I I I
I I I
I
I I I I I
I I
I I I
I
I I I
I I I
I I I I I
I
I I I
I I I
I I I
I I
I I
I
I
I I I I I I
1 1 1 1
1 1 1 1 1 1
1 1 1 1
1 1 1 1
1
1
1

(8.65)
where we held that:
1
o
o
C
C
I
,
2 1 - -
- - -
o
o
C C C C
C
C C C C
C
I I I I I I I
I I
T
1 1 ,
1
C C C C
C
C C C C
C
I I I I I I I I I
I I I
T
- -
o
o
- - 2 1

(8.66)
Now, taking into account the equations in (8.11) and (8.65) we can obtain the constitutive
equation in terms of the principal invariants of C as follows:
(
(

|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o
-
|
|
.
|

\
|
o
o

(
(

|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o
-
o
o

(
(

|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o

o
o

- -
-
2 1
2
1
2
2
2
) (
2
C C
C C
C C
C
C
C
C
C
C
C
C C
C
C
C C
C
C
C
C C
C
C C
C
C C
I I I
I I
I I I
I I I
I I
I I I
I I I
I
I I I I I
I I
I I I
I
I I I
I I I
I I I I I
I
I I I
1 1 1 1
1 1 1 1 1 1
1 1 1 1
1
1
1
1
S
(8.67)
In Chapter 1, in the subsection about the Tensor-Valued Tensor Function, it was shown
that the following relationships are valid:
8 HYPERELASTICITY

435
b C b
b F F b
, , ,
1 1 1
T
(8.68)
Now, if we consider the relationship between the Kirchhoff stress and the second Piola-
Kirchhoff stress tensors,
T
F F S , and the constitutive equation in the reference
configuration (8.11),
C ,
21 S , it is possible to obtain the constitutive equation in the
current configuration as follows:
b b C
b b F F F F
, , ,
2 2 2 + 1 1
T T
S (8.69)
Next, by taking into account the equation J , we can also represent the constitutive
equation for isotropic materials as:
b
b
b b
b
b
o
o

o
o

) (
2
) (
2
1 1
and
b
b
b b
b
b
o
o

o
o

- -
) (
2
) (
2
1 1
1 1
J J (8.70)
Then, in a similar fashion to (8.67), and by considering that J we can obtain:
T T
I I I
I I I I I
I
I I I
J F C C F F F
C
C C
C
C C

(
(

o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o

-1
2
1 1 1 1
1 S
(8.71)
Now, if we consider that F F C
T
and
T
F F b , the above equation becomes:
(
(

o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o
1
b
b b
b
b b
b b I I I
I I I I I
I
I I I
J
1 1 1 1
2
2
(8.72)
Alternatively, we can represent (8.72) by substituting the expression
2
b obtained by means
of the Cayley-Hamilton theorem 0 1 - - -
b b b
b b b I I I I I I
2 3
, (see Chapter 1), i.e.:
1 2
1 2
1 1 1 2 1 3
-
-
- - - -
- -
- - -
- - -
b b b
b b b
b b b b b b b
b b b
b b b
b b b
I I I I I I
I I I I I I
I I I I I I
1
0 1
0 1
(8.73)
Then, the equation in (8.72) can still be written as:
(
(

|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o

-1
2 b b
b
b b
b
b
b
b
I I I
I I I
I I
I I
I I I
I I I
J
1 1 1 1
1
(8.74)
Then, by using equations between stress tensors, (see Chapter 3),
S S P
- -
F F F F F
T T T

(8.75)
the constitutive equation can also be written in terms of the first Piola-Kirchhoff stress
tensor. To do so, let us consider S given in (8.67), and so we obtain:
(
(

o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o

(
(

o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o

- -
-

T T
I I I
I I I I I
I
I I I
I I I
I I I I I
I
I I I
F F F F F F F
C C F
C
C b C
C
C C
C
C C
C
C C
_ _
1
1
1 P
1
1
2
2
1 1 1 1
1 1 1 1

(8.76)
Additionally, by considering that F b F
- -

T T
and the symmetry of b , we obtain:
F
b
b
F b b
C
C C
C
C C

o
o

(
(

o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o

-
) (
2 2
1
1 1 1 1 1
P 1 P I I I
I I I I I
I
I I I
(8.77)
NOTES ON CONTINUUM MECHANICS

436
8.3.1.2 The Constitutive Equation in terms of E
Energy can also be written in terms of the Green-Lagrange strain tensor E , and if we are
dealing with isotropic material the energy constitutive equation can be expressed in terms
of the principal invariants of E :
) , , ( ) (
E E E
E I I I I I I 1 1 (8.78)
where ) (E
E
Tr I , [ ] { ) ( ) (
2
1
2 2
E E
E
Tr Tr - I I , ) (E
E
det I I I . Then, if we consider the
equation in (8.11), i.e.
E
E
o
o

) ( 1
S , we can obtain another one analogous to that obtained in
(8.65), i.e.:
2
2 1 0
E E c c c - - 1 S (8.79)
where the parameters
0
c ,
1
c ,
2
c are given by:
E
E
E E
E
E
E
E E
I I I
c I
I I I I I
c I I
I I I
I
I I I
c
o
o

o
o
-
o
o

o
o
-
o
o
-
o
o

1 1 1 1 1 1
2 1 0
; ;
(8.80)
8.3.2 Series Expansion of the Energy Function
Let us assume that ) (C 1 1 is a continuously differentiable function with respect to the
C -invariants. It is possible, then, to represent 1 by means of infinite power series:
( ) ( ) ( )

- - -
0 , ,
1 3 3 ) , , (
r q p
r q p
pqr
I I I I I I c I I I I I I
C C C C C C
1 1
(8.81)
where the coefficients
pqr
c are independent of the deformation.
Here, we can notice that in an undeformed state we have 1 1 C F , then 3
C
I ,
3
C
I I , 1
C
I I I , which results in 0 1 , as expected, since in said undeformed state strain
energy is zero (normalization condition).
Then, taking into account that
2
U C , where U is the right stretch tensor, it is possible to
express the tensor C in terms of the principal stretches (eigenvalues of U)
1
/ ,
2
/ ,
3
/ ,
(see Chapter 2), so we can use the following spectral representation:


/ /
3
1
) ( ) ( 2 2
3
1
) ( ) (

a
a a
a
a
a a
a
N N U N N U C (8.82)
Next, the principal invariants of C or b in terms of the principal stretches
i
/ are given by:

/ / /
/ / - / / - / /
/ - / - /

(
(
(

/
/
/

2
3
2
2
2
1
2
3
2
1
2
3
2
2
2
2
2
1
2
3
2
2
2
1
2
3
2
2
2
1
2
0 0
0 0
0 0
b C
b C
b C
I I I I I I
I I I I
I I
C
ij ij
U
(8.83)
Then, by substituting the values of (8.83) into the power series (8.81) and after some
mathematical manipulations we obtain:
( ) [ { ( ) ( )]( )

- / / / / - / / - / - / / - / - / / / / /
0 , ,
3 2 1 2 1 3 1 3 2 3 2 1 3 2 1
6 ) , , (
r q p
r
q q p q q p q q p
pqr
a 1 1
(8.84)
8 HYPERELASTICITY

437
where the coefficients
pqr
a are independent of the deformation.
In incompressible materials 1
C
I I I or 1
3 2 1
/ / / is satisfied and the equations in (8.81)
and (8.84) becomes, respectively:
( ) ( )

- -
0 ,
3 3 ) , (
q p
q p
pq
I I I c I I I
C C C C
1 1
(8.85)
( ) [ { ( ) ( )]

- / - / / - / - / / - / - / / / / /
0 ,
2 1 3 1 3 2 3 2 1 3 2 1
6 ) , , (
q p
q q p q q p q q p
pq
a 1 1
(8.86)
8.3.3 Constitutive Equations in terms of the Principal
Stretches
As we have seen before, for isotropic materials, we can express the strain energy function
in terms of the principal stretches
a
/ , 3 , 2 , 1 a , i.e. ) , , (
3 2 1
/ / / 1 1 . Let us now suppose
that
) (

a
N and
) (

a
n are the principal directions (eigenvectors) of the right stretch tensor ( U)
and the left stretch tensor ( V ), respectively, where the following holds:

/
/ /
- / /
3
1
) ( ) ( 1
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) ( 2
3
1
) ( ) (

1
;

;


) 1 (
2
1
;

a
a a
a a
a a
a
a
a a
a
a
a a
a
a
a a
a
n N N n n n V
N N N N U
F F
E
(8.87)
To see how the above relationships are proven, see the Section on Polar Decomposition in
Chapter 2.
Now, the second Piola-Kirchhoff stress tensor ( S ) in terms of the principal stretches
becomes:
|
|
.
|

\
|
o
/ o
/ o
o
-
o
/ o
/ o
o
-
o
/ o
/ o
o

o
/ o
/ o
o

o
o

C C C C C
C
3
3
2
2
1
1
2 2
) (
2
1 1 1 1 1
i
i
S
(8.88)
Then by considering the spectral representation of C given in (8.82), the rate of change of
) (
a
/ C can be evaluated as follows:

/ - / - / /
3
1
) ( ) ( 2 ) ( ) ( 2 ) ( ) (

2
a
a a
a
a a
a
a a
a a
N N N N N N
` `
` `
C (8.89)
Now, if we apply the dot product of
) (

a
N on the right and on the left of both sides of the
equation we have:

|
.
|

\
|
- / - / /



3
1
) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( 2 ) ( ) ( ) ( ) (
) ( ) (

2

a
a a a a a a a a
a
a a a a
a a
a a
N N N N N N N N N N N N
N N
` `
`
`
C
(8.90)
Then, bearing in mind that 1

) ( ) (

a a
N N , and the fact that the rate of change of a vector
with constant magnitude is always orthogonal to itself (Holzapfel (2000)), it follows that
0

) ( ) (

a a
N N
`
and subsequently the above equation becomes:
a a
a a
/ /
` `
2

) ( ) (
N N C (8.91)
NOTES ON CONTINUUM MECHANICS

438
The reader should be aware here that the index 3 , 2 , 1 a is not a dummy index, i.e. we are
not dealing with indicial notation.
Next, using the property ) ( b a T b T a
,
,
,
,
: where a
,
and b
,
are vectors and T is a
second-order tensor, the equation in (8.91) can be rewritten as
a a
a a
/ /
` `
2 )

(
) ( ) (
N N : C ,
and if we also consider that
a
a
a
/
/ o
o
/
` `
C
C ) ( we can obtain:
_
` `
1
) ( ) ( ) ( ) (
) ( ) ( ) ( ) (

2
1
1

2
1
2

2

|
.
|

\
|
o
/ o
|
|
.
|

\
|
/ o
o

|
|
.
|

\
|

/
|
|
.
|

\
|
/ o
o

|
|
.
|

\
|

/
|
|
.
|

\
|
/ o
o

/
/ o
o
/ / /
/ o
o
C
C C C
C C
a
a
a a
a a
a a
a a
a
a a
a
a a
a a
a
a
: : :
: :
N N N N
N N N N

(8.92)
which draws us to the conclusion that:
C o
/ o

/
a a a
a
) ( ) (

2
1
N N
(8.93)
Then, by using the second Piola-Kirchhoff stress tensor expression obtained in (8.88), i.e.

o
/ o
/ o
o

o
/ o
/ o
/ o

3
1
2
) (
2
a
a
a
k
k
a
C C
1 1
S , and by considering the equation obtained in (8.93), we
can express the tenor S as:



/ o
o
/

3
1
) ( ) (
3
1
) ( ) (

1
a
a a
a
a
a a
a a
N N N N S S
1
(8.94)
where
a
S are the second Piola-Kirchhoff stress tensor eigenvalues. Then, by comparing
the equation in (8.94) with the spectral representation of the tensor C , given in (8.82), we
can conclude that in isotropic materials, C and S are coaxial tensors ( C C S S ), i.e.
they have the same principal directions.
Then, as regards the Cauchy stress tensor, we have:
( )
( ) ( )

- -

/ o
o
/

/ o
o
/

/ o
o
/

3
1
) ( ) ( 1
3
1
) ( ) ( 1
3
1
) ( ) ( 1 1

1

1

1
a
a a
a a a
T a a
a a
a
T a a
a a
T
J J
J J
N N N N
N N S
F F F F
F F F F
1 1
1

(8.95)
Moreover, if we take into account the equation
) ( ) (

a
a
a
n N / F , (see subsection 2.8 Polar
Decomposition of F in Chapter 2), we can obtain:


-
o
/ o
o
/
3
1
3
1
) ( ) ( ) ( ) ( 1

a a
a a
a
a a
a
a
J n n n n
1
(8.96)
where
a
o are the Cauchy stress tensor eigenvalues.
Since C and S are coaxial, we can obtain S -eigenvalues by considering the principal
directions of C by means of one of the equations in (8.67):
(
(

/
|
|
.
|

\
|
o
o
- /
|
|
.
|

\
|
o
o
-
o
o
-
|
|
.
|

\
|
o
o

- - 4 2
2
a a a
I I I
I I
I I I
I I I
I I
I I I
C
C
C
C
C
C C
1 1 1 1
S
(8.97)
8 HYPERELASTICITY

439
The first Piola-Kirchhoff stress tensor is given by S P F , then it holds that:



/
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (


a
a a
a
a
a a
a a
a
a a
a
a
a a
a
N n
N n N N N N S P
P
S S S F F F

(8.98)
Thus, we can express the components
a
S as:
(
(

/
|
|
.
|

\
|
o
o
- /
|
|
.
|

\
|
o
o
-
o
o
-
|
|
.
|

\
|
o
o
/
- - 4 2
2
a a a a
I I I
I I
I I I
I I I
I I
I I I
C
C
C
C
C
C C
1 1 1 1
P
(8.99)
We can also express these components in terms of
a
P . To do so, let us consider that:
a a a a
I I I
I I I
I I
I I
I
I
I I I I I I
/ o
o
o
o
-
/ o
o
o
o
-
/ o
o
o
o

/ o
o
C
C
C
C
C
C
C C C
1 1 1 1 ) , , (

(8.100)
Then, the derivatives of
C
I with respect to
a
/ are:
( )
( )
( )
a
a
I
I
I
I
/
/ o
o

/ / - / - /
/ o
o

/ o
o
/ / - / - /
/ o
o

/ o
o
/ / - / - /
/ o
o

/ o
o
2
2
2
2
3
2
3
2
2
2
1
3 3
2
2
3
2
2
2
1
2 2
1
2
3
2
2
2
1
1 1
C
C
C
C

(8.101)
( ) ( )
( ) [ ] ( ) [ ]
( ) ( ) [ ] ( )
4
1
2
1 1
4
1
2
1
2
3
2
2
2
1
2
3
2
2
2
3
2
1
2
2
2
1 1
2
1
2
3
2
2
2
1
2
3
2
2
2
3
2
1
2
2
2
1 1
2
1
2
3
2
1
2
2
2
1 1
2
3
2
2 1
2
3 1
2
2 1
2
3
2
1
2
3
2
2
2
2
2
1
1 1
2 2
2 2
2 2 2
- - - -
- - -
/ - / / / / / / - / / / - / / - / / /
/ / / - / / / - / / - / / / / / / - / / /
/ - / / / / - / / / / - / / - / /
/ o
o

/ o
o
C C
C
I I I I I
I I

(8.102)
which is true for the other principal values, then
( )
4 2
2
- -
/ - / /
/ o
o
a a a
a
I I I I I
I I
C C
C

(8.103)
( ) ( ) ( ) [ ]
2
1 1
2
1
2
3
2
2
2
1 1
2
3
2
2 1
2
3
2
2
2
1
1 1
2 2 2
- -
/ / / / / / / / / / / / /
/ o
o

/ o
o
C
C C
I I I
I I I I I I

(8.104)
which is true for the other principal values, then
2
2
-
/ /
/ o
o
a a
a
I I I
I I I
C
C

(8.105)
Then, by substituting (8.101) into (8.100) we obtain:
(
(

/
|
|
.
|

\
|
o
o
- /
|
|
.
|

\
|
o
o
-
o
o
-
|
|
.
|

\
|
o
o
/
/ o
o
- - 4 2
2
) , , (
a a a
a
I I I
I I
I I I
I I I
I I
I I I
I I I I I I
C
C
C
C
C
C C
C C C
1 1 1 1 1

(8.106)
Additionally, if we compare (8.106) with (8.99) and with (8.97) we can draw the conclusion
that:
a a
a
a
I I I I I I
S P /
/ o
o

) , , (
C C C
1

(8.107)
Then, it holds that:
NOTES ON CONTINUUM MECHANICS

440


/
3
1
) ( ) (
3
1
) ( ) (

a
a a
a
a
a a
a a
N n N n P P S (8.108)
Note that
a
P are not the eigenvalues of P . The Cauchy stress tensor is related to the first
Piola-Kirchhoff stress tensor by means of
T
J F
-
P
1
, after which the eigenvalues of
are given by:
a
a a
J
/ o
o
/ o
-
1
1

(8.109)
which is the same result as that obtained in (8.96). Note that index a does not indicate
summation.
Then, in isotropic materials, the Kirchhoff stress tensor ( ) and the left stretch tensor ( V )
have the same principal directions, and if we consider that J we can obtain:


t /
3
1
) ( ) (
3
1
) ( ) ( 2

a
a a
a
a
a a
a a
n n n n S (8.110)
Now, if we look back at the equations in (8.94), (8.96) and (8.110), we can conclude that
the principal values of the tensors S , , , are interrelated by:
a
a
a
a a a
a
J
t
/
o
/

/ o
o
/

2 2
1 1 1
S
(8.111)
8.4 Compressible Materials
In compressible hyperelasticity materials (which go through a change in volume during the
deformation process), it would be appropriate to separate the motion undergone into
isochoric motion (volume-preserving) and another type characterized by dilatational
transformation (purely volume-change). So, let us consider the multiplicative
decomposition of the deformation gradient, (see Figure 8.4), as follows:
vol
F F F
~

(8.112)
where F
~
show an isochoric transformation (
iso
F F =
~
), and
vol
F describes a dilatational
transformation, (see Figure 8.4). Now let us look back at Chapter 2 subsection 2.13, where
we obtained the following equations:




-
-
-
1
1
1
3
2
3
2
3
2
3
2
3
1
3
1
;
~
;
~
;
~
J J
J J
J J
vol
vol
vol
b b b
C C C
F F F (8.113)
and
J J J J J
vol vol

-
1
3
1
3
1
; 1
~ ~
F F F
(8.114)
8 HYPERELASTICITY

441
Moreover, in Chapter 1 it was proven that
1
2
-

o
o
C
C
J J
, where
b C
I I I I I I J
2
. Likewise,
we can obtain the following relationships:
( )
( ) ( )
( )
1
3
2
1
3
1
3
4
3
4
3
1
3
2

3
1
3
1
3
1
3
1
-
-
-
-
-
- -
-
-
- -
-
o
o
-
o
(

o
o
C C
C
C C C
C
C C
C
C
C
J I I I
I I I I I I
I I I
I I I
I I I
J
T

(8.115)
where we have used
1 - -

o
o
C C
C
C C
C
I I I I I I
I I I
T
, (see Chapter 1). Additionally, we can
obtain:
T
J
J
J J
J
J
J
P
I
I
3
2
1
3
2
1
3
2
3
2
3
2
3
2
3
2
3
1
3
1
) ( ) (
) (
~
-
-
-
-
- -
-
-
-

|
.
|

\
|
-
-
o
o
-
o
o

o
o

o
o
C C
C C
C
C
C
C
C
C
C
C

|
.
|

\
|
-
-
o
o
-
o
o

o
o

o
o
-
-
-
- -
-
-
-
1
3
2
1
3
2
3
2
3
2
3
2
3
2
3
1
3
1
) (
) (
) (
~
kl ij ijkl
kl ij jl ik
kl
ij
kl
ij
kl
ij
kl
ij
C C J
C C J J
C
J
C
C
C
J
C
C J
C
C
I
c c

(8.116)
with which, we introduce the fourth-order tensor P known as the projection tensor with
respect to the reference configuration, (see Holzapfel (2000)):
C C C C - -
- - 1 1
3
1
3
1
I P I P
T
(8.117)












Figure 8.4: Multiplicative decomposition of deformation gradient Kinematic tensors.
B
X
,

x
,

X
,


vol
F
F
~


vol
F F F
~

reference
configuration
current
configuration

0
B
B
pure dilatation
F F C
T

F F C
~ ~ ~

T


T
F F b
1
3
2
J
vol
C

T
vol vol vol
F F b
1
~
F
NOTES ON CONTINUUM MECHANICS

442
8.4.1 The Stress Tensors
Next, we will define the stress tensors in different configurations. To start off we will use
the definitions of the Cauchy stress tensor (current configuration)
T
J
F
C
C
F
o
o

) ( 2 1

and the second Piola-Kirchhoff stress tensor
T
J
- -

o
o
F F
C
C
S
1
) (
2
1
(reference
configuration). Note that we can define a stress tensor, analogous to the Cauchy stress
tensor, in the intermediate configuration ( B ) by means of the transformation
vol
F , i.e.:
T
vol
vol
vol
vol
vol
vol
J
F
C
C
F
o
o

) ( 2 1
(8.118)
Then, if we refer to J J
vol
, 1
3
1
J
vol
F ,
vol
vol
vol
vol
vol
vol
J
J
J
C C
C
o
o
o
o

o
o ) ( ) ( 1 1
, and
1
3
2
1
2 2
-
-

o
o
J
J J J
vol
vol
vol
vol
vol
C
C
, the equation in (8.118) becomes:
1 1 1 1
1 1
1 1
vol
vol vol
vol
vol
vol
vol
vol
vol
vol
vol
vol
vol
vol
vol
vol
T
vol
vol
vol
vol
vol
vol
J
J
J J
J
J
J
J
J
J
J
J
J
J
J
J
J
J
J
J
J J
o
o

|
|
.
|

\
|
o
o

|
|
.
|

\
|
o
o

|
|
.
|

\
|
o
o
o
o

o
o




-
-
) (
2
) ( 2
2
) ( 2
) ( 2 ) ( 2
3
1
3
2
3
1
3
1
1
3
1
3
1
3
1
1 1
1
1 1
C
C
F
C
C
F

(8.119)
Thus,
1
J
J
vol
o
o

) ( 1
(8.120)
We can also define a stress tensor in the intermediate configuration caused by the
transformation F
~
, (see Figure 8.5), as:
C
C
~
)
~
(
2
~
~
o
o

1
S
(8.121)
We will now observe additive decomposition of the strain energy function in two parts,
namely: isochoric and volumetric, i.e.:
) ( )
~
( ) ( ; ) ( )
~
( ) (
~ ~
vol vol vol vol
C C C F F F 1 1 1 1 1 1 - -
(8.122)
Then, if we take the chain rule of derivative of the strain energy function (8.122) we obtain:
J
dJ
J d
t
J
dJ
J d
t
J
vol vol
vol
`
`
`
`
`
) ( ~
~
)
~
( ) (
~
~
)
~
(
) ( )
~
( ) (
~ ~
~ 1 1 1 1
1 1 1 -
o
o

o
o
-
o
o
o
o
- C
C
C C
C
C
C C : :
(8.123)
Now, given that C
C
` `
:
o
o

J
J and
1
2
-

o
o
C
C
J J
, we can obtain C C
` `
:
1
2
-

J
J and the term C
` ~

can be expressed as C C
C
C
C
` `
`
: :
T
J P
3
2 ~
~
-

o
o
. So, the equation in (8.123) can also be
expressed as follows:
8 HYPERELASTICITY

443
C C C
C
C
C C C
C
C
C
` `
` ` `
: : :
: : :
1
3
2
1
3
2
) (
2
~
)
~
(
) (
2
~
)
~
(
) (
~
~
-
-
-
-
-
o
o

-
o
o

dJ
J d J
J
dJ
J d J
J
vol
vol
T
1 1
1 1
1
P
P
(8.124)




















Figure 8.5: Multiplicative decomposition of deformation gradient stress tensors.

In purely elastic materials, internal energy dissipation is zero. Remember that in Chapter 5
in a system with no entropy production, internal energy dissipation in the reference
configuration is given by:
0 ) (
2
1
- C C 1
` `
: S
int
D (8.125)
Then, if we combine the equation in (8.124) with the one above we obtain:
0
2
) (
~
)
~
(
2
0
2
) (
~
)
~
(
2
1
1
3
2
1
3
2
~
~

|
|
.
|

\
|
-
o
o
-
-
o
o
-
-
-
-
-
C
C
C
C
C C C
C
C
C
`
` ` `
: :
: : : :
dJ
J d
J J
J
dJ
J d
J
vol
vol
int
1 1
1 1
P
P
S
S D
(8.126)
1
J
J
vol
o
o

) ( 1

B
X
,

x
,

X
,


vol
F
F
~


vol
F F F
~

reference
configuration
current
configuration

0
B
B
pure dilatation
) ( )
~
( ) (
~
vol vol
C C C 1 1 1 -

T
J
- -

o
o
F F
C
C
S
1
) (
2
1


vol
S S S -
~
with
1
) (
-
o
o
C
J
J
J
vol
vol
1
S , S S
~
~
3
2
: P
-
J

T
J
F
C
C
F
o
o

) ( 2 1

) (
vol vol
F 1
)
~
( )
~
(
~
F F 1 1 =
iso


C
C
~
)
~
(
2
~
~
o
o

1
S
1
J
J
vol
o
o

) ( 1

1
~ ~
J F
NOTES ON CONTINUUM MECHANICS

444
Notice that the above must be satisfied for any admissible thermodynamic process. Let us
now consider that 0 C
`
, so, the only way for (8.126) to be satisfied is if:
1
3
2
) (
~
)
~
(
2
~
-
-
-
o
o
C
C
C
dJ
J d
J J
vol
1 1
: P S
(8.127)
Next, if we take the definition of the tensor S given in (8.11), and use the definition of
energy in (8.122), we can obtain:
[ ]
vol
vol vol
vol vol
S S S -
o
o
-
o
o
-
o
o

o
o

~ ) (
2
)
~
(
2 ) ( )
~
( 2
) (
2
~
~
C
C
C
C
C C
C C
C 1 1
1 1
1

(8.128)
where it holds that:
1 1
) (
2
1 ) (
2
) (
2
) (
2
- -
o
o

o
o

o
o
o
o

o
o
C C
C C
C
J
J
J J
J
J J
J
J
vol vol vol vol vol
vol
1 1 1 1
S
(8.129)
and
S S
~
~
)
~
(
2
~
~
)
~
(
2
)
~
(
2
~
3
2
3
2
~ ~ ~
: : : P P
- -

|
|
.
|

\
|
o
o

o
o
o
o

o
o
J J
C
C
C
C
C
C
C
C 1 1 1

(8.130)
where we have used the definition in (8.121). Additionally, we can verify that the tensor S
~

is in the intermediate configuration, i.e. it is only characterized by a change of shape, (see
Figure 8.5). Then, in summary we have:
vol
S S S -
~

(8.131)
where:
C
C
C
C
~
)
~
(
2
~ ~
~
)
~
(
2
~
3
2
3
2
~
o
o

o
o

- -
iso
with J J
1 1
S S S : : P P

(8.132)
1 1
) (
- -
C C Jp
dJ
J d
J
vol
vol
1
S

(8.133)
In addition, with the constitutive equation for hydrostatic pressure, Holzapfel (2000):
DJ
J
p
vol
D ) ( 1


(8.134)
It is worth mentioning that the operator C C -
-1
3
1
I P given in (8.132) provides the
correct deviatoric operator in the material (Lagrangian) description:
[ ] [ ]
1
) (
3
1
) ( ) , (
-
- - - - C C X :
dev
t
,
(8.135)
Thus,
[ ]
dev
J J J J S S S S S S
~
~
3
1
~ ~
3
1
~
~
3
2
1
3
2
1
3
2
3
2 -
-
-
-
- -

|
.
|

\
|
-
(

- C C C C : : : I P (8.136)
Additionally, it holds that:
[ ] 0
~
C :
dev
S
(8.137)
8 HYPERELASTICITY

445
8.4.2 Compressible Isotropic Materials
In compressible isotropic materials, the energy function decomposition can be given by:
) ( )
~
( ) ( ) (
~
J
vol
1 1 1 1 - b b C (8.138)
Then, as
b
I I I J , we can show that if equations
1
2
-

o
o
C
C
J J
and (8.116) are valid, so are
the following:

|
.
|

\
|
-
o
o
|
.
|

\
|
-
o
o

o
o
-
-
-
-
-
1
3
2
1
3
2
1
3
1
~
3
1
~
;
2
kl ij jl ik
kl
ij
b b J
b
b
J
J J
c c
b b
b
b
b
b
I

(8.139)
Then, if we refer to the constitutive equation for isotropic hyperelastic materials obtained
in (8.70),
b
b
b b
b
b
o
o

o
o

- -
) (
2
) (
2
1 1
1 1
J J , and incorporate the definition of energy
(8.138) into the stress equation, we can obtain:
[ ]
vol
vol
vol
J
J J J J J

-
o
o
-
o
o
-
o
o

o
o

- - - -
~
) (
2
)
~
(
2 ) ( )
~
( 2
) (
2
1 1 1 1
~
~
b
b
b
b
b
b b
b
b
b
b 1 1
1 1
1
(8.140)
Additionally, the volumetric contribution is:
1
J
J J
J
J
J
J
J
J
J
J
J
vol vol vol vol
vol
o
o

o
o

o
o
o
o

o
o

- - - -
) (
2
) (
2
) (
2
) (
2
1 1 1 1
1 1 1 1
b b b
b
b
b

(8.141)
And, the isochoric contribution is:
[ ]
dev
J J
J J J
J J J
1 1

~ ~
3
1 ~
~
)
~
(
2
~ ~
3
1
~
~
)
~
(
2
~
3
1
~
)
~
(
3
1
2
)
~
(
2
~
~
~
~ ~
1 1 1
3
2
1 1
3
2
1
3
2
1
3
2
1 1
|
.
|

\
|
-
o
o
|
|
.
|

\
|
-
o
o
|
|
.
|

\
|
-
o
o
|
.
|

\
|
-
o
o




- - -
-
- -
-
-
-
-
-
- -
: :
:
:
I I
I
I
b
b
b
b b b b
b
b
b
b b b b
b
b
b b b
b
b
b
1
1
1 1

(8.142)
where we have used the relationship b b
~
3
2
J , and it can be proven that if - is a second-
order tensor, then it holds that:
dev
- - |
.
|

\
|
- : 1 1
3
1
I , where
dev
- represents the deviatoric
part of the tensor -, (see Problem 1.26).
Next, we will make the algebraic operations carried out in (8.142) using indicial notation:
NOTES ON CONTINUUM MECHANICS

446
|
|
.
|

\
|
o
o
(

|
.
|

\
|
-
o
o
(

|
.
|

\
|
-
o
o
(
(

|
|
.
|

\
|
-
o
o
(
(

|
.
|

\
|
-
o
o
(
(

|
.
|

\
|
-
o
o
(
(

|
.
|

\
|
-
|
|
.
|

\
|
o
o
o
o

o
o
o
-
- - - -
-
-
- -
-
-
- -
-
-
-
-
-
-
-
-
- -

ts
tq
sq ij qj is
ts
tq
kj ik pq sp qj sp ip
ts
tq
kj ik pq sp sp qj pk ik
tp
sp ts
tq
kj pq qj pk ik
tp
tq
kj pq qj pk ik
pq
kj pq qj pk ik
kj
pq
pq
ik
kj
ik ij
b
b
J
b
b
b b b b b b J
b
b
b b J b b b b J J
b b
b
b b J b J
b
b b J b J
b
b b J b J
b
b
b
b J
b
b J
~
~
)
~
(
2
3
1
~
~
)
~
( ~ ~
3
1 ~ ~
2
~
~
)
~
( ~
3
1 ~
2
~ ~
~
)
~
(
3
1
2
~
)
~
(
3
1
2
~
)
~
(
3
1
2
~
~
)
~
(
2
)
~
(
2
~
~
~
~
~
~
~
~ ~
1
1 1 1 1
1
3
2
1 1
3
2
1
1 1
3
2
1
1
3
2
1
1
3
2
1
1 1
b
b
b
b
b
b
b b
1
c c c c
1
c
1
c c
1
c c
c
1
c c
1
c c
1 1
c
_

(8.143)
which is in accordance with (8.142).
In summary, in compressible isotropic materials, the following is satisfied:
vol
-
~

(8.144)
where
1 1 p
J
J
vol
vol

o
o

) ( 1
,
[ ]
dev
1 1
~ ~
3
1
~
|
.
|

\
|
- : I

(8.145)
in which the constitutive equation for hydrostatic pressure is:
DJ
J
p
vol
D ) ( 1


(8.146)
8.4.2.1 Compressible Isotropic Material in terms of the Invariants
In isotropic materials, the strain energy function can be expressed in terms of the principal
invariants as:
) ( ) , ( ) ( ) , ( ~ ~ ~ ~
~ ~
J I I I J I I I
vol vol
1 1 1 1 1 - -
b b C C
(8.147)
where
vol
1 is a function of the third invariant of the right Cauchy-Green deformation
tensor (
2
J I I I
C
). This function, for an undeformed state, has to fulfill the following:

o
o
0 ; 0 1
C
C
C
I I I
I I I
vol
vol
1
1 1 (8.148)
In Chapter 2, subsection 2.13, we obtained the principal invariants of C
~
in terms of those
of C , i.e.:
8 HYPERELASTICITY

447
1 ; ; ~ ~ ~
3
2
3
4
~ ~
3
3
2
~
- -
b C b
C
C
C
C b
C
C
C
C
I I I I I I I I
I I I
I I
I I J I I I
I I I
I
I J I
(8.149)
Note that the invariants
C
~ I and
C
~ I I are independent of the volumetric deformation.
We can now express the constitutive equation in the material description by means of S :
vol
S S S -
~

(8.150)
where
vol
S is the volumetric part, and S
~
is the isochoric part, both of which are given
respectively by:
S S S
~
~
)
~
(
2
~
;
) (
3
2
3
2
1 1
~
: : P P
- -
- -

o
o
J J Jp
dJ
J d
J
vol
vol
C
C
C C
1 1

(8.151)
where
C
C
~
)
~
(
2
~
~
o
o

1
S can be demonstrated by:
(
(

|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o

o
o
C
C
C
C C
C
C
C C
C
C C C C
~
) , ( ) , ( ) , (
2
~
) , (
2
~
~
~ ~
~
~
~ ~
~
~ ~ ~ ~
~ ~ ~ ~
I I
I I I
I
I I
I I I
I
I I I I I I 1 1 1 1
1 S (8.152)
where we have used one of the relationships obtained in (8.67).
Then, a very simple model for the volumetric part is ) ( ) ( J
vol vol vol
1 1 1 F , where
) (J
vol
1 is given by:
( ) ( ) 1
2
1
2
) (
2
-
x
-
x

C
I I I J J
vol
1 (8.153)
and where x is the bulk modulus. This model at the limit 0 J has no physical meaning
since
2
) 0 (
x -
J
vol
1 . Therefore, we can add the term ( )
2
log
2
) ( J J
vol
x
1 into the
equation, i.e.:
( ) J J J
vol
log 2 1
4
) (
2
- -
x
1 (8.154)
which validated the following: when 0 J , the energy function tends towards infinity, i.e.
) 0 (J
vol
1 . Additionally, in a small deformation regime 1 = J , the term 0 log 2 J .
8.5 Incompressible Materials
Many polymers can be subjected to large deformations without any volume change being
observed, Holzapfel (2000). Hence, these materials can be considered to be incompressible,
i.e. the continuum here is characterized by isochoric motion and the following is fulfilled:
1 ; 1 ) (
2
3 2 1
/ / /
b C
F I I I I I I J J det (8.155)
In incompressible materials, ) 1 ( J , the stress state is not completely determined by the
strain state, because in an incompressible body we can add hydrostatic stress (pressure) to
the current stress state without changing the strain state. Remember that, in isotropic
materials, a hydrostatic state produces only volumetric deformations and because of this,
NOTES ON CONTINUUM MECHANICS

448
the volumetric deformation in incompressible materials is equal to zero for any hydrostatic
state. Note that here, even energy is not affected by the volumetric part, since:
0 ) ( ) ( 1 1 1
vol
F .
According to the equation in (8.113), which is an incompressible case, it holds that F F
~
,
1
vol
F , C C
~
, 1
vol
C .
Then, the hydrostatic stress state is given by:
1 p
hyd
-
ij
hyd
ij
pc - o (8.156)
where p denotes pressure. Then, if we refer to the relationship between the Cauchy stress
and the second Piola-Kirchhoff stress tensors:
T
J F F
-
S
1
, we can obtain the
volumetric part of S :
1 1 1 1

- - - - - - -
- - - - C B F F F F F F p J p J p J p J J
T T T hyd hyd
1 S (8.157)
where B is the Piola deformation tensor, given by
1 1 - - -
C F F B
T
, (see Chapter 2,
subsection 2.6).
Then, for incompressible materials we have:
1

) (
2
-
-
o
o
C
C
C
p J
1
S (8.158)
Next, to solve a problem with a constraint ) 1 ( J , we can introduce the Lagrange
multiplier , which must satisfy the following:
( ) 1 ) ( - - J 1 1 C (8.159)
Now, the second Piola-Kirchhoff stress tensor can be obtained by taking the derivative of
the above equation with respect to C , i.e.:
( ) [ ]
1
1

) (
2
2
2
) (
2 2
) (
2 1 ) ( 2 2
-
-
-
o
o

-
o
o

o
o
-
o
o
- -
o
o

o
o

C
C
C
C
C
C
C C
C
C
C C

1
1
1
J
J J
J S

(8.160)
Now, if we compare (8.160) with (8.158) we can conclude that:
3
kk
p
o
- (8.161)
where p (pressure) is an unknown function to be determined by the incompressibility
condition.
Now, let us consider the Cauchy stress tensor decomposition into a spherical (hydrostatic)
and deviatoric part, i.e. 1
m
dev
o - , then, the second Piola-Kirchhoff stress tensor,
defined as
T
J
- -
F F S
1
, can be split as follows:
hyd dev
m
dev
T
m
T dev T
m
dev T
J
J J J J
S S S
1 1 S
- o -
o - o -
-
- - - - - - - -

1
1 1 1 1
) (
C
F F F F F F F F

(8.162)
Then, by comparing the result above with (8.160) we can clearly identify the deviatoric part
of the second Piola-Kirchhoff stress tensor.
After that, if we refer to the equation in (8.16) we can still write that in (8.160) as:
8 HYPERELASTICITY

449
T T
p p
- - -
-
o
o
-
o
o
F
F
F
F F F
C
C
F F
) (

) (
2
1
1 1
P S (8.163)
and:
1 1
P

) (

) (

) (
p p
p
T
T
T T T T
- |
.
|

\
|
o
o
-
o
o

-
o
o



-
F
F
F F
F
F
F F F
F
F
F
1 1
1

(8.164)
where we have considered the incompressibility condition 1 J . Then, in short, we can
state that the constitutive equation for stress in an incompressible continuum can be given
by:
1 1
P
S

) (

) (

) (

) (
2
1
p p
p
p
T
T
T
- |
.
|

\
|
o
o
-
o
o

-
o
o

-
o
o


-
-
F
F
F F
F
F
F
F
F
C
C
C
1 1
1
1
The constitutive
equation for
hyperelastic
incompressible
materials
(8.165)
8.5.1 Geometrical Interpretation
In this section we will attempt to make a graphic interpretation of the results obtained
previously, (see Bonet&Wood(1997)). To start with, let us consider the internal energy
dissipation equation obtained in (8.9):
0
) (
2
1
|
.
|

\
|
o
o
- C
C
C
`
:
1
S
int
D (8.166)
As we saw before, in incompressible materials, the stress state is not completely defined by
the strain state. Then, the term written in parentheses in the equation in (8.166) is not equal
to zero, which indicates that C
`
is not arbitrary, i.e. it has restrictions. Remember that the
rate of change of the Jacobian determinant is given by C C
` `
:
1
2
-

J
J , (see Problem 2.12 in
Chapter 2). Additionally, if we consider the incompressibility condition 1 J , during
motion 0 J
`
, we obtain:
0
2
1

-
C C
` `
:
J
J (8.167)
The above equation gives us the restrictions on C
`
which the equation in (8.166) has to
satisfy, thereby implying that:
1
2
) (
2
1
-

o
o
- C
C
C J

1
S (unit of stress) (8.168)
where , as seen above, coincides with the hydrostatic pressure module.
Let us now consider an arbitrary plane defined by the normal n . Next, we will project the
following tensors onto this plane, i.e.:
1
) ( 1 ) ( ) (
2

2
;
2
1
;
-
-
|
.
|

\
|
o
o
-
n n n
d n C n
C
d n C
, ,
`
,
`
J J
t S
1
(8.169)
NOTES ON CONTINUUM MECHANICS

450
with which, the equations in (8.166) and (8.167) can be rewritten as:
0
2
; 0
) (
1
) ( ) ( ) (

-
n n n n
d d d
`
, ,
`
, ,
J
t (8.170)
which indicates that
) (n
t
,
and
1
) (
-
n
d
,
are orthogonal to
) (n
d
`
,
, (see Figure 8.6).












Figure 8.6: Incompressibility restriction.
8.5.2 Isotropic Incompressible Hyperelastic Materials
In isotropic incompressible hyperelastic materials, the scalar-valued tensor function
) ( ) ( b C 1 1 1 can be expressed in terms of the invariants
b C
I I and
b C
I I I I :
) , ( ) , (
b b C C
I I I I I I 1 1 1 (8.171)
or in terms of the principal invariants of the Green-Lagrange strain tensor:
) , (
E E
I I I 1 1 (8.172)
Then, the constitutive equation for stress in hyperelastic materials becomes:
E E E E E
C C
C
C
C
C C C
o
o
-
o
o
o
o
-
o
o
o
o

o
o
-
o
o

I I I
p
I I
I I
I
I
I I I
p
I I I 1 1 1 ) , (
S
(8.173)
Next, the relationships between the principal invariants of the tensors E and C (obtained
in Problem 2.10 in Chapter 2) are interrelated by:
( )
( )
( ) 1 2 4 8 1
8
1
3 4 4 3 2
4
1
3 2 3
2
1
- - - - - -
- - - - -
- -
E E E C C C C E
E E C C C E
E C C E
I I I I I I I I I I I I I I I I I I
I I I I I I I I I I
I I I I
Reciprocal
(8.174)
with which we can obtain the following derivatives:
n

1
) ( ) (
2
-

n n
d
, ,
J
t

1
) (
2
-
n
d
,
J

plane normal to n

) (n
d
`
,

plane for possible values of
) (n
d
`
,

8 HYPERELASTICITY

451
( ) 1 2 3 2 -
o
o

o
o
E
C
E E
I
I
, ( ) ( ) 1 1 4 4 3 4 4 - - - -
o
o

o
o
E
E E
E E E
C
I I I I
I I

( ) ( ) 1 1 2 4 8 1 2 4 8
1
- - - - - -
o
o

o
o
-
E E
E E
E E E E E
C
I I I I I I I I I I
I I I

(8.175)
Then the equation in (8.173) becomes:
( ) [ ] ( ) [ ]
( ) [ ] ( ) [ ]
1
1
4 2 1 2 2 1 4 2
2 4 8 4 4 2
-
-
- - - - - -
o
o
-
o
o

- - - - - -
o
o
-
o
o

E E E
E E E
E E E
C C
E E E
C C
I I I I p I
I I I
I I I I p I
I I I
1 1 1
1 1 1 1 1 S
1 1
1 1
(8.176)
Now, if we take into account that 1
C
I I I into the equation in (8.174), we can then express
E
I I I as a function of
E
I and
E
I I , i.e.:
( )
E E E
I I I I I I - - 2
4
1
(8.177)
Then, by using the equation in (8.72), the constitutive equation for isotropic incompressible
materials becomes:
1 p
I I
I
I I I
-
(
(

o
o
-
|
|
.
|

\
|
o
o
-
o
o

2
2 2 b b
b
b
b b
1 1 1

(8.178)
Note that we can still express the constitutive equation for incompressible hyperelastic
materials in terms of principal stretches, by means of the equation in (8.96):
3 , 2 , 1 ,
/ o
o
/ - o a p
a
a a
1

(8.179)
8.5.2.1 Series Expansion of the Energy Function for an Isotropic
Incompressible Hyperelastic Materials
In incompressible materials it holds that 1
C
I I I , and the energy function ) , (
C C
I I I 1 1
can be represented by means a power series, (see equation (8.81)), so:
( ) ( )

- -
0 ,
3 3 ) , (
q p
q p
pq
I I I c I I I
C C C C
1 1
(8.180)
8.6 Examples of Hyperelastic Models
Several models have been developed to simulate the phenomenological behavior of
hyperelastic materials. Here we mention some hyperelastic material models that can be
found in the literature.
Remember that in elastic (or hyperelastic) materials, the only remaining constitutive
equations are the one for energy and that for stress, one of which is redundant, that is, if
we know the energy we can find the stress and vice versa, (see Problem 6.1).
NOTES ON CONTINUUM MECHANICS

452
8.6.1 The Neo-Hookean Material Model
In the Neo-Hookean material model, the strain energy function is given in terms of the
isochoric part of C as follows:
( ) ( ) 3 3
2
1
- -
C C
I c I
j
1 (8.181)
where
2
1
j
c . (The parameter j was originally determined by statistical mechanics,
Treloar (1944)), by T
N
x
2
j , where N is the number of polymer chains per unit of the
reference volume, x is the Boltzmann constant and T is the temperature. Lastly, the
parameter, G j can be determined experimentally and is known as the shear modulus.
Then, the stress constitutive equation for the Neo-Hookean material model becomes:
b
1
2c p - - 1 (8.182)
8.6.2 The Ogden Material Model
This model expresses the strain energy function in terms of the principal stretches and is
given by:
( ) ( ) ( ) [
( ) ] ( )
3 2 1 3 2
1
3 1 2 1
1
3 2 1 3 2 1
3
3 ) , , (
/ / / - - / /
- / / - / / - - / - / - / / / /


h
b a
q
N
q
q q
q
M
p
p p p
p
,
, ,
c c c
1
(8.183)
where
p
a ,
q
b are positive constants, 1
p
a , 1
q
b and h is a one-variable convex
function.
8.6.2.1 The Incompressible Ogden Material Model
This model expresses the strain energy function in terms of the principal stretches and is
given by:
( )

- / - / - / / / /
N
p
p p p
p
p
1
3 2 1 3 2 1
3 ) , , (
c c c
c
j
1
(8.184)
where N ,
p
j ,
p
c are the material constants. In general, the shear modulus j, in the
reference configuration, becomes:
0 2
1
>

p p
N
p
p p
with c j c j j
(8.185)
In the literature, e.g. Holzapfel (2000), we can find the following values for the constants
when 3 p :
2 5
3
2 5
2
2 5
1
3 2 1
/ 10 1 . 0 ; / 10 012 . 0 ; / 10 3 . 6
0 . 2 ; 0 . 5 ; 3 . 1
m N m N m N -
-
j j j
c c c

(8.186)
Then, when 1 N and 2 c , the equation in (8.184) yields:
( ) ( ) 3
2
3
2
2
3
2
2
2
1
- - / - / - /
C
I
j j
1 (8.187)
8 HYPERELASTICITY

453
which is the Neo-Hookean material model given in (8.181).
8.6.2.2 The Hadamard Material Model
This model is a simplified Ogden material model, where it holds that 1 N M and
2
1 1
, c which reduces the equation in (8.183) to:
( ) ( ) ) ( 3 3 ) (
1 1
J h I I b I a - - - -
C C
C 1 1 (8.188)
Then, taking into account the equations in (8.11) and (8.67), the constitutive equation
becomes:
( )
(

o
o
- - -
o
o

- -
C
C C
C
C
C C
J
J h I I I I I b a ) ( 2
) (
2
2 1
1 1
1 S
1
(8.189)
Afterwards, the derivative of the Jacobian determinant with respect to the tensor C can be
evaluated as follows:
( ) ( ) ( )
1 1
2
1
2
1
2
1
2
1
2
1
2
1
- -
- -

o
o

o
o

o
o
C C
C C C
C C
C
C C
J I I I I I I
I I I
I I I I I I
J
(8.190)
Finally, the equation in (8.189) may also be rewritten as:
(

- |
.
|

\
|
- -
- - 2 2
1
1
1 1
) (
2
1
2 C C
C
J b J J h I I b a 1 S (8.191)
8.6.3 The Mooney-Rivlin Material Model
The Mooney-Rivlin material model was originally formulated to simulate rubber-like
materials, today it is also used to simulate biological tissue-like materials.
8.6.3.1 Strain Energy Density
This model has the same energy expression as that provided by (8.184). Then, with the
parameter values 2 N , 2
1
c , 2
2
- c , and with the constraint 1
2
3
2
2
2
1
/ / /
(incompressibility), the strain energy density given in (8.184) becomes:
( ) ( ) 3
2
3
2
) , , (
2
3
2
2
2
1
2 2
3
2
2
2
1
1
3 2 1
- / - / - / - - / - / - / / / /
- - -
j j
1 (8.192)
Note that
C
~
2
3
2
2
2
1
I / - / - / and:
C
C
C
I I
I I I
I I

/ / /
/ / - / / - / /

/
-
/
-
/
2
3
2
2
2
1
2
2
2
1
2
3
2
1
2
3
2
2
2
3
2
2
2
1
1 1 1

(8.193)
since we have the constraint 1
C
I I I we can summarize the strain energy density as follows:
( ) ( ) ( ) ( ) 3 3 3
2
3
2
) (
2 1
2 1
- - - - - -
C C C C
C I I c I c I I I
j j
1 (8.194)
where
2
1
1
j
c and
2
2
2
j
- c . Then, the terms ( ) 3 -
C
I and ( ) 3 -
C
I I ensure that the
strain energy is zero when there is no deformation ( 0 E ), since in this scenario and
according to the equation in (8.172), we will obtain ( ) 3
C
I and ( ) 3
C
I I . Note that in the
NOTES ON CONTINUUM MECHANICS

454
particular case when ( ) 0
2
c we revert to the Neo-Hookean material model given in
(8.181).
8.6.3.2 The Stress Tensor
The second Piola-Kirchhoff stress tensor for the Mooney-Rivlin material model becomes:
( ) [ ] C C
C C
C
C
C
C
C
C C
C
C C
C
C C
2 2 1
1
2 2
2
) (
2
c I c c
I I
I
I I I
I I I
I I I I I
I
I I I
- -
(
(

|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o

(
(

|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o

o
o

-
1 1
1 S
1 1 1
1 1 1 1 1

(8.195)
and the Cauchy stress tensor can be obtained as follows:
1
2 1
2 2
-
- - - b b c c p1 (8.196)
8.6.4 The Yeoh Material Model
8.6.4.1 Strain Energy Density
The Yeoh material model is used to simulate isotropic incompressible materials. Our
starting point here is the series expansion of strain energy density:
( ) ( ) ( )

- - -
N
r q p
r q p
pqr
I I I I I I c I I I I I I
1 , ,
1 3 3 ) , , (
C C C C C C
1 1
(8.197)
Then, by considering the incompressible material, 1
C
I I I , and also by discarding the
second invariant we obtain:
( )

-
3
1
3 ) (
N
p
p
p
I c I
C C
1 1
(8.198)
with which we can obtain the strain energy density for the Yeoh material model as follows:
( ) ( ) ( )
3
3
2
2 1
3 3 3 - - - - -
C C C
I c I c I c 1 (8.199)
where
1
c ,
2
c and
3
c are material constants.
8.6.4.2 The Stress Tensor
In this model the second Piola-Kirchhoff stress tensor becomes:
( ) ( ) [ ]1 1
1 S
3 3 3 2 2 2
2
) (
2
2
3 2 1
1
- - - -
o
o

(
(

|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o

o
o

-
C C
C
C
C C
C
C C
C C
C
C
I c I c c
I
I I I
I I I I I
I
I I I
1
1 1 1 1 1
(8.200)
8.6.5 The Arruda-Boyce Material Model
The Arruda-Boyce Material Model, also called the 8-chain model, takes into consideration
that the shear modulus, j, depends on the strain. This phenomenon is detected in some
polymers.
8 HYPERELASTICITY

455
Here, the strain energy density is given by:
( )

-
-
N
p
p p
p
lock
p
I
c
1
2 2
0
3 ) (
C
C
A
j 1 1
(8.201)
where
0
j is the initial shear modulus,
i
c are constants obtained by statistical theory and
lock
A and N are material constants. Then, if we consider that 3 N we obtain:
( ) ( ) ( )
( ) ( ) ( )
(
(

- - - - -
(
(

- - - - -
27
1050
11
9
20
1
3
2
1
27 9 3 ) (
3
4
2
2
0
3
4
3 2
2
2
1 0
C C C
C C C
C
I I I
I
c
I
c
I c
lock lock
lock lock
A A
j
A A
j 1

(8.202)
8.6.6 The Blatz-Ko Hyperelastic Model
Porous polymers should be considered as compressible materials. Blatz-Ko(1962)
proposed the following strain energy density ) , , (
C C C
I I I I I I 1 based on experimental and
numerical results, with isochoric and volumetric parts:
( ) ( )
(
(

- -
|
|
.
|

\
|
- - -
(

- - -
- -
) 1 (
1
3
2
) 1 ( 1
2
3
2
) , , (
, ,
,
j
,
j j
1
C
C
C
C C C C C
I I I
I I I
I I
f I I I I f I I I I I I
(8.203)
where , is given in terms of the j (shear modulus) and v (Poissons ratio) by
v -
v

2 1
, ,
and [ ] 1 , 0 = f is an interpolation parameter. In the particular case in which 1 f , we obtain:
( ) ( ) 1
2
3
2
) , , ( - - -
-,
,
j j
1
C C C C C
I I I I I I I I I I
(8.204)
In the incompressibility case ( 1
C
I I I ), the equation in (8.203) becomes the Mooney-Rivlin
model, (see equation (8.194)). Then, in the restrictive case where 1 f , and with the
incompressibility condition ( 1
C
I I I ), we revert to the Neo-Hookean incompressible model
given in (8.181).
8.6.7 The Saint Venant-Kirchhoff Model
8.6.7.1 Strain Energy Density
In the Saint Venant-Kirchhoff model, the strain energy density ( ) , (
E E
I I I 1 ) is given by:
( )
E
2
E E E
I I I I I I j j A 1 2 2
2
1
) , ( - - (8.205)
where A and j are material constants.
We can express the strain energy density in terms of the C -invariants. To do so, let us
consider the relationships between the invariants of E and C , (see the equations in
(8.174)):
( ) ( ) ( ) 1
8
1
; 3 2
4
1
; 3
2
1
- - - - - - -
C C C E C C E C E
I I I I I I I I I I I I I I I I (8.206)
NOTES ON CONTINUUM MECHANICS

456
and by substituting them into the equation in (8.205) we obtain:
( ) ( ) ( )
( )( ) ( ) 3 2
2
3 2
8
1
3 2
4
1
2 3
2
1
2
2
1
) , (
2
2
- - - - - -
(

- - - -
(

- -
C C C
C C C C C
I I I I
I I I I I I I
j
j A
j j A 1

(8.207)
8.6.7.2 The Stress Tensor
The derivatives of the function (8.205) with respect to the invariants are:
( ) j
1
j A
1
2 ; 2 -
o
o
-
o
o
E
E
E
I I
I
I

(8.208)
and by using the equation in (8.79), where the parameters are:
( )
j
1 1
j j A
1 1 1
2
2 2
1
0
-
o
o
-
o
o

/ - -
o
o
-
o
o
-
o
o

E
E E
E E E E
E
E
E E
I
I I I I I
c
I I I I I
I I I
I
I I I
c
(8.209)
and by substituting the above parameters into the equation in (8.79), we obtain:
E
E
j A 2 - 1 S I (8.210)
Then we can conclude that the Saint_Venant-Kirchhoff model describes geometric
nonlinearity, but is also characterized by a material linearity, i.e. the stress-strain relationship
is linear.
Then, taking into account that ( ) 1 - C E
2
1
and ( ) 3
2
1
-
C E
I I , we can obtain the tensor
S by means of C as follows:
( ) ( ) ( ) C C E C
C C E
j j
A
j A j A -
(

- - - - - - 1 1 1 1 S S 3
2
2 3 2 ) (
2
1
2
1
I I I
(8.211)
Note that the above equation could have been obtained by means of the constitutive
equation in terms of S given in (8.67), i.e.:
( )( ) ( )
( ) C
C C
C C
C
C
C
C C
C
C
C C
C
C C
C
C C
j j
A
j j j
j A
1 1 1
1 1 1 1 1
-
(

- -
-
)
`

- - - - -
(
(

|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o

(
(

|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o

o
o

-
1
1 1
1 S
3
2
2
3
2 2
3 2
8
2
2 2
2
) (
2
1
I
I I
I I
I
I I I
I I I
I I I I I
I
I I I

(8.212)
8.6.7.3 The Elastic Tangent Stiffness Tensor
The elastic tangent stiffness tensor can be obtained as follows:
( ) ( )
E
E
E
E
E
E
E E
E
E E
o
o
-
o
o
-
o
o
= -
o
o

o
o
j A j A j A 2 2 2 1 1 1
S I
I I
tan
C (8.213)
8 HYPERELASTICITY

457
Note that E is a symmetric tensor, so the result of the operation
E
E
,
is also a symmetric
tensor. Then, the equation in (8.213) in indicial notation becomes:
( ) [ ]
kl
ji ij
ij
kl kl
ij
ij
kl kl
ij
tan
ijkl
E
E E
E
I
E
E
E
I
E o
- o
-
o
o

o
o
-
o
o

o
o

2
1
2 2 j c A j c A
E E
S
C
(8.214)
Thus,
[ ] - -
notation tensorial
jk il jl ik ij kl
tan
ijkl
c c c c j c c A C I 1 1 j A 2 -
tan
C
(8.215)
NOTE: Note that in a small deformation regime, the condition that all stress tensors are
equal is satisfied, i.e. S = , and the same is true for the strain tensors, = E , after which
the constitutive equation for stress in (8.210) becomes:
1

j A 2 - I (8.216)
which is the same constitutive equation for isotropic linear elastic materials as that obtained
in Chapter 7, (see also Problem 6.1). In addition, the elastic tangent stiffness tensor
tan
C
coincides with the elasticity tensor
e
C , (see Chapter 7).
8.6.8 The Compressible Neo-Hookean Material Model
8.6.8.1 Strain Energy Density
In the compressible Neo-Hookean material model, the Helmholtz free energy per unit
reference volume (strain energy density), (see Bonet&Wood (1997)), is defined by:
) ( ) ( ) 3 (
2
) (
2
) (
) ( ) (
2

C C
C
C I J I J J
I J
1 1
j
j
A
1
1 1
- - - -
_ _
ln ln
(8.217)
where A and j are material parameters.
8.6.8.2 The Stress Tensor
The second Piola-Kirchhoff stress tensor, (see Eq. (8.67)), is given by:
C
C C
C C C
C
C
C
C
C
C
C C
C C C
o
o
-
(
(

|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o
-
|
|
.
|

\
|
o
o

|
.
|

\
|
o
o
-
o
o

o
o

- -
) (
2 2
) ( ) , , (
2
) (
2
2 1
J
I I I
I I
I I I
I I I
I I
I I I
J I I I I I I
1 1 1 1 1
1 1 1
1
S

(8.218)
where
2
j 1

o
o
C
I
, 0
o
o
C
I I
1
, 0
o
o
C
I I I
1
and
1
2
1 1 1
) ( 2
2
) (

-
-
o
o

o
o
C
C C
J
J
J
J
J
J
j
A 1
ln . Moreover,
if we consider that
1
2
-

o
o
C
C
J J
, we can conclude that:
( )
1 1 1 1
) (
2
) (
2
2
2
2
- - - -
- -
|
|
.
|

\
|
- - |
.
|

\
|
C C C C J J ln ln A j
j A j
1 1 S
(8.219)
Then, by considering the relationship between the Cauchy stress tensor and the second
Piola-Kirchhoff stress tensor,
T
J F F S , as well as
T - - -
F F C
1 1
, we can obtain:
NOTES ON CONTINUUM MECHANICS

458
( ) [ ]
( ) [ ]
T T T T T
T T T
T T
J
J
J J
F F F F F F F F F F
F F F F F F
F C C F F F



- - - -
- - - -
- -
- -
- -
- -
1 1
1 1
1 1
) (
) (
) (


ln
ln
ln
A j j
A j
A j
1
1 S

(8.220)
Thus,
( ) [ ] 1 1 ) (
1
J
J
ln A j - - b (8.221)
8.6.8.3 The Elastic Tangent Stiffness Tensor
The elastic tangent stiffness tensor, (see Figure 8.3), can be defined as:
C C C o
o

o o
o

S
2

4
2
1
tan
C (8.222)
Then, given the second Piola-Kirchhoff stress tensor equation in (8.219), we obtain:
( ) [ ]
( )
(

o
o
-
o
o
-
o
o
-
o
o

- -
o
o

-
-
-
- -
C
C
C
C
C
C
C
C C
C
1
1
1
1 1
) (ln
) (
2
) ( 2


J
J
J
tan
A A j j
A j
ln
ln
1
1 C
(8.223)
In Chapter 1 we obtained ( ) [ ]
1 1 1 1
1
2
1
- - - -
-
- -
o
o
kr il lr ik
kl
ir
C C C C
C
C
, after which the above equation,
in indicial notation, becomes:
( ) ( )
1 1 1 1 1 1 1 1 1 1
1
1
1
) (
2
1
2
) ( 2
) (
2 2


- - - - - - - - - -
-
-
-
- - - -
o
o
-
o
o
-
o
o
-
kr il lr ik kl ir kr il lr ik
kl
ir
kl
ir
kl
ir tan
irkl
C C C C J C
J
J
C C C C C
C
C
J
C
J
C
C
C
ln
ln
ln
A A j
A A j C
(8.224)
and by simplifying we obtain:
[ ] ( )
1 1 1 1 1 1
) (
- - - - - -
- - -
kl ir kr il lr ik
tan
irkl
C C C C C C J A A j ln C
[ ] ( )
1 1 1 1 1 1
) (
- - - - - -
- - - C C C C C C A A j J
tan
ln C
(8.225)
We can also obtain the tensor L , (see Figure 8.3), which is related to the tensor C by the
equation in (8.32), i.e.:
[ ]( ) {
ds ct st pq sq pt tq ps ap bq
ds ct
tan
pqst ap bq abcd
F F C C C C C C J F F
F F F F
1 1 1 1 1 1
) (
- - - - - -
- - -

A A j ln
C L
(8.226)
Note that:
bc ad wc xd bw ax ds ct qw tw sx px ap bq ds ct tq ps ap bq
F F F F F F F F F F C C F F c c c c c c
- - - - - - 1 1 1 1 1 1
,
bd ac ds ct sq pt ap bq
F F C C F F c c
- - 1 1
,
cd ab ds ct st pq ap bq
F F C C F F c c
- - 1 1
.
Then:
[ ] ( )
cd ab bd ac bc ad abcd
J c c A c c c c A j - - - ) ( ln L
(8.227)
Note that ( )
abcd
sym
abcd bd ac bc ad
I 2 2 - I c c c c , where
sym
I = I is the symmetric fourth-order
unit tensor, with which we can obtain:
8 HYPERELASTICITY

459
[ ] [ ] 1 1 I - - / - - A A j c c A j ) ( 2 ) ( 2 J J
notation Tensorial
cd ab
sym
abcd abcd
ln ln L I L
(8.228)
Now, the tensor A (defined in (8.48)) becomes:
[ ]
1 1 I -
-

J J
J
J
A A j

) (
2
1 ln
L A
(8.229)
Then, if
[ ]
J
J ) ( ln A j
j
-
and
J
A
A , where j and A are the equivalent Lam
constants, it follows that
1 1 I - 2 A j A (8.230)
NOTE: In a small deformation regime, we have 1 = J , with which we obtain j j = and
A A = , and the following also holds:
e tan
C A L C - 1 1 I 2 A j .
Problem 8.1: Consider a motion characterized by dilatation. The continuum is made up of
an isotropic material resembling the compressible Neo-Hookean material model. Find the
Cauchy stress tensor in terms of the Jacobian determinant.
Solution:
In isotropic materials dilation can be characterized by
i i
dX dx / , (see Figure 8.7).














Figure 8.7: Dilatation.
In this scenario we have:
3
1
3
) ( J J / / = / F F det 1
The Cauchy stress tensor can be obtained by means of the equation in (8.221), i.e.:
( ) [ ] 1 1 ) (
1
J
J
ln A j - - b
where the left Cauchy-Green deformation tensor can be evaluated as follows:
1 1 1 1
3
2
2
J
T
/ / / F F b
Therefore, the Cauchy stress tensor becomes:
1 ) ( 1
3
2
(
(

-
|
|
.
|

\
|
- J
J
J
J
ln
A j


2
X

3
X

1
X

1
dX

3
dX

2
dX
dV

0
dV

1
dX /

3
dX /

2
dX /

(
(
(

(
(
(

/
/
/

(
(
(

3
2
1
3
2
1

0 0
0 0
0 0
dX
dX
dX
dx
dx
dx


Stretches
3 2 1
/ / / /
NOTES ON CONTINUUM MECHANICS

460
8.6.9 The Gent Model
Strain energy density in the Gent model, Gent(1996), is characterized by the following
logarithmic function:
|
|
.
|

\
| -
-
-

m
m
I
I
I
3
1 I
2
) (
C
C
ln
j
1
(8.231)
where j and
m
I are material constants, and
m
I is the constant that measures the limit
value of ( ) 3 -
C
I .
8.6.10 The Statistical Model
We can summarize this model as follows.
Strain Energy Density
Let 1
~
be the isochoric part and
vol
1 the volumetric part, then, the energy function is
given by:
) ( ) ( ~
~
C
C
I I I I
vol
1 1 1 -
(8.232)
where
( ) ( ) ( )
( ) ( )
(

(
- - - -

- - - - - -
243
673750
519
81
7000
19
27
1050
11
9
20
1
3
2
1
5
~
4
4
~
3
3
~
2
2
~ ~
~
C C
C C C
I
N
I
N
I
N
I
N
I j 1

(8.233)
( )
2

2
C
I I I
vol
ln
x
1 (8.234)
The Second Piola-Kirchhoff Stress Tensor
1
2 0
-
- C a a 1 S Stress Tensor Statistical Model (8.235)
where the coefficients
0
a ,
1
a and
2
a are given by:
3
4
~
4
3
~
3
2
~
2
~
0
1
134750
519
1750
19
350
11
10
1
2
1
2
C
C C C C
I I I
I
N
I
N
I
N
I
N
a |
.
|

\
|
- - - - - j
(8.236)
C
C
C C C C
I I I I
N
I
N
I
N
I
N
I a
ln
5
~
4
4
~
3
3
~
2
2
~ ~
2
134750
519
1750
19
350
11
10
1
2
1
3
2
x - |
.
|
-

\
|
- - - - -

j

(8.237)
The Elastic Tangent Stiffness Tensor
( ) ( )
1 1 1 1 2 1 1
5 4
1 1
2 1
2
) (
- - - - - - - -
- - - - - -
kj il lj ik kl ij kl ij kl ij kl ij
tan
ijkl
C C C C
a
C C b b C C b b c c c c C
(8.238)
where the parameters are:
|
|
.
|

\
|
|
.
|

\
|
- - - -
3 3
3
~
4
2
~
3
~
2
1
1 1
67375
1038
1750
57
175
11
10
1
2
C C
C C C
I I I I I I
I
N
I
N
I
N N
b j (8.239)
8 HYPERELASTICITY

461
3
4
~
4
3
~
3
2
~
2
~
2
1
26950
519
875
38
350
33
5
1
2
1
3
2
C
C C C C
I I I
I
N
I
N
I
N
I
N
b |
.
|

\
|
- - - - - - j
(8.240)
|
.
|

\
|
- - - - -
5
~
4
4
~
3
3
~
2
2
~ ~
4
26950
519
875
38
350
33
5
1
2
1
9
2
C C C C C
I
N
I
N
I
N
I
N
I b j (8.241)
2
5
x
b (8.242)
C
C
C C C C
I I I I
N
I
N
I
N
I
N
I a
ln
5
~
4
4
~
3
3
~
2
2
~ ~
2
134750
519
1750
19
350
11
10
1
2
1
3
2
x - |
.
|
-

\
|
- - - - -

j

(8.243)
The demonstration of the statistical model can be found in Chaves (2009), (see also
Sansour et al. (2003)).
8.6.11 The Eight-Parameter Model
We summarize this model as follows.
Strain Energy Density
) ( ) , ( ~ ~
~
C
C C
I I I I I I
vol
1 1 1 -
(8.244)
where
2
~ ~
8
4
~
7
2
~
6
3
~
5
~ ~
4
2
~
3
~
2
~
1
~ ~ ) , (
~
C C C C C C C C C C C C
I I I I I I I I I I I I I I I I I c c c c c c c c 1 - - - - - - - (8.245)
( )
2
9

C
I I I
vol
ln c 1
(8.246)
The parameters
1
c ,
2
c , ...,
8
c are the eight material parameters whereas
9
c represents
the bulk modulus. The values these parameters were determined by Sansour (1998) as:
5
8
7
7
3
6
4
5
3
4
2
3 2 1
10 5513 . 0 ; 10 432 . 0 ; 10 8439 . 0 ; 10 3473 . 0
10 3268 . 0 ; 10 1684 . 0 ; 0145 . 0 ; 1796 . 0
- - - -
- -
-
-
c c c c
c c c c

(8.247)
The Second Piola-Kirchhoff Stress Tensor
1
2 1 0
-
- - C C a a a 1 S Stress tensor Eight-parameter model (8.248)
where
(
)
C C C C C C
C C C C C
C
~
2
~
8
2
~
8
3
~
7
~ ~
6
2
~
5
2
~
4
~
4
~
3
~
2 1
3
0
2 4 2
3 2
2
I I I I I I I I I
I I I I I I
I I I
a
c c c c
c c c c c c
- - - -
- - - - - -

(8.249)
( )
3
2
~ ~
8
~
6
~
4 2 1
1
2 2 2
C
C C C C
I I I
I I I I I I a c c c c - - -
(8.250)
(
) ) ( 3 4 2
5 4 3 3 2
3
2

9
2
~
6
~
2
2
~ ~
8
4
~
7
3
~
5
~ ~
4
2
~
3
~
1 2
C
C C
C C C C C C C C
I I I I I I I
I I I I I I I I I I a
ln c c c
c c c c c c
- - -
- - - - - -

(8.251)
The Elastic Tangent Stiffness Tensor
NOTES ON CONTINUUM MECHANICS

462
( ) ( )
( ) ( ) ( )
1 1 1 1 2 1 1
8
1 1
5
1
4
1 1
2 1 0
2 2
- - - - - - - -
- -
- - - - - - -
- - - - -
kj il lj ik kl ij kl ij kl ij il jk jl ik
kl ij kl ij kl ij kl ij kl ij kl ij
tan
ijkl
C C C C
a
C C b C C C C b
a
C C b C C b C C b b
c c c c
c c c c c c C
(8.252)
where
(
)
3
~
8
~ ~
8
2
~
7
~
6
~
6
~
5
~
4 3 2
3
2
0
2 6 12
2 2 6 3 2
2
C C C C
2
C C C C
C
I I I I I
I I I I I
I I I
b
c c c
c c c c c c
- -
- - - - - -

(8.253)
( )
2
~
8
~
8
~
6 4 1
2 2 2
2
C C C
C
I I I I
I I I
b c c c c - - -
(8.254)
(
)
C C C C
C C C C C C C
C
~
2
~
8
2
~
8
3
~
7
~ ~
6
2
~
5
~
4
2
~
4
~
3
~
2 1
3
2
10 5 16
8 9 3 3 4 2
1
3
2
I I I I I I
I I I I I I I I I
I I I
b
c c c
c c c c c c c
- - -
- - - - - - - -

(8.255)
( )
C
C
~
8 6
3
4
4
2 2
2
I
I I I
b c c - , ( )
C C C C
C
~ ~
8
~
6
~
4 2
3
2
5
10 8 3 2
1
3
2
I I I I I I
I I I
b c c c c - - - -
(8.256)
(
)
9
2
~ ~
8
4
~
7
2
~
6
3
~
5
~ ~
4
2
~
3
~
2
~
1 8
9 25 16
16 9 9 4 4
9
2
c c c
c c c c c c
- -
- - - - - -
C C C
C C C C C C C
I I I I
I I I I I I I I I I b
(8.257)
The demonstration of the eight-parameter model can be found in Chaves (2009), (see also
Sansour et al. (2003)).
8.7 Anisotropic Hyperelasticity
Certain materials such as some biological tissues have fibers, and therefore lose their
isotropy. When these fibers are arranged according to a preferential direction,
0

a , we can
approach this material by means of the transversely isotropic material. In other material
such as heart tissue, the fibers are arranged according to two preferential directions, and are
classified as tissue with two families of fibers, (see Figure 8.8). In this subsection we just
describe the model for one family of fibers. Details about two families of fibers can be
found in Holzapfel (2000).








Figure 8.8: Materials with fibers.

0

a

0

a

0

b
b) One family of fibers b) Two families of fibers
8 HYPERELASTICITY

463
8.7.1 Transversely Isotropic Material
As discussed in Chapter 1, the scalar-valued tensor function ) (C 1 1 can be written in
terms of the principal invariants of C , i.e. ) , , (
C C C
I I I I I I 1 1 . Now, if the 1 -arguments
are C and the vector
0

a , i.e. )

, (
0
a C 1 , it can be shown that this function can be written in
terms of the following invariants:
) , , , , ( )

,

, , , ( )

, (
) 5 ( ) 4 (
0
2
0 0 0 0 C C C C C C C C
C C C I I I I I I I I I I I I I I 1 1 1 a a a a a (8.258)
where
) 4 (
C
I and
) 5 (
C
I are the pseudo-invariants of anisotropy. Moreover, if we consider that
the energy, if independent of the sense of the vector
0

a , fulfills that )

, ( )

, (
0 0
a a - C C 1 1 ,
we can then represent the strain energy function by:
)

, (
0 0
a a C 1 1 (8.259)
We can also show that the previous function is objective as follows:
)

, ( )

, ( )

, (
0 0 0 0 0 0
a Q a Q Q Q Q a a Q Q Q a a
T T T
C C C 1 1 1 (8.260)
As the rotated current configuration is defined by the transformation F F Q
*
, (see
Chapter 4), then C F F C
* * *
T
is fulfilled, and a vector in the rotated current
configuration is given by
0
*
0

a Q a , thus:
)

, ( )

, ( )

, (
*
0
*
0
*
0 0 0 0
a a a Q a Q Q Q a a C C C 1 1 1
T
(8.261)
which proves objectivity.
Then, since the material has lost its isotropy, we can not express the energy function just in
terms of the principal invariants. To use the energy function in terms of invariants, Spencer
(1984) obtained two pseudo-invariants of anisotropy, which contribute to the energy
function, (see Holzapfel (2000)), and are given by:
j pj ip i j ij i
C C I C I
I I
0 0
) 5 (
0 0
) 4 (
0
2
0
) 5 (
0 0
) 4 (

;


;

a a a a

C C
C C
C C a a a a

(8.262)
where
2
0 0
) 4 (
0


a
a a / C
C
I , and
0

a
/ is the stretch according to the
0

a -direction (fiber
stretching).
So, the energy function can be expressed as follows:
) , , , , (
) 5 ( ) 4 (
C C C C C
I I I I I I I I 1 1 (8.263)
Now, the derivatives of
) 4 (
C
I and
) 5 (
C
I with respect to the tensor C are given by:
( )
( )
l ik i lj j k jl pk j ip i pl ik pj j i
kl
pj
j ip i
kl
ip
pj j i j pj ip i
kl kl
l k jl ik j i
kl
ij
j i j ij i
kl kl
C C C C
C
C
C
C
C
C C C
C C
I
C
C
C
C C
I
0 0 0 0 0 0 0 0
0 0 0 0 0 0
) 5 (
0 0 0 0 0 0 0 0
) 4 (



a a a a a a a a
a a a a a a
a a a a a a a a
- -
o
o
-
o
o

o
o

o
o

o
o

o
o

o
o
c c c c
c c
C
C

(8.264)
and such equations in tensorial notation become:
( ) ( )
0 0 0 0
) 5 (
0 0
) 4 (

;

a a a a a a -
o
o

o
o
C C
C C
C C
I I

(8.265)
NOTES ON CONTINUUM MECHANICS

464
Then, the constitutive equation for stress (reference configuration) can be represented by:
|
|
.
|

\
|
o
o
o
o
-
o
o
o
o
-
o
o
o
o
-
o
o
o
o
-
o
o
o
o

o
o

C C C C C
C
C
C
C
C
C
C
C
C
C
C
C C C C C
) 5 ( ) 4 (
) 4 (
) 5 ( ) 4 (
2
) , , , , (
2
I
I
I
I
I I I
I I I
I I
I I
I
I
I I I I I I I I
1 1 1 1 1
1
S

(8.266)
Next, if we consider the derivatives in (8.264) and the equation in (8.67) we obtain:
( )
(
(

(
-
o
o
-
o
o

-
|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o
-
|
|
.
|

\
|
o
o


- -
0 0 0 0
) 5 (
0 0
) 4 (
2 1

( )

(

2
a a a a a a
1 S
C C
C C
C C
C
C
C
C
C
C C
I I
I I I
I I
I I I
I I I
I I
I I I
1 1
1 1 1 1

(8.267)
The second derivative of
) 4 (
C
I is:
0
o o
o
C C
C
) 4 ( 2
I

ijkl
kl ij
C C
I
0
o o
o
) 4 ( 2
C

(8.268)
Note that the second derivative of
) 5 (
C
I becomes a fourth-order tensor that features both
major and minor symmetry:
( ) 1 a a
o o
o
0 0
) 5 ( 2

C C
C
I
kl j i
kl ij
C C
I
c
0 0
) 5 ( 2

a a
o o
o
C

(8.269)
If we now consider the directions of the fibers in the current configuration (deformed
configuration), a

, and using the expression of the Kirchhoff stress tensor, (8.71), we


obtain:
( )
T
T
I I
I I I
I I
I I I
I I I
I I
I I I
J
F C C
C C F
F F
C C
C
C
C
C
C
C C


(
(

(
-
o
o
-
o
o
-

-
|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-
o
o
-
|
|
.
|

\
|
o
o


- -
0 0 0 0
) 5 (
0 0
) 4 (
2 1

( )

(

2

a a a a a a
1
S
1 1
1 1 1 1


(8.270)
Then, if we consider that F F C
T
,
T
F F b and a a
a

0

0
/ F , the above equation
then becomes:
( )
(
(

(
-
o
o
/ -
o
o
/ -

-
o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o

a a a a a a
1
a a

)

( )

(

2
) 5 (
2
) 4 (
2
2
0 0

b b
b b
b b
b
b b
b
b b
I I
I I I
I I I I I
I
I I I
J
1 1
1 1 1 1


(8.271)
Moreover, by considering that
2
0 0
) 4 (
0


a
a a / C
C
I , we obtain:
( )
(
(

(
-
o
o
-
o
o
-

-
o
o
-
o
o
-
|
|
.
|

\
|
o
o
-
o
o

a a a a a a
1

( )

(

2
) 5 (
) 4 (
) 4 (
) 4 (
2
b b
b b
b
C
b
C
b
b b
b
b b
I
I
I
I
I I I
I I I I I
I
I I I
J
1 1
1 1 1 1


(8.272)
9 Plasticity











9.1 Introduction
With elastic loading, atomic structures are not affected, which is typical of processes in
which there is no internal energy dissipation. Then, once the load has been removed, the
solid returns to its initial state. In certain types of materials, if we keep loading, a level will
be reached in which the atoms begins to restructure (dislocation at the atomic level), so, in
this way, we have internal energy dissipation (an irreversible process). Most of the
dissipated energy will be used to increase the temperature (heat release), and as a result
there will be an increase in system disorder, i.e. a rise in entropy. A rise in temperature also
involves the dilation phenomenon. At the macroscopic level, in ductile materials, this
atomic restructuring is characterized by permanent deformation (plastic strain). That is, if
the material which has been internally restructured is completely unloaded, it can be
observed that part of the total deformation is regained. This recoverable part is
characterized by the elastic strain, and the permanent deformation by the plastic strain, (see
Figure 9.1), and the constitutive models devised to represent this phenomenon are called
plasticity models or elastoplastic models.
It can be complex to formulate a constitutive model that considers all possible phenomena
that occur during the plasticity process. In general, a process which involves plastic strain
typically involves a large deformation, heat production, and the loss of material isotropy in
the plastic zone due to plastic fibers which are formed in this area. However, in certain
types of materials, the effect of temperature can be discarded (isothermal process), and they
can also be subjected to a deformation state in which the elastic strain is very small when
compared with the plastic strain (small deformation regime). These simplifications have given
rise to the classical Theory of Plasticity.


9
Plasticity
465 , Notes on Continuum Mechanics, Lecture Notes on Numerical
, Methods in Engineering and Sciences 4, DOI 10.1007/978-94-007-5986-2_
International Center for Numerical Methods in Engineering (CIMNE), 2013
E.W.V. Chaves
11
NOTES ON CONTINUUM MECHANICS

466



















Figure 9.1: Tensile testing Plastic behavior.
There were many researchers who endorsed this theory of plasticity, among whom we can
cite: Rankine(1851), Tresca(1864), von Mises(1913), Prandtl(1924), Reuss(1930),
Prager(1945), Hill(1950), Drucker(1950), Koiter(1953), Ziegler(1959), Naghdi(1960),
Mroz(1967).
From a kinematics standpoint, the plasticity theory has been developed considering:
Plasticity with small deformation (infinitesimal strain):
Without the effect of temperature (Classical theory of plasticity);
- With the effect of temperature (Thermoplasticity in infinitesimal strain).
Plasticity with large deformation (Finite strain):
Without the effect of temperature (Plasticity in finite strain);
- With the effect of temperature (Thermoplasticity in finite strain).
In this chapter our approach will be to use plasticity models taking into account the small
and large (finite) deformation regime, (see Figure 9.2), and we will omit the temperature
effect (isothermal process).
However, before formulating these models, we will introduce some concepts that will be
useful in the development of this chapter.



e
r
o
r

Y
o

p
r

p
r - plastic strain

e
r - elastic strain

Y
o

Y
o
I II
III

I
- elastic zone
II - plastification zone
III - complete unloading
II I
9 PLASTICITY

467




















Figure 9.2: Overview of solids mechanics.
9.2 The Yield Criterion
An important concept in the classical plasticity theory (rate independent) is the concept of
the yield surface, which defines a multiaxial stress state at the threshold of plastic strain. If
the current stress state lies within the yield surface, the corresponding mechanical change is
purely elastic. Plastic strain is only possible when the stress state is on the yield surface.
Now, firstly, we will explain what this initial yield surface (yield criterion) is, and then we
will establish how it evolves during plastification.
As discussed in subsection 6.5 in Chapter 6 related to the tensile testing, (see Figure 9.1),
certain materials can exhibit two zones: an elastic zone, in which the upper limit stress is
characterized by
Y
o ; and a plastic zone. Generally speaking in three dimensions cases, if
we include the six independent components of the stress tensor as independent coordinate
axes, the current stress state is defined by a point in the six-dimensional space (hyperspace).
So, we take the infinite possibilities there are for the stress state to trigger plastification at a
material point in order to define a hypersurface, which is known as the yield hypersurface.
Fundamentally, we can state that the yield surface separates the elastic and plastic domains.
Therefore, if the stress state is inside the region delimited by the yield surface, the
corresponding strain change is purely elastic.
Continuum Mechanics
Solids Fluids Multiphysics
Finite strain regime Small strain regime
Plasticity
Viscous models
Hyperelasticity
Hyperplasticity
Damage models, ...
Linear Elasticity
Constitutive Law
Kinematics
Structural Mechanics
NOTES ON CONTINUUM MECHANICS

468
Metaphorically speaking, we can state that the constitutive equation reflects the
personality of the material, i.e. each material (or kind of material) has its own yield
surface and hence the concept of yield criterion appears. That is, the yield criterion establishes
when certain materials start to plastify.
9.2.1 The Yield Surface for Anisotropic Materials
Let us consider a homogeneous material undergoing a purely mechanical process. Then,
the yield surface, or the yield function, can be described mathematically as follows:
0 ) , , , , , , ( ; 0 ) (
13 23 12 33 22 11
o o o o o o o F F
ij

(9.1)
which represent the hypersurface equation in the stress space (six dimensions), in which:
domain plastic
domain elastic
0 ) (
0 ) (
o
< o
ij
ij
F
F

(9.2)
Alternatively, the equation in (9.1) can be rewritten in terms of the eigenvalues (
a
o ) and
eigenvectors (
) (

a
n
,
) of , i.e.:
0 )

, , , (
) 3 ( ) 2 ( ) 1 (
3 2 1
o o o n n n
, , ,
F
(9.3)
As expected, determining the yield criterion for anisotropic material is rather difficult to
achieve. Moreover, because of anisotropy, we need to establish 21 independent constants
in the laboratory (mechanical properties) in order to fully describe elastic material behavior.
9.2.1.1 The Yield Surface Gradient
The yield surface gradient in the stress space is given as follows:
ij
ij
components
o o
o

o
o
= =
) ( ) (

n

F F
F n V
(9.4)
where n is a symmetric second-order tensor or the plastic flow tensor, with six independent
components as well as being a coaxial tensor with .
Then, the Frobenius norm of the tensor n is defined as follows:
n n n

: : F F F
F
F
V V V
V
V
with


(9.5)
which thus shows that the Frobenius norm ( n

) is unitary, i.e. 1

n n n : .
In isotropic materials the yield surface is only dependent on the -eigenvalues, hence, it
can be shown in the space defined by the principal stresses (three dimensions), so the
plastic flow becomes a vector, and because of this some researchers refer to n as the flow
plastic vector.
9.2.2 The Yield Surface for Isotropic Materials
As we saw in Chapter 1, the scalar-valued isotropic tensor function, ) ( F , can be
represented only by the eigenvalues of the arguments, or in terms of the principal
invariants of the argument . From the above, we can draw the conclusion that for
9 PLASTICITY

469
isotropic materials the beginning of plastification does not depend on the principal
directions (eigenvectors) of the Cauchy stress tensor, thus:
0 ) , , (
0 ) , , (
3 2 1

o o o

I I I I I I F
F
Yield surface for isotropic materials (9.6)
where

I ,

I I ,

I I I are the Cauchy stress tensor principal invariants, whose values are
given by ) (

Tr I , [ ] { ) ( ) (
2
1
2 2

Tr Tr - I I , ) (

det I I I . Note that the equations in


(9.6) represent the yield surface in the space defined by the Cauchy stress tensor
eigenvectors. As discussed in the chapter on stress, (see Problem 3.3), the principal
invariants are related to the invariants of the deviatoric part of as follows:
( )
2
2
3
3
1

I I I - - J ; ( )

I I I I I I I 27 9 2
27
1
3
3
- - J (9.7)
where 0
1

dev
I

J ,
dev dev
dev
I I

:
2
1
2
- J , ) (
3
dev
dev
I I I

det J , with which the


yield surface for isotropic materials can still be represented as:
0 ) , , (
3 2
J J

I F Yield surface for isotropic materials (9.8)


Now, in certain kinds of materials such as non-porous metals, experimental evidence
suggests that hydrostatic pressure does not influence the initiation of plastification, that is,
the yield surface is independent of the first principal invariant

I , so it is a function of the
deviatoric part of the stress tensor, s =
dev
:
0 ) , (
3 2
J J F
Yield surface for isotropic materials
independent of hydrostatic pressure
(9.9)
In this case, the yield surface is represented by a prism in the principal stress space, (see
Figure 9.3).
NOTE: Before starting to study the yield criteria, a review of subsection A.4 Graphical
Representation of the Spherical and Deviatoric Parts in Appendix A is recommended.












Figure 9.3: Yield surface for isotropic materials independent of pressure.

3
o

1
o

2
o
H
hydrostatic axis
F

n V =
NOTES ON CONTINUUM MECHANICS

470
In Appendix A we saw that a stress state in the principal stress space (Haigh-Westergaard
stress space) can be represented by means of three variables ( p , q , 0 ) called the Haigh-
Westergaard coordinates, (see Figure 9.4). Thus, the yield surface for isotropic materials
can also be shown as:
0 ) , , ( 0 q p F
Yield surface for isotropic materials (9.10)
and for materials independent of pressure as:
0 ) , ( 0 q F

Yield surface for isotropic materials
independent of pressure
(9.11)


Figure 9.4: Haigh-Westergaard stress space.
By referring to (9.11), it is
sufficient to represent the yield
surface projected onto H -
plane (deviatoric plane that
passes through the origin, also
called Nadais plane), (see
Appendix A). The curve in the
deviatoric plane is called the
yield curve. As illustrated in
Figure 9.5, the yield curve has
triple symmetry and because of
this one need only analyze the
sector
3
0
r
s s 0 .


Figure 9.5: Yield surface projected onto a H -plane.

3
o
2
o

1
o

3 2
o o

2 1
o o

1 3 2
o > o > o

1 2 3
o > o > o

3 2 1
o > o > o

2 3 1
o > o > o

2 1 3
o > o > o
3 1 2
o > o > o
H
3 1
o o
0
[ ]
3
1
3
1
3
1
=
i
n

2
o

3
o

1
o
O
hydrostatic axis
( )
m m m
A o o o , ,
) , , (
3 2 1
o o o P

3 2 1
o o o
H
n
p
q
H
Deviatoric plane
(octahedral plane)
o
o
o

3

e

1

e

2

e
Deviatoric H -plane
(Nadais plane)
9 PLASTICITY

471
9.2.3 The Yield Surface for Materials Independent of
Pressure
In the previous section we saw that, generally speaking, for materials in which the yield
surface is independent of the hydrostatic pressure, the former has a prismatic shape in
which the prismatic and hydrostatic axes coincide. The yield surface cross section is then
defined depending on the model adopted. In this subsection, we will introduce some
models developed to represent material plastification where the hydrostatic pressure has no
influence on plastification, (see Chen&Han(1988)).
9.2.3.1 The von Mises Yield Criterion
The yield criterion of von Mises (1913) assumes that plastification occurs when the second
invariant of the deviatoric tensor,
2
J , reaches a critical value
2
k . We can define this
plastification criterion as:
domain plastic
domain elastic
0
0
2
2
2
2
-
< -
k
k
J
J

(9.12)
where k is a material property (yield stress in pure shear). Remember that in Chapter 1 (see
also Problem 3.3), the second invariant
2
J can be written in terms of the Cauchy stress
tensor components. Thus, the yield surface for von Mises criterion can be written as:
[ ]
2
2
2
13
2
23
2
12
2
33 11
2
33 22
2
22 11
) ( ) ( ) (
6
1
k o - o - o - o - o - o - o - o - o
_
J

(9.13)
or even in terms of the principal stresses:
[ ]
( )
2
3 1 3 2 2 1
2
3
2
2
2
1
2 2
3 1
2
3 2
2
2 1
3
1
) ( ) ( ) (
6
1
k
k
o o - o o - o o - o - o - o
o - o - o - o - o - o

(9.14)
The von Mises surface is defined by a cylinder, (see Figure 9.6), which prismatic axis is
parallel to hydrostatic axis, and does not depend on

I nor 0 . We can obtain the cylinder


radius by k 2 2
2
J q r .










Figure 9.6: Yield surface for von Mises yield criterion (independent of pressure).

3
o

1
o

2
o
H
k 2 r q

3
o
2
o

1
o
von Mises
H
NOTES ON CONTINUUM MECHANICS

472
The parameter k is obtained by means of tensile testing, where it holds that
Y
o o
1
and
0
3 2
o o . Under these conditions the equation in (9.14) becomes:
3
3
2 2 Y
Y
o
o k k
(9.15)
The strain energy density (the energy per unit volume) is given by:

vol dev dev dev
pp kk
dev
ij
dev
ij
pp ii kk
dev
ii kk pp
dev
ii
dev
ij
dev
ij
ij pp ij kk
dev
ij ij kk ij pp
dev
ij
dev
ij
dev
ij
ij pp
dev
ij ij kk
dev
ij
ij ij
1 1
c
c c c c
c c
1
- - r o - r o
r o - r o - r o - r o
r o - r o - r o - r o
|
.
|

\
|
r - r |
.
|

\
|
o - o
r o

) ( ) (
6
1
2
1
6
1
2
1
18
1
6
1
6
1
2
1
18
1
6
1
6
1
2
1
3
1
3
1
2
1
2
1
3 0 0
Tr Tr :

(9.16)
If we now consider the generalized Hookes law, 1 j A 2 ) ( - Tr (isotropic material),
(see Chapter 7), and its deviatoric part,
dev dev dev dev
1 j j A 2 2 ) ( - Tr , (
dev
ij
dev
ij
r o j 2 )
(where it was considered that the trace of any deviatoric tensor is zero) then, the part of the
strain energy density associated with the deviatoric part is given by:
2
2
1
4
1
2
1
J
j j
1 o o r o
dev
ij
dev
ij
dev
ij
dev
ij
dev

(9.17)
Thus, it is possible to interpret the von Mises yield criterion as: plastification begins when
the energy
dev
1 reaches the critical value
j j j 6 2 2
2 2
2 Y
o

k J
.
Another interpretation of the von Mises yield criterion is related to the octahedral shear
stress:
[ ]
[ ]
2
2
2
3 1
2
3 2
2
2 1 2
2
3 1
2
3 2
2
2 1
2
3
2
) ( ) ( ) (
6
1
) ( ) ( ) (
9
1
J
J
t

o - o - o - o - o - o
o - o - o - o - o - o t
oct
oct

(9.18)
whose equations were obtained in Appendix A. Then, by applying the von Mises yield
criterion (9.12), 0
2
2
- k J , we obtain:
k k J
2

3
2
3
2
3
2
2
2
t t
oct oct
(9.19)
This criterion can also be interpreted as that in which material starts to plastify when the
octahedral shear stress reaches the critical value 2
3 3
2
Y
o
k .
Now, by considering the equations in (9.14) and (9.15) we can still write the von Mises
yield criterion as:
2 2
3 2 3 2 3 1
2
3 1
) ( ) )( ( ) (
Y
o o - o - o - o o - o - o - o (9.20)
9 PLASTICITY

473
which shows the ellipse equation in the coordinate system ) (
3 1
o - o - ) (
3 2
o - o , whose
ellipse axis form a 45 angle to the axes ) (
3 1
o - o and ) (
3 2
o - o , (see Figure 9.7).








Figure 9.7: The von Mises yield criterion.
For the state of plane stress ) 0 (
3
o , the yield criterion in (9.20) becomes:
2 2 2
2 2 1
2
1
3
Y
o o - o o - o k (9.21)
which represents an ellipse in the
2 1
o - o -space, (see Figure 9.8), and in uniaxial cases the
yield surface is reduced to a point.










Figure 9.8: The von Mises yield criterion the state of plane stress.
The von Mises yield criterion can also be expressed in terms of the Frobenius norm of the
Cauchy deviatoric stress tensor ( s =
dev
), which is given by:
( )
2 3 1 3 2 2 1
2
3
2
2
2
1
2
3
2
J s s o o - o o - o o - o - o - o
ij ij
s s s : (9.22)
Then, by considering the equations (9.14), (9.15) and (9.22), the von Mises yield criterion
can be expressed as follows:
0
2
3
0 2
2
3
2
o - o -
Y Y
s J (9.23)
We can then summarize the different ways of expressing the yield surface for the von
Mises yield criterion:

Y
o


1
o

2
o

Y
o

uniaxial
pure shear

3
Y
o


3
Y
o
-


Y
o
) (
3 1
o - o
) (
3 2
o - o
45
NOTES ON CONTINUUM MECHANICS

474
0
2
3
) , (
0
2
3
) , (
0 3 ) , (
0
3
2
) , (
0 ) , (
2 2
2
2 2
o - o
o - t o t
o - o
- t t
-
Y Y
Y oct Y oct
Y Y
oct oct
s s F
F
F
F
F
J J
k
J J
k
k k
Yield surface for von Mises yield
criterion
(9.24)
Next, we will calculate the yield surface gradient for the von Mises criterion in the principal
stress space. Starting with the definition

n

o
o
= =
F
F V and by using the function
0 3
2
o -
Y
J F , we obtain:
( )
ij ij ij
Y
ij
ij
o o
o

o o
o

o o
o - o

o o
o

-
) 3 (
3
2
1
) 3 ( ) 3 (
2
2
1
2
2 2
J
J
J J
n
F
(9.25)
Then, if we consider that ( )
3 1 3 2 2 1
2
3
2
2
2
1 2
3
1
o o - o o - o o - o - o - o J , and because we are
working in the principal stress space (
1 11
o o ,
2 22
o o ,
3 33
o o , 0
23 13 12
o o o ) we
can obtain:
ij
ij
s
J

(
(
(
(
(
(

o - o - o
o - o - o
o - o - o

o o
o
3
2
0 0
0
3
2
0
0 0
3
2
) (
2 1 3
3 1 2
3 2 1
2
(9.26)
Now, returning to the equation in (9.25), we can conclude that:
pq pq
ij ij
ij
ij ij
ij
s s
s
J
s
s
J
J
J n
2
3
2
2
3
3
1
2
3 ) 3 (
) 3 (
2
1
2 2
2
2
1
2

o o
o

o o
o

-
F

(9.27)
In tensorial notation the above equation becomes:
s
s
s
s s
s

2
3

2
3
2
3

o
o
= =
:
F
F V
(9.28)
where 1

s s s : holds. The same result as in (9.28) could have been obtained starting
from the yield surface 0
2
3
) , ( o - o
Y Y
s s F , i.e.:
( ) ( ) ( ) ( )

s
s
s
n
s
s
s s
s
o
o

o o
o

o o
o
o
o

o o
o

o o
|
|
.
|

\
|
o - o

o o
o

:
2
3
2
3
2
3
2
3
2
3
kl
ij
kl
ij
ij ij ij
Y
ij
ij
s s
s
n
F

(9.29)
9 PLASTICITY

475
It can be shown that s

s
s
o
o
: , (see Problem 1.39), then it follows that
s
s
n
2
3
and
the reader will be left to verify that the tensors and F

n V = are coaxial, i.e.




F F V V . In other words, they have the same principal directions (eigenvectors).
9.2.3.2 The Tresca Yield Criterion
In the Tresca yield criterion (1864) (also called the criterion of maximum shear stress) plastification
of the material begins when the maximum shear stress reaches the critical value
T
k .
Mathematically, this criterion is represented by:
domain plastic
domain elastic
T max
T max
k
k
t
< t
(9.30)
or more explicitly by:
T
3 1 3 2 2 1
2
,
2
,
2
k
|
|
.
|

\
| o - o o - o o - o
max
(9.31)
Then, by considering the principal stresses such as
III II I
o > o > o , the Tresca yield
criterion is given by:
( ) ( ) 0
2
1
) , , (
2
1
T T T
- o - o o o o - o k k k
III I III I III I
F (9.32)
where we have considered that ( ) 0 > o - o
III I
. Note that the principal stress
II
o has no
influence on the Tresca criterion. Then, the material constant
T
k is evaluated by tensile
testing, where it holds that
Y I
o o , 0 o
III
, thus:
2
T 1
Y
Y
o
o o k (9.33)
Figure 9.9 represents the stress state, at a material point, by means of a Mohrs circle in
stress. In this figure we can appreciate the elastic stress state (before plastification), and we
can verify the evolution of the stress sates from elastic to the initiation of plastification, (see
Figure 9.9).








Figure 9.9: The Tresca yield criterion.
To obtain the Tresca yield surface shape, let us consider the three principal stresses (
1
o ,
2
o ,
3
o ), and according to the Tresca criterion (9.31) we can define the following
equations:
Elastic stress state
Stress state at the beginning of
the plastification

max
t
t

N
o
Plastification zone

T
k
Evolution

max
t

I
o

II
o

III
o

I
o

II
o

III
o

NOTES ON CONTINUUM MECHANICS

476
Y Y Y
Y Y Y
o - o - o o - o - o o - o - o
o o - o o o - o o o - o
1 3 3 2 2 1
1 3 3 2 2 1
; ;
; ;

(9.34)
where we have considered the equation in (9.33). Note that each one represents a plane
equation which is parallel to the hydrostatic axis. The surface generated by putting these
planes together is a prism whose cross-section is defined by a regular hexagon, (see Figure
9.10). As we can verify, the cross section of the prism is the same in shape and size for any
point on the hydrostatic axis.











Figure 9.10: Tresca yield surface.
This cross section could also have been obtained by using the expressions of the principal
stresses obtained in subsection A.4 in Appendix A:
(
(
(

-
- -
(
(
(

o
o
o

(
(
(

o
o
o
r
r
) cos( 0 0
0 ) cos( 0
0 0 cos
3
2
0 0
0 0
0 0
0 0
0 0
0 0
3
2
3
2
2
3
2
1
0
0
0
J
m
m
m

(9.35)
Then, by substituting the stresses
1
o and
3
o from then above equation into the yield
criterion in (9.32) we obtain:
T
3
2
2 2
) cos(
3
2
cos
3
2
2
1
k
|
|
.
|

\
|
- - o -
|
|
.
|

\
|
- o
r
0 0 J J
m m

(9.36)
[ ]
T
3
2
2
) cos( cos
3
1
k - -
r
0 0 J
(9.37)
Then, if we consider the equation [ ] ) sin( ) cos( cos
3
1
3 3
2 r r
- - - 0 0 0 , we can still state that:
T
3
T
3
2
) sin(
2
1
) sin( k k - -
r r
0 0 q J
(9.38)
Let us also remember that in Appendix A the equation
2
2J q was obtained. Then, by
varying the angle 0 from 0 to 60 , we can obtain the yield curve in the deviatoric plane,
(see Figure 9.11). Thus, we can obtain the yield surface in terms of the parameters q and
0 :
) 0 ( 0 2 ) sin( 2 ) , (
3
T
3
r r
s s - - 0 0 0 k q q F (9.39)
von Mises
Tresca

3
o

1
o

2
o

3
o
2
o

1
o
H
H
hydrostatic axis

3 2 1
o > o > o
a) Tresca yield surface.
b) Projection of the yield surface onto
H -plane.

1
o
9 PLASTICITY

477
As expected, we can see that the yield surface is not a function of the hydrostatic pressure
p .











Figure 9.11: Tresca yield surface Projection of the surface onto the deviatoric plane.
As for the state of plane stress, ) 0 (
3
o , the equations in (9.34) become:
Y Y Y
Y Y Y
o - o - o - o o - o - o
o o - o o o o - o
1 2 2 1
1 2 2 1
; ;
; ;

(9.40)
which is represented by a hexagon in the principal stress space
2 1
o - o , (see Figure 9.12).

o
o
o - o
T
T
T
2
2
2
k
k
k
II
I
II I

(9.41)
In Figure 9.13 we can see the von Mises and Tresca criteria for the state of plane stress
( 0
3
o ), which show that the Tresca hexagon is inscribed into the ellipse of von Mises.
The yield surface gradient for the Tresca yield criterion can be obtained by means of the
definition

n

o
o
= =
F
F V . Then, we can use the definition of the Tresca yield surface
given in (9.32), to obtain:
( ) ( ) 0
2
1
) , , (
2
1
T T T
- o - o o o o - o k k k
III I III I III I
F (9.42)
( )
( ) [ ]
(
(
(

o o
o - o o

o o
(

- o - o o

o o
o

-
2
1
2
1
T
0 0
0 0 0
0 0
2
1 2
1
ij
III I
ij
III I
ij
ij
k
F
n
(9.43)
9.2.4 The Yield Criteria for Pressure-Dependent Materials
Porous materials, e.g. soil, rock, concrete, and some porous metals, are affected by
hydrostatic pressure, i.e. these materials are dependent on the first invariant

I , (see
Chen&Han(1988)). Here, can mention some criteria that take into account this
phenomenon, namely: the Mohr-Coulomb criterion, the Drucker-Prager criterion and the
Rankine criterion, among others.

2
o

1
o

3
o
H
0 0 ,
T
3
2
2 k q

3
r
0 ,
T
3
2
2 k q

6
r
0 ,
T
2 k q
NOTES ON CONTINUUM MECHANICS

478










Figure 9.12: Tresca yield curve Plane stress.











Figure 9.13: The yield curves for von Mises and de Tresca Plane stress.
9.2.4.1 The Mohr-Coulomb Criterion
This criterion was formulated by Coulomb in 1773, and was enhanced by Mohr in 1882,
Oller (2001). Mathematically, the Mohr-Coulomb criterion is given by:
) , , , (
2 1
k k
i
o t t (9.44)
where ) , , (
2 1
k k are material constants and the function ) (
i
o t is obtained by means of
laboratory experiments; (see the Triaxial Compression Test in subsection 6.5.12 in Chapter 6).
The function ) (
i
o t corresponds to an envelope curve of the Mohrs circles at the failure
moment, where each Mohrs circle is obtained for a different hydrostatic pressure state.
When the envelope is a straight line, (see Figure 9.14), the Mohr-Coulomb criterion
becomes:
c tg
N
c o - t (9.45)
where t is the shear stress magnitude in the failure plane,
N
o is the normal stress in the
failure plane, c is the cohesion (material property) and c is the angle of internal friction
(material property). In the particular case when 0 c we revert to the Tresca yield
criterion, with c t , and where the cohesion is interpreted as
T
k c .

Y
o

Y
o

1
o

2
o

Y
o -

Y
o -
Tresca yield curve
von Mises yield curve

Y
o

Y
o

1
o

2
o

Y
o -

Y
o o
2


Y
o o
1


Y
o - o
2


Y
o -

Y
o o - o
2 1


Y
o - o - o
2 1


T
2k o
Y

9 PLASTICITY

479









Figure 9.14: Varying the hydrostatic pressure so as to define the Mohr-Coulomb criterion.
Mohr proved by means of a graph that the equation in (9.45) represents a straight line
which is tangent to the greatest circle defined by
I
o and
III
o , (see Figure 9.15). Moreover,
we can also observe that this criterion is independent of the principal stress
II
o .









Figure 9.15: Mohr-Coulomb yield criterion.
Now, to obtain the mathematical equation that represents the Mohr-Coulomb yield
surface, let us consider Figure 9.16 in which we can obtain:
) 2 sin(
2
; ) 2 cos(
2 2
c c
III I A III I III I A
N
o - o
t
o - o
-
o - o
o (9.46)
Then, by substituting the equations in (9.46) into (9.45) we obtain:
( ) ( ) ( )
( ) ( ) ( )
( )( ) ( )
( ) ( ) c c
c c c c
c
c
c
c
c
c
c c c c
c c c
c
sin cos 2
sin cos 2 sin cos
cos
sin
sin
cos
sin
2 cos
)
2
cos( 2 )
2
sin(
) 2 cos(
2 2
) 2 sin(
2
2 2
III I III I
III I III I
III I III I III I
III I III I III I
III I III I III I
N
c
c
c
c
c
c
o - o - o - o
o - o - - o - o
o - o - o - o - o - o
-
r
o - o o - o - -
r
o - o
(

o - o
-
o - o
-
o - o
o - t
- tg tg
tg
tg

(9.47)

I
o
II
o
III
o
I
o
II
o
III
o
t

N
o
c
c
c - cohesion
c - angle of internal friction
c tg
N
c o - t
t

N
o
c

I
o
II
o
III
o
plastification zone
c
c tg
N
c o - t
Stress state at the
beginning of
plastification
c - cohesion
c - angle of internal friction
NOTES ON CONTINUUM MECHANICS

480
where we have considered that c c c c -
r
r -
r
-
2
2
2
2 .












Figure 9.16: Mohr-Coulomb criterion.
Thus, the yield surface is defined as follows:
( ) ( ) 0 sin cos 2 ) , , ( o - o - - o - o c c c
III I III I
c c F
Mohr-Coulomb yield
surface
(9.48)
Then, if we consider the values of
I
o and
III
o given in (9.35), the equation in (9.47)
becomes:
( ) [ ] ( ) [ ] c c 0 0 c 0 0 cos sin cos cos
3
sin cos cos
3
1
3
2
2
3
2
2
c
m
- - - o - - -
r r
J
J (9.49)
Next, by considering that ( ) [ ] ( ) 0 0 0 - - -
r r
3 3
2
sin cos cos
3
1
and
( ) [ ] ( ) 0 0 0 - - -
r r
3 3
2
cos cos cos , we can obtain:
( ) ( ) c c 0 c 0 cos sin cos
3
sin sin
3
2
3
2
c
m
- - o - -
r r
J
J (9.50)
Also by considering that
2
2J q and
m
p o 3 , the equation in (9.49) can still be
expressed as follows:
( ) ( ) c c 0 c 0 cos 6 sin cos sin 2 sin 3
3 3
c q p q - - - -
r r
(9.51)
thereby obtaining the yield surface on the deviatoric plane:
( ) ( ) 0 cos 6 sin cos sin 2 sin 3 ) , , , , (
3 3
- - - - -
r r
c c 0 c 0 c 0 c q p q q p c F
(9.52)
From the above we can verify that for a constant angle 0 the equation is linear with q and
p . Then, when 0 q we have:
c
c
c
c c cotg 3
sin
cos
2
6
cos 6 sin 2 c p c p c p (9.53)

N
o
c

I
o
III
o
c tg o - t c
c 2
t = o
S

R

2
2
III I
III I
C
R
o - o

o - o


C

A
N
o
A

A
t
9 PLASTICITY

481
Note that when 0 0 we obtain:
( ) ( )
( ) ( )
c c c
c c c
c c 0 c 0
cos 6 sin
2
1
sin 2
2
3
3
cos 6 sin cos sin 2 sin 3
cos 6 sin cos sin 2 sin 3
3 3
3 3
c q p q
c q p q
c q p q
- -
- -
- - - -
r r
r r

(9.54)
This line intercepts the q -axis with 0 p , thus:
( ) c
c
c c
sin 3
cos 6 2
cos 6 sin
2
1
2
3
3
-
-
c
q c q q (9.55)
When
3
r
0 , the equation in (9.51) becomes:
( ) ( )
c c c
c c c
cos 6 sin
2
1
sin 2
2
3
3
cos 6 sin cos sin 2 sin 3
3
2
3
2
c q p q
c q p q

-
- -
- -
r r
(9.56)
and this line intercepts the q -axis when 0 p , thus:
( ) c
c
sin 3
cos 6 2
-

c
q (9.57)
In this way we can draw a graph q p - , i.e. to show how pressure varies with the deviatoric
part, (see Figure 9.17).










Figure 9.17: Pressure vs. deviatoric part Mohr-Coulomb yield criterion.
In the principal stress space, the yield surface is represented by a conical surface whose
cross section is shown by an irregular hexagon, (see Figure 9.18).
Next, we calculate the gradient of the Mohr-Coulomb yield surface in the principal stress
space by means of the definition

n

o
o
= =
F
F V . Then, if we consider the expression of
) , , ( c c F given in (9.48), and that c sin 1-
o o
o
I
F
and c sin 1- -
o o
o
III
F
, we can obtain:
p
q
c cotg 3 c p

( ) c
c
sin 3
cos 6 2
-

c
q

( ) c
c
sin 3
cos 6 2
-

c
q
0 0

3
r
0
NOTES ON CONTINUUM MECHANICS

482
(
(
(

- -
-

o o
o

c
c
sin 1 0 0
0 0 0
0 0 sin 1
ij
ij
F
n
(9.58)
The norm of n is given by ( ) ( ) ( ) c c c
2 2 2
sin 1 2 sin 1 sin 1 - - - - - n n n : , with
which we can conclude that:
( )
( )
(
(
(
(
(
(

-
- -
-
-

o o
o
o o
o

c
c
c
c
2
2
sin 1 2
sin 1
0 0
0 0 0
0 0
sin 1 2
sin 1

ij
ij
ij
F
F
n
n
n
n
(9.59)










Figure 9.18: The Mohr-Coulomb yield surface.
9.2.4.2 The Drucker-Prager Yield Criterion
The Drucker-Prager yield criterion modifies that of von Mises, 0 2 ) , ( - k k q q F , by
adding the effect of pressure. In order to do so, we will start from the result obtained in the
Mohr-Coulomb graph made up of the axes q p - when 0 0 , (see Figure 9.17).









Figure 9.19: Pressure vs. deviatoric part Drucker-Prager criterion.
The line equation depicted in Figure 9.19 is given by:
p
q
c cotg 3 c p

( ) c
c
sin 3
cos 6 2
-

c
q
0
2 k q (von Mises)

2
o
3
o

1
o

3
o

1
o

2
o
c cotg 3 c
c
9 PLASTICITY

483
0 2 2 - - - 0 0 tg tg p q p q k k (9.60)
where:
( )
6 6
) sin 3 ( 3
sin 2
) sin 3 (
sin 2 2
3
sin 3
cos 6 2
c
c
c
c
c c
c
0
c

_
cotg
tg
c
c

(9.61)
From the graph with the q p - axes described in Figure 9.19 we can still obtain the
following equations:
( )
( )
( )
( )

c
c
c
c
c
c
c
c
cos 6 2
sin 3 2
sin 3 3
cos 6
sin 3
cos 3 2
sin 3
cos 6 2
2
k
k
k
c
c c
c

(9.62)
Then, by substituting c given above into the equation of p in Figure 9.19 we can obtain:
( ) ( )
c
c
c
c
c
c
c
c
c
3 3
sin 2
sin 3 3
sin
cos
cos 6 2
sin 3 2
3
sin
cos
3
k k k

-
c p
(9.63)
Then, with the above results we can represent the Drucker-Prager yield criterion as:
) sin 3 ( 3
sin 2
0 6 2 ) , , , (
c
c
c c c
-
- - with p q q p k k F
(9.64)
Now, if we consider that
oct
q t 3 2
2
J ,
oct
I
p o 3
3
3

, (see Appendix A), the
yield surface can still be represented as 0
3
3
6 2 2 ) , , , (
2
- -

I
q p c c k k J F , thus:
0 ) , , , (
2 2
- -

I I c c k k J J F (9.65)
We can still express the Drucker-Prager yield surface in terms of the octahedral stresses, i.e.
by means of
3

I
oct
o (normal octahedral stress) and
2
3
2
J t
oct
(tangential octahedral
stress), (see Appendix A). Thus, the equation in (9.65) can be rewritten as:
0 3
2
3
) , , , ( o - - t t o
oct oct oct oct
c c k k F (9.66)
Then, if we consider the expression of k given in (9.62), the Drucker-Prager criterion
becomes:
0 ) , , (
2 2
- - , c c J J

I I F (9.67)
where
( ) c
c
c
sin 3 3
sin 2
-
,
( ) c
c
,
sin 3 3
cos 6
-

c
,
3 2 1
o - o - o

I , c is the cohesion, and c is


the angle of internal friction.
Note that on the yield curve the edge of the Mohr-Coulomb yield curve coincides with the
Drucker-Prager yield curve when 0 0 . The yield surface is not longer a cylinder and
becomes a cone, (see Figure 9.20).

NOTES ON CONTINUUM MECHANICS

484










Figure 9.20: Drucker-Prager yield surface.
Then, in summary the different ways of expressing the Drucker-Prager yield surface are:
0 ) , , (
0 3
2
3
) , , , (
0 ) , , , (
0 6 2 ) , , , (
2 2
2 2


- -
o - - t t o
- -
- -
, c c
c c
c c
c c
J J
J J


I I
I I
p q q p
oct oct oct oct
F
F
F
F
k k
k k
k k
Drucker-Prager yield surface (9.68)
The definition

n

o
o
= =
F
F V give us the yield surface gradient in the stress space, so, for
the Drucker-Prager yield surface, 0 ) , , , (
2 2
- -

I I c c k k J J F , n becomes:
1 s

n
2
2
1
c -
o
o

J
F

(9.69)
where we have considered the equations
( )
s

2
2
2
1
J
J

o
o
and 1

o
oI
.
The Alternative Drucker-Prager Yield Criterion
Note that according to Figure 9.21 and by using the equations obtained previously,
6 c 0 tg ,
oct
q t 3 2
2
J ,
oct m
I
p o o 3
3
3
3

, we can obtain:
K
p
q q
oct
oct

o
t -

3
3 2 k
0 tg (9.70)
whose material parameters are 0 and k . Then, given any stress state, if the point is on the
line, then the material begins plastification. We can now use the definition in (9.70) to find
the norm (measurement of distance in the stress space):
( )
oct oct
K q t - o 3 ) ( (9.71)


3
o

1
o

2
o
Drucker-
Prager

3
o
2
o

1
o
Mohr-
Coulomb
c cot 3 c
c
H
9 PLASTICITY

485









Figure 9.21: Pressure vs. deviatoric part the Drucker-Prager criterion.
To obtain the parameters ( k , 0 ) two different compression tests must be carried out, that
is, a one-dimensional test 1D (
3 2 1
, 0 , 0 o o o ), and a two-dimensional one 2D
(
3 2 1
, 0 o o o ). Both of these must reach the nonlinearity limit. Then, if
-
D
f
1
0
and
-
D
f
2
0

represent the maximum values of the elastic stress (
3
o ) for 1D and 2D respectively, we
obtain:
Test 1D
- -
- t o
D
D
oct
D
D
oct
f f
1
0 2
1
1
0
1
3
2
3
2
;
3
1
) (
3
1
J Tr (9.72)
Test 2D
- -
- t o
D
D
oct
D
D
oct
f f
2
0
2
2
0
2
3
2
;
3
1
) (
3
2
Tr
(9.73)
Then, by using the equation in (9.70) we can obtain:
D
oct
D
oct
D
oct
D
oct
2
2
1
1
3
3 2
3
3 2
o
t -

o
t -

k k
0 tg (9.74)
Moreover, by rearranging the above we have:
( ) ( )
D
oct
D
oct
D
oct
D
oct
D
oct
D
oct D
oct
D
oct
D
oct
D
oct
1 2
1 2 2 1
2 1 1 2
3
2
3 2 3 3 2 3
o - o
o t - o t
t - o t - o k k k
(9.75)
and by substituting the equations (9.72) and (9.73) into the above we obtain:
- -
- -
-

D D
D D
f f
f f
2
0
1
0
2
0
1
0
2 3
1
k (9.76)
The constant K is then defined as follows:
-
-
- -
- -
|
|
.
|

\
|
-
-

o
t -

D
D
D D
D D
D
oct
D
oct
f
f
f f
f f
K
1
0
1
0
2
0
1
0
2
0
1
0
1
1
3
1
2 3
2
3
3 2 k

(9.77)
which after simplifying becomes:
q
p
p
q

3 c
k
p
2 k q
0
NOTES ON CONTINUUM MECHANICS

486
( )
- -
- -
- - -
- - - - - -
-
-

-
- -

D D
D D
D D D
D D D D D D
f f
f f
f f f
f f f f f f
K
2
0
1
0
2
0
1
0
1
0
2
0
1
0
2
0
1
0
1
0
1
0
2
0
1
0
2
2
2
2
2 (9.78)
We can now introduce an auxiliary variable such as:
-
-

D
D
f
f
R
1
0
2
0
0
(9.79)
Then, we can rewrite the equation of the parameter K by:
( )
- - -
- - - - - -
-
-
- -
- -
-
- -

|
|
.
|

\
|
-
-

o
t -

D D D
D D D D D D
D
D
D D
D D
D
oct
D
oct
f f f
f f f f f f
f
f
f f
f f
K
1
0
2
0
1
0
2
0
1
0
1
0
1
0
2
0
1
0
1
0
1
0
2
0
1
0
2
0
1
0
1
1
2
2
2
3
1
2 3
2
3
3 2 k


0
0
2 1
1
2
R
R
-
-

(9.80)
9.2.4.3 The Rankine Yield Criterion
The Rankine yield criterion also known as the maximum-tensile-stress criterion, (see Chen&Han
(1988)), was formulated by Rankine in 1876. This criterion established that material fails
when the maximum principal stress,
I
o , reaches the critical value
max
T
o . Mathematically,
this criterion is represented by:
max
T I
o o (9.81)
where
III II I
o > o > o and 0
max
> o
T
are satisfied. As a result, this criterion is used for
materials that fail only due to traction ( 0 > o
I
). Thus, we can show the Rankine yield
surface as follows:
0 ) , (
max max
o - o o o
T I T I
F (9.82)
If we now consider the three principal stresses
1
o ,
2
o ,
3
o , the Rankine yield surface is
represented in the principal stress space as shown in Figure 9.22.
Then, by substituting the value of the principal stress
I
o given by the equation (9.35) into
the Rankine criterion equation (9.82) we obtain:
0 cos
3
2
3
cos
3
2
0
max
2
max
2
max
o - - o - - o o - o
T T m T I
I
0 0 J J


(9.83)
Thus,
0 3 cos 3 2 ) , , (
max
2 2
o - -
T
I I 0 0 J J

F (9.84)
Then, if we consider the equations
2
2J q and
m
p o 3 , the yield surface can now be
rewritten in terms of the Haigh-Westergaard coordinates, i.e.:
0 3 cos 2 ) , , (
max
o - -
T
q p q p 0 0 F (9.85)
Note that the yield curve shape on the deviatoric plane H (Nadais plane) can be obtained
by using the equation in (9.31) with 0 p , thus:
9 PLASTICITY

487
0
0
cos 2
3
0 3 cos 2
max
max T
T
q q
o
o - (9.86)
Note that the projection of the deviatoric vector q onto the axis
1
o is
2
3
cos
max
T
q
o
0 ,
which is a constant value. When
2
3
max
0
T
q
o
0 and
max
6 60
T
q o 0 .













Figure 9.22: The Rankine criterion.














Figure 9.23: The Rankine yield criterion.

2
o

3
o

1
o

max
T
o

max
T
o

max
T
o

max
3 T
o o

max
1 T
o o

max
2 T
o o

2
o

3
o

1
o
H -plane

2
o

3
o

1
o
arbitrary deviatoric plane
Projection on the Nadais plane
NOTES ON CONTINUUM MECHANICS

488












Figure 9.24: The Rankine yield criterion Nadais plane.
We can now verify that with a constant angle 0 the equation (9.85) is linear with p and q .
Then, when 0 q we obtain:
max
3
T
p o (9.87)
and when 0 p we find:
0 cos 2
3
max
T
q
o
(9.88)
and in the particular case when 0 0 we obtain
2
3
max
T
q
o
, and when 60 0 we have
max
6
T
q o .
Then, in summary, the different ways of expressing the Rankine yield surface are:
0 3 cos 2 ) , , (
0 3 cos 3 2 ) , , (
0 ) , (
max
max
2 2
max max
o - -
o - -
o - o o o
T
T
T I T I
q p q p
I I
0 0
0 0
F
F
F
J J

Rankine yield surface (9.89)
Then, given the definition

n

o
o
= =
F
F V , the gradient of the Rankine yield surface given
in (9.82), 0 ) , (
max max
o - o o o
T I T I
F , can be obtained as follows:
(
(
(

o o
o o

o o
o
o - o o o
0 0 0
0 0 0
0 0 1
0 ) , (
max max
ij
I
ij
ij T I T I
F
F n
(9.90)


H

1
o

2
o

3
o
0
q

2
3
) 0 (
max
T
q
o
0

max
6 ) 60 (
T
q o 0

9 PLASTICITY

489

























Figure 9.25: The Rankine yield surface.
9.2.5 Evolution of the Yield Surface
In order to fully describe material behavior, we must not just look at the onset of
plastification, but rather at its evolution too, i.e. at how such material behavior evolves
when undergoing loading/unloading/loading. From a material point of view, the yield
surface can be altered after the start of plastification, but such changes depend on the type
of material being analyzed. Next, we will discuss some ways of how the yield surface
evolves, Chen&Han (1988).
The simplest model to characterize material behavior during plastification is the so-called
elastic-perfectly plasticity model, which is characterized by a uniaxial stress-strain curve (as
shown in Figure 9.28). In this scenario the yield surface remains unchanged during
plastification. That is, the yield stress value is not affected, and the yield surface remains
unaltered as plastification evolves.

2
o -

1
o -

3
o -
octahedral plane
0
1
I ( H -plane)

II
o -

I
o -

III
o -

III II I
o o o

6
r
-


6
r
-

pure shear
a) Principal stress space.

C
o
q

T
o
p

max
6
T
o

max

2
3
T
o


max
0
3
T
p o

60 0
0 0
b) Pressure vs. deviatoric part.
NOTES ON CONTINUUM MECHANICS

490
Another idealized model used to show how plasticity evolves is known as hardening
plasticity where we can emphasize two basic models: the isotropic and the kinematic. In
uniaxial cases, the isotropic hardening plasticity model is represented in Figure 9.32. As we can
see in said figure, as plastification evolves the elastic range develops symmetrically. In
three-dimensional cases and (if we dealing with the Drucker-Prager yield surface) the
evolution of this yield surface is as shown in Figure 9.26 in which we can observe that it
does not change its shape but, rather, expands symmetrically during plastification.














Figure 9.26: Evolution of the yield surface in the Drucker-Prager model with isotropic
hardening.
The kinematic hardening plasticity model is characterized by the fact that the size and the shape
of the elastic range do not change whilst plasticity evolves, but the elastic range is able to
move. In Figure 9.37 we show a uniaxial stress-strain curve whose yield surface evolution
neither changes its shape nor expands, but rather changes its position. Then, in a
bidimensional case example, in which the initial yield curve is represented by the
circumference in the principal stress space, if we dealing with the kinematic hardening
model the yield curve only moves in this space, (see Figure 9.27).









Figure 9.27: Evolution of the yield curve kinematic hardening behavior.

1
o

2
o
initial yield surface
current yield surface

2
o

3
o

1
o

3 2 1
o o o

3
o
2
o

1
o
Initial yield surface
H
9 PLASTICITY

491
It is possible to formulate more complex models by combining basic models, such as the
isotropic-kinematic hardening plasticity model which considers isotropic and the kinematic
behaviors simultaneously, i.e. the yield surface expands and can also move.
Then, we summarize some criterion to describe how plasticity evolves:

-
-
-
-
moves) and expands surface The
moves just shape, its change not does surface The
lly symmetrica evolves surface The
evolve) not does surface The
Surface Yield
the of evolution of Law
(
) (
) (
(
Plasticity Hardening Kinematic - Isotropic
Plasticity Hardening Kinematic
Plasticity Hardening Isotropic
Platicity Perfect

The following section will deal with the mathematical formulations that govern the models
described above. For simplicity, we will use the one-dimensional case to describe and
formulate the mathematical expressions that characterize each of these models, and then
we will extend these models to the three-dimensional case (3D), (see Simo&Hughes
(1998)).
9.3 Plasticity Models in Small Deformation
Regime (Uniaxial Cases)
9.3.1 Rate-Independent Plasticity Models (Uniaxial Case)
Next, a few one-dimensional plasticity models will be described. However in some cases
these simple models do not describe how most materials really act. Nevertheless, they are
useful to gain and understanding of the plasticity mechanism, and can also be used to
establish more complex models.
9.3.1.1 Perfect Elastoplastic Behavior
Let us assume that at a hypothetical material point, a material undergoing
loading/unloading shows a stress-strain curve as described in Figure 9.28. Such behavior is
known as perfect plasticity.
The loading and unloading steps represented in Figure 9.28 are indicated by the numbers:
1, 2, 3, 4, 5, 6, 7 and we can verify that after unloading (step 4) there is a residual strain,
called the plastic strain denoted by
) 1 (
p
r , and the corresponding recoverable strain is called
the elastic strain,
) 1 (
e
r .
The rheological model (a device used to interpret physical phenomena) used to represent
the perfect elastoplastic behavior is made up of a linear spring and a Coulomb friction
device in series, (see Figure 9.30). The behavior of each device in isolation can be
appreciated in Figure 9.29. Note that until stress
Y
o is reached, the friction device does not
suffer permanent deformation, and when it reaches the stress limit (which is not permitted
to exceed
Y
o ), it starts to deform without increasing stress. Note also that, when subjected
to unloading, it maintains the strain value at the start.

NOTES ON CONTINUUM MECHANICS

492











Figure 9.28: One-dimensional perfect plasticity.
















Figure 9.29: Behavior of some devices.




Figure 9.30: Rheological model for the perfect elastoplastic behavior.
A physical interpretation of the rheological model described in Figure 9.30 follows: in the
beginning all the stress is absorbed by the spring device until the threshold
Y
o is reached.
Afterwards, all the additional strain is absorbed by the Coulomb friction device, without
E

Y
o
o o
6
1
E
o

Y
o
r
1
2
7
4
5
3
1
E
1
E

) 1 (
p
r
) 1 (
e
r

) 2 (
p
r

) 2 (
e
r

Y
o -

e
r
o

p
r
o

Y
o
1
E
E

Y
o
a) Linear spring device b) Coulomb friction device

Y
o -
unloading/loading

) 1 (
p
r
loading
9 PLASTICITY

493
there being any increase in stress (see Figure 9.28, steps 2-3-6). If there is any unloading the
strain undergone by the Coulomb friction device can not be recovered, and the elastic
device starts to recover its elastic strain, (see Figure 9.28 at the branch 4 or 7).
Next, we will establish the mathematical model that characterizes this type of behavior.
We can verify that the following holds:
Additive decomposition of the strain into elastic and plastic parts:
p e p e
r - r r r - r r (9.91)
Constitutive equation for stress:
( )
p e
E E r - r r o (9.92)
The Criterion of Plastification
If by means of the material behavior described in Figure 9.28, we can check that the
magnitude can never be greater than
Y
o , then we can define the admissible stress space,

E , such as:
{ 0 ) ( s o - o o = o
Y
F R E


(9.93)
If we can verify that when the stress state satisfies 0 ) ( < o - o o
Y
F it means that there is
no production of plastic strain, i.e. 0 r
p
` , then:
0 0 ) ( r < o - o o
p
Y
` F (9.94)
We can define the interior of that admissible stress space ( ( )

E int ) as:
( ) { 0 ) ( < o - o o = o
Y
F R E

int
(9.95)
We can also define the yield surface as follows:
{ 0 ) ( o - o o = o o
Y
F R E


(9.96)
where it meets the plastification criteria, and also holds that:
( ) ( ) C o . o

E E E E E ; int int (9.97)
For one-dimensional case, the yield surface is limited by two points, and does not change
during a loading/unloading/loading process, (see Figure 9.31).
Note that, since the stress space does not support the stress value
Y
o >
/
o thus 0 ) ( >
/
o F ,
and also when 0 ) ( < o - o o
Y
F it implies that 0 r
p
` whereas when 0 r
p
` the
implication is that 0 ) ( o - o o
Y
F , i.e. the material undergoes plastification when the
stress state is on the yield surface.
We can now analyze what happens when the stress has negative values ( 0 < o ). To do so,
we will adopt the scalar 0 ` known as the plastic multiplier. Then, it follows that:
0 0
0 0
< o - o s - r
> o o r
Y
p
Y
p
iff
iff

`
`
`
`

(9.98)
Then we can define the flow rule as:
) (o r sign ` `
p
Flow rule for perfect plasticity (9.99)
NOTES ON CONTINUUM MECHANICS

494
where we have introduced the sign function defined as:

< o -
> o -
o
0 1
0 1
) (
if
if
sign
Sign function (9.100)











Figure 9.31: Perfect elastoplastic behavior.
The flow rule defines the rate of change of the plastic strain, and, in general, is given by:
ij
p
ij
o o
o
r
1
` `
(9.101)
where 1 is a potential. In the particular case when 1 is equal to the yield surface, i.e.
F 1 , the flow rule is said to be associated, which is what happens in the case under
consideration:
( ) ( )
) (o
o o
o o

o o
o - o o

o o
o
r sign ` ` ` ` `
Y p
F

(9.102)
Then, in summary we have:
0 0 ) (
0 0 ) (
0 ) ( ; 0
o - o o
< o - o o
s o

`
`
`
Y
Y
if
if
F
F
F

(9.103)
Now, all the conditions above can be unified into one single condition 0 ) ( o F ` . Then,
we can introduce the Kuhn-Tucker conditions:
0 ) ( ; 0 ) ( ; 0 o s o F F ` `
The Kuhn-Tucker conditions (9.104)
Notice that, with respect to ` , we have not yet fully characterized the mathematical model.
Let us now consider a general scenario in which: at a certain instant in time t the variables
) (t r , ) (t
p
r and ) (t o are known, and 0 ) ( o F is fulfilled, and by the fact that the point is
in the plastification process we have 0 > ` . The rate of change of ) (o F
`
can now be
approached by time discretization by means of the intervals t A in such a way that
0 ) ( ) ( ) ( o - o o A
A - t t t
t F F F
`
is satisfied. Note that, for 0 ) ( > o F
`
it implies that
0 ) ( > o
A - t t
F , whose result is inadmissible. That is, for any given stress state, if the point is
1
E
o

Y
o
r

Y
o -

E o
1
E
9 PLASTICITY

495
outside the yield surface, the only way in which the condition 0 ) ( o
A - t t
F is satisfied
occurs if the yield surface evolves, 0 ) ( o F
`
. In perfect plasticity, this point cannot be
outside the yield surface, the only possibility is by moving on the yield surface. Then,
mathematically, we can summarize the previous comment as follows:
0
0 0
0 0

<
>
F
F
F
`
`
`
`
`
`

iff
iff
(9.105)
The latter condition is known as the consistency (or persistency) condition:
0 F
`
` The consistency (or persistency) condition (9.106)
Then, we can summarize perfect elastoplastic behavior, (see Simo&Hughes (1998)), by:
( )
0
condition y Consistenc
0 ) ( ; 0 ) ( ; 0
condition ary complement Tucker - Kuhn
0 ) (
condition Yield
) (
rule Flow
ip relationsh strain - stress Elastic

o s o
s o - o o
o r
r - r r o
F
F F
F
`
`
` `
` `

v.
iv.
iii.
ii.
i.
Y
p
p e
E E
sign
The perfect
elastoplastic model
(9.107)
9.3.1.2 Isotropic Hardening Elastoplastic Behavior
Let us suppose now that a material point (particle) in a hypothetical material is undergoing
the action of loading/unloading/loading in such a way that the stress-strain curve is that
indicated in Figure 9.32. We can describe this behavior as Isotropic Hardening Elastoplasticity.
With regards to this figure, the loading and unloading steps are indicated by the numbers:
1, 2, 3, 4, 5, 6. We will check, once again, that after unloading step 4 there is a residual
strain
) 1 (
p
r (plastic strain) and a corresponding recoverable strain
) 1 (
e
r (elastic strain). The
difference between this and perfect elastoplastic behavior is: as the plastic strain evolves the
elastic limits does so also, but symmetrically, i.e. the yield surface evolves symmetrically. As
seen in Figure 9.32, before plastification begins the elastic limit in the material is defined by
[ ]
Y Y
o - o , . When the stress state reaches point 3 unloading is applied, and we can then
verify we have a new elastic limit defined by ] , [
* *
Y Y
o - o . Then, because the yield surface
expands symmetrically this model is known as the isotropic hardening elastoplastic model.
The rheological model that shows isotropic hardening elastoplastic behavior is made up of
a linear spring, a Coulomb friction device in parallel, and another linear spring device in
series as shown in Figure 9.33.




NOTES ON CONTINUUM MECHANICS

496















Figure 9.32: Isotropic hardening elastoplastic behavior.











Figure 9.33: The rheological model for the isotropic hardening elastoplastic model.
Physically speaking, the rheological model shown in Figure 9.33 can be interpreted as
follows. Initially, we just have the elastic strain,
e
r , which is caused by the spring with the
constant E (since the Coulomb friction device does not deform for stress values less than
Y
o and because of this the spring with the constant K does not undergo deformation
either). This load stage is typical of the initial elastic domain, and is shown by the branch 1
in Figure 9.32. Note that here
e e
E E r o r o ` ` holds, i.e. the secant modulus and the
tangent modulus are the same, i.e. E E E
T S
. Then, when the stress level reaches the
value
Y
o , the Coulomb friction device starts to deform almost freely, i.e. the Coulomb
friction device strain is controlled by the spring of constant K , and at this stage all the
stress is absorbed by the springs. This stage is represented by the branch 3 2 - in Figure
9.32. Note that, in this model, the parameter K is constant during the deformation
1

eq
K

*
Y
o -
1
E
o

Y
o
r
1
2
4
5
3
6
1
E

) 1 (
p
r
) 1 (
e
r

Y
o -

*
Y
o
Initial elastic region
Expanded elastic region
K

Y
o
o o
E

p
r
e
r
9 PLASTICITY

497
process, so,
p p
Y
K K r o r - o o ` ` holds. Here, with 3 2 - , we can state that the
equivalent spring is that which results from those two in series, as shown in Figure 9.34.
The constant of the equivalent spring,
eq
K , is the tangent of the branch slope 6 3 2 - - .












Figure 9.34: Linear springs disposed in series.
Once surpassed the threshold
Y
o , and then at the point 3 of Figure 9.32 a decrease in
stress (unloading) occurs, the plastic strain in the Coulomb friction device,
) 1 (
p
r , is
maintained and consequently the strain in the spring K is also, which causes a new elastic
limit in the material to appear, which is shown here by
) 1 (
* p
Y Y
Kr - o o . At this unloading
stage, (see Figure 9.32, branch 4 ), there is only strain recovery via the spring device E , i.e.
an elastic process.
Next, we can establish the mathematical model that characterizes the behavior described
above. We can verify that the following relations are still valid:
Additive decomposition of the strain into elastic and plastic parts:
p e p e
r - r r r - r r (9.108)
Constitutive relation and its rate of change:
( ) ( )
p change of rate p e
E E E r - r o r - r r o ` ` ` (9.109)
The Elastic domain and Yield surface
For isotropic hardening elastoplastic behavior the yield criterion is given by:
( ) 0 ) , ( s - o - o o c c K K
Y
F (9.110)
where 0 > o
Y
and 0 > K are material constants and where K is known as the plastic
modulus, (see Simo&Hughes(1998)). Note that the term c K has unit of stress, K has the
same unit as E (unit of stress), and 0 c plays the role of plastic strain which is
dimensionless, (see Figure 9.35). Then, by adopting the hardening hypothesis
p
r ` ` c and
by considering that ) (o r sign ` `
p
, we can obtain c ` ` .
We can define the admissible stress space, (see Figure 9.35(a)), as:
o`
K
o` o`
E

p
r`
e
r`

eq
K
o`
r`

r o
r
-
o
o
-

o
-
o
r
r - r r
r - r r
` `
` `
`
` `
`
` ` `
eq
e p
e p
K
E K
KE
KE
E K
E K
) (
) (

where

) ( E K
KE
K
eq
-

NOTES ON CONTINUUM MECHANICS

498
( ) { 0 ) , ( s - o - o o = o c c K K
Y
F R E


(9.111)
the elastic domain as:
( ) ( ) { 0 ) , ( < - o - o o = o c c K K
Y
F R E

int
(9.112)
and the yield surface as:
( ) { 0 ) , ( - o - o o = o o c c K K
Y
F R E


(9.113)
The elastic domain, ( )

E int , together with its boundary,

E o , describes the admissible


stress space

E , i.e.:
( )

E E E o . int (9.114)














Figure 9.35: Isotropic hardening elastoplastic behavior.
The Kuhn-Tucker conditions are still valid:
0 ) , ( ; 0 ) , ( ; 0 o s o c c K K F F ` `
The Kuhn-Tucker conditions (9.115)
as is the consistency condition:
0 F
`
` The consistency condition (9.116)
The consistency condition allows us to obtain the explicit relationship for ` . That is, when
0 > ` , we must satisfy 0 ) , ( o c K F F
` `
:
0 ) )( 1 (
) (
) , ( - - - o
o o
o

o
o
- o
o o
o
o c c
c
c
c `
`
` `
`
K K
Dt
K D
K
K
F F F
F (9.117)
Then, according to the equation in (9.110) we can verify that ) (o
o o
o
sign
F
. In addition, if
K is a constant variable we have 0 K
`
, and by considering ( )
p
E r - r o ` ` ` and c ` ` , the
equation in (9.117) becomes:
c

*
Y
o

Y
o

Y
o -
o

E o
a) Stress space
1
K - Plastic modulus
*
Y
o

Y
o

p
r c
K
b) Evolution of the flow stress
1
K

p
p
Kr o
r o
` `
` `


*
Y
o -
9 PLASTICITY

499
( )

c
` `
` `
`
` ` ` `
`
) ( ) ( ) (
) ( ) ( ) ( ) , (
K E E
K E E K E K
p p
- o o - r o
- r o - r o - r - r o o
sign sign sign
sign sign sign F

(9.118)
Note that [ ] 1 ) (
2
o sign , so the above equation becomes:
- - r o o 0 ) ( ) , ( c ` ` `
`
K E E sign F
( ) K E
E
-
r o

`
`
) ( sign

(9.119)
Then, considering the equation in (9.102) we can obtain:
( ) ( ) K E
E
K E
E
p
-
r
o
-
r o
o r
` `
` `

) (
) (
) ( sign
sign
sign
(9.120)
The Elastoplastic Tangent Stiffness Modulus
The rate of change of the stress-strain relationship can be obtained as follows:
( )
( ) ( )
r r
-

|
|
.
|

\
|
-
r
- r r - r o
` `
`
` ` ` `
ep p
E
K E
K E
K E
E
E E (9.121)
where
ep
E is the elastoplastic tangent stiffness modulus which is the same as the constant of the
equivalent spring obtained in Figure 9.34.







Figure 9.36: Elastoplastic tangent stiffness modulus.
Let us consider the equation in (9.120) for when 0 > ` , then we obtain:
( )
( )
p p
E
K E
K E
E
r
-
r
-
r
r ` `
`
`


(9.122)
and by substituting this into the equation in (9.121), we obtain:
( ) ( )
( )
r
-
-
r
-
o
p
E
K E
K E
K E
K E
K E
` ` `

p
K r o ` `
(9.123)
Note that K is the line slope of the graph
p
r o described in Figure 9.35(b).
The constitutive relations are now defined as follows:
Elastic regime (loading/unloading)
( ) r o = o < o ` ` ; 0 ) , ( E K

E int c F (9.124)
Elastoplastic regime when unloading
( ) r o = o < o ` `
`
; 0 ) , ( E K

E int c F
(9.125)
Elastoplastic regime when loading
1
E
o

Y
o
r
1

ep
E
1
E
r o ` `
ep
E

( ) K E
K E
E
ep
-



NOTES ON CONTINUUM MECHANICS

500

( )
r r
-
o o = o o ` ` `
`
; 0 ) , (
ep
E
K E
K E
K

E c F
(9.126)
Then, we can summarize the mathematical model, (see Simo&Hughes(1998)), as:
( )
( )
0
condition y consistenc The
0 ) , ( ; 0 ) , ( ; 0
conditions ary complement Tucker - Kuhn The
0 ) , (
condition Yield
law hardening Isotropic
) (
rule Flow
ip relationsh strain - stress Elastic

o s o
s - o - o o

o r
r - r r o
F
F F
F
`
`
` `
` `
` `

c c
c c
c

vi.
v.
iv.
iii.
ii.
i.
K K
K K
E E
Y
p
p e
sign
The isotropic
hardening
elastoplastic
model
(9.127)
9.3.1.3 Kinematic Hardening Elastoplastic Behavior
Kinematic hardening elastoplastic behavior was first described by Prager (1955). In this
model the shape of the yield surface, in the principal stress space, is not altered, however it
can move. In uniaxial cases the yield surface is made up of two points, and additionally, in
kinematic hardening the distance between these does not change during plastification, (see
Figure 9.37).
Now, the yield criterion for kinematic hardening elastoplastic behavior is given by:
0 ) , ( o - - o o
Y
q q F (9.128)
The internal variable q is associated with the new position of the yield surface center in the
principal stress space, (see Figure 9.37) as well as being a function of the plastic strain, for
example we can adopt a linear relationship, e.g.:
p
H q r ` ` (9.129)
where H is known as the kinematic hardening modulus (material property) and generally
speaking it is a tangent modulus of the curve
p
q r , (see Figure 9.37). In the case under
consideration we will consider H to be a constant variable.
In this model the rate of change of the plastic strain is defined as follows:

< o - - o < - r
> o - o > r
0 0
0 0
Y
p
Y
p
q if
q if

`
`
`
`
) ( q
p
- o r sign ` `
(9.130)
where

< - o -
> - o -
- o
0 1
0 1
) (
q if
q if
q sign (9.131)
Then, if we consider the equations in (9.129) and (9.130), we obtain:
) ( q H H q
p
- o r sign ` ` ` (9.132)
9 PLASTICITY

501













Figure 9.37: Kinematic hardening elastoplastic behavior.
Then, we can summarize the mathematical model, (see Simo&Hughes(1998)), as:
( )
0
condition y Consistenc
0 ) , ( ; 0 ) , ( ; 0
condition ary complement Tucker - Kuhn
0 ) , (
condition Yield

law hardening Kinematic
) (
rule Flow
ip relationsh strain - stress Elastic

o s o
s o - - o o
r
- o r
r - r r o
F
F F
F
`
`
` `
`
` `
`

vi.
v.
iv.
iii.
ii.
i.
q q
q q
H q
q
E E
Y
p
p
p e
sign

The kinematic
hardening
elastoplastic
model
(9.133)
The Elastoplastic Tangent Stiffness Modulus
As we saw before, the consistency condition, 0 F
`
` , allows us to obtain the explicit
relation for 0 > ` after which we take the rate of change of ) , ( q o F F :
0 ) , (
o
o
- o
o o
o
o q
q
q ` `
`
F F
F
(9.134)
Then, according to the equations given in (9.133), we can verify that
) ( q
q
q
- o
o o
- o o
- o o
o

o o
o
sign
F F
, ) ( q
q
q
q q
- o -
o
- o o
- o o
o

o
o
sign
F F
, and also if we
consider that ( )
p
E r - r o ` ` ` and ) ( q H q - o sign ` ` , the equation in (9.134) becomes:

Y
o -
1

*
o

*
Y
o -
1
E
o

Y
o
r
1
2
4
5
3
6
1
E

) 1 (
p
r
) 1 (
e
r

*
Y
o
Initial elastic region
Displaced elastic region
1

ep
E
o
q
q

p
r
H
NOTES ON CONTINUUM MECHANICS

502
( )
[ ] 0 ) (
0 ) ( ) ( ) (
0 ) ( ) (
0 ) ( ) ( ) (
0 ) ( ) (
0 ) , (
- - r - o
- - o - o - r - o
- r - o - r - o
- o - o - r - r - o
- o - o - o

o
o
- o
o o
o
o
H E E q
H q E q E q
H E q E q
q H q E q
q q q
q
q
q
p
p

`
` `
`
`
`
`
` `
` `
` `
` `
`
sign
sign sign sign
sign sign
sign sign sign
sign sign
F F
F

(9.135)
Thus
[ ]
[ ]
` ` ` `
E q
H E
H E
E q
) (
) (
- o
-
r r
-
- o

sign
sign

(9.136)
after which the rate of change of the plastic strain becomes:
[ ] [ ]
r
-
- o r
-
- o
- o r ` ` ` `
H E
E
q
H E
E q
q
p
) (
) (
) ( sign
sign
sign
(9.137)
Then, by means of the rate of change of the stress-strain relationship we can obtain the
elastoplastic tangent stiffness modulus
ep
E , i.e.:
( )
[ ] [ ]
. r r
-

|
|
.
|

\
|
r
-
- r r - r o
` ` ` ` ` ` `

ep p
E
H E
EH
H E
E
E E
[ ] H E
EH
E
ep
-
(9.138)
Thus, we summarize that:
0 r o ` ` ` if E and
[ ]
0 > r
-
o ` ` ` if
H E
EH

(9.139)
9.3.1.4 Isotropic-Kinematic Elastoplastic Behavior
This model is a combination of the models discussed previously. Here, the yield criterion
can be defined as follows:
( ) 0 ) , , ( s - o - - o o c c K q K q
Y
F (9.140)
We can summarize the mathematical model, (see Simo&Hughes(1998)), as:
( )
( )
0
condition y Consistenc
0 ) , , ( ; 0 ) , , ( ; 0
condition ary complement Tucker - Kuhn
0 ) , , (
condition Yield
) (
law hardening Kinematic - Isotropic
) (
rule Flow
ip relationsh strain - stress Elastic

o s o
s - o - - o o

- o
- o r
r - r r o
F
F F
F
`
`
` `
` `
`
`
`
`

c c
c c
c

vi.
v.
iv.
iii.
ii.
i.
K q K q
K q K q
q H q
q
E E
Y
p
p e
sign
sign

The isotropic-
kinematic
elastoplastic
model
(9.141)
9 PLASTICITY

503
The Elastoplastic Tangent Stiffness Modulus
Once again, we will use the consistency condition, 0 `
`
F , to obtain ` . Firstly, we will
calculate the rate of change of ) , , ( c q o F by means of the chain rule of the derivative, i.e.:
0
) (
) (
) , , ( s
o
o
-
o
o
- o
o o
o
o
Dt
K D
K
q
q
K q
c
c
c
F F F
F ` `
`

(9.142)
Then, according to the equation in (9.141), we will verify that
) ( q
q
q
- o
o o
- o o
- o o
o

o o
o
sign
F F
, ) ( q
q
q
q q
- o -
o
- o o
- o o
o

o
o
sign
F F
,
c
c
c
` K
Dt
K D
K
-
o
o ) (
) (
F
, and also if we consider that ( )
p
E r - r o ` ` ` , c ` ` and
) ( q H q - o sign ` ` , the equation in (9.142) becomes:
( )
[ ]
[ ]
[ ] 0 ) (
0 ) ( ) ( ) (
0 ) ( ) (
0 ) ( ) ( ) (
0 ) ( ) (
0
) (
) (
) , , (

- - - r - o
- - - o - o - r - o
- - r - o - r - o
- - o - o - r - r - o
- - o - o - o

o
o
-
o
o
- o
o o
o
o
K H E E q
K H q E q E q
K H E q E q
K q H q E q
K q q q
Dt
K D
K
q
q
K q
p
p


c
c
c
c
`
` `
`
` `
`
`
`
` `
` `
` `
` `
`
sign
sign sign sign
sign sign
sign sign sign
sign sign
F F F
F
(9.143)
Thus
[ ]
[ ]
` ` ` `
E q
K H E
K H E
E q
) (


) (
- o
- -
r r
- -
- o

sign
sign

(9.144)
after which we can obtain the rate of change of the plastic strain as follows:
[ ] [ ]
r
- -
- o r
- -
- o
- o r ` ` ` `
K H E
E
q
K H E
E q
q
p

) (

) (
) ( sign
sign
sign
(9.145)
Then, by means of the rate of change of the stress-strain relationship we can obtain the
elastoplastic tangent stiffness modulus
ep
E , i.e.:
( )
[ ]
( )
[ ]


. r r
- -
-

|
|
.
|

\
|
r
- -
- r r - r o ` ` ` ` ` ` `
ep p
E
K H E
K H E
K H E
E
E E
( )
[ ] K H E
K H E
E
ep
- -
-


(9.146)
Thus, in summary we have:
( )
[ ]
0

0 > r r
- -
-
o r o ` ` ` ` ` ` ` if E
K H E
K H E
and if E
ep

(9.147)
9.4 Plasticity in Small Deformation Regime
(The Classical Plasticity Theory)
In this section we generalize the theory of plasticity for three-dimensional cases. We will
consider the small deformation regime, (see subsection 2.14 in Chapter 2), and,
additionally, we will make the following assumptions:
NOTES ON CONTINUUM MECHANICS

504
There is material isotropy. That is, during plastification the material does not lose
its isotropy;
The plastic strain produces no change in volume;
The process is isothermal and adiabatic. That is, during plastification, the effect of
temperature is not taken into account, i.e. the constitutive equation is independent
of temperature.
The production of plastic strain is associated with internal energy dissipation, i.e. with
irreversible processes. To put it another way: a process which involves plastic strain is path
dependent, for example, in Figure 9.38 the points A and B are associated with the same
stress but have different strain values, so, here we have a process that depends on the load
history. Now, we will adopt the constitutive equation with internal variables, (see
subsection 6.4.1 in Chapter 6), in order to describe the motion history, albeit indirectly.










Figure 9.38: Plasticity dependence of the load history.
9.4.1 The Infinitesimal Strain Tensor and Constitutive
Equation
With small deformation, we use additive decomposition of the infinitesimal strain tensor
into elastic and plastic parts, i.e.:
p e
-
Additive decomposition of the
infinitesimal strain tensor
(9.148)
where is the infinitesimal strain tensor, whose components in the Cartesian basis are
p
ij
e
ij ij
r - r r , (see Chapter 2 in subsection 2.14). The assumption in (9.148) forms the basis
of the classical theory of plasticity, which is valid in a small deformation regime.
Then, if we consider (9.148) the constitutive equation for stress becomes:
( ) : : :
1
;
-
-
e e p e e e
C C C

(9.149)
where I 1 1 j A 2 -
e
C is the elasticity tensor for isotropic materials, and the elastic
compliance tensor is given by I 1 1
j j A j
A
2
1
) 2 3 ( 2
1
-
-
-

-
e
C , where A, j are the Lam
constants, (see Chapter 7).

) (B
r

e
r

) ( A
r
B A
o
r

B A
o o

p
r
9 PLASTICITY

505
9.4.2 Helmholtz Free Energy
We apply the constitutive equation with internal variables, (see Chapter 6), where, in
general, the Helmholtz free energy, ) , , (
k
T , is described in terms of the
infinitesimal strain tensor ( ), temperature ( T ), and a set of internal variables (
k
), which
could be a scalar, vector, or higher order tensor. Then, if we consider the process to be
isothermal, we have ) , (
k
. We can now reformulate this energy expression, in
order to obtain the elastic part of as a free variable, i.e.: ) , (
k
e
. Now the set of
internal variables
k
do not include the plastic component (
p
) of the strain tensor,
because it is already included in the free variable
p e
- . Then, the following is
satisfied:
p e
o
o
-
o
o

o
o
(9.150)
Now, the rate of change of the Helmholtz free energy ) , (
k
e
becomes:
k
k
e
e

`
`
`

o
o
-
o
o

:
(9.151)
NOTE: The operator is substituted by the number of contractions of the order of .
That is, if is a scalar (zeroth-order tensor), has no contraction; if is a vector (first-
order tensor), there is one contraction, i.e. the scalar (dot) product; if is second-
order tensor, : there are two contractions, i.e. the double scalar product; and so on.
9.4.3 Internal Energy Dissipation and the Evolution of the
Internal Variables
As we discussed in Chapter 6, the entropy inequality is not an additional equation to the
governing equations, but rather it is used to put restrictions on the variables of the
problems. That is, the entropy inequality tells us how the internal variables must evolve
during plastification. Then, by using the Clausius-Duhem inequality, (see Chapter 5), we
obtain:
0 0
1
- - - - j j jj
` `
,
`
D q D : :
int int
T
T
T D D
process
isothermal
x
V
(9.152)
Note that, in isothermal processes the internal energy dissipation is purely mechanical. By
substituting the rate of change of the Helmholtz free energy given in (9.151) into the
Clausius-Duhem inequality, and also if we know that
p e
D
` ` `
`
- = E holds in a small
deformation regime, we can obtain:
0 ) (
(

o
o
-
o
o
- -
k
k
e
e
p e
int

`
` ` `


j

: : D
(9.153)
( ) 0 - - |
.
|

\
|
o
o
-
k k
p e
e
int
`
` `
A

: :

j D (9.154)
where we have denoted by
k
k
o
o
-

j A (the thermodynamic forces). Then, as the
inequality in (9.154) must be valid for any admissible thermodynamic process, (see Chapter
6), so, we can adopt a process in which 0
p
`
and 0
k
` , then the following must hold:
NOTES ON CONTINUUM MECHANICS

506
e

o
o


j (9.155)
Note that is related to the gradient of in the strain space, and
k
A is related to the
gradient of in the space of
k
.
Then, by considering the constitutive equation in (9.155), the internal energy dissipation
(9.154) becomes:
0 -
k k
p
int
`
`
A : D (9.156)
Next, we apply the maximum dissipation principle, which states that the dissipation in a
material reaches a maximum during a change characterized by a dissipative process. Let us
consider the current state ) , (
k
A , which is the current distribution of Cauchy stress tensor
and thermodynamic forces in a body subjected to plastic strain. The principle of maximum
energy dissipation requires that for a change of state, represented by ) , (
* *
k
A , the
following must be satisfied:
0 ) ( ) (
* *
- - -
k k k
p
`
`
A A : (9.157)
The inequality in (9.157) describes an optimization problem with constraint. We can also
maximize the dissipation by minimizing the negative dissipation with the constraint
0 ) , ( s
k
A 1 . To this end, we define the Lagrangian as:
1 1 ` `
`
- - - - -
k k
p
int
A : D L (9.158)
where 0 ` is the Lagrangian multiplier (plastic multiplier) that enforces 0 s 1 .
Then, the flow rule can be obtained by:

0
o
o

o
o
- -
o
o ) , ( ) , (
k p k p
A A 1

1
` `
` `
L
(9.159)
and the evolution of the internal variables is given by:
k
k
k
k
k
A A A o
o

o
o
- -
o
o 1

1
` ` ` ` 0 0
L

(9.160)
If we now make a change of variable such that G 1 , where ) , (
k
A G is a plastic
potential, we obtain:

o
o

G
`
`
p
Plastic flow rule (9.161)
k
k
A o
o

G
` ` Evolution of the internal variables
k
(9.162)
Then to fully define the model we need to introduce the loading/unloading Kuhn-Tucker
conditions:
0 ) , ( ; 0 ) , ( ; 0 s
k k
A A F F ` `
The Kuhn-Tucker conditions (9.163)
and the persistency (or consistency) condition:
0 ) , (
k
A F
`
` The persistency condition (9.164)
where ) , (
k
A F is the yield surface, and ` is the plastic multiplier.
9 PLASTICITY

507
In associated flow, the plastic potential G is equal to the yield surface F , i.e.
F G
Associated flow (9.165)
We can now summarize the elastoplastic model for isothermal processes under the small
deformation regime as:
0 ) , (
condition y Consistenc
0 ) , ( ; 0 ) , ( ; 0
condition ary complement Tucker - Kuhn
s variable internal the of Evolution
rule flow Plastic
stress for equation ve Constituti
ip relationsh strain - stress Elastic

s
o
o

o
o

o
o

-
k
k k
k
k
p
e
p e
A
A A
A


F
F F
G
G
`
`
` `
` `
`
`

vi.
v.
iv.
iii.
ii.
i.

Elastoplastic model for


isothermal small deformation
regime
(9.166)
9.4.4 The Elastoplastic Tangent Stiffness Tensor
In this section we will establish the rate of change of the stress-strain relationship, i.e.:

` `
:
tan_ep
C
(9.167)
with which we can obtain the elastoplastic tangent stiffness tensor,
tan_ep
C , starting off from the
constitutive equation:
e e
: C
(9.168)
where
e
C is the elasticity tensor, which does not vary with time ( O C
e
`
), and
e
is the
elastic part of the infinitesimal strain tensor. Then, by means of the strain tensor additive
decomposition, (see Eq. (9.148)), we can express the elastic part as
p e
- , and by
substituting this into the equation in (9.168) we obtain:
-
change of rate
) (
p e e e
: : C C ) (
p e

` ` `
- : C (9.169)
Next, using the flow rule defined in (9.161), i.e.

o
o

G
`
`
p
, the above equation becomes:
|
|
.
|

\
|
o
o
- -


G
`
` ` ` `
: :
e p e
C C ) (
(9.170)
According to the consistency condition when 0 > ` the implication is that 0 ) , (
k
A F
`
.
The rate of change of ) , (
k
A F F can be evaluated as follows:
0 ) , (
1

o
o
-
o
o

n
a
a
a
k
A
A
A
`
` `
F F
F

: (9.171)
NOTES ON CONTINUUM MECHANICS

508
where n is the number of internal variables. Then, by assuming that
a
A is a function of
plastic strain we can conclude that:
|
|
.
|

\
|
o
o
o
o

o
o
o
o

o
o
o
o

o
o


`
`
`
` `
`
` ` `
`
`
p
p
a
p
p
a
p
p
a p
p
a
a

: : : :
A A A A
A (9.172)
Then, the equation in (9.171) becomes:
0 ) , (
1 1

|
|
.
|

\
|
o
o
o
o
o
o
-
o
o

o
o
-
o
o



n
a
p
p
a
a
n
a
a
a
k

`
`
`
`
`
` `

: : :
A
A
A
A
A
F F F F
F (9.173)
Next, by substituting the equation in (9.170) into the one above we obtain:
0
0
0 ) , (
1
1
1

|
|
.
|

\
|
o
o
o
o
o
o
-
o
o
o
o
-
o
o

|
|
.
|

\
|
o
o
o
o
o
o
-
(

|
|
.
|

\
|
o
o
-
o
o

o
o
-
o
o

n
a
p
p
a
a
e e
n
a
p
p
a
a
e
n
a
a k

`
` `
`
` `
`
`
`
`
`
` `

: : : : :
: : :
: :
A
A
A
A
A
F
G
F F
F
G
F
G
H
F
F
C C
C
(9.174)
with which the plastic multiplier ` can be evaluated as follows:

|
|
.
|

\
|
o
o
o
o
o
o
-
o
o
o
o
o
o

n
a
p
p
a
a
e
e
1

`
`
`
`

: : :
: :
A
A
F
G
F
F
C
C

(9.175)
Then, drawing once more on the equation in (9.170) and if we consider the plastic
multiplier obtained in (9.175), it follows that:
|
|
|
|
|
.
|

\
|
o
o

|
|
.
|

\
|
o
o
o
o
o
o
-
o
o
o
o
o
o
-
|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
-


G
F
G
F
F
G G
n
a
p
p
a
a
e
e
e
e e
1


`
` `
`
`
`
` ` `
: : :
: :
:
: :
A
A
C
C
C
C C
(9.176)
For the sake of simplicity, we will change the variables such as

n
o
o

F
,

m
o
o

G
, where n
and m are symmetric second-order tensors after which the equation in (9.176) can be
rewritten as:
|
|
|
|
|
.
|

\
|

|
|
.
|

\
|
o
o
o
o
o
o
-
-

m n
n

n
a
p
p
a
a
e
e
e
1
`
`
`
` `
: : :
: :
:
A
A
F
C
C
C
(9.177)
Note that the denominator is a scalar, and by denoting the inverse of the denominator by
X we obtain:
( ) m n -
` ` `
: : : :
e e e
C C C X (9.178)
9 PLASTICITY

509
We will now continue by using indicial notation:
( )
( )
( )
( )
( )
st
e
abst ab kl
e
ijkl
e
ijst
st kl
e
abst ab
e
ijkl
e
ijst
st kl td sc
e
abcd ab
e
ijkl tl sk
e
ijkl
kl td sc st
e
abcd ab
e
ijkl tl sk st
e
ijkl
kl cd
e
abcd ab
e
ijkl kl
e
ijkl ij
r -
r -
r -
r - r
r - r o
`
`
`
` `
` ` `






C C C
C C C
C C C
C C C
C C C
n m
m n
m n
m n
m n
X
X
X
X
X
c c c c
c c c c
(9.179)
Then, in tensorial notation the above equation becomes:
( ) n m
` `
: : :
e e e
C C C - X (9.180)

m n
n m

` ` `
`
`
`
: :
: : :
: :
tan_ep
n
a
p
p
a
a
e
e e
e
C
C
C C
C
|
|
|
|
|
.
|

\
|
|
|
.
|

\
|
o
o
o
o
o
o
-

1

A
A
F

(9.181)
Then, we define the elastoplastic tangent stiffness tensor as follows:

|
|
.
|

\
|
o
o
o
o
o
o
-
o
o
o
o
o
o

o
o
-
n
a
p
p
a
a
e
e e
e tan_ep
1
`
`



: : :
: :
A
A
F
G
F
F
G
C
C C
C C
Elastoplastic tangent stiffness
tensor
(9.182)
Note that
tan_ep
C is a fourth-order (but, not necessarily symmetric) tensor, showing only
minor symmetry due to that
`
and
`
. Then, in terms of n and m the elastoplastic
tangent stiffness tensor becomes:

|
|
.
|

\
|
o
o
o
o
o
o
-

-
n
a
p
p
a
a
e
e e
e tan_ep
1
`
`

m n
n m
: : :
: :
A
A
F
C
C C
C C
Elastoplastic tangent stiffness
tensor
(9.183)
In associated flow, i.e. G F , the elastoplastic tangent stiffness tensor becomes:

|
|
.
|

\
|
o
o
o
o
o
o
-

-
n
a
p
p
a
a
e
e e
e tan_ep
1
`
`

n n
n n
: : :
: :
A
A
F
C
C C
C C
Elastoplastic tangent stiffness
tensor for associated flow
(9.184)
In associated flow, the elastoplastic tangent stiffness tensor,
tan_ep
C , is a symmetric tensor,
i.e. it features both major and minor symmetry.
Problem 9.1: Consider a one-dimensional case, find the elastoplastic tangent stiffness
tensor (elastoplastic tangent stiffness modulus) for the following: 1) perfect plasticity; 2)
isotropic hardening plasticity, 3) kinematic hardening plasticity, and 4) isotropic-kinematic
hardening plasticity.
Solution:
We start from the general definition of the elastoplastic tangent stiffness tensor given in
(9.182), in which the associated flow F G takes place. Then:
NOTES ON CONTINUUM MECHANICS

510

|
|
.
|

\
|
o
o
o
o
o
o
-
o
o
o
o
o
o

o
o
-
n
a
p
p
a
a
e
e e
e tan_ep
1
`
`



: : :
: :
A
A
F F F
F F
C
C C
C C
1) Perfect plasticity
With perfect plasticity, the yield surface is given by:
) ( ) ( o
o o
o
o - o o sign
F
F
Y

and the plastic flow is:
) (o r sign ` `
p

Note that, since there are no internal variables we have 0
1

|
|
.
|

\
|
o
o
o
o
o
o

n
a
p
p
a
a
`
`

:
A
A
F
.
Furthermore, as we are dealing with a one-dimensional case, the tensors can be represented
by their only nonzero component, i.e. r o r o ` ` ` `
tan_ep tan_ep
E
11 1111 11
C , and
o o
o

o
o

F F

; ; E E
e tan_ep tan_ep
C C
Thus, as expected:
[ ] [ ]
0
0 ) ( ) (
) ( ) (
-
- o o
o o
- E E
E
E E
E E
tan_ep
sign sign
sign sign

2) Isotropic hardening plasticity
In the case of isotropic hardening plasticity, the yield surface is given by:

( )

-
o
o

o
o
o
o o
o
- o - o o
1
) (
) (
) , (
c
c c
K A
K K
Y
A
F F
F
F
sign

where the internal variable in stress is the scalar c K A .
Then, the plastic flow is:

o
r o
o
o
o
r o
o r
) (
) (
; ) (
sign
sign
sign
p
p
p
c

c
`
` ` `
`
`
The term

|
|
.
|

\
|
o
o
o
o
o
o
n
a
p
p
a
a 1
`
`

:
A
A
F
is evaluated as follows:
( ) K K
K
K
a
p
p
a
p
p
- o o -
|
|
.
|

\
|
o
r o
r o
o
o
o

|
|
.
|

\
|
o
o
o
o
o
o


) ( ) ( 1
1
1
1
1
1
1
sign sign

c
c ` `
`
`
F F

:
A
A

Then, the elastoplastic tangent stiffness modulus becomes:
K E
EK
K E
E
E
K E
E E
E E
tan_ep
-

-
-
- - o o
o o
-
2
) ( ) ( ) (
) ( ) (
sign sign
sign sign


3) Kinematic hardening plasticity
In the case of kinematic hardening plasticity, the yield surface is given by:
9 PLASTICITY

511

- o -
o
o
- o
o o
o
o - - o o
) (
) (
) , (
q
q
q
q q
Y
A
sign
sign
F
F
F
where the internal variable in stress is represented by the scalar q A .
Then, the plastic flow is:
) (

) (
o
o
r o

r
o r
sign
sign

`
` `
` `
`
p
p
p
H q

The term

|
|
.
|

\
|
o
o
o
o
o
o
n
a
p
p
a
a 1
`
`

:
A
A
F
is evaluated as follows:
( ) H q H q
q
q
a
p
p
a
p
p
- - o - o -
|
|
.
|

\
|
o
r o
r o
o
o
o

|
|
.
|

\
|
o
o
o
o
o
o


) ( ) (
1
1
1
1
1
1
sign sign
` `
`
`
F F

:
A
A

Then, the elastoplastic tangent stiffness modulus becomes:
H E
EH
H E
E
E
H E
E E
E E
tan_ep
-

-
-
- - o o
o o
-
2
) ( ) ( ) (
) ( ) (
sign sign
sign sign


4) Isotropic-kinematic hardening plasticity
In the case of isotropic-kinematic hardening plasticity, the yield surface is given by:
( )

-
o
o
- o -
o
o
- o
o o
o
- o - - o o o
1
) (
) (
) , , ( ) , , (
2 1
c
c c
K
q
q
q
K q K q A A
Y
F
F
F
F F sign
sign

Thus:
|
|
.
|

\
|
o
r o
r o
o
o
o
-
|
|
.
|

\
|
o
r o
r o
o
o
o

|
|
.
|

\
|
o
r o
r o
o
o
o

` ` `
` ` `
p
p
p
p
a
p
p
a
a
A
A
A
A
A
A
2
2
1
1
2
1
F F F

The internal variables are q A
1
, c K A
2
, with which we obtain:
( ) ( ) [ ]
K H
q q K q H q
K
K
q
q
A
A
A
A
A
A
p
p
p
p
p
p
p
p
a
p
p
a
a
- -
- o - o - - - o - o -
|
|
.
|

\
|
o
r o
r o
o
o
o
-
|
|
.
|

\
|
o
r o
r o
o
o
o

|
|
.
|

\
|
o
r o
r o
o
o
o
-
|
|
.
|

\
|
o
r o
r o
o
o
o

|
|
.
|

\
|
o
r o
r o
o
o
o

) ( ) ( 1 ) ( ) (
2
2
1
1
2
1
sign sign sign sign

c
c

` `
` ` `
` `
` ` `
F F
F F F

Then, the elastoplastic tangent stiffness modulus becomes:
( )
( )
K H E
K H E
K H E
E
E E
tan_ep
- -
-

- -
-
2


NOTES ON CONTINUUM MECHANICS

512
9.4.5 The Classical
2
J
Flow Theory
9.4.5.1 Perfect Plasticity
Let us consider the following assumptions, (see Simo&Hughes (1998)):
1. Isotropic linear elastic behavior;
2. The Huber-von Mises yield condition
[ ]
Y
R with R o - -
3
2
2
3
1
2
) ( ) ( Tr F
(9.185)
where R is the radius of the yield surface, and
Y
o is the yield stress (elastic limit
stress), and is the Frobenius norm of the Cauchy stress tensor, (see subsection
1.5.7 Norms of Tensors in Chapter 1).
3. There is Levy-Saint_Venant associated plastic flow
4. There is no hardening
Note that ) ( F can be expressed as follows:
[ ] ( )
R
R R R
-
- - - - - -
s
1 1
:
: : : ) ( ) ( ) ( ) (
3
1
3
1
2
3
1
2
Tr Tr Tr F
(9.186)
where
dev
s = , i.e. it is the deviatoric part of the Cauchy stress tensor. Then by substituting
1 s ) (
3
1
Tr - into the above equation we obtain:
( )
R
R R R
-
- - - -
s
s s 1 s s s : : : ) ( ) (
3
1
Tr F
(9.187)
where we have taken into account that the trace of any deviatoric tensor is equal to zero, i.e.
0 ) ( s 1 s Tr : .
We then obtain the gradient of ) ( F in the stress space, the result of which is the plastic
flow tensor n:
[ ] ( ) ( ) s

s
s s

s
s s

s
s s s s

n : : : : : :
o
o
|
.
|

\
|
o
o
-
o
o
-
o
o

o
o

- -
2
1
2
1
2
1 ) (
R
F
(9.188)
Note that s s

s

o
o
: , (see Problem 1.39) with which we obtain:
( )
s
s
s s
s
s

s
s s

n
o
o

o
o

-
:
: : 2
1
) ( F

(9.189)
Notice also that the Frobenius norm of n is unitary, 1 n n n : .
With the above, we can obtain the plastic multiplier such as that defined in (9.175) for the
associated flow case:
0
1
-

|
|
.
|

\
|
o
o
o
o
o
o
-
o
o
o
o
o
o

n n
n

: :
: :
: : :
: :
e
e
n
a
p
p
a
a
e
e
C
C
C
C
`
`
`
`
`

A
A
F
G
F
F

(9.190)
9 PLASTICITY

513
where the elasticity tensor, (see Chapter 7), can be expressed as follows:
[ ] 1 1 I 1 1 I 1 1 - - x -
3
1
2 2 j j A
e
C (9.191)
Thus
( )
( )
( )
n I n n 1 1 n
I n 1 1 n
n I 1 1 n
I 1 1 n
n n
n
: : : :
: : :
: :
: :
: :
: :
j A
j A
j A
j A

2
2
0 2
2
0 -
-

- -
-

` ` `
`
e
e
C
C
(9.192)
Then, if we consider that 0 ) ( n 1 n Tr : (trace of the deviatoric tensor), n n I n
sym
: ,
we obtain:
s
s
n n
n n
n
with
`
`
` :
:
:


(9.193)
We can also find the elastoplastic tangent stiffness tensor for the associated flow case,
given in (9.184):
( )
( ) ( )
( ) n I 1 1 n
I 1 1 n n I 1 1
I 1 1

n n
n n
: :
: :
: : :
: :
j A
j A j A
j A

2
2 2
2
0
1
-
- -
- -
|
|
.
|

\
|
o
o
o
o
o
o
-

_
`
n
a
p
p
a
a
e
e e
e tan_ep
A
A
F
C
C C
C C

(9.194)
Then, by simplifying the above equation we obtain:
[ ]
[ ] n n 1 1 I 1 1
n n I 1 1
- - - x
- -
3
1
2
2
j
j A
tan_ep
C
(9.195)
for when 0 > ` .
9.4.5.2 Isotropic-Kinematic Hardening Plasticity
This model has two internal variables, namely { ) ( ;
2 1
c c K A q A where c (scalar) is the
equivalent plastic strain which defines isotropic hardening behavior and q (a second-order
tensor) defines the center of the von Mises yield surface in the deviatoric stress space. In
this model we have the following hardening law and plastic flow rules:
) ( ) , , (
0 ) ( ;
3
2
2 1
c K A -
-

A
q q

s
F
Tr

3
2
3
2
; ) ( c c

` ` `
`
`

H
p
q


(9.196)
where
dev
s = , ) (c K is the isotropic hardening modulus, and ) (c H is the kinematic
hardening modulus. Then, given that `
`

p
we can obtain the equivalent plastic strain as
follows:
t t
}
d t
p
t
) (
3
2
) (
0

`
c
(9.197)
For the function ) (c K we assume the linear variation:
NOTES ON CONTINUUM MECHANICS

514
c c K K
Y
- o ) ( (9.198)
where K is a constant, and
Y
o is the yield stress.
The plastic multiplier can be found by means of the equation in (9.175) for the associated
flow case, i.e.:


|
|
.
|

\
|
o
o
o
o
o
o
-

|
|
.
|

\
|
o
o
o
o
o
o
-

|
|
.
|

\
|
o
o
o
o
o
o
-

n
a
p
p
a
a
n
a
p
p
a
a
n
a
p
p
a
a
e
e
1
1 1
2
2
2
2

j
j

j
j

`
` `
`
`
`
`
`
`
`

n n
n

n n
n
:
:
: :
:
: : :
: :
A
A
A
A
A
A
F
F F
C
C
(9.199)
where it holds that 1 n . Furthermore, we can verify that the following holds:
K H
K H
K H
K
K
p
p
p
p
p
p
p
p
n
a
p
p
a
a
3
2
3
2
3
2
3
2
) ( ) (
3
2
3
2
) ( ) (
3
2
) (
) (
) (
2
2
1
1 1
- -
- -
|
|
.
|

\
|
-
|
|
.
|

\
|
-
|
|
.
|

\
|
o
o
o
o
o
o
-
|
|
.
|

\
|
o
o
o
o
o
o

|
|
.
|

\
|
o
o
o
o
o
o
-
|
|
.
|

\
|
o
o
o
o
o
o

|
|
.
|

\
|
o
o
o
o
o
o

n n n n

: :
: :
: :
: : :

sign sign sign sign c

c
c

` `
` ` `
` `
` ` `
F F
F F F
q
q
A
A
A
A
A
A

(9.200)
where 1 n n with

.
Then, the plastic multiplier becomes:

- - -

|
|
.
|

\
|
o
o
o
o
o
o
-

K H
n
a
p
p
a
a
3
2
3
2
2
2
2
2
1
j
j

j
j

n
`
`
`
`
`
:
:
:
A
A
F

j

3
1
K H -
-

n
`
`
:

(9.201)
and the elastoplastic tangent stiffness tensor is given by:
(
(
(
(

-
-

- - - x
j
j
3
1
2
3
1
K H
tan_ep
n n
1 1 I 1 1 C (9.202)
9 PLASTICITY

515
9.5 Plastic Potential Theory
Von Mises concluded that the strain tensor, , was related to the stress tensor, , by
means of the elastic potential function (the complementary strain energy),
c
U , as

o
o

c
U
.
Similarly, von Mises suggested there was a plastic potential function, ) ( G , with which the
rate of change of the plastic strain is given by:

o
o

G
`
`
p

ij
p
ij
o o
o
r
G
` `
(9.203)
where ` is the plastic multiplier- a positive scalar. This theory is known as the plastic
potential theory. One possible approach we can take to this is to consider the plastic potential
to be equal to the yield surface F G , with which it is said that the flow is associated.
Otherwise, i.e. when F G , there is said to be non-associated flow rule. In the case of the
former we obtain:

o
o

F
`
`
p

ij
p
ij
o o
o
r
F
` `
(9.204)
The Druckers Stability postulates:
The plastic work done by an external agency, during the application of additional
stress, is positive.
The total work done by an external agency during a cycle can not be negative.
If any of these criteria is not met, the material is said to be unstable.
The Normality
The incremental plastic strain tensor is normal to the yield surface.
We will discuss these rules in the following example. First, let us consider a stress
relationship as shown in Figure 9.39.













Figure 9.39: Plasticity, the load history dependency.

*
) 1 (
o
Unstable
o d

0
o

*
) 1 (
o

) 1 (
o

Y
o
o
r
o

p
r
Stable

*
o
a) b)
NOTES ON CONTINUUM MECHANICS

516
Let us consider the stress state
ij
o which is inside the initial yield surface at the initial
instant of time (
0
t ). At time
1
t , the point is found on the yield surface, so the process
observed between
0
t and
1
t is purely elastic. From
1
t to t t o -
1
, we apply a load
increment described here by o d , and then we apply unloading, as shown in Figure 9.40.









Figure 9.40: Evolution of the yield surface.
The total work done is given by
0
dW dW dW
T n
- , then that done in the process described
in Figure 9.39 can be evaluated as follows:
( )
}
} } } } }
o -
o -
o -
o -
r o
r o - r o r o - r - r o - r o
t t
t
p
ij ij
t t
t
p
ij ij
e
ij ij
t
t t
e
ij ij
t t
t
p
ij
e
ij ij
t
e
ij ij T
dt d
dt d dt d dt d dt d d dt d dW
1
1
1
1
*
1
1
1
1
0

(9.205)
The work done by
*
ij
o , is
}
o -
r o
t t
t
p
ij ij
dt d dW
1
1
*
0
, and therefore,
( )
} } }
o - o - o -
r o - o r o - r o -
t t
t
p
ij ij ij
t t
t
p
ij ij
t t
t
p
ij ij T n
dt d dt d dt d dW dW dW
1
1
1
1
1
1

* *
0

(9.206)
According to Druckers stability criterion, the following must be satisfied
( ) 0 0
1
1
*
> r o - o
}
o - t t
t
p
ij ij ij n
dt d dW
(9.207)
which represents Druckers second postulate and because the above integrand is valid at
any time, it holds that:
( ) 0
*
r o - o
p
ij ij ij
d ( ) 0
*
-
p
: (9.208)
Then, with
*
ij ij ij
d o - o o we obtain:
0 r o
p
ij ij
d d 0
p
d d : (9.209)
Under these conditions the material is said to be plastically stable. For a material with
hardening behavior we obtain:
0 ; 0 > > r o
p p
ij ij
d d d d : (9.210)
(
0
t )
(
1
t )
o o
initial yield surface
o - o d
Evolution of the yield surface
due to o d
( t t o -
1
)
9 PLASTICITY

517
and for a material with perfect plasticity behavior we obtain:
0 ; 0 r o
p p
ij ij
d d d d : (9.211)
Let us now suppose that there is a scalar-valued tensor function ) ( F F called the
plastic potential or yield function. In the elastic regime the following is satisfied:
0 ; 0 < o
o o
o
<
ij
ij
d d
F
F F
(9.212)
and in the plastic zone:
0 ; 0 o
o o
o

ij
ij
d d
F
F F
(9.213)
Then, from the condition in (9.209) the plastic energy becomes:
0 r o
p
ij ij
d d (9.214)
with the restriction 0 o
o o
o

ij
ij
d d
F
F on the yield surface.
Now, to solve a problem with a restriction we will introduce the Lagrange multiplier / d :
0 0 o
|
|
.
|

\
|
o o
o
- r o
o o
o
- r o
ij
ij
p
ij ij
ij
p
ij ij
d d d d d d d
F F
(9.215)
This condition must be fulfilled for any arbitrary value of
ij
do , thus:
;
ij
p
ij
d d
o o
o
r
F

o
o

F
d d
p
Prandtl-Reusss flow rule (9.216)
where d is a positive scalar. The equation (9.216) is called Prandtl-Reusss flow rule.
Note that in an isotropic material the yield surface can be expressed in terms of the three
principal stresses, i.e. ) , , (
3 2 1
o o o F F . In this case,
p
d can be represented by a vector in
the principal stress space, (see Figure 9.41).
Note also that F
F

V =
o
o
is the
gradient of F in the principal stress
space and by definition is normal to the
yield surface. Then, the plastic flow
vector,
p
d , is also normal to the yield
surface, so, together with the normality
condition, we can conclude that the
yield surface must be convex, since by
the condition 0 r o
p
d d the angle
formed by o d and
p
dr cannot be
obtuse.
Figure 9.41: Normality condition principal
stress space.
o d

p
dr
F

V
o d
0 F

1
o

2
o
.
0 < F
NOTES ON CONTINUUM MECHANICS

518
9.6 Plasticity in Large Deformation Regime
Several theories have been developed for describing plasticity in the field of large
deformation, among which we can mention:
That based on the multiplicative decomposition of the deformation gradient
proposed by Lee(1969) in the field of Solids Mechanics:
) , ( ) , ( ) , ( t t t
p e
X F X F X F
, , ,

Multiplicative decomposition of the
deformation gradient
(9.217)
That based on the additive decomposition of the Green-Lagrange strain tensor,
proposed by Green & Naghdi(1965):
) , ( ) , ( ) , ( t t t
p e
X E X E X E
, , ,
-
Additive decomposition of the Green-
Lagrange strain tensor
(9.218)
That based on the additive decomposition of the rate-of-deformation tensor,
proposed by Nemat-Nasser(1982):
) , ( ) , ( ) , ( t t t
p e
x x x
, , ,
D D D -
Additive decomposition of the rate-of-
deformation tensor
(9.219)
In the next subsection we will look at approaching large-deformation plasticity by means of
the multiplicative decomposition of the deformation gradient caused by two
transformations, namely, the elastic and the plastic transformations, Lee (1969), Simo
(1992), Simo&Hughes (1998).
9.7 Large-Deformation Plasticity Based on the
Multiplicative Decomposition of the
Deformation Gradient
9.7.1 Kinematic Tensors
The multiplicative decomposition of the deformation gradient is given by:
) , ( ) , ( ) , ( t t t
p e
X F X F X F
, , ,
Multiplicative decomposition (9.220)
where
e
F is the elastic transformation, and
p
F is the plastic transformation, (see Figure
9.42). Then, according to Figure 9.42 the following is satisfied:
X F F X F x
, ,
,
d d d
p e

(9.221)
Note that, first we make the transformation related to
p
F , thereby defining a new
configuration called the intermediate (or stress-free) configuration in which it holds that
X F X
,
,
d d
p
. Then, we make the transformation associated with
e
F , where X F x
,
,
d d
e

holds, (see Figure 9.42). Next, from the multiplicative decomposition we can obtain the
following relationships:
1
1
1 1
1
-
-
- -
-

p e e p p e
F F F F F F F F F
(9.222)
9 PLASTICITY

519
Next we will establish the kinematic variables in the intermediate configuration B , and
show how these are related to those defined in the reference and current configurations.










Figure 9.42: Multiplicative decomposition of the deformation gradient.
It is simple to show that the intermediate configuration is not unique, since here we can
apply an orthogonal transformation (rotation) which remains in a stress-free state. In this
scenario the deformation gradient can be represented by
p e p e
F F F F F

, where
T e e
Q F F

,
p p
F F Q

, (see Figure 9.43).

















Figure 9.43: Non-uniqueness of the multiplicative decomposition of the deformation
gradient.


p p
F F Q


B
X
,


p
F

T e e
Q F F



p e
F F F
current
configuration
intermediate configuration I
X
,

reference
configuration

0
B
X
,
d
x
,

B
x
,
d
X
,
d
intermediate
configuration II
B
X
,

X
,
d

T
Q

e
F
Q
X
,

x
,


p
F

e
F

p e
F F F
reference
configuration
current
configuration
0
B
B
X
,
d
x
,
d
B
X
,

intermediate
configuration
X
,
d
NOTES ON CONTINUUM MECHANICS

520
9.7.1.1 Deformation and Strain Tensors
Now, remember from Chapter 2 that the right Cauchy-Green deformation tensor
( ) , ( t X C
,
), the left Cauchy-Green deformation tensor ( ) , ( t x b
,
), the Green-Lagrange strain
tensor ( ) , ( t X E
,
), the Cauchy deformation tensor ( ) , ( t x c
,
), the Almansi strain tensor
( ) , ( t x e
,
), the right stretch tensor ( ) , ( t X
,
U ), and the left stretch tensor ( ) , ( t x
,
V ) are related
to each other as shown in Figure 9.44.












Figure 9.44: Kinematic tensors.
The right Cauchy-Green deformation tensor (reference configuration) is defined by
F F X C
T
t) , (
,
, and the left Cauchy-Green deformation tensor by
2
) , ( V
T
t F F x b
,

(current configuration). If we now consider the reference and intermediate configuration
brought about by the transformation
p
F , (see Figure 9.42), we can define the following
tensor:
p
T
p p
t F F X C ) , (
,
Plastic part of the right Cauchy-
Green deformation tensor
(reference configuration)
(9.223)
and its inverse:
T
p p p
t
- - -
F F X C
1 1
) , (
,

(9.224)
with which we can define the plastic part of the Green-Lagrange strain tensor in the
reference configuration with:
( ) |
.
|

\
|
- - 1 1
p
T
p p p
t F F C X E
2
1
2
1
) , (
,
Plastic part of the Green-Lagrange
strain tensor
(reference configuration)
(9.225)
Then, we can also define:
2
) , (
p
T
p p p
t V F F X b
,
Plastic part of the left Cauchy-
Green deformation tensor
(intermediate configuration)
(9.226)
X
,

x
,

F
Current configuration

0
B
B
Reference configuration

( ) 1
U
-


- - -

C X E
C F F X B
F F X C
2
1
) , (
) , (
) , (
1 1
2
t
t
t
T
T
,
,
,


( )
( )
1
1 1
2 1
2
1
2
1
) , (
) , (
) , (
-
- - -
-
-
-

b
c x e
b F F x c
c F F x b
1
1
V
t
t
t
T
T
,
,
,


1
1 1 1
1
- -
- - -
-


F E F e
F C F b
F C F b
T


F e F E
F b F C
F b F C


- - -
-
T
1 1 1
1

9 PLASTICITY

521
Note that ) , ( t
p
X C
,
and ) , ( t
p
X E
,
are defined in the reference configuration while
) , ( t
p
X b
,
is defined in the intermediate configuration, (see Figure 9.45). The Almansi strain
tensor, defined in the intermediate configuration, is given by:
|
.
|

\
|
- |
.
|

\
|
-
- - -

1 1
2
1
2
1
) , (
p
T
p p p
t F F b X e 1 1
,
Plastic part of the Almansi strain
tensor
(intermediate configuration)
(9.227)
In Chapter 2 we obtained the relationship between E and e given by F e F E
T
, (see
Figure 9.44). Then, in comparison, the tensors ) , ( t
p
X E
,
and ) , ( t
p
X e
,
are interrelated, (see
Figure 9.42), by:
p p
T
p p
F e F E

(9.228)
Proof of which follows:
p p p
T
p
p p p
T
p
p
T
p p p
T
p
p p
T
p
T
p p
T
p p p
T
p
p p
T
p
T
p p p
T
p
p
T
p p
p
T
p p
E F e F
E F e F
F F F e F
F F F F F F F e F
F F F F F e F
F F e
F F e


-
-
|
.
|

\
|
-
-
|
.
|

\
|
-





- -
- -
- -
- -
2 2
2
2
2
2
2
1
1
1
1
1
1
1
1
1
1

(9.229)
If we now consider the transformation,
e
F , between the configurations B (intermediate
configuration) and B (current configuration), we can define the following tensors:
2
) , (
e e
T
e e
t U F F X C
,
(intermediate configuration) (9.230)
( ) |
.
|

\
|
- - 1 1
e
T
e e e
t F F C X E
2
1
2
1
) , (
,
(intermediate configuration) (9.231)
T
e e e
t F F x b ) , (
,
(current configuration) (9.232)
|
.
|

\
|
- |
.
|

\
|
-
- - -

1 1
2
1
2
1
) , (
e
T
e e e
t F F b x e 1 1
,
(current configuration) (9.233)
Then, by considering the Almansi strain tensor in the current configuration, ) , ( t x e
,
, we can
define a new tensor E in the intermediate configuration as follows:
e
T
e
F e F E
(9.234)
We can now find the relationship between the tensors E ,
p
E and
e
E starting from the
definition of the Green-Lagrange strain tensor:
( ) ( )
ij kj ki ij
components T
F F E c - -
2
1
2
1
1 F F E (9.235)
NOTES ON CONTINUUM MECHANICS

522
and by considering the multiplicative decomposition,
p
kj
e
ik ij
F F F , the Green-Lagrange
strain tensor becomes:
( )
ij
p
tj
e
kt
p
si
e
ks ij
F F F F E c -
2
1
(9.236)
Then, from the equation in (9.231) we obtain
st
e
st
e
kt
e
ks
E F F c - 2 which by substituting into
the equation in (9.236) yields the following:
( ) [ ] [ ]
[ ]
p
ij
p
tj
e
st
p
si ij
p
sj
p
si
p
tj
e
st
p
si
ij
p
tj
p
si st
p
tj
p
si
e
st ij
p
tj
p
si st
e
st ij
E F E F F F F E F
F F F F E F F E E
- - -
- - - -
c
c c c c
2
1
2
2
1
2
2
1

(9.237)
Then, the above equation in tensorial notation becomes:
p p e
p p e
T
p
E E
E F E F E
-
-
) _ (
(9.238)
where we have defined a new tensor in the reference configuration:
p e
T
p p e
F E F E
) _ (

(9.239)
The rate of change of (9.238) is given by:
p p e
E E E
` ` `
-
) _ (

(9.240)
Let us see what we can obtain from the expression
p e
e E - . Now, if we use the
relationships in (9.231) and (9.227) we obtain:
|
.
|

\
|
- |
.
|

\
|
- - |
.
|

\
|
- -
- - - -

1 1
2
1
2
1
2
1
p
T
p e
T
e p
T
p e
T
e p e
F F F F F F F F e E 1 1 (9.241)
Then, without altering the above outcome, we can apply the dot product of 1
-

T
e
T
e
F F
on the left, and the dot product of 1
-
e e
F F
1
on the right, so, we obtain:
( )
e
T
e e T
T
e
e e p
T
p
T
e e e
T
e
T
e
T
e p e
F e F F F F F
F F F F F F F F F F e E


-
|
.
|

\
|
- -
- -
- - - - - -
1
1 1 1
2
1
2
1
1

(9.242)
Then, from the equations in (9.242) and (9.234) we can conclude that:
p e e
T
e
e E F e F E -

(9.243)
If we now consider the equations in (9.232) and (9.224), we can obtain the relationship
between
1 -
p
C and
e
b , i.e.:
T p T
T
p p
T
p p
T
e e e
F C F F F F F F F F F F F b
- - - - -
|
.
|

\
|

1 1 1 1
(9.244)
T p e
F C F b
-

1

(9.245)
and the trace of
e
b can be evaluated as follows:
9 PLASTICITY

523
( )
) (
) (
1 1
1 1 1 1
- -
- - - -



|
.
|

\
|
|
.
|

\
|

p p
p T p
kp jp jk jp
p
kp ik ij
T p e
C F F F C F
C C C C
C F F F C F b
Tr
Tr
:
: : c 1
(9.246)
Next, we can obtain the relationship between the tensors ) , ( t
p
X e
,
and ) , ( t
e
x e
,
:
( ) ( ) ( )
( ) |
.
|

\
|
-
|
|
.
|

\
|
|
.
|

\
|
-
(

- -
- - - - -
-
- -
- -


1 1 1
1
1
1
2
1
2
1
2
1
2
1
e p
T
p
T
e p e
T
e
T
p
p e
T
p e T
F F F F F F F F
F F F F F F e
1 1
1 1
(9.247)
Now, by considering that ( )
p p
T
p
e F F 2
1
-
- -
1 , (see equation (9.227)), we can obtain:
( )
_
e
e
T
e e p
T
e
e p
T
e e
T
e
e p
T
e e p
T
p
T
e
e
F F F e F
F e F F F
F e F F F F F e

|
.
|

\
|
- -
|
.
|

\
|
- -
|
.
|

\
|
- - |
.
|

\
|
-
- - - -
- - - -
- - - - - -



1 1
1 1
1 1 1
2
1
2
2
1
2
2
1
2
1
1
1
1 1 1

(9.248)
Then, we can define a new tensor:
1
) _ (
- -

e p
T
e e p
F e F e

(9.249)
with which the Almansi strain tensor can also be defined as:
e e p
e e p
T
e
e e
e F e F e
-
-
- -

) _ (
1
(9.250)
Thus
( ) |
.
|

\
|
- |
.
|

\
|
- - - -
-
- -
- 1
1 1
1 ) _ (
2
1
2
1
2
1
b b b b e e e
e e e e p
1 1 (9.251)
Then, from the equation in (9.249), and taking into account that
1 1 - -

p e
F F F , we
obtain:
F e F F e F
F F e F F e
F e F e




- -
- -
p T p e p
T
p
p p T
T
p e p
e p
T
e e p
) _ (
1
) _ (
1
) _ (

(9.252)
and by comparing the above equation with that in (9.228) we can conclude that:
p e p
T
p p
F e F E
) _ (

(9.253)
We can now appreciate all the relationships obtained above in Figure 9.45.




NOTES ON CONTINUUM MECHANICS

524
























Figure 9.45: The kinematic tensors Multiplicative decomposition.
9.7.1.2 Area and Volume Elements Deformation
With the definition of the Jacobian determinant and the multiplicative decomposition of
the deformation gradient, we can obtain:
p e p e p e
J J J ) ( ) ( ) ( ) ( F F F F F det det det det (9.254)
which thus defines the plastic Jacobian determinant
p
J and the elastic Jacobian
determinant
e
J , respectively, as:
[ ]
2
1
2
1
1
) ( ) ( ; ) ( ) (
e e e p p p
J J b F C F det det det det
(


-

(9.255)
Then, the differential volume elements in the respective configurations, (see Figure 9.46),
are given by:
) , ( ) , ( ; ) , ( ) , (
0
t V d J t dV t dV J t V d
e p
X x X X
,
,
,
,

(9.256)

( )
1
1 1
2
2
1
) , (
) , (
) , (
-
- - -
-

b x e
b F F x c
F F x b
1
V
t
t
t
T
T
,
,
,

( )
p e e
T
e
e e
e e
T
e e
t
t
e E F e F E
C X E
F F X C
-
-

1
U
2
1
) , (
) , (
2
,
,

|
.
|

\
|
-


-
- - -

1
1 1
2
2
1
) , (
) , (
) , (
p p
p
T
p p
p
T
p p p
t
t
t
b X e
F F X b
F F X b
1
V
,
,
,

( )
p p e
p e
T
p p e
p p
T
p p p
p
T
p p
t
t
t
E E E
F E F E
C X E
F F X C
F F X C
-

-

- - -
) _ (
) _ (
1 1
2
1
) , (
) , (
) , (
1
,
,
,

B
X
,

X
,


p
F

e
F

p e
F F F
reference
configuration

0
B
intermediate configuration
Current Configuration Reference configuration

( ) 1
U
-


- - -

C X E
C F F X B
F F X C
2
1
) , (
) , (
) , (
1 1
2
t
t
t
T
T
,
,
,

Intermediate configuration
x
,

current
configuration
B

e e p
e p e e p
e e
e e e
F e F e
b e
T
-

|
.
|

\
|
-
- -
-

) _ (
1
) _ (
1
2
1
1


T p
T
e e e
F C F F F b
-

1

Current Configuration
Intermediate configuration
Reference configuration
9 PLASTICITY

525














Figure 9.46: Deformation of the differential volume and area elements.
Let us consider now a differential area element in the reference configuration
0
A
,
d , (see
Figure 9.46), and by considering the multiplicative decomposition we obtain:
0
A F A
,
,
d J d
T
p p

The differential area element in


the intermediate configuration
(9.257)
A F a
,
,
d J d
T
e e

The differential area element in


the current configuration
(9.258)
Remember that the transformation between the
0
A
,
d and a
,
d is given by Nansons formula,
0
A F a
,
,
d J d
T

-
, (see Chapter 2), which can be validated by:
( )( )
0 0
0 0
A F A F F
A F F A F F A F a
, ,
, ,
,
,
d J d J J
d J J d J J d J d
T
T
p e p e
T
p
T
e p e
T
p p
T
e e
T
e e


-
-
- - - - -

|
.
|

\
|

(9.259)
9.7.1.3 The Spatial Velocity Gradient
From the definition of the spatial velocity gradient, i.e.
1 -
F F
`
l , we can introduce the
corresponding tensors that are brought about by the transformations
p
F and
e
F , that is:
1 1
) , ( ; ) , (
- -

e e e p p p
t t F F x F F X
`
,
`
,
l l

(9.260)
Then, also based on the equation l
- -
-
1 1
F F
`
, (see Chapter 2), we can introduce:
e e e p p p
l l
- - - -
- -
1 1 1 1
; F F F F
` `

(9.261)
with which it is also possible to represent the spatial velocity gradient as:

0
A
,
d
B
X
,

x
,

X
,

0
dV J V d
p


e
F

p e
F F F
reference
configuration
current
configuration

0
B
B
intermediate
configuration

p
F

0
dV
dV
V d
V d J dV
e


0
dV J dV
A
,
d

0
A F A
,
,
d J d
T
p p

-

A F a
,
,
d J d
T
e e

-

a
,
d
NOTES ON CONTINUUM MECHANICS

526
( ) ( ) ( )
1 1 1 1 1 1 1
1 1 1
1
- - - - - - -
- - -
-


- -
|
.
|

\
|
-
e p p e e e e p p e e p p e
e p p e p e p e p e
p
Dt
D
F F F F F F F F F F F F F F
F F F F F F F F F F F F
_
` ` ` `
` ` `
l
l

(9.262)
Thus, we can draw the conclusion that:
) _ (
1
e p e e p e e
l l l l l - -
-
F F

(9.263)
Note that,
e
l and
) _ ( e p
l are defined in the current configuration, whereas
p
l is in the
intermediate configuration where the rate-of deformation and spin tensors are also
established:
|
.
|

\
|
- |
.
|

\
|
-
T
p p p
T
p p p
l l l l
2
1
;
2
1
W D (9.264)
Note that
1
) _ (
) , ( ) , (
-

e p e e p
t t F X F x
,
,
l l and its inverse
e e p e p
F F
-

) _ (
1
l l , with
which we can establish the following relationships:
e e
e e e e
e e e e e e e p e p e e
F F
F F F F
F F F F F F
`
`



-
- -
- - -


1
1 1
1
) _ ( ) _ (
1 1
; ; l l l l l l

(9.265)
1
) _ (
1
) _ (
1
; ;
- - -

e e e e e e p e e p e e
F F F F F F l l l l l l
(9.266)
and
( )
p e e
p e e e e p e e e e e e
e e p e e e e e e p e e e e
l l
l l
l l l l l l
-
- |
.
|

\
|
-
- -


- - - - -
- - - -
) _ (
1 1 1 1 1
) _ (
1 1
) _ (
1 1
F F F F F F F F F F
F F F F F F F F
` `

(9.267)
We can define ) , ( t
p
x
,
D in the current configuration as follows:
( )
1 1
1
1 1
1
) _ ( ) _ ( ) _ (
2
1
2
1
2
1
2
1
2
1
- - - -
- -
- - - -
- -




|
.
|

\
|
-
|
.
|

\
|
-
|
.
|

\
|
-
|
.
|

\
|
- |
.
|

\
|
-
e
sym
p e
T
e e e
T
p p e
T
e
e e
T
e
T
p p e
T
e
T
e
e e
T
e
T
p
T
e
T
e e e p e
T
e
T
e
T
e
T
p
T
e e p e
T
e p e p
sym
e p p
F C F F C C F
F F F F F F
F F F F F F F F F F
F F F F
l l l
l l
l l
l l l l l D

(9.268)
We can also verify the following relationship:
( ) ( ) ( )
) _ (
) _ ( ) _ (
) _ ( ) _ (
2
1
2
1
2
1
2
1
e p e
T
e p e p
T
e e
T
e p e e p e T
D D
D
-
|
.
|

\
|
- - |
.
|

\
|
-
(

- - - -
l l l l
l l l l l l

(9.269)
where
e
D can also be represented by:
9 PLASTICITY

527
1
1
2
1
2
1
2
1
- -
- -


|
.
|

\
|
-
|
.
|

\
|
- |
.
|

\
|
-
e e
T
e e
T
e
T
e
T
e
T
e e e
T
e e e
F F F F F F
F F F F
` `
` `
l l D

(9.270)
If we consider that
e
T
e e
T
e e
F F F F C
` `
`
-
(9.271)
we can conclude that:
1
2
1 - -

e e
T
e e
F C F
`
D (9.272)
We can now appreciate all the relationships obtained above in Figure 9.47.

























Figure 9.47: The rate of change of the deformation tensors Multiplicative decomposition.

p e e
e e e e e e e
l l l
l l
-

- -
) _ (
1 1
) _ (
F F F F
`

|
.
|

\
|
-
|
.
|

\
|
-

-

T
p p p
T
p p p
p p p
t
l l
l l
l
2
1
2
1
1
) , (
W
D
F F X
`
,

Intermediate configuration
B
X
,

X
,


p
F

e
F

p e
F F F
reference
configuration

0
B
intermediate configuration
( )
( )
T
T
t
l l
l l
l
-
-

-

2
1
2
1
1
) , (
W
D
F F x
`
,

x
,
current
configuration
B

) _ (
) _ ( ) _ (
2
1
) _ (
2
1
) _ (
1
) _ (
1
) , (
e p e
T
e p e p e p
T
e e e
e p e
e p e e p
e e e
t
D D D
D
D
-
|
.
|

\
|
-
|
.
|

\
|
-
-

-
-

l l
l l
l l l
l l
l
x
F F
F F
,
`

Intermediate configuration
Current configuration
NOTES ON CONTINUUM MECHANICS

528
9.7.1.4 The Oldroyd Rate
By using the Oldroyd rate,
T
l l - - T T T T
`

, (see Chapter 4), we can define the rate of


change of an arbitrary second-order tensor T in the intermediate configuration as:
T
p p
l l - - T T T T
`
Oldroyd rate
(intermediate configuration)
(9.273)
Starting from said definition we can obtain the Oldroyd rate of the elastic part of the left
Cauchy-Green deformation tensor as:
T e e e e
l l - - b b b b
`


(9.274)
NOTE: In the literature, e.g. Marsden&Hughes (1983), we can find that

e
b is denoted by
the Lie derivative of
e
b
`
.
If we now consider that
T p e
F C F b
-

1
, (see equation (9.245)), we can obtain the rate
of change of
e
b as follows:
T p T p T p e
F C F F C F F C F b
` ` ` `

- - -
- -
1 1 1

(9.275)
where F F l
`
, thus:
T e T p e
T T p T p T p e
l l
l l


- -
- -
-
- - -
b F C F b
F C F F C F F C F b
1
1 1 1
`
` `

(9.276)
with which we can obtain:
T e e e T p
l l - -
-
b b b F C F
` `
1

(9.277)
Then, by comparing the above equation with (9.274) we can conclude that the Oldroyd rate
of
e
b is given by:
T p e
F C F b
-

1
`


(9.278)
Note that if
T
l l - b b b
`
, it follows that the Oldroyd rate of b becomes the zero
tensor, (see Problem 4.1), i.e.:
( ) 0 - - - - -
T T T
l l l l l l b b b b b b b b
`


(9.279)
Now, it was proven in Chapter 2 that e e e - -
T
l l ` D , and if we now consider the
definition of

e , (see equation (9.273)), we can conclude that:


( )
( )
D
D
D

- - - -
- - - - - -


T T
T T T
l l l l
l l l l l l
e e e e
e e e e e e e e `


(9.280)
The Oldroyd rate of |
.
|

\
|
- |
.
|

\
|
-
- - -

1 1
2
1
2
1
e e e e
F F b e
T
1 1 is given by:
( )
e
T
e e e e e T e e T e e e e
D D - - - - - - l l l l l l e e e e e e e e `


(9.281)
9 PLASTICITY

529
Thus, by starting from the equation
e e p
e e e -
) _ (
, (see equation (9.250)), we can obtain:
) _ ( ) _ ( e p e e e p
D D D - -

e e e
(9.282)
Then, according to (9.251), i.e. |
.
|

\
|
-
-
-
1
1
) _ (
2
1
b b e
e e p
, we can obtain:
( )

|
.
|

\
|
- |
.
|

\
|

-
-
-

1
1
1
) _ (
2
1
2
1
e e e p
b b b e
0
(9.283)
9.7.1.5 The Cotter-Rivlin Rate
By using the Cotter-Rivlin rate of a tensor, i.e. l l - -
A
T T T T
T
`
, we can define said
rate for ) , ( t X
,
T in the intermediate configuration as follows:
p
T
p
l l - -
A
T T T T
`
Cotter-Rivlin rate
(intermediate configuration)
(9.284)
Then, if we consider both the Cotter-Rivlin rate of the Almansi strain tensor
l l - -
A
e e e e
T
` , and the relationship obtained in Chapter 2 l l - - e e e
T
` D we can
draw the conclusion that D
A
e , (see Problem 4.2).
Another expression obtained in Chapter 2 is
1 - -
F E F
`
T
D , with which we define:
1 - -
A

p p
T
p p p
F E F e
`
D
(9.285)
Then, by starting from the equation
1 - -
F E F
`
T
D we obtain:
1 1
1
- - - -
- -




p
T
p p T
T
p
T
T
F E F F F F F
E F F
F E F
`
`
`
D
D
D
(9.286)
Next, by considering the multiplicative decomposition
1 -

p e p e
F F F F F F , the
above equation becomes:
1 - -

p
T
p e
T
e
F E F F F
`
D
(9.287)
and if we consider that
p p e
E E E
` ` `
-
) _ (
, (see equation (9.240)), into the above expression
we obtain:
( )
A
- -
- - - - - -
-
- -


p p p e
T
p
p p
T
p p p e
T
p p p p e
T
p e
T
e
e F E F
F E F F E F F E E F F F
1
) _ (
1 1
) _ (
1
) _ (
`
` ` ` `
D

(9.288)
where we have applied the equation in (9.285). Then, the term
) _ ( p e
E
`
can be obtained by
means of the expression of
) _ ( p e
E given in (9.239), i.e.:
p e
T
p p e
F E F E
) _ (
, thus:
p e
T
p p e
T
p p e
T
p p e
F E F F E F F E F E
`
`
` `
- -
) _ (

(9.289)
and
NOTES ON CONTINUUM MECHANICS

530
1
1 1 1
) _ (
- -
- - - - - -


-
-
p p e
T
p
T
p
p p e
T
p
T
p p p e
T
p
T
p p p e
T
p
F F E F F
F F E F F F F E F F F E F
`
`
` `

(9.290)
Remember that
1 -

p p p
F F
`
l , so the above equation becomes:
p e e e
T
p p p e
T
p
l l - -
- -
E E E F E F
`
`
1
) _ (

(9.291)
Then, by comparing the above with (9.284) we can conclude that:
A
- -
- -
e p e e e
T
p p p e
T
p
E E E E F E F l l
`
`
1
) _ (

(9.292)
and by substituting the above into (9.288) we obtain:
A A
- -
-
p e p
T
p e
T
e
e E F E F F F
1
`
D
(9.293)
In the equation in (9.243) we obtained
p e e
T
e
e E F e F E - , thus we can conclude
that:
A A A
- -
- E e E F E F F F
p e p
T
p e
T
e
1
`
D

(9.294)
Now, from (9.291) we can express
e
E
`
as follows:
p e e
T
p p p e
T
p e
l l - -
- -
E E F E F E
1
) _ (
`
`

(9.295)
and if we consider that
p p e
E E E
` ` `
-
) _ (
, we obtain:
p e e
T
p p p
T
p p
T
p e
l l - - -
- - - -
E E F E F F E F E
1 1
` `
`

(9.296)
Then, by starting from |
.
|

\
|
- 1
p
T
p p
F F E
2
1
we can obtain its rate of change as follows
|
.
|

\
|
-
p
T
p p
T
p p
F F F F E
` ` `
2
1
, so,
1 1
2
1
2
1 - - - -
-
p p
T
p
T
p p p
T
p
F F F F F E F
` ` `
. Then
the equation in (9.296) becomes:
( ) ( )
(

- -
(

- - - -
|
.
|

\
|
- - - -
|
.
|

\
|
- - - -




- -
- -
- -
- - - -
p e e
T
p p
T
p
p e e
T
p p
T
p
p e e
T
p p
T
p p
T
p
p e e
T
p p p
T
p
T
p p
T
p e
l l
l l
l l l l
l l
C C F E F
E E F E F
E E F E F
E E F F F F F E F E
2
1
2 2
2
1
2 2
2
1
2 2
2
1
1
1
1
1 1
`
`
`
` ` `
`
1 1
(9.297)
Remember that
T
e e
C C , so, the above becomes:
( )
sym
p e p
T
p e
l -
- -
C F E F E
1
`
`

(9.298)
We can also represent the equation in (9.298) as follows:
( )
1 - -

(

-
p p
sym
p e
T
p
T
p e
F F C F E F E l
`
`
(9.299)
9 PLASTICITY

531
9.7.2 The Stress Tensors
Remember from Chapter 3 that the Cauchy stress tensor ) , ( t x
,
, the Kirchhoff stress
tensor ) , ( t x
,
, the first Piola-Kirchhoff stress tensor ) , ( t X
,
P , the second Piola-Kirchhoff
stress tensor ) , ( t X
,
S , the Biot stress tensor ) , ( t X
,
T , and the Mandel stress tensor ) , ( t X
,
M
are related to each other as indicated in Figure 9.48.
Remember that the tensors F and P are two-point tensors (pseudo-tensors), i.e. they are
not defined in any configuration.













Figure 9.48: The stress tensors.
We can define the second Piola-Kirchhoff stress tensor in the intermediate configuration
as:
T
e e
t
- -
F F X
1
) , (
,
S
(9.300)
Then, from the above equation we can obtain:
T T
T
e e
T
e e - - - -
F F F F F F F F
1 1
S S
(9.301)
Then, if we consider the multiplicative decomposition
p e
F F F , we can obtain the
following equations
1
1
1 1
1
-
-
- -
-

p e e p
F F F F F F , thus we can rewrite (9.301) as
follows:
T
p p
T
T
p p
F F
F F F F


- -
- -
S S
S S
1
1
(9.302)
Then, if we consider the Mandel stress tensor,
T T T -
F F F C P S M , (see
Figure 9.48), and the equation in (9.302), we can define the former in the intermediate
configuration as:
S S S M |
.
|

\
|

- -
e e
T
e
T
e
T
e e
T
e
T
e
T
e
C F F F F F F F F
(9.303)

T T
T T T
T T
T T
J
J
J
-
-
-
- - - -




F
F F F C
F F
F F F F
R T
P S M
M
S



1 1


X
,

x
,

F
reference
configuration
current
configuration

0
B
B
Reference configuration

S M
T
S
C X
X
X
) , (
) , (
) , (
t
t
t
,
,
,

) , (
) , (
J t
t
x
x
,
,



T
J
-
F P
Current configuration
NOTES ON CONTINUUM MECHANICS

532















Figure 9.49: The stress tensors Multiplicative decomposition.
9.7.2.1 Stress Tensor Rates
Starting from the equation in (9.302), i.e.
T
p p
F F S S , we can obtain the rate of
change of S as follows:
T
p
T
p
T
p
T
p p p p p
T
p
T
p
T
p p p p
T
p p
T
p p
T
p p
F F F F F F
F F F F F F
F F F F F F



|
.
|

\
|
- -
|
.
|

\
|
- -
- -
- -
- -
l l S S S
S S S
S S S S
`
` ` `
` ` `
`
1
1

(9.304)
where we have considered that
p p p
F F l
`
. In addition, we have:
p p p p p p p p
F F F F F F |
.
|

\
|

- - - 1 1 1
`
l Z
(9.305)
the equation in (9.304) can be rewritten as follows:
T
p
T
p p p
F F |
.
|

\
|
- - Z S S Z S S
`
`

(9.306)
Now, by starting from the relation
T
p p
- -
F F S S
1
we can evaluate S
`
as follows:
T
p p
T
p p
T
p p
- - - - - -
- - F F F F F F
`
`
` `
S S S S
1 1 1

(9.307)
Then, if we consider (9.261), i.e.
p p p
l
- -
-
1 1
F F
`
, and substitute this into the equation in
(9.307), we can obtain:
T
p
T
p p p
T
p
T
p p
T
p p
T
p p p
- - -
- - - - - - -


|
.
|

\
|
- -
- - -
F F
F F F F F F
l l
l l
S S S
S S S S
`
`
`
1
1 1 1

(9.308)
X
,

x
,

X
,


p
F
e
F

p e
F F F
reference
configuration
current
configuration

0
B
B
B
intermediate
configuration

S M
S

- -
e
T
e e
t
t
C X
F F X
) , (
) , (
1
,
,


Reference configuration

S M
T
P
S
C X
X
X
X
) , (
) , (
) , (
) , (
t
t
t
t
,
,
,
,


T
e e
t F F x S ) , (
,

) , (
) , (
J t
t
x
x
,
,


Intermediate configuration
Current configuration
Current configuration
9 PLASTICITY

533
Now, if we remember the Oldroyd rate of S , i.e.
T
p p
l l - - S S S S
`

, the above
expression becomes:
T
p p
- -
F F

S S
1
`

(9.309)
9.7.3 The Helmholtz Free Energy
The Helmholtz free energy (per unit volume) is given by:
) , , ( T F y y (9.310)
For temperature-independent processes the above expression is reduced to:
) , ( F y y (9.311)
Then, if we consider the multiplicative decomposition of the deformation gradient we can
adopt the intermediate configuration to define the energy expression. Note that the
intermediate configuration is a stress-free one, i.e. it is elastically unloaded. Then, the energy
function can be defined as:
) , , (
p
F F y y (9.312)
Considering that ) , (
p e e
F F F F , the energy can be written in terms of
) , (
e
F 1 1 (9.313)
9.7.3.1 Decoupling the Helmholtz Free Energy
The Helmholtz free energy can be approached additively by two parts. One part is caused
by the effect of
e
F and the other part is caused by the effect of , i.e.:
) ( ) (
p e e
1 1 1 - F (9.314)
One advantage of this decoupling is that we can treat the elastic part of the energy,
) (
e e
F 1 , in the same way as when we considered hyperelastic material which was discussed
in Chapter 8.
9.7.3.2 The Objectivity Principle for the Helmholtz Free Energy
As we discussed in Chapter 6, the constitutive equation must satisfy certain principles
including the principle of objectivity and as the energy is a scalar, it satisfies this principle,
i.e.
*
1 1 . However, we can also use this principle to express the energy ) , (
e
F 1 1 in
terms of other parameters and according to it, (see Chapter 4), the following holds:
) , (
*
*
*

e
F 1 1
(9.315)
The free variables
*
can be scalars, vectors, or second-order tensors which fulfills the
following law of transformation c c
*
(scalar),
, ,
Q
*
(vector), and
T
Q Q
*

(Eulerian second-order tensor). For the sake of simplicity, let us consider that
*
is a
scalar
*
. Then, with respect to
*
e
F , remember that the deformation gradient does
not obey the second-order tensor transformation law, and is given by
e e
F F Q
*
. In
NOTES ON CONTINUUM MECHANICS

534
addition, if we consider the polar decomposition of
e
F , (see Figure 9.50), then
e e e e e
R V U R F is valid, thus:
) , ( ) , ( ) , (
*
*

e e e e
U R Q Q 1 1 1 1 F F
(9.316)
As the tensor Q can be any orthogonal tensor, we can adopt
T
e
R Q with which we
obtain:
) , ( ) , (
e e e
T
e
U U R R 1 1 1
(9.317)
That is, the energy function can also be expressed in terms of the right stretch tensor of the
intermediate configuration, (see Figure 9.50). Remember also that the right Cauchy-Green
deformation tensor
e
C is related to
e
U by the following equation
2
e e
U C , so the
energy can also be expressed in terms of tensor
e
C :
) , (
e
C 1 1 (9.318)
Moreover, by considering that the Green-Lagrange strain tensor
e
E is related to
e
C by
) ( 2 1 -
e e
C E , we can still express the energy as:
) , (
e
E 1 1 (9.319)
9.7.3.3 The Isotropic Helmholtz Free Energy
Remember that a scalar-valued isotropic tensor function can be written in terms of the
principal values, i.e.:
) , , , ( ) , (
2
3
2
2
2
1

e e e e
/ / / 1 1 1 C
(9.320)
where
e
1
/ ,
e
2
/ ,
e
3
/ are the principal values (eigenvalues) of
e
U (right stretch tensor). Then,
as the tensors
e
C and
e
b have the same eigenvalues, (see Chapter 2), we can obtain:
) , ( ) , , , ( ) , (
2
3
2
2
2
1

e e e e e
b C 1 1 1 1 / / /
(9.321)
9.7.3.4 The Rate of Change of the Isotropic Helmholtz Free Energy
Let us consider the isotropic Helmholtz energy:
) , , , ( ) , (
2
3
2
2
2
1

e e e e
/ / / 1 1 1 b Isotropic Helmholtz free energy (9.322)
The rate of change of (9.322) is evaluated as follows:

`
` `

o
o
-
o
o

1 1
1
e
e
e
b
b
b : ) , ( (9.323)
NOTE: The operator is replaced by the number of contractions of the order of .
That is, if is an scalar (zeroth-order tensor), does not contract; if is a vector (first
order tensor), contracts once, i.e. the scalar (dot) product; if is second-order
tensor, : contracts twice, i.e. the double scalar product; and so on.
Then, by substituting
e
b
`
given in (9.274) into the equation (9.323) we obtain:
9 PLASTICITY

535
( ) ( )

`
`
`

o
o
-
o
o
-
o
o
-
o
o

o
o
-
(

- -
o
o



1 1 1 1
1 1
1
T e
e
e
e
e
e
T e e e
e
l l
l l
b
b
b
b
b
b
b b b
b
: : :
:


(9.324)
It was proved in Chapter 1 that if 1 is a scalar-valued isotropic tensor function, and if
e
b
is a symmetric second-order tensor, then
e
e e
e
b
b b
b o
o

o
o

1 1
holds, i.e.
e
b o
o1
and
e
b are
coaxial tensosr. Note also that
e
e e
e
b
b b
b o
o

o
o

1 1
results in a symmetric second-order
tensor, since
e
b o
o1
and
e
b are symmetric and coaxial tensors (see subsection 1.5.9 in
Chapter 1).
Note that the following equations are also satisfied:
( ) l l : :
_
symmetric
|
.
|

\
|
o
o

o
o

o
o

o
o

e
e
ik
e
jk
e
ij
e
kj ik
e
ij
e
e
b
b
b
b
b
b
b
b
1 1 1 1
l l
(9.325)
( ) l l : :
_
symmetric
|
.
|

\
|
o
o

o
o

o
o

o
o

e
e
jk
e
ik
e
ji
jk
e
ik
e
ij
T e
e
b
b
b
b
b
b
b
b
1 1 1 1
l l
(9.326)
In Chapter 1, (see Problem 1.16), it was proven that: if A and B are arbitrary second-
order tensors, the following is satisfied:
skew skew sym sym
B A B A B A : : : -
(9.327)
Thus, we can conclude that:

D
0
D
:
: : :
|
.
|

\
|
o
o

|
.
|

\
|
o
o
- |
.
|

\
|
o
o
|
.
|

\
|
o
o

=
e
e
skew
skew
e
e
sym
sym
e
e
e
e
b
b
b
b
b
b
b
b
1
1 1 1
l l l
_

(9.328)
Then, going back to the equation in (9.324), we can conclude that:
( ) ( )

`
`
`
`
`
`

o
o
- |
.
|

\
|
o
o
-
|
|
.
|

\
|
|
.
|

\
|
o
o

o
o
- |
.
|

\
|
o
o
-
|
|
.
|

\
|
|
.
|

\
|
o
o

o
o
- |
.
|

\
|
o
o
-
o
o

o
o
- |
.
|

\
|
o
o
- |
.
|

\
|
o
o
-
o
o

o
o
-
o
o
-
o
o
-
o
o






-
-
-
1 1 1
1 1 1
1 1 1
1 1 1 1
1 1 1 1
1
D
1
: :
: :
: :
: : :
: : :
e
e
e e e
e
e
e
e e e
e
e
e
e e e
e
e
e
e
e
e
e
T e
e
e
e
e
e
b
b
b b b
b
b
b
b b b
b
b
b
b b b
b
b
b
b
b
b
b
b
b
b
b
b
b
2
2
2
1
1
1

l
l
l l
l l

(9.329)
Then, the rate of change of the isotropic Helmholtz free energy becomes:


`
`

o
o
-
(

-
|
|
.
|

\
|
o
o

-

) , (

2
1 ) , (
2 ) , (
1
e
e e e
e
e
e
b
b b b
b
b
b
1 1
1 D

:
(9.330)
NOTES ON CONTINUUM MECHANICS

536


`
`

o
o
-
|
|
.
|

\
|
-
|
|
.
|

\
|
o
o

-

) , (
2
1 ) , (
2 ) , (
1
e
e e e
e
e
e
b
b b b
b
b
b
1 1
1 l

:
(9.331)

















Figure 9.50: Helmholtz free energy.
9.7.4 The Plastic Potential and the Yield Criterion
The plastic potential in the strain space is given by:
) , ( c
e
F G G (9.332)
As we can see, it is a function that depends on the same parameters as the free energy.
The yield surface is defined analogously as follows:
) , ( c
e
F F F (9.333)
The plastic potential and the yield surface have to fulfill the principle of objectivity, so the
following holds:
) , ( ; ) , ( c c
e e
E E F F G G (9.334)
We can also express these functions in the stress space. Then, if we consider that, in the
intermediate configuration, ) (
e
E S S , we then have:
) ), ( ( ; ) ), ( ( c c
e e
E E S S F F G G (9.335)
It is noteworthy that G and F are hypersurfaces as regards the six independent
components of S . When we are dealing with isotropic materials, these functions can be
represented in terms of the three eigenvalues of S .
) , (
e
F 1
X
,

x
,

X
,


p
F

e
F

p e
F F F
reference
configuration
current
configuration

0
B
B
B
intermediate
configuration
) , ( F y y ) , (
e
b 1 1
X
,

B
Intermediate configuration of the
Polar decomposition of
e
F

e e e
U R F

) , (
) , (
) , (

e
e
e
U 1
1
1
E
C


e
U

e
R

T
e
R
) , (
e
b 1 1

e e e
U , , E C

e
b
9 PLASTICITY

537
9.7.5 The Dissipation and the Constitutive Equation
The internal dissipation
int
D in the reference configuration, (see Chapter 5), is given by:
[ ] 0
0
- - j j
` ` `
T
int
E : S D
(9.336)
where is the Helmholtz free energy per unit mass, and j 1
0
is the energy per unit
volume. In the isothermal process, 0 T
`
, this dissipation becomes:
0 - - 1 1
` ` `
D S : : E
int
D
(9.337)
Then, if we consider the equation in (9.301), the above inequality becomes:
0
0
0
- |
.
|

\
|

- -
- |
.
|

\
|
-


1
1 1
1 1
`
` `
` `
e
T
e
e
jp ij
e
ik kp ij
e
jp kp
e
ik
T
e e
int
F F F F
F F
F F
D S
D S D
:
: :
D S D S
D
(9.338)
Then, by considering the equation in (9.294), the dissipation becomes:
0
0
-
-
A
1
1
`
`
E :
:
S
D
int
D
(9.339)
Next, by substituting (9.330) into the dissipation (9.339) we can obtain:
0
2
1
2 2
0
2
1
2
0 ) , (
1
1

o
o
-
|
|
.
|

\
|
- |
.
|

\
|
o
o
- |
.
|

\
|
o
o
-

o
o
-
(

- |
.
|

\
|
o
o
-
-
-
-

`
`
`

1 1 1
1 1
1
e e e
e
e
e
e e e
e
e
int
b b b
b
b
b
b b b
b
b

: :
: :
:

D
D D
D D

(9.340)
As the above inequality must be satisfied for any thermodynamic process, we can deduce
that:
e
e
e
b
b
b

o
o

) , (
2
1
Constitutive equation for stress (9.341)
In addition, by adopting

o
o
-
1
(thermodynamic forces), which represents the
hardening internal forces, we obtain:
0
2
1 1
-
|
|
.
|

\
|
-
-
`
e e
int
b b

: D
(9.342)
Note that in a process that is purely elastic (hyperelasticity), we have 0 ` , and revert to
the following scenario b b
e
, with which we find that the energy dissipation is equal to
zero, i.e. 0
int
D , (as we demonstrated in (9.279) that 0

b with which the constitutive


equation for stress (9.341) is reduced to b
b
b

o
o

) (
2
1
and which is the same expression
obtained as that for an isotropic hyperelastic material, see subsection 8.3.1 in Chapter 8).
NOTES ON CONTINUUM MECHANICS

538
9.7.6 Evolution of the Internal Variables
In order to fully describe the constitutive model we have to establish how the internal
variables evolve.
Firstly, we will define the elastic domain

E in the stress space, and the yield criterion F


in terms of the Kirchhoff stress tensor (current configuration):
{ 0 ) , ( : ) , ( s = F R E

(9.343)
where is the scalar-valued tensor function, denoted by the isotropic hardening function
and F is assumed to be a convex function.
Next, we apply the maximum dissipation principle, which states that dissipation in the
material reaches a maximum during a change characterized by a dissipative process. Let us
consider the current state

E = ) , ( which represents the current distribution of the


Kirchhoff stress tensor and thermodynamic forces in a body subjected to plastic strain. The
maximum dissipation principle requires that for a change of state, show here by

E = ) , (
* *
, the following must be satisfied:
[ ] [ ] 0
2
1
*
1
*
- -
|
|
.
|

\
|
- -
-
`
e e
b b

: (9.344)
The inequality in (9.344) describes an optimization problem with constraint. Here, we can
maximize the dissipation by minimizing the negative dissipation under the constraint
0 ) , ( s 1 . To this end, we define the Lagrangian:
1 1
1
2
1
` ` ` - -
|
|
.
|

\
|
- - - -
-

e e
int
b b

: D L (9.345)
where ` is the Lagrange multiplier that enforces 0 s 1 .
Then, the flow rule can be obtained as follows:
e e e e
b b b b
o
o
-
o
o
-
o
o -

1

1
` ` 2
2
1 1

0 0
L

(9.346)

o
o

o
o
- -
o
o 1

1
` ` ` ` 0 0
L

(9.347)
Next, we can summarize the evolution of the variables as:

o
o

o
o
-
) , (
;
) , (
2

1
` ` `
e e
b b


(9.348)
To fully define the model we introduce the loading/unloading Kuhn-Tucker conditions:
0 ) , ( ; 0 ) , ( ; 0 s F F ` ` (9.349)
and the persistency condition:
0 ) , ( F
`
`
(9.350)
Then, by substituting the equation in (9.278), i.e.
T p e
F C F b
-

1
`

, into (9.346) we can


obtain:
9 PLASTICITY

539
T e p e T p - -
- -
|
.
|

\
|
o
o
-
o
o
- F b F C b F C F

1

1
1 1
2 2 ` `
` `
(9.351)
and if we consider that
T p e
F C F b
-

1
we can conclude that:
1
1
1
1 1
1


2
2 2
-
-
-
-
- - -
-


(

|
.
|

\
|
o
o
-
|
.
|

\
|
o
o
- |
.
|

\
|
o
o
-
p
T T p T e p
C F F
F F C F F F b F C

`
` `
`

(9.352)
Summary
Helmholtz Free Energy (per unit
volume):
) , (
e
b 1 1 (9.353)
Measurement of elastic deformation:
T p
T
e e e
F C F F F b
-

1

(9.354)
Stress:
e
e
b
b

o
o

1
2 (9.355)
Isotropic hardening force:

` o
o
-
1
(9.356)
Yield surface: ) , ( F F (9.357)
Evolution equations:

o
o

|
.
|

\
|
o
o
-
-
-
-

1

` `
`
`
1
1
1
2
p p
C F F C
t
(9.358)
Kuhn-Tucker condition
(loading/unloading):
0 ) , ( ; 0 ) , ( ; 0 s F F ` ` (9.359)
Dissipation: 0
2
1 1
-
|
|
.
|

\
|
-
-
`
e e
int
b b

: D (9.360)
9.7.7 The Elastoplastic Tangent Stiffness Tensors
Let us consider the energy function ) (
e e e
E 1 1 , where ( )
1 - -
-
p p
T
p e
F E E F E ,
(see equation (9.238)), and the second Piola-Kirchhoff stress tensor in the intermediate and
in the reference configuration, respectively, are given by:
E E o
o

o
o

e
e
e
1 1
S S ; (9.361)
where
T
p p
F F S S is fulfilled (see equation (9.302)). The elastoplastic tangent stiffness
fourth-order tensors
e
C and C are introduced, (see Chapter 8), as follows:
E E E E o o
o

o o
o

e
e e
e
e
1 1
2 2
; C C (9.362)
The tensors
e
C and C have minor and major symmetries and they are related to each
other by:
1 1 1 1 - - - -

p
lq
p
kp
e
mnpq
p
jn
p
im ijkl
F F F F C C (9.363)
NOTES ON CONTINUUM MECHANICS

540
The tensor
e
C appears in the following linear relationship:
e e
E
` `
: C S
(9.364)
Then, by substituting the expression of S
`
given by (9.306), and
e
E
`
given by (9.299), into
the equation in (9.364), we obtain:
( )
)
`

- |
.
|

\
|
- -
- -

1
p p
sym
p e
T
p
T
p e
T
p
T
p p p
F F C F E F F F l
`
`
: C Z S S Z S
(9.365)
We then apply the dot product both between
1 -
p
F and (9.365), and also between the
equation in (9.365) and
T
p
-
F , with which we obtain:
( )
( ) |
.
|

\
|
- -
(

-
)
`

- - -


- - - -
T
p p p
sym
p e
T
p
T
p p p
sym
p e
T
p
T
p e p
T
p p
Z S S Z S
Z S S Z S
F C F E
F F F C F E F F
l
l
` `
` `
:
:
C
C
1 1

(9.366)
9.7.7.1 The Elastoplastic Tangent Stiffness Tensor
In the intermediate configuration the following holds:
( ) with
p e e e
e E E
`
` ` `
- : : C C S
e e
e
e
E E o o
o

1
2
C (9.367)
We can express
p
e
`
as follows:
G
G
S
S
S
V
) , (
` `
`

o
o


p
e
(9.368)
where ` is the plastic multiplier. Then, by combining (9.368) with (9.367) we obtain:
( )
|
|
.
|

\
|
o
o
- -
S
S
S
) , ( G
`
`
`
` `
E e E : :
e p e
C C
(9.369)
Next, to obtain the parameter ` we use the consistency condition, i.e. any change in the
intermediate configuration must allow the stress state to remain on the yield surface, thus
for 0 0 > F
`
` we have:
0
0 0 0 ) , (
1
1 1

o
o
-
o
o

o
o
-
o
o

o
o
-
o
o


n
i
i
i
n
i
i
i
n
i
i
i
H
H
c

c
c
c
c
F F
F F F F
F
`
` `
`
` `
`
S
S
S
S
S
S
S
:
: :

(9.370)
For the sake of simplicity we have considered that
i
c are scalars. Then, by combining
(9.369) with (9.370) we obtain:
0
0 0
1
1 1

o
o
-
o
o
o
o
-
o
o

o
o
-
(

|
|
.
|

\
|
o
o
-
o
o

o
o
-
o
o


n
i
i
i
e e
n
i
i
i
e
n
i
i
i
H
H H
c

c

c

F
G
F F
F
G
F F F
` `
` ` `
`
` `
S S S
S S
S
S
: : : :
: : :
C C
C
E
E

(9.371)
9 PLASTICITY

541
|
|
.
|

\
|
o
o
-
o
o
o
o
|
.
|

\
|
o
o

n
i
i
i
e
e
H
1
c

F
G
F
F
S S
S
: :
: :
C
C E
`
`
(9.372)
Next, by substituting the above expression of ` into the equation in (9.369), we find:
S
S S
S
S
S
o
o
|
|
.
|

\
|
o
o
-
o
o
o
o
|
.
|

\
|
o
o
-
o
o
-

G
F
G
F
F
G
:
: :
: :
: : :
e
n
i
i
i
e
e
e e e
H
C
C
C
C C C
1
c

E
E E
`
` ` `
`
(9.373)
Denoting by:
|
|
.
|

\
|
o
o
-
o
o
o
o

n
i
i
i
e
H
1
1
c
F
G
F
K
S S
: : C

(9.374)
the equation in (9.373) becomes:
S S
S
o
o
|
.
|

\
|
o
o
-
G F
K : : : :
e e e
C C C E E
` ` `
kl
ijkl
e
st
e
pqst
pq
kl
ijkl
e
ij
E E
S S
S
o
o
|
|
.
|

\
|
o
o
-
G F
K C C C
` ` `
(9.375)
Then, if we consider the substitution operator property we obtain:
ab
e
pqab
pq kl
ijkl
e
ijab
e
ab
kl
ijkl
e e
pqab
pq
ijab
e
ab
kl
ijkl
e
bt as
e
pqst
pq
bl ak
ijkl
e
kl
ijkl
e
bt as ab
e
pqst
pq
bl ak ab
ijkl
e
ij
E E
E
E E
` `
`
` ` `
(
(

|
|
.
|

\
|
o
o
|
|
.
|

\
|
o
o
-
|
|
.
|

\
|
o
o
o
o
-
|
|
.
|

\
|
o
o
o
o
-
o
o
o
o
-
C C C C C C
C C C
C C C
S S S S
S S
S S
S
F G
K
G F
K
G F
K
G F
K
c c c c
c c c c

(9.376)
Thus:
E E
` ` ` `
: :
: :
: :
ep tan
n
i
i
i
e
e e
e
H
_
1
C
C
C C
C
(
(
(
(
(

|
|
.
|

\
|
o
o
-
o
o
o
o
|
.
|

\
|
o
o

|
|
.
|

\
|
o
o
-

S
S S
S S
S
c
F
G
F
F
G

(9.377)
where we have introduced the elastoplastic tangent stiffness tensor as:
|
|
.
|

\
|
o
o
-
o
o
o
o
|
.
|

\
|
o
o

|
|
.
|

\
|
o
o
-

n
i
i
i
e
e e
e ep tan
H
1
_
c
F
G
F
F
G
S S
S S
: :
: :
C
C C
C C
Elastoplastic tangent stiffness
tensor (intermediate
configuration)
(9.378)
If we consider that
S
m
o
o

G
and
S
n
o
o

F
, where both tensors are symmetric second-order
tensors, then, the above equation can be rewritten as follows:
NOTES ON CONTINUUM MECHANICS

542
( ) ( )
e e e ep tan
C C C C : : n m - K
_
(9.379)
Note that m is the gradient of G in the stress space, n is the gradient of the yield surface
in the same space, i.e. n is normal to the yield surface. Then, in the case of associated flow,
G F , we obtain n m , and the elastoplastic tangent stiffness tensor is symmetric (major
and minor symmetry):
( ) ( )
|
|
.
|

\
|
o
o
-

n
i
i
i
e
e e
e ep tan
H
1
_
c
F
n n
n n
: :
: :
C
C C
C C
Elastoplastic tangent stiffness tensor
for associated flow rule
(9.380)
9.7.8 The Hyperelastoplastic Model with von Mises Yield
Criterion
In this subsection we will formulate a model for finite strain plasticity considering the
2
J
flow theory with the isotropic hardening law, (see Simo&Hughes (1998)).
9.7.8.1 The Helmholtz Free Energy
In this model, Simo&Hughes (1998), the Helmholtz free energy (per unit volume) is given
by:
( ) [ ] 3 )
~
(
2
) ( 1
2
1
2
2
- -
(

- -
x

e
J J b Tr ln
j
1

(9.381)
where ) (F det J , and
e e
J b b
3 / 2
~
-
is the isochoric part of the elastic part of the left
Cauchy-Green deformation tensor, x is the bulk modulus, and G j is the shear
modulus.
Then, starting from the intermediate configuration we apply a multiplicative decomposition
by means of a volumetric transformation brought about by 1
3
1
e
vol
e
J F , (see Figure 9.51),
and another brought about by an isochoric transformation characterized by
e e e
J F F
3
1
~
-
,
(see also Chapter 2) which gives us the following tensors:

3
2
3
1
3
1
e e e
T
vol
e
vol
e
vol
e
J J J |
.
|

\
|
1 1 F F b
(9.382)

( )
e e
T
e e e
T
e e e e
T
e e e
J J J J b F F F F F F b
3
2
3
2
3
1
3
1
~ ~ ~
- - - -

|
|
.
|

\
|
(9.383)







9 PLASTICITY

543























Figure 9.51: Volumetric and isochoric decomposition multiplicative decomposition.
9.7.8.2 The Stress Tensor
The Kirchhoff stress tensor (9.341) becomes
( ) [ ]
( ) [ ]
e e e
e e e
e e
e
e
e
e
J J J
J J
b b
b b b
b b
b
b
b
b

o
o
-
(

o
o
- -
o
o
x
)
`

- -
(

- -
x
o
o

o
o

-
) ( ) ( 1
2
1
3 )
~
(
2
) ( 1
2
1
2
2
) , (
2
3
2
2
2
Tr ln
Tr ln
j
j 1


(9.384)
where ) (
~
)
~
(
3
2
3
2
e e e e e e
J J b b b b Tr Tr
- -
1 1 : : holds and the derivatives of the invariants,
(see Chapter 1), become:

e
e p
vol
J
_
F
1
~ ~

p p
J F

vol
e e
J F
X
,

x
,

X
,


p
F

e
F

p
vol
e e p e
F F F F F F
~

reference
configuration
current
configuration

0
B
B
B
intermediate
configuration
) , ( c 1 1
e
b

e
e
b
b

o
o

1
2
X
,

B
1
3
1
e
vol
e
J F

e e e
J F F
3
1
~
-


p p
J F

e e
J F
1
~ ~

e e
J F
volumetric elastic
intermediate
configuration

e p p
vol
e e
J J J F F F F
~

X
,

B
volumetric plastic
intermediate
configuration

p e p p e
e p
vol
J J J F F F
~
3
1
3
1
3
1
_

-

1
3
1
p
vol
p
J F

p p p
J F F
3
1
~
-


vol
p
vol
p
J F
NOTES ON CONTINUUM MECHANICS

544
1
2
1 ) ( -

o
o
e e
e
e
J
J
b
b
,
1 1
2
1
2
1 ) ( ) ( - -

o
o

o
o

o
o
e e e p
e
e
p
e
p e
e
J J J
J
J
J J J
b b
b b b
,
[ ] 1 1
2
1
2
1 1 ) ( 1 ) ( - -

o
o

o
o
e e
e e
J
J
J
J
J
b b
b b
ln
,
[ ]
1
o
o
e
e
b
b ) ( Tr
,
( )
1
2
1
2
2
1
2 2 1
- -

o
o
-
o
o
e e
e e
J JJ
J
J J b b
b b
,
[ ]
1
3
2
1
3
5
3
2
3
5
3
2
) (
2
1
3
2 ) (
) (
3
2
) (
-
-
- - - -
-
-

o
o
-
o
o -

o
o
e e e e
e
e
e e
e
e e e
e
J J J J
J
J J b b
b
b
b
b
b
b
Tr
Tr
Tr Tr
Then, by incorporating the above derivatives into the equation in (9.384) we obtain:
( ) [ ]
[ ]
[ ]
[ ]
[ ]
(

-
-
- -
x

-
-
- -
x

-
-
- -
x

-
-
- -
x

-
-
-
(

- x

o
o
-
(

o
o
- -
o
o
x

-
- -
-
- -
-
-
-
-
-
-
-
- -
-
e e
e e e e e e e
e e e e e
e e e e e e
e e e e e e e
e e e
e e e
J
J J
J J
J J J
J J J J
J J J
b b
b b b b b b
b b b b
b b b b
b b b b b
b b
b b b
~
)
~
(
3
1
1
2
)
~
(
3
1
1
2
)
~
(
3
1
1
2
) (
3
1
1
2
) (
2
1
3
2
2
1
2
1
) ( ) ( 1
2
1
2
3
2
1 1
2
3
2
1 1
2
3
2
1
3
2
1
2
3
2
1
3
5
1 1
2
3
2
2
1 1
1
1
1
1
Tr
Tr
Tr
Tr
Tr
Tr ln
j
j
j
j
j
j

(9.385)
where [ ]
dev
e e e
b b b
~ ~
)
~
(
3
1
-
-
1 Tr . Then, the constitutive equation for stress becomes:
[ ] [ ]
dev
e
J b
~
1
2
2
j - -
x
1 The constitutive equation for stress (9.386)
9.7.8.3 Formulation Considering the Transformation
p
F as an
Isochoric Transformation
We will next consider the plastic deformation
p
F to be purely isochoric:
e e p p
J J ) ( ) ( 1 )
~
( )
~
( F F C F det det det det (9.387)
With this simplification the energy and stress equations become:
[ ]
[ ]
dev
e e
e e e
J
J J
b
b
~
1
2
3 )
~
(
2
) ( 1
2
1
2
2
2
j
j
1
-
(

-
x

- -
(

- |
.
|

\
|
-
x

1
Tr ln
(9.388)
As seen in Chapter 2 subsection 2.13, the following holds:
9 PLASTICITY

545
1 ; ; ~ ~
3
2
~ ~
3
~ ~
C b
b
b
C b
b
b
C b
I I I I I I
I I I
I I
I I I I
I I I
I
I I
(9.389)
Then, considering the equation in (9.383), i.e.
e e e
J b b
3
2
~
-
, we can show that:
) ( ) ( )
~
(
2
3
2
~
e e e e
J J I I I b b b
e
b
e
det det det
-
-

(9.390)
Afterwards, if we consider that
2
) ( ) ( J C b det det , we obtain
2
) ( ) (
e e e
J C b det det ,
which give us:
1 ) ( )
~
(
2 2
3
2
~
-
-
e e e e
J J J I I I b b
e
b
e
det det
(9.391)
We can also show that, if the relationships in (9.389) hold, so do
1 ; ; ~ ~
3
2
~ ~
3
~ ~
e e
e
e
e e
e
e
e e
C b
b
b
C b
b
b
C b
I I I I I I
I I I
I I
I I I I
I I I
I
I I
(9.392)
9.7.8.4 The Rate of Change of the Helmholtz Free Energy
The rate of change of the Helmholtz free energy, (see equation (9.388)), is given by
[ ]
(

-
(

-
x

)
`

- -
(

- |
.
|

\
|
-
x
1 1 : :
e e
e
e e e e e
J
J
J J J J
Dt
D
b b
`
` ` `
~
2
1
2
2
1
2
3
~
2
) ( 1
2
1
2
2 j j
1 ln (9.393)
Remember that F F C C E C
` ` ` `
: : :
T
J
J
J J J
- - -

1 1
2
) (D Tr , (see Problem 2.12), and
also that:
|
.
|

\
|
- -
-
- -
)
`


- - - - -
T
e e e e e e e e e e e e e e e e
J J J J J J J
Dt
D
l l b b b b b b b
3
2
3
5
3
2
3
5
3
2
3
2
3
2 ~
` ` `
`
(9.394)
Note that
T
l l - b b b
`
holds, (see Chapter 2). Similarly, we can prove that the
following relationship
T
e e e e e
l l - b b b
`
is valid.
Then, by substituting (9.394) into (9.393) we obtain:

|
.
|

\
|
- - - -
(

-
x

|
.
|

\
|
- - - -
(

-
x

-
(

-
x



- -
- -
1 D D
1
1
:
:
:
T
e e e e e e e e e e e
e
e
T
e e e e e e e e e
e
e
e e
e
e e
J J J J
J
J
J J J J
J
J
J
J
J J
l l
l l
b b b
b b b
b
3
2
3
5
3
2
3
5
) ( ) (
3
2
2
) (
1

2
3
2
2
1

2
~
2
1
2
2
1
2
Tr Tr Tr
j
j
j
1
` `
`
` ` `
(9.395)
Furthermore, if ( )
e
kj
e
jk
e
kj
e
jk ij
e
jk
e
ik
e
kj
e
ik
T
e e e e
b b b b D 2 2 - |
.
|

\
|
- l l l c 1 : l l b b holds.
Then:
NOTES ON CONTINUUM MECHANICS

546
[ ]
e
e
dev
e e e e e e
e e e e e e e e
J J
J J J
D
D 1 D 1 1
D D D
:
: :
:

(

- |
.
|

\
|
-
x

|
.
|

\
|
- - - |
.
|

\
|
-
x

- - -
(

-
x

- -
b b b
b b
~
1
2
~
)
~
(
3
1
1
2
2 ) ( ) (
3
2
2
) ( 1
2
2 2
3
2
3
2
2
j j
j
1
Tr
Tr Tr Tr
`

(9.396)
9.7.8.5 Yield Criterion and Evolution of the Internal Variables
Let us consider the Mises-Hubers yield condition formulated in terms of the Kirchhoff
stress tensor as:
[ ] 0
3
2
) , ( s - o - c K
Y
dev
F (9.397)
where
Y
o is the yield stress, 0 > K is the isotropic hardening modulus, and c is the
hardening parameter.
The plastic flow rule in the current configuration is given by:
dev
dev
e
dev
e

) (
3
2
b b Tr ` -
|
|
.
|

\
|


(9.398)
The Isotropic Hardening Law and the Loading/Unloading Conditions
We assume that the rate of change of the hardening is given by:
c ` `
3
2
(9.399)
where ` is the consistency parameter subjected to Kuhn-Tucker (loading/unloading)
conditions:
0 ) , ( ; 0 ) , ( ; 0 s F F ` ` (9.400)
which together with the persistency condition:
0 ) , ( F
`
`
(9.401)
complete the model formulation.











10 Thermoelasticity












10.1 Thermodynamic Potentials
Remember from Chapter 5 that the Clausius-Duhem inequality can be expressed as
follows:
0
1

1 1
) , (
2
- - - T
T
u
T T
t
x
x
,
,
`
,
` V q D j j j :
Clausius-Duhem inequality
(current configuration)
(10.1)
0
1 1 1
0
1 1 1
0
2
0 0
0
2
0 0
- - -
- - -

T
T
u
T T
o
T
T
u
T T
X
X
F
E
,
,
,
`
`
,
`
`
`
`
V
V
q P
q S
j j j
j j j
:
:
Clausius-Duhem inequality
(reference configuration)
(10.2)
Note that 0 s T
x
,
,
V q , since the sense of the heat flux vector ( q
,
) is always opposite to that
of the temperature gradient ( T
x
,
V ). Thus, we can formulate the heat conduction inequality,
(see Chapter 5), as follows:
0 - T
x
,
,
V q (current configuration)
Heat conduction inequality
0
0
- T
X
,
,
V q (reference configuration)
(10.3)
Then, by imposing the restriction (10.3) into the Clausius-Duhem inequality (10.1) and
(10.2) we are lead to the Clausius-Planck inequality:
10
Thermoelasticity
547 , Notes on Continuum Mechanics, Lecture Notes on Numerical
Methods in Engineering and Sciences 4, DOI 10.1007/978-94-007-5986-2_12,
International Center for Numerical Methods in Engineering (CIMNE), 2013
E.W.V. Chaves
NOTES ON CONTINUUM MECHANICS

548
0 ) , (
1 1
) , ( - - t u
T T
t
int
x x
,
`
,
` j j j D : D (current configuration)
Clausius-Planck
inequality
0 ) , (
1 1
) , (
0
1 1
0 0
0 0
- -
- -
t u
T T
t
or
u
T T
int
int
X F X
E
,
`
`
,
`
`
`
`
j j j
j j j
:
:
P
S
D
D
(reference configuration)
(10.4)
where
int
D is the internal energy dissipation (or the local entropy production), which must
be positive throughout the continuum at any point and time, i.e. 0
int
D .
Then, in a reversible process we have 0
int
D , with which we obtain:
j
j
j j j ` `
`
` `
`
T u u
T T
int
- - - E E : : S S
0
0 0
1
0
1 1
D
(10.5)
10.1.1 The Specific Internal Energy
The equation in (10.5) indicates that the specific internal energy ( u ) is a thermodynamic
potential in terms of j , E (independent state variables) when evaluating T , S (the state
function), (see Asaro&Lubarda (2006)). In fact, if ) , ( j E u its rate of change becomes:
j
j
j `
`
`
o
o
-
o
o

u u
u E
E
E : ) , (
(10.6)
Then, by comparing the equations (10.5) and (10.6) we can draw the conclusion that:
0 E
E
E
E

|
|
.
|

\
|
o
o
|
.
|

\
|
o
o

`
`
j
j j
j
j
) , (
;
) , (
0
0
u
T
u
S
(10.7)
10.1.2 The Specific Helmholtz Free Energy
Now, another thermodynamic potential is the specific Helmholtz free energy denoted by
, which is defined as:
j T u -
Specific Helmholtz free energy
(

kg
J

(10.8)
We now need to verify that we are working with specific energy, i.e. energy per unit mass:
[ ] [ ] [ ]
kg
J
kgK
J
K T u j and then the rate of change of (10.8) is given by:
j j `
` `
` T T u - - (10.9)
Then, by substituting the rate of change of the specific internal energy given in (10.5) into
the above equation we obtain:
j
j
j j j
j
j j T T T T T T u
` ` ` ` `
` ` ` `
`
- - - - - - E E : : S S
0 0
1 1

(10.10)
10 THERMOELASTICITY

549
where is a thermodynamic potential in terms of ( T , E ) (independent state variables)
when evaluating ( j , S ). Moreover, by calculating the rate of change of ) , ( T E , we obtain:
T
T
T
` ` `
o
o
-
o
o


E
E
E : ) , ( (10.11)
Then, by comparing the equations (10.10) and (10.11) we can conclude that:
0 E
E
E
E
E
E

|
.
|

\
|
o
o
- |
.
|

\
|
o
o

` ` T
T
T
T
T
T
) , (
) , ( ;
) , (
) , (
0
0

j

j S
(10.12)
Now, the rate of change of entropy ) , ( T E j becomes:
T
T
T T
T
T
` `
` `
`
0 E
E
E
E
E
E

|
.
|

\
|
o
o
- |
.
|

\
|
o
o

) , ( ) , (
) , (
0
j j
j :
(10.13)
Furthermore, let us imagine a process where E E E d - , dT T T - , and j j j d - .
It then holds that:
dT
T
T
d
T
T d
T 0 E
E
E
E
E
E

|
.
|

\
|
o
o
- |
.
|

\
|
o
o

` `
) , ( ) , (
) , (
0
j j
j :
(10.14)
We can also express the equation (10.6) by means of differentials as follows:
j
j
j d T d du
1
) , (
0
- E E : S
(10.15)
Then, by substituting (10.14) into (10.15) we obtain:
dT
T
T
T d
T
T
dT
T
T
d
T
T d du
T
T
0 E
0 E
E
E
E
E
E
E
E
E
E


|
.
|

\
|
o
o
-
(

|
.
|

\
|
o
o
-
(

|
.
|

\
|
o
o
- |
.
|

\
|
o
o
-
` `
` `
) , (

) , ( 1
) , ( ) , (

1
0
0
0
0
j j
j
j j
j
:
: :
S
S

(10.16)
The necessary and sufficient condition for du to be a total differential is guaranteed by:
(

|
.
|

\
|
o
o
o
o

|
.
|

\
|
o
o
-
o
o
0 E
E
E E
E
` ` T
T
T
T
T
T
T
) , (

) , ( 1
0 0
j j
j
S (10.17)
which give us the following equation:
|
.
|

\
|
o
o
-
o
o

0
S
0
) , ( 1
0 T
T
T ` E
E j
j
0
) , (
0

|
.
|

\
|
o
o
- |
.
|

\
|
o
o
T
T
T ` ` E
E
0 E
j
j
S
(10.18)
Then, we can introduce a new second-order tensor as follows:
0
) , (
0

|
.
|

\
|
o
o
- |
.
|

\
|
o
o

T
T
T ` ` E
E
M
0 E
j
j
S
The thermal stress tensor
(

K
Pa
(10.19)
10.1.3 The Specific Gibbs Free Energy
We will now introduce a new thermodynamic potential: the specific Gibbs free energy ( G)
which is a potential in terms of stress and temperature:
NOTES ON CONTINUUM MECHANICS

550
E E : S S
0
1
) , ( ) , (
j
- T T G
Specific Gibbs free energy
(

kg
J

(10.20)
whose rate of change becomes:
E E E
` ` ` ` `
: : : S S S
S
0 0
1 1
) , (
j j
- -
o
o
-
o
o
T T
T
G G

(10.21)
Then, by substituting the rate of change of the Helmholtz free energy given in (10.10) into
the above equation, we can obtain:
E
E E E
: :
: : : :
S S
S
S S S S
S
`
` `
`
` ` ` ` ` `
0
0 0 0
1
1 1 1
j
j
j j
j
j
- -
o
o
-
o
o

- - -
o
o
-
o
o
T T
T
T T
T
G G
G G
(10.22)
with which we can draw the conclusion that:
0 S
S
S
S
S
S

|
.
|

\
|
o
o
- |
.
|

\
|
o
o
-
` ` T
T
T
T
T
T
) , (
) , ( ;
) , (
) , (
0
0
G G
j j j E E
(10.23)
and if we consider the relationship between the specific Helmholtz free energy and specific
internal energy, j j T u T - ) , ( ) , ( E E , we can obtain the following equations:
j
j
j
j
j j
j
T T u T u T T - - - - - E E E E E E : : : S S S S S
0 0 0
1
) , ( ) , (
1
) , (
1
) , ( ) , ( G G
(10.24)
10.1.4 The Specific Enthalpy
Next, we will introduce the specific enthalpy (H) which is a thermodynamic potential in
terms of stress and entropy, such that:
j
j
j j T T u - - ) , (
1
) , ( ) , (
0
S S S G H E E :
Specific
Enthalpy
(

kg
J

(10.25)
Now, by evaluating the rate of change of the above equation, we obtain:
E E E
` `
`
`
` : : : S S S
S
0 0
1 1
) , (
j j
j j
j
- -
o
o
-
o
o
u
H H

(10.26)
Then, by substituting j
j
`
`
` T u - E : S
0
1
, (see equation (10.5)), into the above, we find:
E
E E E
: :
: : : :
S S
S
S S S S
S
` `
` ` ` `
` `
` `
0
0 0 0
1
1 1 1
j
j j
j
j j
j
j
j
j
-
o
o
-
o
o

- - -
o
o
-
o
o
T
T
H H
H H
(10.27)
with which we can draw the conclusion that:
0 S
S
S
S

|
|
.
|

\
|
o
o
|
.
|

\
|
o
o
-
`
`
j
j j
j
j
) , (
;
) , (
0
0
H H
T E
(10.28)
10 THERMOELASTICITY

551
Now, let us suppose that we make a change in the system characterized by the following
process dT T T - , S S S d - , j j j d - , with which the equation in (10.27) can be
expressed as follows:
S d Td d : E
0
1
j
j - H
(10.29)
Then, taking the differential of ) , ( T S j j , we obtain:
dT
T
d d
T
: :
0 S
S
S

|
.
|

\
|
o
o
- |
.
|

\
|
o
o

` `
j j
j
0

(10.30)
and combining (10.30) with (10.29) gives us:
dT
T
T d T
d dT
T
T d T d
T
T
: :
: : :
0 S
0 S
S
S
S S
S


|
.
|

\
|
o
o
-
(

- |
.
|

\
|
o
o

- |
.
|

\
|
o
o
- |
.
|

\
|
o
o

` `
` `
j
j
j
j
j j
E
E
0
0
1
1
0
0
H
(10.31)
Then, the necessary and sufficient condition for H d to be a total differential is guaranteed
by:
(

|
.
|

\
|
o
o
o
o

- |
.
|

\
|
o
o
o
o
0 S
S S ` ` T
T T
T
T
j
j
j
E
0
1
0
(10.32)
the result of which is:

o
o
- |
.
|

\
|
o
o

0
S T
T
E
0
1
0
j
j
`
0 S
S

|
.
|

\
|
o
o
|
.
|

\
|
o
o
`
` T
T
E
0
1
0
j
j
(10.33)
Next, we can define a new tensor: the thermal expansion tensor, by:
0
0

|
.
|

\
|
o
o
|
.
|

\
|
o
o

T
T `
` S
0 S
j
j
E
A
The thermal expansion tensor
(

K
1
(10.34)
Now, we can make a summary of all the thermodynamic potentials in Table 10.1 by means
of which we can easily show that 0 ) ( ) ( - - - H G u .
Table 10.1: Thermodynamic potentials.
Specific internal energy
Specific Helmholtz free
energy
Specific Gibbs free energy Specific enthalpy
) , ( j E u ) , ( T E ) , ( T S G ) , ( j S H
j
j
j j
o
o

o
o

u
T
u
) , (
) , (
0
E
E
E S

T
T
T
o
o
-
o
o

j
) , (
) , (
0
E
E
E S

T
T
T
o
o
-
o
o
-
G
G
) , (
) , (
0
S
S
S
j
j E

j
j
j j
o
o

o
o
-
H
H
) , (
) , (
0
S
S
S
T
E
E : S
0
1
j
j - - T u G

j T u -

j
j

T -
-
H
G E : S
0
1

j
j
T
u
-
-
G
H E : S
0
1

NOTES ON CONTINUUM MECHANICS

552
10.2 Thermomechanical Parameters
10.2.1 Isothermal and Isentropic Processes
Isothermal processes are characterized by having a constant temperature ( 0 T
`
) during a
system change. Good approximations of isothermal processes are those found in materials
that are good heat conductors (e.g. metals) and which are subjected to quasi-static
processes. We can describe Isentropic processes as those with constant entropy ( 0 j` ) during
a system change. A good approximation of an isentropic process is when the continuum is
a poor heat conductor and quantities (velocity, stress, strain) vary rapidly.
Let us now return to some of the expressions obtained previously:

|
|
.
|

\
|
o
o

|
.
|

\
|
o
o

o
o
-
o
o

0 E
E
E
E
E
E
E
E
E
`
`
`
`
`
j
j
j
j
j j
j
j
j
j
) , (
) , (
) , (
) , (
) , (
0
0
u
T
u
u u
u
ise
S
: (10.35)
and from the rate of change of the specific Helmholtz free energy we obtain:

|
.
|

\
|
o
o
-
|
.
|

\
|
o
o

o
o
-
o
o

0 E
E
E
E
E
E
E
E
E
`
`
` ` `
T
T
T
T
T
T
T
T
T
isoT
) , (
) , (
) , (
) , (
) , (
0
0

S
:
(10.36)
which gives us two ways to obtain the stress tensor, namely:
0 0
) , (
) , ( ;
) , (
) , (
0 0

|
.
|

\
|
o
o
|
.
|

\
|
o
o

T
T
T
u
isoT ise
`
`
E
E
E
E
E
E

j
j
j j
j
S S
(10.37)
Then, by calculating the rate of change of ) , ( T E S , (see Table 10.1), we obtain:
T T
T
T
T
T
e
isoT
T
T
` ` ` `
` `
`
` `
` `
M E
E
E
E E
E
E
E
0 E
0 E
-
|
|
.
|

\
|
o o
o
-
|
|
.
|

\
|
o o
o

|
.
|

\
|
o
o
- |
.
|

\
|
o
o



: :
:
C

j
2
0
2
0
0
0
) , (
S S
S
(10.38)
where we have introduced the symmetric fourth-order tensor:
The isothermal elastic tangent stiffness tensor:
0
0
) , ( ) , (
2
0

|
|
.
|

\
|
o o
o
|
.
|

\
|
o
o

T
T
T T
e
isoT
`
` E E
E
E
E
j
S
C
The isothermal elastic
tangent stiffness tensor
[ ] Pa (10.39)
Then, calculation of the rate of change of ) , ( T S E yields:
T T
T
T
T
T
e
isoT
T
T
`
`
`
`
`
`
`
` `
` `
A
E E
E
-
|
|
.
|

\
|
o o
o
-
|
|
.
|

\
|
o o
o
-
|
.
|

\
|
o
o
- |
.
|

\
|
o
o



S
S
S
S S
S
S
S
0 S
0 S
: :
:
D
G G
2
0
2
0
0
0
) , (
j j
(10.40)
10 THERMOELASTICITY

553
where
1 -

e
isoT
e
isoT
C D and M A :
1 -
-
e
isoT
D holds.
Now, if we calculate the rate of change of ) , ( j E S , (see Table 10.1), we can obtain:
j j
j
j j
j
j
j
j
j
` `
`
` `
` `
`
`
`
`
M E
E
E
E E
E
E
E
0 E
0 E
-
|
|
.
|

\
|
o o
o
-
|
|
.
|

\
|
o o
o

|
|
.
|

\
|
o
o
- |
.
|

\
|
o
o

: :
:
e
ise
u u
C
2
0
2
0
0
0
) , (
S S
S
(10.41)
where we have introduced a new symmetric fourth-order tensor:
The adiabatic elastic tangent stiffness tensor:
0
0
) , ( ) , (
2
0

|
|
.
|

\
|
o o
o
|
.
|

\
|
o
o

j
j
j
j
j
`
`
E E
E
E
E u
e
ise
S
C
The adiabatic elastic
tangent stiffness tensor
[ ] Pa (10.42)
Remember that in isotropic linear elastic materials the elasticity tensor is expressed in terms
of the Lam constants as follows: I 1 1 j A 2 -
e
C . Likewise, we can define the adiabatic
and isothermal elasticity tensors for isotropic linear elastic materials as:
I 1 1
ise ise
e
ise
j A 2 - C Adiabatic elasticity tensor for isotropic materials (10.43)
I 1 1
isoT isoT
e
isoT
j A 2 - C Isothermal elasticity tensor for isotropic materials (10.44)
where (
adi adi
j A , ), (
isoT isoT
j A , ) are the Lam constants for isentropic and isothermal
processes, respectively.
10.2.2 Specific Heats and Latent Heat Tensors
According to Asaro&Labarda(2006), the ratio of the absorbed amount of heat and the
temperature increase is called heat capacity. The heat capacity per unit mass is denoted by
the specific heat capacity or simply the specific heat. To understand this concept, we can
make an analogy. For example, we can take a dry sponge and put it under a tap (faucet)
which we then open. We will observe that the sponge is able to retain a certain amount of
water until it is fully saturated after which it will no longer be able to retain the water. We
can also observe that the amount of water flowing out from the tip of the sponge varies
over time. We will observe that, depending on the characteristics of the sponge, that is, if it
has fewer of more holes, it will retain less or more water. The same is true with heat:
materials have the ability to retain a certain amount of thermal energy (internal energy
store).
By the fact that the increase of heat is not a perfect differential, the specific heat depends
on the system change path, and then the specific heat can be measured under various
conditions. We then set two types of transformations, one at a constant stress (pressure),
and the other at a constant strain (volume), (see Asaro&Lubarda(2006)).
Specific heat at a constant strain: it is a scalar that corresponds to the heat supplied to a
unit of mass so as to achieve a unit temperature change whilst maintaining the strain
constant, ( 0 E
`
):
NOTES ON CONTINUUM MECHANICS

554
0 E
0 E 0 E
E


|
|
.
|

\
|
o o
o
- |
.
|

\
|
o
o
|
.
|

\
|
o
o

`
` ` T T
T
T
u
T
T c
j
2
Specific heat at a
constant volume
(

kgK
J

(10.45)
Next, we can also check the SI unit: [ ] [ ]
[ ]
[ ]
[ ]
kgK
J
T
T
T
T c
(

|
.
|

\
|
o
o
j
j j
E
.
Specific heat at a constant stress: it is a scalar that corresponds to the heat required for a
unit temperature change while the stress is maintained constant ( 0 P
`
):
0 S 0 S
S

|
.
|

\
|
o
o
|
.
|

\
|
o
o

` ` T T
T c
H j
Specific heat at a constant
stress
(

kgK
J

(10.46)
The entropy here can be expressed in terms of:
) , ( ) , ( T T S j j j E (10.47)
In Figure 10.1 we can appreciate the graph entropy vs. temperature for water, where we can
verify that during the phase changes there is a jump in entropy without there being any
temperature variation. In said graph we can also verify that ) (T c c
E E
is temperature
dependent.
We can define the latent heat tensor of change of strain as the heat that must be provided at the
material point so as to achieve a unit strain change while the temperature is maintained
constant:
0
|
.
|

\
|
o
o

T
T
` E
L
E
j

(

kg
J
kgK
J
K

(10.48)
Note that in Figure 10.1 in
sl
T or
lg
T there is an entropy jump, i.e., we are providing heat
with no temperature change, since at these points there is a phase change.
We can define the latent heat tensor of change of stress as the heat that must be provided
at the material point so as to achieve a unit stress change while the temperature is
maintained constant:
0
|
.
|

\
|
o
o

T
T
` S
S
j
L
(

kg
m
N
m
kgK
J
K
3 2

(10.49)
Note that
E
L and
S
L are symmetric second-order tensors, since E and S are symmetric
tensors too. Then, if we consider the following relationships, (see equations (10.19) and
(10.34)) we have:
A
E
M
E
0 E
|
.
|

\
|
o
o
|
.
|

\
|
o
o
- |
.
|

\
|
o
o
- |
.
|

\
|
o
o
0 S
S
S
`
` ` ` T T
T T 0 0
0 0
;
j
j
j
j
(10.50)
and the latent heat tensors can be rewritten as follows:
M
E
L
0 E
E
0 0 0
j j
j T
T
T
T
T
- |
.
|

\
|
o
o
- |
.
|

\
|
o
o


` `
S
Latent heat tensor of change of strain (10.51)
A
E
L
0 0 0
j j
j T
T
T
T
T
|
.
|

\
|
o
o
|
.
|

\
|
o
o

0 S
S
S
`
`
Latent heat tensor of change of stress (10.52)


10 THERMOELASTICITY

555












Figure 10.1: Entropy vs. temperature (water).
If we now take the derivative of ) , ( ) , ( T T S j j j E with respect to temperature we
obtain:
T
T
T
T
T
T
T
T
T
T
T
T
T
T
T
c c
o
o
o
o
-
o
o

o
o

o
o
o
o
-
o
o

o
o E
E
E E E
E
E E
E
: :
) , ( ) , ( ) , ( ) , ( ) , ( ) , ( j j j j j j
_ _
S
S S

(10.53)
Note that
S
0 S
L
E
0
j |
.
|

\
|
o
o

` T
T ,
T
T
E
L
E
|
.
|

\
|
o
o
0
`
j
, thus:
E E
L L :
S S
T
c c
0
j
-

(10.54)
Additionally, the following holds:
E
E
E
E E
E
E
o
o

o
o

o
o
o
o

o
o - - ) , ( ) , ( ) , ( ) , ( 1 1 T T T
T
T
e
isoT
e
isoT
j j j j
: : : C C
S
S
(10.55)
or:
e
isoT
e
isoT
e
isoT
T T
C C C : : :
S S
S
L L L L
L L
E E
E
=
- - 1 1
(10.56)
S S S S
L L L L L L L L
E E E E
: : : : : :
e
isoT
e
isoT
e
isoT
C C C
- - 1 1

(10.57)
Then, by using the definition in (10.54) we can draw the conclusion that:
( )
S S S
L L L L
E E E
: : : :
e
isoT
e
isoT
T T
c c C C
0
1
0
j j
|
.
|

\
|
-
-
(10.58)
and as
1 -
e
isoT
C is a positive definite tensor, i.e. 0
1
>
-
E E
L L : :
e
isoT
C is satisfied, so, the
following must be met:
E
c c >
S
(10.59)
Furthermore, if we return to the equation in (10.13) we can also express this as follows:
) (T j
solid
liquid
gas
) (K T
sl
T
lg
T T
curve tangent

T
T c ) (
E

o
o
T
j

Third Law of
Thermodynamics
0 ) 0 ( j j T
NOTES ON CONTINUUM MECHANICS

556
T
T
c
T
T
T T
T
T
` ` ` `
` `
`
E
E M
E
E
E
E
E
0 E
- - |
.
|

\
|
o
o
- |
.
|

\
|
o
o


: :
0
1 ) , ( ) , (
) , (
0
j
j j
j
(10.60)
where we have used the equations in (10.50) and (10.45).
Next, we will obtain the relationship between
e
ise
C and
e
isoT
C , (see Holzapfel(2000)). Note
that in isentropic processes, 0 j` , the equation in (10.60) becomes:
E M
E
` `
:
c
T
T
0
j

(10.61)
Then, if we take the rate of change of ) , ( T E S , (see equation (10.12)), we are given:
T T
T
T
e
isoT
T
` ` ` ` `
` `
) , (
0
M E E
E
E
0 E
- |
.
|

\
|
o
o
- |
.
|

\
|
o
o


: : C
S S
S
(10.62)
Next, by substituting (10.61) into (10.62), we obtain:
E M M
E M M E M E E
E
E
`
` ` ` `
`
:
: : :
|
|
.
|

\
|
-
|
|
.
|

\
|
- -
c
T
c
T
T T
e
isoT
e
isoT
e
isoT
0
0
) , (
j
j
C
C C S

(10.63)
and by comparing this with the equation in (10.41) we can conclude that:
M M
E
-
c
T
e
isoT
e
ise
0
j
C C (10.64)
10.3 Linear Thermoelasticity
10.3.1 Linearization of the Constitutive Equations
As discussed in Chapter 6, the constitutive equations for simple thermoelastic materials can
be expressed as follows:
) , , (

) , (
) , (
) , (
) , (
0 0
0
T T
T
T
T
T
T
X
E
E
E
E
E
E
,
, ,
V q q
S

o
o
-
o
o

j

The constitutive equations for simple
thermoelastic materials
(Reference configuration)
(10.65)
where is the specific Helmholtz free energy (per unit mass), S is the second Piola-
Kirchhoff stress tensor, j is the specific entropy (per unit mass),
0
q
,
is the heat flux
vector, E is the Green-Lagrange strain tensor, T denotes temperature, T
X
,
V is the
temperature gradient, and F is the deformation gradient.
To make the linearization of the constitutive equations, (Nowacki(1967), ilhav(1997),
Pabst(2005)), we will use the Taylor series expansion, where the following condition holds:
given a function ) (x f , said function can be approximated by using the Taylor series:
10 THERMOELASTICITY

557

-
o
o

0
) (
) (
!
1
) (
n
n
n
n
a x
x
a f
n
x f , applied at point a (the application point). We will now apply
this same definition but applied to tensors, (see Chapter 1).
10.3.1.1 The Linearized Piola-Kirchhoff Stress Tensor
The Piola-Kirchhoff stress tensor can be represented by means of the Taylor series in
which we will consider up to linear terms:
) (
) , (
) (
) , (
) (
) , (
) (
) , (
) , ( ) , (
2
0
2
0
terms
order Higher
O
O O
O
O O
O
O
O O
O
O O
O O
T T
T
T T
T T
T
T T
T T
-
o o
o
- -
o o
o
- =
- -
o
o
- -
o
o
-
E
E
E E
E E
E
E
E E
E
E
E E

j :
:
S
S S
S S
_


(10.66)
where we have considered that
E
E
E
o
o

) , (
) , (
0
O O
O O O
T
T

j S S . Note that we have used
the subscript O to indicate the variable value at the application point, so as to differentiate
this from the subscript 0 which is used to identify variables in the reference configuration.
Note that the linearized constitutive equation for stress has a linear relationship with strain
and temperature, but also considers large deformation kinematics.
Now, if we consider the equation in (10.66), we can identify these material properties:
The isothermal elasticity tensor (
e
isoT
C ) (reference configuration):
E E
E
E
E
o o
o
|
.
|

\
|
o
o

) , ( ) , (
2
0
0
O O O O e
isoT
T T
T

j
`
S
C The isothermal elasticity tensor (10.67)
The thermal stress tensor ( M):
E
E
E
E E
M
0 E
o o
o
=
o o
o
|
.
|

\
|
o
o

T
T
T
T
T
T
O O O O O O
) , ( ) , ( ) , (
2
0
2
0

j

j
`
S The thermal stress
tensor
(10.68)
If we then consider the equations S P F ,
E
E
F E
o
o

) , (
) , (
0
T
T

j P ,
F
F
F
o
o

) , (
) , (
0
T
T

j P , where P is the first Piola-Kirchhoff stress tensor, we can obtain:
F
F F
M
o o
o

o
o

T
T
T
T ) , ( ) , (
2
0

j
P

(10.69)
Then, returning to the equation in (10.66) and by considering that 0 S
O
and 0
O
E , the
linearized constitutive equation for stress becomes:
) (
O
e
isoT
T T - - M E : C S (10.70)
Note that the following holds:
0 S - E M with T T
O
) ( (10.71)
That is, the thermal stress tensor provides stress in the absence of strain ( 0 E ).
We can also define:
The Latent heat tensor of change of strain (
E
L )
NOTES ON CONTINUUM MECHANICS

558
The thermal stress tensor is closely linked to the latent heat tensor
E
L (symmetric second-
order tensor) which can be expressed as follows:
T O T O
O O
O
O O O
T
T T
T
T
T
T
T T
F
F
F M
E
E E
L
E

o o
o
- -
o o
o
-
o
o
-

j j

j
2
0 0
2
0
) , ( ) , ( S
Latent heat tensor of change of
strain
(10.72)
Now, we can relate the latent heat tensor (
E
L ) to the thermal stress tensor ( M) by:
M
E
E
L
E
0
2
) , (
j

O O O
O
T
T
T
T -
o o
o
- (10.73)
Next, we will define some material parameters which are related to deformation, i.e. those
which are associated with the inverse of the equation in (10.70).
The thermal expansion tensor ( A)
Now, based on the stress expression given in (10.70), ) (
O
e
isoT
T T - - M E : C S , we can
obtain the inverse relationship as follows:
) (
) (
1 1
1 1 1
O
e
isoT
e
isoT
O
e
isoT
e
isoT
e
isoT
e
isoT
T T
T T
- -
- -
- -
- - -
M E
M E
: :
: : : :
C C
C C C C
S
S
I
_

(10.74)
where
S o
o

-
E
1
) (
e
isoT
C holds. Note that if the body can deform freely, the implication is that
there is no stress 0 S and here the equation in (10.74) becomes:
) ( ) (
1
O O
e
isoT
T T T T - - -
-
A M E : C
(10.75)
Therefore, we can define the thermal expansion tensor, denoted by A, as follows:
M A :
1 -
-
e
isoT
C
The thermal expansion tensor (10.76)
Then, the equation in (10.74) can be rewritten as follows:
) (
1
O
e
isoT
T T - -
-
A E S : C
(10.77)
NOTE: Although we have defined the tensors
E
L and A, in thermal stress analysis, we
need only know the tensor M. However, as regards practice in laboratory measurement, it
is more convenient, from a practical standpoint, to obtain the thermal expansion tensor
( A) and then we can obtain the thermal stress tensor by means of the equation
A M :
e
isoT
C - .
10.3.1.2 The Linearized Heat Flux Vector
The linearization if the heat flux vector can be obtained as follows:
T
T
T
T
T T T
O O O
X
X
X
X
X X
E
,
,
,
,
, ,
,

,
, , ,
V
V
V
V
V V
o
o
= -
o
o
-
0 0
0 0 0
) , , ( ) (
q q
q q q
(10.78)
where we have taken into account that 0 0 q q
, ,
, ,
,
) , , (
0 0
T T
O O O X
E V . Next, we can
define the following thermal material properties:
10 THERMOELASTICITY

559
The thermal conductivity tensor ( K ):
The thermal conductivity tensor is given by the following equation:
T
X
,
,
V o
o
-
0
0
q
K
The thermal conductivity tensor
(Reference configuration)
(

mK
W
(10.79)
If we now consider the units: [ ]
s m
J
2
0
q
,
, [ ]
m
K
T
X
,
V , it is easy to show that:
[ ]
[ ]
[ ] mK
W
K
m
s m
J
T

2
0
0
X
,
,
V
q
K .
Then, we can rewrite the linearized heat flow vector as follows:
T T
X X
, ,
,
V V -
0 0
) ( K q (10.80)
Remember from Chapter 5 that the heat conduction inequality in the reference configuration is
given by 0
0
- T
X
,
,
V q . Then, by substituting the linearized heat flow vector given in
(10.80) into the heat conduction inequality we obtain:
( ) 0 0
0 0
- - - T T T
X X X
, , ,
,
V V V K q (10.81)
which in indicial notation becomes:
( ) [ ] ( ) ( ) ( ) 0 0 - -
j
ij
i i j
ij
T T T T
X X X X
, , , ,
V V V V K K (10.82)
or which is the same as:
0
0
T T
X X
, ,
V V K
(10.83)
Thus if we can conclude that the thermal conductivity tensor (
0
K ) is a semi-positive
definite tensor, then, the
0
K -eigenvalues are greater than or equal to zero, i.e. 0
1 0
K ,
0
2 0
K , 0
3 0
K .
Let us now consider a non-symmetric tensor
0
K (anisotropic material) which we will then
split into a symmetric (
sym
0
K ) and antisymmetric (
skew
0
K ) part, so that the inequality in
(10.83) becomes:
( ) 0
0 0 0 0
- - T T T T T T T T
skew sym skew sym
X X X X X X X X
, , , , , , , ,
V V V V V V V V K K K K K
Note that ( ) 0
0 0
T T T T
skew skew
X X X X
, , , ,
V V V V : K K , since the double scalar product
between a symmetric and antisymmetric tensor is equal to zero. Then the above equation
becomes:
0
0
T T
sym
X X
, ,
V V K (10.84)
That is, the antisymmetric part of the thermal conductivity tensor has no influence on
entropy evolution or on the second law of thermodynamics, (Powers (2004)).
We will now take this opportunity to introduce the thermal diffusivity tensor as follows:
0
0
0
1
K D
E
c j

The thermal diffusivity tensor


(Reference configuration)
(

s
m
2
(10.85)
NOTES ON CONTINUUM MECHANICS

560
10.3.1.3 Linearized Entropy
Entropy linearization can be obtained as follows:
) (
) , (
) (
) , (
) (
) , (
) (
) , (
) , (
2 2
O
O O
O
O O
O
O
O O
O
O O
O
T T
T T
T
T
T
T T
T
T T
T
-
o o
o
- -
o o
o
-
- -
o
o
- -
o
o
-
=
E
E E
E
E
E
E E
E
E
E

j
j j
j j
:
:
(10.86)
We can the define the following material properties:
Specific heat (
o
c
E
) at a constant volume, (see equation (10.45)):
T T
T
T
T
T
T c
O O
O
O O
O
o
o o
o
-
o
o

) , ( ) , (
2
E E
E
j
Specific heat (10.87)
Then, by considering both 0
O
E , ) (
O
T T T - A and the material parameters in (10.87)
and (10.68), the linearized entropy becomes:
T
T
c
T
O
o
O
A - -
1
) , (
0
E
E M E :
j
j j (10.88)










Figure 10.2: Curve entropy vs. temperature (linearization).
10.3.1.4 The Helmholtz Free Energy Approach
Note that to achieve consistency that
E
E
o
o


j
0
) , ( T S and
T
T
o
o
-

j ) , (E are linear
functions, the approximation of the Helmholtz free energy must present quadratic terms,
so:
- -
o o
o
- -
o o
o
- - -
o o
o
-
- -
o
o
- -
o
o
-
2
2
2 2
) (
) , (
2
1
) (
) , (
) ( ) (
) , (
) (
2
1
) (
) , (
) (
) , (
) , ( ) , (
O
O O
O
O O
O O
O O
O
O
O O
O
O O
O O
T T
T T
T
T
T
T T
T
T T
T
T T
T T
E
E E
E
E
E E
E E
E
E E
E
E E
E
E
E E




: : :
:


(10.89)
T
j

O
T T

O
j
j
slope:
O
O
T T T -
-

o
o j j j


T
T T
O O
o
o
- -
j
j j ) (

O
O
T
T c ) (
E


T
T c ) (
E

10 THERMOELASTICITY

561
Then, by considering both the mechanical and thermal properties defined in the previous
subsections as well as 0
O
E and ) (
O
T T T - A , we can express the approximation of the
Helmholtz free energy as follows:
2
0 0 0

2 2
1
) , (
1
) , ( T
T
c T
T T T
O
o
e
isoT O
O
O O
A -
A
- - A - -
E
E M E E E E : : : :
j j
j
j
C
_
S
0 S
(10.90)
We must also consider that 0 S
O
at the application point with which the above equation
becomes:
2
0 0

2 2
1
) , ( T
T
c T
T T
O
o
e
isoT O O
A -
A
- - A -
E
E M E E E : : :
j j
j C (10.91)
10.3.1.5 Linearization of the Constitutive Equations
By considering the above in the previous subsections, we can sum up the linearized
constitutive equations for simple thermoelastic materials as follows:
Linearized Constitutive Equations for Simple Thermoelastic Materials
(Reference Configuration)
Constitutive equation for energy
2
0
0

2
2
1
) , (
T
T
c T
T T
O
o
e
isoT O O
A -
A
-
- A -
E
E M
E E E
:
: :
j
j
j C
(10.92)
Constitutive equations for stress T T
e
isoT
A - ) , ( M E E : C S (10.93)
Constitutive equation for entropy T
T
c
T
O
o
O
A - -
1
) , (
0
E
E M E :
j
j j (10.94)
Constitutive equations for heat flux T T
X X
, ,
,
V V -
0 0
) ( K q (10.95)
where
e
isoT
C , M,
o
c
E
,
0
K are the thermo-mechanical material properties, which are
obtained in the laboratory.
10.3.1.6 Linear Thermoelasticity in a Small Deformation Regime
When we are dealing with small deformation regime (infinitesimal strain), the displacement
gradient is very small when compared with the unitary. Moreover, there is no distinction
between the reference and current configurations (see Chapter 7). It is also true that the
Green-Lagrange strain tensor E (reference configuration) and the Almansi strain tensor e
(current configuration) collapse into the infinitesimal strain tensor, i.e. e E = = , and the
same happens to the stress tensors S P = = . Here, we will define the following tensors:
L
M M = (the linear thermal stress tensor) and
L
A A = (the linear thermal expansion
tensor). Note, in the small deformation regime, it is also true that q q
, ,

0
,
0
j j = ,
- - = - V V V
x
X
, ,
. Now, considering all the previous observations, the constitutive
equation for stress given in (10.93) becomes:
) (
0
T T
L e
isoT
- - M : C Duhamel-Neumann equations (10.96)
which is the generalized Hookes law for thermoelastic material, which is also known as the
Duhamel-Neumann equation.
NOTES ON CONTINUUM MECHANICS

562
The linear thermal stress tensor (
L
M ) provides stress in absence of strain, i.e.:
) ( ) ( 0 - with T T
O
L
M (10.97)
Now, the reciprocal of (10.96) is given by:
) (
) (
1 1
1 1 1
O
L e
isoT
e
isoT
O
L e
sym
e
isoT
e
isoT
e
isoT
T T
T T
- -
- -
- -
- - -
M
M
: :
: : : :
C C
C C C C
I


_
(10.98)
Then, considering that
L e
isoT
L
M A :
1 -
- C , (see equation (10.76)) we obtain:
) (
1
O
L e
isoT
T T - -
-
A : C
(10.99)
Next, the linear thermal expansion tensor (
L
A ) provides strain in the absence of stress, i.e.:
) ( ) ( 0 - with T T
O
L
A (10.100)
Note that, here, the material point (particle) can expand freely.
Then, the constitutive equations for heat flux ( q q
, ,
=
0
) can be expressed by Fouriers law of
thermal conduction, which is given by:
T V - K q
,
(10.101)
where K is the thermal conductivity tensor.
Next, the constitutive equation for entropy becomes:
:
L
O
O
o
O
T T
T
c
M
E
0
1
) (
j
j j - - - (10.102)
and the constitutive equation for energy becomes:
2
) (
2
) (
2
1

O
O
o
L O e
isoT O O
T T
T
c T T
T - -
-
- - A -
E
M : : :
j j
j C (10.103)
10.3.1.7 Linear Thermoelasticity in a Small Deformation Regime
In elastic materials, the fourth-order tensor
e
isoT
C , in general, is anisotropic and features 21
independent constants to be determined in the laboratory. For isotropy the number of
constants reduces to 2 and the elasticity tensor can be expressed as follows
I 1 1
isoT isoT
e
isoT
j A 2 - C , where
isoT
A and
isoT
j are isothermal Lam constants. Then, the
constitutive equation for stress, which considers isotropic linear elastic materials and an
isothermal processes, is given by
1
isoT isoT
j A 2 ) ( - Tr (10.104)
Remember that the isotropic second-order tensor is already spherical tensor and vice-versa.
Then, in isotropic materials, the thermal tensors can be represented by the following:
1 K 1 1 K ; ; c
L L
m A M (10.105)
where K is the coefficient of thermal conductivity, m is the thermal stress coefficient and
c is the coefficient of thermal expansion. We can now find the relationship between c
and m by means of the definition in (10.76), thus:
10 THERMOELASTICITY

563
( ) ( )
( ) ( )1 1 1
1 I 1 1 1 1 I 1 1 1
isoT isoT isoT isoT
isoT isoT isoT isoT
L e
isoT
L
m
j A c j A c
j A c c j A
2 3 2 3
2 2
- - - -
- - - -
-
: : :
: A M C
(10.106)
with which we can draw the conclusion that:
c c j A c
) 2 1 (
3 ) 2 3 (
isoT
isoT
isoT isoT isoT
E
m
v -
- x - - -
(10.107)
where we have introduced the isothermal bulk modulus:
3
) 2 3 (
isoT isoT
isoT
j A -
x and
where the following is also true
) 2 1 ( 3
isoT
isoT
isoT
E
v -
x .
Then, the constitutive equations for stress, (see equation (10.96)), becomes:
1 1
1 1
1 1
3 2 ) (
) )( 2 3 ( 2 ) (
) ( 2 ) (
T
T T
T T m
isoT isoT isoT
O isoT isoT isoT isoT
O isoT isoT
A x - -
- - - -
- - -
c j A
j A c j A
j A
Tr
Tr
Tr
(10.108)
When 0 we have:
1 1 ) ( T m T T m
O
A - (10.109)
which represents the spherical state brought about by the temperature change (pressure) in
isotropic materials.
Then, the inverse of the constitutive equation for stress is given by (10.98) and by
considering isotropic materials said equation becomes:
1
1 1
T T
e
isoT
L e
isoT
A - A -
- -
c : : C C A
(10.110)
Then, considering that I 1 1
isoT isoT isoT isoT
isoT e
isoT
j j A j
A
2
1
) 2 3 ( 2
1
-
-
-

-
C , the above
becomes:
1 1
2
1
) (
) 2 3 ( 2
T
isoT isoT isoT isoT
isoT
A - -
-
-
c
j j A j
A
Tr
(10.111)
and in stress-free cases we obtain:
) ( 0 1 A with T c (10.112)
which represents the uniform dilatation case.
Then, the constitutive equations for heat flux (Fouriers Law of thermal conduction) in
isotropic materials become:
T T T V V V K K - - - 1 K q
,
(10.113)
Then, the constitutive equation for entropy (Biots law), in isotropic materials, becomes:
) (
1
1 Tr
j
j
j
j
j
j j
m
T
T
c
m T
T
c m
T
T
c
O
o
O
O
o
O
L
O
o
O
- A - - A - - A -
E E E
M : : (10.114)
and, the constitutive equation for energy turns into:
NOTES ON CONTINUUM MECHANICS

564
[ ]
2 2
2

2
) ( ) (
2
) (
2
1
T
T
c T
m
O
o
isoT
isoT
A -
A
-

-
E

Tr Tr
Tr
j
j
A
j
(10.115)
Then, we can sum up the linearized constitutive equations obtained previously:
Linearized Constitutive Equations for Isotropic Thermoelastic Materials in a
Small Deformation Regime

Constitutive equation for energy
[ ]
2
2
2
2
) (
) (
2
) (
2
1

T
T
c T
m
T
O
o
isoT
isoT
O O
A -
A
-

- - A -
E

Tr
Tr
Tr
j
j
A
j
j

(10.116)
Constitutive equations for stress
1 1
1 1
) 2 3 ( 2 ) (
2 ) (
T
T m
isoT isoT isoT isoT
isoT isoT
A - - -
A - -
j A c j A
j A
Tr
Tr

(10.117)
Constitutive equation for
entropy
) ( Tr
j
j j
m
T
T
c
O
o
O
- A -
E
(10.118)
Constitutive equations for heat
flux
T V K - q
,
(10.119)
Next, we will obtain the relations between the isothermal and adiabatic Lam constants. To
do so, we will use the equations in (10.43), (10.44), and (10.64):
I 1 1 I 1 1
1 1 I 1 1 I 1 1
isoT
o
isoT ise ise
o
isoT isoT ise ise
L L
o
e
isoT
e
ise
c
T m
m m
c
T
c
T
j
j
A j A
j
j A j A
j
2 2
) ( ) ( 2 2
0
2
0
0
-
|
|
.
|

\
|
- -
- - -
-
E
E
E
M M C C

(10.120)
with which we can conclude that:
o
isoT ise
c
T m
E 0
2
j
A A -
and
isoT ise
j j (10.121)
In real materials 0 >
o
c
E
holds with which we can conclude that
isoT ise
A A > .
Now, if we start from the equations in (10.121), we can define other parameters such as the
adiabatic bulk modulus:
( )
o
isoT
o
isoT isoT
isoT
o
isoT
ise ise
ise
c
T m
c
T m
c
T m
E E
E
0
2
0
2
0
2
3 3 3
2 3
3
2 3
3
) 2 3 (
j j
j A
j
j
A
j A
- x -
-

|
|
.
|

\
|
- -

-
x
(10.122)
The latent heat tensors, (of change of strain
E
L , and of change of stress
S
L ), (see
equations (10.51) and (10.52)), in isotropic materials, are given by:
1
0 0
m
T T
L
j j
- - M L
E
; 1 1
S

) 2 3 (
0 0 0 isoT isoT
L
mT T T
j A j
c
j j -
- A L
(10.123)
where we have considered the equation
isoT isoT isoT
m x - - - c j A c 3 ) 2 3 ( , (see Eq.
(10.107)). Additionally, we can obtain:
10 THERMOELASTICITY

565

) 2 3 (
3
) 2 3 (

) 2 3 (
0
2
3 0
2
0 0
0 0
isoT isoT isoT isoT
isoT isoT
T m T m
mT
m
T
T T
c c
j A j j A j
j A j j
j j
-

|
|
.
|

\
|
-
-
|
|
.
|

\
|
- -

1 1
1 1
S S
:
: :
E E
L L
(10.124)
Then, if we start from (10.122), we can obtain:
o
isoT isoT
ise
o
isoT isoT
isoT
isoT
ise
c
T m
c
T m
E E 0
2
0
2
3
1
3 j j x
-
x
x

x
-
x
x

x
x

(10.125)
and by dividing that given in (10.124) by
E
c , we have:
o
isoT isoT isoT
o o
o
isoT isoT
o o
o o
c
T m
c
T m
c
c
c
T m
c
c
c
c
E E E
E E
E
E
0
2
0
2
0
2
3 ) 2 3 (
3
1
) 2 3 (
3
j j A j
j A j
x

-
-
-
-
S
S

(10.126)
with which we can draw the conclusion that:
isoT
ise
o
o
c
c
x
x

E
S

(10.127)
10.4 The Decoupled Thermo-Mechanical
Problem in a Small Deformation Regime
In certain structures (solids) whose temperature variation is not sufficiently high (in the
sense that their mechanical properties do not vary significantly) we can tackle this situation
by means of the decoupled thermo-mechanical problem. That is, at any given time, we can
carry out a thermal analysis without taking into consideration any deformation and then we
can solve the mechanical problem considering initial deformations caused by temperature
change, (see Figure 10.3, and Figure 10.4 and subsection 7.10 in Chapter 7).
As seen previously, the governing equations for simple thermoelastic materials are given by:
Basic Equations of Continuum Mechanics
(Reference Configuration)
The continuity equation 0 ) ( j J
Dt
D
(10.128)
The equations of motion
( ) u b S
u b P
` `
,
`
,
,
` `
,
`
,
`
,
,
,
,
0 0 0 0
0 0 0 0 0
j j j
j j j j
-
-

V F
V V
X
X
V
V
(10.129)
The second Piola-Kirchhoff stress tensor
T
S S or
T T
P P F F
(10.130)
The energy equation
) , ( ) , (
0 0 0
t r t u X E X
X
,
,
`
,
`
,
j j - - q S V :
or ) , ( ) , (
0 0 0
t r t u X F X
X
,
,
`
,
`
,
j j - - q P V :
(10.131)
The Clausius-Plack inequality 0
1
0 0
- - u
T
int
`
`
` j j j E : S D (10.132)
NOTES ON CONTINUUM MECHANICS

566
















Figure 10.3: The decoupled thermo-mechanical problem.

















Figure 10.4: Flowchart of the decoupled thermo-mechanical problem.

No
THERMAL PROBLEM
Obtain the temperature T
Input database
MECHANICAL PROBLEM

i
t

=

i
t
+
1

AR tolerance?
Yes
Yes
No
End
T A

t

=

t
+

t

Load increment
it = 0 (it-iteration)
New load
increment?

_


*
q
,

B

S
u
S
) (
*
x
,
,
t
n
) ( x
,
,
b j
T A
r j
T
B

S
u
S
) (
*
x
,
,
t
n
) ( x
,
,
b j

*
q
,

B
r j

*
T
a) Thermal conduction problem b) Mechanical problem
+
T A
t
t t
dV
dV
dV
dV
10 THERMOELASTICITY

567
Constitutive Equations for Simple Thermoelastic Materials
(Reference Configuration)
The constitutive equation for energy ) , ( T E (1 equation) (10.133)
The constitutive equations for stress
E
E
o
o

) , (
0
T
j S (6 equations) (10.134)
The constitutive equation for entropy
T
T
T
o
o
-
) , (
) , (
E
E

j (1 equation) (10.135)
The constitutive equations for heat flux
) , , (

0 0
T T
X
E
,
,
,
V q q (3 equations)
(10.136)
Then, if we consider a small deformation regime we can make the following
simplifications: S P = = , u
,
sym
V = = e E ,
0
j j = , V V V = =
x
X
, ,
. Additionally, as the
mass density is no longer an unknown the mass continuity equation plays no role. Then, we
can summarize the basic equations for linear thermoelastic materials as follows:
Basic Equations for Linear Thermoelastic Solids
The equations of motion
u b
` `
,
`
,
`
,
,
j j j j - V V V
(10.137)
The energy equation r u j j - - q
,
`
` V : (10.138)
The Clausius-Planck inequality 0
1
- - u
T
int
`
`
` j j j : D (10.139)
without forgetting the linearized constitutive equations obtained previously:
Linearized Constitutive Equations for Isotropic Thermoelastic Materials in a
Small Deformation Regime
The constitutive equation for energy
[ ]
2
0
0
0
2
2
) (
2
) (
) (
) (
2
) (
2
1
T T
T
c T T
m - -
-
-

-
E

Tr
Tr
Tr
j
j
A
j

(10.140)
The constitutive equations for stress
1 1
1 1
) )( 2 3 ( 2 ) (
) ( 2 ) (
0
0
T T
T T m
- - - -
- - -
j A c j A
j A
Tr
Tr

(10.141)
The constitutive equation for entropy
) ( ) (
0
0
Tr
j
j
m
T T
T
c
- -
E

(10.142)
The constitutive equations for heat flux T V - K q
,
(10.143)
10.4.1 The Purely Thermal Problem
Next, let us consider a purely thermal problem in which the equations of motion play no
role. Then, by discarding the terms that are related to strain or stress we obtain:
Basic Equations for Linear Thermal Problems
The energy equation r u j j - - q
,
` V (10.144)
The Clausius-Planck equation
(conservative process)
j j j j j j ` ` ` ` T u u
T
int
- 0
1
D (10.145)

NOTES ON CONTINUUM MECHANICS

568
Linearized Constitutive Equations for Thermal Problems in Small Deformation
Regimes (isotropic material)

The constitutive equation for energy
2
0
0
) (
2
T T
T
c
- -
E

(10.146)
The constitutive equation for entropy
) (
0
0
T T
T
c
-
E
j
(10.147)
The constitutive equations for heat flux T V - K q
,
(10.148)
Then, by substituting the expression ( u` j ), given by the equation in (10.145) into (10.144),
we can obtain the following:
r T j j j - - q
,
` V (10.149)
Next, the rate of change of the entropy
T
T
o
o
-
) (
j can be obtained as follows:
T
T T
T
`
`
o o
o
-
) (
2

j
(10.150)
and by substituting (10.150) into (10.149) we obtain:
r T
T T
T
T
) (
2
j

j - -
o o
o
- q
,
`
V
(10.151)
Additionally, by considering that
T T
T c
o o
o
-

2
E
, we have:
r T c j j - - q
,
`
V
E
(10.152)
Finally, by substituting the constitutive equations for heat flux given in (10.148) into
(10.152) we obtain ( ) r T T c j j - - - V V K
`
E
, or:
( ) T c r T
`
E
j j - V V K The heat flux equation (10.153)
with which we have one equation and one unknown (temperature). Remember from
Chapter 5 that the above equation is the same as that obtained when starting directly from
the principle of conservation of energy, where we also defined the variable r Q j . Now,
to fully describe this problem (which was already discussed in subsection 5.12.1.4 in
Chapter 5) we must add the boundary and initial conditions.
10.4.2 The Purely Mechanical Problem
The effect of temperature in mechanical problems can be addressed by means of the initial
stress/strain as discussed in subsection 7.10. The mechanical problem statement was
described in subsection 7.2, the only difference here lies in the constitutive equations for
stress where we include the thermal effect by means of the initial stress, (see equation in
(10.141)), i.e.:
The constitutive equations for stress (isotropic linear elastic materials):
ij ij ij kk ij
T T
T T
c j A c j c A
j A c j A
) )( 2 3 ( 2
) )( 2 3 ( 2 ) (
0
0
- - - r - r o
- - - - 1 1 Tr
(10.154)
10 THERMOELASTICITY

569
10.5 The Classical Theory of Thermoelasticity in
Finite Strain (Large Deformation Regime)
The classical theory of thermoelasticity in finite strain considers two configurations,
namely: the reference (or initial) configuration
0
B and the current configuration B, (see
Vujoevi&Lubarda (2002) as well as Figure 10.5). The hallmarks of the initial
configuration denoted by
0
B are a stress-free state and an initial temperature distribution
) (
0
X
,
T . In the current configuration (deformed) the stress state is characterized by the
Cauchy stress tensor ) , ( t x
,
and by a temperature distribution denoted by ) , ( t T x
,
.
We have defined the following tensors in the reference configuration: the Green-Lagrange
strain tensor ) , ( t X E
,
, the second Piola-Kirchhoff stress tensor ( ) , ( t X
,
S ), and the heat flux
vector ) , (
0
t X
,
,
q . Then, in the current configuration we have defined: the Almansi strain
tensor ( ) , ( t x e
,
), the Cauchy stress tensor ) , ( t x
,
and the heat flux vector ) , ( t x
,
,
q , (see
Figure 10.5).








Figure 10.5: Reference and current configurations.
Remember from Chapter 5, (see Figure 5.9), that the heat flux vector (conduction) in the
current configuration is related to the heat flux vector in the reference configuration by:
T T
J J F F
- -
=
0
1
0
q q q q
, , , ,
(10.155)
Next, the principle of conservation of energy, by means of the Lagrangian variables, is
given by:
) , ( ) , (
0 0 0
t r t u X E X
X
,
,
`
,
`
,
j j - - q S V : (10.156)
Then, by considering the continuum without any internal heat source, the energy equation
becomes:
0 0
) , ( q S
,
`
,
`
,
-
X
E X V : t u j (10.157)
Next, by considering the Clausius-Duhem inequality we can obtain:
0
1 1
0
0
2
0
0
0 0
0
0 0
- - -
|
|
.
|

\
|
- -
>

_
, ,
_
,
, , ,
` ` T
T T T
r
T T
r
X X X
V V V q q
q
j j j j j j
(10.158)
Note that, here, the entropy production is only caused by the heat flux. Thus
X
,

x
,

F
reference configuration
current configuration

0
B
B

) , (
) , (
) , (
t
t
t
x
x
x e
,
,
,
,
q


) , (
) , (
) , (
0
t
t
t
X
X
X E
,
,
,
,
q
S


) (
) 0 , (
0
X
X
,
,
T
t 0

) , (
) , (
t T
t
x
x
,
,


NOTES ON CONTINUUM MECHANICS

570
0 0
1
q
,
,
` -
X
V
T
j j (10.159)
Therefore, in reversible processes, we have:
0
0
1
q
,
,
` -
X
V
j
j T
(10.160)
Remember that an alternative way to express the Clausius-Planck inequality is as follows:
0
0

(

- -
Dt
D
Dt
DT
int

j j E
`
: S D (10.161)
where is the specific Helmholtz free energy (per unit mass).
Then, for processes with no internal energy dissipation 0
int
D (reversible), we can state
that:
T
` ` `
j
j
- E : S
0
1

(10.162)
where ) , ( T E is a thermodynamic potential used to obtain the stress tensor ( S ) and
entropy ( j ):
T
T
` ` `
o
o
-
o
o


E
E
: (10.163)
Now, by comparing the equations (10.162) and (10.163), we can conclude, as expected,
that:
T
T T
o
o
-
o
o

) , (
;
) , (
0
E
E
E
j

j S (10.164)
10.5.1 The Coupled Heat Flux Equation
Let us now suppose that the heat flux vector is governed by Fouriers law of thermal
conduction:
T T T T
X
x
F
, ,
, ,
V V - - ) , ( ; ) (
0 0
K q K q (10.165)
where the thermal conductivity tensors K and
0
K are interrelated by:
T
T T
- -
F F F F ) ( ) ( ) , (
1
0
K K det (10.166)
We can now prove the above equation is valid by starting from the following equation:
A a
,
,
,
,
d d
0
q q

(10.167)
where A
,
d and a
,
d are the differential area elements in the reference and current
configuration, respectively, and which are related to each other by means of the Nansons
formula A F a
,
,
d J d
T

-
where F J . The following is then satisfied:
0
0
0
q q
q q
q q
, ,
,
,
,
,
,
,
,
,

-
-



T
T
J
d d J
d d
F
A A F
A a

0
q q
, ,

-

T
J F
k k ki i
k k k ki i
k k i i
dA F J
dA dA F J
dA da
0
1
0
1
0
q q
q q
q q

-
-

k ki i
F J
0
1
q q
-

(10.168)
10 THERMOELASTICITY

571
Then, by substituting the equation in (10.165) into the above we obtain:
T J T
T J T
J
T
T
x
X
x
X
F
F
F
, ,
, ,
, ,
V V
V V



-
-
-

- -




1
0
0
0
K K
K K
q q

k jk ij Q IQ
ij k jk Q IQ
ij j I
T JF T
F T J T
F J
, ,
, ,
1
0
1
0
1
0
K K
K K
q q
-
-
-

- -


(10.169)
Note that the following is valid:
( )
1
1 1
,
) , (
) , (
) ), , ( (
,
) , (
1
-
- -

o
o

o
o
o
o
=
o
o

-
F
X
X
X x x
X
x
x
T T
F T F
X
t T
x
t X
X
t t T
T
x
t T
T
Qk Q Qk
q
F
k
q
q
k
k
k
qk
, ,
,
,
_
,
,
, ,
V V
V
(10.170)
with which the equation in (10.169) becomes:
T J T
T J T
T J T
T
X X
X X
x
X
F F
F F
F
, ,
, ,
, ,
V V
V V
V V



- -
- -
-


) (

1
0
1 1
0
1
0
K K
K K
K K

Q Qk jk ij Q IQ
Qk Q jk ij Q IQ
k jk ij Q IQ
T F JF T
F T JF T
T JF T
, ) ( ,
, ,
, ,
1 1
0
1 1
0
1
0
- -
- -
-

K K
K K
K K
(10.171)
and:
T
J
- -
F F K K
1
0
(10.172)
which thus proves the equation in (10.166) is valid.
Then, by substituting Fouriers law (10.165) into the equation in (10.160) we obtain:
[ ] [ ]
QI IQ Q I IQ
I
Q IQ I I
T T T q T
, 0 , , 0
0
,
, 0
0
, 0
0
1 1 1
K K K - - - -
j j j
j`
(10.173)
where [ ]
iq
IQ
q i i q
QI
T T
X X
T
X X
T
T ) (
,
2 2
,
X X
, ,
V V
|
|
.
|

\
|
o o
o

|
|
.
|

\
|
o o
o
is satisfied and
( ) T T
T X
T
X
T
T X
T
X
T
q i
iq
q i
iq
Q I IQ X X
, ,
V V
o
o

o
o
o
o
o
o

o
o
o
o
:
0
0 0
, , 0
K
K K
K with which the equation in
(10.173) becomes:
( ) [ ]
)
`

-
o
o
) (
1
0
0
0
T T T
T
T
X X X X
, , , ,
` V V V V : : K
K
j
j
(10.174)
It can now be shown that the previous equation in the current configuration is given by:
( ) [ ]
)
`

-
o
o
) (
1
T T T
T
T
x x x x
, , , ,
` V V V V : : K
K
j
j
(10.175)
and the rates of change of the equations in (10.164) are given by:
|
|
.
|

\
|
o o
o
-
o o
o
-
o
o
-
|
|
.
|

\
|
o o
o
-
o o
o

o
o

T
T T
T
T
T
T
T
T
T
T T T
rate
rate
` `
` ` `
`
) , ( ) , ( ) , (
) , ( ) , ( ) , (
2 2
2 2
0 0
E
E
E
E E
E
E
E
E E
E
E
E

j

j

j

j
:
: S S

(10.176)
which yields:
NOTES ON CONTINUUM MECHANICS

572
T
T
T
T
T T
T T
T T
` ` ` `
`
2
2
2
2 2
o
o
- |
.
|

\
|
o
o
o
o
-
|
|
.
|

\
|
o
o
-
o o
o
-

j E
E
E
E
: : (10.177)
Then, if we consider that
E
E
o
o

) , (
0
T
j S , the above equation becomes:
( ) T
T
T
T
T
T
T
T
T
T T
` ` ` `
`
2
2
0
2
2
o
o
-
o
o
-
o
o
- |
.
|

\
|
o
o
o
o
-

j

j E E
E
: : S (10.178)
We can now define a second-order tensor denoted by the latent heat tensor of change of
strain, (see equation (10.51)), and given by:
E
L
0 E
E
o o
o
- |
.
|

\
|
o
o
-

T
T
T
T
j
2
0
`
S
Latent heat tensor of change of strain (10.179)
and the specific heat at a constant volume, (see subsection 10.2.2):
0 E
E

|
|
.
|

\
|
o
o
-
`
2
2
T
T c

Specific heat at a constant volume (10.180)
with which we can rewrite the equation in (10.178) as follows:
T c T
` `
`
E E
E L - : j (10.181)
In Chapter 2 we obtained the relationship between the rate of change of the Green-
Lagrange strain tensor ( E
`
) and the rate-of-deformation tensor ( D ) as follows:
F F E D
T
`
, which is expressed in indicial notation as
qj pq pi ij
F F E D
`
with which we
can conclude that ( ) D :
T
pq qj ij pi qj pq pi ij ij ij
F L F F F L E L F L F
E E E E
D D
`
. Additionally,
by starting from the definition of
E
L given in (10.179) we can draw the conclusion that
( )
T
T
T
T
o
o -



F F
F L F
E
S
0
j
. Also, in Chapter 2 we obtained the relationship between
the second Piola-Kirchhoff stress tensor ( S ) and the Cauchy stress tensor ( ) as follows:
S J
T
F F , where
0
j j J holds and by taking all of the above into consideration the
equation in (10.181) can be rewritten as:
( )
( )
T c
T
T
T c
T
JT
T c
T
T
T c T c T
T
T
` `
` ` ` `
`
E E
E E E E E
F F
F L F E L
-
|
|
.
|

\
|
o
o
- -
|
|
.
|

\
|
o
o
-
-
|
|
.
|

\
|
o
o
- - -


D

D
S
D
: :
: : :
j j
j
j
0
0
(10.182)
10.5.2 The Specific Helmholtz Free Energy
Let us now assume that the stress tensor S varies linearly with the strain tensor E , and
also that the specific heat and the latent heat tensor vary linearly with temperature
according to:
0
0
0
0
; ) (
E E E E
L L
T
T
T T c c c - -
(10.183)
where c is a constant while
0
E
c and
0
E
L are the values of
E
c and
E
L , respectively, at
0
T T and 0 E . Then, we can represent the Helmholtz free energy (per unit mass):
10 THERMOELASTICITY

573
( ) ( )
2
0
0
0 0
0
0
0 0
0
0
) (
2 2
1
T T
c
T
T
T T T cT c
T
T T
e
- -
(
(

|
|
.
|

\
|
- - - -
-
- ln
E E
E L E E : : : C
j

(10.184)
where
e
0
C is the isothermal elasticity tensor (a symmetric fourth-order tensor).
The constitutive equations for stress are given by:
( )
0
0
0
0
0
0 E
L E
T
T T
T
e
-
-
j
: C S
(10.185)
and the constitutive equation for entropy by:
( ) ( )
0
0
0
0 0
0

1
T T c
T
T
cT c
T
- -
|
|
.
|

\
|
- - ln
E E
E L : j
(10.186)
Then, if we consider:
0
0
0
0 0 E
L A
T
e
j
: C
(10.187)
where
0
A is the thermal expansion tensor defined in (10.76), we can obtain:
( )
0 0
1
0
T T
e
- -
-
A E S : C
(10.188)
Next, in isotropic linear elastic materials, the following is satisfied:
1 1 I 1 1
0 0
0
0
0
0 0 0 0 0
3 ; ; 2 x - c
j
c j A
E
L A
T
e
C
(10.189)
Then, the constitutive equations for stress and entropy become:
1 1
1 1 S
) )( 2 3 ( 2 ) (
) ( 3 2 ) (
0 0 0 0 0 0
0 0 0 0 0
T T
T T
- - - -
x - - -
j A c j A
c j A
E E
E E
Tr
Tr

(10.190)
( ) ( )
0
0
0
0
0 0
0
) (
3
T T c
T
T
cT c - -
|
|
.
|

\
|
- - x ln Tr
E
E c
j
j
(10.191)
and the Green-Lagrange strain tensor is given by:
1 1 S S ) ( ) (
3 2
1
0 0
0
0
0
T T - -
(

x
- c
A
j
Tr E
(10.192)
10.6 Thermoelasticity based on the
Multiplicative Decomposition of the
Deformation Gradient
Now, we will tackle thermoelasticity in finite strain, in this subsection, by multiplicative
decomposition of the deformation gradient into elastic (
e
F ) and thermal parts (
0
F ),
Lubarda(2004), (see Figure 10.6). According to Vujoevi&Lubarda(2002), this approach to
describing thermal problems was first introduced by Stojanovi. The first transformation is
caused by
0
F , which then defines the intermediate configuration
0
B , which is
characterized by the absence of stress. After the transformation
0
F takes place, we can
NOTES ON CONTINUUM MECHANICS

574
carry out the transformation caused by
e
F . In this way, the deformation gradient can be
represented by:
0
F F F
e
(10.193)











Figure 10.6: Multiplicative decomposition.
Thus, as proven in the chapter on plasticity, (see Chapter 9), multiplicative decomposition
of the deformation gradient into either a plastic and elastic part (
p e
F F F ) or thermal
and elastic parts is not unique.
10.6.1 Kinematic Tensors
Then, if we consider the transformations
e
F and
0
F , we can define the elastic and
thermal Lagrangian strain tensors as follows:
|
.
|

\
|
- |
.
|

\
|
- 1 1
0 0 0
F F E F F E
T
e
T
e e
2
1
;
2
1
(10.194)
thus, the rate of change of
0
E can be evaluated as such:
|
.
|

\
|
-
0 0 0 0 0
F F F F E
` ` `
T T
2
1
(10.195)
Then, by considering that ( ) 1 - F F E
T
2
1
we can obtain the following equation:
( ) |
.
|

\
|
- |
.
|

\
|
- - - -
0 0 0 0 0
F F F F F F F F E E
T
T
T
T
2
1
2
1
2
1
1 1 (10.196)
and by using the multiplicative decomposition shown in (10.193),
0
F F F
e
, the above
can still be written as:
OBS.: In this subsection we use the parameter 0 instead of T to represent
temperature so as to avoid confusion with the transpose
T
F .
Inter. Conf.
X
,

x
,

X
,


0
F

e
F

0
F F F
e

reference
configuration
current
configuration

0
B

0
B
B
intermediate
configuration

0
) , ( t X
,
0

Ref. Conf.
0
) , ( ); , ( J t t x x
, ,

) (
0
X
,
0
- Cauchy stress tensor
- Kirchhoff stress tensor
10 THERMOELASTICITY

575
( ) ( ) ( )
( )
0 0
0 0 0 0 0 0
0 0 0 0 0 0 0
F E F
F F F F F F F F F F
F F F F F F F F F F E E


|
.
|

\
|
- -
- - -
e
T
e
T
e
T T
e
T
e
T
T
e
T
e
T
T
2
2
1 (10.197)
with which we can draw the conclusion that:
0 0 0
F E F E E -
e
T

( )
1 - -
-
0 0 0
F E E F E
T
e

(10.198)
Then, the rate of change of
e
E can be evaluated as follows:
( ) ( ) ( )
1 1 1 - - - - - -
- - - - -
0 0 0 0 0 0 0 0 0
F E E F F E E F F E E F E
` ` ` ` `
T T T
e

(10.199)
Remember that the rate of change of the deformation gradient is given by F F l
`
,
where l is the spatial velocity gradient, with which the equation l
- -
-
1 1
F F
`
holds.
Likewise, we can define the following tensors as:
0 0 0
0 0 0
0 0 0
l
l
l

- -
-
-



1 1
1
F F
F F
F F
`
`
`
and
(10.200)
Now, returning to (10.199) and by taking into account the previous equations in (10.200),
we can state that:
( ) ( )
( )
0 0 0 0
0 0 0 0 0 0 0
l
l


- -
- - - -
-
- - - - -
1
1 1
F E E F
F E E F F E E F E
T
T T T
e
` ` `
(10.201)
We can now substitute the term ( )
0
E E - given by the equation in (10.198) into the above:
( )
0 0 0 0 0
0 0 0 0 0 0 0 0
l
l


- -
- - - -
- - - -
1
1 1
F F E F F
F E E F F F E F F E
e
T T
T
e
T T T
e
` ` `
(10.202)
the result of which is:
( )
0 0 0 0 0 0 0
0 0 0 0 0
l l
l l


- - - -
- - - -
- - - -
- -
e
T T
e
T
e
T
e
T
e
E F E F F E F E
E F E E F E E
1 1
1
` `
` ` `
(10.203)
Then, regarding the term
1 - -

0 0 0
F E F
`
T
, we can substitute
0
E
`
by means of the
definition in (10.195), thus:
0 0 0 0
0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0 0
D =
(

-
(

|
.
|

\
|
- |
.
|

\
|
|
.
|

\
|
-
|
.
|

\
|
-
|
.
|

\
|
-
- - - -
- - - -
- - - -



sym T
T
T T
T T T T
T T T T
l l l
2
1
2
1
2
1
2
1
2
1
1 1 1
1 1
1 1
F F F F F F F F
F F F F F F F F
F F F F F F F E F
` ` ` `
` `
` ` `
(10.204)
Hence, the equation in (10.203) becomes:
sym
e e
T T
e 0 0 0 0 0
l l l - - -
- -
E E F E F E
1
` `

(10.205)
NOTES ON CONTINUUM MECHANICS

576
















Figure 10.7: Kinematic and stress tensors in different configurations.
10.6.2 The Stress Tensor
Remember from Chapter 3 that the second Piola-Kirchhoff stress tensor is related to the
Cauchy stress tensor by means of the equation
T - -
F F F S
1
. Then, by using the
multiplicative decomposition
0
F F F
e
we can obtain:
( ) ( )
T
e
T T
e e e
T
e e e T
- - - - - -
- -
- -




0 0 0 0 0 0
0 0 0
F F F F F F F F F
F F F F F F F F F
S
S
1 1 1
1
1
(10.206)
where we have introduced:
T
e e e e
- -
F F F S
1
(10.207)
10.6.3 Area and Volume Elements
If we now consider the definition of the Jacobian determinant and the multiplicative
decomposition of the deformation gradient, we can draw the conclusion that:
0 0 0
J J J
e e e
) ( ) ( ) ( ) ( F F F F F det det det det (10.208)
Therefore, we can define the thermal Jacobian determinant
0
J and the elastic Jacobian
determinant
e
J , respectively, as follows:
[ ]
2
1
2
1
1
) ( ) ( ; ) ( ) (
0 0 0 0
b F C F det det det det
(


-
e e
J J
(10.209)
X
,

x
,

X
,


0
F
e
F

0
F F F
e

reference
configuration
current
configuration

0
B

0
B
B
intermediate
configuration

0
0

0

1 -
F F
`
l

( )
T
T
- -

-
F F F
F F E
S
1
1
2
1

|
.
|

\
|
- 1
0 0 0
F F E
T
2
1


T
e e e e
e
T
e e
- -

|
.
|

\
|
-
F F F
F F E
S
1
1
2
1



sym
0 0
0 0 0
l
l
=

D
1
F F
`


( )
1 - -
-
0 0 0
F E E F E
T
e

10 THERMOELASTICITY

577
Then, the differential volume elements in the respective configurations, (see Figure 10.8),
are given by:
) , ( ) , ( ; ) , ( ) , (
0
t V d J t dV t dV J t dV
e
X x X X
,
,
,
,

0 0

(10.210)
Remember from Chapter 2 that the material time derivative of the volume element is given
by dV dV
Dt
dV D
) ( ) (
) (
D Tr Tr l . Likewise, we can obtain the rate of change of
0
dV as:
0 0 0 0
0
dV dV
Dt
dV D
) ( ) (
) (
D Tr Tr l
(10.211)
Let us now consider a differential area element in the reference configuration
0
A
,
d , (see
Figure 10.8). Next, we will define the differential area elements in different configurations:
0
A F A
, ,
d J d
T

0 0 0
Differential area element in the
intermediate configuration
(10.212)
and
0
A F a
,
,
d J d
T
e e

Differential area element in the


current configuration
(10.213)
Remember from Chapter 2 the transformation between
0
A
,
d and a
,
d is given by the
equation
0
A F a
,
,
d J d
T

-
(Nansons formula). Hence, the following is valid:
( )( )
0 0
0 0
A F A F F
A F F A F F A F a
, ,
, , ,
,
d J d J J
d J J d J J d J d
T
T
e e
T T
e e
T T
e e
T
e e


-
-
- - - - -

|
.
|

\
|

0 0
0 0 0 0 0
(10.214)
















Figure 10.8: Area and volume elements deformations.

0
A
,
d
X
,

x
,

X
,

0
dV J dV
0 0


e
F

0
F F F
e

reference
configuration
current
configuration

0
B
B
intermediate
configuration

0
F

0
dV
dV

0
dV

0
dV J dV
e


0
dV J dV

0
A
,
d

0
A F A
, ,
d J d
T

0 0 0


0
A F a
,
,
d J d
T
e e

-

a
,
d

0
B
NOTES ON CONTINUUM MECHANICS

578
10.6.4 Isotropic Materials
Let us now consider an isotropic material in which the thermal deformation gradient is
represented by:

0
-
(inverse)
1
(rate)
1
1
1 1
1
0
0 0
0
0
0
0
0
0
0
0
F
F
F
` `
d
d
dt
d
d
d
(10.215)
where the scalar ) (0
0 0
0 0 is the thermal stretch coefficient. Note, an expression for
0
0 was
proposed by Lu&Pister(1975) which is given by:
|
|
|
.
|

\
|
}
0
0
0
0 0 c
0
0

( d
exp
(10.216)
where ) (0 c c
0 0
is the linear thermal expansion coefficient, and where
0
0 and 0 show
the temperature in the configurations
0
B and B, respectively.
Now, if we consider the equation in (10.215) we can express
0
E as follows:
( ) 1 1 1 1 1 ) 1 (
2
1
2
1
2
1
2
- 0 - 0 0 |
.
|

\
|
-
0 0 0
0 0 0
F F E
T
(10.217)
and its rate of change by:
1 1 1 2
2
1
) 1 (
2
1
2
0
0
0
0
0
0 0
0 0
0
` `
0
0

0
0
(

- 0
d
d
dt
d
d
d
dt
d
E
(10.218)
Then, if we consider the equation in (10.198) we are given:
e e e
T
E E F E F E E
2
0 0 0
0 0 0
0 0 0 - 1 1
(10.219)
or
( )
0
0
E E E -
0

2
1
e

(10.220)
Next, by substituting the expression of
0
E given in (10.217) into the above equation we
obtain:
( ) |
.
|

\
|
- 0 -
0
-
0
1 ) 1 (
2
1 1 1
2
2 2
0
0 0
0
E E E E
e

(10.221)
or
( ) 1 1 1 1 - - 0 - 0 - 0 E E E E 2 2 2 2
2 2 2 e e
0 0 0
(10.222)

( )
e
E E 2 2
2
- 0 - 1 1
0
(10.223)
Then, the relationship between
0
0 and the thermal expansion coefficient,
0
c , is given by:
0
0 c c
0
0
0 0
d
d0
0

1
) (
(10.224)
Therefore, we can express the rate of change of the elastic strain tensor,
e
E
`
, by means of:
10 THERMOELASTICITY

579
( ) ( ) ( )
( ) ( )
( ) ( ) [ ] ( ) [ ]
( ) [ ] 0
0 0
0 0
0
0
0
0
0 0
0
0 0 0
0
0 0
0
0 0 0 0
0 0
0
0
0
0
0 0 0
c
c c
c c
0
0
0 0 0
` `
` ` ` `
` ` `
` ` ` ` ` `
2
1
2
1
2
1

1
2
1

1 1
2
1 1
2
2
2 2
2
2
2
2
2
2 3 2 2
e
e
e
d
d
d
d
dt
d
E E
E E E E E
E E E
E E E E E E E E
- 0 -
0

0 - 0 -
0
0 - - -
0

0 - - - 0
0
-
0

|
.
|

\
|
0
0
-
0
- -
0
0
- -
0
- -
(
(

1
1 1
1
1
(10.225)
and by using the equation in (10.223), we can obtain:
( ) [ ] 0 c
0
0
` ` `
E E E 2
1
2
- -
0
1
e

(10.226)
Now, in isotropic materials, the spatial velocity gradient
0
l , (see equation (10.200))
becomes
1 1 1
1 1

1
0
0
0
0
0
0 0
0 0 0 0
` ` `
d
d
d
d 0
0

|
|
.
|

\
|
0
|
.
|

\
| 0

-
F F l (10.227)
where we have used the expressions of
0
F
`
and
1 -
0
F given in (10.215) with which we can
obtain the rate of change of the differential volume element (
0
dV ) as follows:
0 0 0 0
0
0 0
0
0
0
0
c dV dV
d
d
dV
Dt
dV D
` `
3
3
) (
) (

0
0
l Tr (10.228)
10.6.5 The Constitutive Equations
10.6.5.1 The Constitutive Equation for Energy
Here, we will assume the specific Helmholtz free energy, (per unit mass), is as follows:
) ( ) , ( 0 0
0
-
e e
E (10.229)
where
e
is given in terms of the strain tensor
e
E and temperature 0 , (see Figure 10.9),
and
0
can be adjusted according to the experimental results of specific heat, (see
Lubarda(2004)). Next, we will evaluate the rate of change of the specific Helmholtz free
energy:
0
0

0
0

0
` ` ` `
d
d
e
e
e
e
-
o
o
-
o
o
E
E
: (10.230)
Now, by substituting (10.226) into the above equation we obtain
( ) [ ] 0
0

0
0

0 c

0
0

0
0

0 0
0
0
` ` ` ` ` ` ` `
d
d
d
d
e
e
e e
e
e
e
-
o
o
- - -
0 o
o
-
o
o
-
o
o
E E
E
E
E
2
1
2
1 : : (10.231)
and on simplifying we find:
( ) 0
0

0
c

0
0
0
0
` ` `
(
(

-
o
o
- -
o
o
0
-
o
o
0

d
d
e
e
e
e
e
E
E
E
E
2
1
2 2
1 : : (10.232)
NOTES ON CONTINUUM MECHANICS

580
Then, starting from the Clausius-Planck inequality, we can express the rate of change of the
energy as follows:
0 j
j

0 j j j j 0 j j
` `
` ` ` `
`
` `
- s
- - - - -
E
E E
:
: :
S
S S
0
0 0 0 0
1
0
(10.233)
and by comparing the equations (10.233) and (10.232), we obtain:
( )
0

0
c
j
j
0
0
0
0
d
d
e
e
e
e
e
-
o
o
- -
o
o
0

o
o
0

E
E
E
2
2
2
0
1
S
:
(10.234)















Figure 10.9: Energy, stress and mass density in different configurations.
10.6.5.2 The Constitutive Equations for Stress
If we now consider the relationship between mass density
0
j (reference configuration) and
0
j (intermediate configuration):

> 0 0

0
3
0
0
0 0
0 0
0
0
j j
j j
1 F
F
J
J
J
J
(10.235)
we can obtain
0
j j
0
3
0
0 (10.236)
Now, the constitutive equations for stress (10.234) can be rewritten as follows:
X
,

x
,

X
,


0
F
e
F

0
F F F
e

reference
configuration
current
configuration

0
B

0
B
B
intermediate
configuration
0

0
j

e
S S
0
0

e
e
e
E o
o


j
0
S
0
j
) (0
0

) , ( 0
e e
E
0
j j
0
3
0
0
0
0
10 THERMOELASTICITY

581
_
e
e
e
e
e
S
S
E E o
o
0
o
o
0


j
j
0 0
0

2
0

(10.237)
or:
e
e
e e
with
E o
o
0

j
0 0
S S S

(10.238)
Note that the equation
e
S S
0
0 could have been directly obtained by means of that in
(10.206), i.e.:
( ) ( )
e T e
T
e
S 1 S 1 1 S S
0 0 0 0
0 0 0
0 0 0 0
- -
- -

1
1
F F F (10.239)
Let us now suppose that
e
is a quadratic function in terms of the elastic strain tensor as
follows:
[ ]
e e e e
E E E :
0 0
j A j
0
-
2
) (
2
1
Tr (10.240)
where ) (0 A A
0 0
and ) (0 j j
0 0
are the Lam constants which are dependent on
temperature, and:
I 1 1 S
0 0 0 0
j A 2 -
e e e e
with C C E : (10.241)
where ) (0
0 0
e e
C C is the thermal elasticity tensor, and I is the symmetric fourth-order unit
tensor whose components are ( )
jk il jl ik ijkl
c c c c -
2
1
I .
Now, going back to the equation in (10.241) we can obtain:
[ ]
e e
sym
e
e
e
e e e e e
E E
E E E E
E
E
0 0
0 0 0 0 0
j A
j A j A
2 ) (
2 2
) (
-
- -

1
I 1 1 I 1 1 S
Tr
Tr
_ _
: : : : C
(10.242)
and if we then consider the expression of
e
E given in (10.221) the following is valid:
(

- 0 -
0
|
.
|

\
|
- 0 -
0
) 1 (
2
3
) (
1
) ( ) 1 (
2
1 1
2
2
2
2
0
0
0
0
E E E E Tr Tr
e trace e
1
(10.243)
Then, by substituting the above into the equation in (10.242) and by considering that
e
S S
0
0 we obtain:
[ ]
[ ] [ ]1 1
1 1
1 S S
2 3
) 1 (
2
1
2 ) (
1
) 1 (
2
1 1
2 ) 1 (
2
3
) (
1
2 ) (
2
2
2
2
2
0 0
0
0
0 0
0
0
0
0 0
0
0 0
0 0 0 0
j A j A
j A
j A
-
0
- 0
- -
0

- 0 -
0
-
(

- 0 -
0
0
- 0 0
E E
E E
E E
Tr
Tr
Tr
e e e

(10.244)
Now, if we consider that the bulk modulus is dependent on temperature ( ) (0
0 0
x x ),
which is given by
0 0 0
j A 2 3 3 - x , the above then becomes:
NOTES ON CONTINUUM MECHANICS

582
[ ] 1 1 S
0
0
0
0 0
0
j A x
0
- 0
- -
0

) 1 (
2
3
2 ) (
1
2
E E Tr (10.245)
If we then take the following approach ) ( 1 ) (
0 0
0 0 c 0
0 0
- - = 0 0 (where
0
c is the linear
thermal expansion coefficient) and if in addition to that we assume that the Lam constants
are independent of temperature, the stress equation given in (10.245) becomes:
1 1
1 1 S
) )( 2 3 ( 2 ) (
) ( 3 2 ) (
0 0 0 0 0 0
0 0 0 0 0
0 0 j A c j A
0 0 c j A
- - - -
x - - -
E E
E E
Tr
Tr

(10.246)
10.6.5.3 The Constitutive Equation for Entropy
According to (10.234) we can express the constitutive equation for entropy as:
( ) ( )
( ) [ ]
0

j
c
j
c
0

j
c
0

j
c
0

0
c
j
0 0
0 0
0 0 0 0
0
0
0
0
0
d
d
d
d
d
d
d
d
e
e
e
e
e e
e
e
-
o
o
-
0
-
0
-
o
o
- - 0
-
o
o
- - -
o
o
- -
o
o
0

E E
E E
E
: : :
: :
S 1 S 1 S
1 S 1
0
2
0
2
2
0
0
2
2
2
2 2
(10.247)
where we have used the equation in (10.223): ( )
e
E E 2 2
2
- 0 - 1 1
0
. Note also that the
following holds:
[ ]
[ ]
e e e
e e
e e e e
E E E
E E
E E
:
: :
1
1 S 1 S 1 S
0 0 0 0
0 0
0 0
0
j A
j A
j A
x x -
-
-
0
3 ) ( 3 ) ( 2 3
) ( 2 ) ( 3
2 ) ( ) (
1
Tr Tr
Tr Tr
Tr Tr Tr

(10.248)
Now, returning to the equation in (10.247) we obtain:
0

j
c
j
c
0

j
c
j
c
j
0
0
0 0 0 0 0
0 0 0 0
d
d
d
d
e
e e
e
e
-
o
o
-
0
-
x 0

-
o
o
-
0
-
0

E E
E
: :
: :
S 1
S 1 S
0
2
0
3
0
2
0
2
2 3
2
(10.249)
Note, the above term
0

o
o
e
can be obtained by starting from the energy equation:
e e e e e e
E E : : S S
2 2
1
3
0
0
j j
0
0

(10.250)
where we have used the equation in (10.236). Next, we will obtain the derivative of
e

with respect of the temperature 0 :
e
e
e e
e
d
d
E E : :
0 0 0
0 0
0

j
o
o 0
-
0
0
o
o S
S
2 2
3
3
2
0

(10.251)
Now, if we consider the equation in (10.224) and
e
S S
0
0 , we can state that:
e
e
e e
e
e e
e
E E E E : : : :
0 0 0
0
0 0
0
0 0
c c

j
o
o 0
- 0
o
o 0
- 0
o
o S
S
S
S
2 2
3
2 2
3
3
2
3
3
0

(10.252)
and by using (10.221), the above equation becomes:
10 THERMOELASTICITY

583
e
e
e
e e
E E
E E
: :
: :
0
0 0
0
0 0
0
0
0
0 0
c
c

j
o
o 0
-
(

- 0 -
o
o 0
-
(
(

|
.
|

\
|
- 0 -
0
0
o
o
S
S S
S
1 S
2
) ( ) 1 (
2
1
2
3
2
) 1 (
2
1 1
2
3
3
2
3
2
2
2
0
Tr
(10.253)
Now, by substituting (10.252) into the entropy equation given in (10.249) we obtain:
0

c c
j
0

c c
j
0

c c c
j
0

j
c
j j
c
j
c
0

c
j j
c
j
c
0

j
c
j
c
j
0
0
0
0
0
0
0
0
0
0
0
0 0 0
0
0 0 0 0 0 0
0 0 0 0 0 0 0
0
0 0
0 0 0 0 0
0
0 0
0 0 0 0 0
0 0 0 0 0
d
d
d
d
d
d
d
d
d
d
d
d
e
e
e
e
e
e
e
e
e
e
e
e e e
e
e e
-
(

o
o
- - x
0

-
(

o
o
0 - 0 - x 0
-
(

o
o
0 - 0 - 0 - x 0
-
(
(

o
o 0
- 0 -
0
-
x 0

-
(
(

o
o 0
- 0 -
0
-
x 0

-
o
o
-
0
-
x 0

E
E
E
E
E E E E
E E
:
:
:
:
: : : :
: :
S
S 1
S
S 1
S
S S 1
S
S S 1
S
S S 1
S 1
6
2
6
2
1
3 4 6
2
1
2 2
3 2 3
2 2
3 1 2 3
2 3
0
3
3 2 3
0
3 2 2 3
0
0
3
2
0 0
2
0
3
3
2
0 0
2
0
3
0
2
0
3
(10.254)
In addition, by considering that
0
j j
0
3
0
0 , the above equation becomes:
0

c c
j
j
0
0
0
0 0 0
d
d
e
e
e
-
(

o
o
- - x E :
S
S 1 6
2
1
(10.255)
10.7 Thermoplasticity in a Small Deformation
Regime
In this subsection we will extend the classical theory of plasticity, (see Chapter 9), in which
temperature is included as a free variable.
10.7.1 The Specific Helmholtz Free Energy
In a small deformation regime, the specific Helmholtz free energy is a function of the
following free variables: the infinitesimal strain tensor , the temperature T , and the
internal variables
k
, i.e.:
) , , (
i
T (10.256)
As we have discussed in subsection 9.4.1, we can reformulate the energy equation in order
to obtain the elastic part of strain
e
as a free variable, i.e.:
) , , (
i
e
T (10.257)
NOTES ON CONTINUUM MECHANICS

584
Now the set of internal variables
i
do not include the plastic part
p
of the strain tensor,
since this is already included in the free variable
p e
- . Then, the following is
satisfied:
p e
o
o
-
o
o

o
o
(10.258)
Finally, the rate of change of the free energy given in (10.257) becomes:
k
k
e
e
T
T

`
` `
`

o
o
-
o
o
-
o
o

:
(10.259)
10.7.2 Internal Energy Dissipation
Let us consider the Clausius-Duhem inequality:
0
1
- - - T
T
T
int
V q D
,
` `
j jj : D (10.260)
Now, by substituting the rate of change of energy given in (10.259) into the above
inequality and by considering that
p e
D
` ` `
`
- = E (in a small deformation regime), we
obtain:
0
1
) ( -
(

o
o
-
o
o
-
o
o
- - - T
T
T
T
T
k
k
e
e
p e
int
V q


,
`
`
`
` `
`



j jj : : D
(10.261)
which on simplifying yields:
0
1
- - |
.
|

\
|
o
o
- - - |
.
|

\
|
o
o
- T
T
T
T
k k
p e
e
int
V q

,
`
` `
` A

j j

j : : D (10.262)
where
k
k
o
o
-

j A are the thermodynamic forces. As the above inequality must be valid
for any thermodynamic process, (see Chapter 6), we obtain
e

o
o


j ,
T o
o
-

j . Then, we
can summarize this model as follows:
k
i
e
k
i
e
e
i
e
T
T
T T


o
o
-
o
o

o
o

) , , (
;
) , , (
;
) , , (

j A (10.263)
and the internal energy dissipation becomes:
0
1
- - T
T
k k
p
int
V q
,
`
` A : D (10.264)
which we can the split into mechanical and thermal parts, i.e.:
thermal mechanical int
D D D - (10.265)
where:
0 -
k k
p
mechanical
`
`
A : D and 0
1
- T
T
thermal
V q
,
D (10.266)
Now, by regrouping the variables we obtain:
10 THERMOELASTICITY

585
{ { 0
O
O
T
k
p
k mechanical
_
`
_
`
`
b
B
A D
and
{

0
)
`

-
_
,
`
b
B
thermal
T
T
q
V D
(10.267)
where
O
B and B include all forces, and
O
b
`
, b
`
include all variables related to flux. Then,
using the dissipative potential ) , , (
T
k
p q

,
`
`
, , the complementary laws can be rewritten as
follows:
|
|
.
|

\
|
o
o
-
o
o

o
o

T
T
k
k
p
q

,
` `
, , ,
V ; ;

A
(10.268)
or by means of the flux variables as:
) (
; ;
* * *
T T
k
k
p
V o
o
-
o
o

o
o

, , , q

,
`
`
A
(10.269)
where ) , , (
*
T
k
V A , is the dual of ) , , (
T
k
p q

,
`
`
, and is given by:
)
`

|
|
.
|

\
|
-
|
|
.
|

\
|
- -
T T
T T
k
p
k k
p
T
k
k
p
q

q

q

,
`
,
`
` `
,
` `
, , ) , , (
, ,
*

V V A A : Sup ,
(10.270)
We will now introduce the potential d such that:
O
o
d o

O
O
` `
`
;
B
b
(10.271)
and, according to the maximum plastic dissipation principle, the following is satisfied:

o
o

F
`
`
p
Plastic flow rule (10.272)
where ` is the plastic multiplier, F is the yield surface (which is a convex function) and
the plastic flow direction is normal to the surface F , (see Chapter 9).














11 Damage Mechanics











11.1 Introduction
The term Continuum Damage Mechanics has been used to models materials which are
characterized by loss of stiffness, i.e. by a decrease in their stiffness modulus. Damage
models have also been used to simulate different materials (fragile and ductile), which are
fundamentally characterized by irreversible material degradation. Physically speaking, we
can describe the degradation of mechanical material properties as processes in which the
initiation and growth (propagation) of micro-defects such as micro pores and microcracks
take place.
In the pioneering work of Kachanov (1958) the concept of effective stress was introduced,
and by using continuum damage he solved problems related to creep in metals. Rabotnov
(1963) gave the problem physical meaning by suggesting we measure how the sectional area
has reduced by means of the damage parameter. Nowadays, Continuum Damage
Mechanics has become an important tool and is a consistent theory based on irreversible
thermodynamic processes (the Clausius-Duhem inequality). Thermodynamic formalism
was developed by Lemaitre&Chaboche (1985) and among important contributors to our
knowledge about damage mechanics we can include: Mazars (1986), Mazars&Pijaudier-
Cabot (1985), Chaboche (1979), Simo&Ju (1987 a,b), Ju(1989), Oliver et al. (1990) and Oller
et al. (1990).
The continuum damage models, from a computational point of view, are very attractive
since these present simple algorithms and are satisfactory for solving large problems.
In this chapter we will present some basic damage models used to study the failure
mechanism after which we can develop more complex ones.
11
Damage Mechanics
587 , Notes on Continuum Mechanics, Lecture Notes on Numerical
Methods in Engineering and Sciences 4, DOI 10.1007/978-94-007-5986-2_13,
International Center for Numerical Methods in Engineering (CIMNE), 2013
E.W.V. Chaves
NOTES ON CONTINUUM MECHANICS

588
11.2 The Isotropic Damage Model in a Small
Deformation Regime
Continuum damage models have been widely accepted for simulating the behavior of
materials whose mechanical properties are degrading due to the presence of small cracks
that propagate during loading. To fully describe this phenomenon, we will first use a one-
dimensional model (1D) which we will then extrapolate to three dimensional ones (3D).
With regard to continuum kinematics, our study in this section will be carry out in a small
deformation regime, and will be based on the lecture notes of Prof. Javier Oliver,
Universitat Politcnica de Catalunya.
11.2.1 Description of the Isotropic Damage Model in Uniaxial
Cases
Let us now suppose that a material point is subjected to the stress state as shown in Figure
11.1, whose apparent stress ( o) acts on the section s and due to the presence of faults
(microcracks), only the undamaged region will be considered, i.e. the effective section ( s )
on which the effective stress ( o ) acts.











Figure 11.1: Continuum with microcracks.
Then, if we consider the force balance in Figure 11.1, we obtain:
o o s s (11.1)
The equation (11.1) can also be rewritten without altering its outcome as follows:
o |
.
|

\
|
- o |
.
|

\
| -
- |
.
|

\
|
-
-
o o o 1 1 1
s
s
s
s s
s
s s
s
s
d

(11.2)
where
d
s is the damaged section.
Note, the expression
s
s
d
represents the amount of the original section which is corrupted,
which in extreme cases, assumes the follows values:
s
P
B
o - effective stress
o- apparent stress
s
o
o
o
micro
crack
material point
11 DAMAGE MECHANICS

589
o o 0 0
s
s
s
d
d
- The section is not damaged;
0 1 o
s
s
s s
d
d
- The section is completely damaged.
The amount
d
s depends on the stress state o or indirectly on r . The dimensionless ratio
s
s
d
represents the damage variable and is denoted by
s
s
d
d
. Then, the equation in (11.2)
can be written as:
( ) 1 0 ; 1 s s o - o d d
(11.3)
where o is the effective stress.
11.2.1.1 The Constitutive Equation
The effective stress o and strain, in the undamaged area element, are interrelated by
Hookes law as:
r o E (11.4)
where E is Youngs modulus. Then, by substituting (11.3) into (11.4) we can obtain the
constitutive equation for stress in the one-dimensional isotropic damage model:
( ) r - o 1 E d 1 0 s s d (11.5)
We can now verify that as the damage variable evolves, the state no longer returns to its
original value. Physically speaking, we can interpret this as once the material has suffered
damage this will be permanent. Hence, we can conclude that 0 d
`
, which characterizes an
irreversible process. Now, the equation in (11.5) can still be written as:
( ) E d E E
sec_d sec_d
1- r o with (11.6)
where
sec_d
E is the damage secant stiffness modulus with which we can observe that the damage
variable can be interpreted as a measure of the loss of stiffness modulus of the material.
In general, materials have a yield stress that separates the elastic (reversible process) from
the inelastic zone (irreversible process). In the strain space, we can represent the elastic
limit by the variable
0
r , (see Figure 11.2), in which the damage process has not yet begun,
i.e.:
0 d if
0
r < r (11.7)
In the elastic region, the following is satisfied:
- 0 0 d s
d
`
` ;
-
0
r s r where
0
r is the threshold that defines the elastic region.
In a representative stress-strain curve, (see Figure 11.2), during an unloading process
( ) 0 d
`
, the secant modulus is given by E d E
sec_d
) 1 ( - and after unloading is complete,
there is no residual (permanent) strain, (see Figure 11.2), although the material has suffered
some internal damaged.
We can now summarize the basic features of the one-dimensional damage model as
follows:
NOTES ON CONTINUUM MECHANICS

590
( ) ) 1 0 ( ; 1 s s r - o d E d
0 d if
0
r < r
(11.8)
Now, by starting from the above equation we can obtain the energy equation in the system
as follows:
( ) ( )
e
e
d E d 1 1 1
1
- r r - ro

1
2
1
1
2
1
_

(11.9)
where
e
1 is the elastic strain energy.







Figure 11.2: Stress-strain curve.
11.2.2 The Three-Dimensional Isotropic Damage Model
The basis of this damage model is to define a transformation between physical (real) and
fictitious spaces (effective space) in which the material is undamaged, (see Figure 11.3).








Figure 11.3: Real and fictitious spaces.
NOTE: As described above, this model depends on a single variable: the damage parameter
d , which means that we are assuming a mechanical behavior in which the degradation is
independent of the orientation, and because of this, this model is referred to as the Isotropic
Damage Model. As a result of this, the fourth-order damaged elasticity tensor remains an
isotropic tensor.
11.2.2.1 Helmholtz Free Energy
Let us now consider the Helmholtz free energy function ) , , (
k
T F 1 1 , or simply free
energy, which is a function of the deformation gradient ( F ), temperature (T ) and the set
1
E d E
sec_d
) 1 ( -
1
o
E
r
Loading(damage)
Unloading / elastic loading
Elastic limit
Y
o

0
r
Physical space (real) Effective space (fictitious)


o - o
o
) 1 (

d
stress effective

r
1
r
1
o
o
o
o
11 DAMAGE MECHANICS

591
of internal variables (
k
). Let us also consider there is a process independent of
temperature, and the internal variable associated with the problem is characterized by the
damage variable d . Furthermore, as seen in previous chapters, as the Helmholtz free
energy must satisfy the principle of objectivity (see Chapter 6), we can express 1 in terms
of the Green-Lagrange strain tensor ( E ), which in turn collapses with the infinitesimal
strain tensor in a small deformation regime, i.e. = E . Then, if we consider all of the
above, the Helmholtz free energy can be expressed in terms of:
( ) , d 1 1 (11.10)
or explicitly as follows:
( ) ( )
2
1
1 1 : :
e e
d d C - - 1 1
Helmholtz free energy for
isotropic damage model
(11.11)
where ) (
e
1 is the elastic strain energy density, which is a function of strain only, and
e
C
is the elasticity tensor (or elastic stiffness tensor).
11.2.2.2 Internal Energy Dissipation and the Constitutive Equations
The damage model has thermodynamic consistency, and so, entropy inequality is fulfilled.
One way to express this entropy inequality is by means of the alternative form of the
Clausius-Planck inequality, (see Chapter 5), which is expressed by:
[ ] 0 - - 1 j
` `
T
int
D : D
(

3
m
J
(11.12)
Note that the terms D : , T
`
j , 1
`
have the unit of energy per unit volume (density
energy). In a small deformation regime D
`
= holds, and by considering the isothermal
process we have 0 T
`
, so, the equation in (11.12) becomes:
0 - 1
`
`
:
int
D (11.13)
Then, the rate of change of the free energy ( ) , d 1 1 can be evaluated as follows:
( ) d
d
d
`
`
`
o
o
-
o
o

1 1
1

: , (11.14)
Next, by substituting (11.14) into the internal energy dissipation given in (11.13) we obtain:
0 ) , (
o
o
-
(

o
o
-
o
o
-
o
o
- - d
d
d
d
d
int
`
`
`
` ` `
`
1 1 1 1
1

: : : : D (11.15)
Note that the above inequality must hold for any admissible thermodynamic process, so, let
us assume there is one where 0 d
`
. Here, we obtain 0
(

o
o
-

`
:
1
int
D , which in turn
must also be true for any process. Additionally, if we have a process such that
` `
- , the
only way for the entropy inequality to be satisfied is when

o
o

1
holds with which we
obtain the constitutive equation for stress. Thus, the entropy inequality becomes:
0
o
o
-
o
o
-
(

o
o
-

d
d
d
d
int
` `
`
_
1 1 1

0
: D
(11.16)
NOTES ON CONTINUUM MECHANICS

592
Now, if we consider the energy equation ( )
e
d 1 1 1- , we obtain
e
d
1
1
-
o
o
, thus
0 d
e
int
`
1 D (11.17)
where by definition 0
e
1 . Then, to satisfy the inequality (11.17), the rate of change of the
damage parameter must satisfy:
0 d
`

(11.18)
Then, by means of thermodynamic considerations we can draw the conclusion that:
0 ;
o
o
d
`

1
(11.19)
We can also express the rate of change of the Helmholtz free energy by means of the
equation in (11.11), i.e.:
( ) ( )
e e e e e
d d d d d 1 1 1 1 1 1 1
`
`
`
`
` ` `
- - - - - : : : C (11.20)
Next, the rate of change of the elastic strain energy,
2
1
: :
e e
C 1 , was obtained as
follows:
( )
2
1

`
`
`
`
: : : : : :
e e e e
C C C - - 1 (11.21)
where 0
e
C
`
, since
e
C is constant, and as the elasticity tensor features major symmetry
(
e
ijkl
e
klij
C C ), the equation in (11.21) becomes:
( ) ( )
( )

` ` `
`
` ` ` `
`

1
1

2
1
2
1
: : : :
d
e
kl
e
ijkl ij
kl
e
ijkl ij ij
e
klij kl kl
e
ijkl ij kl
e
ijkl ij
e
-
r r
r r - r r r r - r r
C C
C C C C 1

(11.22)
Note that due to the major symmetry of
e
C , : :
e e
C C is fulfilled.
Then, starting from the equation in (11.11) we can obtain the stress by taking the derivative
of the strain energy with respect to strain, i.e.:
( )
( ) ( ) [ ]
( )
( )
( ) [ ] ( ) [ ]
( ) ( ) ( )
( ) ( ) ( ) ( ) {
e
pqij pq
e
ijkl kl
e
pqji
e
pqij pq
e
jikl
e
ijkl kl
kj li lj ki pq pj qi qj pi kl
e
pqkl
ij
lk kl
pq
ij
qp pq
kl
e
pqkl
ij
kl
pq
ij
pq
kl
e
pqkl
kl pq
ij
e
pqkl kl
e
pqkl pq
ij ij
ij
d d
d
d
d
d d
d
C C C C C C
C
C
C
C C
r - r -
)
`

- r - - r -
)
`

- r - - r -

r o
r - r o
r -
r o
r - r o
r -
|
|
.
|

\
|
r o
r o
r -
r o
r o
r -
r r
r o
o
-
(

r r -
r o
o

r o
o
o
2
1
1
2
1
2
1
2
1
1
2
1
2
1
2
1
1
2
1
1
2
1
1

2
1
1
2
1
1
,
2
1
2
1
c c c c c c c c
1

(11.23)
where we have taken into account the minor symmetry of the elasticity tensor, i.e.
e
jikl
e
ijkl
C C ,
e
pqji
e
pqij
C C . Note also that the indexes p , q are dummy indexes, so we can
11 DAMAGE MECHANICS

593
exchange them for k and l without altering the expression. Additionally, by taking into
account the major symmetry of the elasticity tensor,
e
klij
e
ijkl
C C , we obtain:
( )
kl
e
ijkl ij
d r - o C 1 (11.24)
which in tensorial notation becomes:
( )
( ) ( )

1 1
,
d d
d
e
- -
o
o
: C
1 The constitutive equations for isotropic
damage model
(11.25)
where is the effective Cauchy stress tensor and is defined as:
:
e
C The effective Cauchy stress tensor (11.26)
and
e
C is the elasticity tensor (fourth-order definite positive tensor) which contains the
elastic mechanical properties. Remember that
e
C can be represented in terms of the Lam
constants ( A, j) as follows:
( )
jk il jl ik kl ij
e
ijkl
e
c c c c j c Ac j A ; 2 - - - C I 1 1 C (11.27)
where 1 is the second-order unit tensor, and
sym
I = I is the symmetric fourth-order unit
tensor, whose components are expressed in terms of the Kronecker delta (
ij
c ) as follows:


j i if
j i if
ij ij
0
1
) ( c 1 ; ( )
jk il jl ik ijkl ijkl
sym
ijkl
c c c c
2
1
) ( - = I I (11.28)
Then, by analyzing the constitutive equation in (11.25) we can put in evidence the
following sentences:
- Since the damage parameter is a scalar, the stiffness degradation is isotropic;
- We can calculate the stress immediately once we know the current values of
(strain) and d (internal variable);
- We can interpret the equation in (11.25) as the sum of elastic and inelastic parts, i.e.:
( )
i e
inelastic
e
elastic
e e
d d - - -
_ _
1 : : : C C C
(11.29)
The Elastic-Damage Secant Stiffness Tensor
We can then define the elastic-damage secant stiffness tensor for the isotropic damage
model as:
( )
e sec_d
d C C 1- The elastic-damage secant stiffness tensor (11.30)
Let us now consider a uniaxial case, (see Figure 11.4), where the material is loaded until the
stress state reaches the point P represented in Figure 11.4, after which unloading occurs,
with the unloading path being that indicated by the slope E d E
sec_d
) 1 ( - defined in
Figure 11.4.
11.2.2.3 Ingredients of the Damage Model
The damage constitutive model is completely determined when the damage variable
t
d is
known at each time step t of the loading/unloading process. Then, we can define the
following elements of the constitutive equation:
NOTES ON CONTINUUM MECHANICS

594
The energy norm of the stress (or strain) tensor;
The damage surface and damage criterion. The damage surface defines the elastic
limit, and the damage criterion establishes when the material is in a loading or in a
elastic process, and;
A set of evolution laws for internal variables.








Figure 11.4: Stress-strain curve.
The Energy Norm in the Stress/Strain Space
The norm is a measure of distance and so is a scalar. Next, we will define a simple norm in
the stress space denoted by

t (equivalent stress), and in the strain space denoted by

t .
The latter is also known as the equivalent strain:



t
t t
t ) 1 (
2 ;
1
1
d
e e e
e e
-


-
-
_
1 : : : : C C
C C
(11.31)
Note that

t and

t are surface equations (ellipsoids) that characterize the stress state at


the current point (see Figure 11.5). The proof of (11.31) now follows:
( ) ( ) ( )
( )



t t
t
t
d
d d d
e
e
-


- - -
-
1
1 1 1
2
1
: : :
: : : : :
C
C

(11.32)
In order to better describe material behavior, others norms will be introduced (see
subsection 11.2.4).
The Damage Criterion
Next we will define the damage criterion in the stress and strain space:
_
space

stress
r q q 0 ) ( ) , ( s -

t t F and
_
space strain
r r 0 ) , ( s -

t t G
(11.33)
where r is an internal variable (current damage threshold), and q is a stress-like
hardening/softening variable which is a function of r . Note that each material in its
undamaged state is characterized by the initial value of r which is denoted by
0
r (the
material parameter), which defines the initial yield in the strain space. Then, the material
o
E
r
E d E
sec_d
) 1 ( -

Y
o
Dissipated energy
P
1
1
11 DAMAGE MECHANICS

595
starts to fail (initial damage) when the energy norm exceeds the value
0
r . Later we will
relate the variables r and q to the damage variable.










Figure 11.5: Strain and stress state in the principal space.
The damage criterion requires that the current stress state must be on or inside the damage
surface. When the stress state lies inside said damage surface, the material shows elastic
behavior, which can be elastic loading or unloading.
Then we can define the admissible strain space as follows:
{ 0 ) , ( : s r

t G E (11.34)
and the admissible stress space as:
{ 0 ) , ( : s q

t F E (11.35)
When it holds that 0 ) , ( q

t F , in the stress space, the stress state is on the surface as


indicated in Figure 11.5(b).
The stress space (

E ), (see Eq. (11.35)), can be decomposed into the inner domain
( )

E int (when the stress state is inside the surface), and other by the surface itself,

E o .
We can define then the elastic region in strain and stress respectively as:
{ { 0 ) , ( : ; 0 ) , ( : s < q r

t t F G E E int int (11.36)
and the elastic limit (damage surface):
{ { 0 ) , ( : ; 0 ) , ( : o o q r

t t F G E E (11.37)
where it holds that:
( ) ( ) C o o .

E E E E E int int ; (11.38)
Note that ( )

E int is the same as 0 ) , ( < q

t F which describes the elastic region, and

E o
is the damage surface. Note that when the stress state is at a point inside of the space

E it
will also be inside the space

E , and when the stress state is on the surface

E o it will also
be on the surface

E o . Hence, we can use either the stress or strain space to describe how
the damage evolves, proof of which follows:
0 0 ) 1 ( ) 1 ( 0 ) ( s - s - - - s - r r d d r q

t t t (11.39)

1
r

2
r

3
r

t
a) Strain principal space
0 ) , ( r

t G


t t ) 1 ( d -

1
o

2
o

3
o

t
b) Stress principal space
0 ) , ( q

t F
NOTES ON CONTINUUM MECHANICS

596
Said damage evolves when the norm

t exceeds the maximum value reached by r . Then,


considering (11.33) and (11.31) we can also conclude that:
r d r q ) 1 ( ) ( -
(11.40)
In uniaxial cases, damage starts when

t exceeds the first damage threshold value


0
r .
Then, from the equation in (11.31) and by means of Figure 11.2, we can obtain:
E
r r
E
E
E
E E
Y
Y Y e
o
-
o

o
r r r
0 0
0 0 0
uniaxial
0



t
t t : : C
(11.41)
where
Y
o is the yield stress (obtained in the laboratory). Then, ) , (
0
E r
Y
o can be
interpreted as a material mechanical property also obtained in the laboratory.


























Figure 11.6: The evolution of

t and r over time t .


2(
0
) 2 (
r r )
3(

t
) 3 (
r )
4(

t
) 4 (
r )
r
o

Y
o

0
r

1
r
1 0 d
1 2 3
4
5(
) 4 ( ) 5 (
r r )

3
r

1
r
1 2 3 4, 5

0
r
5

5 4
r r


` 0 r

t
r

0
r

0
r
t

t
1
2
3
4 5
1
2
3
4
5
6
6


` 0

> r

11 DAMAGE MECHANICS

597
The Internal Variable Evolution Law. The Kuhn-Tucker and Consistency Conditions
The constitutive equation described above uses three types of variables, namely: the free
variable { ; the internal variable { r ; the dependent variables { ) ( ), , ( ), , ( r d d r 1 .
Now, to establish how the internal variable r evolves, let us take the example described in
Figure 11.6. As we can observe, the discretized r` between points 2-3 and 3-4 are positive
and between points 1-2 and 4-5 are equal to zero, so we can conclude that r is a
monotonically increasing function, i.e.:
0 r` (11.42)
Graphically, we can see in Figure 11.6 how the variables r and

t evolve. Furthermore,
we can also verify that in the range between the points 4-6 0 ) , ( < - r r

t t G holds, i.e.
there is an elastic regime.
Thus, we can establish that at time t ,
t
r is given by the following equation:
( ]

=
s
t s
t
r r

t
,
0
, max max
(11.43)
As we saw, the Helmholtz free energy is a function of ( ) ) ( , r d 1 1 , where now the
damage variable is in terms of the variable r (the internal variable), thus:
( ) [ ] ( )
r
r
r d
t
r
r
r d
d
r d r d r d
e
e
`
`
`
`
o
o
-
o
o

o
o
o
o
o
o
-
o
o
-
) (
) (
) ( , ) ( ) ( 1 ) ( ,
1
1
1 1
1 1 1


:
:

(11.44)
where we have considered that
e
d
1
1
-
o
o
. Then, by using the above equation, the internal
energy dissipation becomes:
0
) ( ) (
0
) (
) , (

o
o

o
o
-
(

o
o
-

o
o
-
o
o
- -

r
r
r d
r
r
r d
r
r
r d
r
e e
e
int
` `
`
_
`
` ` `
`
1 1
1
1
1
1


0
:
: : : D

(11.45)
If we compare the above inequality with the one obtained in (11.17) we can conclude that:
r d r
r
r d
d ` `
`
) , (
) (

t H
o
o

(11.46)
where H is the continuum hardening/softening modulus. The evolution laws for r and
for d (damage variable) are then given by:
) , ( ; ) , ( d d r r

t H
` ` `
`
(11.47)
where 0
`
is the damage parameter consistency (damage multiplier).
Let us now consider a time t in which the process is characterized by
t
and r . Then, we
can observe the following: if at any time both 0 ) , (
t
r

t G and 0 ) , ( > r

t G
`
hold, this
implies that 0 ) , ( >
A - t t
r

t G , which thereby violates the condition { t r


t
0 ) , ( V s

t G , so
0 0 > G
` `
must be satisfied. Another possible situation is when the current state is
NOTES ON CONTINUUM MECHANICS

598
inside the damage surface, i.e. 0 ) , ( <
t
r

t G , and if in the next loading step 0 ) , ( <


A - t t
r

t G
is satisfied, this implies that 0 ) , ( 0 < r r
`
`
`
G . We can gather these previous
conditions by means of the loading/unloading condition, also called the Kuhn-Tucker conditions:
0 ) , ( ; 0 ) , ( ; 0 s r r

t t G G
` `
The Kuhn-Tucker conditions (11.48)
and by the consistency (persistency) condition:
0 ) , ( r

t G
` `
The consistency condition (11.49)
If we are undergoing loading, this implies that 0 >
`
, then by means of the Kuhn-Tucker
conditions 0 ) , ( r

t G must be fulfilled. Here, the value of


`
can be obtained by means
of the consistency condition:
r r r `
`
`


t t t 0 ) , ( ) , ( G G (11.50)
Schematically, we can summarize the above loading/unloading states as follows:






(11.51)

NOTE: If the parameter ) , ( d

t H , given in (11.46), is not a function of d , we can


express it by means of

t
t
t
o
o

) (
) (
G
H , where we have introduced the scalar function G
which is a monotonically increasing function, which has proven to be a convenient way to
express the damage criteria:
( ) ( )
( ) ( ) 0 ; 0 ) , (
0 ; 0 ) , (
V s -
V s -
t q F F q
t r G G r


t t
t t
F
G

(11.52)
Here the loading/unloading condition becomes:
r
r
d r r
o
o

) , (
; ) , (

t G

` ` `
` (11.53)
0 ) , ( ; 0 ) , ( ; 0 s r r

t t G G
` `
The Kuhn-Tucker conditions (11.54)
0 ) , ( r

t G
`
`

The consistency condition (11.55)


The Damage Variable
The parameter q is the stress-like hardening/softening parameter, and is defined in terms
of r as follows:
r
r q
r d r d r q
) (
1 ) ( ) 1 ( ) ( - - (11.56)

<

<
0
0
0
0
G
G
G
G
`
`

>

0
0

`
`




0
`

0
`


0 d
`

0 d
`

0 d
`

0 > d
`

(elastic)
(unloading)
(neutral loading)
(loading)
11 DAMAGE MECHANICS

599
Now, by using the equations in (11.56) and (11.25) we can obtain:

r
r q ) (
(11.57)
in which the following holds:
[ ] = s s , 1 0
0
r r d (11.58)
Note that with the new definition of the damage parameter given in (11.56), we can
restructure the equation in (11.46) as follows:
r
r q
r d
) (
1 ) ( - (11.59)

(
(
(
(

o
o
-

-
o
o

o
o
r
r
r
r q
r q
r
r
r q
r
r
r
r d
d ` ` `
`
2
) (
) (
) (
1
) (
r
r
r r q
d
d
`
`
(

2
) ( ) ( H

(11.60)
where we have defined a new parameter
) (
) (
r
r q
r
d
o
o
H , which is the hardening/softening
parameter.
11.2.2.4 The Hardening/Softening Law
The expression
r
r q
o
o ) (
defines the hardening/softening parameter, thus:
[ ) [ ] ; , ; ) 1 ( ), 0 ( ; ) (
0 0 0 0
E
r q a r q d d r r r r q
Y d
o
= = ` ` H
(11.61)
where
d
H is the continuum hardening/softening parameter and which is characterized by:
0 ) ( Softening with Damage
0 ) ( Damage Perfect
0 ) ( Hardening with Damage
<

>
r
r
r
d
d
d
H
H
H
(11.62)
Here, we will consider the relationship between q and r to be linear or exponential.
The Linear Hardening/Softening Law
Now, by assuming that q varies linearly with r , we have:

> <
/
- -
s

0 0 0
0
0 ) (
0
) (
r r r r r
r r
r q
d
H

(11.63)
Then, taking into account the equation in (11.56) we can still state that:
( )

> >
/
- - -
s
-
0
0
1 1 1
0
1
0 0
r r
r r
r
q
d
r
r d
r
r
H

(11.64)


NOTES ON CONTINUUM MECHANICS

600







Figure 11.7: The linear hardening/softening law.
The Exponential Hardening/Softening Law
The exponential law is described by Figure 11.8. Then we can express ) (r q as follows:
( ) 0 ) (
0
1
0
> - -
|
|
.
|

\
|
-

A with r q q r q
r
r
A
exp
(11.65)
in addition to this we have:
( ) |
|
.
|

\
|
-

o
o
0
1
0
0
) (
r
r
A
r
r q
A
r
r q
exp
(11.66)








Figure 11.8: The exponential hardening/softening law.
Table 11.1: Summary of the Isotropic Damage Model in a small deformation regime
described in the strain space.
ISOTROPIC DAMAGE MODEL IN A SMALL DEFORMATION REGIME
Helmholtz free energy ( ) [ ] ( ) : :
e e e
r d r C
2
1
with ) ( 1 , - 1 1 1 (11.67)
Damage parameter [ ] [ ] 1 , 0 ; , , ; 1 ) (
0
= = - d a a r q
r
q
r d (11.68)
The constitutive equations ( ) ( )

:
e
d d C 1 1 - -
o
o

1
(11.69)
Evolution law
`
` r
[ )

o

=

E
r r
r r
Y
t 0
0
0
,

(11.70)
Damage criterion
( ) r r r
e
- -

: : C t , G
(11.71)
) (r q

0
r

0
r q <

r r
0
r

0
r q >

q

0
r
) 1 ( d
r
0 >
d
H
) (r q

0
r
0 <
d
H
r
11 DAMAGE MECHANICS

601
Hardening Law ( ) 0 ) ( ; ) ( s r q r r q
d d
H H ` ` (11.72)
Loading/unloading condition
0 ; 0 ; 0 < G G
` `
(11.73)
Consistency condition
0 G
` `
(11.74)
11.2.3 The Elastic-Damage Tangent Stiffness Tensor
Next, we will obtain the elastic-damage tangent stiffness tensor, which gives us an
advantage, from a computational point of view, when we are dealing with the incremental-
iterative solution procedures and as a result of this, convergence is improved considerably.
The relationship between
`
and
`
give us this tensor denoted by
tan_d
C , i.e.
` `
:
tan_d
C .
Now, by considering the equation in (11.25), ( ) 1 :
e
d C - , we can obtain the rate of
change of the stress as follows:
( ) ( ) ( )
( ) d d
d d d d d
d
d
e
e e e
`
`
`
`
`
`
`
` `
- -
- - - -
o
o
-
o
o


1
1 1 ,
:
: : : :
C
C C C

(11.75)
in which there is the following:
a) A process with elastic loading or unloading
0 0 d
`
, thus the equation in (11.75) becomes ( ) ( )
` `
:
e
d d C 1 , - , with which the
elastic-damage tangent stiffness tensor coincides with the elastic-damage secant stiffness
tensor when we are dealing with elastic loading:
( ) ) 1 ( 1 d where d
e e tan_d sec_d
- - C C C C (11.76)
b) A process with damage loading
r r ` `


t t , and the rate of change of the damage parameter ) (r d d becomes:

t
t
` `
`

o
o

o
o

o
o
o
o

d
r
r
d
t
r
r
d
d
(11.77)
where the rate of change of

t can be evaluated as follows:


( ) ( )





` ` `
` ` `
: : : : :
: :
: : : : : : : :
t t
t t
1 1 1
2
1

2
1

-
-
e e
e
e e e e
C C
C
C C C C

(11.78)
Now, by substituting (11.78) into the equation in (11.77) we obtain:


`
`
:
t t
1
o
o

d
d
(11.79)
Then, taking into account the equations (11.79) and (11.75), we can find the relationship
between the rates of stress and strain change:
( ) ( )
( ) ( )



`
` `
`
` `
:
: : :
(

o
o
- -

o
o
- - - -
t t
t t
1
1
1
1 1
d
d
d
d d d
e
e e
C
C C
(11.80)
NOTES ON CONTINUUM MECHANICS

602
which thus defines the elastic-damage tangent stiffness tensor:
( ) ( )
(

o
o
- -

t t
1
1
d
d
e tan_d
C C (11.81)
and by considering that in a loading process r
r
t holds we then find:
3 2
) ( 1 ) ( 1 1
r
r r q
r r
r r q
r r
d d
d d
H H -

|
|
.
|

\
|
-

o
o

o
o

t t
(11.82)
where we have taken into account the equation in (11.61),
2
) (
r
r r q
r
d
d
H -

o
o
.
Then, by substituting the equation in (11.82) into that in (11.81) we can obtain
tan_d
C in
terms of q and r :
|
|
.
|

\
|

|
|
.
|

\
|
-
- -
_ _


e e
d
e tan_d
r
r r q
d C C C C : :
) (
) 1 (
3
H
(11.83)
Now, the general equation for the elastic-damage stiffness tensor
tan_d
C (symmetric
fourth-order tensor) is given by:
( )
_
`
0 0


K
K
r
loading unloading
0) (d with elastic
e e e
e
tan_d
C C C
C
C
: :
The elastic-damage stiffness
tensor for isotropic damage
model
(11.84)
where,
3
) (
r
r r q
d
H
K
-
and
( )
d - 1 .
11.2.4 The Energy Norms
Next, we will define some energy norms, which together with the damage criteria, play an
important role in defining the yield damage surface.
In order to adequately represent the materials different norms will need to be defined so as
to describe how these materials really behave. For example, in a simple model for concrete,
if we only want to simulate the process of failure caused by tension, the tension-only damage
model is used which means that it cannot capture the other type of failure caused by
compression. Next, we will define some models used in the isotropic damage process.
11.2.4.1 The Symmetrical Damage Model (Tension-Compression)
Model I
This type of model shows when the material behave the same both with tension or and
compression. The energy norm of this model is then represented by:

: : : ) 1 (
1
d
e I
-
-
C t
(11.85)
We can also define the energy norm of the strain tensor (also known as the equivalent
strain), proposed by Simo&Ju(1987), (see equation (11.31)):
e e I
1 2

: : C t
(11.86)
11 DAMAGE MECHANICS

603
To better illustrate this model, let us consider the state of plane stress ( 0
3
o
i
). In this
case, the yield surface is represented by an ellipse, (see Figure 11.9), where 0 > o
Y
is the
stress limit for tension and compression and the damage surface evolves symmetrically.










Figure 11.9: Damage surface in 2D and the uniaxial stress-strain curve for model I.
11.2.4.2 The Tension-Only Damage Model Model II
The tension-only damage model does not take into account failure by compression, i.e. the
material can only fail by tension and here we can define the following stress field:
2


-
) (
-

(11.87)
where
2
- - -
(-)
def
is the Macaulay bracket whose graphical representation can be
appreciated in Figure 11.10.





Figure 11.10: Ramp function.
Now, by means of spectral representation, we can represent the stress tensor in terms of
eigenvalues (principal stresses) and eigenvectors as follows:
) ( ) (
3
1

a a
a
a
n n o

(11.88)
thus:
) ( ) (
3
1

a a
a
a
n n o

-
(11.89)
Note, the relationship between the real and effective stress remains valid, i.e.:
- -
- ) 1 ( d (11.90)
a) Norm in the principal stress space-2D.
r
o

0
r

t

Y
o

Y
o -

Y
o

Y
o

Y
o -

1
o

2
o
E
b) Stress-strain curve
1
Elastic
region
) (x
x

<
) (
0
0 0
x if x
x if
x
NOTES ON CONTINUUM MECHANICS

604
Then, the norm for the isotropic damage model defined previously becomes:

: : :
e e
C 1 2 t
(11.91)
Next, in the tension-only damage model
-
, it follows that:

: : : : : : :
1 1
2
1
) 1 (
1
) 1 (
1 -
-
-
-
-
- -
-

-

e e e II
d d
C C C t
(11.92)
Then, if we consider the equation in (11.31), we can conclude that:

: :
1 -
-

e II
C t
(11.93)
Finally, in Figure 11.11 we can visualize the damage surface for two-dimensional cases
(2D).
11.2.4.3 The Non-Symmetrical Damage Model Model III
The non-symmetrical damage model is useful to simulate materials, such as concrete,
whose tension domain differs with respect to compression. This model uses the following
norm:

: :
1

1 -
|
.
|

\
| -
-
e III
n
C
0
0 t (11.94)
where the parameter 0 is the weight factor dependant on the stress state which is given
by:

o
) o (

3
1
3
1
i
i
i
i
0
(11.95)
The parameter n is defined by means of the ratio of the compression elastic limit
c
Y
o to
the tension elastic limit
t
Y
o , i.e.:
t
Y
c
Y
n
o
o

(11.96)
In the case of concrete n is approximately equal to 10 = n .









Figure 11.11: Damage surface in 2D and the uniaxial stress-strain curve for model II.
r
o
Elastic region

Y
o

1
o

2
o
E
a) Norm in the principal stress space-2D. b) Stress-strain curve.
1

0
r

t Y
o

Y
o
11 DAMAGE MECHANICS

605












Figure 11.12: Damage surface in 2D and the uniaxial stress-strain curve for model III.
11.3 The Generalized Isotropic Damage Model
Note that the elasticity tensor
e
C can be written in terms of the following sets of
mechanical parameters ) , ( j A , ) , ( v E , ) , ( G x :
_
_
part isochoric
part
volumetric

3
1
2
) 1 ( ) 2 1 )( 1 (
2
(

- - x
v -
v
-
v - v -
v
- 1 1 I 1 1 I 1 1 I 1 1 j j A
E E
e
C
(11.97)
where ) (E =Youngs modulus, ) (v =Poissons ratio, ) , ( j A =Lam constants, ) (x =bulk
modulus, and j G is the shear modulus.
In the isotropic damage model the elastic-damage secant stiffness tensor can be
represented as follows:
I 1 1 I 1 1
) 1 ( ) 2 1 )( 1 ( ) 1 (
) 1 (

) 2 1 )( 1 (
) 1 (
) 1 (
v -
v
-
v - v -
v

v -
- v
-
v - v -
- v
-
sec_d sec_d
e sec_d
E E E d E d
d C C
Note that, in this model the damage variable affects only one of the mechanical parameters,
namely, the Youngs modulus. We can also verify that the same damage parameter equally
affects both the spherical and deviatoric part:
(

- - - x - -
3
1
2 ) 1 ( ) 1 ( ) 1 ( 1 1 I 1 1 j d d d
e sec_d
C C (11.98)
Another model described by Carol et al. (1998) generalizes the isotropic damage model by
considering independent degradation of the spherical and deviatoric parts and because of
this the model requires two independent damage variables.
Now, the elasticity tensor components can be expressed by means of their spherical and
deviatoric parts as follows:
( )
(

- - - x
kl ij jk il jl ik kl ij
e
ijkl
c c c c c c j c c
3
1
2
1
2 C (11.99)
r
o
Elastic
region

t
Y
o
t
Y
o

t
Y
o

1
o

2
o
E

c
Y
o -

c
Y
o -

t
Y
c
Y
n o - o -
a) Norm in the principal stress space-2D. b) Stress-strain curve.

0
r

t
NOTES ON CONTINUUM MECHANICS

606
Then, with
kl ij
V
ijkl
c c
3
1
P and ( )
V
ijkl jk il jl ik
D
ijkl
P P - - c c c c
2
1
, the above equation becomes:
D
ijkl
V
ijkl
e
ijkl
P P C 2 3 j - x
D V e
P P C 2 3 j - x (11.100)
Let us now consider that the material parameters x and j can be degraded by means of
the variables
V
d and
D
d , respectively, and according to the following equations:
0 0
) 1 ( ; ) 1 ( j j
D V
d d - x - x (11.101)
with which the elastic-damage secant stiffness tensor becomes:
D e
ijkl
D V e
ijkl
V D
ijkl
D V
ijkl
V sec_d
ijkl
d d d d
_ _
0 0
) 1 ( ) 1 ( ) 1 ( 2 ) 1 ( 3 C C P P C - - - - - x - j (11.102)
where we have introduced:
( )

- -

x
x
kl ij jk il jl ik
D
ijkl
D e
ijkl
kl ij
V
ijkl
V e
ijkl
c c c c c c j
j
c c
3
1
2
1
2
2
;
3
0
0
_
0
0
_
P C
P C

(11.103)
11.3.1 The Strain Energy Function
Now, if we consider (11.100), the elastic strain energy function can be rewritten as follows:
( ) ( ) ( )
dev e vol e
D V D V e e
_ _
2
2
1
3
2
1
2 3
2
1
2
1
1 1
j j 1
-
- x - x : : : : : : : : P P P P C

(11.104)
where we have introduced:
( )
( )
dev e vol e e
D e D vol e
V e V vol e
_ _
_ _
_ _
) (
2
1
2
2
1
2
1
3
2
1
1 1 1
j 1
1
-



: : : :
: : : :
C P
C P
(11.105)
after which it becomes:
[ ]
) , ( ) , ( ) 1 ( ) 1 (
2
1
) 1 (
2
1
) 1 (
) 1 ( ) 1 (
2
1
2
1
) , , (
_ _
_ _
_ _
D dev V vol
dev
dev e D
vol
vol e V
D e D V e V
D e D V e V sec_d D V
d d d d
d d
d d d d



1 1 1 1
1
1 1
- - - -
- - -
- - -

_ _
: : : :
: : : :
C C
C C C

(11.106)
Additionally, the following holds:
) , ( ) , ( ) 1 ( ) 1 (
2
1
) 1 (
2
1
) 1 (
2
1
) 1 (
2
1
) 1 ( ) , , (
_ _
_ _
_ _
D dev dev V sph vol
dev
dev e D
vol
vol e V
dev D e dev D sph V e sph V
D e D V e V D V
d d d d
d d
d d d d



1 1 1 1
1
1 1
- - - -
- - -
- - -

_ _
: : : :
: : : :
C C
C C

(11.107)
11 DAMAGE MECHANICS

607
11.3.2 Spherical and Deviatoric Effective Stress
Note that the following equations hold:
dev D sph V D e D V e V sec_d
d d d d ) 1 ( ) 1 ( ) 1 ( ) 1 (
_ _
- - - - - - : : : C C C (11.108)
where
sph
,
dev
are the spherical and deviatoric effective stresses, respectively and where
the following is valid:
dev sph
dev D dev
sph V sph
d
d



-

-
-
) 1 (
) 1 (
(11.109)
It is noteworthy that the following equations hold:
( ) ( )
dev D e D sph V e V
dev sph D e D dev sph V e V
D e D V e V
d d
d d
d d



: :
: :
: :
_ _
_ _
_ _
) 1 ( ) 1 (
) 1 ( ) 1 (
) 1 ( ) 1 (
C C
C C
C C
- - -
- - - - -
- - -
(11.110)
Then, the relationship between stress and strain in rate is given by:
( )

-
- -

dev d tan dev
sph d tan sph
dev d tan sph d tan
dev sph d tan dev sph
d tan





` `
` `
` `
` ` ` `
` `
:
:
: :
:
:
_
_
_ _
_
_
C
C
C C
C
C

(11.111)
where
d tan _
C is the elastic-damage tangent stiffness tensor.
11.3.3 Thermodynamic Considerations
In a small deformation regime D
`
= holds and in isothermal processes 0 T
`
is satisfied,
so, it then follows that the expression for internal energy dissipation given in (11.13)
becomes:
0 - 1
`
`
:
int
D (11.112)
Then, by evaluating the rate of change of the strain energy function given in (11.106),
D e D V e V
d d
_ _
) 1 ( ) 1 ( 1 1 1 - - - , we can obtain:
D D e V V e D D e V V e
D D e D D e V V e V V e
d d d d
d d d d
` `
` `
` `
` ` `
_ _ _ _
_ _ _ _
) 1 ( ) 1 (
) 1 ( ) 1 (
1 1 1 1
1 1 1 1 1
- - - - -
- - - - -

(11.113)
and by using the stress equation given in (11.108) we have:
[ ]


` `
` `
: : : :
: : : :
D e D V e V
D e D V e V
d d
d d
_ _
_ _
) 1 ( ) 1 (
) 1 ( ) 1 (
C C
C C
- - -
- - -

(11.114)
Note that
`
`
: :
V e V e _ _
C 1 and
`
`
: :
D e D e _ _
C 1 , thus:
D e D V e V
D e D V e V
d d
d d
_ _
_ _
) 1 ( ) 1 (
) 1 ( ) 1 (
1 1
` `
` ` `
- - -
- - - : : : : : C C

(11.115)
Then, together the equations (11.115), (11.113), and the internal energy dissipation given in
(11.112), yields:
NOTES ON CONTINUUM MECHANICS

608
0
0 ) 1 ( ) 1 ( ) 1 ( ) 1 (
0
_ _
_ _ _ _ _ _
-
- - - - - - - - -
-
D D e V V e
D D e V V e D D e V V e D e D V e V
int
d d
d d d d d d
` `
` `
`
` ` ` `
`
1 1
1 1 1 1 1 1
1 : D

(11.116)
Since (11.116) must be satisfied for any admissible thermodynamic process, it follows that:
0 ; 0
D V
d d
` `
(11.117)
where we have taken into account that 0
_

V e
1 and 0
_

D e
1 .
11.3.4 The Elastic-Damage Tangent Stiffness Tensor
Initially we adopt the following norms:
sph V e sph sph sph sph V e V

: : : :
_ _
2 C 1 t
(11.118)
dev D e dev dev dev dev D e D

: : : :
_ _
2 C 1 t
(11.119)
where the following holds:
( ) ( ) ( )

` ` ` `
: : : :
: :
sph
V
sph sph
V
sph V e sph
sph V e sph
V
t t
t
1 1 1
_
_
C
C

(11.120)
( )

` `
:
dev
D
D
t
t
1

(11.121)
Next, we obtain the rate of change of the Cauchy stress tensor:
dev sph
D
D
dev V
V
sph
D
D
V
V
dev sph D
D
V
V
D V
d
d
d
d
d
d
d
d
d
d
d
d
d d


` `
`
`
`
`
` `
` `
` `
` `
-
|
.
|

\
|
o
o
-
o
o
- |
.
|

\
|
o
o
-
o
o

o
o
-
o
o
- -
o
o

o
o
-
o
o
-
o
o

: :
: : ) ( ) , , (

(11.122)
where the following holds, (see equation (11.109)):
dev
D
sph
V
d d

-
o
o
-
o
o
; (11.123)
and
D
D
D
D
D
D D
D
D
D V
V
V
V
V
V V
V
V
V
d
r
r
d
t
r
r
d
d
d
r
r
d
t
r
r
d
d

t
t
t
t
` ` `
`
`
`
o
o

o
o

o
o
o
o

o
o

o
o

o
o
o
o
;
(11.124)
Then, we can express the rates of change
sph

`
and
dev

`
as follows:
11 DAMAGE MECHANICS

609
( )
sph sph sph
V V
V
V e V
sph sph
V V
V
V e V
sph
V V
V
sph V e V
V
V
V
sph V e V V
V
sph sph
d
d
d
d
d
d
d
d d
d



`
`
` `
`
`
` ` `
:
:
: :
: :
(
(

o
o
- -
(
(

o
o
- -
o
o
- -
o
o
- -
o
o
-
o
o

t t
t t
t t
t
t
1
) 1 (
1
) 1 (
1
) 1 (
) 1 (
_
_
_
_
C
C
C
C

(11.125)
and
( )
dev dev dev
D D
D
D e D
dev dev
D D
D
D e D
dev
D D
D
dev D e D
D
D
D
dev D e D D
D
dev dev
d
d
d
d
d
d
d
d d
d



`
`
` `
`
`
` ` `
:
:
: :
: :
(
(

o
o
- -
(
(

o
o
- -
o
o
- -
o
o
- -
o
o
-
o
o

t t
t t
t t
t
t
1
) 1 (
1
) 1 (
1
) 1 (
) 1 (
_
_
_
_
C
C
C
C

(11.126)
with which we can define the following equation:


` `
:
(
(

o
o
-
o
o
- - - -
sph sph
V V
V
dev dev
D D
D
V e V D e D
d d
d d
t t t t
1 1
) 1 ( ) 1 (
_ _
C C
(11.127)
and by comparing the above with (11.111), we can conclude that:
sph sph
V V
V
dev dev
D D
D
V e V D e D d tan
d d
d d


o
o
-
o
o
- - - -
t t t t
1 1
) 1 ( ) 1 (
_ _ _
C C C
(11.128)
11.4 The Elastoplastic-Damage Model in a
Small Deformation Regime
The classical theory of damage has been modified and extended in order to include residual
(plastic) strain. Among the researchers who worked in this area we can mention:
Bazant&Kim (1979), Dragon&Mrz (1979), Ortiz (1985) and Simo&Ju (1987a,b).
Next, we will discuss the elasto-plastic damage model by considering there is an isothermal
process under a small deformation regime.
Fundamentally, we can describe an elasto-plastic damage model as one that presents
residual strain (plastic strain) and also where degradation of the secant stiffness tensor
occurs, (see Figure 11.13).

NOTES ON CONTINUUM MECHANICS

610







Figure 11.13: Stress-strain curve for elasto-plastic damage model.
Several elastoplastic-damage models have been developed, a few of which we mention
below. We will start off from the strain energy function statement:
One of the models: the elastoplastic-damage decoupled model, considers additive
decomposition of the energy into elastic and plastic parts, where both types of
energy are functions of the damage parameter:
) , ( ) , ( d d
p p e
c 1 1 1 - (11.129)
The next model considers the above plus an additional term which is only used in
terms of the damage variable:
) ( ) , ( ) , (
d d p p e
d d c 1 c 1 1 1 - - (11.130)
The next model considers energy decomposition into the elastic part ) , ( d
e
1 and
) (
p p
c 1 which in turn is only used in terms of the plastic strain:
) ( ) , (
p p e
d c 1 1 1 - (11.131)
11.4.1 The Elasto-Plastic Damage Model by Sim&Ju (1987)
in a Small Deformation Regime
Now, we will examine the isotropic damage model proposed by Simo&Ju (1987a,b) in
which the following equation is still valid:
( )

1
1

d -

The effective stress tensor (11.132)


11.4.1.1 Helmholtz Free Energy
The free energy adopted for this model is given by:
) , ( ) ( ) 1 ( ) , , , (
0 p p p
d d q q E - - - : 1 1 1 Helmholtz free energy (11.133)
where d is the damage parameter, q is a set of plastic internal variables,
p
is the plastic
relaxation stress tensor, ) , (
p
q E is the plastic potential, and ) (
0
1 is elastic strain
function (energy density). In the particular case when the constitutive relationship is linear,
we have : :
e e
C
2
1
) ( ) (
0
1 1 .
1
o
E
r
Loading with degradation
Unloading / Elastic loading
E d) 1 ( -
Elastic limit
Y
o

p
r
1
11 DAMAGE MECHANICS

611
11.4.1.2 Internal Energy Dissipation. Constitutive
Equations. Thermodynamic Considerations
Once again, let us consider the alternative form of the Clausius-Planck inequality, (see
equation (11.12)) and if we consider the process to be purely mechanical we obtain:
0 ) , , , ( - d
p
int
q 1
`
`
: D

(11.134)
Then, taking into account the Helmholtz free energy we can obtain its rate of change as
follows:
d
d
d
p
p
p
`
`
` `
`
o
o
-
o
o
-
o
o
-
o
o

1 1 1 1
1 q
q
q

: : ) , , , (

(11.135)
and by substituting (11.135) into (11.134) we obtain:
0
(

o
o
-
o
o
-
o
o
- |
.
|

\
|
o
o
- d
d
p
p
`
`
` `
1 1 1 1
q
q

: :

(11.136)
Next, as the above inequality must be satisfied for any admissible thermodynamic process,
we find:

o
o

1
(11.137)
and by considering the strain energy equation given in (11.133) we can obtain:


-
o
o
-
o
o
p
d
0
) 1 (
1 1
,
p
p
p

o
E o
- -
o
o ) , (q 1
,
q
q
q o
E o

o
o ) , (
p
1
,
0
1
1
-
o
o
d
(11.138)
Now, substituting (11.138) in (11.136) yields:
0
) , ( ) , (
) 1 (
0
) , ( ) , (
) 1 (
0
0
0
0
-
|
|
.
|

\
|
o
E o
-
|
|
.
|

\
|
-
o
E o
-
|
|
.
|

\
|
-
o
o
- -

(
(

-
|
|
.
|

\
|
o
E o
-
|
|
.
|

\
|
o
E o
- - -
(
(

|
|
.
|

\
|
-
o
o
- -
d d
d d
p
p
p
p
p
p
p
p
p
p
`
`
` `
`
`
` `
1
1
1
1
q
q
q q
q
q
q q

: :
: :

(11.139)
Note that the above inequality must hold for any admissible thermodynamic process, so, let
us consider one in which we have 0
p
`
, 0 q` , 0 d
`
, so the only way to fulfill the
entropy inequality in (11.139) is when:
p p
d d

- -
o
o
-
o
o
-
0
0
) 1 ( ) 1 (
1 1

(11.140)
or what is the same:
p
-
0

(11.141)
Then, by substituting the equation in (11.140) into the inequality (11.139) we find:
0
) , ( ) , (
0
-
|
|
.
|

\
|
o
E o
-
|
|
.
|

\
|
-
o
E o
- d
p
p
p
p
`
`
`
1 q
q
q q

: (11.142)
Now, we can assume a purely damage process, and another purely plastic process, with
which we obtain the following restrictions:
NOTES ON CONTINUUM MECHANICS

612
0
) , ( ) , (
0
0

|
|
.
|

\
|
o
E o
-
|
|
.
|

\
|
-
o
E o
- q
q
q q
`
`
`
p
p
p
p
and d

: 1
(11.143)
where 0
0
1 holds, then we can conclude that 0 d
`
.
11.4.1.3 Damage Characterization
Simo&Ju(1987) adopted the following norm:
( ) ( )




` `
` `
: :
0
0
0
2
1
0 0
1 1
) ( 2 2
2
1
) ( 2
t t
t t
o
o

-
1
1 1 1
rate
(11.144)
Note that we are dealing with a symmetrical norm, and also that when the constitutive
equation is linear we obtain
e e
1 2

: : C t t which is the same as that
outlined in the isotropic damage model, (see equation (11.31)).
We can describe the damage state in the material by means of the damage criteria in the
strain space by:
0 ) , ( s - r r

t t G (11.145)
Next, we can define the evolution of the damage variable:
) , ( ; ) , (
) , (
r r d
r
r
d


` ` `
`
`

o
o
t
t
H
G

(11.146)
where
`
is the damage consistency parameter which defines the loading/unloading
condition (the Kuhn-Tucker conditions):
0 )) , ( ; 0 ) , ( ; 0 s r r r

t t G G
` `
` Kuhn-Tucker condition (11.147)
and the consistency (persistency) condition:
0 ) , ( r

t G
`
`

The consistency condition (11.148)


11.4.1.4 The Elastic-Damage Tangent Stiffness Tensor
The elastic-damage tangent stiffness tensor, with no plastic phenomena ( 0
p
`
), can be
obtained similarly to that obtained for the isotropic damage model (see equation (11.81)).
The only difference is that in this situation we have strain energy density ) (
0
1 , not the
elastic strain energy density ) (
e
1 .
Then, by evaluating the rate of change of ) , ( d we obtain:
d
d
d
`
` `

o
o
-
o
o

: ) , ( (11.149)
Remember that according to the equation in (11.140) we have

o o
o
-
o
o
0 2
) 1 (
1
d ,
0

-
o
o
d
and according to (11.146) we have ) , ( d d

t H
` `
. Additionally, when
11 DAMAGE MECHANICS

613
undergoing plastic loading we have ) , ( r r
`
` . So, here, the damage consistency parameter
) , ( r
`
can be determined by the persistency condition,

`
` `
:
0
1
) , (
t
t r , whereby:

`
`
:
0
) , (
t
t d
d
H

(11.150)
Then, given all the above considerations, the equation in (11.149) can be rewritten as:




`
` ` `
:
: :
(

-
o o
o
-
-
o o
o
-


0 0
0 2
0 0
0 2
) , (
) 1 (
) , (
1
) 1 ( ) , (
t
t
t
t
d
d
d d d
H
H
1
1
(11.151)
with which we have obtained the elastic-damage tangent stiffness tensor:
(

-
o o
o
-
0 0
0 2
) , (
) 1 (

t
t d
d
tan_d
H 1
C
(11.152)
Note that the tensor
tan_d
C features major and minor symmetry.
11.4.1.5 Characterization of the Plastic Response. The Elastoplastic-
Damage Tangent Stiffness Tensor
Characterization of the plastic response will be formulated in the effective stress space
and
p
, then the following holds:
p p

-
o
o

) (
) , (
0
1

(11.153)
We can also postulate the yield function in the effective stress space, so that the elastic-
damage domain is characterized by 0 ) , ( s q F .
Then, by assuming there is an associated flow rule, the constitutive equations for plastic
response are given by:

o
|
|
.
|

\
|
-
o
o
o

q ,
) (
0
p
p
1

F
`
`
Plastic flow rule

|
|
.
|

\
|
-
o
o
q q h ,
) (
0

p

1
` ` Plastic hardening law
0 ,
) (
0
s
|
|
.
|

\
|
-
o
o
q
p

1
F Yield criterion
(11.154)
where
p

`
is the rate of change of the plastic relaxation stress tensor, ` is the plastic
consistency parameter, and h is the hardening law and the loading/unloading conditions
are given by:
0 ,
) (
; 0 ; 0 ,
) (
0 0

|
|
.
|

\
|
-
o
o
s
|
|
.
|

\
|
-
o
o
q q
p p

1

1
F F ` `

Kuhn-Tucker
conditions
(11.155)
NOTES ON CONTINUUM MECHANICS

614
Now, to obtain the value of the plastic consistency parameter 0 > ` we turn to the
consistency condition, with 0 ) , ( q F
`
. Then, the rate of change of ) , ( q F is given by:
( ) [ ] 0 , ) , (
o
o
-
o
o

o
o
-
o
o
q
q
q
q
q h

`
`
`
`
`
F F F F
F : :
(11.156)
Note that:
( ) ( )

o o
o
o
o

o
o
o
o

o
|
|
.
|

\
|
-
o
o
o

) ( , ), , (
,
) (
0 2
0
1

1
: :
q q
q
F F
F
` ` `
`
p
p
p

(11.157)
The rate of change of
`
is then evaluated as follows:
|
.
|

\
|
o
o
-
o o
o
-
o o
o
-
o
o

1 1 1
`
`
`
`
`
: :
) ( ) ( ) (
0 2 0 2 0
p p
(11.158)
Next, substituting the above equation into that in (11.156) yields:
( ) [ ]
( )
( ) 0 ,
) ( ) (
0 ,
) ( ) (
0 ,
) ( ) (
0 ) , (
0 2 0 2
0 2 0 2
0 2 0 2

o
o
-
o
o
o o
o
o
o
-
o o
o
o
o

o
o
-
o
o
o o
o
o
o
-
o o
o
o
o

o
o
-
(

|
.
|

\
|
o
o
o o
o
-
o o
o
o
o

o
o
-
o
o

q
q
q
q
q
q
q
q
q
h
h
h

F F F F
F F F F
F F F
F F
F
: : : :
: : : :
: : :
:
1

1

1 1
`
` `
` `
`
`
`
`
`
`

(11.159)
Then, ` can be obtained as follows:
( ) q
q
h ,
) (
) (
0 2
0 2

o
o
-
o
o
o o
o
o
o
o o
o
o
o

F F F
F
: :
: :
1
1

`
`
(11.160)
and by substituting (11.160) into (11.158), the result is:
( )

` `
`
` :
: :
: :
:
(
(
(
(
(

o
o
-
o
o
o o
o
o
o
|
|
.
|

\
|
o o
o
o
o

|
|
.
|

\
|
o
o
o o
o
-
o o
o
|
.
|

\
|
o
o
-
o o
o

q
q
h ,
) (
) ( ) (
) ( ) (
0 2
0 2 0 2
0 2 0 2
F F F
F F
F
1
1 1
1

1

(11.161)
Note, the development of the equation (11.161) in indicial notation is similar to that seen in
Chapter 9 in subsection 9.4.4 The Elastoplastic Tangent Tensor.
Then, the effective elastoplastic tangent stiffness tensor is defined as follows:
( ) q
q
h ,
) (
) ( ) (
) (
0 2
0 2 0 2
0 2







o
o
-
o
o
o o
o
o
o
|
|
.
|

\
|
o o
o
o
o

|
|
.
|

\
|
o
o
o o
o
-
o o
o

F F F
F F
: :
: :
1
1 1
1
tan_p
C
(11.162)
11 DAMAGE MECHANICS

615
Next, if we begin with the equation in (11.132), ( ) d - 1 , we can obtain the rate of
change of the Cauchy stress tensor as:
( ) ( ) 1 1 d d d d
tan_p
`
`
` `
`
- - - - : C (11.163)
Additionally, if we consider that

`
`
:
0
) , (
t
t d
d
H
, (see equation (11.150)), we can
obtain:
( ) ( )


) , (
1 1
0
- - - -
` `
`
` `
: : :
t
t d
d d d
tan_p tan_p
H
C C
(11.164)
Then, by representing the above equation in indicial notation we can conclude that:
( ) ( )
kl kl ij
tan_p
ijkl ij kl kl kl
tan_p
ijkl ij
d
d
d
d r
(

o o - - o r o - r - o ` ` ` `
0 0
) , (
1
) , (
1

t
t
t
t H H
C C
(11.165)
which in tensorial notation becomes:
( )

` ` `
: :
p - tan_d tan_p
d
d C C
(

- -
0

) , (
1
t
t H

(11.166)
where we have introduced the elastoplastic-damage tangent stiffness tensor:
( )
0

) , (
1

- -
t
t d
d
tan_p p - tan_d
H
C C
The elastoplastic-damage
tangent stiffness tensor
(11.167)
which is non-symmetric tensor, since
p - tan_d
C does not feature major symmetry.
11.5 The Tensile-Compressive Plastic-Damage
Model
The next model has two independent internal variables used to describe material
degradation caused by tension and compression. Next, we will describe this model
according to Faria&Oliver (1993).
In this model we assume that the effective stress tensor can be decomposed as follows:
- -
-
(11.168)
where
-
and
-
are the tensile and compressive effective stress tensors, respectively,
and are defined by means of the spectral representations as follows:


-

-
o o
3
1
) ( ) (
3
1
) ( ) (
;
a
a a
a
a
a a
a
n n n n (11.169)
where - is the Macaulay brackets, and where - - - - holds. For example, let us
consider the following effective stresses 0
1
> o , 0
2
> o , 0
3
> o , and also let us suppose
that the Cauchy stress tensor in the principal stress space is given by ) , , (
3 2 1
o - o o . We can
now decompose this tensor as follows:
NOTES ON CONTINUUM MECHANICS

616
_ _
- -
o o
(
(
(

o -
-
(
(
(

o
o

(
(
(

o -
o
o
o
ij ij
ij
3
2
1
3
2
1
0 0
0 0 0
0 0 0
0 0 0
0 0
0 0
0 0
0 0
0 0

(11.170)
We can also verify that the following properties hold:
) ( ) ( ) ( ) (
- - - - - -
- - - 1 Tr Tr Tr Tr : (11.171)
0
- -
:
(11.172)
11.5.1 Helmholtz Free Energy
In this model the free energy is a function of the strain , the plastic strain
p
, and the
parameters
-
d and
-
d and is expressed with:
( ) ( ) ( )
-
-
-
- - -
- - -
e e p
d d d d 1 1 1 1 1 1 , , ,
(11.173)
where
( ) ( ) : : : :
1 1
2
1
) , ( ;
2
1
) , (
-
-
- - -
-
- -

e p e e e p e e
C C 1 1 1 1 (11.174)
In Chapter 7 we verified
e e
D C =
-1
to be the inverse of the elasticity tensor (symmetric
fourth-order tensor), whose components are given by:
[ ] 1 1 I I 1 1 v - v - -
-
-

-
) 1 (
1
2
1
) 2 3 ( 2
1
E
e
j j A j
A
C
( )
)
`

v -
(

- v - =
-
kl ij jk il jl ik
e
ijkl
e
ijkl
E
c c c c c c
2
1
) 1 (
1
1
D C
(11.175)
We can now represent
-
e
1 as:
[ ]
) ( ) (
2 2
) 1 (
) 1 (
2
1
2
1 1

1 1 I
Tr Tr
- -
-
-
-
-
v
-
v -

v - v -
E E
E
e e
:
: : : : C 1

(11.176)
Then, taking into account (11.171), the equation in (11.176) can be rewritten as follows:
( ) [ ]
[ ] ) ( ) (
2
) (
2 2
) 1 (
2
) 1 (
) ( ) ( ) (
2 2
) 1 (
) ( ) (
2 2
) 1 (
2
0
- - - - - - -
- - - - - -
- -
-
v
-
v
-
v -
-
v -

-
v
- -
v -

v
-
v -




Tr Tr Tr
Tr Tr Tr
Tr Tr
E E E E
E E
E E
e
_
: :
:
: 1

(11.177)
Afterwards,
-
e
1 can be expressed in the following ways:
[ ]
) ( ) (
2 2
1
) ( ) (
2 2
) 1 (
) ( ) (
2
) (
2 2
) 1 (
1
2
- - -
-
- - -
- - - - -
-
v
-
v
-
v -

v
-
v
-
v -



Tr Tr Tr Tr
Tr Tr Tr
E E E
E E E
e
e
: : :
:
C
1

(11.178)
11 DAMAGE MECHANICS

617
Likewise, we can obtain
-
e
1 as:
( )
[ ]
-
-
- - - - -
-
-
- - -
-
-
- -
-
- - -
-
-
-
-
-
-
v
-
v -

- v - v -
- -


1 1 I

: : :
: : : :
: : : : : : : :
1
1
1 1 1 1
2
1
) ( ) (
2 2
) 1 (
2
1
) 1 (
2
1
2
1
2
1
2
1
2
1
0
e
e
e e e e e
E E
E
C
C
C C C C
Tr Tr
_
1

(11.179)
Then, taking into account that 0 ) ( <
-
Tr and 0 ) ( >
-
Tr we can guarantee that:
0
2
1
) ( ) (
2
0 0
1
-
v
-

-
-
- - -
-
_ _
: :
e e
E
C Tr Tr 1
(11.180)
Now, the damage variable values lie between the following ranges:
1 0 ; 1 0 s s s s
- -
d d (11.181)
Next, by considering the Helmholtz free energy ( )
- -
d d
p
, , , 1 1 , its rate of change is
evaluated as follows:
-
-
-
-
o
o
-
o
o
-
o
o
-
o
o
d
d
d
d
p
p
` `
`
`
1 1 1 1
1

: : (11.182)
Remember from subsection 6.4.1 Constitutive Equations with Internal Variables in Chapter 6,
that the terms
i
i
A
o
o
-

1
were denoted by the thermodynamic forces, where
i
are the
set of internal variables. Therefore, with the denotations
-
d
1
c ,
-
d
2
c ,
-
A A
1
and
-
A A
2
, the thermodynamic forces becomes:
-
-
-
-
-
-

o
o
-
o
o
-
e e
d d
1
1
1
1
A ; A (11.183)
where we have taken into account that:
( ) ( )

-
o
o
-
o
o
- - -
-
-
-
-
-
-
-
-
e
e
e e
d
d
d d
1
1
1
1
1 1 1 1 1 (11.184)
11.5.2 Damage Characterization
In order to fully define this model we need to characterize the loading/unloading/loading
process.
Next, we will define the norm in the tension stress space:
-
-
- -
: :
1
e
C t
(11.185)
In compression, the norm (based on the Drucker-Prager model and which was obtained by
Faria&Oliver (1993)) can be expressed in terms of the normal octahedral effective stress
(
-
o

oct
) and octahedral tangential stress (
-
t
oct
) and is given by:
NOTES ON CONTINUUM MECHANICS

618
( )
- - - -
t - o o
oct oct
K q 3 ) ( t
(11.186)
proof of which can be found in the chapter on Plasticity (the Drucker-Prager Criterion).
We will now introduce two damage criteria denoted by
-
and
-
:
0 ) , ( ; 0 ) , ( s - s -
- - - - - - - - - -
r r r r t t t t (11.187)
where
-
r and
-
r are the current damage thresholds which serve to remind us where the
damage surface is during the loading/unloading/loading process and whose initial values
are represented by
-
0
r and
-
0
r , respectively.
Then, by defining
-
D
f
1
0
and
-
D
f
1
0
as the stresses after a visible non-linearity in a uniaxial
tensile and compression tests, we obtain
-
o
D
D
oct
f
1
0
1
3
1
) (
3
1
Tr ,
-
- t
D
D
oct
f
1
0 2
1
3
2
3
2
J ,
(see Chapter 9 in the subsection: 9.2.4.2 The Drucker-Prager Yield Criterion), from which
we define the following elastic thresholds:
- -
-
- - -
-
D
D
D D
f K r
E
f
f
E
f r
1
0 0
1
0
1
0
1
0 0
) 2 (
3
3
;
1

(11.188)
11.5.3 Evolution of the Damage Parameters
When observing how the damage parameters evolve the following equations in rates are
considered:
Tension:
- -
-
- -
- -

o
o
r d `
` ` `
;
) (
t
t
(11.189)
Compression:
- -
-
- -
- -

o
o
r d `
` ` `
;
) (
t
t
(11.190)
where
-
G and
-
G are monotonically increasing functions (obtained from experimental
observations), and
-

`
and
-

`
are the damage consistency parameters. Then, the Kuhn-
Tucker conditions are:
Tension
0 ; 0 ; 0 s
- - - -

` `
(11.191)
Compression
0 ; 0 ; 0 s
- - - -

` `
(11.192)
We can verify that when 0 <
-
damage ceases and from the Kuhn-Tucker conditions we
obtain 0
-

`
. When 0 >
-

`
there is damage evolution and here the Kuhn-Tucker
conditions hold if 0
-
, that is, providing that the current state is on the damage surface.
In these conditions it is also true that
- -
r` t and:
( ) {
- - -
t max max ,
0
r r
(11.193)
likewise, for
-
r :
11 DAMAGE MECHANICS

619
( ) {
- - -
t max max ,
0
r r
(11.194)
Then, as the damage evolves, we can state that:
0
) (
; 0
) (

o
o

o
o

-
-
- -
- - -
-
- -
- -

` ` ` ` ` `
t
t
t
t
t
t d d (11.195)
11.5.4 Evolution of the Plastic Strain Tensor
The plastic strain tensor evolution law, adopted by Faria&Oliver (1993), is defined as
follows:



:
:
: 1
) (
-
-
) (

e p
d EH C
`
`
`

(11.196)
where E is the Youngs modulus and 0 is also a material parameter that will control
the rate intensity of plastic strain. The value 0 is equivalent to the elastic damage case.
) (
-
d H
`
is the Heaviside function of the compression damage rate, which was introduced to
stop plastic evolution during compression unloading. Also, note that
1

:

(11.197)
with which the equation in (11.196) can be rewritten as:


) ( ) (
) ( ) (
1 1
1 1
: :
:
: :
:
:
: :
:
:
:
:
-
-
-
-
-
-
-
-
) ( ) (
) (

) (

e e
e e p
d EH d EH
d EH d EH
C C
C C
`
`
`
`
`
`
`
`
`
(11.198)
11.5.5 Internal Energy Dissipation
As how we have proceeded in previous models, we start from the Clausius-Planck
inequality to put restrictions on the thermodynamic variables:
0 ) , , , ( -
- -
d d
p
int
1
`
`
: D (11.199)
Then, by evaluating the rate of change of ) , , , (
- -
d d
p
1 we obtain:
-
-
-
-
- -
o
o
-
o
o
-
o
o
-
o
o
d
d
d
d
d d
p
p
p
` `
` `
`
1 1 1 1
1

: : ) , , , ( (11.200)
and by substituting this into the internal energy dissipation given in (11.199), we find:
0
o
o
-
o
o
-
o
o
- |
.
|

\
|
o
o
-
-
-
-
-
d
d
d
d
p
p
` `
` `
1 1 1 1

: : (11.201)
Then, as the above inequality must be satisfied for any admissible thermodynamic process,
we can draw the conclusion that:

o
o

1

(11.202)
which is the constitutive equation for stress.
NOTES ON CONTINUUM MECHANICS

620
In a small deformation regime, we can decompose additively the infinitesimal strain tensor
into elastic (
e
) and plastic (
p
) parts, i.e.:
p e p e
- - (11.203)
In this way, we can replace two variables (
p
, ) with one
e
.
Then, the free energy defined in (11.173) can be expressed in terms of:
( ) ( ) ( ) ) ( 1 ) ( 1 , ,
e e e e e
d d d d
-
-
-
- - -
- - - 1 1 1 1
(11.204)
whereby the stress is:
( ) ( )
( ) ( )

o
o
o
o
- -
o
o
o
o
-
(

- - -
o
o

o
o

-
-
-
-
-
-
-
-
e
e
e e e
e
e e
e e e e
d d
d d
: :
) (
1
) (
1
) ( 1 ) ( 1
1 1
1 1
1
(11.205)
Then, taking into account that I


o
o
e
(symmetric fourth-order unit tensor), the above
equation becomes:
( ) ( ) ( ) ( )
- - - - - -
-
-
-
-
- - - -
o
o
- -
o
o
-


1 1
) (
1
) (
1 d d d d
e
e e
e
e e
1 1

(11.206)
thereby defining the following effective stresses:
( ) ( )
-
-
-
- -
-
-
-
-

o
o

o
o

d d
e
e e
e
e e
1
1 ) (
;
1
1 ) ( 1 1

(11.207)
In addition, we can express the energy function as:
( ) ( ) ( ) ( )
( )



: : :
: : : : : :
: : : :
2
1
2
1

2
1
2
1

2
1

2
1
1
2
1
1 1 1
1
1 1 1
1 1

- -
- - - - - -
-
-
- -
-
-
-
-
-
- -
-
- -
-
-
-
-
e
e e e
e e e e
d d d d
C
C C C
C C 1 1 1
(11.208)
where we have taken into account the expressions of
-
e
1 and
-
e
1 given in (11.174).
Then, if we consider the internal energy dissipation in (11.201) and the constitutive
equation in (11.202), we can conclude that:
0 - -
o
o
-
o
o
-
o
o
-
o
o
-
-
-
-
-
-
-
-
-
d d d
d
d
d
e e p
p
p
p
int
` `
`
` `
`
1 1
1 1 1 1

: : D (11.209)
where 0
-
e
1 and 0
-
e
1 are positive by definition. Moreover if we consider there be a
process with neither plastic evolution nor the evolution of the parameter
-
d , we can
conclude that 0
-
d
`
. Likewise, we can obtain 0
-
d
`
. Thus:
0 ; 0
- -
d d
` `
(11.210)
If we now consider there to be a purely plastic process, the internal energy dissipation
becomes:
11 DAMAGE MECHANICS

621
( ) 0
) ( ) ( ) (

o
o
-
o
o
-
o
o
o
o
-
o
o
-
p
e
e
p
e
e
p
p
e
e
e
p
p

` ` ` `
: : : : : :
1 1 1 1
(11.211)
Note, the above term
e
e

o
o ) ( 1
can be obtained directly from the equation in (11.208):
( ) ( ) I



- - |
.
|

\
|
o
o
-
o
o

o
o

o
o
2
1
2
1
2
1
2
1
: : : : :
e
e e e e
C
1
(11.212)
Then, restructuring evolution law of
p

`
given in (11.196) we have:
0 ) (
1

) (

-
-



:
:
:
`
`
`
d EH a with a
e p
C (11.213)
and by incorporating the equations (11.212) and (11.213) into the internal energy
dissipation in (11.211), we can obtain:
0 2 ; 0 ; 0
) ( 1

o
o -
1
1
a a
e p
e
e

: : : C
`
(11.214)
after which the internal energy dissipation becomes:
0 2 - -
-
-
-
-
d d a
e e
int
` `
1 1 1 D
(11.215)
Further details about the numerical implementation of the tensile-compressive plastic-
damage model are described in Faria&Oliver (1993).
11.6 Damage in a Large Deformation Regime
The classical hyperelastic models (large deformation regime) discussed in Chapter 8 are not
capable of describing how certain polymers characterized by loss of stiffness behave. This
dissipation phenomenon is known as the Mullins effect which was studied by several
researchers, among whom we can cite: Bueche (1960), (1961), Mullins (1969) and Souza
Neto et al. (1998). In the uniaxial cyclic test, the Mullins effect is phenomenologically
characterized by degradation of the elastic properties, (see Figure 11.14). Let us now
consider the stress-strain curve described in Figure 11.14. During loading (branch [ ] 1 0 - ),
the path is A and unloading is done according to path B . Then, after the unloading is
completed, the material fully recovers its initial state. The second loading will take place
according to path B and follow on to path C . Note that according to the classical
hyperelastic models, loading will take place according to path C A - and unloading would
occur along the same path A C - .








NOTES ON CONTINUUM MECHANICS

622









Figure 11.14: Mullins effects.
11.6.1 Gurtin & Francis One-Dimensional Model
Gurtin&Francis(1981) proposed a simple one-dimensional model in which the current state
of the damage variable is characterized by the maximum axial strain
m
r :
{ ) ( ) (
0
s t
t s
m
r r
s s
max
(11.216)
In this model Gurtin&Francis adopted as the constitutive equation for stress, o, by means
of the current strain state and the damage variable as follows:
) ( ) (
m
g f r o (11.217)
where ) (
m
g r is called the virgin curve and is the relative strain:
m
r
r
(11.218)
The function ) ( f , called the damage master curve, defines the loss of stiffness and satisfies:
1 ) 1 ( f (11.219)
Then, when the maximum strain takes place in the current time ) ( r r
m
, the uniaxial stress
is given by:
) (
m
g r o

(11.220)
Then, the function g defines the uniaxial stress-strain curve obtained from a
monotonically increasing/decreasing uniaxial test. In Figure 11.14, this function is defined
according to path ACE .
Now, to fully describe the material parameters in this model, we need to determine both
the virgin curve and the damage master curve ) ( f . This latter curve is obtained from the
uniaxial test.
11.6.2 The Rate Independent 3D Elastic-Damage Model
Based on the concepts of the Gurtin&Francis model, Souza Neto et al. (1994), (1998)
extended this model to the 3D model which will be explained below.
r
) 2 (
r
) 1 (
r
A
B
C
D
E
o
0
1
2
11 DAMAGE MECHANICS

623
Let us now consider a isotropic hyperelastic material, (see Chapter 8), which is governed by
the free energy,
0
1 , described in terms of the principal stretches (
1
/ ,
2
/ ,
3
/ ). Now, the
Kirchhoff stress tensor eigenvalues, in terms of principal stretches, can be expressed as
follows:
) , , ( :
3 2 1
0
/ / /
/ o
o
/ t
a
a
a a
g
1
(11.221)
The above equation is valid only when we are dealing with virgin material during loading.
Then, the general form of (11.221) can be expressed as follows:
) , , ( ) (
3 2 1
/ / / t
a a
g f (11.222)
As when we looked at the 1D case, we will define a function dependent on the relative
strain, , in 3D, in which the following remains valid:
1 ) 1 ( f
(11.223)
11.6.3 The Damage Variable. Damage Evolution
We will now define the damage variable d , which records the level of damage suffered by
the material during the loading history, as:
{ ) ( ) (
0
0
s t d
t s
1
s s
max
(11.224)
We can now define the relative strain ) ( as follows:
d
0
:
1

(11.225)
If we draw an analogy with the yield surface from classical plasticity, we can define a damage
surface in the principal stretch space as follows:
0 ) , , ( : ) , , , (
3 2 1
0
3 2 1
- / / / / / / d d d 1 (11.226)
For a fixed value of d , the damage surface limits a region in the principal stretches space
where the material behavior is purely elastic, i.e. where there is no damage evolution. As
with plasticity, the damage variable evolution is characterized by the loading/unloading
condition:
0 ; 0 ; 0 d s d d d
` `
(11.227)
So, we can summarize this model as follows
{
0 0 0
0 ) , , ( : ) , , , (
:
) , , ( ) (
) ( ) (
3 2 1
0
3 2 1
0
3 2 1
0
0
d s d
- / / / / / / d

/ / / t

s s
d d
d d
d
g f
s t d
i i
t s
` `
Criterion loading Loading/Un
Surface Damage
Equation ve Constituti The
Variable Damage
iv.
iii.
ii.
i.
1
1
1 max
(11.228)
NOTES ON CONTINUUM MECHANICS

624
11.6.4 The Plastic-Damage Model by Sim & Ju (1989)
We will now discuss the plastic-damage model in a large deformation regime (finite strain)
proposed by Simo&Ju (1989). Note that, the way in which the plastic-damage model by
Simo&Ju (1987a,b) in a small deformation regime was described, the extension of this
model to large deformation regime is almost trivial.
11.6.4.1 Specific Helmholtz Free Energy
The Helmholtz free energy (per unit mass) in the reference configuration is given by:
( ) ) , (
2
1 1
) ( ) 1 (
) , (
1
) ( ) 1 ( ) , , , (
0
0
0
0
p p
p p p
d
d d
S S 1
S S S
A C C
A E C A C
E -
(

- - -
E - - -
:
:
j

j

(11.229)
where
T
F F C is the left Cauchy-Green deformation tensor, F is the deformation
gradient, ( ) 1 - C E
2
1
is the Green-Lagrange strain tensor, S is the second Piola-
Kirchhoff stress tensor, A is the set of internal plastic variables, d is the damage variable,
p
S is the plastic relaxation stress tensor, and ) , (
p
S A E is the plastic potential function.
Moreover, it is noteworthy that
0
,
p
S : E
0
1
j
and ) , (
p
S A E have the unit of energy per
unit mass.
11.6.4.2 Internal Energy Dissipation. Constitutive
Equations. Thermodynamic Considerations
Remember that in Chapter 5 the alternative form of the Clausius-Planck inequality in the
reference configuration is given by:
[ ] 0
0
- - j j
` ` `
T
int
E : S D
(11.230)
in which all variables are described in the reference configuration, and
0
j is the mass
density, j is the specific entropy, T is temperature, and is the specific Helmholtz free
energy (per mass unit). In isothermal processes we have 0 T
`
, in which the Clausius-Planck
inequality becomes:
0
0
- j
` `
E : S
int
D (11.231)
Then, by evaluating the rate of change of the Helmholtz free energy, ) , , , ( d
p
A C S , we
obtain:
d
d
d
p
p
p
` ` ` ` `
o
o
-
o
o
-
o
o
-
o
o


A
A
C
C
A C S
S
S : : ) , , , ( (11.232)
Next, if we consider the expression of the Helmholtz free energy given in (11.229) we find:
p
d S
0
0
2
1 ) (
) 1 (
j

-
o
o
-
o
o
C
C
C
,
p
p
p
S
S
S o
E o
- -
o
o ) , ( 1
0
A
E
j

,
A
A
A o
E o

o
o ) , (
p
S
,
) (
0
C

-
o
o
d

(11.233)
11 DAMAGE MECHANICS

625
After that, substituting (11.232) into the entropy inequality (11.231), and by considering the
equations in (11.233), we have:
0 ) (
1
2
1 ) (
) 1 (
0
0 0
0
0

-
o
E o
-
(

-
o
E o
-
(

-
o
o
- - d d
p
p
p
` ` ` ` `
C A
A
E C
C
C
E
j j

j S
S
S S : : :

(11.234)
or:
0 ) (
1
2
1
) (
) 1 (
0
0 0
0
0
0
0
-
o
E o
-
(

-
o
E o
-
-
o
o
- -
d
d
p
p
p
` ` ` `
` `
C A
A
E C
C
C
C
E
j j
j
j

j
S
S
S
S
: :
: :

(11.235)
Remember from Chapter 8 (Hyperelasticity) that
E
E
C
C
o
o

o
o ) ( ) (
2
0 0

, and in addition
( ) C E C E
` `
2
1
2
1
- 1 holds, so we can now rewrite the entropy inequality as:
0 ) (
) (
) 1 (
0
0 0 0
0
0
-
o
E o
-
(

-
o
E o
-
(

-
o
o
- - d d
p
p
p
` ` `
`
C A
A
E E
E
E
j j j

j S
S
S S : : (11.236)
which must be satisfied for any admissible thermodynamic process, so, by considering the
following process 0 S
p
`
, 0 A
`
, 0 d
`
, we obtain:
0
) (
) 1 (
0
0

(

-
o
o
- - E
E
E
`
:
p
d S S

j (11.237)
Furthermore, if we consider two processes where 0 > E
`
and 0 < E
`
, the only way of
enforcing the inequality is when the term within the brackets is equal to zero, i.e.:
p p
d d S S S S - - -
o
o
-
0
0
0
) 1 (
) (
) 1 (
E
E
j

(11.238)
where
E
E
o
o

) (
0
0
0

j S is the non-damage stress tensor. Additionally, if we take into
account that S S ) 1 ( d - and
p p
d S S ) 1 ( - , where
p
S is the plastic relaxation effective
stress tensor, we obtain:
p
p
p
d d d
d
S S S
S S S
S S S
-
- - - -
- -
0
0
0
) 1 ( ) 1 ( ) 1 (
) 1 (

(11.239)
Then, substituting the constitutive equation in (11.238) into the inequality in (11.236)
yields:
0 ) (
0
0 0 0
-
o
E o
-
(

-
o
E o
- d
p
p
` ` `
C A
A
E j j j S
S
: (11.240)
Similarly, if there is a pure damage process and then another pure plastic process, we can
obtain the following restrictions:
NOTES ON CONTINUUM MECHANICS

626
0
) , ( ) , (
; 0 ) (
0 0
0
0

o
E o
-
(

-
o
E o
- A
A
A
E
A
C
` ` `
p
p
p
p
d
S
S
S
S
j j j :
(11.241)
11.6.4.3 Damage Characterization
Here, we will adopt the following energy norm:
( ) ( ) E E
E
E E
E E
E E
` ` ` `
: :
0
0
0
2
1
0 0
1 1
) ( 2 2
2
1
) ( 2 S
t t
t t
o
o

-
1
1 1 1
rate
(11.242)
We emphasize here that ) ( ) (
0
0
0
E E j 1 has the unit of energy per unit volume (energy
density).
We can then characterize the state of damage in the material by means of the damage
criterion defined in the principal strain space as:
0 ) , ( s - r r
E E
t t G (11.243)
Then, we can define the damage evolution law as follows:
) , ( ; ) , (
) , (
r r d
r
r
d E
E

` ` `
`
`

o
o
t
t
H
G


(11.244)
where
`
is the damage consistency parameter which defines the loading/unloading Kuhn-
Tucker conditions:
0 ) , ( ; 0 ) , ( ; 0 s r r r
E E
t t G G
` `
` Kuhn-Tucker conditions (11.245)
and the persistency (consistency) condition:
0 ) , ( r
E
t G
`
`

The consistency condition (11.246)


11.6.4.4 The Hyperelastic-Damage Tangent Stiffness Tensor
In the absence of plastic phenomena we have 0 S
p
`
, which along with the rate of change
of ) , ( d E S gives us:
d
d
d
d
d
` ` ` `
`

o
o
-
o
o
=
o
o
-
o
o

S S S S
S E
E
E
E
E : : ) , ( (11.247)
Then, taking into account that
0
S
S
-
o
o
d
, d
d
d
` `
`
o
o
-
o
o

S S
S E
E
E : ) , ( , ) , ( d d
E
t H
` `
, (see
equation (11.244)), and also that in plastic loading ) , ( r r E
`
` holds, in which the damage
consistency parameter ) , ( r E
`
can be evaluated by means of the consistency condition
E E
E
E
` ` `
:
0
1
) , ( S
t
t r , the equation in (11.247) can be rewritten as follows:
) , (
1 ) (
) 1 (
) , (
) (
) 1 ( ) , (


0 0
0 2
0
0 2
0
d d
d
d
d d
E
E
E
E E
E E
E
E
E E
E
E
t
t
t
H
H
` `
`
` `
: :
:
S S
S
S
-
o o
o
-

o
o
-
o o
o
-

j

11 DAMAGE MECHANICS

627
E
E E
E
E
E
E
`
`
:
(

-
o o
o
-
0 0
0 2
0
) , ( ) (
) 1 ( ) , ( S S S
t
t d
d d
H
j (11.248)
Thus, we can define the hyperelastic-damage tangent stiffness tensor in the reference
configuration as:
0 0
0 2
0
) , ( ) (
) 1 ( S S -
o o
o
-
E
E
E E
E
t
t d
d
tan_d
H
j C (11.249)
The tensor
tan_d
C features major and minor symmetry, which is, primarily, due to the fact
that the norm adopted is symmetrical.
11.6.4.5 Characterization of the Plastic Response. The Effective
Elastoplastic-Damage Tangent Stiffness Tensor
Characterization of the plastic response is formulated in the effective stress spaces S and
p
S after which the following holds:
p
S S -
o
o

E
E) (
0
0

j
(11.250)
We will now postulate the yield function in the effective stress space, such that the elastic-
damage domain is characterized by 0 ) , , ( s A E
p
S F .
Then, if we assume the associated flow rule holds, the constitutive equations for the plastic
response are given by:
0 ) , , ( ; ) , , ( ;
) , , (
s
o
o
A E A E A
E
A E
H
p p
p
p
S S
S
S F
F
` `
`
`

(11.251)
where
p
S
`
is the rate of change of the plastic relaxation effective stress tensor, ` is the
plastic consistency parameter, and H is the hardening law.
Now, the loading/unloading condition can be expressed as follows:
0 ) , , ( ; 0 ; 0 ) , , ( s A E A E
p p
S S F F ` ` Kuhn-Tucker conditions (11.252)
Then, to obtain the value of the plastic consistency parameter 0 > ` we turn to the
consistency condition, which requires that 0 ) , , ( A E
p
S F
`
. Then the rate of change of
) , , ( A E
p
S F is given by:
( ) [ ] 0 ,
) , , (
0 ) , , (

o
o
-
(

o
o
o
o
-
o
o

o
o
-
o
o
-
o
o

A
A E
A E
E
E
A
A
E
E
A E
H S
S
S
S
S
S
` `
`
`
`
` `
F F F F
F F F
F
p
p
p
p
p
: :
: :

(11.253)
Next, ` can be obtained as follows:
( ) A
A E
E
E
H , S
S
o
o
-
o
o
o
o
o
o
-

F F F
F
:
:
p
`
`
(11.254)
The rate of change of S is given by:
NOTES ON CONTINUUM MECHANICS

628
p p
S S S S
`
`
`
-
o o
o
-
o
o
E
E E
E
E
E
:
) ( ) (
0 2 0
1 1

(11.255)
Then, substituting (11.254) into (11.255) yields:
E
A E
E E
E E
E
E
A E
E
E
E
E E
E
E
E
E E
E
E
E E
E
H
H
`
`
`
`
`
`
`
`
:
:
:
:
:
: :
(
(
(
(

o
o
-
o
o
o
o
o
o
o
o
-
o o
o

o
o
|
|
|
|
.
|

\
|
o
o
-
o
o
o
o
o
o
-
-
o o
o

o
o
-
o o
o
-
o o
o

F F F
F F
F
F F F
F
F
p
p
p
S
S
S S
) (
) (
) ( ) (
0 2
0 2
0 2 0 2
1
1

1 1

(11.256)
Next, we will define the effective elastoplastic tangent stiffness tensor as follows:
H
A E
E E
E E
E
o
o
-
o
o
o
o
o
o
o
o
-
o o
o


F F F
F F
:
p
tan_p
S
) (
0 2
1
C
(11.257)
11.6.4.6 The Elastoplastic-Damage Tangent Stiffness Tensor
Starting from the equation S S ) 1 ( d - , the rate of change is evaluated as follows:
( ) S S S S 1 ) 1 ( d d d d
tan_p
` ` `
`
`
- - - - E : C
(11.258)
In addition, taking into account that E
E
E
` `
:
0
) , (
S
t
t d
d
H
, we obtain:
( ) ( )
( ) E
E E E
E
E
E
E
`
` ` ` ` `
:
: : :
(

- -
- - - -
0
0

) , (
1

) , (
1 1
S S
S S S S
t
t
t
t
d
d
d
d d d
tan_p
tan_p tan_p
H
H
C
C C
(11.259)
or:
E
` `
:
p - tan_d
C S
(11.260)
where we have introduced the elastoplastic-damage tangent stiffness tensor:
( )
0

) , (
1 S S - -
E
E
t
t d
d
tan_p p - tan_d
H
C C
The elastoplastic-damage
tangent stiffness tensor
(11.261)
which is a non symmetric tensor, due to the lack of major symmetry.
11.6.5 The Plastic-Damage Model by Ju(1989)
We will now discuss the formulation of damage based on strain coupled with the strain
elastoplastic model described by Ju (1989). This formulation is based on the multiplicative
decomposition of the deformation gradient into an elastic (
e
F ) and plastic (
p
F ) part,
11 DAMAGE MECHANICS

629
where
p e
F F F holds. For further details on multiplicative decomposition see Chapter
9 in subsection 9.6 Large-Deformation Plasticity.
11.6.5.1 Helmholtz Free Energy
The Helmholtz free energy per unit volume in the reference configuration is given by:
[ ] ) ( ) , ( ) 1 ( ) , , ( ) 1 ( ) , , , (
0 0 0
A E E A E E A E E
p p e p p
d d d 1 1 1 1 1 - - -
(11.262)
where ( ) 1 - C E
2
1
is the Green-Lagrange strain tensor,
T
F F C is the right Cauchy-
Green deformation tensor, F is the deformation gradient, and
p
E is the plastic part in the
reference configuration.
11.6.5.2 Internal Energy Dissipation. Constitutive
Equation. Thermodynamic Considerations
Remember that in Chapter 5 the alternative form of the Clausius-Planck inequality in the
reference configuration is given by:
[ ] 0
0
- - 1 j j
` ` `
T
int
E : S D

(11.263)
where
0
j is the mass density, j is the specific entropy (per unit mass), T is temperature,
and 1 is the Helmholtz free energy (per unit volume). In isothermal processes we have
0 T
`
, with which the Clausius-Planck inequality becomes:
0 - 1
` `
E : S
int
D (11.264)
Then, by taking the rate of change of ) , , , ( d
p
A E E 1 , we obtain:
d d d
d
d
d
p
p
p
p
p e p e
p
p
p
` ` ` `
` ` ` `
`
`
) , , (
) ( ) , (
) 1 (
) , (
) 1 (
) , , , (
0
0 0 0
A E E A
A
A
E
E
E E
E
E
E E
A
A
E
E
E
E
A E E
1
1 1 1
1
1 1 1 1
1
-
o
o
-
o
o
- -
o
o
-
o
o
-
o
o
-
o
o
-
o
o

: :
: :

(11.265)
Next, substituting the above equation into the internal energy dissipation given in (11.264),
yields:
0 ) , , (
) ( ) , (
) 1 (
) , (
) 1 (
0
) , , ( ) ( ) , (
) 1 (
) , (
) 1 (
0
0 0 0
0 0 0 0
-
(
(

o
o
-
o
o
- - -
|
|
.
|

\
|
o
o
- -

o
o
-
o
o
-
o
o
- -
o
o
- -
d d d
d
d
d d
p
p
p
p
p e p e
p p
p
p
p e p e
` ` ` `
` ` ` ` `
A E E A
A
A
E
E
E E
E
E
E E
A E E
A
A
A
E
E
E E
E
E
E E
E
1
1 1 1
1 1 1 1
: :
: : :
S
S
(11.266)
Then, as the above inequality must satisfy for any admissible thermodynamic process, we
obtain:
E
E E
o
o
-
) , (
) 1 (
0
p e
d
1
S
(11.267)
Thus the entropy inequality becomes:
NOTES ON CONTINUUM MECHANICS

630
0 ) , , (
) (
) 1 (
) , (
) 1 (
0
0 0
-
o
o
- -
o
o
- - d d d
p
p
p
p
p e
` ` `
A E E A
A
A
E
E
E E
1
1 1
: (11.268)
If only we have a process characterized by damage, we have:
0 ) , , (
0
d
p
`
A E E 1 (11.269)
and if there is a purely plastic process, the following must be satisfied:
0
) ( ) , (
0 0

o
o
-
o
o
- A
A
A
E
E
E E
` `
p
p
p
p e
1 1
: (11.270)
Then we can define the effective stress tensor as follows:
E
E E
o
o

) , (
) 1 (
1
0
p e
d
1
S S (11.271)
11.6.5.3 Characterization of Damage. The Tangent Damage
Hyperelasticity Tensor
The equation in (11.269) leads us to adopt the free energy ) , , (
0
A E E
p
1 so as to
characterize the loading/unloading conditions. That is, we will adopt the following variable:
) , , (
0
A E E
p
1 = (11.272)
Then, the damage criterion is represented by:
0 ) , ( s - =
t t t t
r r (11.273)
where ) , (
t t
r represent the current values of ) , ( r , at time t , where
t
r is the current
damage threshold, and by adopting
0
r as the material elastic threshold
0
r r
t
holds.
The evolution of the variables d and r are then given, respectively, by:

` `
`
`
E
t t t
t
r a s d d ; ) , , , ( (11.274)
where the Kuhn-Tucker conditions hold:
0 ) , ( ; 0 ) , ( ; 0 s
t t t t
r r r
` `
` Kuhn-Tucker conditions (11.275)
We can then obtain the parameter
`
by imposing loading:
>
`
`
` `
0 ) , ( ) , ( 0
t t t t
r r (11.276)
where - is the Macaulay brackets.
Then, the variable
t
r is defined as follows:
[ ]
{
t t t t t t
r r r r
t s
>
=
max max , ; 0 ) , ( ) , ( 0
0
, 0
`
`
` `

(11.277)
11.6.5.4 The Elastic-Damage Tangent Stiffness Tensor
In the absence of plastic phenomena we have 0
p
E
`
, thus by evaluating the rate of
change of the stress tensor given in (11.267),
E
E E
o
o
-
) , (
) 1 (
0
p e
d
1
S , we obtain:
11 DAMAGE MECHANICS

631
d d d
p e p e
` ` `
E
E E
E
E E
E E
E
o
o
-
o o
o
-
) , ( ) , (
) 1 ( ) , (
0 0
2
1 1
: S

) , (
) 1 (
) , (
) 1 (
) , (
) 1 ( ) , (
0
2
0
2
0
2
E -
o o
o
-
-
o o
o
- = -
o o
o
-
` `
` ` ` ` `
S
S S S
E
E E
E E
E
E E
E E
E
E E
E E
E
:
: :
p e
p e p e
d
d d d d d
1
1 1
(11.278)
Note that during loading E E
`
`
`
r d is valid, and also based on the expression
) , , (
0
A E E
p
1 = , the non-plastic process, we obtain:
E E
E
` ` `
: : S
o
o

0
1
(11.279)
Then
E
E E
E E
E E
E E
E E
E
E E
E E
E
`
` ` ` ` `
:
: : :
(
(

E -
o o
o
-
E -
o o
o
- E -
o o
o
-
S S
S S S S
) , (
) 1 (
) , (
) 1 (
) , (
) 1 ( ) , (
0
2
0
2
0
2

p e
p e p e
d
d d d
1
1 1

(11.280)
where we have introduced the elastic-damage tangent stiffness tensor:
S S E -
o o
o
-
E E
E E ) , (
) 1 (
0
2
_
p e
d tan
d
1
C
(11.281)
11.6.5.5 Characterization of Plastic Response. The elastoplastic
Tangent Stiffness Tensor.
Now, by assuming there is an associated flow rule, the constitutive equation for the plastic
response is given by
E
A E E
E
o
o
-
-
) , , (
1

p
p
F
: M `
`
, where
p p
p e
E E E
E E
o
o

o o
o

S ) , (
0
2
1
M .
Now, in the stress space we have:
E
A E E
E
o
o
-
) , , (

p
p p
F
`
`
`
: M S
(11.282)
Then, the law of evolution for the variable A is given by:
) , , ( A E E A H
p
`
`
(11.283)
where H is the generalized hardening law.
The loading/unloading conditions are then given by:
0 ) , , ( ; 0 ; 0 ) , , ( s A E E A E E
p p
F F ` ` Kuhn-Tucker conditions (11.284)
Next, to obtain the plastic consistency parameter 0 > ` we turn to the consistency
condition which requires that 0 ) , , ( A E E
p
F
`
. Then the rate of change of ) , , ( A E E
p
F is
given by 0 ) , , (
o
o
-
o
o
-
o
o
A
A
E
E
E
E
A E E
` ` ` `
F F F
F
p
p
p
: : , which can be expressed as:
[ ] 0 ) , , (
1

o
o
-
(

o
o
-
o
o
-
o
o

-
H
A E E
E
E
A E E ` `
` `
F F F F
F : : : M
p
p
(11.285)
NOTES ON CONTINUUM MECHANICS

632
Afterwards, ` can be obtained as follows:
H
A E E
E
E
o
o
-
o
o
o
o
o
o

-
F F F
F
: :
:
1
M
p
`
`
(11.286)
and the rate of change of
E
E E
o
o

) , (
0
p e
1
S is evaluated as:
E
E
E E
E E
E
E E
E E
E E
E E
E E
E
E E
E E
E
E E
E E
o
o
-
o o
o
-
o o
o

-
o o
o

o o
o
-
o o
o

F ) , ( ) , (
) , ( ) , ( ) , (
0
2
0
2
0
2
0
2
0
2

1 1
1 1 1
`
`
`
`
` ` ` `
`
: :
: : : :
p e
p
p e
p
p e
p
p
p e p e
S
S M
(11.287)
Then, if we consider that
E E
E E
o o
o

) , (
0
2
0
p e
e
1
C , and that the value of ` is given by the
equation in (11.286), we obtain:
E
H
A E E
E E
E
H
A E E
E
E
E
E
E
E
E
E E
E E
`
`
`
` `
`
` `
:
: : : :
:
:
: :
(
(
(
(

o
o
-
o
o
o
o
o
o
o
o
-
o
o
|
|
|
|
.
|

\
|
o
o
-
o
o
o
o
o
o
-
o
o
-
o
o
-
o o
o

- -



1
0
1
0
0
0
2
) , (
F F F
F
F F F
F F
F
F F
M
C
M
C
C
p
e
p
e
e
p e

1
S

(11.288)
thus we can define the effective elastoplastic tangent stiffness tensor (
p tan _
C ) as:
H
A E E
E E
o
o
-
o
o
o
o
o
o
o
o
-
-

F F F
F F
: :
1
0
_
M
C C
p
e p tan
(11.289)
11.6.5.6 The Elastoplastic-Damage Tangent Stiffness Tensor
The elastoplastic-damage tangent stiffness tensor is defined according to the relationship
E
` `
:
p d tan _ _
C S . Thus, we can start from the rate of change of S S ) 1 ( d - :
S S S S ) 1 ( ) 1 (
_
E - - - -
` ` `
`
`
E :
p tan
d d d C
(11.290)
Then, the rate of change of (11.272), ) , , (
0
A E E
p
1 = , is given by:
(

o
o
-
o
o
o
o
-
o
o
-
o
o
-
o
o
=
-
H
A E E
E
A
A
E
E
E
E
A E E
0 1 0
0 0 0
0
) , , (
1 1

1 1 1
1
F
: : :
: : :
M
p
p
p
p
`
`
`
` ` ` `
S
(11.291)
or
11 DAMAGE MECHANICS

633
[ ] Y
X
F
F F F
F
F
F

0 1 0
1
0 1 0
|
|
|
|
.
|

\
|
o
o
-
(

o
o
-
o
o
o
o
|
|
|
|
.
|

\
|
o
o
-
o
o
o
o
o
o
-
(

o
o
-
o
o
o
o
-
-
-
-

E
E
E
H
A E E
A E E
E
E
E
H
A E E
E
H
`
`
`
`
` `
`
:
:
: :
: :
:
:
: : :
S
S
S
1 1
1 1

M
M
M
p
p
p

(11.292)
where we have introduced the following scalars H
A E E o
o
-
o
o
o
o

-
F F F
X : :
1
M
p
,
H
A E E o
o
-
o
o
o
o

- 0 1 0
1 1 F
Y : : M
p
. Then, substituting (11.292) into (11.290) yields:
[ ]
S S S
S S S S
) 1 (
) 1 ( ) 1 (
_
_ _

o
o
|
|
.
|

\
|
E - E - -

(
(
(
(

|
|
|
|
.
|

\
|
o
o
- E - - E - -
E
E
E E
E
E
E E E
` ` `
`
` ` ` `
`
: : :
:
: : :
F
X
Y
Y
X
F
p tan
p tan p tan
d
d d
C
C C

(11.293)
in indicial notation:
kl ij
kl
ij kl
p tan
ijkl
ij kl
kl
ij kl kl kl
p tan
ijkl ij
E
E
d
E
E
E E d
`
` ` ` `
(

o
o
|
|
.
|

\
|
E - E - -
o
o
|
|
.
|

\
|
E - E - -
S S S
S S S S
F
X
Y
F
X
Y
) 1 (
) 1 (
_
_
C
C
(11.294)
which in tensorial notation becomes:
E
E
E
A E E
H
A E E
E
E
H
`
`
`
`
:
:
: :
: :
:
p d tan
p
p
p tan
p tan
d
d
_ _
1
0 1 0
_
_
) 1 (
) 1 (
C
M
M
C
C

(
(
(
(

o
o

|
|
|
|
.
|

\
|
o
o
-
o
o
o
o
o
o
-
o
o
o
o
E - E - -
(

o
o

|
|
.
|

\
|
E - E - -
-
-
F
F F F
F
X
Y
F
S S S
S S S S
1 1

(11.295)
where we have introduced the elastoplastic-damage tangent stiffness tensor:
E
A E E
H
A E E
H
o
o

|
|
|
|
.
|

\
|
o
o
-
o
o
o
o
o
o
-
o
o
o
o
E - E - -
-
-
F
F F F
F
S S S ) 1 (
1
0 1 0
_ _ _
: :
: :
M
M
C C
p
p
p tan p d tan
d
1 1
(11.296)


12 Introduction to Fluids










12.1 Introduction
In this chapter, we will introduce an important branch of continuum mechanics: fluid
mechanics with which we intend to study fluids in motion or at rest. These can be classified
into:

Gases
Liquids
Fluids
There are several areas where fluids mechanics can be applied, e.g. meteorology,
oceanography, aerodynamics, hydrodynamics and engineering, among others.
Fundamentally, we can state that solids can resist shear stress while liquids have very low
(viscous fluids, e.g. oil) or no resistance to it (non-viscous fluids, e.g. water).
Both gases and liquids are materials consisting of molecules (an agglomeration of two or
more atoms) colliding with each other. To treat fluids with assumption of continuum
mechanics properties (e.g. mass density, pressure and velocity) are treated as continuous
functions. Then, treating a system of molecules as a continuous medium is valid when
comparing the mean free path of molecules ( A ) (average distance particles travel before
colliding with each other) with the characteristic physical length scale (
C
).For example,
for solids and liquids we have cm
7
10
-
= A and for gases cm
6
10
-
= A , Chung (1996). Then,
the ratio
C

A
is known as the Knudsen number ( Kn ). If this number is much smaller than
unity, the domain can be treated as a continuum, otherwise we must use statistical
mechanics to obtain the governing equations of the problem with which we can establish
that:
12
Introduction to Fluids
635 , Notes on Continuum Mechanics, Lecture Notes on Numerical
Methods in Engineering and Sciences 4, DOI 10.1007/978-94-007-5986-2_14,
International Center for Numerical Methods in Engineering (CIMNE), 2013
E.W.V. Chaves
NOTES ON CONTINUUM MECHANICS

636
approach c microscopi
approach c macroscopi
>
A

<<
A

1
1
C
C
Kn
Kn

(12.1)
Let us now consider a fluid is between two surfaces separated by a distance d , (see Figure
12.1). The lower surface is fixed whilst the upper surface is moving at the constant velocity
0
v . We can then observe that the force required to maintain this motion is given by:
d
v
A
F
0
j (12.2)
where A is the surface area and j is the fluid viscosity. The above equation indicates that
the shear stress
A
F
is proportional to how the velocity varies with distance (the velocity
gradient).








Figure 12.1: Motion of the plate.
12.2 Fluids at Rest and in Motion
12.2.1 Fluids at Rest
By means of experiments, it can be proven that a fluid at rest or with uniform flow is free
of tangential stresses, i.e. the shear stress components are zero. Then, the traction vector on
the surface element is only a function of pressure whereby we conclude that the traction
vector
)

(n
t
,
which acts on a surface is collinear with the normal n

and is given by:


Tensorial notation components
n t
n

0
)

(
p -
,

i i
p n t

0
)

(
-
n

(12.3)
where
0
p is the hydrostatic pressure.
The traction vector can also be expressed in terms of the Cauchy stress tensor, (see
Chapter 3), as n n t
n

)

(
p -
,
, where p is given in terms of the mean stress as p
m
- o .
Then, in the case of fluids at rest or with uniform motion we have:
3
kk
p p
o
- (12.4)
F A

0
v v
0 v
d
fixed
in motion
12 INTRODUCTION TO FLUIDS

637
In this case, any direction is a principal direction and the stress tensor, ) (
0
p , is
represented as follows:
Tensorial notation Components
1
0
p -
ji ij
p c
0
- o

(12.5)
The constitutive law above was described by Bernoulli for a non-viscous fluid.
Unfortunately this equation is not valid for any fluid, for instance a fluid in motion. The
negative sign indicates a compressive stress with a positive pressure value.
Figure 12.2 shows the hydrostatic state by means of the Mohrs circle, here reduced to a
single point. Note that for a fluid at rest the maximum tangential (shear) stress is zero and
here the stress tensor is spherical.








Figure 12.2: Hydrostatic pressure.
NOTE: In general, for a fluid in motion, the parameter ( p ) and the hydrostatic pressure
(
0
p ) do not match. However, as we saw previously, for a fluid at rest p p
0
holds.
12.2.2 Fluids in Motion
For a fluid in motion, the shear stress components are, generally speaking, nonzero and the
stress tensor, ) , ( p , is usually decomposed into:
- - 1 p
ij ji ij
p t - - o c
(12.6)
where p is the thermodynamic pressure and is the viscous stress tensor.
NOTE: The thermodynamic pressure ( p ) is a variable that is related to other
thermodynamic variables such as mass density ( j ) and absolute temperature ( T ) by means
of the equations of state 0 ) , , ( T p f j .
Note, in general, the thermodynamic pressure is different from the hydrostatic pressure, i.e.
0
p p .
12.3 Viscous and Non-Viscous Fluids
Real fluids are compressible and viscous, although, in many practical cases this viscosity or
compressibility can be overlooked. So, traditionally, they have been classified into viscous


3 2 1
o o o

N
o

S
o
b) Mohrs circle a) Point under hydrostatic pressure

0
p

0
p

0
p
NOTES ON CONTINUUM MECHANICS

638
and non-viscous fluids, e.g. water (an incompressible non-viscous fluid), air (a compressible
non-viscous fluid) and oil (incompressible viscous fluid).
12.3.1 Non-Viscous Fluids (Perfect Fluids)
A non-viscous fluid is free of shear stress (negligible), so the viscous stress tensor is zero:
0 = Non-viscous fluids (12.7)
In this case we have a perfect fluid and the equation in (12.3) holds.
12.3.2 Viscous Fluids
With viscous fluid in motion, resistance to tangential movement cannot be ignored, so, we
have 0 . Then, by taking the equation in (12.6) and multiplying both sides of it by the
Kronecker delta,
ij
c , i.e. by taking the trace of , we obtain:

ii ii
ii ii
ij ij ij ij ij ij
p
p
p
m
t - - o
t - - o
t - - o
o
3
1
3
1
3
c c c c

3
) (
3
) (
) ( 3 ) (

: : :
Tr Tr
Tr Tr
- -
- -
- -
p
p
p

1 1 1 1

(12.8)
Then, with regard to hydrostatic pressure we can state that:
- For viscous or non-viscous fluids at rest 0 t
ij
, thus:
p p p
0
(12.9)
- For an incompressible fluids, p is an independent mechanical variable
- Compressible Fluids
In general, pressure is a function of mass density and temperature which are
related by means of the equation of state:
) , ( T p p j Constitutive equation for pressure. (12.10)
where p is pressure, j is mass density, and T is absolute temperature. For example, the
equation of state for ideal gases is RT p j , where R is the gas constant.
When temperature is not included in the equation of state, ) ( p j j , then the state change
is called barotropic.
It can then be shown that in reversible adiabatic processes, an ideal gas is governed by the
barotropic relationship:
v v
p
c
R
c
c
p
-

1 ; constant
j

(12.11)
where
p
c is specific heat at constant pressure,
v
c is the specific heat at a constant volume,
R is the gas constant and a perfect incompressible fluid is governed y the barotropic
equation:
constant j
(12.12)
12 INTRODUCTION TO FLUIDS

639
where we have
0
p p .
12.4 Laminar and Turbulent Flow
Fluid flow is considered to be laminar when the various fluids layers move in a parallel,
uniform and regular fashion, (see Figure 12.3). As we will see later, the Navier-Stokes
equations are only valid for laminar fluids. For such flows, it is well established that the
shear stress ( ) is proportional to the velocity gradient. Therefore, the Navier-Poisson
constitutive equations describe laminar flow behavior well.
A laminar flow is identified generally by so named Reynolds number (
e
R -dimensionless), and
is given by:
* *

v

C C

, ,
v v
j
j
e
R
(12.13)
where v
,
is the mean velocity of the object relative to the fluid;
C
is the characteristic
length;
*
j is the dynamic viscosity, and
*
v is the kinematic viscosity which is given by
j
j
*
*
v . Then, the SI units of these are [ ] s m /
2 *
v and [ ] s Pa
ms
kg
s
m
m
kg

2
3
*
j .






Figure 12.3: Laminar flow.
We state that a fluid flow is turbulent when stresses and velocities at each point randomly
fluctuate over time, (see Figure 12.4).





Figure 12.4: Turbulent flow.



Laminar flow
Turbulent flow
NOTES ON CONTINUUM MECHANICS

640
12.5 Particular Cases
Reminder:
Remember that the material time derivative of velocity provides us with the acceleration,
i.e.:
k i k
i
i
i
i
v v
t
t v
v
Dt
Dv
t a
t
t
t
Dt
t D
t
,
) , (
) , (
) , (
) , (
) , (
) , (
-
o
o
=
-
o
o
=
x
x
v v
x v
x v
x v
x a
x
,
`
,
, ,
, ,
,
`
,
, ,
, ,
,
V

(12.14)
The spatial velocity gradient ( l = v
x
,
,
V ) can be split into a symmetric and an antisymmetric
part as follows W D- l , where D is the rate-of deformation tensor (symmetric tensor)
and W is the spin tensor (antisymmetric tensor). Then, the acceleration can also be
expressed as:
v v
x v
v
x v
v v
x v
x v x a
x
, ,
, ,
,
, ,
, ,
, ,
,
`
, , ,
,


- -
o
o

-
o
o

-
o
o

W D
t
t
t
t
t
t
t t
) , (
) , (
) , (
) , ( ) , (
l
V
(12.15)
Then, because of the antisymmetric tensor property, the equation v w v
, , ,
r W holds,
where w
,
is the axial vector associated with W. In addition, ( ) v v w
x
, , , ,
,
r V ) ( 2 rot
holds, where
,
is the vorticity vector, ( ) v v v
x
, , ,
,
r r V
2
1
W . Next, the term v
,
D can be
represented as follows:
( ) v v v v
x
, , , ,
,
- ) (
2
1
sym T
V l l D ( )
j i j j i j ij
v v v v
, ,
2
1
- D (12.16)
Then, the acceleration can still be represented as follows:
( )
v v v
x v
v v v v
x v
v v
x v
x a
x
x x
, , , ,
, ,
, , , ,
, ,
, ,
, ,
, ,
,
, ,
r - -
o
o

r r - -
o
o
- -
o
o

W D
2
1
) (
) , (
2
1
) (
) , ( ) , (
) , (
sym
sym
t
t
t
t
t
t
t
V
V V

(12.17)
Finally, remember in Chapter 2 it was shown that the following equation is valid:
( ) v v
x v
v v
x v
x a
x x x
, ,
, ,
, ,
, ,
, ,
, , ,
r r - -
o
o
r - -
o
o
V V V ) (
2
1 ) , (
) (
2
1 ) , (
) , (
2 2
v
t
t
v
t
t
t rot (12.18)
12.5.1 Incompressible Fluids
As we saw in Chapter 2, an incompressible medium is characterized by isochoric motion, it
then follows that 0
kk
D v
x
,
,
V is satisfied for any incompressible fluid. Moreover,
taking into account the mass continuity equation, 0 ) ( - v
x
,
,
V j
j
Dt
D
, we can conclude that
for any incompressible fluid the following is valid:
0
0 0 ) ( j j
j

Dt
D
D Tr v
x
,
,
V Incompressible fluids (12.19)
12 INTRODUCTION TO FLUIDS

641
12.5.2 Irrotational Flow
A flow is said to be irrotational when the spin tensor vanishes at any point in the fluid:
( ) [ ] 0 W -
|
|
.
|

\
|
o
o
-
o
o

T
ij
i
j
j
i
ij
x
v
x
v
v v
x x
, ,
, ,
V V
2
1
;
2
1
0 W
0 0 W
,
,
,
r v
x
V

Irrotational flow (12.20)
An incompressible irrotational flow is characterized by:
0
,
, ,
, ,
r v v
x x
V V and 0 Incompressible irrotational flow (12.21)
When the flow is irrotational the acceleration given in (12.17) becomes:
v v
x v
v v v
x v
x a
x
x
, ,
, ,
_
, , , ,
, ,
, ,
,
,
,

-
o
o

r - -
o
o

) (
) , (
2
1
) (
) , (
) , (
sym
sym
t
t
t
t
t
V
V
0


( )
sym
k i k
i
i
v v
t
t v
t a
,
) , (
) , ( -
o
o

x
x
,
,

(12.22)
Note that ( )
k i k k i k i k k
v v v v v v
, , ,
- holds and due to the symmetry of
k i i k
v v
, ,
, the
following is valid ( ) ( )
i k k k i k
v v v v
, ,
2 , i.e. ) (
2
1
) (
2
1
2
v
x x
v v v v
, ,
, , , ,
V V D l , with which
the acceleration for an irrotational flow becomes:
) (
2
1 ) , (
) (
) , (
) , (
2
v
t
t
t
t
t
sym
x x
x v
v v
x v
x a
, ,
, ,
, ,
, ,
, ,
V V -
o
o
-
o
o

Acceleration for an
irrotational flow
(12.23)
12.5.3 Steady Flow
A steady flow, (see Chapter 2), is characterized by:
0
,
,

o
o
t
v
Steady flow (12.24)
Remember that the material time derivative of the velocity is given by:
v v
x v
v
v
x
, ,
, ,
`
,
,
,
-
o
o
= V
t
t
Dt
D ) , (

k k i
i
i
i
v v
t
t v
v
Dt
Dv
,
) , (
-
o
o
=
x
,
` (12.25)
and in steady flow the rate of change of the velocity becomes
k k i i
v v v
,
; `
, ,
`
,
,
v v v
x
V Rate of change of velocity for steady flow (12.26)

Problem 12.1: Demonstrate whether the following statements are true or false:

a) If the velocity field is steady, then the acceleration field is also;
b) If the velocity field is uniform, the acceleration field is always equal to zero;
c) If the velocity field is steady and the medium is incompressible, the acceleration is always
zero.
Solution:
NOTES ON CONTINUUM MECHANICS

642
a) In a steady velocity field we have 0
,
, ,

o
o
t
t) , ( x v
whereby the acceleration field becomes:
_
, , , , , , , ,
, ,
`
, ,
_
,
`
, ,
time of t Independen
) ( ) ( ) ( ) (
) (
) , (
, ,
0
x v x v x v x v
x v
v a
x
x x
-
o
o

-
o
o

V V
t
v v v v
t
t v
v a
k k i k k i
i
i i
i

Then, assumption (a) is TRUE.
b) A uniform velocity field implies that ) ( ) , ( t t v x v
, , ,
, whereby:
t
t
t t
t
t
o
o
-
o
o

) , (
) , ( ) , (
) , ( x v
x v x v
x v
v a
x
, ,
, ,
_
, ,
, ,
`
, ,
,
,
0
V
Then, assumption (b) is FALSE.
c) A steady velocity field implies that ) ( ) , ( x v x v
, , , ,
t and an incompressible medium means
that 0 ) , ( t x v
x
, ,
,
V , so, we can conclude that:
) ( ) ( ) ( ) (
) (
x v x v x v x v
x v
v a
x x
, , , , , , , ,
, ,
`
, ,
, ,
-
o
o
V V
t

Then, assumption (c) is FALSE.

12.6 Newtonian Fluids
The viscous stress tensor ( ) is associated with the internal energy dissipation brought
about by the viscosity. Remember that in Chapter 6 in Subsection 6.4 in viscoelastic
materials the dynamic part of the stress tensor is either a function of F
`
or a function of D
(rate-of-deformation tensor), so:


Fluid) Newtonian - (Non hip relations nonlinear a If
Fluid) (Newtonian lationship linear re a If
) (D
(12.27)
where D is the symmetric part of the spatial velocity gradient (see Chapter 2), whose
components are given by:
|
|
.
|

\
|
o
o
-
o
o

i
j
j
i
ij
x
v
x
v
2
1
D (12.28)
The equation in (12.27) is, general speaking, nonlinear, which is characteristic of the
Stokesian (non-Newtonian) fluid. Blood, paints and sauces are all examples of this.
When the relationship in (12.27) is linear we have what we term Newtonian fluids which are
described in the following format:
kl ijkl ij
D K t ) (D : K
Stress constitutive equation for
Newtonian fluid
(12.29)
where K is the tensor containing the viscosity coefficients.
Additionally, the Cauchy stress tensor is represented by:
- - 1 p

ij ji ij
p t - - o c
(12.30)
12 INTRODUCTION TO FLUIDS

643
Then, to directly obtain an expression for , we can make an analogy with the stress
constitutive equation for isotropic solid materials, (see Chapter 7), in which:
Isotropic solids Fluids
ij kk ij ij ij kk ij ij
kl ijkl ij kl ijkl ij
D D
D
* *
2 2 j c A j Ac - t r - r o
t r o K C


(12.31)
where
*
A is the viscous dilatational coefficient and
*
j is the viscous tangential coefficient
and, generally speaking, these variables are associated with other thermodynamics variables,
i.e.:
) , ( ; ) , (
* * * *
T T j j j j A A (12.32)
Now, by substituting the viscous stress tensor given in (12.31) into the equation in (12.30)
we obtain:
D 1 D 1
* *
* *
2 ) (
2
j A
j c A c
- - -
- - - o
Tr
D D
p
p
ij ij kk ij ij Navier-Poisson law
(Newtonian fluid)
(12.33)
These equations provide us with the Navier-Poisson law of a Newtonian fluid.
In incompressible fluids we have 0 ) ( D Tr v
x
,
,
V , (see equation (12.19)), whereby the
Navier-Poisson law becomes:
D 1
*
2j - - p
Navier-Poisson law
(Incompressible Newtonian fluid)
(12.34)
Then, by multiplying the equation in (12.33) by
ij
c we obtain:
kk kk
kk kk kk
kk ii kk kk kk
ij ij kk ij ij ij ij ij ij
p
p
p
p
D
D D
D D
D D

3
2
3
1
2 3 3
2
2
* *
* *
* *
* *
|
.
|

\
|
- - - o
- - - o
- - - o
- - - o
j A
j A
j c A c
c j c c A c c c


- - o
kk kk
k p D
*
3
1

kk
k p p D
*
-

(12.35)
which thus defines the bulk viscosity coefficient (also called the volume or second viscosity)
(
*
x ) as:
3
2
*
* *
j
A - x
(12.36)
It may be interesting to express the Navier-Poisson equations in terms of the deviatoric
parts. To do so, we split the Cauchy stress tensor and the rate-of-deformation tensor D
into a deviatoric and spherical part:

-
-

o
- o o
-
ij
kk dev
ij ij
dev
ij
kk dev
ij ij
dev
c c
3
3
) (
;
3
3
) (
D
D D
Tr Tr
1
D
D D 1



(12.37)
which, substituting into the constitutive equation in (12.33), yields:
NOTES ON CONTINUUM MECHANICS

644

_
_
change shape
to Related
change volume to Related
dev
ij ij kk ij
ij
kk dev
ij kk ij ij ij
kk dev
ij
ij kk ij ij ij
p
p
p
p
D D
D
D D
D D
*
*
*
* *
* *
2
3
2
3
2
3
2
j c
j
A c
j c A c c
j c A c
-
|
|
.
|

\
|
- - -
|
.
|

\
|
o - - - -
o
- o
- - - o
-

(12.38)
the result of which is:
( )
( )
dev
ij ij kk ij
dev
ij
dev dev
p p
p p
D D
Tr
* *
* *
2
2 ) (
j c c
j
- x - - o
- x - - D 1 D 1
(12.39)
The above can be decomposed into two sets of equations. Then, we can consider the
equations in (12.35) in which ( )
kk
p p D
*
x - - holds, which, substituted into the equation
in (12.39), yields:
dev
ij
dev
ij
dev dev
D
* *
2 ; 2 j j o D (12.40)
Then, the equations in (12.39) can be replaced with the following set of equations:

x -

) (
2
*
*
D
D
Tr p p
dev dev
j
(12.41)
Now, let us remember that:
kk
D Tr Tr Tr - ) ( ) ( ) ( D W D l (12.42)
where l is the spatial velocity gradient, and the trace of D can be expressed in terms of
velocity divergence by:
v
x
,
,

o
o
-
o
o
-
o
o
V
3
3
2
2
1
1
x
v
x
v
x
v
kk
D
(12.43)
and by considering the mass continuity equation, (see Chapter 5), the following remains
valid:
kk
Dt
D
Dt
D
D - -
j
j
j
j 1
0 v v
x x
, ,
, ,
V V
(12.44)
Then, by substituting (12.44) into the equation ) (
*
0
D Tr x - p p p we obtain:
Dt
D
p p
j
j
1
*
0
x -
(12.45)
The above equation indicates that the relationship p p
0
will only be fulfilled when:
1) The rate-of-deformation tensor trace is equal to zero (incompressible fluid, 0 v
x
,
V ):
0
kk
D i.e. 0
Dt
Dj
(Incompressible fluid) (12.46)
2) The bulk viscosity coefficient is equal to zero:
0
*
x
(Stokes condition) (12.47)
12 INTRODUCTION TO FLUIDS

645
The latter condition is known as the Stokes condition.
Then, in incompressible cases, ( p p p p x - ) (
*
D Tr ), the set of equations in (12.41)
becomes:
dev dev
D
*
2j
Constitutive equations for
incompressible Newtonian fluid
(12.48)
12.6.1 The Stokes Condition
The Stokes condition is met when:
0
3
2
*
* *
- / x
j
The Stokes condition (12.49)
This condition ensures us that the pressure p is defined as the average of the normal
stresses, i.e.:

p p p p
m ii
m
- o x - - o
o
) (
3
1
*
D Tr
(12.50)
12.7 Stress, Dissipated and Recoverable Powers
By considering the equation in (12.33),
ij kk ij ij ij
p D D
* *
2j c A c - - - o , the mechanical
(stress) power ( D : ) can be rewritten as follows:
[ ] D D D D
D D D 1 D D 1 D
:
: : : :
* 2 *
* *
2 ) ( ) (
2 ) (
j A
j A
- - -
- - -
Tr Tr
Tr
p
p
ij ij ii kk ii
ij ij ij kk ij ij ij ij ij
p
p
D D D D D
D D D D D D
* *
* *
2
2
j A
j c A c
- - -
- - - o

(12.51)
Then, by splitting the rate-of-deformation tensor into a deviatoric and spherical part,
(
ij kk
dev
ij ij
c D D D
3
1
- ), and by substituting them into the equation in (12.51) we can obtain:
[ ]
[ ]
[ ]
[ ]
dev dev
dev dev
dev dev
p
p
p
D D D D
D D
D
D D
1
D
D 1
D
D D D D
:
:
: :
* 2
*
*
2
* 2 *
* 2 *
2 ) (
3
2
) (
3
) (
2 ) ( ) (
3
) (
3
) (
2 ) ( ) (
j
j
A
j A
j A
-
|
|
.
|

\
|
- - -
|
|
.
|

\
|
- - - -
|
.
|

\
|
- |
.
|

\
|
- - - -
Tr Tr
Tr
Tr Tr
Tr Tr
Tr Tr

(12.52)
or in indicial notation:
|
|
.
|

\
|
- - - -
|
|
.
|

\
|
- |
.
|

\
|
- - - - o
dev
ij
dev
ij
pp kk
ii kk ii
ij
pp
dev
ij ij
kk dev
ij ii kk ii ij ij
p
p
D D
D D
D D D
D
D
D
D D D D D
3
2
3 3
2
* *
* *
j A
c c j A

(12.53)
NOTES ON CONTINUUM MECHANICS

646
where the deviatoric tensor trace is equal to zero, 0 1 D :
dev
. Then, by restructuring the
above equation we obtain:
dev
ij
dev
ij ii kk ii
dev
ij
dev
ij ii kk ii ij ij
p
p
D D D D D
D D D D D D
* *
*
*
*
2
2
3
2
j
j
j
A
- x - -
-
|
|
.
|

\
|
- - - o

(12.54)
Next, the stress power can be expressed as follows:
[ ]
_
D
dev dev
W
p
2
* 2 *
2 ) ( ) ( D D D D D : : j - x - - Tr Tr
Stress power (12.55)
The term ( ) (D Tr p - ) is related to elastic energy, so, it is recoverable and because of this we
have the definition:
) (D Tr p -
Recoverable power (12.56)
which does not contribute to internal entropy generation in the system. We can also define
the dissipated power per unit volume (
D
W 2 ), which is associated with internal energy
dissipation:
[ ]

- x
- x
dev dev
D
dev
ij
dev
ij ii kk D
W
W
D D D :
* 2 *
* *
2 ) ( 2
2 2
j
j
Tr
D D D D
Dissipated power (12.57)
NOTE: As we can verify by looking at the dissipated power, all dissipated energy is
brought about by viscosity when the fluid is in relative motion between particles, (see
Figure 12.5).








Figure 12.5: Viscous fluid in motion.
Now, by considering the second law of thermodynamics (nonnegative dissipation), we can
conclude that:
( 0
*
> j ) and
3
2
0
*
* *
j
A x
(12.58)
For a fluid without viscosity, the dissipated power is zero, i.e. 0 2
D
W , and the stress
power becomes:
v
x
,
,
- - V p p ) (D D Tr :
Stress power for non-viscous fluid (12.59)
Dissipated energy
viscous fluid in motion
particle
12 INTRODUCTION TO FLUIDS

647
12.8 The Fundamental Equations for Newtonian
Fluids
The fundamental equations for Newtonian fluids
(Current configuration)
The mass continuity
equation:
0 ; 0 ) (
,
- -
i i
v
Dt
D
Dt
D
j
j
j
j
v
x
,
,
V (12.60)
The equations of motion:
i i j ij
v`
`
,
,
,
j j j - o p - b ;
,
v
x
b V (12.61)
The energy equation:
[ ] r p u
or
r u r u
D
dev dev
i i ij ij
W
j j j
j j
j j
- - - x - -
- - o - -

q D D D D
q D
,
_
`
`
,
`
,
,
x
x
V
V
2
* 2 *
,
2 ) ( ) (
1 1
;
:
:
Tr Tr
q D

(12.62)
where u is the specific internal energy, r is the heat generated by internal sources and q
,
is
the heat flux vector (non-convective).
The mass continuity equation, the equations of motion and the energy equation give us five
equations in total. The unknowns are: velocity v
,
(three components), temperature T ,
mass density j , the Cauchy stress tensor (six components), specific internal energy u ,
the heat flux vector q
,
(three components), entropy j , and pressure p , making a total of
17 unknowns.
For the problem to be well-posed 12 equations must be added to the system, as discussed
in Chapter 6, these equations are the so-called constitutive equations:
The constitutive equations
for stress:
ij kk ij ij ij
p
p
D D
Tr
* *
* *
2
2 ) (
j c A c
j A
- - - o
- - - D 1 D 1
(12.63)
for heat conduction:
T T
j ij i x
,
,
V - - K q ; , K q (Fouriers law)
(12.64)
for entropy:
) , ( T j j j
(12.65)
) , ( T p p j
The equations of state:
) , ( T u u j
(12.66)
where K is the thermal conductivity tensor, (see Chapter 10). So, the problem results in a
system of 17 equations with 17 unknowns.
For fluids in which is independent of temperature, the pressure can be expressed in term of
mass density, ) (j p p and the internal energy ) (j u u . So, the mechanical problem can
be represented by the following equations:
The fundamental equations for barotropic Newtonian fluids
The mass continuity equation 0 ) ( - v
x
,
,
V j
j
Dt
D
(12.67)
NOTES ON CONTINUUM MECHANICS

648
The equations of motion: v
x
`
,
,
,
p - b j V (12.68)
The constitutive equations
for stress: D 1 D 1
* *
2 ) ( j - / - - Tr p (12.69)
the equation of state:
) (j p p
(12.70)
which results in a system with 11 equations and 11 unknown.
12.8.1 The Navier-Stokes-Duhem Equations of Motion
The Navier-Stokes-Duhem equations of motion are a combination of the equations of
motion (12.61) and the constitutive equations (12.63). Then, by considering
ij kk ij ij ij
p D D
* *
2j c A c - - - o obtained in (12.33), the Cauchy stress tensor divergence
(
x
,
V ) can be evaluated as follows:
( )
j ij j kk ij j ij
j
kk ij j ij j ij
p
,
*
,
*
,
*
,
*
, ,
2 2 D D D D j c A j c A c - - - - o
(12.71)
In addition, by considering
i j j i ij
v v
, ,
2 - D and
k k k k k k kk
v v v
, , ,
2 2 - D , we obtain:
kj k j kk
ji j jj i ij j jj i j ij
v
v v v v
, ,
, , , , ,
2

- -
D
D
(12.72)
whereby the equation in (12.71) becomes:
( ) ( )
( ) ( )
jj i ji j ji j jj i ji j
ji j jj i ki k ji j jj i kj k ij j ij j kk ij j ij
v v v v v
v v v v v v
,
*
,
* *
, ,
*
,
*
, ,
*
,
*
, ,
*
,
*
,
*
,
*
,
2
j j A j A
j A j c A j c A
- - - -
- - - - - o D D
(12.73)
Then, by substituting the equation in (12.73) into the equations of motion
(
i i j ij
v` j j - o b
,
), (see equation (12.61)), we obtain:
v v v
x x x x
, ,
,
`
,
`
, , , ,
2 * * *
,
*
,
* *
,
) ( ) (
) (
V V V V j j A j j
j j A j j
- - - -
- - - -
p
v v p v
jj i ji j i i i
b
b Navier-Stokes-Duhem
equations of motion
(12.74)
which are the Navier-Stokes-Duhem equations of motion. The terms on the right of the
equation in (12.74) represents force terms, b
,
j represents force per unit volume and
( p
x
,
V - ) is the pressure gradient and represents force per unit volume brought about by
thermodynamic pressure. Finally, the remaining terms represent the viscous force per unit
volume:
v v f
x x x
, ,
,
, , ,
2 * * *
) ( ) ( V V V j j A - -
vis
The viscous forces (12.75)
12.8.1.1 Alternative Form of the Fundamental Equations for Newtonian
Fluids
With the above, the fundamental equations for Newtonian fluids can also be expressed as
follows:


12 INTRODUCTION TO FLUIDS

649
The fundamental equations for Newtonian fluids
The mass continuity
equation:
0 ; 0 ) (
,
- -
i i
v
Dt
D
Dt
D
j
j
j
j
v
x
,
,
V (12.76)
The Navier-Stokes-Duhem
equations:
v v v
x x x x
, ,
,
`
,
`
, , , ,
2 * * *
,
*
,
* *
,
) ( ) (
) (
V V V V j j A j j
j j A j j
- - - -
- - - -
p
v v p v
jj i ji j i i i
b
b

(12.77)
The energy equation:
[ ] r p u
D
dev dev
W
j j j - - - x - - q D D D D
,
_
`
,
x
V
2
* 2 *
2 ) ( ) ( : Tr Tr
(12.78)
The mass continuity equation, the Navier-Stokes-Duhem equations of motion and the
energy equation give us five of these in total. The unknowns are: velocity v
,
(three
components), temperature T , mass density j , specific internal energy u , the heat flux
vector q
,
(three components), entropy j , and pressure p , which makes a total of 11
unknowns.
Then, for the problem to be well-posed, six equations must be added to the system,
namely:
The constitutive equations
for heat conduction:
T
T
j ij i
x
,
,
V -
-
K q
, K q
(Fouriers law)
(12.79)
for entropy:
) , ( T j j j
(12.80)
) , ( T p p j
The equations of state:
) , ( T u u j
(12.81)
12.8.1.2 The Fundamental Equations for Incompressible Newtonian
Fluid
With incompressible fluids 0 ) ( v
x
,
,
V holds, then:
The fundamental equations for incompressible Newtonian fluids
The mass continuity
equation:
0
0 j j
j

Dt
D
(12.82)
The Navier-Stokes-Duhem
equations of motion:
v v
x x
,
,
`
,
`
, ,
2 *
,
*
,
V V j j j
j j j
- -
- -
p
v p v
jj i i i i
b
b

(12.83)
The energy equation:
[ ] r p u
D
dev dev
W
j j j - - - x - - q D D D D
,
_
`
,
x
V
2
* 2 *
2 ) ( ) ( : Tr Tr
(12.84)

The constitutive equations
for heat conduction:
T
T
j ij i
x
,
,
V -
-
K q
, K q
(Fouriers law)
(12.85)
for entropy:
) , ( T j j j
(12.86)
) , ( T p p j
The equations of state:
) , ( T u u j
(12.87)
NOTES ON CONTINUUM MECHANICS

650
12.8.2 The Navier-Stokes Equations of Motion
If we have the Stokes condition |
.
|

\
|
-
* *
3
2
j A , the Navier-Stokes-Duhem equations of
motion becomes the Navier-Stokes equations of motion. Then, by substituting
* * *
3
1
) ( j j - / into the equation in (12.77) we obtain:
v v v
x x x x
, ,
,
`
,
`
, , , ,
2 * *
,
*
,
*
,
) (
3
1
3
1
V V V V j j j j
j j j j
- - -
- - -
p
v v p v
jj i ji j i i i
b
b
The Navier-Stokes equations
of motion (Compressible fluid)
(12.88)
when the fluid is incompressible ( 0
,

j j
v v
x
,
,
V ) the above equation becomes:
v v
x x
,
,
`
,
`
, ,
2 *
,
*
,
V V j j j
j j j
- -
- -
p
v p v
jj i i i i
b
b
The Navier-Stokes equations
of motion (Incompressible
fluid)
(12.89)
Note that for in incompressible fluids the Navier-Stokes equations of motion (see (12.89))
and the Navier-Stokes-Duhem equations of motion (see (12.83)) coincide.
12.8.3 The Euler Equations of Motion
In a non-viscous fluid (perfect fluid) there is no viscous forces, i.e. 0
, ,

vis
f , so, the
equations of motion become:
p
p v
i i i
x
v
,
,
`
,
`
V -
-
b j j
j j
,
b
The Euler equations of motion
(Non-viscous incompressible fluid)
(12.90)
which are known as the Euler equations of motion.
Then, by considering a perfect fluid and an isothermal and adiabatic process, we have:
The fundamental equations for perfect fluid (isothermal and adiabatic process)
The mass continuity
equation:
0 ) ( - v
x
,
,
V j
j
Dt
D
(12.91)
The Navier-Stokes-Duhem
equations of motion:
p
t
t
or p
x x x
v v
x v
v
, , ,
,
, ,
, ,
,
`
,
V V V
j
j j
1 ) , (
- -
o
o
- b b
(12.92)
The energy equation: ) ( ) ( v
x
,
`
,
- - V p p u D Tr j (12.93)
which results in a total of five equations and six unknowns, namely: v
,
, p , u , j . Then, to
complete the set of equation we have to add:
The constitutive equation
The equation of state
) , ( u p p j
(12.94)
Then, in the particular case when velocity is equal to zero, the equations of motion in
(12.98) become:
12 INTRODUCTION TO FLUIDS

651
p
x
,
,
V b j i i
p
,
b j
(12.95)
which describes the hydrostatic equilibrium. Then, if we assume there is a barotropic
condition, ) ( p j j , it is possible to define the pressure function as:
dp p P
p
p
}

0
1
) (
j

(12.96)
12.8.3.1 Non-Viscous and Incompressible Fluids
In incompressible fluids, 0 ) ( v
x
,
,
V , mass density is constant and is no longer an
unknown. In addition to this, if we consider there is an isothermal and adiabatic process,
we have:
The fundamental equations for incompressible perfect fluids
The mass continuity equation: 0 v
x
,
,
V (12.97)
The Euler equations of motion: p p v
i i i x
v
,
,
`
,
` V - - b j j j j ;
,
b (12.98)
which results in four unknowns and four equations. Here, we can verify that the Navier-
Stokes-Duhem equations of motion become the Euler equations of motion (12.98). In
addition to this, we can consider the body force field to be conservative, so, we can express
it by means of the potential , as follows ,
x
,
,
V - b whereby the Euler equations of
motion become:
p
p
x x
x
v
, ,
,
,
`
,
V V
V
- -
-
, j
j j b

i i
i i i
p
p v
, ,
,
- -
-
j,
j j b `

(12.99)
Then, by using the material time derivative, we can express the Eulerian velocity
components as follows:
j j i
i
j
j
i i i
i
v v
t
v
v
x
v
t
v
Dt
Dv
v
,
-
o
o

o
o
-
o
o
= `
(12.100)
Note, the resulting components of the operation
j j i
v v
,
are the same as those of v v
x
, ,
,
) (V ,
(see Chapter 1) and it was shown that the following holds:
) (
2
1
) (
2
1
) ( ) (
2
1
) ( ) (
2 2
v v
x x x x x x
v v v v v v v v v
, , , , , ,
, , , , , , , , , ,
V V V V V V - r - r r - r r (12.101)
where v
,
v , and ( ) v v
x
, , ,
,
r = = V ) ( rot . Then, the velocity field can be expressed as
follows:
) (
2
1

2
v
t
x
v
v
v
,
, ,
,
`
,
V - r -
o
o
(12.102)
Then, returning to the equation in (12.99) we can affirm that:
0

,
, ,
,
, ,
,
, , ,
, , ,
- - - r -
o
o

- - - r -
o
o
p v
t
p v
t
x x x
x x x
v
v
v
v
V V V
V V V
j
,
, j j j j
1
) (
2
1

) (
2
1

2
2

(12.103)
NOTES ON CONTINUUM MECHANICS

652
Now, if we consider that the mass density field is homogenous, i.e. it does not very with the
position vector, the following holds:
_
, , ,
0
1
) (
1

|
|
.
|

\
|
-
|
|
.
|

\
|
j j j
x x x
V V V p p
p

(12.104)
whereby the equation in (12.103) can be rewritten as follows:
0
,
, ,
,
, ,

|
|
.
|

\
|
- - - r r -
o
o
j
,
p
v
t
2
2
1
) (
x x
v v
v
V V
(12.105)
If we then consider the velocity field to be conservative, that means that the curl of the
field is zero, 0
,
, , ,
,
r v v
x
rot V . Furthermore, a conservative field can be represented
by a potential, thus:
|
.
|

\
|
o
o
o
o
|
.
|

\
|
o
o
o
o

o
o

o
o
=
t t t
c c c
c
x x
v
x
v
x
, ,
,
,
,
,
) ( V (12.106)
whereupon the equation in (12.105) becomes:
i i
al Irrotacion
p
v
t
p
v
t
0 0
|
|
.
|

\
|
- - -
o
o

|
|
.
|

\
|
- - - r r -
o
o

j
,
c
j
c
2 2
2
1
2
1
) (
x x x
v v
v
, , ,
,
_
,
,
,
V V V
0

(12.107)
Thus, we can conclude that:
) (
2
1
2
t C
p
v
t
- - -
o
o
j
,
c

(12.108)
where ) (t C is a constant and time-dependent.
12.8.3.2 Bernoullis Equation
If the velocity field is stationary, 0
,
,

o
o
t
v
, and if the velocity field is irrotational,:
0
,
, ,
,
r v v
x
rot V , the equation in (12.107) becomes:
constant
v p v p
i
- -
|
|
.
|

\
|
- -
2 2
2 2
j
,
j
, 0
x
,
V (12.109)
Then, by considering the potential gh , , where g is the acceleration of gravity, and h is
the piezometric height, the equation in (12.109) becomes:
constant
v p
gh - -
2
2
j
Bernoullis equation (12.110)
Let us now check the SI units: [ ]
2
2 3
2
2
2 s
m
kg
J
kg
Nm
kg
m
m
N p v
gh
(

j
, which is the
unit of specific energy, i.e. energy per unit mass, (see Figure 12.6).



12 INTRODUCTION TO FLUIDS

653













Figure 12.6: Energy vs. the Bernoulli theorem.
12.8.4 The Equation of Vorticity
Next, we will establish the equation of vorticity. To do so, we will start from the Navier-
Stokes-Duhem equations of motion given in (12.74):
v v v
x x x x
, ,
,
`
,
`
, , , ,
2 * * *
,
*
,
* *
,
) ( ) (
) (
V V V V j j A j j
j j A j j
- - - -
- - - -
p
v v p v
jj i ji j i i i
b
b

(12.111)
Then, by taking into account the expression of velocity given in (12.102) the above
equation becomes:
0 b
b
b
,
, ,
,
, ,
,
, ,
,
, ,
,
, ,
,
`
,
`
, , , , ,
, , , , ,
, , , ,
-
-
- - - - r -
o
o

- - - - |
.
|

\
|
- r -
o
o

- - - -
- - - -

v v v
v
v v v
v
v v v
x x x x x
x x x x x
x x x x
2
* * *
2
2 * * * 2
2 * * *
,
*
,
* *
,
) (
) ( 1
) (
2
1

) ( ) ( ) (
2
1

) ( ) (
) (
V V V V V
V V V V V
V V V V
j
j
j
j A
j
j j A j j
j j A j j
j j A j j
p v
t
p v
t
p
v v p v
jj i ji j i i i
b
(12.112)
Next, we can take the curl of the above equation, the result of which is:
0 b
,
, ,
,
, ,
,
, , , , , ,

(

-
-
- - - - r -
o
o
r v v v
v
x x x x x x
2
* * *
2
) (
) ( 1
) (
2
1
V V V V V V
j
j
j
j A
j
p v
t

(12.113)
In Chapter 1 we proved that the following holds:
[ ] 0
,
, ,
r ) (
2
v
x x
V V , [ ] 0
,
, ,
r p
x x
V V , [ ] 0
,
,
, , ,
r ) ( v
x x x
V V V ;

[ ] [ ]
[ ] [ ]
, , , , , , , ,
, , , , , , , ,
, , , ,
, , , , , , , ,


- - r r
r - r - r r r r
v v v v
v v v v v v v v
x x x x
x x x x x x x x
V V V V
V V V V V V V V
) (
) ( ) ( ) )( ( ) (
;
[ ] [ ] [ ]
, , , ,
, , , , , , , ,
) (
2 2 2
x x x x x x x x
v v v V V V V V V V V r r r r - r ;

A
p
j


A
gh

A
v
2
2

A
B
constant energy
h

B
p
j


B
v
2
2


B
gh
energy at A
energy at B

B A
v p
gh
v p
gh
|
|
.
|

\
|
- -
|
|
.
|

\
|
- -
2 2
2 2
j j

NOTES ON CONTINUUM MECHANICS

654
[ ]
t t t o
o
r
o
o

o
o
r

,
,
,
, ,
v
v
x x
V V ;
If we consider that the field b
,
is conservative, and the curl of any conservative
vector field is zero, we have 0 b
, ,
,
r
x
V .
Then, given the above, we can conclude that:
0

,
, , , , , , ,
,
, , , ,
- - - -
o
o
) ( ) (
2
*
x x x x
v v v V V V V
j
j
t
(12.114)
Note that the following relationships hold:
j j i j j i j j i j j i j j i j j i j j i
i j i i j i j i i i j i j i i i j i
v v v v v v v
v v v v v v
), ( , ), ( , , , ), (
, ), ( , , , ), (
c c - c c c - c c
c - c c c - c c
(12.115)
which is the same in tensorial notation:
[ ]
[ ] [ ]

, , , , , , , ,
, , , , , ,
, , , ,
, , ,
-
-


v v v v
v v v
x x x x
x x x
V V V V
V V V
) ( ) (
) ( ) (
(12.116)
where we have applied the definition that the divergence of the curl of a vector is zero, i.e.
0 ) ( r v
x x x
, ,
, , ,
V V V . Then, by considering (12.116), the equation (12.114) becomes:
[ ] [ ]
[ ] [ ]
[ ] 0

,
, , , , ,
,
,
, , , , ,
,
,
, , , , , , , , ,
,
,
, , , , , , ,
,
, ,
, , ,
, , , , ,
, , , ,
- - -
o
o

- - -
o
o

- - - - -
o
o

- - - -
o
o






) ( ) (
) ( ) (
2
*
2
*
2
*
2
*
x x
x x x
x x x x x
x x x x
v v
v v
v v v v
v v v
V V
V V V
V V V V V
V V V V
j
j
j
j
j
j
j
j
t
t
t
t

(12.117)

[ ] 0

,
, , ,
,
, ,
- -
o
o
) ( 2
2
*
x x
v V V
j
j
skew
t
The equation of vorticity (12.118)
Note that to obtain the equation in (12.118), the only simplification made was for the fluid
to be Newtonian, and the b
,
-field to be conservative.


Problem 12.2: Prove that the Cauchy deviatoric stress tensor
dev
is equal to
dev
, where
ij ij ij
p t - - o c .
Solution
If we consider that
kk kk
p t - - o 3 we can obtain:
( )
dev
ij ij
kk
ij ij
kk
ij ij ij
kk
ij
dev
ij
p
p t
t
- t
t - -
- t - - o
o
- o o c c c
3 3
3
3



12 INTRODUCTION TO FLUIDS

655
Problem 12.3: Let us consider a body immersed in a Newtonian fluid. Find the total
traction force E
,
acting on the closed surface S which delimits the volume V . Consider
that the bulk viscosity coefficient to be zero.














Solution: We know that the following holds:
dS dE
i i
)

(n
t
The total traction force is given by the following integral:
dV dS dS E
V
j ij
S
j ij
S
i i
} } }
o o
,
)

(

n t
n

where we have used the relationship
)

n
i j ij
t n o .
Then, if the bulk viscosity coefficient is zero, we have:
* * *
3
2
0 j A - x (Stokes
condition).
Next, by considering the stress constitutive equation for Newtonian fluids, we obtain:
_
dev
ij
ij
kk
ij ij ij kk ij ij ij kk ij ij ij
p p p
D
D
D D D D D |
.
|

\
|
- - - - - - - - - o c j c j c j c j c A c
3
2 2
3
2
2
* * * * *

dev
ij ij ij
p D
*
2j c - - o
Then
dS p E
S
j
dev
ij ij i
}
- - n D

) 2 (
*
j c
and by applying the Gauss theorem, we obtain:
( ) ( ) dV p dV p dV p E
V
dev
j ij i
V
dev
j ij ij j
V
j
dev
ij ij i
} } }
- - - - - - ) 2 ( 2 2
,
*
, ,
*
,
,
*
D D D j j c j c
where we have considered that
j j
0
*
,
j , i.e.
*
j is a homogenous scalar field (homogenous
material). Then, the above equation in tensorial notation becomes:
[ ]
}
- -
V
dev
dV p ) ( 2
*
D
x x
E
, ,
,
V V j (12.119)





)

(n
t
,

n


V
NOTES ON CONTINUUM MECHANICS

656
Problem 12.4: Let us consider a fluid at rest which has the mass density
f
j . Prove
Archimedes Principle: Any body immersed in a fluid at rest experiences an upward buoyant force
equal to the weight of the volume fluid displaced by the body.
If mass density in the body is equal to
s
j and the body force (per unit mass) is given by
3 i i
gc - b , obtain the resultant force and acceleration acting on the body.
Solution:
In Problem 12.3 we showed that [ ]
}
- - dV p
dev
) ( 2
*
D
x x
E
, ,
,
V V j . If the fluid is at rest
0 D
dev
holds, and the thermodynamic pressure is equal to the hydrostatic pressure, i.e.
0
p p whereby we have:
[ ]
}
-
V
dV p
0 x
E
,
,
V (12.120)


















The weight of the fluid volume displaced by the body is given by:
}

V
f f
dV b
,
,
j W (12.121)
Then, by applying the equilibrium equations we have:
b
b 1
b
0 b
,
,
,
, ,
,
,
,
,
f
f
f
f
p
p
j
j
j
j

- -
-
-

0
0
) (
x
x
x
x
V
V
V
V

i
f
i
i
f
j ij
i
f
j ij
i i
f
j ij
p
p
b
b
b
0 b
j
j c
j
j

- -
- o
- o
, 0
0
,
,
), (

(12.122)
Next, by considering both (12.120) and (12.121), we can conclude that:
E W
x
,
,
,
,
-
} }
V V
f f
dV p dV
0
V b j (12.123)
which thereby proves Archimedes principle.
Now, the body weight, with mass density
s
j , can be obtained as follows:

s
W
,

n 0
p
E
,

V -volume

1
x

2
x

3
x
12 INTRODUCTION TO FLUIDS

657
}

V
s s
dV b
,
,
j W
and the resultant force acting on the body is given by:
} } }
- - - -
V
f s
V
s
V
f
s
dV dV dV b b b
, , ,
, , ,
) ( j j j j W E R
whose components are:
(
(
(
(
(
(

-
- - -
}
} }
V
s f
V
i
f s
V
i
f s
i
dV g
dV g dV R
) (
0
0
) ( ) (
3
j j
c j j j j b
thereby verifying that: if the body has a mass density lower than fluid mass density, e.g. if
the body is a gas, the body rises, i.e. 0
, ,
> > R
s f
j j , and if not the body falls. Moreover,
if we consider that a R
,
,
s
m , where
s
m is the total mass of the submerged body, we can
obtain the acceleration of the body as:
s
s f
s
V
s
s
s f
s
V
s
s
s f
s
V
s f
s
g
m
dV
g
m
dV g
m
dV g
m
R
a
j
j j
j
j
j j
j
j
j j j j
) (
) (
) ( ) (
3
3
-

-

} } }

NOTE: It is interesting to note that if the medium ( f ) is such that 0
f
j we have
g a -
3
, i.e. the acceleration is independent of the mass. Here we have clearly seen, as did
Galileo, by means of a simple experiment, that a freely falling body was independent of the
mass. For example, on the moon where we can consider that the mass density of air is
equal to zero, two bodies with different masses in free fall, e.g. a feather and a hammer, will
have the same acceleration and will reach the moon surface at the same time.

Problem 12.5: Obtain the one-dimensional mass continuity equation for a non-viscous
incompressible fluid flow through a pipeline. Then, consider the volume V between the
two arbitrary cross sections A and B .











Solution:

In an incompressible medium, the mass density is independent of time 0 = j
j
`
Dt
D
.
Moreover, here we can consider the mass continuity equation 0 ) (
,
- v
x
,
,
` V j j j
k k
v ,
where 0 v
x
,
,
V or 0
,

k k
v holds. Then, by considering the entire volume we obtain:

) (

A
n

) (

B
n
B
A
V
NOTES ON CONTINUUM MECHANICS

658
} }

V
k k
V
dV v dV 0 ; 0
,
v
x
,
,
V
and by applying the divergence theorem (Gauss theorem) we obtain:
} }

S
k k
S
dS v dS 0

; 0

n n v
,

Thus:
} }
-
B A
S
B B
S
A A
dS dS 0

n n v v
, ,

Next, the velocities at the cross sections
A
S and
B
S can be expressed as follows:
B B B A A A
v v n n

;

- v v
, ,

and by substituting the velocity into the integral we can obtain:
} }
- -
B A
S
B B B
S
A A A
dS v dS v 0

n n n n
B B A A
S v S v


Problem 12.6: Determine the conditions needed for mean normal pressure
m
kk
p o -
o
-
3
to be equal to thermodynamic pressure p for a Newtonian fluid.
Solution:
It was deduced that:

kk
p
kk
ii kk
dev
ij
dev
ij
p p k p D D D
* * *
3
;
3
1
; 2 x - - -
o
- - o o
-
j
Thus, p p holds when the following is satisfied:
* * *
3
2
0 ) (
0
0 j A -

x or or
ii
D Tr
D


















ALEXANDER, H. (1968). A constitutive relation for rubber-like material. Int. J. Eng. Sci., Vol.
6, pp. 549-563.
ANTMAN, S.S. (1995). Nonlinear Problems of Elasticity. Springer-Verlag, New York.
ARRUDA, E.M. & BOYCE, M.C. (1993). A three-dimensional constitutive model for the large
stretch behavior of rubber elastic materials. J. Mech. Phys. Solids, Vol. 41, N. 2, pp. 389-
412.
ASARO, R.J. & LUBARDA, V.A. (2006). Mechanics of solids and materials. Cambridge University
Press, New York, USA.
BAAR, Y. & WEICHERT (2000). Nonlinear continuum mechanics of solids: fundamental concepts and
perspectives. Springer Verlag. Berlin.
BATRA, R. C. (2006). Elements of Continuum Mechanics. John Wiley & Sons Ltd., United
Kingdom.
BAANT, Z.P. & KIM, S.-S. (1979). Plastic-Fracturing Theory for Concrete. J. Engng. mech.
Div. ASCE, 105, 407.
BAANT, Z.P. & PLANAS, J. (1997). Fracture and size effect in concrete and other quasibrittle
materials, CRC Press LLC, USA.
BIGONI, D. (2000). Bifurcation and instability of non-associative elastoplastic solids. CISM
Lecture Notes. Material Instabilities in elastic an plastic Solids, H. Petryk (IPPT, Warsaw)
Coordinator.
BIOT, M.A. (1954). Theory of stress-strain relations in anisotropic viscoelasticity and
relaxation phenomena. Jour. Appl. Phys., pp. 1385-1391.
BIOT, M.A. (1955). Variational principles in irreversible thermodynamics with application to
viscoelasticity. The Physical Reviwe 97, pp. 1463-1469.
BIOT, M.A. (1956). Thermoelasticity and irreversible thermodynamics. Jour. Appl. Phys., pp.
240-253.
BLATZ, P.J. & KO, W.L. (1962). Application of finite elasticity theory to the deformation of
rubbery material. Transactions of the Society of Rheology, Vol. VI, pp. 223-251.
BONET, J. & WOOD, R.D. (1997). Nonlinear continuum mechanics for finite element
analysis. Cambridge University Press, USA.
BUECHE, F. (1960). Molecular basis for the Mullins effect. J. Appl. Poly. Sci., 4(10), 107-114.

Bibliography
659 , Notes on Continuum Mechanics, Lecture Notes on Numerical
n Engineering and Sciences 4, DOI 10.1007/978-94-007-5986-2, Methods i
International Center for Numerical Methods in Engineering (CIMNE), 2013
E.W.V. Chaves
NOTES ON CONTINUUM MECHANICS

660
BUECHE, F. (1961). Mullins effect and rubber-filler interaction. J. Appl. Poly. Sci., 5(15), 271-
281.
CAROL, I. & WILLAM, K. (1997) Application of analytical solutions in elasto-plasticity to
locaization analysis of damage models. In Owen,. D.R.J., Oate, E., and Hinton, E.
(Eds.), Computational Plasticity (COMPLAS V), pp. 714-719, Barcelona. Pineridge Press.
CAROL, I.; RIZZI, E. & WILLAM, K. (1998) On the formulation of isotropic and anisotropic
damage. Computational Modelling of Concrete Structures. de borst, Biani, Mang & Meschke
(eds.). Balkema, Rotterdam, ISBN 9054109467. pp. 183-192.
CERVERA, M. & BLANCO DAZ, E. (2001). Mecnica de Estructuras Libro 1 Resistencia de
materiales. Edicions UPC, Barcelona. Eapaa.
CHABOCHE, J.L. (1979). Le concept de contrainte effective appliqu llasticit et la
viscoplasticit en presence dun endommagement anitrope. Colloque EUROMECH
115, Grenoble Edition du CNRS.
CHADWICK, P. (1976). Continuum mechanics concise theory and problems. George Allen & Unwin
Ltd.Great Britain.
CHANDRASEKHARAIAH, D.S. & DEBNATH, L. (1994). Continuum mechanics. Academic Press,
San Diego (CA, U.S.A.).
CHAVES, E.W.V.(2007). Mecnica del Medio Continuo: Conceptos Bsicos. CIMNE (Centro
Internacional de Mtodos Numricos en la Ingeniera)-Barcelona. ISBN: 978-84-
96736-38-2.
CHAVES, E.W.V.(2009). Mecnica del Medio Continuo: Modelos Constitutivos. CIMNE (Centro
Internacional de Mtodos Numricos en la Ingeniera)-Barcelona. ISBN: 978-84-
96736-68-9.
CHEN, A. & CHEN, W.F.(1975). Constitutive relations for concrete. J. Eng. Mech.-ASCE,
101:465-481.
CHEN, W.F. & HAN, D.J.(1988). Plasticity for Structural Engineers. Springer-Verlag New Yor
Inc.
CHEN, W.F. (1982). Plasticity in reinforced concrete. McGraw-Hill, Inc. USA.
CHUNG, T.J. & KIM, J.Y. (1984). Two-dimensional, combined-mode heat transfer by
conduction, convective and radiation in emitting, Absorbing, and scattering media.
ASME Trans. J. Heat Transfer 106, pp.:448-452.
CHUNG, T.J. (1996). Applied continuum mechanics, Cambridge University Press.
COLEMAN, B.D. & GURTIN, M.E. (1967). Thermodynamics with internal sate variables. J.
Chem. Phys. 47, pp. 597-613.
COLEMAN, B.D. (1964). Thermodynamics of materials with memory. Arch. Rat. Mech. Anal.
17, pp. 1-46.
CRISFIELD, M.A. (1997). Non-Linear Finite Element Analysis of Solids and Structures, volume 1,2.
John Wiley and Sons, New York, USA.
CRISTENSEN, R. M. (1982). Theory of Viscoelasticity, second edition. Dover Publications, Inc.,
New York, USA.
DESAI, C.S. & SIRIWARDANE, H.J. (1984). Constitutive laws for engineering materials with emphasis
on geological materials. Prentice-Hall, Inc.USA.
DAZ DEL VALLE, J. (1984). Mecnica de los Medios Continuos I. Servicio de publicaciones
E.T.S. de Ingenieros de Caminos, C. y P. Santander.
BIBLIOGRAPHY

661
DOBLAR, M. & ALARCN, M. (1983). Elementos de Plasticidad. Sevicio de publicaciones de la
E.T.S.I. Industriales. Madrid.
DRAGON, A. & MRZ, Z. (1979). A Continuum Model for Plastic-Brittle Behavior of Rock
and Concrete. Int. J. Engng. Sci., 17, 121.
DRUCKER, D.C. (1950). Stress-strain relations in the plastic range A survey of the theory
and experiment. Report to the Office of Naval Research, under contact N7-onr-358,
Division of Applied Mathematics, Brown University, Providence, RI.
FARIA, R. & OLIVER, X. (1993). A rate dependent plastic-damage constitutive model for
large scale computations in concrete structures. Monograph CIMNE N17, International
Center for Numerical Methods in Engineering, Barcelona, Spain.
FELIPPA,C.A. (2002). Introduction to Finite Element Methods. Course Notes, see World Wide
Web.
FINDLEY, W.N.; LAI, J.S. & ONARAN, K.(1989). Creep and relaxation of nonlinear viscoelastic
materials: with an introduction to linear viscoelasticity. Dover Publications, Inc. NY.
FUNG, Y.C. (1965). Foundations of solids mechanics. Prentice Hall Inc., New Jersey.
FUNG, Y.C. (1977). A first course in continuum mechanics. Prentice Hall Inc., New Jersey.
GENT, A.N. (1996). A new constitutive relation for rubber. Rubber Chem. Technol., Vol. 69.
pp. 59-61.
GREEN, A.E. & NAGHDI, P.M. (1965). A general theory of an elasto-plastic continuum.
Arch. Rat. Mech. Anal., 18.
GREEN, A.E. & NAGHDI, P.M. (1971). Some remarks on elastic-plastic deformation at finite
strain. Int. J. Engng. Sci., Vol. 9, pp. 1219-1229.
GURTIN, M. E. (1996). An introduction to continuum mechanics. NY: Academic Press, Inc.
GURTIN, M.E. & FRANCIS, E.C. (1981). Simple rate-independent model for damage. J.
Spacecraft, 18(3) pages: 285-286.
HAUPT, P. (2002). Continuum mechanics and theory of materials. Springer-Verlag, Gemany.
HILL, R. (1950). The mathematical theory of plasticity. Oxford University Press.
HILL, R. (1959). Some Basic Principles in the Mechanics of Solids without a natural Time. J.
Mech. Phys. Solids, 7, 209.
HOLZAPFEL, G.A. (2000). Nonlinear solid mechanics. John Wiley & Sons Ltd. England.
IORDACHE, M.-M. (1996). Failure Analysis of Classical and Micropolar Elastoplastic Materials.
Ph.D. Dissertation, University of Colorado at Bolder.
JIRSEK, M. (1998). Finite elements with embedded cracks. LSC Internal Report 98/01,
April.
JU, J.W. (1989). Energy-based coupled elastoplastic damage models at finite strains. Journal
of Engineering Mechanics, Vol.115, No. 11, pp-2507-2525.
KACHANOV, L. M. (1986). Introduction to Continuum, Damage Mechanics. Nijhoff, Dordrecht,
The Netherlands.
KOITER, W.T. (1953). Stress-strain relations, uniqueness and variational theorems for
elastic-plastic material with singular yield surface. Q. Appl. Math. 11, 350-54.
LAI, W.M.; RUBIN, D. & KREMPL, E.(1978). Introduction to Continuum Mechanics. Pergamon
Press.
NOTES ON CONTINUUM MECHANICS

662
LANCZOS, C. (1970). The variational principles of mechanics. Dover Publications, Inc., New
York.
LEE, E.H. (1969). Elastic-Plastic deformation at finite strains. Journal of Applied Mechanics,
Vol. 36, pp. 1-6.
LEE, E.H. (1981). Some comments on elastic-plastic analysis. Int. J. Solids Structures, Vol. 17,
pp. 859-872.
LEMAITRE, J. & CHABOCHE, J.-L. (1990). Mechanics of Solids materials. Cambridge University
Press, Cambridge.
LINDER, C. (2003). An arbritary lagrangian-Eulerian finite element formulation for dynamics and finite
strain plasticity models. Masters Thesis. University Stuttgart.
LOVE, A.E.H. (1944). A treatise on the Mathematical Theory of Elasticity. Cambridge University
Press, London.
LU, S.C.H. & PISTER, K.S. (1975). Descomposition of deformation and representation of
the free energy function for isotropic solids, Int. J. of Solids and Structures, 11, pages: 927-
934.
LUBARDA, V.A. & BENSON D.J. (2001). On the partitioning of the rate of deformation
gradient in phenomenological plasticity. Int. J. of Solids and Structures, 38 pages: 6805-
6814.
LUBARDA, V.A. & KRAJCINOVIC, D. (1995). Some fundamental issues in rate theory of
damage-elastoplasticity. Int. J. of Plasticity, Vol. 11, N 7, pp. 763-797.
LUBARDA, V.A. (2004). Constitutive theories based on the multiplicative decomposition of
deformation gradient: Thermoelasticity, elastoplasticity, and biomechanics. Appl. Mech.
Rev., Vol. 57, no 2, March.
LUBLINER, J. (1990). Plasticity Theory. Macmillan Publishing Company, New York.
MALVERN, L.E. (1969). Introduction to the mechanics of a continuous medium. Prentice-Hall, Inc.
Englewood Cliffs, New Jersey.
MARSDEN, J. E. & HUGHES, T. J.R. (1983). Mathematical foundations of elasticity. Dover
Publications, Inc., New York.
MASE, G.E. (1977). Mecnica del Medio Continuo. McGraw-Hill, USA.
MAZARS, J. & LEMAITRE, J. (1984). Application of continuous damage mechanics to strain
and fracture behavior of concrete. Advances of Fracture Mechanics to Cementitious
Composites, NATO Advanced Research Workshop, 4-7 September 1984, Northwestern
University (Edited by S.P. Shah), pp. 375-388.
MAZARS, J. (1982). Mechanical damage and fracture of concrete structures. Advances in
Fracture Research (Fracture 81), Vol. 4, pp. 1499-1506. Pergamon Press, Oxford.
MENDELSON, A. (1968). Plasticity: Theory and application. New York, Robert E. Krieger
Publishing Company.
MOONEY, M. (1940). A theory of large elastic deformation. Journal of Applied Physics, Vol.
11, pp. 582-92.
MORMAN, K.N. (1986). The generalized strain measure with application to
nonhomegeneous deformation in rubber-like solids. J. Appl. Mech., Vol. 53, pp. 726-
728.
MORZ, Z. (1967). On the description of anisotropic work hardening. J. Mech. Phys. Solids,
15, 163.
BIBLIOGRAPHY

663
MULLINS, L. (1969). Softening of rubber by deformation. Rubber Chemistry and Tecnology, 42,
339-351.
NAGHDI, P.M. (1960). Stress-strain relations in plasticity and thermoplasticity. in E.H. Lee
& P. Symonds, eds., Plasticity. Pergamon, Oxford, pp. 121-69.
NEMAT-NASSER, S. (1982). On finite deformation elasto-plasticity. Int. J. Solids Structures,
Vol. 18, pp. 857-872.
NOWACKI, W. (1967). Problems of thermoelasticity. Progress in Thermoelasticity, VIIIth
European Mechanics Colloquium, Warszawa.
OGDEN, R.W. (1984). Non-linear elastic deformations. Dover Publications, Inc., New York.
OLIVER, J. & AGELET DE SARACBAR, C. (2000). Mecnica de medios continuos para ingenieros.
Ediciones UPC, Barcelona, Espaa.
OLIVER, J. & AGELET DE SARACBAR, C. (2000). Cuestiones y problemas de mecnica de medios
continuos. Ediciones UPC, Barcelona, Espaa.
OLIVER, J. (2002). Topics on Failure Mechanics. Monograph CIMNE N 68, International
Center for Numerical Methods in Engineering, Barcelona, Spain.
OLIVER, J.O. ; CERVERA, M. ; OLLER, S. & LUBLINER, J. (1990). Isotropic damage models
and smeared crack analysis of concrete, SCI-C Second Int. Conf. On Computer Aided Design
of Concrete Structure, Zell am See, Austria. Pg 945 957.
OLLER, S. (1988). Un modelo de dao continuo para materiales friccionales, Ph.D. thesis.
Universidad Politcnica de Catalua, Barcelona, Espaa.
OLLER, S. (2001). Fractura mecnica. Un enfoque global. CIMNE, Barcelona, Espaa.
OLLER, S.; OATE, E.; OLIVER, J. & LUBLINER, J. (1990). Finite element non-linear analysis
of concrete structures using a plastic-damage model. Engineering Fracture Mechanics Vol.
35, pp. 219-231.
OATE, E. (1992). Clculo de estructuras por el mtodo de elementos finitos anlisis esttico lineal.
Centro Internacional de Mtodos Numricos en Ingeniera. Barcelona- Espaa.
ORTIZ BERROCAL, L. (1985). Elasticidad. E.T.S. de Ingenieros Industriales. Litoprint. U.P.
Madrid.
ORTIZ, M. (1985). A Constitutive theory for inelastic behavior of concrete. Mech. Mater.,
Vol. 4, 67.
PABST, W. (2005). The linear theory of thermoelasticity from the viewpoint of rational
thermomechanics. Ceramics - Silikty, 49 (4) 242-251.
PARKER, D.F. (2003). Fields, Flows and Waves: An introduction to continuum models. Springer-
Verlag London, UK.
PILKEY, W. & WUNDERLICH, W. (1992). Mechanics of structures. Variational and computational
methods. CRC Press, Inc., Florida, USA.
POWERS, J.M. (2004). On the necessity of positive semi-definite conductivity and Onsager
reciprocity in modeling heat conduction in anisotropic media. Journal of Heat Transfer,
Vol.126, pp. 670-675.
PRAGER, W. (1945). Strain hardening under combined stress. J. Appl. Phys. 16, 837-40.
PRAGER, W. (1955). The theory of plasticity: A survey of recent achievements. Proc. Inst.
Mech. Eng. 169:41.
NOTES ON CONTINUUM MECHANICS

664
PRANDTL, L. (1924). Spannungverteilung in plastischen korpen, in Proceedings of the First
International. Congress of Applied Mechanics, Delft, The Netherlands, Vol. 43.
RABOTNOV, Y.N. (1963). On the equations of state for creep. Progress in Applied Mechanics,
Prager Anniversary Volume, page 307, New York, MacMillan.
RANKINE, W.J.M. (1951). Laws of the elasticity of solids bodies. Cambridge Dublin Math.
J. 6, 41-80.
REUSS, A. (1930). Berrucksichtingung der elastischen formanderungen in der
plastizitatstheorie, Z. Angew. Math. Mech. 10, 266.
ROMANO, A.; LANCELLOTA, R. & MARASCO, A. (2006). Continuum Mechanics using
Mathematica: fundamentals, applications, and scientific computing. Birkhauser Boston. USA.
RUNESSON, K. & MROZ, Z. (1989). A note on nonassociated plastic flow rules. Int. J.
Plasticity, 5, 639-658.
RUNESSON, K. & OTTOSEN, N. (1991). Discontinuity bifurcation of elastic-plastic solutions
at plane stress and plane strain. Int. J. Plasticity, 7:99-121.
SANSOUR, C. (1998). Large strain deformations of elastic shell. Constitutive modeling and
finite element analysis. Comp. Mech. Appl. Mech. Engrg. 161, pp1-18.
SANSOUR, C.; FEIH, S. & WAGNER, W. (2003). On the performance of enhanced strain
finite elements in large strain deformations of elastic shells. Comparison of two classes
of constitutive models for rubber materials. Report Universitat Karlsruhe, Institut fr
Baustatik
SCECHLER, E. (1952). Elasticity in Engineering. John Willey & Sons, Inc. new York.
ILHAV, M. (1997). The mechanics and thermodynamics of continuous media. Springer-
Verlag, Germany.
SIMO, J. & HUGHES, T.J.R. (1998). Computational Inelasticity. Springer-Verlag, New York.
SIMO, J.C. & JU, J.W. (1987a). Strain and Stress Based Continuum damage Models I.
Formulation. International Journal Solids Structures, Vol. 23, pp. 821-840.
SIMO, J.C. & JU, J.W. (1987b). Strain and Stress Based Continuum damage Models II.
Computational aspects. International Journal Solids Structures, Vol. 23, pp. 841-869.
SIMO, J.C. (1988a). A framework for finite strain elastoplasticity based on maximum plastic
dissipation: Part I. Computer Methods in Applied Mechanics and Engineering, Vol. 66, pp.
199-219.
SIMO, J.C. (1988b). A framework for finite strain elastoplasticity based on maximum plastic
dissipation: Part II. Computer Methods in Applied Mechanics and Engineering, Vol. 68, pp. 1-
31.
SIMO, J.C. (1992). Algorithms for static and dynamic multiplicative plasticity that preserve
the classical return mapping schemes of the infinitesimal theory, Computer Methods in
Applied Mechanics and Engineering, Vol. 99, pp. 61-112.
SOKOLNIKOFF, I.S. (1956). Mathematic theory of elasticity. New York, McGraw-Hill.
SOUZA NETO, E.A.; PERI, D. & OWEN, D.R.J. (1998). A phenomenological three-
dimensional rate-independent continuum damage model for highly filled polymers:
Formulation and computational aspects. J. Mech. Phys. Solids. 42(10), pp. 1533-1550.
SOUZA NETO, E.A.; PERI, D. & OWEN, D.R.J. (1998). Continuum Modelling and
Numerical Simulation of Material Damage at Finite Strains, Archives of Computational
Methods in Engineering, Vol.5,4, pp. 311-384.
BIBLIOGRAPHY

665
SPENCER, A.J.M. (1980). Continuum Mechanics, Longmans, Hong-Kong.
TIMOSHENKO, S. & GOODIER, J.N. (1951). Theory of elasticity, 2
nd
edition, McGraw-Hill.
TRELOAR, L.R.G. (1944). Stress-strain data for vulcanized rubber under various types of
deformation. Proc. of the Faraday Soc, 40:59-70.
TRELOAR, L.R.G. (1975). The physics of rubber elasticity. Clarendon Press, Oxford.
TRESCA, H. (1864). Mmire sur lEcoulement des corps solids soumis a de fortes pressions
comptes rendua academie de sciences. Paris, France, Vol.59, p.754..
TRUESDELL, C.A. & NOLL, W. (1965). The non-linear field theories of mechanics, in
Handuch der Physik, Vol. III/3, S. Flgge (Ed.), Springer-Verlag, Berlin.
UGURAL, A.C. (1981). Stress in Plates and Shells. McGraw Hill.
VALVERDE GUZMN, Q.M.(2002). Elementos estabilizados de bajo orden en mecnica de slidos.
PhD Thesis, Universitat Politecnica de Catalunya, Espaa.
VON MISES, R. (1930). ber die bisherigen Anstze in der lassischen MEchanik der
Kontinua. in Proceedings of the Third Intenational Congress for Applied Mechanics.
Vol.2, pp1-9.
VUJOEVI, L. & LUBARDA, V.A. (2002). Finite-Strain thermoelasticity based on
multiplicative decomposition of deformation gradient. Theoretical and Applied Mechanics,
vol. 28-29, pp. 379-399, Belgrade.
WILLAM, K. (2000). Constitutive models for materials: Encyclopedia of Physical Science & Technology,
3
rd
edition. Academic Press.
YEOH, O.H. (1993). Some forms of the strain energy function for rubber. Rubber Chem.
Technol., Vol. 66, pp. 754-71.
ZIEGLER, H. (1959). A modification of Pragers hardening rule. Q. Appl. Math. 17, 55-64.
ZIENKIEWICZ, O.C. & TAYLOR, R.L. (1994a). El mtodo de los elementos finitos. Volumen 1:
Formulacin bsica y problemas lineales. CIMNE, Barcelona, 4 edicin.
ZIENKIEWICZ, O.C. & TAYLOR, R.L. (1994b). El mtodo de los elementos finitos. Volumen 2:
Mecnica de slidos y fluidos. Dinmica y no linealidad. CIMNE, Barcelona, 4 edicin.

























A
acceleration
angular............................................................. 214
Eulerian ........................................................... 158
Lagrangia......................................................... 158
vector............................................................... 152
addition
vector................................................................. 18
additive decomposition
Green-Lagrange strain tensor .......................... 518
infinitesimal strain tensor ................................ 504
rate-of-deformation tensor............................... 518
adjugate of a Tensor.......................................... 42, 48
almansi strain tensor ......................181, 185, 199, 243
angle change
small deformation............................................ 232
angular momentum............................................... 302
angular velocity vector.................................. 204, 214
angular velocity tensor.......................................... 203
anisotropic tensor.................................................... 79
anisotropy (material)............................................. 381
antisymmetric tensor......................................... 36, 38
Archimedes Principle .......................................... 656
area Element ......................................................... 215
rate................................................................... 217
associated flow...................................................... 507
auxetic materials ................................................... 399
axial vector ................................................. 37, 39, 42
axiom of Impenetrability .............................. 153, 219
B
balance of mechanical energy............................... 310
barotropic.............................................................. 638
Bauschinger effect ................................................ 366
bijective function .................................................. 149
body force............................................................. 246
Boltzmann postulate ............................................. 303
Brazilian Test........................................................ 366
bulk modulus
adiabatic .......................................................... 564
bulk modulus ........................................................ 394
bulk viscosity coefficient ...................................... 643
C
Cauchy deformation tensor................................... 181
Cauchy heat flux................................................... 313
Cauchy stress tensor ......................................248, 250
effective (damage)........................................... 593
Cauchys first equation of motion......................... 298
Cauchys fundamental postulate........................... 248
Cauchys second law of motion............................ 303
Cauchys vorticity formula................................... 225
Cayley-Hamilton theorem ...................................... 77
change of angle
small deformation............................................ 232
characteristic determinant ....................................... 66
characteristic polynomial........................................ 67
circulation............................................................. 224
circulation preserving ........................................... 224
Clausius-Duhem inequality ...........................321, 346
Clausius-Planck inequality ................................... 321
coaxial tensors ........................................................ 81
coefficient of thermal expansion........................... 408
cofactor matrix........................................................ 49
cofactor tensor ........................................................ 42
cohesion................................................................ 368
commutative ........................................................... 81
compliance tensor ................................................. 391
component
normal ............................................................... 34
tangential ........................................................... 34
component transformation law............................... 54
compressibility factor ........................................... 394
compressibility modulus....................................... 394
compressible hyperelasticity material................... 440
compression test ................................................... 366
triaxial ............................................................. 368
conduction ............................................................ 328
configuration
current ............................................................. 149
deformed ......................................................... 149
initial ............................................................... 149
reference.......................................................... 149
conservation Law.................................................. 291
consistency condition ........................................... 495
constitutive equations ........................................... 341

Index
667 , Notes on Continuum Mechanics, Lecture Notes on Numerical
n Engineering and Sciences 4, DOI 10.1007/978-94-007-5986-2, Methods i
International Center for Numerical Methods in Engineering (CIMNE), 2013
E.W.V. Chaves
NOTES ON CONTINUUM MECHANICS

668
for heat conduction.......................................... 343
for simple thermoelastic material .................... 347
for stress (linear).............................................. 376
for stress .......................................................... 343
thermoviscoelastic material ..................... 353, 354
with internal variables ..................................... 355
constitutive equations for stress
hyperelastic...................................................... 427
continuity equation ....................................... 291, 338
contraction
double................................................................ 30
single ................................................................. 30
control surface ...................................................... 221
control volume...................................................... 221
convection............................................................. 328
convection-diffusion equation .............................. 336
convective rate...................................................... 279
coordinate
material............................................................ 151
spatial .............................................................. 151
coordinate system................................................... 16
Cartesian............................................................ 16
Cotter-Rivlin rate.......................................... 280, 281
coulombic frictional device .................................. 373
creep ..................................................................... 372
criterion of maximum shear stress ........................ 475
critical value
von Mises ........................................................ 472
cross product........................................................... 12
cubic symmetry..................................................... 389
curl........................................................................ 113
current configuration ............................................ 145
D
damage criterion ................................................... 594
damage master curve ............................................ 622
damage models ..................................................... 361
damage parameter consistency ............................. 597
damage variable............................................ 589, 598
Darcys law........................................................... 334
dashpot device ...................................................... 373
decomposition
additive.............................................................. 40
deformation
area element..................................................... 215
volume element ....................................... 215, 218
deformation gradient..................................... 163, 198
material............................................................ 165
spatial ...................................................... 166, 174
deformation of the volume element ...................... 218
deformation tensor
Cauchy............................................................. 181
left Cauchy-Green ............................177, 198, 212
Piola ........................................................ 177, 184
right Cauchy-Green ..........................177, 183, 191
density .................................................................. 285
derivative with tensors............................................ 84
derivative ................................................................ 94
descriptions
Eulerian ........................................................... 152
Lagrangian....................................................... 152
determinant of a tensor ........................................... 45
deviatoric plane..................................................... 139
deviatoric tensor
Voigt notation.................................................. 101
deviatoric tensor ................................................54, 87
diagonalization ........................................................69
differential .............................................................106
dilatancy ................................................................368
dilatation................................................................220
small deformation.............................................235
dilatational transformation.....................................440
Dirichlet boundary condition.................................333
displacement
vector................................................................152
displacement gradient ............................................228
material tensor..........................................169, 178
spatial tensor ............................................169, 182
dissipated power ....................................................646
dissipative pseudo-potential...................................356
divergence theorem ...............................................117
divergence .............................................................111
dot product...............................................................30
double scalar product...............................................30
Druckers stability postulates ................................515
ductile materials ....................................................364
Duhamel-Neumann equations ...............................561
dyadic ......................................................................28
E
effective stress ...............................................587, 589
eigenvalue........................................................66, 127
eigenvector ..............................................................66
Einstein notation......................................................20
elastic acoustic tensor ............................................398
elastic compliance tensor...............................391, 397
elastic limit ............................................................363
elastic modulus......................................................363
elastic potential......................................................424
elastic process................................................357, 360
elastic pseudomoduli .............................................431
elastic stiffness tensor............................................377
elasticity tensor.............................. 358, 377, 390, 397
isothermal.........................................................557
tangent..............................................................427
elasticity tensor components..................................381
elastoplastic model
isotropic-kinematic...........................................502
kinematic hardening.........................................501
small deformation.............................................507
ellipsoid
tensor................................................................138
energy density........................................................286
strain.................................................................357
energy equation .....................................................315
Eulerian............................................................315
Lagrangian .......................................................317
with discontinuities ..........................................317
engineering notation..............................................375
engineering strain ..........................................233, 243
enthalpy.........................................................322, 343
entropy...................................................................319
entropy inequality..................................................319
with discontinuity.............................................325
equation of vorticity ..............................................654
equations of motion.......................................297, 376
Eulerian............................................................298
Lagrangian .......................................................298
with discontinuities ..........................................301
with discontinuities ..........................................302
INDEX

669
equations of state .................................................. 343
equilibrium equations
Eulerian ........................................................... 298
Lagrangian....................................................... 299
equivalent plastic strain ........................................ 513
Euclidean norm....................................................... 11
Euler angles ............................................................ 65
Eulerian
variables .......................................................... 153
Eulerian description.............................................. 157
Eulerian finite strain tensor................................... 181
Eulerian stretch tensor .......................................... 195
Eulerian variable
material time derivative................................... 158
external mechanical power ................................... 308
F
Ficks law of diffusion.......................................... 335
field
conservative............................................. 115, 120
scalar ............................................................... 105
second-order tensor ......................................... 105
stationary......................................................... 159
vector............................................................... 105
finite strain............................................................ 176
first law of thermodynamics ......................... 307, 314
first Piola-Kirchhoff stress tensor ................. 261, 262
flow plastic vector................................................. 468
flow rule
associated ........................................................ 494
perfect plasticity .............................................. 493
Prandtl-Reuss................................................... 517
fluids..................................................................... 369
flux........................................................................ 286
energy.............................................................. 287
mass................................................................. 287
flux problem ................................................. 327, 338
forces .................................................................... 245
body................................................................. 246
density ............................................................. 246
gravitational..................................................... 246
internal............................................................. 245
surface ............................................................. 245
thermodynamic........................................ 356, 505
viscous............................................................. 648
Fouriers equation................................................. 333
Fouriers law of heat conduction .......... 325, 328, 329
fourth-order projection tensor ................................. 54
fragile materials .................................................... 364
Frobenius norm....................................................... 79
fundamental equations of continuum mechanics .. 342

G
Gauss theorem..................................................... 117
generalized Hookes law....................................... 377
Gibbs free energy.......................................... 322, 343
gradient ................................................................. 106
Green elasticity..................................................... 423
Greens first identity............................................. 123
Greens second identity ........................................ 123
Greens theorem.................................................... 121
Green-Lagrange strain tensor........................ 178, 243
Green-Naghdi Rate............................................... 282
H
Haigh-Westergaard stress space ........................... 470
hardening plasticity............................................... 490
hardening/softening modulus
continuum........................................................ 597
heat capacity......................................................... 325
heat conduction inequality.............................321, 547
heat flux equation ..........................................325, 332
heat flux.........................................................313, 329
heat source............................................................ 314
Helmholtz free energy ...................................322, 343
isotropic damage model................................... 591
hexagonal symmetry............................................. 386
homogeneous deformation.................................... 191
homogeneous........................................................ 105
continuum........................................................ 146
Hookes law...........................................372, 377, 392
plane strain ...................................................... 414
plane stress ...................................................... 412
Voigt notation...........................................378, 379
hydrostatic axis 140
hydrostatic pressure .......................................368, 636
hyperelasticity................................................361, 423
I
identity tensor ......................................................... 44
incompressibility restriction ................................. 450
incompressible...............................................220, 295
index
dummy .............................................................. 21
free .................................................................... 21
inelastic behavior.................................................. 361
infinitesimal strain tensor ..............................230, 375
infinitesimal strain theory..................................... 229
initial boundary value problem............................. 339
linear elasticity .........................................357, 376
initial configuration .............................................. 145
initial strain........................................................... 412
instantaneous elastic moduli ................................. 431
integration by parts ............................................... 117
internal energy...................................................... 313
internal force......................................................... 245
internal friction..................................................... 368
internal variables .................................................. 506
invariant.......................................................43, 68, 86
irreversible process ............................................... 362
irrotational .....................................................173, 224
isentropic process ..........................................319, 552
isochoric motion ....................................220, 225, 295
isochoric transformation....................................... 440
isothermal Lam constants ................................... 562
isothermal processes ............................................. 552
isotropic material ...........................................359, 390
isotropic tensor ....................................................... 80
isotropic tensor function ......................................... 92
isotropic-kinematic hardening plasticity............... 491
isotropy................................................................. 381
J
Jacobian determinant .....................153, 174, 175, 220
NOTES ON CONTINUUM MECHANICS

670
rate........................................................... 171, 185
Jaumann-Zaremba rate.................................. 280, 281
K
Kelvin-Stokes theorem........................................ 121
kinematic equations .............................................. 377
kinematic hardening plasticity.............................. 490
kinematic tensors .......................................... 183, 197
rate................................................................... 208
objectivity........................................................ 273
kinetic energy ....................................................... 307
rigid body motion............................................ 312
Knudsen number................................................... 635
Kronecker delta............................................22, 44, 80
Kuhn strain ........................................................... 244
Kuhn-Tucker conditions ....................................... 494
L
Lagrangian
variable............................................................ 153
Lagrangian description ......................................... 156
Lagrangian stretch tensor...................................... 195
Lam constants ..............................359, 363, 390, 397
isentropic......................................................... 553
isothermal ........................................................ 553
Laplaces equation........................................ 333, 335
Laplacian operator ................................................ 112
latent heat tensor....................................554, 557, 572
left Cauchy-Green deformation tenso................... 198
left stretch tensor .................................................. 195
Levi-Civita pseudo-tensor................................. 45, 80
Lie derivative........................................................ 528
linear elasticity...................................................... 361
linear elasticity theory........................................... 375
linear momentum.................................................. 297
rigid body motion............................................ 304
linearly dependent................................................... 13
M
Macaulay bracket.................................................. 603
mass continuity equation .............................. 224, 291
Eulerian ........................................................... 292
incompressible medium................................... 295
Lagrangian....................................................... 294
with discontinuities.......................................... 295
mass density...................................146, 150, 285, 286
material curve ........................................169, 194, 220
material frame indifference................................... 270
material point........................................................ 146
material surface..................................................... 220
material time derivative ........................................ 156
material volume .................................................... 221
maximum tangential component 129
mean stress............................................................ 636
mechanical power ................................................. 309
mechanical properties ........................................... 398
mixed product ......................................................... 12
Mohrs circle ................................................ 134, 137
moments of inertia................................................ 305
monoclinic symmetry ........................................... 383
motion with deformation ...................................... 147
Mullins effect........................................................ 621
multiplication
scalar ..................................................................29
multiplicative decomposition
deformation gradient ........................................518
volumetric and isochoric ..................................225
N
nabla symbol..........................................................106
Nansons formula ..........................................216, 262
Navier-Lam equations..........................................410
Navier-Poisson law................................................643
negative definite (tensor) .........................................52
Neumann boundary condition................................333
Newtons law of cooling........................................330
Newtonian fluids ...................................................371
non-linear elastic ...................................................361
nonlinear elasticity.................................................423
non-Newtonian fluids ............................................371
normal component .................................................127
normal octahedral vector .......................................139
normal vector.........................................................125
normalization condition.................................427, 436
norms of tensors ......................................................79
notation
indicial..........................................................20, 34
symbolic.............................................................34
tensorial..............................................................34
O
objective rates........................................................278
objectivity of tensors .............................................270
octahedral plane.............................................139, 140
octahedral vector
tangent..............................................................139
Oldroyd rate...........................................................279
operator
Laplacian..........................................................112
orthogonal matrix ..................................................148
orthogonal tensor.............................................51, 148
improper .............................................................51
proper .................................................................51
orthogonal transformation .......................................51
orthogonality ...........................................................12
orthonormal basis ..............................................16, 26
orthotropic material ...............................................404
orthotropic symmetry ............................................384
ortogonalidad...........................................................51
P
parallel axis theorem..............................................306
particle...................................................................146
path line.................................................................146
permutation symbol .........................................23, 113
permutation tensor ...................................................45
persistency
condition...........................................................495
Piola deformation tensor................................177, 184
plane strain ....................................................236, 410
plane stress ............................................................410
plastic flow rule.....................................................506
plastic flow tensor..................................................468
plastic multiplier....................................................493
INDEX

671
plasticity
perfect.............................................................. 491
plasticity models ................................................... 361
elastic-perfectly ............................................... 489
Poissons equation ................................................ 332
Poissons ratio....................................................... 394
polar decomposition........................................ 82, 195
rates ................................................................. 203
rotation tensor.................................................. 198
polar rate............................................................... 282
pore pressure......................................................... 369
position vector ...................................................... 147
positive definite tensor.................................... 52, 126
power extended..................................................... 310
principal invariants ........................................... 67, 68
derivative........................................................... 94
deviatoric stress tensor..................................... 259
principal space ................................................ 67, 128
principle
conservation of angular momentum ................ 285
conservation of energy............................. 285, 307
conservation of linear momentum................... 285
conservation of mass ....................................... 285
determinism..................................................... 343
dissipation........................................................ 343
equipresence.................................................... 343
irreversibility ................................................... 285
limited memory ............................................... 344
local action ...................................................... 343
objectivity........................................................ 343
of action and reation........................................ 249
of objectivity.................................................... 270
Saint-Venant.................................................... 406
superposition ................................................... 407
product
tensor ................................................................. 28
double scalar...................................................... 30
products of inertia................................................. 305
projection tensor ................................................... 441
proportionality limit.............................................. 363
pseudo-invariants of anisotropy............................ 463
pseudo-tensor
Levi-Civita ........................................................ 24
R
radiant heat constant ............................................. 314
radiation................................................................ 328
ramp function........................................................ 603
rate
convective........................................................ 279
Cotter-Rivlin.................................................... 280
Green-McInnis................................................. 282
Green-Naghdi .................................................. 282
Jaumann-Zaremba ........................................... 280
Oldroyd.................................................... 279, 281
polar................................................................. 282
Truesdell.......................................................... 283
rate independent.................................................... 467
rate of change
convective........................................................ 157
local ................................................................. 157
rate of the material rotation tensor ........................ 203
rate-of-deformation tensor ............................ 172, 204
recoverable power................................................. 646
reflection tensor ...................................................... 51
relaxation.............................................................. 372
reversible process ..........................................323, 427
Reynolds number.................................................. 639
Reynolds transport theorem................................. 287
with discontinuities ......................................... 290
rheological models................................................ 372
isotropic hardening elastoplastic ..................... 496
perfect elastoplastic......................................... 492
right Cauchy-Green deformation tensor ........177, 183
right stretch tensor ................................................ 195
rigid body motion .. 147, 148, 173, 182, 192, 214, 327
rotation tensor......................................................... 51
infinitesimal..................................................... 230
polar decomposition ........................................ 198
rate................................................................... 203
rotor ...................................................................... 113
rupture strength point............................................ 363
S
scalar......................................................................... 9
scalar multiplication ............................................... 18
scalar product.........................................11, 18, 22, 30
Voigt notation.................................................... 98
scalar product trace................................................. 43
scalar triple product .......................................... 12, 18
scalar-valued tensor ................................................ 91
second law of Thermodynamics ........................... 319
second Piola-Kirchhoff stress tensor .................... 262
second-order tensor
projection......................................................... 125
second-order-valued tensor..................................... 91
shear modulus....................................................... 394
sign function......................................................... 494
simple thermoelastic materials.............................. 344
singular
tensor................................................................. 46
skew tensor ............................................................. 36
small deformation theory...................................... 229
spatial velocity gradient........................................ 213
specific heat ...................................................560, 572
constant stress.................................................. 554
constant volume............................................... 554
specific internal energy......................................... 343
spectral representation
Voigt notation.................................................. 100
spectral representation .................................73, 74, 81
spherical axis ........................................................ 140
spherical part .......................................................... 54
spherical tensor..........................................68, 86, 138
spin tensor............................................................. 172
infinitesimal..................................................... 230
spring device......................................................... 373
Stefan-Boltzmann law .......................................... 330
Steiners theorem.................................................. 306
stiffness modulus
damage secant ................................................. 589
stiffness tensor
elastic .............................................................. 377
elastic-damage secant ...................................... 593
Stokes condition.................................................. 644
Stokesian fluids .................................................... 371
strain
engineering...................................................... 233
equivalent ........................................................ 594
strain energy density......................357, 400, 402, 424
NOTES ON CONTINUUM MECHANICS

672
strain gauge........................................................... 416
strain localization.................................................. 364
strain rosette.......................................................... 416
strain tensor
Almansi ........................................................... 181
Biot.................................................................. 242
Eulerian finite.................................................. 181
Green-Lagrange....................................... 178, 191
Green-St_Venant ............................................. 178
Hencky ............................................................ 241
infinitesimal............................................. 230, 375
linear Almansi ................................................. 229
linear Green-Lagrange..................................... 229
logarithmic ...................................................... 241
strain-displacement equations
linear................................................................ 377
streamlines............................................................ 161
stress constitutive equation
Newtonian fluid............................................... 642
stress power .................................................. 309, 424
Newtonian fluids ............................................. 646
stress tensors......................................................... 265
Biot.................................................................. 265
Cauchy..............................................248, 250, 426
effective................................................... 587, 610
first Piola-Kirchhoff .........................261, 262, 426
Kirchhoff ................................................. 262, 426
Mandel..................................................... 265, 426
nominal............................................................ 262
objective rates.................................................. 282
objectivity........................................................ 275
second Piola-Kirchhoff.................................... 262
thermal..................................................... 549, 557
true .................................................................. 250
viscous............................................................. 637
stretch ................................................................... 164
principal........................................................... 190
small deformation............................................ 231
stretch ratio........................................................... 164
stretch tensor......................................................... 187
left ................................................................... 195
right ......................................................... 191, 195
substitution operator ............................................... 23
subtraction
vector................................................................. 18
summation convention............................................ 20
surface force ................................................. 245, 246
Swaiger strain ....................................................... 244
symbol
permutation........................................................ 24
symmetric tensor..................................................... 36
symmetry
major ................................................................. 36
minor ................................................................. 36
symmetry planes ................................................... 382
system
material............................................................ 149
spatial .............................................................. 149
T
tangent stiffness modulus
elastoplastic............................................. 499, 501
tangent stiffness pseudo-tensor
elastic ...................................................... 431, 432
tangent stiffness tensors........................................ 433
adiabatic elastic ................................................553
elastic ...............................................................427
elastic-damage..................................................602
elastoplastic..............................................507, 509
elastoplastic-damage ........................................615
instantaneous elastic.................................430, 431
isothermal elastic..............................................552
material elastic .........................................427, 428
spatial elastic............................................428, 430
tangential component.............................................128
maximum, minimum........................................128
tangential vector ....................................................125
tasa de Cotter-Rivlin..............................................529
Taylor series ............................................................91
temperature............................................................328
tensile modulus......................................................363
tensile testing.........................................................362
tensor .........................................................................9
additive decomposition ......................................53
Cauchy stress....................................................248
cofactor ..............................................................42
components ........................................................32
first-order .............................................................9
identity ...............................................................44
negative definite.................................................52
psotive definite...................................................52
rate-of-deformation ..........................................172
rate-of-rotation .................................................172
second-order...................................................9, 28
skew-symmetric .................................................36
spin...................................................................172
transpose.............................................................34
zeroth-order..........................................................9
tensor field.............................................................105
tensor product ..........................................................28
tensor series .............................................................91
tetragonal symmetry ..............................................386
theorem
Reynolds' transport...........................................288
thermal conduction................................................328
thermal conductivity tensor ...................325, 329, 559
thermal convection transfer ...................................330
thermal deformation ..............................................408
thermal diffusivity .................................................332
thermal diffusivity tensor.......................................559
thermal expansion tensor ...............................551, 558
thermal power................................................313, 314
thermal strain.........................................................415
thermal stretch coefficient .....................................578
thermodynamic forces ...........................356, 505, 584
thermodynamic potential .......................322, 549, 551
thermodynamic pressure........................................637
thermodynamics
second law........................................................319
thermoelastic materials ..................................344, 345
consitutive equations........................................349
thermoviscoelastic material ...................................351
time derivative.........................................................87
total derivative.......................................................106
trace
tensor..................................................................42
traction vector........................................245, 252, 254
pseudo ..............................................................261
traction...................................................................245
trajectory................................................................146
transformation
linear ..................................................................14
INDEX

673
transformation law
elasticity tensor................................................ 381
transformation law............................................ 54, 58
transformation matrix ..........................56, 62, 99, 380
transport equations........................................ 223, 338
transpose................................................................. 34
transversely isotropic material .............................. 405
transversely isotropic symmetry ........................... 388
triclinic materials .................................................. 382
triple scalar product ................................................ 26
true strain...................................................... 241, 243
true stress tensor ................................................... 250
Truesdell stress rate .............................................. 283
U
ultimate strength point .......................................... 363
undeformed configuration..................................... 145
uniform................................................................. 105
unit extension................................................ 164, 188
small deformation............................................ 233
small deformation............................................ 231
unit tensor
fourth-order ................................................. 44, 75
second-order ...................................................... 44
V
vector ........................................................................ 9
angular-velocity............................................... 204
axial ................................................................... 37
norm of .............................................................. 18
octahedral 139
product......................................12, 18, 25, 32, 113
projection........................................................... 11
triple product ............................................... 13, 19
unit............................................................... 11, 18
vorticity ........................................................... 172
zero.................................................................... 11
zero.................................................................... 18
velocity
angular............................................................. 214
Eulerian ........................................................... 158
Lagrangian....................................................... 158
vector............................................................... 152
velocity gradient
spatial ...................................................... 172, 213
virgin curve........................................................... 622
viscoelastic materials ............................................ 371
viscosity................................................................ 370
viscous forces ....................................................... 648
Voigt notation................................................... 96, 97
transfromation law............................................. 99
unit tensors ........................................................ 97
volume element
rate................................................................... 219
volume element deformation ................................ 218
volume ratio
small deformation............................................ 235
vorticity vector...................................... 173, 224, 640
Y
yield condition
Huber-von Mises ............................................. 512
Mises-Huber.................................................... 546
yield criterion........................................................ 468
Alternative Drucker-Prager ............................. 484
yield curve ............................................................ 470
yield point............................................................. 363
yield surface
Drucker-Prager ................................................ 484
isotropic material............................................. 469
Mohr-Coulomb................................................ 480
Rankine ........................................................... 488
von Mises ........................................................ 474
Youngs modulus.................................................. 363

Você também pode gostar