You are on page 1of 391

Modeling and Simulation in Science,

Engineering and Technology

Alan D. Freed

Soft
Solids
A Primer to the Theoretical
Mechanics of Materials
Modeling and Simulation in Science, Engineering and Technology

Series Editor
Nicola Bellomo
Politecnico di Torino
Torino, Italy

Editorial Advisory Board

K.J. Bathe P. Koumoutsakos


Department of Mechanical Engineering Computational Science & Engineering
Massachusetts Institute of Technology Laboratory
Cambridge, MA, USA ETH Zrich
Zrich, Switzerland
M. Chaplain
H.G. Othmer
Division of Mathematics
Department of Mathematics
University of Dundee
University of Minnesota
Dundee, Scotland, UK Minneapolis, MN, USA

P. Degond K.R. Rajagopal


Department of Mathematics, Department of Mechanical Engineering
Imperial College London, Texas A&M University
London, United Kingdom College Station, TX, USA

A. Deutsch T.E. Tezduyar


Center for Information Services Department of Mechanical Engineering &
and High-Performance Computing Materials Science
Rice University
Technische Universitt Dresden
Houston, TX, USA
Dresden, Germany
A. Tosin
M.A. Herrero Istituto per le Applicazioni del Calcolo
Departamento de Matematica Aplicada M. Picone
Universidad Complutense de Madrid Consiglio Nazionale delle Ricerche
Madrid, Spain Roma, Italy

For further volumes:


http://www.springer.com/series/4960
Alan D. Freed

Soft Solids
A Primer to the Theoretical Mechanics
of Materials
Alan D. Freed
Department of Mechanical Engineering
Saginaw Valley State University
University Center, MI, USA

ISSN 2164-3679 ISSN 2164-3725 (electronic)


ISBN 978-3-319-03550-5 ISBN 978-3-319-03551-2 (eBook)
DOI 10.1007/978-3-319-03551-2
Springer Cham Heidelberg New York Dordrecht London
Library of Congress Control Number: 2014931652

Mathematics Subject Classification (2010): 15A72, 65L06, 65R20, 74-01, 74A05, 74A10, 74A20,
74B20, 74D10, 74L15, 74S30

Springer International Publishing Switzerland 2014


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part
of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or
information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed. Exempted from this legal reservation are brief ex-
cerpts in connection with reviews or scholarly analysis or material supplied specifically for the purpose
of being entered and executed on a computer system, for exclusive use by the purchaser of the work.
Duplication of this publication or parts thereof is permitted only under the provisions of the Copyright
Law of the Publishers location, in its current version, and permission for use must always be obtained
from Springer. Permissions for use may be obtained through RightsLink at the Copyright Clearance
Center. Violations are liable to prosecution under the respective Copyright Law.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publi-
cation does not imply, even in the absence of a specific statement, that such names are exempt from
the relevant protective laws and regulations and therefore free for general use.
While the advice and information in this book are believed to be true and accurate at the date of
publication, neither the authors nor the editors nor the publisher can accept any legal responsibility for
any errors or omissions that may be made. The publisher makes no warranty, express or implied, with
respect to the material contained herein.

Printed on acid-free paper

Springer is part of Springer Science+Business Media (www.birkhauser-science.com)


Dedicated to the love of my life:
Karen
In the fields of observation,
chance favors only the prepared mind.
Louis Pasteur
Lecture, University of Lille (7 December 1854)
Preface

This textbook is intended for seniors and masters-level graduate stu-


dents whose studies are in the fields of engineering, physics, or
applied mathematics. The intent of this book is to give you an opportu-
nity to come to an understanding of the physical principles and concepts
pertinent to the mathematical modeling of soft materials used in engineer-
ing practice and to gain competence by applying some basic mathematical
techniques to physical problems that arise in the characterization of mate-
rials, thereby enhancing your overall understanding and skill level.
This text is not intended to provide lengthy derivations of, e.g., the con-
servation laws and other such things. Rather, their results are stated and
used in an attempt to provide you with an overall understanding of the topic.
Advanced derivations and treatments are relegated to graduate courses, for
which there are numerous textbooks, some being Bowen (1989), Chadwick
(1976), Gurtin (1981), Gurtin et al. (2010), Holzapfel (2000), Lai et al.
(1974), Lodge (1964), Ogden (1984), Sokolnikoff (1964), Truesdell and
Noll (2004), and Truesdell and Toupin (1960).
Those of you who have studied the calculus up through differential
equations and who have had exposure to linear algebra will have sufficient
mathematical skills for taking this course. Any other mathematics that you
may need will be taught to you along the way. A casual inspection of this
text may give you cause for pause, but be assured, the authors intention
is to educate you in your understanding of concepts, not in your ability to
reproduce derivations and developments. The authors intention is that you
enhance your mathematical knowledge by osmosis.
Seven BVPs are studied over the course of this book, each pertaining
to an experiment used for material characterization, some classic, some
not. Four are done as examples, and three are left for you to learn from.

ix
x Preface

These seven BVPs are visited at the end of each chapter, applying what you
just learned in the current chapter and intermingling it with what you have
already learned from prior chapters, thereby extending the development
of each experiment and your understanding of them topic by topic. The
four worked-out BVPs increase with complexity. The three exercise BVPs
left for your development do not require higher-level mathematical skills.
All questions are designed to build upon your understanding of the basic
concepts and principles that are being taught to you at that particular time.
By the end of the course, your overall exposure to the mathematics in-
volved should make you more comfortable with them. Mathematics is a
language, and in this textbook, the author intends that you learn it by im-
mersion rather than by the more wrought and formal approach of lemmas,
theorems, and proofs. That can come later, enhanced by your experiences
and intuitions gained after taking this primer course.

University Center, MI, USA Alan D. Freed

Acknowledgments
Many people, in one way or another, helped me bring this textbook into
reality. Foremost is my family who willingly made the greatest sacrifice
that anyone can offer: lost time together.
Formidable were my mentors at the University of WisconsinMadison
who inspired me, and to whom I am indebted. May this be a partial in-
stallment on the debt I owe to Prof. Bela I. Sandor, my thesis advisor, who
introduced me to the topic of mechanics of materials and who taught me
how to write and how to learn from adversity; Prof. Millard W. Johnson, Jr.
(19282009), who taught me classic continuum mechanics, then opened
my eyes to it through applications; and to Prof. Arthur S. Lodge (1922
2005), who patiently taught me the rigor of mathematics and the physics
that underlie continuum mechanics, that one should aspire to think outside
the box and, by example, that good theory and good experiment go hand in
hand.
Prof. Raymond Ogden from the University of Glasgow and Prof. K. R.
Rajagopal from Texas A&M provided critical reviews of earlier drafts of
this book, which helped me sharpen its focus. Prof. Rajagopal was the ad-
vocate through which publication became possible. Dr. Allen Mann was
my managing editor at Birkhuser Science, and Prof. Beth Jorgensen from
Preface xi

Saginaw Valley State University provided technical editing of the draft


copy. Three anonymous reviewers provided useful critical assessment of
the draft copy.
Detailed conversations with Prof. John C. Butcher from the University
of Auckland and years of collaboration with Prof. Kai Diethelm, who holds
joint appointments at GNS Gesellschaft fr Numerische Simulation mbH
and the Technische Universitt Braunschweig, have resulted in numerical
tools that have allowed me to explore many aspects of material modeling
and behavior that would have otherwise been inaccessible to me, two of
which are presented in Appendices D and E. Prof. Yuri Luchko from the
Beuth Hochschule fr Technik Berlin derived the algorithm used to com-
pute the Mittag-Leffler function, presented in Appendix F, which further
enabled my study of fractional-order viscoelasticity. Guidance from Dr.
Daniel R. Einstein from Pacific Northwest National Laboratory has greatly
facilitated my recent studies of soft tissues. Dr. Randall Schmidt of Dow
Corning Corporation provided experimental data for three varieties of their
PDMS silicon elastomers. Prof. Ian LeGrice from the University of Auck-
land provided experimental data for porcine ventricular myocardium. Dr.
E. Malcolm Field, MD, endowed the chair in engineering that I hold, which
he named after his uncle, Clifford H. Spicer, founder of Spicer Engineering
in Saginaw. Dr. Fields financial support made it possible for me to write
this book. I am indebted to you all.
Mr. Chandler Benjamin, my first SVSU student, was my sounding board
while this book was being forged. I am grateful as well to all of the students
who have followed since, pointing out errors and offering their advice.
This document was typeset in LATEX. TX fonts,1 created by Young
Ryu and distributed under the GNU license, were used to typeset the text.
MathTimeTM Professional II fonts,2 created by Walter Schmidt, were used
to typeset the mathematical expressions. Computer modern fonts were used
to typeset the text within equations, like sin in sin x, and the figure and table
captions. For the creation of figures, vector illustrations were drawn using
Xfig ,3 while data illustrations were drawn using Grace.4 Data taken from

1 TX fonts come prepackaged with most TEX distributions.


2 MathTime is a trademark of Publish or Perish, Inc. MathTime Professional II
(a.k.a. MTPro II) fonts are commercially available from http://www.pctex.com.
3X
fig is publicly available from http://www.xfig.org.
4 Grace is publicly available from http://plasma-gate.weizmann.ac.il/Grace/.
xii Preface

published sources were re-digitized by scanning an image of the required


figure and then digitizing the individual data points from the scanned image
using the g3data software of Jonas Frantz.5

5 g3data is publicly available from http://www.frantz.fi.


Contents

Preface ix
Nomenclature xxi
Introduction xxix

Continuum Fields 1
1. Kinematics 5
1.1 Motion . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.1.1 Homogeneous Motions . . . . . . . . . . . . . 7
1.2 Velocity and Acceleration . . . . . . . . . . . . . . . . . 8
1.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3.1 Uniaxial Extension . . . . . . . . . . . . . . . 10
1.3.2 Equi-biaxial Extension . . . . . . . . . . . . . 11
1.3.3 Simple Shear . . . . . . . . . . . . . . . . . . . 12
1.3.4 Homogeneous Planar Membranes . . . . . . . . 14
1.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.4.1 Pure Shear . . . . . . . . . . . . . . . . . . . . 18
1.4.2 Biaxial Extension . . . . . . . . . . . . . . . . 19
1.4.3 Extension Followed by Simple Shear . . . . . . 20
1.4.4 Other Problems . . . . . . . . . . . . . . . . . 22

2. Deformation 23
2.1 Homogeneous Deformation . . . . . . . . . . . . . . . . 25
2.2 Conservation of Mass . . . . . . . . . . . . . . . . . . . 25
2.2.1 Isochoric Deformation . . . . . . . . . . . . . . 26
xiii
xiv Contents

2.3 Deformation Gradient as a Mapping Function . . . . . . 27


2.4 Stretch and Rotation . . . . . . . . . . . . . . . . . . . . 28
2.5 Rate Fields . . . . . . . . . . . . . . . . . . . . . . . . 30
2.5.1 Lagrangian Velocity Gradient . . . . . . . . . . 31
2.5.2 Isochoric Deformations . . . . . . . . . . . . . 32
2.6 Numerical Implementation . . . . . . . . . . . . . . . . 32
2.6.1 Angular Velocity . . . . . . . . . . . . . . . . . 33
2.6.2 Rotation . . . . . . . . . . . . . . . . . . . . . 34
2.6.3 Stretch . . . . . . . . . . . . . . . . . . . . . . 37
2.7 Examples . . . . . . . . . . . . . . . . . . . . . . . . . 38
2.7.1 Uniaxial Extension . . . . . . . . . . . . . . . 38
2.7.2 Equi-biaxial Extension . . . . . . . . . . . . . 39
2.7.3 Simple Shear . . . . . . . . . . . . . . . . . . . 40
2.7.4 Homogeneous Planar Membranes . . . . . . . . 42
2.8 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . 44
2.8.1 Pure Shear . . . . . . . . . . . . . . . . . . . . 44
2.8.2 Biaxial Extension . . . . . . . . . . . . . . . . 44
2.8.3 Extension Followed by Simple Shear . . . . . . 45
2.8.4 Other Problems . . . . . . . . . . . . . . . . . 45

3. Strain 47
3.1 Deformation . . . . . . . . . . . . . . . . . . . . . . . . 47
3.2 Measures of Strain . . . . . . . . . . . . . . . . . . . . 49
3.2.1 Hencky Strain . . . . . . . . . . . . . . . . . . 51
3.2.2 Infinitesimal Strain/Rotation Relationships . . . 52
3.3 Geometric Interpretations of Strain . . . . . . . . . . . . 53
3.3.1 An Areal Interpretation . . . . . . . . . . . . . 56
3.4 Strain Rates . . . . . . . . . . . . . . . . . . . . . . . . 57
3.4.1 Integration . . . . . . . . . . . . . . . . . . . . 60
3.5 Strain and Strain-Rate Fields for Numerical Analysis . . 61
3.5.1 Formulation in Terms of Green Strain . . . . . . 61
3.5.2 Formulation in Terms of Lodge Strain . . . . . 64
3.6 Examples . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.6.1 Uniaxial Extension . . . . . . . . . . . . . . . 67
3.6.2 Equi-biaxial Extension . . . . . . . . . . . . . 68
3.6.3 Simple Shear . . . . . . . . . . . . . . . . . . . 69
3.6.4 Homogeneous Planar Membranes . . . . . . . . 71
3.7 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . 73
3.7.1 Pure Shear . . . . . . . . . . . . . . . . . . . . 73
Contents xv

3.7.2 Biaxial Extension . . . . . . . . . . . . . . . . 73


3.7.3 Extension Followed by Simple Shear . . . . . . 73
3.7.4 Other Problems . . . . . . . . . . . . . . . . . 73

4. Stress 77
4.1 Kirchhoff Stress . . . . . . . . . . . . . . . . . . . . . . 79
4.2 Conservation of Momenta . . . . . . . . . . . . . . . . 79
4.3 First PiolaKirchhoff Stress . . . . . . . . . . . . . . . 81
4.4 Second PiolaKirchhoff Stress . . . . . . . . . . . . . . 83
4.5 Stress Rates . . . . . . . . . . . . . . . . . . . . . . . . 83
4.5.1 Integrate for Stress . . . . . . . . . . . . . . . . 84
4.6 The Extra Stress . . . . . . . . . . . . . . . . . . . . . . 85
4.7 Hills Constitutive Inequalities . . . . . . . . . . . . . . 86
4.7.1 Incompressible Materials . . . . . . . . . . . . 86
4.7.2 Eulerian Formulations . . . . . . . . . . . . . . 87
4.8 Stresses for Numerical Analysis . . . . . . . . . . . . . 88
4.9 Examples . . . . . . . . . . . . . . . . . . . . . . . . . 90
4.9.1 Uniaxial Extension . . . . . . . . . . . . . . . 90
4.9.2 Equi-biaxial Extension . . . . . . . . . . . . . 92
4.9.3 Simple Shear . . . . . . . . . . . . . . . . . . . 95
4.9.4 Homogeneous Planar Membranes . . . . . . . . 97
4.10 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . 101
4.10.1 Pure Shear . . . . . . . . . . . . . . . . . . . . 101
4.10.2 Biaxial Extension . . . . . . . . . . . . . . . . 101
4.10.3 Extension Followed by Simple Shear . . . . . . 102
4.10.4 Other Problems . . . . . . . . . . . . . . . . . 102

Constitutive Equations 105


5. Explicit Elasticity 109
5.1 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . 110
5.1.1 Green Elastic Solid . . . . . . . . . . . . . . . 110
5.1.2 Lodge Elastic Solid . . . . . . . . . . . . . . . 112
5.1.3 Properties . . . . . . . . . . . . . . . . . . . . 113
5.2 Isotropic Theory . . . . . . . . . . . . . . . . . . . . . 114
5.2.1 Isotropic Green Solid . . . . . . . . . . . . . . 114
5.2.2 Isotropic Lodge Solid . . . . . . . . . . . . . . 117
5.2.3 A Property of Explicit Elastic Solids . . . . . . 119
xvi Contents

5.3 A Collection of Material Models . . . . . . . . . . . . . 119


5.3.1 Incompressible Green Materials . . . . . . . . . 120
5.3.2 Incompressible Lodge Materials . . . . . . . . 121
5.3.3 Material Stability . . . . . . . . . . . . . . . . 121
5.3.4 Incompressible Material Models of Renown . . 122
5.3.5 Compressible Green Materials . . . . . . . . . 124
5.3.6 Compressible Lodge Materials . . . . . . . . . 126
5.3.7 Compressible Material Model of Renown . . . . 127
5.4 Numerical Implementation . . . . . . . . . . . . . . . . 127
5.4.1 Incompressible Materials . . . . . . . . . . . . 128
5.4.2 Compressible Materials . . . . . . . . . . . . . 131
5.5 Examples . . . . . . . . . . . . . . . . . . . . . . . . . 135
5.5.1 Uniaxial Extension . . . . . . . . . . . . . . . 136
5.5.2 Equi-biaxial Extension . . . . . . . . . . . . . 139
5.5.3 Simple Shear . . . . . . . . . . . . . . . . . . . 142
5.5.4 Homogeneous Planar Membranes . . . . . . . . 145
5.6 Applications . . . . . . . . . . . . . . . . . . . . . . . . 146
5.6.1 Guidelines for Selecting an Elastic Material
Model . . . . . . . . . . . . . . . . . . . . . . 149
5.7 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . 150
5.7.1 Pure Shear . . . . . . . . . . . . . . . . . . . . 150
5.7.2 Biaxial Extension . . . . . . . . . . . . . . . . 150
5.7.3 Extension Followed by Simple Shear . . . . . . 151
5.7.4 Other Problems . . . . . . . . . . . . . . . . . 151

6. Implicit Elasticity 161


6.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . 162
6.1.1 Attempts at Capturing Fungs Law Using
Explicit Elasticity . . . . . . . . . . . . . . . . 165
6.2 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . 167
6.2.1 Dilatoric/Deviatoric Formulation . . . . . . . . 170
6.3 Modulus and Compliance . . . . . . . . . . . . . . . . . 171
6.3.1 Incompressible Materials . . . . . . . . . . . . 173
6.3.2 Compressible Materials . . . . . . . . . . . . . 174
6.3.3 Eulerian Formulations . . . . . . . . . . . . . . 175
6.3.4 Stability . . . . . . . . . . . . . . . . . . . . . 177
6.3.5 Plane-Stress Formulation . . . . . . . . . . . . 177
6.4 Isotropic Materials . . . . . . . . . . . . . . . . . . . . 179
6.4.1 Implicit Hookean Solid . . . . . . . . . . . . . 179
Contents xvii

6.5 Rajagopal Elastic Solids . . . . . . . . . . . . . . . . . 183


6.5.1 Material A . . . . . . . . . . . . . . . . . . . . 184
6.5.2 Material B . . . . . . . . . . . . . . . . . . . . 186
6.5.3 Material C . . . . . . . . . . . . . . . . . . . . 189
6.6 Examples . . . . . . . . . . . . . . . . . . . . . . . . . 190
6.6.1 Uniaxial Extension . . . . . . . . . . . . . . . 190
6.6.2 Equi-biaxial Extension . . . . . . . . . . . . . 192
6.6.3 Simple Shear . . . . . . . . . . . . . . . . . . . 194
6.6.4 Homogeneous Planar Membranes . . . . . . . . 196
6.7 Applications . . . . . . . . . . . . . . . . . . . . . . . . 197
6.8 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . 199
6.8.1 Pure Shear . . . . . . . . . . . . . . . . . . . . 199
6.8.2 Biaxial . . . . . . . . . . . . . . . . . . . . . . 199
6.8.3 Extension Followed by Simple Shear . . . . . . 199
6.8.4 Other Problems . . . . . . . . . . . . . . . . . 199

7. Viscoelasticity 209
7.1 1D Viscoelastic Solid . . . . . . . . . . . . . . . . . . . 210
7.1.1 Clocks . . . . . . . . . . . . . . . . . . . . . . 212
7.1.2 Stress Relaxation Experiment . . . . . . . . . . 212
7.1.3 Causal Deformations . . . . . . . . . . . . . . 214
7.1.4 Memory Kernel Formulation . . . . . . . . . . 215
7.1.5 Additive Strain Formulation . . . . . . . . . . . 216
7.1.6 Quasi-Linear Viscoelasticity . . . . . . . . . . 218
7.2 Viscoelastic Kernels . . . . . . . . . . . . . . . . . . . 220
7.2.1 IOV Kernel . . . . . . . . . . . . . . . . . . . 221
7.2.2 FOV Kernel . . . . . . . . . . . . . . . . . . . 222
7.2.3 BOX Kernel . . . . . . . . . . . . . . . . . . . 229
7.2.4 Implementing a Physical Kernel:
The MCM Kernel . . . . . . . . . . . . . . . . 230
7.3 Additive Strain Fields . . . . . . . . . . . . . . . . . . . 234
7.3.1 Lagrangian Strains . . . . . . . . . . . . . . . . 235
7.3.2 Eulerian Strains . . . . . . . . . . . . . . . . . 236
7.3.3 Field Transfer . . . . . . . . . . . . . . . . . . 237
7.4 K-BKZ Viscoelasticity . . . . . . . . . . . . . . . . . . 238
7.4.1 Viscoelastic Lodge Solid . . . . . . . . . . . . 239
7.4.2 Viscoelastic Green Solid . . . . . . . . . . . . 241
7.4.3 Viscoelastic MooneyRivlin Solid . . . . . . . 242
xviii Contents

7.5 Quasi-Linear Viscoelasticity . . . . . . . . . . . . . . . 243


7.5.1 Guth Strains for Explicit Elastic Solids . . . . . 245
7.5.2 Guth Strains for Implicit Elastic Solids . . . . . 246
7.5.3 Bulk/Shear Split . . . . . . . . . . . . . . . . . 246
7.5.4 Tangent Moduli . . . . . . . . . . . . . . . . . 249
7.6 Examples . . . . . . . . . . . . . . . . . . . . . . . . . 251
7.6.1 Uniaxial Extension . . . . . . . . . . . . . . . 252
7.6.2 Equi-biaxial Extension . . . . . . . . . . . . . 253
7.6.3 Simple Shear . . . . . . . . . . . . . . . . . . . 254
7.6.4 Planar Membranes . . . . . . . . . . . . . . . . 255
7.7 Applications . . . . . . . . . . . . . . . . . . . . . . . . 256
7.7.1 Selecting a Kernel . . . . . . . . . . . . . . . . 256
7.7.2 Selecting a Constitutive Equation . . . . . . . . 264
7.8 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . 265
7.8.1 Pure Shear . . . . . . . . . . . . . . . . . . . . 265
7.8.2 Biaxial Extension . . . . . . . . . . . . . . . . 265
7.8.3 Extension Followed by Simple Shear . . . . . . 266
7.8.4 Other Problems . . . . . . . . . . . . . . . . . 266

Appendix A Linear Algebra 277


A.1 Arrays and Vectors . . . . . . . . . . . . . . . . . . . . 277
A.2 Matrices and Tensors . . . . . . . . . . . . . . . . . . . 279

Appendix B Covariant and Contravariant Issues:


Configuration Physics 289
B.1 Invariant Observer . . . . . . . . . . . . . . . . . . . . 290
B.2 Field Transfer . . . . . . . . . . . . . . . . . . . . . . . 292
B.3 Material Derivatives . . . . . . . . . . . . . . . . . . . 297
B.4 Lie Derivatives . . . . . . . . . . . . . . . . . . . . . . 298
B.4.1 Integration . . . . . . . . . . . . . . . . . . . . 302
B.5 Weighted Field Transfer . . . . . . . . . . . . . . . . . 307

Appendix C Kronecker Products


A Tensor-to-Array Mapping Scheme 309
C.1 The vec Operator . . . . . . . . . . . . . . . . . . . . . 310
C.2 Kronecker Product . . . . . . . . . . . . . . . . . . . . 311
C.3 The ten Operator . . . . . . . . . . . . . . . . . . . . . 312
C.4 Coordinate Transformations . . . . . . . . . . . . . . . 314
Contents xix

Appendix D Solver for a First-Order ODE 317


D.1 Estimate Error . . . . . . . . . . . . . . . . . . . . . . . 319
D.2 IRKS Methods . . . . . . . . . . . . . . . . . . . . . . 320

Appendix E Solver for Convolution Integrals 329

Appendix F Solver for the MittagLeffler Function 341

Bibliography 343
Index 357
Nomenclature

General
B Body
E1 Exponential integral
E, Two-parameter Mittag-Leffler function
H Heaviside step function
P Particle
S Surface
Y Material surface
R2 Two-dimensional subspace of R3
R3 Ambient Euclidean 3-space
 Gamma function
 Configuration
m Machine precision

Scalar Fields
Ai Gage areas, i = 1, 2, 3
d Depth
D Dissipation function
E Youngs modulus
f Imposed experimental force
G Relaxation modulus for shear response
G Relaxation kernel for shear response
h Height
H Relaxation spectrum
Ii Elastic invariants, i = 1, : : : , 10
J Dilation, Jacobian of deformation
K Relaxation kernel for bulk response
` Length
xxi
xxii Nomenclature

`i Gage lengths, i = 1, 2, 3
M Viscoelastic memory function
M Viscoelastic memory kernel
p Hydrostatic pressure
} Lagrange multiplier
t Time
T Temperature
Tg Glass transition temperature
Tm Melting temperature
T Engineering stress
T Viscoelastic tangent modulus
U Internal energy
U Internal energy function
V Gage volume
w Width
W Mechanical work
W Elastic strain-energy function
: Guth strain
Fractional order in viscoelastic kernel
Fung parameter for bulk response
Fungs 1D parameter
Fung parameter for shear response
 Magnitude of shear
Transverse stretch
 Dilatation
 Engineering strain
 Angle of rotation
 Bulk modulus
Lams modulus
Stretch

Areal stretch
Shear modulus
Poissons ratio
Inhomogeneous stretch
 Extent of rigid-body rotation
% Mass density
, & Shear stress
 Characteristic time in viscoelastic kernel
 Proportionality parameter in MooneyRivlin model
Nomenclature xxiii

Material Scalar Fields


dA Differential element of area
dH Differential separation between two neighboring material surfaces
dS Differential separation between two neighboring particles
dV Differential element of volume
dY Difference between parametric values of neighboring material
surfaces
Spatial Scalar Fields
da Differential element of area
dh Differential separation between two neighboring surfaces
ds Differential separation between two neighboring locations
dv Differential element of volume
Vector Fields
{G } Array representation of vector G in basis (e1 , e2 , e3 )
{G}I Ith component of vector {G }
u Infinitesimal displacement vector
Vector Mapping Fields
T Pseudo traction vector
 Motion map
 Inverse motion map
 Motion map for translation
Material Vector Fields
A Acceleration of a particle
dA Normal to differential element of area
dX Vector connecting two neighboring particles
eI , e I Base vector in the Ith material direction
N Normal to a surface
V Velocity of a particle
X Particle coordinates
Spatial Vector Fields
a Acceleration of a particle at a location
b Body force
da Normal to differential element of area
xxiv Nomenclature

df Force of traction acting on an infinitesimal area


dx Vector connecting two neighboring places
ei , e i Base vector in the ith spatial direction
fi Homogeneous force of traction in the ith spatial direction
n Normal to a surface
r Axis of rotation
t Traction vector
v Velocity of a particle at a location
x Location of a particle, its place
! Angular velocity vector

Tensor Fields
F History function for the deformation gradient F
g Riemannian metric of ambient space
[H ] Matrix representation of tensor H in basis (e1 , e2 , e3 )
[H]IJ IJth component of matrix [H ]
ij Kronecker delta
 Infinitesimal strain tensor
! Infinitesimal rotation tensor

Tensor Mapping Fields


I Identity map
dFx Deviatoric change in deformation gradient
f , F 1 Reverse (Eulerian) deformation gradient
F (Lagrangian) deformation gradient
G Infinitesimal displacement gradient
P First PiolaKirchhoff stress
Q Spatial rotation
R Material rotation
RH Hencky rotation
Motion map for homogeneous deformation

Material Tensor Fields


C Green deformation
C 1 Cauchy deformation
Cx Green distortion
Cx 1 Cauchy distortion
dEx Deviatoric change in Green strain
dEx Deviatoric change in Lodge strain
D Lagrangian stretching
Nomenclature xxv

E Green strain
Ex Distortional Green strain
E Lodge strain
Ex Distortional Lodge strain
Ey Truesdell strain
Ey Biot strain
EH Hencky strain
I Identity tensor
L Lagrangian velocity gradient
S Second PiolaKirchhoff stress
Sx Deviatoric second PiolaKirchhoff stress
U Right stretch
W Lagrangian vorticity
Y Lagrangian Guth volumetric strain
Z Lagrangian Guth shear strain
Lagrangian extra stress
Spatial Tensor Fields
b Finger deformation
b1 Piola deformation
d Stretching or strain rate
e Signorini strain
eN Distortional Signorini strain
e Almansi strain
eN Distortional Almansi strain
eO BellEricksen strain
I Identity tensor
l Velocity gradient
s Kirchhoff stress
T Cauchy stress
v Left stretch
w Vorticity
z Eulerian Guth strain
 Eulerian extra stress
Angular velocity
Third-Order Tensor Field
ijk Permutation operator
xxvi Nomenclature

Fourth-Order Material Tensor Fields


A Lagrangian resilience
B Lagrangian Fung adjustment
C Lagrangian tangent compliance
D Lagrangian deviatoric tangent modulus
I Kronecker identity tensor
J Kronecker permutation tensor
M Lagrangian tangent modulus
T Lagrangian viscoelastic tangent modulus
V Lagrangian volumetric tangent modulus

Fourth-Order Spatial Tensor Fields


a Eulerian resilience
b Eulerian Fung adjustment
c Eulerian tangent compliance
m Eulerian tangent modulus
t Eulerian viscoelastic tangent modulus

Greek Alphabet
A () alpha
B (b) beta
 (g) gamma
 (d) delta
"E (e) epsilon
Z (z) zeta
H (Na) eta
# (th) theta
I (Ne) iota
K (k) kappa

(l) lambda
M (m) mu
N (n) nu
 (ks) xi
oO (o) omicron
$ (p) pi
%P (r) rho
& (s) sigma
T (t) tau
 (, oo) upsilon
Nomenclature xxvii

' (f) phi


!X (H) chi
) (ps) psi
! (No) omega

Acronyms
ASTM American society for testing and materials
BOX Box model, a.k.a. the QLV model
BVP Boundary value problem
CCM ColeCole model
CFD Computational fluid dynamics
FE Finite elements
FEA Finite element analysis
FOV Fractional-order viscoelastic
GL General linear
IC Initial condition
IOV Integer-order viscoelastic
IRKS Inherent RungeKutta stable
IVP Initial value problem
K-BKZ KayeBernstein, Kearsley, and Zapas
KWW KohlrauschWilliams and Watts
MCM Maxwell chain model
MPL Modified power law
ODE Ordinary differential equation
QLV Quasi-linear viscoelaticity
PDMS Polydimethylsiloxane
PMMA Polymethyl methacrylate
RK RungeKutta
1D, 2D, 3D Spatial dimensions
Introduction

In his Herbert Spencer lecture delivered at Oxford University on June 10,


1933, Einstein (1933, pp. 1011) said:
It can scarcely be denied that the supreme goal of all theory is to
make the irreducible basic elements as simple and as few as pos-
sible without having to surrender the adequate representation of
a single datum of experience.
In short, Make it as simple as possible, but no simpler!
Your journey will follow down a path mapped out by Einstein in his
lecture. On pp. 1415, he gives his opinion on the simplicity of Nature:
Our experience up to date justifies us in feeling sure that in Na-
ture is actualized the ideal of mathematical simplicity.
In other words, if a mathematical representation of Nature is too complex,
it is likely to be flawed. Earlier in his lecture, on p. 10, he addressed the
separate roles of human thought and physical observation:
Reason gives the structure to the system; the data of experience
and their mutual relations are to correspond exactly to conse-
quences in the theory.
He later embellished upon this statement, on p. 15, by adding:
Experience can of course guide us in our choice of serviceable
mathematical concepts; it cannot possibly be the source from
which they are derived; experience of course remains the sole
criterion of the serviceability of a mathematical construction for
physics, but the true creative principle resides in mathematics.
This was Einsteins contention:
Your mind is your greatest asset!
xxix
xxx Introduction

Overview
The mathematical theory that treats materials composed of discrete atoms
and molecules as if they were a smeared-out continuous medium is a true
triumph of human intellect: continuum mechanics. Our confidence in this
theory has been strengthened through our extensive use of engineering
tools like finite elements (FE) and computational fluid dynamics (CFD).
These are giant applications of continuum mechanics where the physi-
cal laws are satisfied over user-specified meshes created to solve bound-
ary value problems (BVPs) of interest, all in accordance with associated
constitutive theories like elasticity, Newtonian and non-Newtonian fluids,
viscoelasticity, creep, plasticity, viscoplasticity, and others, whose models
are used to describe various material behaviors.
That the nonlinear theories in these disciplines are, at present, neither
simple nor entirely accurate representations of Nature only means that hu-
man reason has not yet come to complete terms with physical reality. Non-
linear material modeling remains a fertile topic for creative thought.
This text is a primer for students who are interested in studying how ma-
terial behavior is modeled in engineering applications using mathematics
in accordance with physical laws. Such knowledge is fundamental to your
understanding of modern FE and CFD outputs. The focus of this textbook
is on ideas and concepts and the needed mathematics necessary to come to
such an understanding. Full mathematical rigor is sacrificed from time to
time so as not to intimidate the inexperienced reader. The intent of this text
is to provide upper-level undergraduate and entry-level graduate students
with a basic skill set needed to comprehend our theoretical approach to the
mechanics of continuous materials so that upon entering the work force
they can converse with their colleagues in an intelligent manner and to also
prepare them for further studies, if they are so inclined.
This book is not so much a textbook on continuum mechanics, as it is an
introduction into how engineers use the continuum framework to construct
mathematical relationships for various classes of materials, illustrated via
the material class known as soft solids.
You can only embark on your journey once you have an appreciation
for where you have been and an inspiration as to where you want to go.
This text will help you attain a compass heading, thereby starting your own
journey down this wondrous pathway to adventure and discovery. Each
journey starts with a single step. Only you can decide if you are going to
take that step. Most who venture this step will be sightseers along the way.
Introduction xxxi

A few of you will become earnest practitioners of the craft. Such is the
outcome of making choices along the timeline of your life.
Notation
Throughout this book, special notations are used as visual aids to help you
discern what type of field you are looking at. A field is any characteristic
of a particle of mass that also exists at every other particle throughout a
body, e.g., its mass density, temperature, or elastic modulus. Fields often
require physical units to give them meaning. The notation adopted here
is close to the notation used by Holzapfel (2000) in his classic text on the
subject, whose origins trace back to the encyclopedic works of Truesdell
and Toupin (1960) and Truesdell and Noll (2004).
A function of a variable is expressed using a script font, e.g., y = F (x).
When Roman fonts are used, a scalar field (or number) appears in an italic
font, e.g., a; a vector field (or array) appears in a slant blackboard bold
font, e.g., a; while a tensor field (or matrix) is expressed in a bold italic
font, e.g., a. When Greek fonts are used, a scalar looks like, e.g., ; a
vector field looks like, e.g., ; and a tensor field looks like, e.g., . Fourth-
order tensor fields are typeset in a bold calligraphy font, e.g., a . They are
not as common. They arise when transforming our theories into forms that
are more readily implemented into software. Notational exceptions are kept
to a minimum and typically arise for historical reasons.
Two frames of reference are commonly used in mechanics: material and
spatial. Both reside within the infinite ambient space that we call our world
or universe. The material frame is often referred to as the reference, initial,
or undeformed frame, while the spatial frame also goes by the name of the
current or deformed frame. A field that associates with a material frame is
typeset in an uppercase character from its respective font set, e.g., S , and is
often referred to as its Lagrangian representation. The same physical field,
when associated with a spatial frame, will be typeset in its lowercase char-
acter, i.e., s, and is referred to as its Eulerian representation. Tensors used
to map between these two frames, which have a footing in both frames,
are typeset in an uppercase, upright, bold font, e.g., F. Notational excep-
tions are kept to a minimum and typically arise for historical reasons, e.g.,
Cauchy stress T does not follow this notational dictate; it is an Eulerian
field expressed in a Lagrangian notation.
When expressing vectors and tensors in component form, it has become
accepted practice to write the indices that associate with material coordi-
nates in uppercase, while the indices that associate with spatial coordinates
xxxii Introduction

are written in lowercase. So, for example, material vectors are written as
A = AI eI , while spatial vectors look like a = ai ei , where, in both cases,
the repeated indices I and i are summed from 1 to 3 in accordance with Ein-
P
steins summation convention, e.g., a = a1 e1 + a2 e2 + a3 e3 = 3i=1 ai ei
is written in the shorthand notation of Einstein as a = ai ei , where the sum
over i is implicitly implied. Tensors, on the other hand, would be expressed
in material coordinates as A = AIJ eI eJ and in spatial coordinates like
a = aij ei e j , where denotes a dyadic product. Components are type-
set in an upright san serif font. Components of Greek fields are typeset
in an upright Greek font, for example,  = i ei or = IJ eI eJ .
By this notation one means that the components, say AI of A, are evalu-
ated in a coordinate system specified by base vectors eI . Throughout this
text it is considered that ei and eI are coincident base vectors, specifically,
ei = I eI = Ii eI and eI = I ei = Ii ei , where Ii and Ii denote the Kronecker
delta function: 1 if I = i, otherwise 0.
Cartesian tensors are used throughout this text, as the metric of ambient
space is g = ij e i e j . Ambient space is said to be flat. Cartesian tensors
have three possible indical positions. Using indices c and r to denote col-
umn and row, the matrix notation of a tensor, say a, includes arc , arc , and
arc , with its transpose aT having matrix components of reversed order, viz.,
acr , acr , and acr . Vectors can have components of ar and ar with transposed
components of ac and ac .
Additional points on notation include the following: whenever the iden-
tity tensor has mixed indices, viz., one is a material index and the other is
a spatial index, then it is displayed as either I = Ii ei e I or I = Ii eI e i ;
otherwise, it is displayed as either I = ij e i e j , I = ij ei e j , or as
I = ij ei e j . In Cartesian tensor analysis, which is the form of tensor
analysis used in this text, the contravariant base vectors ei  eI are coaxial
with the covariant base vectors e i  e I , which is not true of general tensor
analysis, e.g., cf. Holzapfel (2000) and Sokolnikoff (1964).
The notational feature of using subscripts and superscripts to desig-
nate between covariant and contravariant component indices, respectively,
is adopted; it is used in general tensor analysis.1 Covariant/contravariant
properties of a tensor field are plain when written in the Lagrangian frame;

1 The author did not use subscripts and superscripts as a means to distinguish
between covariant and contravariant components in the earlier drafts of his text,
as is common practice among texts that use Cartesian tensors. It was a request
from his students that this notational enhancement be incorporated.
Introduction xxxiii

however, ambiguities can and do arise when mapping Lagrangian fields


into Eulerian fields whenever Cartesian tensors are used, as they are here.
These mappings are not always one to one. For this reason, there are Eu-
lerian fields (e.g., strain rate) that can result from either a covariant or a
contravariant mapping from the Lagrangian domain. The use of either
subscripts or superscripts, applied to the same field, can occur in Eulerian
expressions. Because of this loss of tensor quality in Eulerian analysis, this
author prefers to develop his ideas in the Lagrangian frame and then map
them into the Eulerian frame where we can use them in analysis.
A Look Ahead
Before embarking on your adventure, you ought to read through Appendix
A and become familiar with the content and mathematical results presented
there. This will be your toolbox. Maybe you do not know or understand,
right now, all that is found therein, but you should by the end of this course.
Your formal study of the mechanics of continuous media begins with
the topic of kinematics discussed in Chap. 1. This is the study of mo-
tion. In this chapter, seven BVPs are presented that are used in laboratories
around the world to understand and characterize materials. These BVPs
will be followed and studied throughout this text. They are uniaxial ex-
tension, equi-biaxial extension, simple shear, homogeneous planar mem-
branes, pure shear, biaxial extension, and an axial extension followed by a
simple shear. This last experiment is the juxtaposition of two of the previ-
ously studied experiments. For some of the fields that you will be studying,
the outcome will be a straightforward linear superposition, but in other cal-
culations, it will not. In these cases, the history of deformation becomes
important. From a study of this last BVP over the duration of this course,
you will gain an appreciation of both the power and potential limitations of
linear superposition.
The next chapter, Chap. 2, addresses deformation and how it is quanti-
fied. Here the author builds upon your classic training in the mechanics of
motion by considering the fact that the shape of an object can change over
time, too, i.e., move relative to itself. After the deformation gradient has
been introduced in Sect. 2.3, the instructor ought to proceed to Appendix
B before finishing out Chap. 2. Appendix B is vital to your understand-
ing of the mechanics of continuous media, and although the rigors of the
mathematics and physics that underlie this appendix will likely be beyond
your capabilities, gaining an intuitive understanding of what this appendix
is all about is essential for you to grasp and comprehend the influence that
xxxiv Introduction

motion and deformation have on the mathematics needed to describe them.


The remainder of Chap. 2 introduces a number of different fields that have
value when quantifying various aspects of a deformation. Also included
are formul that relate these finite deformation fields to the commonplace
infinitesimal strain and rotation fields used in classical elasticity theory.
At this point, the author introduces the fields used to describe and quan-
tify material behavior. The first such field is strain, which is discussed in
some detail in Chap. 3. Unlike stretch, which is defined in the prior chapter,
strain is not unique. Two admissible strain fields are put forward. The first
is a measure of the change in separation between two neighboring mate-
rial particles. The second is a measure of the change in separation between
two, neighboring, material surfaces that, under isochoric conditions, equate
with the change in area of a material surface. Both strain measures have
strong ties to thermodynamics and statistical mechanics. Both are Rie-
mannian descriptors of geometric change, and both are used in the latter
chapters where material models are presented.
Chapter 4 finishes our discussion of the physical fields used in the study
of deformable continua. Here the idea of stress is introduced. Strain is a
mechanical effect that is caused by stress; stress is a mechanical effect that
is caused by applied loads. Several definitions for stress are encountered in
applications, the most commonly used of which are discussed in some de-
tail in this chapter, along with explanations of when one definition of stress
ought to be selected over another. The chapter wraps up with a discussion
of material stability, and how stress and strain can affect stability, which is
a mathematical property, not a physical property.
At this juncture, you should be prepared to understand how materials
behave under various conditions. Physical laws have been addressed from
time to time throughout the text, as the need arises for you to understand
a concept being presented, but physical laws cannot tell us how a mate-
rial will respond to a given input. Constitutive equations do that, and al-
though physical laws constrain what one can propose as a constitutive law,
they cannot establish these laws. Constitutive theory is an important topic
within the discipline of continuum mechanics. Three constitutive theories
are presented in this text; no universal constitutive law exists. Each class of
materials is described by its own constitutive law, which must be discov-
ered by, and quantified through, experiments.
Two different classifications for elastic materials are presented. The
first is a stress/strain theory, which finds application with rubbers, elas-
tomers, and synthetic polymers and is addressed in Chap. 5. The second is
Introduction xxxv

a stress-rate/strain-rate theory, which finds applications with hard materi-


als, some soft synthetic solids, and soft biological tissues, and is addressed
in Chap. 6. The implicit theory of tissue elasticity is much more broad
and inclusive than the explicit theory of rubber elasticity. Because of the
inadequacies of the theory of rubber elasticity to satisfactorily describe bi-
ological tissues, a more general theory was sought to describe tissue elas-
ticity. These theories arise from explicit and implicit functions of state,
respectively, that are used to describe the elastic strain energy of a ma-
terial, which is an energy stored internally by the stretching of chemical
bonds within a material. Simple models are constructed for both theories
and contrasted against experimental data. A technique by which these mod-
els can be transformed into array/matrix equations suitable for numerical
analysis is presented in Appendix C, and a numerical method suitable for
solving the resulting ODEs is presented in Appendix D.
The final chapter, Chap. 7, addresses the fact that materials typically
dissipate energy as heat during their deformations (elastic solids, by defi-
nition, do not dissipate energy), which affects their mechanical response.
Both classes of elastic solids are analytically continued into a viscoelastic
domain in this chapter, and applications are considered. Two qualities of a
viscoelastic material are discussed. The first addresses how these materials
relax, and a variety of relaxation functions are introduced and discussed.
The second addresses the tensorial nature of a viscoelastic material, which
is the by-product of an analytic continuation procedure. Viscoelasticity
is introduced as a Volterra integral equation of the second kind. A nu-
merical method for solving such convolution integrals is presented in Ap-
pendix E. Appendix F provides an algorithm for computing the Mittag-
Leffler function, a relaxation function derived from statistical mechanics.
The focus of this text is on viscoelasticity as a material model, rather than
viscoelasticity as a mechanical system. See, e.g., the textbooks by Lakes
(1998), Mainardi (2010), and Wineman and Rajagopal (2000) if you are in-
terested in more of a solid-mechanics/mechanical-engineering approach to
viscoelasticity.
Viscoelasticity is one of the simpler constitutive theories for describ-
ing dissipating materials. Other theories include, for fluids, Newtonian and
non-Newtonian fluid mechanics and, for crystalline solids, creep, plasticity,
and viscoplasticity. Other topics in mechanics like composites, porous me-
dia (mixtures), fracture, and damage are not discussed here, either, but are
equally important. Constitutive theory, as a discipline, is vast and interest-
ing. This textbook offers only an introductory glimpse into this fascinating
xxxvi Introduction

field, which resides at an intersection between engineering, physics,


applied mathematics, and, more recently, biology and medicine.
Noticeably missing from this text is the linear (or classic) theory of
elasticity, although finite-strain Hookean theories are derived. The author
excluded the linear theory of elasticity on purpose, as it has been his ex-
perience that when young engineers are first schooled in linear elasticity,
they are prone to make potentially grave errors in judgment when extending
their linear intuition and understanding of hard solids in their attempts to
describe nonlinear soft-solid behavior. By first studying finite deformation
analysis, instead of infinitesimal deformation analysis, and by first studying
nonlinear material behavior, instead of linear material behavior, the author
hopes that many of these bad habits will not become ingrained in the minds
of young practicing engineers. Many academics would argue that finite de-
formation theory is too difficult to learn first. I argue that, in most aspects,
finite deformation analysis is no more difficult to understand than the in-
finitesimal case, especially for incompressible isotropic materials and, in
many regards, finite analysis is simpler once one has grasped the notion of
configurations and the various kinds of maps that can exist between them,
which is why their discussion is so prominent in the presentation of this
textbook.
PART 1
Continuum Fields
More physics, Al, less mathematics.

From the last conversation between the author


and his beloved mentor, Arthur S. Lodge.
These words continue to sear my conscience
with an intensity as if spoken yesterday.

The student has listened!


Chapter 1

Kinematics

Kinematics is the study of motion. This text addresses the kinematics of an


idealized material called a continuum, also referred to as a body.
Consider a body B occupying a region  in ambient space R3 at time t.
This body is comprised of an infinite set of particles P , written as B = {P },
each possessing a unique location x with coordinates (x1 , x2 , x3 ) quantified
in a right-handed Cartesian coordinate system drawn in R3 , with an origin
O and base vectors ei , i = 1, 2, 3, as shown in Fig. 1.1.
Matter is composed of discrete atoms and molecules. Every particle
P in body B is considered to be an ensemble of atoms/molecules whose
nano-sized volume is large enough so that the response of the ensembles
average is taken to be the response of P . Yet the nano-sized volume that is
P is small enough that gradients across it can be neglected without intro-
ducing measurable error. Conceptually, this is what is meant by the notion
of a particle in a continuum.
At some prior moment in time t0 , body B occupied region 0 in R3 .
This embedding of body B in space R3 at some past time t0 is referred to
as the reference configuration and associates with its undeformed shape,
whereas its occupation of region  in R3 at current time t, t  t0 , denotes
the current configuration, which is said to be its deformed shape. In its
undeformed configuration 0 at time t0 , particle P was located at place
X with coordinates (X1 , X2 , X3 ) using coordinate axes eI , I = 1, 2, 3, that,
in this text, are taken to be coaxial with the coordinate directions ei used
above to locate P in  at time t, as seen in Fig. 1.1.
Because time is relative, reference time t0 is usually set to zero, i.e., t0 =
0. The notations of 0 and t0 for reference time are used interchangeably
throughout this text.

A.D. Freed, Soft Solids: A Primer to the Theoretical Mechanics of Materials, 5


Modeling and Simulation in Science, Engineering and Technology,
DOI 10.1007/978-3-319-03551-2_1, Springer International Publishing Switzerland 2014
6 Soft Solids

X x @ t
@ t0
e2
0
e1
e3

Fig. 1.1 Body B is deformed from its original shape at time t0 , i.e., from cong-
uration 0 , into a nal deformed shape at time t, viz., into conguration . Par-
ticle P was located at X = (X1 , X2 , X3 ) in 0 and is now located at x = (x1 , x2 , x3 )
in , quantied in a common Cartesian basis (e1 , e2 , e3 ) in R3

1.1 Motion

Assertion 1.1. Law of Continuous Media: The motion of every particle P


within body B moves through an ambient space R3 with coordinate frame
(e1 , e2 , e3 ) and, in so doing, obeys a pair of invertible maps
x = (X , t) = i (X1 , X2 , X3 , t) ei ,
(1.1)
X =  (x, t) = I (x1 , x2 , x3 , t) eI
where map  is the inverse mapping to map . Motions  and  are
considered to be smooth, in the sense that they are both continuous and
dierentiable (at least once) functions of their positional arguments.
The first formula in Eq. (1.1) uses material coordinates X for its inde-
pendent positional variables (the spatial coordinates x are the dependent
variables), which constitutes a Lagrangian description for the motion. In
contrast, the second formula uses spatial coordinates x as its independent
positional variables (the material coordinates X are now the dependent
variables), which constitutes an Eulerian description for the motion.1
1 Lagrangian formulations were studied before Lagrange by Euler, and Eulerian
formulations were used before Euler by DAlembert (private communication with
Prof. K.R. Rajagopal, 2011). The incorrect attribution of an idea to one person,
which actually belongs to another, pervades our literature and is made rampant by
authors propagating these inconsistencies without taking the time to actually read
the literature for themselves (Rajagopal 2011b). Throughout this text the author
takes exception, from time to time, to commonly held terminologies whenever
Kinematics 7

In a material description, attention is paid to a particle, and we


observe what happens to the particle as it moves. : : : In the
spatial description, attention is paid to a point in space, and we
study what happens at the point as time changes. (Holzapfel
2000, p. 60)
A subtle but important implication from the law of continuous media is
that spatial coordinates locate a particle along its path of trajectory through
ambient space and, therefore, depend upon time, i.e., x(t), whereas mate-
rial coordinates are fixed and are, therefore, independent of time, viz., X .
A dependence upon time that has been introduced into a motion by map 
is removed by its inverse map  .
Lagrangian formulations are usually preferred for describing solids
where a reference configuration can be selected, whereas Eulerian formu-
lations are usually preferred for describing fluids where a reference config-
uration has no physical meaning. The fields of interest in this text will be
described in terms of both formulations.

1.1.1 Homogeneous Motions


The experiments of this text belong to a special subclass of admissible
motions, viz., those that are homogeneous. A motion is said to be homog-
eneous if it can be described by the vector equation
(t)] {X } + {(t)}
{x(t)} = {(X , t)} = [ (1.2)
or in component notation as
xi = Ii XI + i (1.3)
whose inverse motion is, therefore, described by
1 (t)] {x(t) (t)}
{X } = { (x, t)} = [ (1.4)
or in component notation as
 
XI = [1 ]Ii x i i (1.5)
where is a matrix describing the distortion of a motion, while  is a
vector describing the translation of that motion. Existence of the map 
ensures that matrix is invertible, i.e., not singular. Matrix (t) and vector
(t) are functions of time only. If either depends on a spatial variable, be
such inconsistencies, in his mind, seem to exist. This author is by no means a
historian, just a person who has a passion for the history of his beloved science,
but whose thirst is throttled back by his dyslexia.
8 Soft Solids

it x or X , then the motion is said to be inhomogeneous. Matrix has


components Ii , while matrix 1 has components [1 ]Ii . The superscript
index associates with the row, while the subscript index associates with the
column.

1.2 Velocity and Acceleration

The velocity and acceleration fields belonging to a particle P as it moves


through space R3 over time t are obtained by dierentiating the motion 
of that particle P at a fixed material location X so that

P , t) = @(X , t) = @ (X , t) eI ,
I
V (X , t) = x(X
P , t) = (X (1.6)
@t @t
@2 (X , t) @2 I (X , t)
A(X , t) = x(X
R , t) = (X
R , t) = = eI (1.7)
@t2 @t2
where V (X , t) and A(X , t) are material descriptions for the velocity and
acceleration of particle P . Notations P and R mean @  /@t and @2  /@t2 ,
respectively, for any field . Velocity and acceleration are contravariant
vector fields, i.e., they are physical fields described by tangent vectors to
curves of trajectory in ambient space; cf. Appendix B.
Deriving the spatial descriptions for the velocity and acceleration of a
particle as it moves through space R3 requires a bit more care; specifically,
 
@  (x, t), t
v(x, t) =  V (X , t), (1.8)
 @t 
@2   (x, t), t Dv(x, t)
a(x, t) = =
@t 2 Dt
@v(x, t) @v(x, t) @x
= +  (1.9)
@t @x @t
= vP (x, t) + grad v(x, t)  v(x, t)  A(X , t)
where Dv/Dt is called the material derivative of the velocity field. This
time derivative follows the motion of a material particle whenever the
operand is expressed as a spatial field. The acceleration vector a follows
from an application of the chain rule. The first term, i.e., vP = @v/@t, de-
scribes the local acceleration of particle P , while the second term, viz.,
grad(v)  v = (@v/@x)  v = (@v i /@x j )v j ei , describes the convective acceler-
ation being experienced by particle P ; cf. Appendix B.
Kinematics 9

A motion is said to be planar if its velocity can be described by


v(x1 , x2 , t), i.e., if a coordinate frame can be chosen such that v3 = 0. Sim-
ple shear is the only planar motion studied in this text. Although all mo-
tions studied here belong to a class of deformations known as plane stress,
most are not truly planar motions because their thickness varies over the
motion, i.e., v3 0, except for simple shear.

1.3 Examples

Throughout this book, seven BVPs are examined that can be used to per-
form experiments for the purpose of material characterization. Each chap-
ter will build upon the prior ones by incrementally extending analysis to
include the topics covered in the most recent chapter, thereby increasing
your understanding of these important BVPs as we progress.
The four cases of simple extension, equi-biaxial extension, simple
shear, and homogeneous planar membranes will be worked out as exam-
ples in the text, while the three cases of pure shear, biaxial extension, and
axial extension followed by simple shear are left as exercises for you to
develop your skills. None of these BVPs require curvilinear tensor anal-
ysis. These important but more challenging BVPs are relegated to high-
er-level courses; e.g., cf. the textbooks of Bird et al. (1987a), Bird et al.
(1987b), Lodge (1974), Malvern (1969), Nicholson (2008), Truesdell and
Noll (2004), and Truesdell and Toupin (1960).
The targeted materials for application in this book include natural and
synthetic rubbers, elastomers, soft polymers, and soft biological tissues.
Collectively, they are referred to as soft solids. Although some materials
belonging to this class are anisotropic, a large percentage are mechanically
isotropic (or nearly isotropic) in that their stiness does not depend upon
direction even though their microstructures may be highly organized. An
assumption of material isotropy is imposed throughout this introductory
text. Anisotropy is left for advanced graduate study.
Another assumption frequently imposed throughout this book is that
the materials of interest are (or are nearly) incompressible. Conceptually,
any material whose bulk modulus greatly exceeds its shear modulus (typ-
ically by a factor of one hundred or more, like most soft solids, except,
e.g., foams and lung parenchyma) can be treated as being incompressible
in a mathematical sense even though no material is truly incompressible
in a physical sense. Consequently, for the most part, motions considered
10 Soft Solids

in this text are constructed to be isochoric, i.e., volume preserving in a


mathematical sense.
The restrictions of considering incompressible, isotropic, soft solids
greatly simplify the mathematical treatment and ensuing material models
that are put forward in this text. Soft solids constitute a relatively simple
class of material models from which you can learn about actual material
behavior without falling into many of the misconceptions and traps that
often arise when a student first learns linear elasticity and then attempts to
extrapolate that understanding into the domain of nonlinear material behav-
ior, often with disastrous results. In short, this author advocates learning
about nonlinear materials and their behaviors before studying the classic
discipline of linear elasticity.
Two experimental classifications are addressed in this book: shear-free
motions and shear motions; cf. Lodge (1974, pp. 8184). The motion of
a material element is said to be shear free if the principle directions (i.e.,
the eigenvectors) of the deformation do not rotate within the body over
the history of its motion, starting at some reference time t0 and ending
at current time t for any t  t0 . All other motions are said to be shear-
ing motions. A general shear-free motion is depicted in Fig. 1.2. Sim-
ple extension, equi-biaxial extension, pure shear, and biaxial extension are
each shear-free motions, whereas simple shear, homogeneous planar mem-
branes, and extension followed by simple shear are each shearing motions.

1.3.1 Uniaxial Extension


Without a doubt, uniaxial extension is the single most important experi-
ment done on materials for the purpose of their characterization. An iso-
choric uniaxial extension of an isotropic incompressible material has a mo-
tion that is described by
1 1
x1 = X 1 , x2 = /2 X2 , x3 = /2 X3 (1.10)

whose inverse motion is readily determined to be


1 1
X1 = 1 x1 , X2 = /2 x2 , X3 = /2 x3 (1.11)

where (t) = `(t)/`0 is the stretch, `0 = `(t0 ) is the gage length, and `(t)
is the current length of extension; therefore, stretch is normalized so that
(t0 ) = 1. This shear-free motion is a special case of the deformation
illustrated in Fig. 1.2, which is redrawn in Fig. 1.3.
Kinematics 11

l1 (t)

l1 (t0)

l3 (t0)
l2 (t)

e2 e2
e1 e1
l2 (t0)
e3 e3
l3 (t)

t0
t

Fig. 1.2 Shear-free motions (Lodge 1974) take an elemental cube and deform it
into a rectangular prism. Such deformations are quantied by their three prin-
ciple stretches: 1 = `1 (t)/`1 (t0 ), 2 = `2 (t)/`2 (t0 ), and 3 = `3 (t)/`3 (t0 ). These
stretches associate with an isochoric deformation, i.e., volume preserving, when-
ever 1 2 3 = 1. Lengths `1 (t0 ), `2 (t0 ), and `3 (t0 ) are the gage lengths in their
respective 1-, 2-, and 3-directions

The velocity and acceleration vectors for this motion are determined

 
to be
P 1
X
P 2 ,
{V }  {v} = 12 3/2 X (1.12)

 
1 3/2 P
X3 2
R 1
X
{A}  {a} = 1 3/2  3 1 
P 2 R X2 (1.13)
1 3/2  3 1 
2 2
P 2 R X3
2 2

where the notation {v} means components v i of vector v are being as-
signed in a coordinate basis of ei , i = 1, 2, 3, as drawn in Fig. 1.3.

1.3.2 Equi-biaxial Extension


An equi-biaxial experiment simultaneously stretches a material in two or-
thogonal directions by the same amount and at the same rate in both di-
rections, viz., setting  1 = `1 (t)/`1 (t0 ) and  2 = `2 (t)/`2 (t0 ),
as established in Fig. 1.2, from which we arrive at the motion drawn in
Fig. 1.4. The isochoric equi-biaxial extension of an isotropic incompress-
12 Soft Solids

w
d

w0

d0 l

l0

t0 e1 t
e3

e2

Fig. 1.3 The uniaxial extension of a cube into a rectangular prism whose gage
section in the 1-direction stretches from a length of `0 to `, with w and d denoting
width and depth. p Assigning thepprincipal stretch as = `/`0 , constancy of volume
requires w = w0 / and d = d0 /

ible material has a shear-free motion described by


x1 = X 1 , x2 = X 2 , x3 = 2 X3 (1.14)
whose inverse motion is given by
X1 = 1 x1 , X2 = 1 x2 , X3 = 2 x3 , (1.15)


implying that velocity and acceleration obey
P 1
X

{V }  {v} = P 2
X , (1.16)


2 X3
3 P
R 1
X
R 2
 X
{A}  {a} = (1.17)

2 3 P 2 R X3
3 1

for that particle P whose original location was at place (X1 , X2 , X3 ).

1.3.3 Simple Shear


Simple shear is a fundamentally dierent experiment from the previous
two experiments and from the pure-shear problem to follow. In simple
Kinematics 13

w
d

w0

d0
l

l0

t0 e1 t
e3

e2

Fig. 1.4 The equi-biaxial extension of a cube into a rectangular prism whose
gage section in the 12 plane stretches from area `0 w0 to area `w such that
 1 = `/`0 and  2 = w/w0 where, from constancy of volume, d = d0 / 2

shear, the eigenvectors of the deformation rotate in the body because of the
deformation (Lodge 1964), whereas in the previous two experiments, or
any other shear-free extension, these eigenvectors remain aligned over the
deformation history (Lodge 1974). This is one reason why the simple-shear
experiment is so important, yet it is seldom performed on solids.
Simple shear has a planar shearing motion, i.e., v3 = 0, described by
x1 = X 1 +  X 2 , x2 = X 2 , x3 = X 3 (1.18)
whose inverse motion is given by
X 1 = x1  x2 , X 2 = x2 , X 3 = x3 (1.19)
with  (t) being the magnitude of shear, as seen in Fig. 1.5, where  /2 is
often referred to as the shear strain, initialized so that  (t0 ) = 0. This
motion has a velocity and acceleration of
P X 
2
{V }  {v} = 0 , (1.20)
R X 
0
2
{A}  {a} = 0 . (1.21)
0
Simple shear is planar, viz., v3 = 0. This deformation is isochoric indepen-
dent of whether the material being sheared is incompressible or not.
14 Soft Solids

h
e2 e2
e1 e1

t0 t

Fig. 1.5 Simple shear takes a square and deforms it into a parallelogram of the
same area by shearing surfaces X2 = constant in the 1-direction by an amount
x1 =  X2 , relative to the bottom surface X2 = 0 where x1 = X1 8 t

1.3.4 Homogeneous Planar Membranes


All motions that are studied in this text are special cases of the planar mem-
brane. General membrane experiments are sometimes used to characterize
tissues (Sacks 2000). Dierent definitions for the shear displacements are
used here from those that are employed therein and elsewhere, e.g., Freed
et al. (2010) and Humphrey (2002a, Sect. 5.2.2).
A membrane is said to be planar whenever it is flat, i.e., its surface has
no curvature. A planar membrane comprised of an isotropic incompress-
ible material has a homogeneous isochoric motion described by2
x1 = 1 X 1 +  1 2 X 2 , x2 =  2 1 X 1 + 2 X 2 , x3 =
1 X3 (1.22)
where 1 and 2 are stretches in the 1- and 2-directions, respectively, while
1 and 2 are their associated shears, as illustrated in Fig. 1.6. Ensuring
that this motion is isochoric requires that
 

= 1 2 1 1 2 , (1.23)
which is the areal stretch, i.e., a ratio of the deformed area to the initial
area for any representative volume element during a homogeneous motion
of a planar membrane.

2 Inthe literature, e.g., Sacks (2000) and Freed et al. (2010), 1 2 is commonly
written as 1 and 2 1 is written as 2 . Choosing the description that we did here
means that 1 and 2 retain their physical interpretation of being magnitudes
of shear that is otherwise lost with the choice of 1 and 2 , which embed 2 and
1 within them. Mathematically, nothing is wrong with choosing 1 and 2 . It is
in their physical interpretation that confusion can arise.
Kinematics 15

( 1 + 1 2 , 2 + 2 1)
2
(1 2 , 2 )

e2

( 1 , 2 1)

e1 1

Fig. 1.6 A planar membrane is deformed homogeneously from a square into a


parallelogram by stretches 1 and 2 and by shears 1 and 2 in the 1- and 2-
directions, respectively, according to the motion described in Eq. (1.22). The
shear deformations 1 and 2 are taken to be positive in the sense that they are
drawn here

 
The inverse of this motion is gotten by inverting the system of equations
2 3
x1 1 1 2 0 X1
x2 = 4  2 1 2 0 5 X 2 , (1.24)
x3 0 0
1 X 3

 
yielding [cf. Eqs. (A.53)(A.56)]
2 3
X1 2 1 2 0 x1
1
X2 = 42 1 1 0 5 x2 (1.25)

X3 0 0
2 x3

where the matrices in these equations are the [] and [


1 ] matrices in
Eqs. (1.2) and (1.4) describing the distortion of a homogeneous motion.
Observe that
is the determinant of the upper-left 2  2 submatrix in
Eq. (1.24). From Eq. (1.25), the inverse motion of a homogeneous planar
membrane is uniquely described by the formul
2 1 2 2 1 1
X1 = x1 x2 , X2 = x1 + x2 , X3 =
x3 . (1.26)



P 
The motion of a planar membrane has a velocity and acceleration of

1 X1 + 1P 2 X2
{V }  {v} = 2P 1 X1 + P 2 X2 , (1.27)

2
X
P 3
16 Soft Solids

 
R 1 X1 + 1R 2 X2
{A}  {a} = 2R 1 X1 + R 2 X2 (1.28)

2 2
1
P 2
R X3
for that particle P whose original location was at place (X1 , X2 , X3 ). Ob-
viously, the motion of a planar membrane is not planar in the sense that
v3 0; it is planar in the sense of being flat. By 1P 2 , e.g., we mean
P1 2 + 1 P 2 , i.e., the product rule of dierentiation is applied.

1.3.4.1 2D Quad Element


Membrane analysis provides an example through which one can derive the
simplest of finite elements: the two-dimensional (2D) quadrilateral ele-
mentthe quad element. A study of planar membranes enables a bridge
of understanding to be built between the discipline that you are studying,
viz., the mechanics of continuous materials, and the engineering discipline
known as structural mechanics where FEA is king in modern-day practice.
The linear 2D quad element of finite elements is not restricted to
homogeneous deformations. This FE membrane element serves as an ex-
ample of an inhomogeneous motion. It is described by
x1 = 1 X 1 +  1 2 X 2 + 1 X 1 X 2 ,
x2 =  2 1 X 1 + 2 X 2 + 2 X 1 X 2 , (1.29)
1
x3 = X3
where 1 and 2 are the homogeneous stretches in the 1- and 2-directions,
respectively, while 1 and 2 are the inhomogeneous stretches aligned with
these directions, and 1 and 2 are their associated shears, as illustrated in
Fig. 1.7. Ensuring that this motion is isochoric requires that
     
= 1 2 1  1  2 + 1 2  2 1 X 1 + 2 1  1 2 X 2 , (1.30)
which is the areal stretch of a linear planar membrane. This motion can


also be written in matrix form as
x1
2
1 + 1 X 2  1 2 + 1 X 1 0
3 X 
1
x2 = 42 1 + 2 X2 2 + 2 X1 0 5 X2 (1.31)
x3 0 0 1 X3
where inhomogeneity manifests itself through the presence of X1 and X2
in the matrix term.
Kinematics 17

( 1 + 1 2 + 1 ,
2 + 2 1 + 2 )
2 ( 1 2 , 2 )

( 1 + 1 2 ,
2 + 2 1 )
e2

( 1, 2 1 )

e1 1

Fig. 1.7 A planar membrane is deformed from a square into a quadrilateral by


homogeneous stretches 1 and 2 , by inhomogeneous stretches 1 and 2 , and
by shears 1 and 2 , all in the 1- and 2-directions, respectively, according to the
motion described in Eq. (1.29). The shear deformations 1 and 2 are taken to
be positive in the sense that they are drawn here. The homogeneously deformed
quadrilateral (i.e., 1 = 2 = 0) is drawn as a dashed parallelogram

w0
d
w0
d0
l
l0

t0 e1 t
e3

e2

Fig. 1.8 The pure-shear extension of a rectangular prism into a dierent rect-
angular prism whose gage section in the 12 plane stretches from an area of `0 w0
into an area of `w0 such that  1 = `/`0 and 2 = 1, while d = d0 / from the
constancy of volume

1.4 Exercises

In this section, and in like sections throughout this book, you will find
problems that your instructor may choose to assign as homework problems.
18 Soft Solids

Shear Strain

ln 0 ln

Normal
Strain

Fig. 1.9 Mohrs circle in true strain for the pure-shear experiment

1.4.1 Pure Shear


Pure-shear experiments are done on soft solids in uniaxial test systems
where the aspect ratio of the specimen (width to height between grips) is
quite large, typically greater than three in the final deformed state, with
the grips imposing a lateral kinematic constraint to enforce x2  X2 , as
illustrated in Fig. 1.8. The greater this aspect ratio, the better the kinematic
constraint from gripping will approximate the idealized motion being pre-
scribed here.
A pure-shear deformation is defined by a Mohrs circle centered at
the origin with equally sized inner circles: one in the negative half plane
and the other in the positive half plane (cf. any introductory textbook on
strength of materials). The isochoric experiment considered here for an
isotropic incompressible material produces a Mohrs circle for pure shear
in strain, provided that it is expressed in terms of the true strain of Hencky
(1928), as shown in Fig. 1.9. Pure shear, so defined, associates with the
motion
x1 = X 1 , x2 = X 2 , x3 = 1 X3 (1.32)
whose inverse motion is given by
X1 = 1 x1 , X 2 = x2 , X 3 = x3 . (1.33)
This deformation belongs to the class of shear-free motions displayed in
Fig. 1.2, i.e., the triad of eigenvectors orienting deformation remains fixed
in a material over the duration of the experiment.
Pure shear is not a shearing motion!
Kinematics 19

1.4.1.1 Thin-Walled Tubes


The motion described in Eq. (1.32) also represents the inflation of a
thin-walled tube held to a fixed length. Here = r/R where R and r are
the radii to its mid-surface in the reference and current states, respectively.
In this setting, the 1-direction associates with the circumferential direction,
while the 2-direction aligns along the axis of the tube, and the 3-direction
associates with the radial or cross-wall direction. By being thin walled,
we mean that the radial eects due to curvature can be neglected, making
the problem essentially planar. This is a somewhat crude yet reasonable
first-order approximation for boundary conditions describing the descend-
ing aorta.
Rule of Thumb: The thin-walled assumption can be applied whenever a
ratio of the radius to wall thickness exceeds about ten for linear responses
or about seven for nonlinear responses that strain soften.

1.4.1.2 Problems
Another way to imagine pure shear is to rotate a square by 45 so that
the coordinate axes go through the four corners of the square. Now, pull
on a pair of opposing corners while allowing the two adjacent corners to
contract by an amount that preserves area. Obviously, this coordinate frame
does not rotate over the motion. In what plane of the motion described by
Eq. (1.32), as drawn in Fig. 1.8, does this visual image apply?
Derive the components of the velocity V and acceleration A vectors
that describe pure shear.

1.4.2 Biaxial Extension


Biaxial extension is a generalization of equi-biaxial extension. The spec-
imen is still stretched in opposing orthogonal directions, but with distinct
stretches in that 1 = `1 (t)/`1 (t0 ) and 2 = `2 (t)/`2 (t0 ) are now dierent,
as drawn in Fig. 1.10.
The isochoric biaxial extension of an isotropic incompressible material
has a shear-free motion described by
x1 = 1 X 1 , x2 = 2 X 2 , x3 = 1 1
1 2 X3 (1.34)
whose inverse motion is given by
X1 = 1
1 x1 , X2 = 1
2 x2 , X 3 = 1 2 x3 . (1.35)
20 Soft Solids

w
d

w0

d0
l

l0

t0 e1 t
e3

e2

Fig. 1.10 The biaxial extension of a cube into a rectangular prism whose gage
section in the 12 plane stretches from area `0 w0 to area `w with 1 = `/`0 and
2 = w/w0 where, from constancy of volume, d = d0 /( 1 2 )

1.4.2.1 Problem
Derive the components of the velocity V and acceleration A vectors that
describe biaxial extension. Show that they reduce to the velocity and ac-
celeration vectors of equi-biaxial extension given in Eqs. (1.16) and (1.17)
for the special case where 1 = 2 .

1.4.3 Extension Followed by Simple Shear


When performing simple-shear experiments on membranes, it is often use-
ful to prestretch a test sample prior to imposing a simple shearing onto it
in order to keep the whole membrane under a state of tension so as to al-
leviate the onset of wrinkling. In a certain sense, the initial deformation
caused by an axial extension imposed in, say, the 2-direction serves as an
intermediate state of reference onto which a new motion of simple shear
is to be superimposed. The juxtaposition of these two motions produces a
motion that is illustrated in Fig. 1.11, which is described by
x1 = X1 , x2 =  X1 + n X2 , x3 = n1 X3 (1.36)
whose inverse motion is given by
X1 = 1 x1 , X2 =  n x1 + n x2 , X3 = 1n x3 . (1.37)
Kinematics 21

w w
d d
l

w0

d0
l l

l0

t0 t1 t

e1
e3

e2

Fig. 1.11 Juxtaposition of an extension followed by a simple shear. Here `0 ,


w0 , and d0 are the dimensions of length, width, and depth of a gage section that
is rst extended to a rectangular prism with dimensions `, w, and d and later
sheared by some extent 

In the motion displayed in Fig. 1.11, stretch varies over the time interval
[t0 , t1 ] and is held constant thereafter, while shear  is held fixed at 0 over
the time interval [t0 , t1 ] and then varies thereafter, i.e., for all t  t1 . The
parameter n accounts for the aspect ratio of the sample being tested, in that
(
1/2 whenever height/width  1,
n= (1.38)
0 whenever height/width 1
where the specimens height aligns with the 1-direction and its width aligns
with the 2-direction, as drawn in Fig. 1.11. Reality lies somewhere in
the interval 0
n
1/2 where n = 0 depicts pure shear in the sense of
Eqs. (1.32) and (1.33), while n = 1/2 depicts uniaxial extension in the sense
of Eqs. (1.10) and (1.11).

1.4.3.1 Problem
Derive the components of the velocity V and acceleration A vectors that
describe the second stage of this experiment, viz., the shearing. How
do they compare with those of the simple-shear experiment given in
Eqs. (1.20) and (1.21)?
22 Soft Solids

1.4.4 Other Problems


1. Show that the equations describing the motions of (a) uniaxial exten-
sion, (b) equi-biaxial extension, (c) simple shear, (d) pure shear, (e)
biaxial extension, and (f) extension followed by simple shear are each
a special case of Eq. (1.22), which describes the motion of a homoge-
neous planar membrane.
2. Show that the components of the velocity and acceleration vectors
for (a) uniaxial extension, (b) equi-biaxial extension, (c) simple shear,
(d) pure shear, (e) biaxial extension, and (f) extension followed by
simple shear are each a special case of Eqs. (1.27) and (1.28), which
describe their counterparts for an isotropic, homogeneous, planar
membrane.
3. Derive the components for the velocity and acceleration vectors
described by the motion in Eq. (1.29) that belongs to the linear 2D quad
element of finite elements. How does the inhomogeneity of this defor-
mation enter into its velocity and acceleration vectors?
Chapter 2

Deformation

The velocity and acceleration vectors derived in the preceding chapter are
important kinematic fields, but, in and of themselves, they are not capable
of describing how a body B deforms; they only describe how any particle P
within B moves through ambient space R3 . In order to study deformation,
one needs to quantify the change in shape of a body B as it is transformed
from some initial configuration 0 into its final configuration  over some
interval [t0 , t] in time, which is the topic of this chapter.
Motion is described by a position vector, but not deformation. How-
ever, the dierence between two position vectors associated with a pair of
neighboring particles is such a measure. Through the relative motions of
such dierences, deformations can be quantified. Consider two particles P
and P 0 that are neighbors to one another in body B. Let the incremental
displacement vector connecting particle P to particle P 0 in configuration
0 be denoted by dX = X 0 X and let the displacement vector that con-
nects these same two particles in the current configuration  be denoted by
dx = x 0 x, as shown in Fig. 2.1. Vectors dX and dx point from material
particle P to material particle P 0 in configurations 0 and , respectively.
Applying the chain rule to the law of continuous media given on p. 6
produces a linear transformation or mapping between dX = dXI eI in 0
and dx = dx i ei in  that is expressed as

@i @i
dx i = dXI or dx i = FIi dXI with FIi = (2.1)
@XI @XI

which, in accordance with Appendix B, maps the Lagrangian vector dX


into its Eulerian vector dx; specifically, it obeys the field-transfer operator
A.D. Freed, Soft Solids: A Primer to the Theoretical Mechanics of Materials, 23
Modeling and Simulation in Science, Engineering and Technology,
DOI 10.1007/978-3-319-03551-2_2, Springer International Publishing Switzerland 2014
24 Soft Solids

dx
dX

X x @t
@ t0
e2 x
X
0
e1
e3

Fig. 2.1 Body B is deformed from its original shape at time t0 , i.e., from con-
guration 0 , into a nal deformed shape at time t, viz., into conguration .
Neighboring particles P and P 0 were at positions X and X 0 in 0 and are now
located at x and x 0 in . The incremental vector dX = X 0 X maps into vector
dx = x 0 x according to Eq. (2.1), whose return mapping maps dx back into
dX via Eq. (2.3). All elds are quantied against a common Cartesian basis
(e1 , e2 , e3 ) in R3

of a contravariant vector field described in Eq. (B.9) so that in matrix


notation
@(X , t)
{dx(x, t)} = [F(X , t)] {dX (X )} with F(X , t) = (2.2)
@X
where tensor F(X , t) is the deformation gradient. Tensor F = FIi ei e I
is a two-state field with row index i in FIi associating with configuration
and column index I in FIi associating with configuration 0 . Tensor F
is the fundamental field describing deformation. F is normalized so that
F(X , t0 ) = I = Ii ei e I where I is the identity tensor with components
Ii that are the Kronecker delta, which equal 1 whenever i = I and are 0
otherwise. Here ei and eI are the contravariant base vectors in  and 0 ,
while e i and e I are the covariant base vectors in  and 0 . In Cartesian
tensor analysis, ei  eI  e i  e I . This is not the case in general tensor
analysis [cf. Sokolniko (1964)]. The matrix representation [F] of tensor
F contains nine nonsymmetric components.
In like manner, a deformation gradient exists for the inverse motion.
Applying the chain rule to the law of continuous media also allows one to
write
@I i @I
dXI = dx or dX I
= f I
i dx i
with f I
i = (2.3)
@x i @x i
which, in accordance with Appendix B, maps the Eulerian vector dx into
its Lagrangian vector dX ; specifically, it obeys the field-transfer operator
Deformation 25

of a contravariant vector field described in Eq. (B.10) so that in matrix


notation
@ (x, t)
{dX (X )} = [f (x, t)] {dx(x, t)} with f (x, t) = . (2.4)
@x
Like F = FIi ei e I , f = f Ii eI e i is a two-state tensor field.
Tensor F maps the tangent vector dX from 0 into its associated tan-
gent vector dx in . Tensor f maps the tangent vector dx from  into its
associated tangent vector dX in 0 , which is the reverse of mapping F.

2.1 Homogeneous Deformation

From the definition of a homogeneous motion given in Eq. (1.2), it follows


that its spatial gradient is
@x @(X , t) @X
F= = = (t) = (t) I = (t) (2.5)
@X @X @X
and, therefore, any distortion that contributes to a homogeneous motion,
i.e., , equates with a deformation gradient F that depends only upon
time t.
A deformation is said to be inhomogeneous whenever its deformation
gradient F depends upon position X as well as upon time t, e.g., the 2D
quad element whose deformation gradient is contained within Eq. (1.31).

2.2 Conservation of Mass

From the conservation of massa law of physicscomes the governing


field equation (Holzapfel 2000, pp. 131135)
D%
+ div(v) % = 0 (2.6)
Dt
where % is the mass density and div(v) = @v i /@x i is the divergence of the
velocity vector, which is scalar valued. The solution to this dierential
equation can be written in the form of
dv %0
det F = = >0 (2.7)
dV %
where dv/dV ratios the volume of an infinitesimal material element (whose
centroid is located at particle P ) between its current value dv in  to its
26 Soft Solids

reference value dV in 0 , while the reciprocal ratio %0 /% is a quotient of


that volume elements respective mass densities.
A direct consequence of the conservation of mass is that the deforma-
tion gradient cannot be singular, i.e., det F 0, consequently, its inverse
must exist
 
@(X , t) 1 @X @ (x, t)
F 1 = = = = f (x, t) (2.8)
@X @(X , t) @x

whose components are

@I
[F1 ]Ii = = f Ii . (2.9)
@x i

So one observes that F has components FIi , while its inverse F 1 has com-
ponents [F1 ]Ii . The notation [F1 ]Ii does not mean 1/FIi . The superscript
(or row) index of F 1 associates with the material frame, while the sub-
script (or column) index belongs to the spatial frame, which is the opposite
index pairing present in F.
Regarding deformation gradients, the superscript index associates with
the argument in the numerator (the dependent variable of the motion map),
while the subscript index associates with the argument in the denominator
(the independent variable of the motion map). Whenever the independent
variable is X (i.e., Lagrangian), then FIi are the components of F; likewise,
whenever the independent variable is x (viz., Eulerian), then [F1 ]Ii are the
components of F 1 .
In what follows, F 1 will be used in place of f , as they are two expres-
sions of the same mapping. When describing fluids, f is preferred over F
because one places oneself in . When characterizing solids, which is the
focus of this text, F is preferred over f because one places oneself in 0 .

2.2.1 Isochoric Deformation


A deformation is said to be isochoric if it preserves its volume, which is
also referred to in the thermodynamics literature as an isometric or iso-
volumetric process. Consequently, imposing a constraint of det F = 1,
which implies that dv = dV from Eq. (2.7), ensures that the corresponding
deformation is isochoric.
Deformation 27

2.3 Deformation Gradient as a Mapping Function

Tensor F is a two-state field, i.e., it has one index in the material configura-
tion 0 , while the other index resides in the spatial configuration . This
is because the deformation gradient F is a transformation mapping. Matrix
[F] maps a tangent vector {dX } from the reference configuration 0 into
{dx}, which is this tangent vectors representation in the current config-
uration , according to Eq. (2.2). Physically, these are the same tangent
vectors, but, mathematically, they are distinct.
Similarly, matrix [F 1 ]T maps the normal vector {N } = {dS(X )/dX }
to some material surface S(X ) from a reference configuration 0 into a
normal vector {n} = {dS( (x, t))/dx} that resides within the current con-
figuration . Applying the chain rule to S( (x, t)), incorporating the law
of continuous media, allows one to write
dS dS @I
= or ni = [F1 ]Ii NI . (2.10)
dx i dXI @x i
which, in accordance with Appendix B, maps the Lagrangian vector dS/dX
into its Eulerian vector dS/dx; specifically, it obeys the field-transfer oper-
ator of a covariant vector field described in Eq. (B.13) in that
(  )    
dS  (x, t) dS(X ) d (x, t) dS(X ) h 1 i
= = F (2.11)
dx dX dx dX
or, equivalently,
{n} = {N } [F 1 ] so {n(x, t)} = [F 1 (X , t)]T {N (X )}. (2.12)
The dual mapping matrices of [F] and [F 1 ]T are the kinematic corner-
stones of continuum mechanics; cf. Appendix B. The notation for the de-
formation gradient F is made special, viz., it is typeset in an upright font
instead of a slanted font like other tensor fields, precisely because the field-
transfer operators [F] and [F 1 ]T map vector and tensor fields between the
two configurations of a deformation marked by the end points of a motion
over some interval in time, say [t0 , t].
Vector fields that are pushed forward from 0 into  via [F], in accor-
dance with Eq. (2.2), can be pulled back from  into 0 with the reverse
mapping [F 1 ]. Such vectors are called contravariant vector fields. Like-
wise, vector fields that are pushed forward from 0 into  via [F 1 ]T , in
accordance with Eq. (2.12), can be pulled back from  into 0 with the re-
verse mapping [F]T . Such vectors are called covariant vector fields (Lodge
28 Soft Solids

1974; Marsden and Hughes 1983; Sokolniko 1964). These mappings are
discussed in more detail in Appendix B, wherein Figs. B.2 and B.3 are use-
ful illustrations showing how these mappings apply to the transfer of field
from one configuration into another.
Coming to an understanding of the concepts that are outlined in Ap-
pendix B, viz., configuration physics, is essential before you can come to a
physical understanding of the mechanics that materials incur during finite
deformations. Please study Appendix B before advancing.

2.4 Stretch and Rotation

At any particle P in body B, the deformation gradient F that connects


a reference configuration 0 with its current configuration  will have a
unique polar decomposition, as illustrated in Fig. 2.2, that is given by (Noll
1958)
F = RU = vR (2.13)
where tensors U (X , t) and v(x, t) are called the right and left stretch ten-
sors, so named because they reside at the right and the left of the rotation
tensor R. Mathematically, tensors U and v are positive definite with matrix
representations that are always symmetric, while tensor R is orthogonal, all
of which are consequences of the polar decomposition theorem from linear
algebra. Consequently, [U ] = [U ]T , [v] = [v]T , and RT R  RRT = I
implying that [R]T = [R1 ] that, because det F > 0, requires det R = 1
since stretches U and v are positive definite, i.e., rotation R is a proper
orthogonal tensor. In component form, this decomposition is written as
j
FIi = RJi UIJ = v ji RI (2.14)

where R = RIi ei e I , U = UIJ eI e J , and v = v ji ei e j where in each


the superscript is the row indexer and the subscript is the column indexer.
Furthermore, R1 = [R1 ]Ii eI e i and, in accordance with Eq. (A.26),
j
RT = ij RJ JI eI e i (2.15)
j
so that [R1 ]Ii = ij RJ JI . The transpose of a mixed tensor field, like
RT , is a bit tricky to work with in component notation, requiring extra
care in its handling [cf. Appendix A and Marsden and Hughes (1983,
pp. 4849, 137)].
Deformation 29

R
U

F @ t
@ t0

R v

Fig. 2.2 Body B is deformed from its original shape at time t0 , i.e., from con-
guration 0 , into a nal deformed shape at time t, viz., into conguration .
Considering the eld transfer of a contravariant vector illustrated in Fig. B.2, the
deformation gradient F can be decomposed in one of two ways: F = RU applies
a Lagrangian stretch of U from 0 that is followed by a rotation of R into ,
whereas F = vR rotates out of 0 via R after which an Eulerian stretch of v
places B into . The intermediate congurations (drawn as dashed ellipses) are
not physically realizable unless the deformation is uniform

Contracting Eq. (2.13) with R1 from the right gives


v = RU R1 or v ji = RIi UIJ [R1 ]Jj , (2.16)
while contracting Eq. (2.13) with R1 from the left produces
j
U = R1 vR or UIJ = [R1 ]Ii v ji RJ . (2.17)
An absence of rotation implies that R1  R = I. An absence of stretch
implies that U  v = I. Consequently, an absence of both rotation and
stretch implies that F 1  F = I.
What Eq. (2.13) implies is that any deformation F from 0 to  can
be decomposed into two dierent mapping sequences whose intermediate
configurations are, in general, nonphysical, viz., they are not realizable. In
the case of F = vR, a rotation takes place first, followed by a stretch of
v. In the other case of F = RU , a stretch of U is imposed first, with a
rotation following. Both sequence maps begin in the reference configura-
tion 0 , and both eventually end up in the spatial configuration ; they
just traverse dierent paths to get there, as illustrated in Fig. 2.2. This is a
consequence of the fact that matrix multiplication does not commute. The
30 Soft Solids

polar decomposition of the deformation gradient plays an important role


in many theories describing the mechanics of materials (Holzapfel 2000;
Ogden 1984; Simo and Hughes 1998), but they are not directly employed
in the constitutive theories put forward in this text.
Stretch is a unique descriptor of deformation. After one selects a frame
of reference, viz., material or spatial, unique components can be assigned
to stretch. Consequently, great value exists in measuring and reporting
stretches, not strains, when documenting experiments done on soft solids.
Given components for one strain measure, it is not always possible to ar-
rive at components for another strain measure without also knowing the
components of stretch.
j
Contraction RT R = I has covariant components RIi ij RJ = IJ . Con-
j
traction RRT = I has contravariant components RIi IJ RJ = ij . Kro-
necker deltas ij and IJ are the covariant and contravariant metrics of
Cartesian space in the Eulerian and Lagrangian frames, respectively. An
j
application of Eq. (A.42) to RT R specifies components IK RKi ij RJ with
the first Kronecker delta being able to contract out, thereby proving the
j
stated result RIi ij RJ . Likewise, an application of Eq. (A.44) to RRT
specifies components RIi IJ RJk kj with the second Kronecker delta being
j
able to contract out, thereby proving the stated result RIi IJ RJ .

2.5 Rate Fields

Taking the material derivative of Eq. (2.2), viz., taking the time derivative
of the transfer of field {dx} = [F] {dX } at a fixed particle P and, therefore,
at fixed material coordinates X so that dX P = 0, one determines from the
product rule that
P = [FP ] {dX } = [FP ] [F 1 ] [dx] = [l] {dx}
{dx} (2.18)
where {dX } = [F 1 ] {dx} follows from Eqs. (2.4) and (2.8). This expres-
sion defines the Eulerian velocity gradient tensor1

P 1 = @(X , t) @ (x, t) = @v = grad(v)


P
l(x, t) = FF (2.19)
@X @x @x
1 The Eulerian velocity gradient l = FF
P 1 is typically denoted as L in the litera-
ture. That notation is reserved here for the Lagrangian velocity gradient dened
in Eq. (2.24), which is a eld not found in the literature to the best of the authors
knowledge.
Deformation 31

where
P 1
l = FF has spatial components ` ij = FP Ii [F1 ]Ij , (2.20)
which plays the analogous role in rate-based theories that F plays in
deformation-based theories. Like the deformation gradient F, the veloc-
ity gradient l is not symmetric. Unlike F, the inverse of l does not exist.
It is useful to decompose the velocity gradient, not as a product as in
the polar decomposition of F, but rather as a sum, viz.,
l = d +w (2.21)
wherein
   
d = 12 l + l T has components d ij = 12 ` ij + jk ` k` `i , (2.22)
   
w = 12 l l T has components w ji = 12 ` ij jk ` k` `i (2.23)
where d (x, t) is the symmetric stretching tensor (also referred to in the lit-
erature as the rate-of-deformation tensor or the strain-rate tensor) whose
components come from Eq. (A.27), while w(x, t) is the skew-symmetric
vorticity tensor (also referred to in the literature as the rate-of-rotation ten-
sor or the spin tensor) whose components come from Eq. (A.29).

2.5.1 Lagrangian Velocity Gradient


It is useful to define what is, essentially, the Lagrangian velocity gradient
tensor L(X , t) via the formula
L = F 1 FP with material components LIJ = [F1 ]Ii FP Ji , (2.24)
which, by an inspection of its uppercase indices, is seen to be a material
field defined over the reference configuration 0 . Recall that l = FF
P 1 . It
is a trivial matter to show that L pushes forward into l, and l pulls back
into L between configurations 0 and  according to the paired mappings
[l] = [F] [L] [F 1 ] or ` ij = FIi LIJ FJj = FP Ji [F1 ]Jj ,
j
(2.25)
[L] = [F 1 ] [l] [F] or LI = FI ` i F = [F1 ]I FP i ,
J i j J i J
which are the field-transfer operators for a mixed tensor field defined in
Eqs. (B.23)(B.26).
As in the Eulerian frame, the velocity gradient of the Lagrangian frame
can be additively decomposed into symmetric and skew-symmetric parts
L=D +W (2.26)
32 Soft Solids

wherein
   
D = 12 L + LT with DIJ = 12 LIJ + JK LLK LI , (2.27)
   
W = 12 L LT with WJI = 12 LIJ JK LLK LI (2.28)
where D and W are the respective Lagrangian stretching and vorticity
tensors. It follows that [l T ] = [F] [LT ] [F 1 ] and [LT ] = [F 1 ] [l T ] [F]
and therefore
[d] = [F] [D] [F 1 ] and [D] = [F 1 ] [d ] [F], (2.29)
1 1
[w] = [F] [W ] [F ] and [W ] = [F ] [w] [F]. (2.30)

2.5.2 Isochoric Deformations


Recall that an isochoric deformation requires that det F = 1. Applying
Eq. (A.83) to F leads to
 
detP F = det(F) tr FP F 1 = det(F) tr(l) (2.31)
or, equivalently, from Eq. (A.84), one arrives at
P F) = tr l.
ln(det (2.32)
If det F = 1 then detP F = 0, thereby requiring that det(F) tr(l) = 0, but
det F = 1 so tr l must equal 0. Furthermore, from their definitions in
Eqs. (2.19) and (2.24) and a property of the trace given in Eq. (A.70), it
P 1 ) = tr(F 1 FP ) = tr L.
also follows that tr l = tr(FF
Consequently, det F = 1 (or det F 1 = 1) and tr L = 0 (or tr l = 0) are
equivalent constraints for imposing isochoric deformations. The former
pair applies to formulations where the deformation gradients F or F 1 play
the role of the independent variable for deformation, while the latter pair
applies to formulations where the velocity gradients L or l play the role
of the independent variable for deformation. The first occurrence in each
pairing applies for Lagrangian analysis, while the second occurrence in
each pairing applies for Eulerian analysis.

2.6 Numerical Implementation

The polar decomposition theorem tells us that the rotation R and stretch
tensors U and v exist, but it does not tell us how to compute them. That is
the topic of this section. Here we consider interfacing with, e.g., an updated
Deformation 33

Lagrangian finite element code where the velocity gradient l is the inde-
pendent kinematic variable [cf. Belytschko et al. (2000)]. The algorithms
that follow employ a two-step AdamsBashforth predictor followed by a
trapezoidal corrector for numerical integration,2 which is in keeping with
the target application of an algorithm that resides within a finite element
code.

2.6.1 Angular Velocity


The skew-symmetric angular velocity = ij ei e j is defined by
P 1
= RR ij = RP Ii [R1 ]Ij ,
with components (2.33)
which has the same mathematical structure as l = FP F 1 . From the defi-
nitions for a polar decomposition of the deformation gradient and for the
velocity gradient given in Eqs. (2.13) and (2.19) one arrives at a dierential
equation governing stretch, in particular
vP = lv v (2.34)
whose skew-symmetric contribution describes the expression
v + v = vw + wv + z with z = dv vd (2.35)
where w = w ji ei e j,
z= z ij ei e j,
and = ij ei ej
are each skew
symmetric. Because these three tensors are skew, they can be expressed as
vectors w = w i ei , z = z i ei , and ! = i ei defined so as to obey maps
j j j
w i =  ikj wk , z i =  ikj zk , and i =  ikj k , (2.36)
ij
wherein  k is the cyclic permutation operator with values: 1 if indices
ijk are unique and cyclically ordered, i.e., 123, 231, or 312; 1 if they are
unique with reverse cyclic ordering, viz., 132, 213, or 321; or 0 whenever
ij
two or three indices have the same indical value. Note:  ikj =  k . From
here, Dienes (1979) derived the formula
 1  j j 
! = w + tr(v)I v z wherein z i =  ikj d` vk` v` d`k , (2.37)
where ! is the angular velocity vector.
These formul comprise Algorithm 2.1, which returns the angular ve-
locity vector ! given the velocity gradient l and stretch v tensors in their
2 One should not use forward Euler as ones numerical method for integrating
ODEs, even though it was likely the method taught to you in your class on
dierential equations. This method has serious stability issues and has been the
cause of numerous engineering failures, even disasters.
34 Soft Solids

Algorithm 2.1 Vector of angular velocity !


function AngularVelocity ( l, v)
d ij 1 (` i + ` k `i )
2 j jk `
w ji 1 (` i ` k `i )
2 j jk `
j
wi  j wk
ik
 j j 
zi  ikj d` vk` v` d`k
 1
! w + tr(v)I v z
return !
end function

Eulerian form. The velocity gradient l is the known, independent, kine-


matic variable in such formulations, while a predicted estimate for the
stretch tensor v comes from Algorithm 2.3.

2.6.2 Rotation
The issue with applying a conventional numerical technique to integrate
RP = R [from Eq. (2.33)] for updating the rotation tensor R, without
addressing the intrinsic character of R, is an unavoidable degradation in the
orthogonality of R with increasing numbers of integration steps n taken, no
matter how fine a step size t is used (Atluri and Cazzani 1995; Flanagan
and Taylor 1987). The reason for this degradation is that the sum of two
orthogonal tensors is not orthogonal. Algorithm 2.2 is free from this defect.
It multiplies rotations, thereby maintaining orthogonality. The product of
two orthogonal tensors is an orthogonal tensor, within numerical roundo
error.
Because the rotation tensor R arising from a polar decomposition of the
deformation F is orthogonal, it can be described in an alternative form of
R = eQ so Rr = r with [Q] ij =  ikj rk , krk = 1, (2.38)
where Q is a skew-symmetric tensor representation for the vector that is
the axis r about which rotation R occurs, with r being the only real-valued
eigenvector of R, and with  corresponding to the angle of this rotation.
From the eigenvector property of Rr = r, one determines that
R = I + sin( ) Q + (1 cos  ) Q2 , (2.39)
while from the evolution of R described by R = R one arrives at
P
! = P r + sin( ) rP + (1 cos  ) r  r,
P (2.40)
Deformation 35

which are Eqs. (2.16) and (8.25) of Atluri and Cazzani (1995), neither be-
ing simple to derive. Equation (2.39) is the matrix representation for a set
of formul originally derived by Euler in 1775 (Cheng and Gupta 1989).
Following a suggestion by Dienes (2003), Eq. (2.40) can be rewritten
as the matrix equation

A rP = ! P r so that rP = A 1 (! P r) (2.41)

with tensor A having components


2 3
sin  (1 cos  ) r3 (1 cos  ) r2
[A] = 4 (1 cos  ) r3 sin  (1 cos  ) r1 5 . (2.42)
(1 cos  ) r2 (1 cos  ) r1 sin 

By definition r  r = 1 and, therefore, 2 r  rP = 0 that, from Eq. (2.41),


requires r A 1 (! P r) = 0. But it can be shown that r A 1 = csc( ) r
0, which reflects the singularity of A present at  = 0, while r  A =
sin( ) r. Consequently, it is sucient to require (! P r) = 0 to ensure
r  A 1 (! P r) = 0, thereby producing the anticipated result
!
! = P r ) P = r  ! = k!k with r = (2.43)
k!k

which enables the rotation tensor R to be quantified via Eq. (2.39). To


the best of the authors knowledge, ! = P r has been suggested in the
literature, but not rigorously proven.
The axis of rotation r cannot be oriented in an absence of rotation, viz.,
whenever P = k!k = 0. This condition associates with the point singular-
ity of tensor A 1 , with A being defined in Eq. (2.42). This special case is
readily handled, because R = I in the absence of rotation. Parameter tol
in Algorithm 2.2 handles this condition, where tol is set to a small positive
number, e.g., machine precision m .
The formul of this section present a strategy whereby the rotation ten-
sor R of Eq. (2.39), which is the rotation tensor within F = RU = vR,
can be acquired, given the angular velocity vector ! is known a priori,
which follows from Algorithm 2.1. This strategy is implemented in Algo-
rithm 2.2, where a two-step AdamsBashforth predictor and a trapezoidal
corrector are used to numerically integrate the angle of rotation  and to
assign the axis of rotation r. A forward Euler step is used to start the pre-
dictor.
36 Soft Solids

Algorithm 2.2 Updating the rotation tensor R


var !n1 , Rn1 F stored variables

procedure InitRotation (t, !0 , var R0 , var R1 ) Fn=0


R0 I
P k!0 k
P > tol then
if || F forward Euler predictor
r !0 /P
[Q] ji  ikj r k
 
R1 I + sin(P t) Q + 1 cos(P t) Q2
else
R1 I
end if
!n1 !0 F update stored variables
Rn1 R0
end procedure

procedure Rotation (t, !n , var Rt , var R+t , var Rn , var Rn+1 )
P k 12 (!n + !n1 )k
if |P | > tol then
F run corrector over previous increment
r 1 (! + !
n n1 )/P
2
[Q ] ij  ikj {r }k
 
Rt I + sin(P t) Q + 1 cos(P t) Q  Q
else
Rt I
end if
Rn Rt Rn1
P + k 12 (3 !n !n1 )k
if |P + | > tol then F run predictor over next increment
r+ 1
(3 ! n ! n1 )/P+
2
[Q+ ] ij  ikj {r+ }k
 
R+t I + sin(P + t) Q+ + 1 cos(P + t) Q+  Q+
else
R+t I
end if
Rn+1 R+t Rn
!n1 !n F update stored variables
Rn1 Rn
end procedure
Deformation 37

Algorithm 2.2 requires, as input, the step size of integration  t and the
axis of angular velocity at the beginning of the step !n . The algorithm
returns a corrected estimate for the rotation at the beginning of the step
Rn and a predicted estimate for the rotation at the end of the step Rn+1 .
It also provides estimates for the incremental rotations Rt and R+t oc-
curring over the prior [tn t, tn ] and next [tn , tn + t] integration steps,
respectively. Rotation Rt is evaluated according to the integration rule
of the corrector, while R+t is evaluated according to the integration rule
of the predictor. These two incremental rotations find application when
numerically integrating Eq. (2.34) for stretch.

2.6.3 Stretch
The left stretch tensor v is governed by dierential equation (2.34), whose
solution can be acquired via, e.g., Algorithm 2.3. Stretch v is required in-
put for computing the axis vector for angular velocity ! using Algorithm
2.1. All numerical integrations taking place in Algorithm 2.3 are done
in the updated Lagrangian configuration n aliated with time tn in ac-
cordance with Appendix B. Stretches belonging to the prior configuration
n1 are pushed forward into n via the map Rt , with the final result
then being pushed forward from the updated Lagrangian frame n into the
next Eulerian frame n+1 using the map R+t . It follows from Eq. (2.13)
that v = RU R1 , recalling that R1 = RT , which makes our algorithm for
integrating stretch rate fundamentally dierent from, say, an algorithm for
integrating stress rate.
Algorithm 2.3 requires, as input, the step size of integration  t , the an-
gular velocity vector at the beginning of the step !n , the velocity gradient
at the beginning of the step ln , the incremental rotations over the previous
Rt and next R+t integration steps, and the left stretch tensor at the begin-
ning of the step vn . The algorithm returns a corrected estimate for vn and
a predicted estimate for vn+1 for use in the next calling of Algorithm 2.1.
A dierent approach for quantifying the fundamental deformation
fields of stretch and rotation, which is based upon the spectral decompo-
sition theorem with the deformation gradient being the known kinematic
variable, can be found in Simo and Hughes (1998, pp. 241244).
38 Soft Solids

Algorithm 2.3 Updating the left stretch tensor v


var vn1 , vP n1 F stored variables

procedure InitLeftStretch (t, l0 , var v0 , var v1 ) Fn=0


v0 I 
[Pv0 ]ij 1 [l ]i + [l ]k `i 
2 0 j jk 0 `
v1 I + t vP 0 F forward Euler predictor
vn1 v0 F update stored variables
vP n1 vP 0
end procedure

procedure LeftStretch (t, !n , ln , Rt , R+t , var vn , var vn+1 )


[n ]ij  ikj {n }k
vP n ln vn vn n   F evaluate
vn Rt vn1 (Rt )T + t
2 vP n + R vP (R
t n1 t
)T F correct
vP n ln vn vn n  F reevaluate
Un,n+1 t 3 vP R vP (R )T F predict
2 n t n1 t
vn+1 R+t (vn + Un,n+1 )(R+t )T F push from n into n+1
vn1 vn F update stored variables
vP n1 vP n
end procedure

2.7 Examples

A fair number of fields can be used to describe deformation, each having its
own purpose. The examples below will determine the deformation gradient
F and its inverse F 1 ; its associated polar fields U , v, and R; and from F,
P
the aliated rate fields of L, l, d , D, w, and W as they apply to the BVPs
studied in this text.

2.7.1 Uniaxial Extension


For the case of an isochoric uniaxial extension of an isotropic incompress-
ible material, whose motion is given in Eqs. (1.10) and (1.11), the defor-
mation gradient
2 and its inverse
3 are determined2to 1 have components
3
0 0
0 0
[F] = 4 0 /2 0 5 and F 1 = 4 0 1/2 0 5
1
(2.44)
0 0 1/2 0 0 /2
1
Deformation 39

whose polar decomposition produces stretches and a rotation of3


2 3 2 3
0 0 100
[U ]  [v] = 4 0 /2 0 5 while [R] = 40 1 05
1
(2.45)
0 0 1/2 001
and whose rate fields, quantified via
2P 3
0 0
P = 4 0 /2
[F] P 3/2
0 5, (2.46)
0 0 P
/2 /2
3

become
2P 3
/ 0 0
[L]  [l]  [D]  [d ] = 4 0 /2
P 0 5. (2.47)
0 P
0 /2
No vorticity occurs because w = l d = 0; likewise, W = L D = 0. As
a check, det F = 1 and tr L = 0, so the prescribed deformation is isochoric.

2.7.2 Equi-biaxial Extension


For the case of an isochoric equi-biaxial extension of an isotropic incom-
pressible material, whose motion is given in Eqs. (1.14) and (1.15), the
deformation gradient and its inverse are determined to have components
2 3 2 1 3
0 0
1 0 0
[F] = 4 0 0 5 and F = 4 0 1 0 5 (2.48)
00 2
0 0 2

whose polar decomposition produces stretches and a rotation of


2 3 2 3
0 0 100
[U ]  [v] = 4 0 0 5 while [R] = 40 1 05 (2.49)
0 0 2 001
and whose rate fields, quantified via
2P 3
0 0
[FP ] = 4 0 P 0 5, (2.50)
P
0 0 2 / 3
3 Shear-free motions experience no rotation, by denition, therefore, R = I for
all deformations belonging to this kinematic class of motions.
40 Soft Solids

become
2P 3
/ 0 0
4 P
[L]  [l]  [D]  [d ] = 0 / 0 5, (2.51)
0 P
0 2 /
implying that no vorticity occurs, i.e., w  W = 0. As a check, det F = 1
and tr L = 0, so this deformation is isochoric, too.

2.7.3 Simple Shear


From the planar motion describing a simple shear given in Eqs. (1.18) and
(1.19), the deformation gradient and its inverse are determined to have
components
2 3 2 3
1 0
1 1  0
[F] = 40 1 05 and F = 40 1 05 . (2.52)
001 0 0 1
Derivation of the polar fields that associate with this deformation takes a
bit of doing. Begin by considering a clockwise angle of rotation  in the
12 plane described by
2 3
cos  sin  0
[R] = 4 sin  cos  05 , (2.53)
0 0 1

which is easily shown to be proper orthogonal, i.e., RT R = I with


det R = 1, and is, therefore, an admissible rotation. From the definition
for stretch U = R1 F obtained from the polar decomposition of the defor-
mation gradient, i.e., Eq. (2.13), noting that [R1 ] = [R]T , one gets
" #" # " #
cos  sin  0 1 0 cos  sin  +  cos  0
[U ] = sin  cos  0 0 1 0 = sin  cos  +  sin  0 . (2.54)
0 0 1 001 0 0 1

However, because [U ] = [U ]T , it follows that U12 = U21 and, as such,


sin  = sin  +  cos  , which can be solved for  yielding  = tan1 ( /2 ),
/2

/2 . Recalling that tan  = rise/run , it follows that the rise = 
  1/
and the run = 2, so the hypotenuse is 4 +  2 2 and, therefore, cos  =
  1/   1/
2/ 4 +  2 2 and sin  =  / 4 +  2 2 . With the sin  and cos  now known
Deformation 41

t0 t
e2

e1

Fig. 2.3 Simple shear takes a square of dimension h (for height) and deforms it
into a quadrilateral of the same area by a magnitude  and angle  of shearing

in terms of the deformation variable  , whose geometric interpretation is


drawn in Fig. 2.3, the rotation tensor is found to have components
2 3
2  0
1 4 2 5
[R] = 0 (2.55)
(4 +  2 ) 1/2 0 0 (4 +  2 ) 1/2

from which the two stretch tensors are determined to have components
2 3
2  0
1 4 2 +  2 5
[U ] = 0 (2.56)
(4 +  )
2 1/2
0 0 (4 +  ) 2 1/2

and
2 3
2 + 2  0
1 4  2 5
[v] = 0 (2.57)
(4 +  2 ) 1/2
0 0 (4 +  )
2 1/2

with their dierences residing in the locations of the second-order term  2


found in the normal components.
Arriving at the velocity gradients is much more straightforward, viz.,
2 32 3 2 3

1  0 0 
P 0 0 
P 0
[L] = F 1 FP = 40 1 05 40 0 05 = 40 0 05 (2.58)
0 0 1 000 000
and
2 32 3 2 3

0 P 0 1  0 0 P 0
1
[l] = FF
P = 40 0 05 40 1 05 = 40 0 05 (2.59)
00 0 0 0 1 00 0
42 Soft Solids

where [L]  [l] is uncommon for deformations with o-diagonal terms,


although that is the case here for simple shear. The symmetric and
skew-symmetric parts of the velocity gradients are therefore given by
2 3 2 3
0 P 0 0 P 0
1 1
[D]  [d] = 4P 0 05 and [W ]  [w] = 4P 0 05 . (2.60)
2 000 2 0 00

Consequently, from a kinematic perspective, simple shear finds distinction


between its various deformation fields, important distinctions that are not
present in uniaxial and biaxial deformations and which you can use to gain
understanding when seeking insight into a materials behavior.

2.7.4 Homogeneous Planar Membranes


For the case of an isotropic planar membrane undergoing the isochoric
homogeneous motion specified in Eqs. (1.22) and (1.26), the deformation
gradient and its inverse are determined to have components
2 3 2 3
1  1 2 0
1 1 2 1 2 0
[F] = 42 1 2 0 5 & F = 42 1 1 0 5 (2.61)
0 0
1
0 0
2
whose time rate of change is
2 3
P 1 1P 2 0
6 P 7
[F ] = 42 1
P
P 2 0 5, (2.62)
0 0
/

P 2
which follow straightaway from Eqs. (1.24) and (1.25), while recalling that
the areal stretch of Eq. (1.23) is given by
= 1 2 (11 2 ) and, therefore,
 

P = P 1    P .
1 2 1 2 1 2 1 2 (2.63)
The motion maps defined in Eqs. (1.22) and (1.26) are isochoric because
det F  det F 1 = 1 due to how F33 is defined.
Following the same line of reasoning that was used to derive the rota-
tion and stretch tensors for simple shear, Freed et al. (2010) arrived at like
values applicable for planar membranes, which, in the notation of this text,
lead to polar descriptions where
2 3
1 4 1 + 2 1 2 2 1 0
[R] = (1 2 2 1 ) 1 + 2 0 5 (2.64)
 0 0 
Deformation 43

wherein
q
 2  2
= 1 + 2 + 1 2 2 1 (2.65)
normalizes the rigid-body rotation, i.e., it ensures that det R = 1. This
expression for the rotation R allows the right stretch tensor U = RT F to
be written as
2
( + 2 ) 2 1 (1 2 2 1 )
1 4 1 1
[U ] = 1 2 (1 + 2 )
 0
3
1 2 (1 + 2 ) 0
2 ( 1 + 2 ) + 1 2 (1 2 2 1 ) 0 5 , (2.66)
0
1 
while the left stretch tensor v = FRT has components
2
( + 2 ) + 1 2 (1 2 2 1 )
1 4 1 1
[v] = 1 22 + 2 21

0
3
1 22 + 2 21 0
2 ( 1 + 2 ) 2 1 (1 2 2 1 ) 0 5 (2.67)
0
1 
where [U ] and [v] are similar, yet distinct.
The components of the velocity gradients are easily acquired, too, being


[L] = F 1 FP
2 3
2 P 1 1 2 2P 1 2 1P 2 1 2 P 2 0
1 6 7
= 4 1 2P 1 2 1 P 1 1 P 2 2 1 1P 2 0 5 (2.68)

0 0
P
and


[l] = FP F 1
2 3
2 P 1 2 1 1P 2 1 1P 2 1 2 P 1 0
1 6 7
= 4 2 2P 1 2 1 P 2 1 P 2 1 2 2P 1 0 5 (2.69)

0 0
P
whose components are obviously distinct in this case. A little eort, along
with Eq. (2.63), leads to a verification of the isochoric constraints, viz.,
tr l = tr L = 0. The symmetric part of l, i.e., the stretching, is given by
44 Soft Solids

2
2 P 1 2 1 1P 2
1 6 
[d] = 4 1  P +  P  P  P 

2 1 1 2 2 2 1 1 2 1 2 1 2
0
3
1   P +  P  P  P


0
2 1 1 2 2 2 1 1 2 1 2 1 2
7
P P
1 2 1 2 2 1 0 5 , (2.70)
0
P

while its skew-symmetric part, viz., the vorticity, is described by

2
0
1 4 1 P P 
[w] = 1 1 2 + 2 2 1 + 1 2 P 1 2 1 P 2

2
0

1  P  P   3
P
1 2 1 + 2 1 2 0
P
2 1 1 2 2 2 1
0 0 5 . (2.71)
0 0

Components for their Lagrangian counterparts D and W are left as an


exercise.

2.8 Exercises

2.8.1 Pure Shear


Derive the components of the deformation gradient F and its inverse F 1 ;
their associated polar fields, i.e., U , v, and R; and the rate fields that de-
scribe this deformation, viz., F,
P L, l, D, d , W , and w, for the motion of
pure shear defined in Eqs. (1.32) and (1.33).

2.8.2 Biaxial Extension


Derive the components of the deformation gradient F and its inverse F 1 ;
their associated polar fields, i.e., U , v, and R; and the rate fields that de-
scribe this deformation, viz., F,
P L, l, D, d , W , and w, for the motion of
biaxial extension defined in Eqs. (1.34) and (1.35).
Deformation 45

2.8.3 Extension Followed by Simple Shear


Derive the components of the deformation gradient F and its inverse F 1 ;
their associated polar fields, i.e., U , v, and R; and the rate fields that de-
scribe this deformation, viz., F,
P L, l, D, d , W , and w for the second phase
of the motion, viz., that of simple shearing, which follows the initial phase
of an extension, as described in Eqs. (1.36) and (1.37). Here, in the second
phase, the aspect ratio n and the stretch are held fixed, i.e., nP = P = 0.

2.8.4 Other Problems


1. In the two-dimensional case, show that for any two orthogonal tensors
R1 and R2 described by angles of rotation 1 and 2 , their product
R1 R2 is another orthogonal tensor, but their sum R1 + R2 does not
produce an orthogonal tensor.
2. Compute the components for the stretching D and vorticity W tensors
of the Lagrangian frame for the homogeneous membrane. How do they
compare with their Eulerian counterparts d and w found in Eqs. (2.70)
and (2.71)?
3. Show that the components of the deformation gradient F and its inverse
F 1 ; their associated polar fields, i.e., U , v, and R; and the rate fields
that describe deformation, viz., F, P L, l, d , and w, for (a) uniaxial ex-
tension, (b) equi-biaxial extension, (c) simple shear, (d) pure shear, (e)
biaxial extension, and (f) extension followed by simple shear are each a
special case of Eqs. (2.61)(2.71), which describe their counterparts for
planar membranes.
4. Show that the transpose of a covariant tensor is a covariant tensor and
that the transpose of a contravariant tensor is a contravariant tensor.
(Hint: Take the transpose of their respective field-transfer laws.) Also
show that the inverse of a covariant tensor is a contravariant tensor and
the inverse of a contravariant tensor is a covariant tensor, if they exist.
5. Show that the inverse of a mixed tensor is a mixed tensor. (Hint: Take
the inverse of the field-transfer law for mixed tensors.) Show that the
transpose of the field-transfer law for mixed tensors is not a mixed ten-
sor in the sense of Eqs. (B.25) and (B.26). What does this imply? [It
is for this reason that only one type of mixed tensor field is considered
here, viz., right covariant, and it is why their transposes have tensor
components defined according to Eq. (A.26)].
6. Show that, like the Lagrangian velocity gradient L, the Lagrangian
stretch U and its inverse U 1 obey the field-transfer operator of a mixed
46 Soft Solids

tensor field with U and U 1 pushing from 0 forward into  as v


and v1 according to Eq. (B.25). Conversely, show that the Eulerian
stretches v and v1 pull back from  into 0 as mixed tensor fields
producing U and U 1 according to Eq. (B.26). Do the stretches U and
v and the velocity gradients L and l all obey the same field-transfer
law? Are they the same or dierent from the transformation law be-
tween U 1 and v1 ?
7. The components of the rotation tensor in Eq. (2.55) for simple shear
were derived assuming a clockwise rotation in Eq. (2.53). Show that
you would get the same result if, instead, Eq. (2.53) were to describe a
counterclockwise rotation, i.e.,  .
8. Consider the motion of a compressible isotropic solid that is uniaxially
stretched according to the mapping
x1 = a X 1 , x2 = t X 2 , x3 = t X 3
where a and t are the axial and transverse stretches, respectively.
What is this motions deformation gradient F? Under what condition
would this motion be isochoric?
Chapter 3

Strain

Strain has been defined and quantified a number of different ways over the
past two centuries. Essentially, strain is a measure of the change in shape
of a localized region in a body B that has been deformed from its reference
configuration 0 (where strain is typically normalized to be zero) into its
current configuration . Strain is a two-state property; it depends upon 0
and .
Unlike stretch, which is uniquely defined, strain is not unique. In fact,
Hill (1968) has shown that one can choose from an infinity of admissible
strains. Nevertheless, only a few have physical significance and are prac-
tical to work with. All measures of strain are equivalent to the infinitesi-
mal strain tensor of linear elasticity (the strain measure you were taught in
your introductory strength of materials course) whenever the magnitudes
of strain are infinitesimal (kk . 1 %, typically, cf. Eq. 3.22). Their dier-
ences become noticeable at larger strains (say, kk & 4 %).
Rule of Thumb: Linear strain  is in error about 1 1/2 % per percent strain.
So how does one choose a suitable strain measure for soft materials
where strains routinely exceed 10 % and frequently exceed 100 %? That is
the topic of this chapter.

3.1 Deformation

The deformation gradient F is the fundamental kinematic field quantifying


shape change. The tricky aspect is that the deformation gradient has a non-
physical rotation R embedded within it, as seen in Eq. (2.13). But any
rotation of an observer, by the principles of physics, cannot cause a change
in shape to occur within any body B, cf. Appendix B, which suggests that
A.D. Freed, Soft Solids: A Primer to the Theoretical Mechanics of Materials, 47
Modeling and Simulation in Science, Engineering and Technology,
DOI 10.1007/978-3-319-03551-2_3, Springer International Publishing Switzerland 2014
48 Soft Solids

strain constructions ought to be based upon stretch tensors U or v and


that selecting stretch tensors over the deformation gradient ought to be the
preferred approach to take. But in practice, solving for stretch is a more
involved process than solving for deformation, as was found to be the case
in Sect. 2.6. So, ideally, we seek formul written in terms of F that can be
expressed in terms of either U or v with the eects from R canceling out.
It just so happens that this can be done. For Lagrangian constructions,
one can define a deformation tensor C as
C = F T F = U T RT RU = U T U (3.1)
where, from Eqs. (2.15) and (A.26), its transposed fields are described by
j
F T = ij FJ JI eI e i and U T = JK ULK LI eI e J . (3.2)
Component notation of Eq. (3.1) is1
j j
CIJ = FIi ij FJ = UIK RKi ij RL ULJ = UIK KL ULJ , (3.3)
j
where U = UIJ eI e J , recalling from Eq. (2.15) that RKi ij RL = KL . The
inverse C 1 of deformation tensor C is
C 1 = F 1 F T = U 1 R1 RT U T = U 1 U T (3.4)
that in component notation becomes
[C1 ]IJ = [F1 ]Ii ij [F1 ]Jj = [U 1 ]K
I
[R1 ]Ki ij [R1 ]Lj [U 1 ]LJ
= [U 1 ]K
I KL
[U 1 ]LJ . (3.5)
These two deformation tensors, which are independent of rotation R, have
dyadic notations of C = CIJ e I e J and C 1 = [C1 ]IJ eI eJ .
For Eulerian constructions, one defines a dierent deformation tensor
b = FF T = vRRT vT = vvT with vT = jk v`k `i ei e j (3.6)

1 As observed for the rotation tensor R stated on p. 30, contraction F T F = C has


j
covariant components FIi ij FJ = CIJ , while contraction FF T = b has contravariant
j
components FIi IJ FJ = b ij . Kronecker deltas IJ and ij are the contravariant and
covariant metrics of Cartesian space in the Lagrangian and Eulerian frames, re-
j
spectively. An application of Eq. (A.42) to F T F species components IK FKi ij FJ
with the rst Kronecker delta being able to contract out, thereby yielding the
j
stated result FIi ij FJ . Likewise, an application of Eq. (A.44) to FF T species
components FI FJ kj with the second Kronecker delta being able to contract
i IJ k
j
out, yielding its stated result FIi IJ FJ .
In general tensor analysis, the metrics ij and IJ of Cartesian space would be
replaced by their Riemannian counterparts, viz., g ij and GIJ [cf., e.g., Sokolniko
(1964)].
Strain 49

that in component notation becomes


bij = FIi IJ FJ = vki RIk IJ R`J v` = vki k` v`
j j j
(3.7)

recalling that RIk IJ R`J = k` . Its inverse b1 is


b1 = F T F 1 = vT RT R1 v1 = vT v1 (3.8)
that in component notation becomes

[b1 ] ij = [F1 ]Ii IJ [F1 ]Jj = [v1 ]ki [R1 ]Ik IJ [R1 ]J` [v1 ]`j
= [v1 ]ki k` [v1 ]`j (3.9)

where b = bij ei e j and b1 = [b1 ] ij e i e j are independent of any


rotation.
Because rotation R is a mixed orthogonal tensor, it follows that RT R 
RRT = I. Consequently, using stretch tensors is not necessary if strain
measures can be constructed from any of these four distinct pairings of
the deformation gradient squared. In fact, these pairings lead to the most
commonly used strain fields. It turns out that b and C and their inverses are
symmetric positive-definite tensor fields (Holzapfel 2000; Lodge 1964).
They are the deformation tensors introduced by Finger (1894) b, Piola
(1833) b1 , Green (1841) C , and Cauchy (1827, pp. 6069) C 1 .2 Fields b
and C 1 map as contravariant tensors [cf. Eqs. (B.17) and (B.18)], whereas
fields C and b1 map as covariant tensors [cf. Eqs. (B.21) and (B.22)].

3.2 Measures of Strain

Each of these pairings of the deformation gradient squared is a measure of


deformation in one configuration that becomes the identity tensor I when
mapped into the other configuration, as determined from their field-transfer
properties, cf. Appendix B. This is a desirable characteristic for construct-
ing strain measures. It means that, in one configuration, a deformation
tensor is a measure of the shape change that has occurred between 0
and , while in the other configuration, it becomes the identity tensor I,

2 The literature commonly refers to C as the right CauchyGreen deformation


tensor and b as the left CauchyGreen deformation tensor. Rather than adopting
these terminologies, this author chooses to name each deformation tensor after
its originator.
50 Soft Solids

indicating that it is in its reference state there. This is a theoretical analog to


the concept of a gage length used for quantifying strain in experiments. As
a consequence, and in accordance with Appendix B, one determines that
C = F T F pushes forward into I because C is covariant; C 1 = F 1 F T
pushes forward into I because C 1 is contravariant; b = FF T pulls back
into I because b is contravariant; and b1 = F T F 1 pulls back into I be-
cause b1 is covariant (Lodge 1964, 1974). With this information at hand,
useful measures for strain can now be proposed.
In the geometry of Riemann, which is the geometry that describes our
universe, distance is measured via a quadratic form. So, considering strain
measures that are quadratic in stretch, it follows that two covariant strain
fields can be defined, one in each configuration. In , one has
   
e(x, t) = 12 I b1 = 12 I F T F 1 , (3.10)
which pulls back into 0 as
   
E (X , t) = 12 C I = 12 F T F I (3.11)
and whose components are quantified by
   
e ij = 12 ij [b1 ] ij and EIJ = 12 CIJ IJ . (3.12)
Likewise, two contravariant strain fields can be defined, one in each con-
figuration. In , one has
   
e(x, t) = 12 b I = 12 FF T I , (3.13)
which pulls back into 0 as
   
E(X , t) = 12 I C 1 = 12 I F 1 F T (3.14)
and whose components are quantified by
   
e ij = 12 bij ij and E IJ = 12 IJ [C1 ]IJ . (3.15)
Here, e is the Almansi (1911) strain tensor, E is the Green (1841) strain
tensor, e is the Signorini (1930) strain tensor3 , and E is the Lodge (1964)

3 This strain measure is sometimes credited to Finger in the literature, but


nowhere in Fingers documents (that this author has found) does he dene such
a strain eld.
Strain 51

strain tensor4 , which Lodge expressed in terms of convected (embedded or


body) tensor fieldsthe likely reason for its obscurity.
Covariant tensors map between configurations  and 0 according to
formul (B.21) and (B.22) obeying a field transfer of
[e(x, t)] = [F 1 ]T [E (X , t)] [F 1 ],
(3.16)
[E (X , t)] = [F]T [e(x, t)] [F].
Therefore, the strains in Eqs. (3.10) and (3.11) represent the same physical
measure for strain; they are just defined over dierent frames of reference.
Likewise, contravariant tensors map between configurations  and 0
according to formul (B.17) and (B.18) obeying a field transfer of
[e(x, t)] = [F] [E(X , t)] [F]T ,
(3.17)
[E(X , t)] = [F 1 ] [e(x, t)] [F 1 ]T .
Therefore, in like manner to the covariant strains, the contravariant strains
in Eqs. (3.13) and (3.14) are the same physical measure for strain; they are
just defined over dierent frames of reference.
The covariant and contravariant strains are physically dierent from one
another. They are said to be dual measures of strain (Lodge 1964, 1974).

3.2.1 Hencky Strain


One often hears of the terminology true strain, a strain measure explored
and developed by Hencky (1928, 1931) whose conceptual origin he at-
tributes to Ludwik. It is a logical 1D strain measure, but it is not a practical
3D strain measure. Hencky strain is defined as EH = ln U with an associ-
ated rotation tensor of RH = ln R because ln F = ln(RU ) = ln R + ln U =
RH + EH . These deformation fields are presented here only for the purpose
of informing the reader about their existence. Taking the logarithm of a
matrix is not easily done (Fitzgerald 1980), nor is taking its rate (Hoger

4 This strain measure is rarely found in the literature, but when it is, it is often
credited to Piola, although in no document of his (that this author has found) does
Piola dene such a strain eld. I credit this strain measure to one of my mentors,
the late Prof. Arthur S. Lodge, not because this strain measure likely originated
with him, but because he was apparently the rst to study it; specically, its
geometric interpretation originates in his textbook entitled Elastic Liquids (Lodge
1964, Chap. 2). It is also the strain measure that arises in the rubberlike liquid,
which is a viscoelastic material model that he derived from molecular statistical
mechanics (Lodge 1956, 1958).
52 Soft Solids

1986). Its inverse operation, however, is described by a well-behaved con-


vergent series: the exponential of a matrix. Specifically, one can write the
identities5
U = exp (EH ) = I + EH + 12 EH2 + 16 EH3 +    ,
U 1 = exp (EH ) = I EH + 12 EH2 16 EH3 +    ,
2 3
(3.18)
R = exp (RH ) = I + RH + 12 RH + 16 RH +  ,
R1 = exp (RH ) 1 2 1 3
= I RH + 2 RH 6 RH +   
from which one derives a useful pair of approximations
1 U U 1  = E + O  1 E 3 ,
2 H 3 H
1 R R1  = R + O  1 R3 
(3.19)
2 H 3 H
that provide third-order accurate estimates for EH and RH . Obviously,
Hencky strain is a mixed tensor field (Freed 1995), because U and U 1 are
both mixed tensor fields.

3.2.2 Infinitesimal Strain/Rotation Relationships


At this juncture in your study of finite deformations and their ensuing strain
measures, it is natural to ask: How do the linear strains and rotations that
are taught in courses on strength of materials and linear elasticity dier
from their finite cousins? Linear strain fields result whenever one considers
an infinitesimal displacement vector
u = x  (x, t) (3.20)
and its associated displacement gradient
@u i
G = I F 1 with components G ij = (3.21)
@x j
through which one expresses the classic tensor fields defining infinitesimal
strain
 
 = 12 G + G T (3.22)
and infinitesimal rotation
 
! = 12 G G T , (3.23)
5 This clever formulation was shown to the author many years ago by Prof.
Arkady Leonov. The author has not seen it published in the literature.
Strain 53

wherein the strain tensor  is symmetric, i.e.,  = T , while the rotation


tensor ! is skew-symmetric, viz., ! = !T . These are the strain and
rotation fields used in the classical theory of linear elasticity.
Strain  is linear in F, i.e.,  = I 12 (F 1 + F T ), while the strain
measures used in this text are quadratic in F, e.g., E = 12 (I F 1 F T ), the
latter of which is in accordance with Riemannian geometrythe geometry
of our physical worldas we shall now demonstrate.

3.3 Geometric Interpretations of Strain

Consider two neighboring particles P and P 0 in some body B that are con-
nected by an infinitesimal vector dX in the reference configuration 0 and
by another infinitesimal vector dx in the current configuration , as de-
picted in Fig. 2.1. Let the ambient space be flat so that its metric g is
the identity tensor I (Sokolniko 1964). Then, in accordance with Rie-
mannian geometry, the lengths of these two vectors are described by their
Euclidean norms

(dS)2 = dX  I dX = dXI IJ dXJ (3.24)

and
(ds)2 = dx  I dx
= [F] {dX }  I [F] {dX }
(3.25)
= {dX }  [F]T [I] [F] {dX } = {dX }  [F T F] {dX }
= dX  C dX = dXI CIJ dXJ ,
j
recalling that F T F = FIi ij FJ e I e J and that dS = dS(X ) and ds =
ds(x, t). Here, Eq. (2.2) was used to convert dx into dX in the second line
of the expression. An examination of formula (ds)2 = dX  C dX justifies
referring to C as the Lagrangian metric of deformation.
By contracting the covariant strain tensor of Green (1841) defined in
Eq. (3.11) with the contravariant unit vector dX /dS from both the left and
the right, one arrives at the following quadratic form

(ds)2 (dS)2 dX dX dXI dXJ


=  E (X , t) = EIJ . (3.26)
2(dS)2 dS dS dS dS
54 Soft Solids

dH dx
dX dh ds
dS

t0 t

Fig. 3.1 Material surfaces S and S0 , containing particles P and P 0 , deform from
an initial conguration at time t0 into a nal conguration at time t. The initial
separation of these surfaces dH deforms to dh in a manner that is distinct from
the separation of the particles going from dS to ds as dX goes to dx

 of Green strain E , then one


Consequently, if one knows the tensorial state
can quantify a scalar measure (ds)2 (dS)2 /2(dS)2 of the strain acquired
in any unit direction dX /dS directed away from particle P . This result is
well known and agrees with our understanding of Riemannian geometry.
The vast majority of continuum mechanicians use the Green strain tensor
in their applications.
Now let us consider that these same two particles P and P 0 reside on a
pair of nonintersecting material surfaces S and S0 that belong to a family
of surfaces Y with values Y(X ) = Y and Y(X + dX ) = Y + dY whose
descriptions are independent of time, cf. Fig. 3.1. The increment in height
dH that initially separates surfaces S and S0 at P in the direction of normal
dY/dX (cf. Sect. 2.3), as established in the reference configuration 0 , is
quantified by (Lodge 1964, pp. 2632, p. 318)

 
dY 2 dY(X ) dY(X ) dY IJ dY
= I = (3.27)
dH dX dX dXI dXJ

where by the chain rule

dY(X ) dY
dY =  dX = dXI , (3.28)
dX dXI

while the incremental height dh that currently separates these same two
material surfaces, now located in the spatial configuration , is given by
Strain 55

     
dY 2 @Y  (x, t) @Y  (x, t)
= I
dh @x @x
   
dY(X ) @ (x, t) dY(X ) @ (x, t)
= I
dX @x dX @x
  "    # 
dY(X ) @ (x, t) @ (x, t) T dY(X )
= I
dX @x @x dX
   
dY(X ) dY(X ) (3.29)
= [F 1 ]  I [F 1 ]T
dX dX
   
dY(X ) 1 1 T dY(X )
=  [F ] [I] [F ]
dX dX
   
dY(X ) dY(X )
=  [F 1 F T ]
dX dX
dY(X ) dY(X ) dY 1 IJ dY
=  C 1 = [C ]
dX dX dXI dXJ

where F 1 F T = [F1 ]Ii ij [F1 ]Jj eI eJ and where an application of


the chain rule has been made, in accordance with dependencies dY(X ),
dH(X ), and dh(x, t).
By contracting the contravariant strain tensor of Lodge (1964) defined
in Eq. (3.14) with a covariant unit vector of dH dY (dY(X )/dX ) from both the
left and the right, one arrives at the following quadratic form

(dh)2 (dH)2 dH dY(X ) dH dY(X )


2
=  E(X , t)
2(dh) dY dX dY dX (3.30)
dH dY IJ dH dY dH IJ dH
= I
E J
= E .
dY dX dY dX dXI dXJ
Consequently, if one
 knows the tensorial
 state of strain E, one can quantify
a scalar measure (dh)2 (dH)2 /2(dh)2 of the strain occurring in any unit
direction dH/dX directed away from particle P . This result is not as well
known as its dual, Eq. (3.26), yet E is often the appropriate strain measure
to use. In fact, it is the strain measure derived from classical statistical
mechanics when describing the elastic and viscoelastic responses of the
simplest of polymers and elastomers (Lodge 1964, 1974, 1999).
The strain tensors of Green E and Lodge E are dual tensor fields,
both giving a full description of strain but each with a dierent physical
interpretation. Green strain correlates with changes in the length of arc be-
tween neighboring material particles, whereas Lodge strain correlates with
changes in the separation between neighboring material surfaces on which
56 Soft Solids

these two particles reside, as illustrated in Fig. 3.1. Given that Y is chosen
so that 0 < dH = dS in 0 , it follows from the triangle inequality that
0 < dh
ds in . Greens strain tensor maps as a covariant field, while
Lodges strain tensor maps as a contravariant field. They are each viable
strain measures, both physically and mathematically.

3.3.1 An Areal Interpretation


An alternate geometric interpretation to strain can be gotten through an
investigation of Cauchys metric C 1 . Consider an infinitesimal element of
area dA whose unit normal is N , collectively expressed as the vector field
dA = N dA in its Lagrangian representation, or as da = n da in its Eulerian
representation, assuming Cartesian tensor analysis is being utilized. These
two vector fields represent the same physical feature, they are just defined
over two dierent configurations, viz., 0 and .
As proven by Truesdell (1958), the dierentiated areas associated with
these two vector fields are described by Euclidean norms with measure
(dA)2 = dAI IJ dAJ (3.31)
and
[C1 ]IJ
(da)2 = dAI dAJ (3.32)
det C 1
thereby suggesting a strain measure of6
 
Ey = 1 I det(C ) C 1 , (3.33)
2
which is a contravariant tensor field with a weight of 2, cf. Appendix B.
Contracting strain Ey from the left and right by the unit vector dA/dA
leads to
(dA)2 (da)2 dA y dA dAI y IJ dAJ
= E = E . (3.34)
2(dA)2 dA dA dA dA
Consequently, if one knows the tensorial y one can quan-
state of strain E,
2 2
 2
tify a scalar measure (dA) (da) /2(dA) of strain occurring in any unit
direction dA/dA directed away from particle P . This direction aligns with
the normal to its surface element in that dA/dA = N . Regarding this mea-
sure of deformation/strain, Truesdell (1958) wrote:
6 Truesdell (1958) did not propose this strain measure, only a physical inter-
pretation for Cauchys deformation eld C 1 from which such a strain measure
naturally follows.
Strain 57

Theorem 3.1. The elements of area suering extremal changes are normal
to the principal directions of strain, and the greatest (least) change of area
occurs in the plane normal to the axis of least (greatest) stretch; in fact,
if the principal stretches dx/dX satisfy 1  2  3 the corresponding
ratios da/dA satisfy 2 3
3 1
1 2 .

For isochoric deformations, as studied herein, strain E y is equivalent


to Lodges strain measure E defined in Eq. (3.14), because det C 1 = 1.
Equation (3.34) therefore provides an alternate geometric interpretation for
Lodge strain whenever it is used to describe incompressible materials.

3.4 Strain Rates

Before deriving the various strain rates, it is useful to state the material
derivative of the deformation gradient in its various forms; specifically,
they include

FP T = F T l T P P
FP = lF F 1 = F 1 l F T = l T F T
(3.35)
= FL, = LT F T , = LF 1 , = F T LT
where the first two came from rearranging Eqs. (2.19) and (2.24), while the
last two came from dierentiating the identity equation FF 1 = I. These
four identities are used in the following derivations of strain rate.
The material derivative of Greens covariant description for strain E =
1 (F T F I), when expressed in terms of the Eulerian velocity gradient l,
2
is determined to be
 
EP (X , t) = 12 FP T F + F T FP ,
 
[EP (X , t)] = 12 [F]T [l T ] [F] + [F]T [l] [F] (3.36)
T1 T  T
= [F] 2 [l + l] [F] = [F] [d ] [F].
From Eq. (B.43), the Lie derivative for Almansi strain is given by
O O
[ e(x, t)] = [F 1 ]T [EP (X , t)] [F 1 ] with e(x, t)  d . (3.37)
O
From Eq. (B.54) we have e = De/Dt + l Te + el so that
O @e ij k
e ij = eP ij + v + ` ki e kj + e ik ` kj = [F1 ]Ii EP IJ [F1 ]Jj = d ij (3.38)
@xk
where here d = d ij e i e j , i.e., it takes on covariant components.
58 Soft Solids

The material derivative of Lodges contravariant description for strain


E = 12 (I F 1 F T ), when expressed in terms of l, is determined to be
 
P , t) = 1 FP1 F T + F 1 F PT
E(X 2
 
P , t)] = 1 [F 1 ] [l] [F 1 ]T [F 1 ] [l T ] [F 1 ]T
[E(X (3.39)
2
 
= [F 1 ] 12 [l + l T ] [F 1 ]T = [F 1 ] [d ] [F 1 ]T
and, therefore, from Eq. (B.39), the Lie derivative for Signorini strain is
described by7
M M
[e(x, t)] = [F] [EP (X , t)] [F]T with e(x, t)  d . (3.40)
M
From Eq. (B.53), one has e = De/Dt le el T so that
M ij @e ij k j j
e = eP ij + v `ki e kj e ik `k = FIi EP IJ FJ = d ij (3.41)
@x k

where now d = d ij ei e j , i.e., it takes on contravariant components.


In Eq. (2.22), the stretching tensor d is defined as having mixed com-
ponents, i.e., d ij . In Eq. (3.40) for strain rate has contravariant components,
i.e., d ij , while for the strain rate defined in Eq. (3.37), d has covariant com-
ponents, viz., d ij . All are correct. This quandary requires some extra care
when using Cartesian tensors in Eulerian formulations.
The Lie derivatives of both spatial strain measures e and e equate with
the stretching tensor d defined in Eq. (2.22) and, for this reason, d is often
referred to as the strain-rate tensor. This further attests to the claim that
the strain measures of Almansi (3.10) and Signorini (3.13) are dual tensor
fields. It also demonstrates that, in Cartesian tensor analysis, maps between
0 and  need not be bijective (one-to-one and onto) for all fields. Rates
O M
e and e both equate with the mixed stretching tensor d and, therefore,
the mapping from 0 to is many-to-one and onto with respect to strain
rates, which is an exception. In Cartesian analysis, most physical fields and
their rates obey bijective maps between 0 and  but not strain rate.

7 This result proves that the integral equation describing strain derived by the
author (Freed 2010) is an alternative expression of the Signorini (1930) strain
measure, as the derivative of the integral equation reported therein reproduces
the result given in Eq. (3.40).
Strain 59

The material derivative of Greens covariant strain E = 12 (F T F I),


expressed now in terms of the Lagrangian velocity gradient L, is given by
 
EP (X , t) = 12 FP T F + F T FP
 
= 12 LT F T F + F T FL (3.42)
 
= 12 LT C + CL

that in component notation looks like


 
EP IJ = 12 LIK CKJ + CIK LK
J . (3.43)

It is important to point out that EP D = 12 (L + LT ).


Likewise, the material derivative of Lodges contravariant measure for
strain E = 12 (I F 1 F T ) can also be expressed in terms of L with outcome

 
P , t) = 1 FP1 F T + F 1 F PT
E(X 2
 
= 12 LF 1 F T F 1 F T LT (3.44)
 
= 12 LC 1 + C 1 LT

that in component notation becomes


 I 1 KJ J .
EP IJ = 12 LK [C ] + [C1 ]IK LK (3.45)

Like above, we point out that EP D. These strain rates are symmetric
constructions based upon the Lagrangian velocity gradient L where the
deformation metrics of Green C and Cauchy C 1 are required so that the
resulting rates for EP and EP map between 0 and  as covariant and con-
travariant fields, respectively, just as their associated strains do.
From Eqs. (3.42) and (3.44), it is readily shown that

EP = C 1EP C 1 P ;
or EP = C EC (3.46)

consequently, just one physical measure for strain rate exists, it being EP
when expressed as a contravariant field or EP when expressed as a covariant
field. Notably, although EP = C 1EP C 1 , E C 1EC 1 . This will have
ramifications later when constitutive equations are considered that can be
based on either strain or strain rate.
60 Soft Solids

3.4.1 Integration
In applications, one may need to integrate strain rate in order to obtain
strain. In accordance with Eqs. (3.37) and (B.79), Almansi strain e follows
from the integral equation
Z t
1 T
[e(x, t)] = [F ] [EP (X , t0 )] dt0 [F 1 ]
0
Z t
= [F 1 ]T [F((X , t0 ))]T [d ((X , t0 ), t0 )] [F((X , t0 ))] dt0 [F 1 ],
0
(3.47)
while, from Eqs. (3.40) and (B.78), Signorini strain e follows from the
integral equation (Freed 2010)
Z t
P , t0 )] dt0 [F]T
[e(x, t)] = [F] [E(X
0
Z t
= [F] [F 1 ((X , t0 ))] [d ((X , t0 ), t0 )] [F 1 ((X , t0 ))]T dt0 [F]T ,
0
(3.48)
which are equivalent expressions to their definitions given in Eqs. (3.10)
and (3.13); however, these integral equations are not as easy to work with.
The previous integrals specify how the Eulerian strains e and e could
be solved for by integration. Simpler to evaluate would be the integrals that
return the Lagrangian strain measures, viz.,
Z
1 t T
E= L (X , t0 ) C (X , t0 ) + C (X , t0 ) L(X , t0 ) dt0 ,
2 0
Z (3.49)
1 t 0 1 0 1 0 T 0

0
E= L(X , t ) C (X , t ) + C (X , t ) L (X , t ) dt
2 0
although, in most applications, it would be far easier to just compute them
directly via Eqs. (3.11) and (3.14).
No initial conditions are associated with the above integrals for strain
because strain is typically normalized to be zero in the reference frame,
i.e., at time t = 0; hence, their ICs are typically set to 0, which is tacitly
assumed in the above formul.
The results of this section are not as useful, per se, from an application
point of view, as they are essential from a theoretical point of view. They
bring the theory for strain full circle with respect to the calculus.
Strain 61

3.5 Strain and Strain-Rate Fields for Numerical Analysis

Numerical stability issues often arise with material models whenever they
are incompressible or nearly incompressible (Simo and Hughes 1998).
Thus, a variety of variational principles for FE analysis have been de-
veloped (Belytschko et al. 2000; Bonet and Wood 1997) whose various
techniques revolve around the splitting of strain into dilatoric (pertaining
to volume change) and deviatoric (pertaining to shape change) parts. The
Lagrangian frame is selected for constructing our theory for strain. Algo-
rithms are provided to compute the covariant and contravariant distortions
from which distortional strain measures are constructed.
In accordance with the conservation of mass, cf. Sect. 2.2, dilation J
is defined as the ratio of volumes belonging to an infinitesimal volume
element evaluated at two instances in time, specifically
dv
J= = det F so that d ln J = tr(F 1 dF) (3.50)
dV
where J is the Jacobian of the field-transfer map, cf. Appendix B, hence its
notation. Dilatation , as put forward by Hencky (1928), is defined as
 = 13 ln(det F) = 13 ln J with d = 13 tr(F 1 dF) (3.51)
where the rate relationship results from Eq. (A.84). Stretch = d`/dL,
areal stretch
= da/dA, and dilation J = dv/dV are all normalized ra-
tios of geometric objects constructed from incremental extents in length
of line, area of surface, and volume of space, respectively, whose values
in the current configuration  are proportioned by their respective values
from the initial or reference configuration 0 . The notion that the loga-
rithm of dilation (a volumetric stretch) describes dilatation (a volumetric
strain) was developed by Hencky (1928) from an idea that he credits to
Ludwik (Hencky 1931) in Henckys modeling of the high-pressure data of
Bridgman (1923).

3.5.1 Formulation in Terms of Green Strain


It is supposed that a dierential change in the deformation gradient dF
can be split into dilatoric and deviatoric parts; specifically, let (Freed and
Einstein 2013)
dF
x = dF F d  
1
with tr F 1 dF x =0 (3.52)
= dF 3 F d ln(det F)
62 Soft Solids

where 13 F d ln J is the dilatoric part of dF, while dF


x is the deviatoric part
of dF. Tensor dF x is said to be deviatoric because it is traceless, i.e.,
tr(F dF ) = tr(F 1 dF) 3 d = 0 due to Eqs. (2.32) and (3.51). Ten-
1 x
sor dFx represents a deviatoric increment in deformation, with the current
state of deformation F serving as its state of reference.
From the definition for Greens deformation tensor, viz., C = F T F
from Eq. (3.1), comes dC = dF T  F + F T  dF from the product rule,
from which one can advance a definition for the deviatoric rate of Green
deformation as
x T  F + F T  dF
dCx = dF x
 
= dC 2 C d with tr C 1 dCx = 0 (3.53)
= dC 13 C d ln(det C )
where the second line follows from a substitution of Eq. (3.52) into the
first line whose trace is tr(C 1 dCx ) = tr(C 1 dC ) 6 d = tr(F T dF T ) +
tr(F 1 dF) 6 d = 0 because of Eqs. (2.32), (3.51), (A.69), and (A.70).
Consequently, dCx is a deviatoric measure for the rate of deformation dC
with 2 C d being its dilatoric counterpart. From Eqs. (3.51) and (A.74),
one determines that ln J = ln(det F) = 12 ln(det C ), which takes us from the
second line to the third because of Eq. (3.50).
The Green distortion tensor Cx can be solved for by integrating
Eq. (3.53), thereby producing
Z t
x
C =C 2 C (t0 ) d(t0 )
Z0 (3.54)
2 t  0 0 0
=C tr l(t ) C (t ) dt ,
3 0
where use has been made of Eqs. (2.19), (3.51), and (A.84). It follows that
Cx is not deviatoric in the sense that tr(C 1 Cx ) 0, even though its rate dCx
is deviatoric, viz., tr(C 1 dCx ) = 0, which is why Cx is called a distortion.
Dierentiating Eq. (3.54) returns Eq. (3.53).
Algorithms 2.12.3 compute the rotation R and left stretch v tensors
given that the velocity gradient l is the independent kinematic variable, as
is the case with updated Lagrangian FE codes. The deformation gradient
can then be constructed via F = vR in accordance with Eq. (2.13). With
this information available at the beginning and end of an integration step,
information supplied by Algorithms 2.12.3, Algorithm 3.1 integrates the
second line in Eq. (3.54) using the trapezoidal rule to approximate Green
distortion Cx in a manner that is consistent with Algorithms 2.12.3.
Strain 63

Algorithm 3.1 Updating the Green distortion tensor Cx


var tr(ln ) Cn , An F stored variables

function InitGreenDistortion (tr(l0 )) Fn=0


Cx0 I
tr(ln ) Cn tr(l0 ) I F update stored variables
An 0
return Cx0
end function

function GreenDistortion (t, tr(ln+1 ), Fn+1 )


Cn+1 Fn+1
T F
n+1
 
An+1 An + t2 tr(l n ) Cn + tr(l n+1 ) Cn+1 F integrate
Cxn+1 Cn+1 23 An+1 F Green distortion
tr(ln ) Cn tr(ln+1 ) Cn+1 F update stored variables
An An+1
return Cxn+1
end function

In accordance with the above definitions, one can readily write down
dEx = dE C d  1 
1
with tr C dEx = 0 (3.55)
= dE 6 C d ln(det C )
because from Eq. (3.11), i.e., from E = 12 (C I), comes dE = 12 dC with
the deviatoric, Green, strain rate dEx = 12 dCx following by analogy.
For those constitutive theories that select {, Ex } instead of {E } as
their set of independent state variables, one is required to be able to com-
pute the distortional Green strain Ex described by
Z t
x
E =E C (t0 ) d(t0 )
0
  (3.56)
= E 12 C Cx
 
= 1 Cx I ,
2
where the second line follows from the first as a consequence of the first
line in Eq. (3.54), with the third line following from the definition of
Green strain in Eq. (3.11). Like Cx , Ex is not deviatoric in the sense that
tr(C 1 Ex ) 0, even though its rate dEx is deviatoric, which is why Ex
64 Soft Solids

is called a distortional strain and not a deviatoric strain. Dierentiating


Eq. (3.56) returns Eq. (3.55).

3.5.1.1 Field Transfer


The identity tensor I, the Green deformation tensor C , and the Green strain
tensor E = 12 (C I) push forward as covariant tensor fields from 0 into
 to become the Piola deformation tensor b1 , the identity tensor I, and
the Almansi strain tensor e = 12 (I b1 ), respectively, obeying

1
1 T

b = F [I] F 1 ,

T

[I] = F 1 [C ] F 1 ,

T

[e] = F 1 [E ] F 1
according to their definitions and the field-transfer operation of Eq. (B.21).
Consequently, a distortional Almansi strain eN can be constructed from the
mapping

T


[Ne] = F 1 Ex F 1 . (3.57)
In the updated Lagrangian FEA codes, e.g., constitutive equations will of-
ten be expressed in terms of Almansi strain e, not Green strain E .

3.5.2 Formulation in Terms of Lodge Strain


A similar decomposition into dilatoric and deviatoric parts exists for a dif-
ferential change in the inverse of the deformation gradient dF 1 in that
dFx 1 = dF 1 + F 1 d  
 x 1 = 0
 with tr F dF (3.58)
1 1 1
= dF 3 F d ln det F 1

whose derivation requires the use of identity dF 1 = F 1  dF  F 1 from


Eq. (3.35), which is why the dilatoric part, 13 F 1 d ln J, is preceded by a
plus sign here, whereas it is preceded by a minus sign in Eq. (3.52). To
go from the first line to the second requires a property (A.75) of the deter-
minant, i.e., det F 1 = 1/ det F, plus the identity ln(1/x) = ln x. Tensor
x 1 , like dF
dF x , represents a deviatoric increment in deformation, with the
current state of deformation F 1 serving as its state of reference. Beware,
x 1 = d(F
dF x 1 ) (dFx )1 and, therefore, dF
x  dF
x 1 I.
From the deformation tensor of Cauchy, viz., C 1 = F 1 F T from
Eq. (3.4), comes dC 1 = dF 1F T +F 1dF T from which one can advance
Strain 65

the following definition for a deviatoric rate of the Cauchy deformation


tensor
dCx 1 = dF
x 1  F T + F 1  dF
x T
 
= dC 1 + 2 C 1 d with tr C dCx 1 = 0 (3.59)
 1 
= dC 1 13 C 1 d ln det C
whose derivation parallels that of Eq. (3.53).
The Cauchy distortion tensor Cx 1 can be solved for by integrating
Eq. (3.59), thereby producing
Z t
x 1
C =C +2 1
C 1 (t0 ) d(t0 )
Z0 t (3.60)
1 2  0  1 0 0
=C + tr l(t ) C (t ) dt ,
3 0
where use has been made of Eqs. (2.19), (3.51), and (A.84). It follows
that Cx 1 is not deviatoric in the sense that tr(Cx 1 C ) 0, even though its
rate dCx 1 is deviatoric, viz., tr(C dCx 1 ) = 0, which is why Cx 1 is called
a distortion. Dierentiating Eq. (3.60) returns Eq. (3.59). Algorithm 3.2
solves for the Cauchy distortion. Note: Cx 1 (Cx )1 .
From the definition for Lodges strain measure, viz., E = 12 (I C 1 ),
from Eq. (3.14), it follows straightaway that
x = dE C 1 d
dE  
 x =0
tr C dE
1 
1 1
with (3.61)
= dE + 6 C d ln det C

because dE = 12 dC 1 . Therefore, dE x = 1 dCx 1 is the deviatoric part


2
and C 1 d is the dilatoric part of the Lodge strain rate dE. Integrating
this formula establishes the distortional Lodge strain
Z t
x =E
E C 1 (t0 ) d(t0 )
0
  (3.62)
= E 2 Cx 1 C 1
1
 
= 1 I Cx 1 ,
2

where it immediately follows that E x is not deviatoric in the sense that


x ) 0, like Ex , even though its rate dE
tr(EC x is deviatoric, like dEx . This is
x
why E is referred to as a distortional strain instead of as a deviatoric strain.
66 Soft Solids

Algorithm 3.2 Updating the Cauchy distortion tensor Cx 1


var tr(ln ) Cn1 , Bn F stored variables

function InitCauchyDistortion (tr(l0 )) Fn=0


Cx01 I
tr(ln ) Cn1 tr(l0 ) I F update stored variables
Bn 0
return Cx01
end function

function CauchyDistortion (t, tr(ln+1 ), Fn+1 )


1
Cn+1 Fn+1
1 F T
n+1 
Bn+1 Bn + t 1 1
2 tr(ln ) Cn + tr(ln+1 ) Cn+1 F integrate
x 1
Cn+1 1 2
Cn+1 3 Bn+1 F Cauchy distortion
1
tr(ln ) Cn 1
tr(ln+1 ) Cn+1 F update stored variables
Bn Bn+1
return Cxn+1
1
end function

3.5.2.1 Field Transfer


The identity tensor I, the Cauchy deformation tensor C 1 , and the Lodge
strain tensor E = 12 (I C 1 ) push forward as contravariant tensor fields
from 0 into  to become the Finger deformation tensor b, the identity
tensor I, and the Signorini strain tensor e = 12 (b I), respectively, obeying
[b] = [F] [I] [F ]T ,


[I] = [F] C 1 [F ]T ,
[e] = [F] [E] [F ]T
according to their definitions and the field-transfer operation of Eq. (B.17).
Consequently, a distortional Signorini strain eN can be constructed from the
mapping


[e]
N = [F] E x [F ]T . (3.63)
In the updated Lagrangian FEA codes, e.g., constitutive equations will
likely be expressed in terms of Signorini strain e instead of Lodge strain E.
One may question why F and F 1 serve as the mapping fields when
pushing the distortional Lodge and Green strains forward into the Eulerian
Strain 67

frame, instead of using F x and F


x 1 . The reason for this is because distortion
is defined in Eqs. (3.52) and (3.58) via the rates dF x and dF x 1 in which F
and F , not F
1 x and Fx , serve as their respective states of reference. Hence,
1
all fields derived from dFx and dFx 1 will map according to the field-transfer
operations outlined in Appendix B.

3.6 Examples

In the example problems of this text, the spatial gradient grad() = @  /@x
of any field  is considered to be negligible, i.e., the motions considered
produce spatially homogeneous deformation fields. As such, the material
derivative D  /Dt at particle P has no local convective part and reduces
to just its local contribution of @  /@t = P . Consequently, all convective
contributions belonging to field  in the presence of an homogeneous de-
formation must come from handling the global transport of  embedded
within the current configuration  as it moves through space over a con-
tinuum in time; they enter via Oldroyds convective terms in his various
Lie derivatives, cf. Appendix B.

3.6.1 Uniaxial Extension


From the values for the deformation gradient F and its inverse F 1 stated
in Eq. (2.44) for uniaxial extension, the Finger b and Green C deformation
fields have components described by
2 2 3
0 0
[b]  [C ] = 4 0 1 0 5 , (3.64)
0 0 1

from which it follows that the Green E and Signorini e strain tensors pre-
sented in Eqs. (3.11) and (3.13) have like components of
2 2 3
1 1 0 0
[E ]  [e] = 4 0 (1 )/ 0 5, (3.65)
2
0 0 (1 )/
while the Almansi e and Lodge E strain tensors have like components of
2 2 2 0
3
1 4( 1)/ 0
[e]  [E] = 0 1 0 5, (3.66)
2
0 0 1
as defined in Eqs. (3.10) and (3.14).
68 Soft Solids

For their time rates of change, governed by Eqs. (3.36) and (3.39), using
the components of the stretching tensor d found in Eq. (2.47) for uniaxial
extension, one determines that the various strain fields evolve according to
2 P 3

2 0 0
1
EP  [e] P = 4 0 / P 2 0 5 (3.67)
2 P
0 0 / 2

and
2 P 3 3

1 2 / 0 0
[Pe]  EP = 4 0 P 0 5 . (3.68)
2
0 0 P
The Lie derivatives of these Eulerian strain measures are equivalent be-
cause of Eqs. (3.37) and (3.40) with
2P 3
/ 0 0
O M
[ e]  [e]  [d ] = 4 0 /2
P 0 5 (3.69)
0 P
0 /2
in accordance with Eq. (2.47).
Obviously, regarding how strain is processed as a response by a given
material, uniaxial extension experiments cannot distinguish between con-
stitutive eects whose origins stem from strain being either covariant or
contravariant in nature. Another experiment is needed for that purpose.

3.6.2 Equi-biaxial Extension


From the values for F and F 1 stated in Eq. (2.48) for equi-biaxial ex-
tension, the Finger b and Green C deformation fields have components
described by
2 2 3
0 0
[b]  [C ] = 4 0 2 0 5 , (3.70)
0 0 4

from which it follows that the Green E and Signorini e strain tensors put
forward in Eqs. (3.11) and (3.13) have like components of
2 2 3
1 0 0
14 5,
[E ]  [e] = 0 2 1 0 (3.71)
2
0 0 (1 )/
4 4
Strain 69

while the Almansi e and Lodge E strain tensors have like components of
2 2 3
( 1)/ 2 0 0
14
[e]  [E] = 0 ( 2 1)/ 2 0 5 , (3.72)
2
0 0 1 4

as defined in Eqs. (3.10) and (3.14).


For their time rates of change, governed by Eqs. (3.36) and (3.39), us-
ing the components of the stretching tensor d found in Eq. (2.51) for equi-
biaxial extension, one determines that these strain measures evolve accord-
ing to
2 P 3

2 0 0
1
EP  [e]P = 4 0 2 P 0 5 (3.73)
2 P
0 0 4 / 5
and
2 P 3 3

1 2 / 0 0
[Pe]  EP = 4 0 2 /
P 3 0 5. (3.74)
2
0 0 4 3 P
The Lie derivatives of these Eulerian strain measures are equivalent be-
cause of Eqs. (3.37) and (3.40) with
2P 3
/ 0 0
O M
[ e]  [e]  [d ] = 4 0 /
P 0 5 (3.75)
P
0 0 2 /
in accordance with Eq. (2.51).
Like uniaxial extension, an equi-biaxial extension experiment, by itself,
cannot distinguish between constitutive eects that may be attributed to
strain as being either covariant or contravariant in character.

3.6.3 Simple Shear


From the planar values for F and F 1 stated in Eq. (2.52) for simple shear,
the Finger b and Green C deformation fields are described by
2 3 2 3
1 + 2  0 1  0
[b] = 4  1 05 and [C ] = 4 1 +  2 05 (3.76)
0 01 0 0 1
70 Soft Solids

whose inverses, the Piola b1 and Cauchy C 1 deformations, have compo-


nents
2 3 2 3

1 1  0
1 +  2  0
b = 4 1 +  2 05 and C 1 = 4  1 05 , (3.77)
0 0 1 0 0 1
from which it follows that the covariant strains of Green (3.11) and Al-
mansi (3.10) have components of
2 3 2 3
0  0 0  0
14 2 5 14
[E ] =  0 and [e] =   2 05 , (3.78)
2 2
0 0 0 0 0 0
while the contravariant strains of Lodge (3.14) and Signorini (3.13) have
components of
2 2 3 2 2 3
1   0 1   0
[E] = 4  0 05 and [e] = 4  0 05 . (3.79)
2 2
0 00 0 00
Consequently, unlike uniaxial and biaxial extensions, simple shear does
provide a clear distinction between covariance and contravariance in the
states of strain it produces. However, this eect only manifests itself when-
ever a simple-shear deformation becomes finite in the sense that  2 has an
eect.
For their time rates of change, governed by Eqs. (3.36) and (3.39), using
the components of the velocity gradient l found in Eq. (2.59) for simple
shear, one determines that the covariant strains have rates of
2 3 2 3

1 0 P 0 1 0 P 0
EP = 4P 2 P 05 and [Pe] = 4P 2 P 05 , (3.80)
2 0 0 0 2 0 0 0

while the contravariant strains have rates of


2 3 2 3

1 2 P P 0 1 2 P P 0
EP = 4 P 0 05 and [e] P = 4 P 0 05 . (3.81)
2 0 00 2 0 00
The Lie derivatives of these Eulerian strain measures are equivalent be-
cause of Eqs. (3.37) and (3.40) with
2 3
0 P 0
O M 14
[ e]  [e]  [d ] = P 0 05 (3.82)
2 000
in accordance with Eq. (2.60).
Strain 71

The choice of strain type, viz., covariant or contravariant, is a charac-


teristic property of a material and, therefore, of its constitutive law. Its
selection is not a consequence of any physical law, although both strain
measures have been argued from physical grounds. Hence, simple shear,
as a characterization experiment, has great importance and value; however,
it is not a simple experiment to execute in the laboratory. A better experi-
ment, in this regard, is the homogeneous planar membrane experiment.

3.6.4 Homogeneous Planar Membranes


From the values for F and F 1 stated in Eq. (2.61) for planar membranes,
the covariant Green deformation C of the material configuration 0 has
components
2 2 3
1 (1 + 22 ) 1 2 (1 + 2 ) 0
[C ] = 4 1 2 (1 + 2 ) 22 (1 + 12 ) 0 5, (3.83)
0 0
2

while the contravariant Finger deformation b of the spatial configuration


 has components
2 2 3
1 + 12 22 1 22 + 2 21 0
[b] = 41 22 + 2 21 22 + 22 21 0 5 (3.84)
0 0
2

whose dierences are subtle, but nonetheless real. The inverse of the Green
deformation tensor [which is the measure for deformation actually used by
Cauchy (1827)] is contravariant and has components
2 2 (1 +  2 ) ( +  ) 0 3

1 2 1 1 2 1 2
1
C = 2 4 1 2 (1 + 2 ) 21 (1 + 22 ) 0 5, (3.85)

0 0
4

while the inverse of Fingers deformation tensor is Piolas covariant de-


scription for deformation, which has components
2 2 3

1 2 + 22 21 1 22 2 21 0
1
b = 2 41 22 2 21 21 + 12 22 0 5 , (3.86)

0 0
4
72 Soft Solids

wherein
= 1 2 (1 1 2 ) is the areal stretch defined in Eq. (1.23).
From these deformation fields, the covariant strain tensor of Green, E =
1 (C I), defined in the material frame  , has components
2 0
2 2 3
1 (1 + 22 ) 1 1 2 (1 + 2 ) 0
1
[E ] = 4 1 2 (1 + 2 ) 22 (1 + 12 ) 1 0 5, (3.87)
2
0 0
1
2

while the contravariant strain tensor of Lodge, E = 12 (I C 1 ), which is


also defined over the reference configuration 0 , has components
2 2 3

22 (1 12 ) 1 2 (1 + 2 ) 0
1 4
[E] = 1 2 (1 + 2 )
2 21 (1 22 ) 0 5. (3.88)
2
2
0 0

4
2

Likewise, the covariant strain tensor of Almansi, e = 12 (I b1 ), defined


in the spatial frame , has components
2 2 3

22 22 21 1 22 + 2 21 0
1 4
[e] = 1 22 + 2 21
2 21 12 22 0 5, (3.89)
2
2
0 0

2 4

while the contravariant strain tensor of Signorini, e = 12 (bI), also defined


over the current configuration , has components
2 2 3
+  2 2 1  2 +  2 0
1 4 1 2 1 2 2 21 2 2 22 1
[e] = 1 2 + 2 1 2 + 2 1 1 0 5. (3.90)
2
0 0
1
2

Recall that the covariant strains E and e are measures of separation be-
tween neighboring material particles, while the contravariant strains E and
e are measures of separation between neighboring material surfaces, or,
equivalently, because the deformation is isochoric, E and e are measures
of their change in area. One measure for each type is given in the mate-
rial configuration 0 , and one measure for each type is given in the spatial
configuration .
The time rates of change of these strain fields become a bit messy and,
therefore, are not written out here in component form. Even so, they are not
dicult to derive. The Lie derivatives of the Eulerian strain measures obey
O M
[ e]  [e]  [d] because of Eqs. (3.37) and (3.40), whose components are
listed in Eq. (2.70).
Strain 73

3.7 Exercises

3.7.1 Pure Shear


Derive the components for the deformation tensors b and C and their in-
verses b1 and C 1 for the motion of pure shear defined in Eqs. (1.32) and
(1.33). Then construct the four strain tensors E , E, e, and e and their time
O M
P eP , e,
rates of change EP , E, P e, and e. Do this using the deformation, F
and F 1 , and velocity, l and L, gradients that you derived in last chapters
exercises.

3.7.2 Biaxial Extension


Derive the components for the deformation tensors b and C and their
inverses b1 and C 1 for the motion of biaxial extension defined in
Eqs. (1.34) and (1.35). Then construct the four strain tensors E , E, e,
O M
P eP , e,
and e and their time rates of change EP , E, P e, and e. Do this using the
deformation, F and F 1 , and velocity, l and L, gradients that you derived
in last chapters exercises.

3.7.3 Extension Followed by Simple Shear


Derive the components for the deformation tensors b and C and their in-
verses b1 and C 1 for the motion of a simple shear following an axial
prestretch, as defined in Eqs. (1.36) and (1.37). Then construct the four
O
P eP , e,
strain tensors E , E, e, and e and their time rates of change EP , E, P e,
M
and e. Do this using the deformation, F and F 1 , and velocity, l and L,
gradients that you derived in last chapters exercises.

3.7.4 Other Problems


(1) The strain tensor introduced by Biot (1939) is often referred to as the
engineering strain tensor of finite deformation analysis, as it produces
a strain of  = 1 in simple extension, in accordance with how strain
is taught in a strength of materials course to engineers. Biot strain is
a mixed tensor field defined in the reference configuration 0 by the
formula Ey (X , t) = U I. Another linear strain tensor was introduced
by Bell (1983), at the suggestion of Ericksen, to describe the plastic
74 Soft Solids

response of metals. BellEricksen strain is also a mixed tensor field. It


is defined in the current configuration  by the formula eO (x, t) = v I.
Prove that these are the same physical strain measure and that they are
just defined over dierent configurations. (Hint: Push Ey forward from
0 into .)
(2) Show that the components of the deformation tensors b and C , their
inverses b1 and C 1 , and the four strain tensors E , E, e, and e, for
i) uniaxial extension, ii) equi-biaxial extension, iii) simple shear, iv)
pure shear, v) biaxial extension, and vi) extension followed by sim-
ple shear are special cases of Eqs. (3.83)(3.90) which describe their
counterparts for planar membranes.
(3) Show that when C 1EC 1 is pushed forward from 0 into  via the
contravariant mapping (B.15), it produces the Almansi strain e, which
is a covariant tensor. Similarly, show that when C EC is pushed for-
ward from 0 into  via the covariant mapping (B.19) it produces the
Signorini strain e, which is a contravariant tensor. Comment on what
the Cauchy C 1 and Green C deformation tensors are doing here.
(4) Is the linear strain of Eq. (3.22) valid for applications of small displace-
ments with large rotations? Justify your answer. Think about a blade
of grass blowing in the wind or the motion of a jump rope.
(5) Take the formul in Sect. 3.5 that decompose strain into their isotropic
and deviatoric parts, where they have been defined for the Lagrangian
frame 0 , and push them forward into their counterparts of the Eule-
rian frame .
(6) Robert Hooke (16351703), the man credited with being the founding
father of solid mechanics, first proposed a mechanical model for the
springing of solid bodies, from which he then conjectured an em-
pirical law, i.e., a Rule or Law of Nature. Hooke expressed both
his model, which was constructed from experimental observations,
and his law, which was deduced by reason, in terms of words, not
mathematics. When his words are translated into formul, using the
notations and terminology of today, one arrives at (Moyer 1977)

f = k(1 1 ) Hookes model,


f = k( 1) Hookes law,

where f is the force applied, k is the spring constant of the material


or structure, and is the stretch or ratio of deformed to undeformed
lengths. Moyer (1977), a scientific historian, concluded that:
Strain 75

In general, his (Hookes) mechanical model is quantitatively in-


compatible with his empirical law.
Demonstrate why Moyer was able to come to this conclusion by draw-
ing the responses of Hookes model and law in a common figure
and explain what you observe. For the Lagrangian frame, speculate
on what finite-strain measure Hookes law suggests. Would the same
strain measure apply to Hookes model? If not, what one would? Un-
der what physical limitations, if any, could Hookes model and his
law become equivalent?
Chapter 4

Stress

The concept of stress traces back nearly two centuries to the published
works of Cauchy (1827). Cauchy generalized Eulers concept of pres-
sure and the hydrodynamic laws that Euler derived some 70 years ear-
lier. Cauchy made the notion of stress precise. He surmised that a
body responds to externally applied loads by transmitting forces internally
throughout the body via a matrix valued field that now bears his name:
Cauchy stress. Not only did Cauchy develop the concept of stress, but he
also derived the physical conservation laws that apply to stress. In doing
so, he generalized Eulers theory for an inviscid fluid.
Once again, let us consider a body B whose reference configuration 0
is affiliated with an initial time t0 . Over the course of time, body B is sub-
jected to externally applied forces that cause it to deform into its current
configuration , which is affiliated with present time t. In Fig. 4.1, a parti-
cle P residing on a material surface S belonging to body B is investigated
by (fictitiously) cleaving the body along S while simultaneously applying
a resultant force distribution over surface S sufficient to keep the cleaved
body in equilibrium with its surroundings, as if it were whole.
Consider a distributed resultant force df (t, da) acting at time t over
some infinitesimal area da that surrounds particle P and belongs to surface
S in . Cauchy (1827) postulated that this distributed force of infinitesimal
extent is equivalent to a point force of traction t applied at the particle P ,
scaled by the dierential area da(x, t) over which df acts; this is Cauchys
postulate
df = t da or df i = t i da where t = t(n ; x, t), (4.1)

A.D. Freed, Soft Solids: A Primer to the Theoretical Mechanics of Materials, 77


Modeling and Simulation in Science, Engineering and Technology,
DOI 10.1007/978-3-319-03551-2_4, Springer International Publishing Switzerland 2014
78 Soft Solids

n t
N T

da
dA

0
e2

t0 e1 t
e3

Fig. 4.1 Body B is deformed from its original shape at time t0 , i.e., from cong-
uration 0 , into a nal deformed shape at time t, viz., into conguration . This
sets up a traction vector t at particle P acting on a surface S over an innitesimal
area da with unit normal n. All elds are quantied in a common Cartesian basis
(e1 , e2 , e3 ) in R3 . Vector T is the pseudo traction vector dened on 0 , often
referred to as the nominal traction

wherein n(x, t) is the unit normal to surface da, cf. Eq. (2.11), while the
traction vector t has physical units of force per unit deformed area.
Cauchys fundamental theorem for stress establishes stress as a linear
operator on  that maps a normal to a surface into a vector of traction
acting on that surface, all of these fields being located at particle P , i.e.,
t(n ; x, t) = T (x, t) n(x, t) (4.2)
or in component notation
t i = T ij n j , (4.3)
where T is the Cauchy (or true) stress tensor1 which, like the traction
vector t, has units of force per unit deformed area.
It can be readily argued that stress maps as a contravariant tensor field
by examining Eqs. (4.2) and (4.3). Normal n maps as a covariant vector
field because of Eq. (2.11). Traction t behaves like a tangent to a curve, so
its index maps like a contravariant vector, as in Eq. (2.2). Consequently,
1 Cauchy stress is expressed in an uppercase font, as if it were a Lagrangian eld,
but it is not; Cauchy stress is an Eulerian eld. The notation T is adopted for
historical reasons. T is the commonly accepted notation for Cauchy stress when
written in a roman font [cf. Truesdell and Noll (2004)]. The engineering stress
that associates with the innitesimal strain  of Eq. (3.22) is typically denoted
as  , which appears in the linear theory of elasticity.
Stress 79

the conservation of indices from tensor analysis requires that the indices
of stress T ij must map in a contravariant manner with the jth contravariant
index of T ij contracting with the covariant index of n j , leaving the ith
contravariant index of T ij to associate with the contravariant index of t i
on the left-hand side of the equation. But the field-transfer properties of T
are a bit more complicated than this. Cauchy stress also has a field-transfer
property known as weight, which is discussed in more detail in Sect. 4.4.

4.1 Kirchho Stress

Another stress tensor defined over  is often selected for use. In fact, it is
the stress tensor used in this text when working in the spatial configuration
, viz., the Kirchho (1852) stress tensor defined by
s(x, t) = det(F) T or s ij = det(F) T ij , (4.4)
which has units of force per unit undeformed area (as if it was quantified
in the reference configuration 0 ), but the tensor field s, itself, resides in
the deformed configuration .
Whenever materials are incompressible, as they are assumed to be in
this text, for the most part, then the stress tensors of Cauchy and Kirchho
become numerically equivalent because det F = 1 follows from the iso-
choric constraint. Throughout this text, the Kirchho stress s will be the
stress field used most often when working in the Eulerian frame.

4.2 Conservation of Momenta

The stress equations of motion derived by Cauchy (1827) are a continuum


representation of Newtons second law of motion, i.e., f = ma. The stress
equations of motion state that2
div T + b = % a (4.5)
or, in component notation,
@T ij
+ b i = % a i, (4.6)
@x j
2 Seventyyears before Cauchys work, Euler derived a eld equation that de-
scribes Newtons laws of motion for an inviscid uid, i.e.,
grad(p) + b = % a,
80 Soft Solids

where b is a body force, e.g., % g where g is the acceleration due to gravity


and % is the mass density, while a is the acceleration of particle P through
space R3 . The divergence of stress produces a vector field that is defined
by div T = @T ij /@x j ei .
This physical constraint is satisfied, to a good approximation, for the
motions addressed in this text. Here the loading rates are considered to
be quasi-static in that % a  0. Furthermore, the extent of the body force
b = % g is typically much smaller than the magnitudes of traction t that
are imposed onto the test samples. This leaves the conservation of mo-
mentum for static equilibrium, viz., div T = 0, which is trivially satisfied
because the loading histories considered herein are taken to be spatially
homogeneous, i.e., they are without divergence. Although this conserva-
tion law has but a minor role to play in your current study of the mechanics
of materials and their characterization, its importance takes center stage in
applications addressing the mechanics of structures, e.g., in FE and CFD
codes.
The conservation of linear momentum combines the formula governing
the conservation of mass (2.6) with the stress equation of motion (4.5) to
produce a single governing field equation described by
Dp
+ div(v) p = b + div T , (4.7)
Dt
where p(x, t) = %v is the momentum per unit volume at particle P and
time t. This equation simplifies to
@p
+ div(v) p = b + div T (4.8)
@t
for fluids and soft solids because @p/@x = 0, which follows from %P = 0
and Eq. (2.6) because of the assumption of material incompressibility. The
above expressions are often used in CFD codes as the physical law gov-
erning fluids, whereas Eq. (4.5) is more commonly used in FE codes as the
physical law governing solids.
where p is the internal hydrodynamic pressure, which Cauchy described in terms
of stress as p = 13 tr T . Seventy additional years passed before Cauchys genius
was able to generalize Eulers governing equation for hydrodynamics to that of a
general material. Of this, Truesdell (1961) wrote:
Nothing is harder to surmount than a corpus of true but too specic
knowledge; to reforge the tradition of his forebears is the greatest origi-
nality a man can have.
Stress 81

The conservation of angular momentum has greater significance


pertaining to the topics covered in this text. It requires that the Cauchy
stress must be symmetric
T = TT so T ij = Tji , ) s = sT so s ij = sji . (4.9)
The need for Cauchy and Kirchho stress tensors to be symmetric does
not imply that all useful stress measures must be symmetric, too. One very
important stress tensor is not symmetric, which is now introduced.

4.3 First PiolaKirchho Stress

The Kirchho stress s(x, t) is a useful measure for stress in applications of


theory, but it is a dicult measure to directly quantify via experiment. One
vector field in Fig. 4.1 has not yet been discussed, viz., T . This pseudo
traction vector, drawn as a dashed line in Fig. 4.1, lies in the same direc-
tion as the actual traction vector t. Their magnitudes, however, may vary
slightly, at least for compressible materials, because {T } = det(F) {t}.
This nominal traction vector is defined by Ogden (1984):
df = T dA where T = T (N ; X , t) (4.10)
consequently, from Eqs. (4.1) and (4.10), one has
t i da = T i dA, (4.11)
where N is the unit normal to surface S defined in 0 . Notice that
T (X , t) = T i ei ; in other words, field T depends on material coordinates
via its dependence upon X ; however, because it represents a force in the
current frame, it has a spatial index associated with it, viz., i. Vector T
is, in a certain sense, a two-state field, which is why it is typeset in an up-
right blackboard bold font instead of a slanted one (cf. Holzapfel (2000,
pp. 109113) for a detailed discussion of the subtleties involved here).
The normal to a surface is a covariant vector and must map accordingly,
viz., via Eqs. (2.11), (B.8), and (B.12); that is,

{n(x, t)} = [F 1 ]T {N (X )} or n i = [F1 ]Ii NI . (4.12)


However, the juxtaposition of a normal vector with an infinitesimal element
of area, as a single entity, behaves somewhat dierently; it maps according
to Nansons formula given by
82 Soft Solids

{n(x, t)} da(x, t) = det(F) [F 1 ]T {N (X )} dA(X )


[F 1 ]T
or {da(x, t)} = {dA(X )} (4.13)
det F 1
that in component notation becomes

[F1 ]Ii
n i da = det(F) [F1 ]Ii NI dA or dai = dAI , (4.14)
det F 1

with like terms appearing in the Truesdell strain E y of Eq. (3.33). Field-
transfer mappings that depend upon the determinant of the transforma-
tion operator, also known as the Jacobian of the transformation, as occurs
here, produce what are called weighted fields (Lodge 1974; Oldroyd 1950;
Sokolniko 1964). Equation (4.13) contains the only vector-valued field
transfer with weight addressed in this text, cf. Appendix B.
From the above results, one is led to the first PiolaKirchho stress ten-
sor P (Kirchho 1852; Piola 1833). As with Cauchys theorem, P is a lin-
ear operator that maps one vector into another, specifically, a normal vec-
tor into a traction vector, but here the mapping takes place in a Lagrangian
frame instead of the Eulerian one used by Cauchy, viz.,

T (N ; X , t) = P (X , t) N (X ), (4.15)

with components

T i = P iI NI . (4.16)

We see that, like the deformation gradient F(X , t), stress P (X , t) has one
index in the reference configuration 0 and the other in the current config-
uration , i.e., it, too, is a transformation map, which is why it is generally
easier to quantify via experiment than, say, the Kirchho stress s. This is
why it is typeset in an upright font, like F.
The first PiolaKirchho stress relates to the Kirchho stress via the
formula
j
s = PFT or s ij = P iI FI ) P F T = FP T , (4.17)

where the latter relationship follows from the conservation of angular mo-
mentum, viz., Eq. (4.9); consequently, P is not a symmetric field (neither
are F, L, or R), although its elements may be symmetric from time to time.
Stress 83

4.4 Second PiolaKirchho Stress

The Kirchho stress s(x, t) is an absolute, contravariant, tensor field; con-


sequently, from Eq. (B.16), it pulls back from  into 0 according to the
mapping
[S (X , t)] = [F 1 ] [s(x, t)] [F 1 ]T ) S (X , t) = F 1 P (X , t) (4.18)
or, in component notation, according to the formula
SIJ = [F1 ]Ii s ij [F1 ]Jj ) SIJ = [F1 ]Ii P iJ , (4.19)
where S (X , t) is the so-called second PiolaKirchho stress tensor (Kirch-
ho 1852; Piola 1833) that, after Cauchy stress T , is the most popular
measure for stress in use today. It is the Lagrangian stress tensor.
To complete the set of possible mappings between stress tensors, it fol-
lows from Eqs. (4.4) and (4.18) that
[S (X , t)] = det(F) [F 1 ] [T (x, t)] [F 1 ]T , (4.20)
which is the pull-back operator of a weighted contravariant tensor field,
in this particular case, the mapping of Cauchy stress into the second
PiolaKirchho stress; cf. Appendix B.
In Eq. (4.18), S maps as an absolute contravariant field, whereas in
Eq. (4.20), S maps as a relative (with weight 1) contravariant field. How
can this be? This fact is one of the reasons why the topic of stress is of-
ten dicult for students to grasp. The dierence between Eqs. (4.18) and
(4.20) resides in the physical units that are assigned to the Eulerian fields:
T has units of force per unit deformed area, whereas s has units of force
per unit undeformed area, like S . To change the units of area in a mapping
from 0 to  requires a weighted field-transfer operation. If the physical
units are the same in both configurations, as is the case between s and S ,
then an absolute field-transfer operation applies.
Again, because our interest in this book resides predominantly with
incompressible materials, it is sucient to employ stress fields that obey
an absolute field-transfer map in our applications, viz., the Kirchho and
second PiolaKirchho stresses present in Eq. (4.18).

4.5 Stress Rates

Because the Kirchho stress s and second PiolaKirchho stress S are the
same physical measure of stress, although defined in dierent frames, and
because they are contravariant tensor fields, it necessarily follows that their
84 Soft Solids

time rates of change are related according to the maps given in Eqs. (B.39)
and (B.40), in other words, via
M
[ s(x, t)] = [F] [SP (X , t)] [F]T
M
or [SP (X , t)] = [F 1 ] [ s(x, t)] [F 1 ]T , (4.21)
M
where s(x, t) = Ds/Dt ls sl T is a Lie derivative for stress, a.k.a. Ol-
droyds (1950; 1970) upper-convected stress rate.

4.5.1 Integrate for Stress


M M M
Often s is specified by a constitutive equation , e.g., s = s(s, F 1 , l), that
will need to be solved to determine the actual state of stress s. Solutions can
be acquired by one of two pathways. The most straightforward approach
would be to (numerically) integrate the ODE:
M @s
sP (x, t) = s(s, F 1 , l; x, t)  v + ls + sl T . (4.22)
@x
Alternatively, usually in the older literature, one can encounter solutions
cast as integral equations that, in our construction, would look like (cf.
Eq. B.78)
[s(x, t)] = [F] [S0 ] [F]T
Z t
M
+ [F] [F 1 (x 0 )] [s(s, F 1 , l; x 0 , t0 )] [F 1 (x 0 )]T dt0 [F]T , (4.23)
t0
where x 0 = (X , t0 ) locates particle P at time t0 with S0  s(X , t0 ) be-
ing the initial condition (IC). This is Oldroyds (1950) convected integra-
tion, i.e., the objective integration of a Lie rate which, in this case, is for
M
Kirchho stress s. Recall that the Lie derivative of stress s, which appears
in the integrand, is to be supplied by an outside function, viz., some con-
M
stitutive equation s(s, F 1 , l), so the problem becomes well posed in the
end.
The above integral equation is written for the spatial configuration .
Often, especially for solids, it is easier to map the constitutive equation into
the material configuration 0 and perform the integration there, in which
case Eq. (4.23) becomes
Z t
S (X , t) = S0 + P , F, L; X , t0 ) dt0 ,
S(S (4.24)
t0
Stress 85

where SP = SP (S , F, L) is the constitutive equation with S0 = S (X , t0 )


being the initial value. Usually, but not always, S0 = 0. Once S is known,
s can be calculated via its field transfer, viz., [s] = [F] [S ] [F]T .

4.6 The Extra Stress

When analyzing incompressible materials, as we are doing in this text, the


constitutive equation will only provide part of the state of stress. Curi-
ously, the other part comes from the BVP actually being solved. How this
works is that a Lagrange multiplier, denoted herein as }, is introduced
into the formulation. This Lagrange multiplier produces a pseudo state of
hydrostatic pressure, which in turn induces a pseudo dilation, so that the
net outcome becomes a state of stress which associates with an isochoric
deformation, i.e., the material behaves as if it were incompressible, in a
mathematical sense.
The complete state of stress caused by any deformation history imposed
on an incompressible material is called the extra stress and is defined by
= S + } C 1 or equivalently  =s+}I (4.25)
or in terms of their components
IJ = SIJ + } [C1 ]IJ or equivalently ij = s ij + } ij , (4.26)
where } is a Lagrange multiplier introduced to ensure a condition of ma-
terial incompressibility, viz., det C = 1 or tr L = 0. Like the stresses s and
S , a contravariant mapping exists between the Eulerian extra stress  in
 and the Lagrangian extra stress in 0 in that
[ ] = [F 1 ] [] [F 1 ]T or [] = [F] [ ] [F]T (4.27)
and, therefore, they map according to Eqs. (B.15) and (B.16).
This means that their rates must map as contravariant fields, too, i.e.,
according to Eqs. (B.39) and (B.40), implying that
M M
[P ] = [F 1 ] [] [F 1 ]T and [] = [F] [P ] [F]T , (4.28)
where these extra-stress rates relate to their physical stress rates via the
formul
M M
P = SP + }P C 1 2} EP and  = s + }P I 2} d (4.29)
M
wherein [ s] = [F] [SP ] [F]T and [d ] = [F] [EP ] [F]T , which are Eqs. (4.21)
and (3.40), respectively, and where use has been made of the identity CP 1 =
2EP that follows from the definition of Lodges strain tensor in Eq. (3.14).
86 Soft Solids

4.7 Hills Constitutive Inequalities

For an admissible pair of conjugate thermodynamic variables, e.g., the


second PiolaKirchho stress S of Eq. (4.18) and the Green strain E of
Eq. (3.11) (Ogden 1984), an incremental change in the work being done on
some particle P within body B is given by
 
dW = tr S dE = SIJ dEJI per unit reference volume, (4.30)
subject to the thermodynamic requirement that
I I
dW = S W dE  0, (4.31)

where an integration can take place over any closed cycle in strain. A
perpetual motion machine is any hypothetical contraption that violates this
integral inequality, which, by its very definition, cannot occur in Nature.
Stability and uniqueness of a solution are important mathematical prop-
erties that lie beyond the confines of thermodynamics or physics in general.
A by-product of such a mathematical analysis leads to what is referred to as
a constitutive inequality. Drucker (1959) and Hill (1957, 1958) were pio-
neers in this field of study. Whenever S and E are taken to be the defining
conjugate pair, and the material is simple in the sense of Noll (1958, 1972),
then the associated mathematical criterion for uniqueness and stability of a
solution is straightforward to implement. It is Hills (1957; 1968) constitu-
tive inequality:
 
tr dS dE > 0 8 dE 0. (4.32)
The choice of a stress/strain conjugate pair is not unique (cf. Ogden (1984)
for a thorough treatment of this topic). The choice made here is the one
selected most often by mechanicians.
Convexity of potential surfaces in state space is another means for en-
suring mathematical stability of a solution. As this book was going to
press, Nicholson (2013) constructed a technique by which convexity could
be tested for a given potential function, previously, a daunting problem.

4.7.1 Incompressible Materials


Because the extra stress is the actual stress occurring at a particle P
within a deforming, incompressible, material body B, the actual work
being done at P is
     
dW = tr dE = tr S dE + } tr C 1 dE (4.33)
Stress 87

per unit (reference) volume, with its associated constitutive inequality


being
   
tr d dE = tr (dS + C 1 d} + } dC 1 ) dE
     
= tr dS dE + tr C 1 dE d} + } tr dC 1 dE > 0 (4.34)
for all dE 0.
From the incompressibility constraint of tr L = tr l = tr d = 0, one gets
tr(C 1 dE ) = 0, because tr(C 1 EP ) = tr([F 1 ][I][F 1 ]T  [F]T [d ][F]) =
tr d = 0 and, as such, Eq. (4.33) reduces to
   
dW = tr dE = tr S dE , (4.35)
while Eq. (4.34) reduces to
     
tr d dE = tr dS dE 2} tr dE dE > 0, (4.36)
recalling that dC 1 = 2 dE. In other words, the Lagrange multiplier }
has no eect on the work that is being done; however, it can aect the
constitutive inequality that designates the boundary between a stable and
an unstable material response. Hill was the first to point this out.

4.7.2 Eulerian Formulations


Equations (4.30) and (4.32), written in the reference configuration 0 , are
for the rate of working and for material stability. They push forward, in
rate form, into the current configuration  as
 
WP = tr s d per unit (reference) volume, (4.37)
and
M 
tr s d > 0 8 d 0. (4.38)
Equation (4.37) is from the manipulation W P = tr(S EP ) = tr([F][S ][F]T 
O
[F 1 ]T [EP ][F 1 ]) = tr(s e) = tr(s d ), while Eq. (4.38) comes from the field
MO M
transfer tr(SP EP ) = tr([F][SP ][F]T  [F 1 ]T [EP ][F 1 ]) = tr( s e) = tr( s d )
where Eqs. (3.37), (4.18), (4.21), (A.70), (B.15), (B.19), (B.39), and (B.43)
have been used along with the identities F 1 F = F T F T = I.
Equation (4.37) applies for both compressible and incompressible ma-
terials, because of Eq. (4.35), but Eq. (4.38) only applies for compressible
materials. For incompressible materials, one must push Eq. (4.36) forward,
88 Soft Solids

in its rate form, from the reference configuration 0 into the current con-
figuration , which leads to the inequality (Hill 1968):
M  M   
tr  d = tr s d 2} tr d 2 > 0 8 d 0. (4.39)
This is the correct form of Hills constitutive inequality for one to use in
Eulerian formulations. This follows from trace manipulations of tr(P EP ) =
MO M
tr([F][P ][F]T  [F 1 ]T [EP ][F 1 ]) = tr( e) = tr( d ) and from tr(SP EP ) +
} tr(CP 1 EP ) = tr([F][SP ][F]T  [F 1 ]T [EP ][F 1 ]) + } tr([F][CP 1 ][F]T 
MO
P
[F 1 ]T [EP ][F 1 ]) = tr( s e) 2} tr([F][E][F] T  [F 1 ]T [EP ][F 1 ]) =
MO MO M
tr( s e) 2} tr(e e) = tr( s d ) 2} tr(d 2 ) where Eqs. (3.1), (3.4), (3.37),
(3.40), (4.18), (4.21), (4.28), (A.61), and (A.70) have been used.
Note: Incompressibility requires that tr d = 0. This, however, does not
imply that tr(d 2 ) is also zero; in fact, tr(d 2 )  0 for isochoric motions.

4.8 Stresses for Numerical Analysis

Strain rates were decomposed into additive dilatoric/deviatoric splits in


Sect. 3.5. The reason for introducing such a split is to be able to adapt ones
theories to variational principles that are employed by FE and CFD codes
so that they can better handle numerical instability issues that otherwise
arise from constitutive equations for incompressible or nearly incompress-
ible materials (Belytschko et al. 2000; Bonet and Wood 1997). A hydro-
static/deviatoric split in stress is presented here that is compatible with the
dilatoric/deviatoric splits in strain rate introduced in Eqs. (3.55) and (3.61).
From these compatible splits in stress and strain rate, one obtains alterna-
tive representations for the thermodynamic rate of working. Again, our
development here is in the Lagrangian frame, the preferred frame for the
analysis of solids.
Consider a hydrostatic/deviatoric split of the second PiolaKirchho
stress such that
Sx = S + p C 1 with p = 13 tr(SC ), (4.40)

where p is the hydrostatic pressure and Sx is the deviatoric stress, both


of which have units of force per unit undeformed area. Pressure is as-
signed with a negative value as a matter of convention. It aligns with our
intuition that an increasing gauge pressure is positive when, in actuality, it
Stress 89

designates a compressive state of isotropic stress. It follows from these def-


initions that tr(Sx C ) = 0, i.e., stress Sx is traceless and therefore deviatoric.
Note the similarity between this definition and the definition for extra
stress given in Eq. (4.25). Their dierences lie in their intended applica-
tions. The above hydrostatic/deviatoric split applies for compressible and
nearly incompressible materials, whereas the split that leads to extra stress
only applies for incompressible materials. The Lagrange multiplier } is in-
troduced to ensure that the volume remains constant, whereas hydrostatic
pressure p is the physical cause for a volumetric dilatation . Scalar } is
mathematical in origin, whereas scalar p is physical in origin.
With these definitions in place the rate at which work is being done, as
defined in Eq. (4.30), can be expressed in two alternative forms, the first
being
dW = tr(S dE )
(4.41)
= tr(Sx dEx ) 3p d
and the second being
dW = tr(CSC dE)
(4.42)
x 3p d.
= tr(C Sx C dE)
To derive the second line from the first, in the first expression, substitute
the decomposition of strain rate listed in Eq. (3.55) into the expression for
work giving dW = tr(S dEx ) + tr(SC ) d = tr(S dEx ) 3p d using the
definition for pressure in Eq. (4.40). Replacing S with Eq. (4.40) leads
to tr(S dEx ) = tr(Sx dEx ) p tr(C 1 dEx ) = tr(Sx dEx ) because the deviatoric
strain rate is traceless, i.e., tr(C 1 dEx ) = 0 from Eq. (3.53), which leads
to the stated result. Derivation of the second expression follows a similar
path after an application of the strain-rate identity of Eq. (3.46).
In other words, Eqs. (4.41) and (4.42) imply that the rate of work caused
by stressing can be separated into two independent contributions using ei-
ther of two dierent sets of state variables. One contribution relates to the
work associated with isochoric changes in shape. The other contribution
relates to the work associated with isotropic changes in volume. One set
of state variables adopts the Green strain and the second PiolaKirchho
stress. The other set of state variables adopts the Lodge strain and the
covariant form of the second PiolaKirchho stress. An important point
here is that for either choice of state variables, the dilatroic and deviatoric
rates of working are independent of each other, i.e., they are completely
90 Soft Solids

separable. This is a consequence of physics; it is independent of material


constitution.

4.9 Examples

The theories developed in this text, for the most part, have stress as the
dependent variable and strain as the independent variable. To be able to
compare theory with experiment requires that one be able to extract the
individual components of stress from the measured quantities in an exper-
iment: forces, moments, stretches, and/or rotations. Acquiring such rela-
tionships for the seven BVPs addressed in this book is the remaining topic
of this chapter.
In the prior chapters, you studied how to represent deformation and
were, therefore, consumed by how a body changes its shape. In this chap-
ter, the physical cause of such changes is investigated. Fortunately, ac-
counting for shape change in ones assessment of stress and its components
is not always necessary. Applying the definition for nominal traction, as it
pertains to the first PiolaKirchho stress established in Eqs. (4.10) and
(4.15), allows for an analysis of stress to take place in the reference config-
uration 0 where the boundary conditions and specimen dimensions are
presumed known and are in forms that are usually easier to work with.
After the first PiolaKirchho stress P has been quantified through exper-
imentally measured variables, P can be mapped into the Kirchho stress s
via Eq. (4.17) for use in Eulerian constructions, or into the second Piola
Kirchho stress S via Eq. (4.18) for use in Lagrangian constructions.

4.9.1 Uniaxial Extension


From this, the stalwart of all experiments, experimentalists have obtained
the vast majority of our knowledge about material behavior. Here, an
isotropic specimen is pulled or pushed along its longitudinal axis, as shown
in Fig. 4.2.3 The nominal traction vector then has components

3 The force of reaction is understood to exist and to be present, but it is not


drawn so as to keep the schematic simple and uncluttered. This is true of most
schematics drawn for this chapter. This force is carried through a clamped bound-
ary condition, which is drawn as a hatched surface.
Stress 91

e2
A(t0)

e1 T (t)
e3

Fig. 4.2 The deformation of axial extension illustrated in Fig. 1.3 is caused by
a distributed force f acting over an initial area A(t0 ) whose normal N lies in the
1-direction in the reference conguration 0 . From Kirchhos expression (4.10)
of Cauchys postulate (4.1) a nominal traction vector T (t) = f (t)/A(t0 ) acts at the
centroid of area A(t0 ) and aligns with the 1-direction. This traction is opposed by
a clamped boundary condition acting along the shaded surface

 
T(t) f (t)/A(t )
df 0
{T }e1 = H) 0 = 0 , (4.43)
dA e1 0 0

where uniformity of the distributed force f over area A in 0 allows


df (t)/dA to be expressed in terms of a nominal traction vector T (t) =
f (t)/A(t0 ) located at the centroid of area A(t0 ). Because the specimen is
traction free in the two principal directions normal to e1 , it follows that
{T }e2 = {0} and {T }e3 = {0}. (4.44)
These are implications of Kirchhos (1852) representation (4.10) of
Cauchys (1827) postulate (4.1). The first PiolaKirchho stress, there-
fore, has a construction of
T

2
T00
3
1

{T }e1 = [P ]{N }e1 H) 0 = 4 0 0 05 0 , (4.45)
0 0 00 0
2
T00
0
3
0

{T }e2 = [P ]{N }e2 H) 0 = 4 0 0 05 1 , (4.46)
0 0 00 0
and
2
T00
0
3
0

{T }e3 = [P ]{N }e3 H) 0 = 4 0 0 05 0 , (4.47)
0 0 00 1
where traction is being applied in only one direction, viz., N |e1 .
With the first PiolaKirchho stress now known, an application of the
deformation gradient F quantified in Eq. (2.44) allows Eq. (4.17) to be
92 Soft Solids

e3

T2 (t)
e2

e1
A2 (t0)

A1 (t0)
T1 (t)

Fig. 4.3 The deformation of an equi-biaxial extension illustrated in Fig. 1.4 is


caused by a pair of distributed forces f1 and f2 acting over initial areas A1 (t0 )
and A2 (t0 ) whose normals N |e1 and N |e2 align with the 1- and 2-directions, re-
spectively, when drawn in the reference conguration 0 . From Kirchhos ex-
pression (4.10) of Cauchys postulate (4.1), orthogonal nominal traction vectors
T1 (t) = f1 (t)/A1 (t0 ) and T2 (t) = f2 (t)/A2 (t0 ) act at the centroids of their respec-
tive areas A1 (t0 ) and A2 (t0 ). These tractions are opposed by a clamped boundary
condition acting along the shaded surfaces

solved for the Kirchho stress, which in turn allows Eq. (4.18) to be solved
for the second PiolaKirchho stress, producing components
2 3 2 3
T 0 0 T/ 0 0
[s] = 4 0 0 05 and [S ] = 4 0 0 05 (4.48)
0 00 0 00
whose rates, from Eq. (4.21), are
2 3 2 3
P + T P 0 0
T P T P 0 0
T


M
sP = 4 0 0 05 so s =4 0 0 05 (4.49)
0 00 0 00
and
2 3
P 200
P T /
T/


SP = 4 0 0 05 (4.50)
0 00
all of which are distinct from one another.

4.9.2 Equi-biaxial Extension


In this class of experiments, an isotropic specimen is simultaneously pulled
in two orthogonal directions in equal measure, as shown in Fig. 4.3. In
Stress 93

such loadings two nominal traction vectors are created simultaneously with
components
 
T (t) f (t)/A (t )
df 1 1 1 0
{T }e1 = H) 0 = 0 (4.51)
dA e1 0 0
and
  0

0

df
{T }e2 = H) T2 (t) = f2 (t)/A2 (t0 ) (4.52)
dA e2 0 0
with
{T }e3 = {0}, (4.53)
where f1 (t) and f2 (t) are two uniformly distributed forces imposed on
the specimen, which act on surfaces whose initial areas are A1 (t0 ) and
A2 (t0 ). Ideally, forces f1 and f2 are adjusted so that both tractions
T1 (t) = f1 (t)/A1 (t0 ) and T2 (t) = f2 (t)/A2 (t0 ) equal T(t), i.e., the stressed
loading is equi-biaxial with T1 = T2 = T at every moment t. The first
PiolaKirchho stress must resolve both of these external tractions and, as
such, obeys

T1
2
T00
3
1

{T }e1 = [P ]{N }e1 H) 0 = 4 0 T 05 0 (4.54)
0 0 00 0
and
2 0
T00
3
0

{T }e2 = [P ]{N }e2 H) T2 = 4 0 T 05 1 (4.55)
0 0 00 0
with
T00
0
3
0
2 
4
{T }e3 = [P ]{N }e3 H) 0 = 0 T 0 5 0 , (4.56)
0 0 00 1
where tractions are now being applied in two directions, viz., N |e1 and
N |e2 .
With the first PiolaKirchho stress now known, assuming an isotropic
material, an application of the deformation gradient F quantified in
Eq. (2.48) allows Eq. (4.17) to be solved for the Kirchho stress which, in
94 Soft Solids

f2R
f2L
T2 t2
T1 t1
h
f1
h h
w f1 w
T1 t1
T2 e2
t2 f2L

0 f2R
e1

Fig. 4.4 The deformation of simple shear illustrated in Fig. 2.3, caused by an
applied force f1 , is imposed on a specimen of dimensions w (width) by h (height)
by d (depth, not shown) thereby producing a traction of t1 = f1 /wd. This sets up
reaction forces f2R and f2L producing a traction of t2 = (f2R + f2L )/wd. Tractions
t1 and t2 of the spatial conguration  associate with pseudo tractions T1 and
T2 in material conguration 0

turn, allows Eq. (4.18) to be solved for the second PiolaKirchho stress,
thereby producing components
2 3 2 3
T 0 0 T/ 0 0
[s] = 4 0 T 05 and [S ] = 4 0 T/ 05 (4.57)
0 0 0 0 0 0

whose rates, from Eq. (4.21), are


2 3 2 3
T P + T P P T P
T

0 0
M 0 0
sP = 4 0 TP + T P 05 , s =4 0 P T P 05
T (4.58)
0 0 0 0 0 0

and
2 3
T/ P 2
P T / 0 0


P
S = 4 0 P P 2
T/ T / 05 . (4.59)
0 0 0

Here the assumption is that both the loading and the deformation are equi-
biaxial. A tacit requirement is that the material must be isotropic to be able
to fulfill this assumption.
Stress 95

4.9.3 Simple Shear


Simple shear is not simple! It is actually a dicult experiment to execute.
From symmetry of this BVP with respect to coordinate axes (e1 , e2 , e3 )
and (e1 , e2 , e3 ), cf. Fig. 4.4, Lodge (1964, pp. 6265) has shown that the
most general expression for a state of stress caused by the simple shear of
an isotropic material is given by
2 3
s11 s12 0
[s] = 4s21 s22 0 5 , (4.60)
0 0 s33

where s12 = s21 from the symmetry of s. The Kirchho stress and first
PiolaKirchho stress relate to each other according to Eq. (4.17), from
which it follows that the first PiolaKirchho stress has components of
2 3 2 3
P11 P12 0 s11  s12 s12 0
[P ] = 4P21 P22 0 5 = 4s21  s22 s22 0 5 , (4.61)
0 0 P33 0 0 s33

where P12 P21 , in general.


The rectilinear simple shear of an isotropic material is illustrated in
Fig. 4.4. In the deformed configuration  of the right graphic, tractions
t1 = f1 /wd and t2 = (f2R + f2L )/wd are directed through the centroids
of the upper- and lowermost shear planes denoted by x2 = h and x2 =
0. Summing moments about the lower left corner leads to the constraint
f1 = wh (f2R f2L ) +  (f2R + f2L ) where forces f1 , f2R , and f2L are the
actual forces applied to the material sample by the experimental apparatus.
Introducing two reaction forces, viz., f2R and f2L , that are pinned at the
sample corners ensures that the moment hf1 caused by shearing the sample
can be canceled out, thereby maintaining equilibrium in this test fixture.
The simple shear of an isotropic planar material drawn in Fig. 4.4
has a nominal traction vector in the 2-direction whose components are
described by
 
T (t) f1 (t)/wd

df 1
{T }e2 = H) T2 (t) = (f2R (t) + f2L (t))/wd , (4.62)
dA e2 0 0

where T1 (t) and T2 (t) are the two components of traction T (t) imposed on
a surface whose initial area is A(t0 ) = wd; they are the shear traction T1
96 Soft Solids

and the normal traction T2 . The normal N to this surface aligns with the

T1

2-direction. From Eqs. (4.15) and (4.61) it readily follows that
2
s11  s12 s12 0
3
0 s12
 
4
{T }e2 = T2 = s21  s22 s22 0 5 1 = s22 (4.63)
0 0 0 s33 0 0
and, therefore, s12 = T1 and s22 = T2 .
The gripping surfaces along the shear planes of X2 = 0 and X2 = h carry
internal tractions of T1in and T2in along material surfaces whose normals
align with the 1-direction in that
 
T 
df 1in (t)
{T }e1 = H) T2in (t) , (4.64)
dA e1 0
so that
T  2
s11  s12 s12 0
3
1

1in
{T }e1 = T2in = 4s21  s22 s22 0 5 0
0 0
s 0 s33 0
 s12
 unknown

11
= s21  s22 = (f2R f2L )/hd , (4.65)
0 0
where the internally carried traction T2in = (f2R f2L )/hd follows from the
previously determined stress components of s12 = f1 /wd and s22 = (f2R +
f2L )/wd along with the expression guaranteeing equilibrium of moments,
viz., f1  (f2R + f2L ) = wh (f2R f2L ). In contrast, the internal traction T1in
cannot be resolved from these boundary conditions.
Finally, planes whose normals align with the 3-direction are considered
to be traction freethe plane-stress assumptionleading to
 
0
df
{T }e3 = H) 0 , (4.66)
dA e3 0
so that
0 2
s11  s12 s12 0
3 
0
0
{T }e3 = 0 = 4s21  s22 s22 0 5 0 = 0 (4.67)
0 0 0 s33 1 s33
and, therefore, s33 = 0.
Stress 97

With the Kirchho stress now known, assuming an isotropic material,


an application of the deformation gradient F quantified in Eq. (2.52) allows
Eq. (4.18) to be solved for the second PiolaKirchho stress, thereby
producing components
" # 2 3
s11 T1 0 s11 2 T1 +  2 T2 T1  T2 0
[s] = T1 T2 0 & [S ] = 4 T1  T2 T2 05 (4.68)
0 0 0 0 0 0
with the rate of Kirchho stress having components of
2 3 2 3
P
s P1 0
T P 1 T2 P
sP11 2T1 P T 0

11
M
4 P
sP = T1 T2 0 P 5 4
so s = T1 T2 P
P P2
T 05 . (4.69)
0 0 0 0 0 0
From here, the time rate of change of the second PiolaKirchho stress SP
can be determined via Eqs. (2.52) and (4.21) or by direct dierentiation of
Eq. (4.68).
Component s11 cannot be extracted from experimental data for rectilin-
ear simple shear.4 Multiplication commutes between the Kirchho stress
s and the Finger deformation b for all explicit elastic solids (cf. Chap. 5),
i.e., [s] [b]  [b] [s], from which it follows that P11 = P22 during simple
shear and, therefore, s11 = s22 +  s12 implying that T1in = s22 . However,
this result only applies for constitutive equations where the Kirchho stress
and Finger deformation commute. This is not a general result.

4.9.4 Homogeneous Planar Membranes


Simple shear is not an easy BVP to realize in an experimental apparatus,
yet, on paper, it introduces states of shear into a material sample that would
be desirable from the perspective of material characterization. A BVP that
can be more readily realized in an experimental apparatus, which also com-
bines states of shear and tension, is presented here that, to the best of the
authors knowledge, has not been reported on in the literature.
Like simple shear, the nonzero components of Kirchho stress are de-
scribed by Eq. (4.60) for membranes, too. In fact, what one means by a
membrane is any structure that cannot support an out-of-plane shear stress,
4 To capture s11 in uid experiments, shear ows are set up along planes with
curvature. Analysis of these BVPs exceeds the scope of this book. The interested
reader is referred to the texts by Bird et al. (1987a), Ferry (1980), and Lodge
(1974).
98 Soft Solids

viz., s13 = s31 = s23 = s32 = 0, which is the case in Eq. (4.60). If the
plane-stress assumption applies, then one also has s33 = 0.5
From the deformation gradient that describes an isotropic,
homogeneous, planar membrane, viz., Eq. (2.61), and from the relation-
ship between the Kirchho and first PiolaKirchho stresses given in
Eq. (4.17), one determines its components as being
2 3
2 (s11 1 s12 ) 1 (s12 2 s11 ) 0
1
[P ] = 4 2 (s21 1 s22 ) 1 (s22 2 s21 ) 0 5 , (4.70)

0 0
s33
2

where 1 and 2 are the stretches, 1 and 2 are the in-plane shears, and

= 1 2 (1 1 2 ) is the areal stretch.


The homogeneous stretching of an isotropic planar membrane is illus-
trated in Fig. 4.5. In the deformed configuration of , tractions t1 and t2
are applied, separated by an angle of /2 1 2 , with their lines of action
passing through the centroids of their areas of traction, and through the
centroid of the volume. Consequently, no moments exist, which is a sig-
nificant simplification over the boundary conditions that pertain to simple
shear. When tractions t1 and t2 are placed into the initial configuration 0
according to Kirchhos postulate, Eq. (4.15) and Fig. 4.1, they become
the pseudo tractions T1 and T2 displayed in the left graphic of Fig. 4.5.
These tractions are imposed at angles of incidence of 2 and 1 from their
respective normals in 0 . In contrast with equi-biaxial extensions, the two
tractions of this BVP are not required to be orthogonal to one another, nor
need they be of equal magnitude.
In this experiment, the material is allowed to both stretch and rotate
when subjected to a pair of planar forces. Consequently, like equi-biaxial
extension, two nominal traction vectors are created, but here they have
components of
 

T1 cos 2

f1 cos(2 )/hd

df
{T }e1 = H) T1 sin 2 = f1 sin(2 )/hd (4.71)
dA e1 0 0

5 In the presence of curvature, e.g., for a balloon, a pressure s


33 acting normal
to the surface of a membrane can exist. In such cases, this normal pressure is
balanced by the four in-plane stress components sij , i, j = 1, 2 that are carried along
the curved membrane surface. A study of curvature lies beyond the intended scope
of this introductory text.
Stress 99

t2

T2
1
t1
1
2 T1
h
2
2 t1
T1 w

1
T2 t2
e2

0
e1

Fig. 4.5 An isotropic, incompressible, planar membrane with initial dimensions


of w (width) by h (height) by d (depth, not shown) is deformed according to
Fig. 1.6 through the application of tractions t1 and t2 in the current conguration
 from which are resolved pseudo tractions T1 and T2 residing in the initial
conguration 0 . Tractions t1 and t2 pass through the centroids of their respective
areas, in accordance with Cauchys postulate, and through the centroid of the
volume, so as to conserve angular momentum. Pseudo tractions T1 and T2 are
coaxial with the applied tractions t1 and t2 , respectively, because of Kirchhos
version of Cauchys postulate, viz., Eqs. (4.10) and (4.11)

T  f 
and
  2 sin 1 2 sin(1 )/wd
df
{T }e2 = H) T2 cos 1 = f2 cos(1 )/wd , (4.72)
dA e2 0 0
while normal to the surface of the membrane
  0

df
{T }e3 = H) 0 , (4.73)
dA e3 0
where f1 (t) and f2 (t) are the two distributed forces imposed on the speci-
men causing tractions t1 and t2 , which act on surfaces whose initial areas
are A1 (t0 ) = hd and A2 (t0 ) = wd, cf. Fig. 4.5.
The first PiolaKirchho stress must resolve these two external trac-
tions, and, as such, from {T }e1 = [P ]{N }e1 , one gets
100 Soft Solids

T  2
2 (s11 1 s12 ) 1 (s12 2 s11 ) 0
3 
1 cos 2 1
1
T1 sin 2 = 4 2 (s21 1 s22 ) 1 (s22 2 s21 ) 0 5 0

0
s 0
1 s12
 0
2 s33 0

2 11
= s21 1 s22 (4.74)

0
and from {T }e2 = [P ]{N }e2 , one gets

T2 sin 1

2
(s 1 s12 ) 1 (s12 2 s11 ) 0
3
0

1 4 2 11
T2 cos 1 = 2 (s21 1 s22 ) 1 (s22 2 s21 ) 0 5 1

0 0
s 2 s11
0

2 s33 0

1 12
= s22 2 s21 (4.75)

0

while from {T }e3 = [P ]{N }e3 , one gets



0
2
(s 1 s12 ) 1 (s12 2 s11 ) 0
3
0

1 4 2 11
0 = 2 (s21 1 s22 ) 1 (s22 2 s21 ) 0 5 0 (4.76)

0 0 0
2 s33 1

and, therefore, s33 = 0.


Equations (4.74)(4.76) produce the system of equations
2
1 0 1 0
3
s11

T1 cos(2 )/ 2

1 6 7
6 0 1 0 1 7 s21 = T1 sin(2 )/ 2 (4.77)

42 0 1 0 5 s12 T2 sin(1 )/ 1
0 2 0 1 s22 T2 cos(1 )/ 1

s T
whose solution is
2 3
11 1 0 1 0 1 1 cos 2
s21 60 1 0 1 7 1 T1 sin 2
=6
42
7 , (4.78)
s12 0 1 05 2 T2 sin 1
s22 0 2 0 1 2 T2 cos 1

where
= 1 2 (1 1 2 ) establishes the extent of isochoric areal stretch
defined in Eq. (1.23).
Stress 101

A constraint equation follows from the symmetry of Kirchho stress,


i.e., s21 = s12 , viz.,
   
2 cos 2 sin 2 1 T1 = 1 cos 1 sin 1 2 T2 . (4.79)
One might, e.g., choose to control a ratio of tractions 1 T1 / 2 T2 = const
separated by an angle /2 1 2 from which 1 and 2 would follow.
Given that a material model is in hand, the tractions predicted by that
model come from solving Eq. (4.77) resulting in
q
2
T1 = (s11 1 s12 )2 + (s21 1 s22 )2 , (4.80)

q
1
T2 = (s12 2 s11 )2 + (s22 2 s21 )2 , (4.81)

which provides a mechanism whereby model predictions can be compared


against experimental data.

4.10 Exercises

4.10.1 Pure Shear


The forces imposed on this experiment are of the same type as those im-
posed in the equi-biaxial extension test shown in Fig. 4.3, except T1 T2 .
In practice, f1 will be able to be measured experimentally, but not f2 ,
as it is carried internally via the clamping constraint of the grips (cf.
simple-shear experiment). From these uniformly distributed forces, de-
termine their associated tractions T1 and T2 . Now, given the deformation
F and velocity l gradients that you derived previously, determine com-
ponents for the first PiolaKirchho stress P , the second PiolaKirchho
M
stress S , and the Kirchho stress s, plus their rates SP , sP , and s. Leave
the traction in the 2-direction T2 as an unknown. T2 can only be predicted
once an appropriate constitutive equation has been determined. This will
become apparent in the remaining chapters.

4.10.2 Biaxial Extension


The forces imposed on this experiment are of the same as those imposed
in the equi-biaxial extension test shown in Fig. 4.3, except T1 T2 .
Given the deformation F and velocity l gradients that you derived previ-
ously, derive components for the first PiolaKirchho stress P , the second
102 Soft Solids

PiolaKirchho stress S , and the Kirchho stress s, plus their rates SP , sP ,


M
and s. Show that your results reduce to those of equi-biaxial extension
whenever kT1 k = kT2 k.

4.10.3 Extension Followed by Simple Shear


The forces imposed on this experiment are of the same type as those im-
posed in the simple-shear experiment shown in Fig. 4.4. This experiment
prestretches the sample before shearing it. The force caused by this pre-
stretch of in the 2-direction can be thought of as an initial condition to
the ensuing shear deformation. First, quantify the initial state of stress S0
incurred by prestretching. From this prestretched state, using the deforma-
tion F and velocity l gradients that you derived previously, extending the
stress analysis of simple shear, determine components for the first Piola
Kirchho stress P , the second PiolaKirchho stress S , and the Kirchho
M
stress s, plus their rates SP , sP , and s that arise during the ensuing shear
deformation. How do they compare with those of simple shear? What can
you say about s11 ?

4.10.4 Other Problems


1. From the deformation gradient F given in Eq. (2.61), determine the
components for the second PiolaKirchho stress S as they pertain to
a homogeneously deformed membrane.
M
2. Express the components for SP , sP , and s for the homogeneous mem-
brane using the deformation F and velocity l gradients given in
Eqs. (2.61) and (2.69) along with the definition for Oldroyds stress
rate, which appears in Eq. (4.22).
3. Push the isotropic/deviatoric split of the second PiolaKirchho stress
stated in formula (4.40) forward from the Lagrangian frame 0 into the
Eulerian frame .
4. Express the isotropic/isochoric split for the rate of working (4.41) in
terms of Eulerian fields.
5. Prove that tr(CSC dE) = tr(S dE ). If E is the strain measure selected
for use, then CSC will be its thermodynamic conjugate stress.
Stress 103

6. How does the Biot stress Ty relate to the other stresses introduced in this
text? Biot stress Ty is the thermodynamic conjugate to Biot strain Ey
which is defined in Prob. 3.7.4(1). In other words, determine Ty such
that tr(Ty dEy ) = tr(S dE ).
PART 2
Constitutive Equations
In theory,
there is no difference between
theory and practice.
But in practice, there is.
Yogi Berra
Albert Einstein
Chapter 5

Explicit Elasticity

The theory of constitutive equations is steeped in physics and


mathematics and is an important discipline within the topic of the me-
chanics of continuous media. Physical laws have already been introduced.
The conservation of mass was addressed in the discussion on deformation,
which led to a constraint equation for isochoric deformations. Newtons
laws of motion entered into discussion in the previous chapter on stress via
the conservation of momenta, with the conservation of angular momen-
tum requiring a symmetric stress tensor. Constitutive equations have their
association with physical laws through the conservation of energy.
As is often the case in this field of study and, in keeping with the texts
theme of being a primer for mechanics, attention is not paid to energetic
contributions that are chemical, electromagnetic, or thermal in origin, al-
though this is not necessary, but is instead focused on only the mechanical
energies that associate with deformation. These energies appear in a state-
ment of the ClausiusPlanck inequality for continuous media (Holzapfel
2000, pp. 166173)
 
D = tr S dE dU where D  0 (5.1)
with the inequality being a consequence of the second law of thermo-
dynamics. State function D(X , t) accounts for the rate of dissipation oc-
curring at particle P , while state function U(X , t) represents the energy
that is in P s possession. In other words, this expression says that the rate
at which work is being done on  a particle
 caused by external forces acting
upon the body, i.e., dW = tr S dE , minus any accumulation in energy
stored internally at the particle, viz., dU, establishes a loss rate due to dis-
sipation D at P , a rate that cannot be negative because of the second law.

A.D. Freed, Soft Solids: A Primer to the Theoretical Mechanics of Materials, 109
Modeling and Simulation in Science, Engineering and Technology,
DOI 10.1007/978-3-319-03551-2_5, Springer International Publishing Switzerland 2014
110 Soft Solids

For many soft materials, dU accumulations are entropic in origin (e.g.,


rotations of chemical bonds) while, for hard materials, and some soft ma-
terials, too, dU accumulations are energetic in origin (e.g., stretching of
chemical bonds). Such discussions have interesting thermodynamic impli-
cations but they lie beyond the scope of this text. The interested reader is
referred to, e.g., the texts of Holzapfel (2000, Chap. 7), Lodge (1999), and
Treloar (1975).

5.1 Theory

Explicit elasticity, also known as Green (1841) elasticity, hyperelasticity


(Truesdell and Noll 2004), and rubber elasticity (Treloar 1975), is a
finite-strain theory for describing the elastic response of materials. It has
found application with natural and synthetic rubbers, elastomers, and many
plastics. Here stress is described in terms of strain through a potential func-
tion of thermodynamic origin.
Deformation without dissipation is a thermodynamic definition for an
elastic response. Several definitions for elasticity are used throughout the
literature; cf. Rajagopal (2011a). Deformation without dissipation, in the
opinion of many, including this author, is the definition that most closely
aligns with the underlying physics.

5.1.1 Green Elastic Solid


Green strain E = 12 (C I) is a relative measure of the change in distance
separating two neighboring particles, as established in Eq. (3.26). Here we
derive an elastic theory based upon this measure of strain.
Because elastic materials do not dissipate energy, by definition, the in-
ternal energy U in Eq. (5.1) can be interpreted as being a work function,
i.e., U = W (E ), where W is the so-called elastic strain-energy function.
Considering the elastic strain energy at P to be a dierentiable function of
the Green strain E allows the chain rule to be applied to Eq. (5.1), which
in turn leads to the expression
 
  dW (E )
D = tr S dE dW = S W dE = 0 (5.2)
dE
Explicit Elasticity 111

out of which we can extract an elastic constitutive equation describing the


second PiolaKirchho stress, viz.,
 
dW (E ) IJ 1 dW (E ) dW (E )
S = or S = + (5.3)
dE 2 dEIJ dEJI
where assurance of symmetry in stress, as required by the conservation
of angular momentum, is apparent in the component form of this classic
theory. The resulting constitutive relationship resides within the large par-
enthetical brackets found in Eq. (5.2). By setting this bracketed quantity to
zero, the imposed equality will always be satisfied for every possible value
of dE . Equation (5.3) is commonly referred to as the Green elastic solid,
so named because Green strain E is the independent variable.

5.1.1.1 Hydrostatic/Deviatoric Form


Material models that are incompressible, or nearly incompressible, which
typify the majority of soft solids, cause numerical issues when imple-
mented into finite element codes, e.g., element locking. For this reason
it is advantageous to split the material response into separate isotropic and
isochoric responses that in this form lend themselves to variational formu-
lations that can alleviate such issues (cf. Belytschko et al. 2000).
The Green elastic solid can be modified by employing the isotropic/
isochoric split of the work being done dW = 3p d + tr(Sx dEx ) found in
Eq. (4.41). Furthermore, because the state variables  and Ex and their
associated strain energies act independently of one another, it necessarily
follows that the elastic strain energy is a sum of two separable functions,
one for each independent variable, allowing us to write our constitutive
assumption as W (E ) = Wb () + Ws (Ex ) so that
 
D = 3p d dWb () + tr Sx dEx dWs (Ex )
  !
dWb () dW s (Ex) (5.4)
= 3p + d + Sx W dEx = 0
d dEx
where Wb and Ws are the strain-energy contributions governing the elastic
bulk and shear responses, respectively. In order for the two parenthetical
terms to not dissipate energy for any possible set of paired values d and
dEx , which are independent rates, it is sucient that each bracketed term
be zero resulting in two separate constitutive equations, viz.,
1 dWb ()
p= (5.5)
3 d
112 Soft Solids

and
dWs (Ex )  
Sx = so that tr Sx C = 0. (5.6)
dEx
These quantify the hydrostatic and deviatoric contributions to stress in a
Green elastic solid through which the total state of stress can then be estab-
lished from Eq. (4.40), viz., S = p C 1 + Sx . Recall that the distortional
strain Ex need not, in general, be deviatoric, i.e., traceless, whereas the
deviatoric stress Sx must be traceless, by definition, hence the constraint
equation tr(Sx C ) = 0.

5.1.2 Lodge Elastic Solid


Lodge strain E = 12 (I C 1 ) is a relative measure of the change in distance
separating two, neighboring, nonintersecting, material surfaces, as estab-
lished in Eq. (3.30). In the case of incompressible materials, Truesdell
y becomes equivalent to Lodge strain E, defined in Eqs. (3.33) and
strain E
(3.14), respectively, with the former being a relative measure of the change
in area of the associated material surface, as established in Eq. (3.34). Here
we derive an elastic theory based upon such a measure for strain.
Equation (3.46) determined that dE = C  dE  C and, therefore, the
two physical strain rates dE and dE are not independent, even though their
associated strains E and E are unique. Consequently, the rate of working
dW = tr(S dE ) can also be written as dW = tr(S C  dE  C ). From a
property of the trace listed in Eq. (A.70), this expression for the rate of
working can be altered to read as dW = tr(CSC dE) and, as such, the
internal energy becomes U = W (E) with Lodge strain E now being the
thermodynamic variable. This allows Eq. (5.2) to be rewritten as
 
  dW (E)
D = tr CSC dE dW = CSC W dE = 0 (5.7)
dE
out of which an alternative elastic constitutive equation is extracted for the
second PiolaKirchho stress, viz.,
dW (E) 1
S = C 1 C
dE  
IJ IK 1 dW (E) dW (E)
or S = C KL
+ LK
CLJ (5.8)
2 dE dE
where here the strain measure E = 12 (I C 1 ) of Lodge (3.14) replaces the
strain measure E = 12 (C I) of Green (3.11) present in Eq. (5.2) as the
Explicit Elasticity 113

independent variable. Equation (5.8) is the Lodge elastic solid, so named


because Lodge strain E is the independent variable.

5.1.2.1 Hydrostatic/Deviatoric Form


For the same reason that it is advantageous to create Green elastic solids as
being comprised of separate isotropic and isochoric constituents, so too is it
advantageous to create Lodge elastic solids that are comprised of separate
bulk and shear contributions.
An isotropic/isochoric split of the work being done exists for this choice
of state variable, too, in that dW = 3p d + tr(C Sx C dE), x which fol-
lows from inserting the expressions for stress S from Eq. (4.40) and strain
rate dE from Eq. (3.61) into the work expression dW = tr(CSC dE),
cf. Eq. (4.42), where now our constitutive assumption becomes W (E) =
x so that
Wb () + Ws (E)
 
D = 3p d dWb () + tr C Sx C dE x dWs (E) x
  !
dWb () dW s ( x
E)
= 3p + d + C Sx C x =0
W dE (5.9)
d x
dE
out of which come the constitutive expressions
1 dWb ()
p= (5.10)
3 d
and
x
dWs (E)  
Sx = C 1 C 1 so that tr Sx C = 0 (5.11)
x
dE
where S = p C 1 + Sx updates the state of stress. Again, recall that the dis-
x need not, in general, be deviatoric, i.e., traceless, whereas
tortional strain E
the deviatoric stress Sx must be traceless, by definition, hence the constraint
equation tr(Sx C ) = 0.

5.1.3 Properties
The Green and Lodge elastic solids are explicit constitutive equations in
the sense that their strain-energy functions W depend only upon strain;
they do not depend upon both strain and stress. Such models are discussed
in the next chapter.
114 Soft Solids

Green and Lodge elastic solids are isotropic theories because the
tensorial dependence of their strain-energy functions W depends only upon
the metric of deformation C . Recall that E = 12 (C I) and E = 12 (I C 1 ).
Green and Lodge elastic solids are objective in the presence of an arbi-
trary rotation Q whenever W (E ) = W (QE QT ) for the Green solid, and
whenever W (E) = W (QEQT ) for the Lodge solid, given that QQT =
QTQ = I (cf. Appendix B).

5.2 Isotropic Theory

Invariant theory (Rivlin and Smith 1969; Spencer 1972) is called upon
to simplify the process of constructing viable energy functions. Invari-
ant theory is an outgrowth of the CayleyHamilton theorem, Eqs. (A.87)
and (A.88). It allows the tensorial arguments of an isotropic function to be
replaced with a set of scalars, called invariants, that are unique character-
istics of the tensorial arguments belonging to that function. A vector has
one invariant: its magnitude or length. A tensor in R2 has two invariants:
its trace and its determinant. While a tensor in R3 has three invariants:
its trace (the sum of its three 1  1 minor determinants), the sum of its
three 2  2 minor determinants, and its 3  3 determinant. The theory of
determinants plays a vital role in invariant theory.

5.2.1 Isotropic Green Solid


Invariant theory allows the tensorial arguments of an admissible strain-
energy function, e.g., the Green elastic solid of Eq. (5.3) described by
W (E ), to be expressed in terms of an associated set of scalar invariants,
viz., W (I, II, III). These principal invariants from the CayleyHamilton
theorem, viz., I, II, and III from Eq. (A.88), are swapped with their three
basis functions, i.e., the simplest terms present within I, II, and III, viz.,
their moment invariants, which for W (E ) include
 
I1 = tr C 1E ,
 
I2 = tr C 1EC 1E , (5.12)
 1 1 1 
I3 = tr C EC EC E

where the contravariant indices of Cauchys metric C 1 contract with


the covariant indices of Greens strain E in accordance with general
Explicit Elasticity 115

tensor analysisa quality retained by the Lagrangian frame 0 for ex-


ample, I1 = [C1 ]IJ EJI . Because C 1 maps contravariantly and E
maps covariantly, it follows from Eqs. (B.17) and (B.21) that the first
invariant pushes forward into the Eulerian frame as I1 = tr(C 1 E ) =
tr([F][C 1 ][F]T [F 1 ]T [E ][F 1 ]T ) = tr(Ie) = tr(e), i.e., the first invariant
is equivalent to the trace of Almansi strain. Likewise, it can be shown that
I2 = tr(e2 ) and I3 = tr(e3 ).
Because QQT = QTQ = I for any admissible rotation Q, and from a
property of the trace listed in Eq. (A.70), it follows that
   
I1 = tr C 1E = tr QC 1E QT ,
   
I2 = tr C 1EC 1E = tr QC 1EC 1E QT , (5.13)
   
I3 = tr C 1EC 1EC 1E = tr QC 1EC 1EC 1E QT

and, therefore, these three scalar invariants are objective measures of strain
for any given tensorial state of Green strain E .
When substituted into Eq. (5.3), along with an application of the chain
rule, the strain-energy function W (E )  W (I1 , I2 , I3 ) leads to a general
constitutive theory of

S = W,1 C 1 + 2W,2 C 1EC 1 + 3W,3 C 1EC 1EC 1 (5.14)

where the coecients, W,1 = @W (I1 , I2 , I3 )/@I1 , etc., are scalar functions
of the invariants I1 , I2 , and I3 . This theory becomes a material model
whenever a strain-energy function W = W (I1 , I2 , I3 ) is assigned.
Each term in the above formula is contravariant and, therefore, this
expression is admissible in the sense of general tensor analysis. Because it
is contravariant, it pushes forward from 0 into  according to Eq. (B.15),
leading to its Eulerian form

s = W,1 I + 2W,2 e + 3W,3 e2 (5.15)

where s is the Kirchho stress from Eq. (4.4) and e is the Almansi strain
from Eq. (3.10). All sense of covariance and contravariance is lost in the
Eulerian version of this theory, which illustrates why the author derives his
theories in the Lagrangian frame where these qualities provide a framework
that assists in the overall construction of admissible tensor equations.
116 Soft Solids

5.2.1.1 Incompressible Materials


A large number of soft solids have a bulk modulus  that is much stier
(greater than 100 times, typically) than its shear modulus . The pres-
sure response p under these conditions can be eectively absorbed by the
Lagrange multiplier } present in the extra stresses and  defined in
Eq. (4.25). Hence, and  replace the second PiolaKirchho S and
Kirchho s stresses, respectively, in their constitutive formul.
An incompressible Green elastic solid is, therefore, described by either
= 2W,2 C 1EC 1 + 3W,3 C 1EC 1EC 1 (5.16)
or
 = 2W,2 e + 3W,3 e2 , (5.17)
depending upon the frame of reference chosen, where and  are the
Lagrangian and Eulerian extra stresses from Eq. (4.25). The strain-energy
function simplifies to W = W (I2 , I3 ) for this class of materials.

5.2.1.2 Hydrostatic/Deviatoric Form


In this formulation of the Green elastic solid, Green strain E is re-
placed with its dilational  and distortional Ex counterparts according to
Eqs. (3.51), (3.56), and (5.6) so that dilation is governed by Eq. (5.5) while
distortion is quantified through its three invariants
 
IN1 = tr C 1Ex ,
 
IN2 = tr C 1Ex C 1Ex , (5.18)
 1 
IN3 = tr C Ex C 1Ex C 1Ex .
Recall that tr(C 1 dEx ) = 0 by definition; however, tr(C 1 Ex ) 0. Substi-
tuting these dependencies into Eq. (5.6) leads to the constitutive expression
Sx = Ws,1 C 1 + 2Ws,2 C 1Ex C 1 + 3Ws,3 C 1Ex C 1Ex C 1 where, accord-
ing to Eq. (5.6), the constraint
Ws,1 = 2 IN1 Ws,2 IN2 Ws,3
3 (5.19)
ensures that tr(Sx C ) = 0, i.e., that Sx is deviatoric. The outcome is the
general constitutive equation
 
Sx = 2Ws,2 C 1Ex C 1 13 tr C 1Ex C 1
 
+ 3Ws,3 C 1Ex C 1Ex C 1 13 tr C 1Ex C 1Ex C 1 (5.20)
Explicit Elasticity 117

with S = p C 1 + Sx establishing the total state of stress. The coecients


Ws,i = @Ws (IN1 , IN2 , IN3 )/@INi , i = 1, 2, 3, which are scalar functions of the three
invariants IN1 , IN2 , IN3 of Eq. (5.18), are constrained by Eq. (5.19), thereby
restricting this class of admissible isotropic Green elastic solids.

5.2.2 Isotropic Lodge Solid


The Lodge elastic solid (5.8) distinguishes itself from the Green elastic
solid (5.3) through its choice of an independent variable, viz., E instead of
E , to which are attributed the following set of invariants
I1 = tr(EC ),
I2 = tr(EC EC ), (5.21)
I3 = tr(EC EC EC )
where the contravariant indices of Lodges strain E contract with the
covariant indices of Greens metric C in accordance with general ten-
sor analysisa quality retained by the Lagrangian frame 0 but lost to
the Eulerian frame for example, I1 = E IJ CJI . Because E maps
contravariantly and C maps covariantly, it follows from Eqs. (B.17) and
(B.21) that the first invariant pushes forward into the Eulerian frame as
I1 = tr(EC ) = tr([F][E][F]T [F 1 ]T [C ][F 1 ]T ) = tr(eI) = tr(e), i.e., the
first invariant is equivalent to the trace of Signorini strain. Likewise, it can
be shown that I2 = tr(e2 ) and I3 = tr(e3 ).
Because QQT = QTQ = I for any admissible rotation Q, and from a
property of the trace listed in Eq. (A.70), it follows that
   
I1 = tr EC = tr QEC QT ,
   
I2 = tr EC EC = tr QEC EC QT , (5.22)
   
I3 = tr EC EC EC = tr QEC EC EC QT
and, therefore, these three scalar invariants are objective measures of strain
for any given tensorial state of Lodge strain E.
When substituted into Eq. (5.8), along with an application of the chain
rule to the strain-energy function W (E)  W (I1 , I2 , I3 ), this theory arrives
at
S = W,1 C 1 + 2W,2 E + 3W,3 EC E (5.23)
where the coecients, W,1 = @W (I1 , I2 , I3 )/@I1 , etc., are scalar functions
of the invariants I1 , I2 , and I3 , as defined in Eq. (5.21). The theory becomes
a model whenever a strain-energy function W = W (I1 , I2 , I3 ) is assigned.
118 Soft Solids

Each term in the above formula is contravariant and, therefore, this


expression is admissible in the sense of general tensor analysis. Because it
is contravariant, it pushes forward from 0 into  according to Eq. (B.15)
leading to its Eulerian form
s = W,1 I + 2W,2 e + 3W,3 e2 (5.24)
where s is the Kirchho stress from Eq. (4.4) and e is the Signorini strain
from Eq. (3.13). This formula has the same structure as the Eulerian Green
elastic solid of Eq. (5.15) but with a dierent strain measure; the Eulerian
Lodge elastic solid is expressed in terms of Signorini strain e, while the
Eulerian Green elastic solid is expressed in terms of Almansi strain e.

5.2.2.1 Incompressible Materials


The isotropic contributions to stress get absorbed by the Lagrange multi-
plier in the extra stress for this class of materials, so the incompressible,
Lodge, elastic solid is described by either
= 2W,2 E + 3W,3 EC E or  = 2W,2 e + 3W,3 e2 (5.25)
depending upon your choice of reference frame. The strain-energy function
simplifies to W = W (I2 , I3 ) for this class of materials.

5.2.2.2 Hydrostatic/Deviatoric Form


In this formulation of the Lodge elastic solid, Lodge strain E is re-
placed with its dilational  and distortional E x counterparts according to
Eqs. (3.51), (3.62), and (5.11) so that dilation is governed by Eq. (5.10)
while distortion is quantified through its three invariants
 
x ,
IN1 = tr EC
 
x EC
IN2 = tr EC x , (5.26)
 
x EC
IN3 = tr EC x ECx .

Substituting these dependencies into Eq. (5.11) leads to the constitutive


x + 3Ws,3 EC
expression Sx = Ws,1 C 1 + 2Ws,2 E x E x where the constraint

Ws,1 = 23 IN1 Ws,2 IN2 Ws,3 (5.27)

ensures that tr(Sx C ) = 0 according to Eq. (5.11), i.e., that Sx is deviatoric.


The outcome is the constitutive equation
Explicit Elasticity 119

  1
Sx = 2Ws,2 Ex 1 tr EC
x C
3
  1
+ 3Ws,3 EC x 1 tr EC
x E x EC
x C (5.28)
3

with S = p C 1 + Sx establishing the total state of stress. The coecients


Ws,i = @Ws (IN1 , IN2 , IN3 )/@INi , i = 1, 2, 3, which are scalar functions of the three
invariants IN1 , IN2 , IN3 of Eq. (5.26), are constrained by Eq. (5.27), thereby
restricting this class of admissible, isotropic, Lodge, elastic solids.

5.2.3 A Property of Explicit Elastic Solids


The stress and deformation tensors commute in an explicit elastic solid,
e.g., [S ][C ] = [C ][S ] and [s][b] = [b][s] or, equivalently, [ ][C ] =
[C ][ ] and [][b] = [b][]. This property is valid only in a mathe-
matical sense, which is why it is written as a matrix equation, because
right-covariant (e.g., C ) and left-covariant (e.g., C ) mixed tensor
fields belong to disjoint vector (i.e., Hilbert) spaces spanning 0 (Lodge
1974). In this text, all mixed tensor fields are considered to be right co-
variant. Moon and Truesdell (1974) exploited this property as an aid for
solving BVPs for hyperelastic solids. Commuting matrices are an uncom-
mon occurrence in continuum mechanics.

5.3 A Collection of Material Models

Stable material models are likely to be odd functions of strain, at least


that is the premise supposed in the development of a catalog of material
models constructed for your consideration. This implies that the elastic
strain-energy function will be an even function in strain. Included are elas-
tic models that are of first, third, and fifth order in strain. Certainly this
is not exhaustive, but it should be illustrative. The discussion begins with
incompressible material models, with a discussion of their compressible
counterparts following thereafter. Both Green and Lodge material mod-
els are put forward. Choice of a particular material model, be it Green or
Lodge, and be it first, third, fifth, or some other order will depend upon the
experimental data that you intend the selected model should describe.
120 Soft Solids

5.3.1 Incompressible Green Materials


Here material models are constructed from Eqs. (5.16) and (5.17) using
the invariants from Eq. (5.12). Recall that W = W (I2 , I3 ) for this class
of materials, because the first invariant I1 gets absorbed by the Lagrange
multiplier } in the incompressible case.

5.3.1.1 First-Order Model


Gradients of the strain-energy function for a first-order Green elastic solid
are W,2 = 1 and W,3 = 0, i.e., W = 1 I2 , so that

= 2 1 C 1EC 1 or  = 2 1 e (5.29)

where 1 is the first-order shear modulus.

5.3.1.2 Third-Order Model


Gradients of the strain-energy function for a third-order Green elastic solid
are W,2 = 1 + 3 I2 and W,3 = 0, i.e., W = 1 I2 + 12 3 I22 , so that

= 2( 1 + 3 I2 ) C 1EC 1 ,
(5.30)
 = 2( 1 + 3 I2 ) e

where 1 and 3 are the first- and third-order shear moduli.

5.3.1.3 Fifth-Order Model


Gradients of the strain-energy function for a fifth-order Green elastic solid
are W,2 = 1 + 3 I2 + 5,2 I22 and W,3 = 5,3 I3 , i.e., W = 1 I2 + 12 3 I22 +
1 I 3 + 1 I 2 , so that
3 5,2 2 2 5,3 3
 
= 2 1 + 3 I2 + 5,2 I22 C 1EC 1
+ 3 5,3 I3 C 1EC 1EC 1 , (5.31)
 
 = 2 1 + 3 I2 + 5,2 I22 e + 3 5,3 I3 e2

where 5,2 and 5,3 are the fifth-order shear moduli associated with the
second and third invariants, respectively. Theory cannot discern between
5,2 and 5,3 , only experiments can do so.
Explicit Elasticity 121

5.3.2 Incompressible Lodge Materials


Here material models are constructed from Eq. (5.25) using the invariants
from Eq. (5.21). Again, recall that W = W (I2 , I3 ) for this class of materials,
because the first invariant I1 gets absorbed by the Lagrange multiplier } in
the incompressible case.

5.3.2.1 First-Order Model


Gradients of the strain-energy function for a first-order Lodge elastic solid
are W,2 = 1 and W,3 = 0, i.e., W = 1 I2 , so that
= 2 1 E or  = 2 1 e (5.32)
where 1 is the first-order shear modulus.

5.3.2.2 Third-Order Model


Gradients of the strain-energy function for a third-order Lodge elastic solid
are W,2 = 1 + 3 I2 and W,3 = 0, i.e., W = 1 I2 + 12 3 I22 , so that
= 2( 1 + 3 I2 ) E or  = 2( 1 + 3 I2 ) e (5.33)
where 1 and 3 are the first- and third-order shear moduli.

5.3.2.3 Fifth-Order Model


Gradients of the strain-energy function for a fifth-order Lodge elastic solid
are W,2 = 1 + 3 I2 + 5,2 I22 and W,3 = 5,3 I3 , i.e., W = 1 I2 + 12 3 I22 +
1 I 3 + 1 I 2 , so that
3 5,2 2 2 5,3 3

= 2( 1 + 3 I2 + 5,2 I22 ) E + 3 5,3 I3 EC E,


(5.34)
 = 2( 1 + 3 I2 + 5,2 I22 ) e + 3 5,3 I3 e2
where 5,2 and 5,3 are the fifth-order shear moduli associated with the
second and third invariants, respectively.

5.3.3 Material Stability


From the Hill (1968) constitutive inequality (4.39), the first-order Green
and Lodge models will be stable if
1 > 0 (5.35)
122 Soft Solids

M O
which follows from tr( d ) = 2 1 tr( e d ) = 2 1 tr(d 2 ) > 0 and from
M M
tr( d ) = 2 1 tr(e d ) = 2 1 tr(d 2 ) > 0 where Eqs. (3.37) and (3.40) have
been used. From this logic and the fact that I2  0, the third-order Green
and Lodge models will be stable if
1 > 0 and 3 > 0. (5.36)
In contrast, the fifth-order Green and Lodge elastic solids have the potential
to become unstable whenever tr(d 2 e) 0 for the Green solid or whenever
tr(d 2 e) 0 for the Lodge solid. Even so, their magnitudes will not likely
exceed those of the remaining positive valued terms so as to violate their
respective overall inequalities.

5.3.4 Incompressible Material Models of Renown


5.3.4.1 Neo-Hookean Solid
The simplest rubber-elastic solid in use today is the incompress-
ible, first-order, Lodge solid of Eq. (5.32), commonly known as the
neo-Hookean elastic solid. This material model is usually expressed as
T = } I + b (Holzapfel 2000), which is an equivalent representation
to Eq. (5.32) because } is a Lagrange multiplier, not a material parameter,
and because no distinction exists between the stresses of Cauchy T and
Kirchho s when addressing incompressible materials, i.e.,  = s + } I 
T + } I.
An interpretation for the shear modulus 1 in the neo-Hookean solid
(5.32) has been derived from a molecular network theory for cross-linked
elastomers whose polymer chain mechanics are described by a random
walker under the control of a Gaussian probability density function ac-
cording to the principles of statistical mechanics (James and Guth 1943,
1944, 1947; Lodge 1999; Treloar 1975). The associated energy function
is expressed in terms of an ensemble averaging of the distances between
cross-links in the molecular network, viz., between neighboring material
points.
A derived result from their analysis is that 1 = %RT/Mc [cf. Treloar
(1975, p. 65)] where % is the mass density of the rubber, R is the universal
gas constant (8.3145 J/K mol), T is the absolute temperature (in Kelvin),
and Mc is the (number averaged) molecular weight of the elastomer. Unlike
crystalline solids, elastomers (above their glass transition temperature Tg )
actually get stier with increasing temperaturea fact first observed by
Explicit Elasticity 123

Joule (cf. James and Guth 1944; Treloar 1975). This is because rubbery
deformations tend to be entropic, whereas glassy deformations tend to be
energetic, in a thermodynamic sense.

5.3.4.2 MooneyRivlin Solid


The neo-Hookean solid is a remarkable theory, especially since it has been
derived from statistical mechanics. Yet, from time to time, it falls short of
adequately describing the experimental response displayed by soft solids,
so improvements have been sought over the years.
As a possible remedy, Mooney (1940) proposed the constitutive equa-
tion  = C1 b C2 b1 where C1 and C2 are his material parameters with
b and b1 being the respective deformation tensors of Finger and Piola.
This model is referred to today as the MooneyRivlin elastic solid after the
works of Mooney (1940) and Rivlin and Saunders (1951).
The MooneyRivlin material model can be constructed as a mixture of
the two first-order models of Lodge and Green. In the Lagrangian frame
0 , it can be expressed as
 
= 2 1 (1 ) E +  C 1EC 1 (5.37)
while in the Eulerian frame , it can be written as
 
 = 2 1 (1 ) e +  e (5.38)
where  2 [0, 1] proportions between the Lodge and Green solids. En-
hancing other models through mixture theory exceeds the scope of this
work.
Wang and Guth (1952) derived a theory of rubber elasticity from statis-
tical mechanics that incorporates a non-Gaussian chain mechanics.1 Their
model produces a dependence upon b1 ; however, they found C1 and C2
to be not entirely independent of one another.2
What is expressed here as 1 (1 ) is usually referred to as either C1
or C10 in the literature, and what is expressed here as 1  is referred to
as either C2 or C01 . The former terminology is due to Mooney (1940).
The latter terminology is due to Rivlin and Saunders (1951). These param-
eters are often provided by manufacturers of elastomeric and polymeric
materials to firms that want to use their products in applications, as the
1 The models of James and Guth (1943, 1944) implement Gaussian chain me-
chanics in their derivation of the neo-Hookean solid.
2 If someone has succeeded in deriving the MooneyRivlin model from a statis-
tical mechanics analysis, this author is not aware of their work.
124 Soft Solids

MooneyRivlin model has become the bread-and-butter material model for


soft elastic solids in that it is built into all commercial large-strain FE codes
(known to the author). This allows their clients to perform FE analyses on
their designs using model parameters that the material manufacturer has
confidence in. Elaborate materials, like those used in todays tire indus-
try, require more sophisticated material models, models that lie beyond the
intended scope of this text.
The contribution of 1 I that arises from strains e and e in Eq. (5.38)
is tacitly absorbed by the Lagrange multiplier }I within  in Mooneys
formulation. This does not change the physics of our constitutive equation,
i.e.,  = 2 1 ((1 ) e +  e) vs.  = C1 b C2 b1 , only our interpreta-
tion of it, which seems to be more easily grasped in Eqs. (5.37) and (5.38)
where stress is written as a function of strain rather than in Mooneys clas-
sic form where stress is expressed as a function of deformation. In fact,
Eqs. (5.37) and (5.38) are written in terms of the two dual strain measures
that associate with Riemannian geometry, as they apply to the Lagrangian
and Eulerian frames, respectively.
An application of the Hill (1968) constitutive inequality (4.39) to the
MooneyRivlin solid (5.38) imposes a constraint of
1 > 0 for stable behavior (5.39)
M O M
which follows from tr( d ) = 2 1  tr( e d ) + 2 1 (1 ) tr(e d ) =
2 1 tr(d 2 ) > 0 where Eqs. (3.37) and (3.40) have been used. No constraint
on  arises from Hills stability condition.

5.3.5 Compressible Green Materials


Here a collection of material models are constructed from Eqs. (5.5),
(5.19), and (5.20) using their associated invariants listed in Eq. (5.18).

5.3.5.1 Hydrostatic Pressure


The dilatoric response of a typical Green elastic solid is described by a bulk
strain-energy function of the form Wb = 92 2 such that when substituted
into Eq. (5.5) one gets
 
p = 3 =  ln(dv/dV) = 12  ln det C 1 , (5.40)

where dilatation  = 13 ln det F is defined in Eq. (3.51). This result is


in accordance with the simpler of Henckys (1928, 1931) two constitutive
Explicit Elasticity 125

equations for describing the bulk response. He used these models to ex-
plain the experimental data of Bridgman (1923).
In the presence of large dilatations, Hencky (1931) found it necessary to
introduce a limiting state of dilation in order to be able to describe Bridg-
mans data. Such limiting states would be better handled through the im-
plicit theory of elasticity presented in the next chapter. For our purposes
here, higher-order models for pressure are not usually required because di-
latations tend to be of infinitesimal extent in soft-solid applications, foams
and lung parenchyma withstanding (e.g., cf. Freed and Einstein 2013).
For the deviatoric contributions to stress, we construct first-, third-, and
fifth-order models replicating the approach followed for their incompress-
ible counterparts.

5.3.5.2 First-Order Model


A first-order Green elastic solid has an elastic strain-energy function for
shear of Ws = 1 (IN2 13 IN12 ), which satisfies the constraint equation (5.19),
so that
 
Sx = 2 1 C 1Ex C 1 13 tr C 1 Ex C 1 (5.41)
where 1 is the first-order shear modulus.

5.3.5.3 Third-Order Model


A third-order Green elastic solid has an elastic strain-energy function for
shear of Ws = 1 (IN2 13 IN12 ) + 12 3 (IN22 23 IN12 IN2 + 19 IN14 ), which satisfies the
constraint equation (5.19), so that
  1  
Sx = 2 1 + 3 IN2 1 IN2 3 1 C Ex C 1 1 tr C 1 Ex C 1
3 (5.42)
where 3 is the third-order shear modulus.
If the strain energy is to be a polynomial in its invariants, then no
third-order contribution can arise from W,3 that is also compatible with
constraint equation (5.19). If one were to consider an integral strain-
energy expression, one could then advance such a constitutive expression
that would be compatible with Eq. (5.20).

5.3.5.4 Fifth-Order Model


A fifth-order Green elastic solid has an elastic strain-energy function
for shear of Ws = 1 (IN2 13 IN12 ) + 12 3 (IN22 23 IN12 IN2 + 19 IN14 ) + 13 5
126 Soft Solids

(IN23 IN12 IN22 + 13 IN14 IN2 27


1 IN 6 ), which satisfies the constraint equation (5.19),
1
so that
   
Sx = 2 1 + 3 IN2 13 IN12 + 5 IN22 23 IN12 IN2 + 19 IN14
 
 C 1Ex C 1 13 tr C 1 Ex C 1 (5.43)
where 5 is the fifth-order shear modulus. Like the third-order model,
Ws,3 = 0 whenever the strain-energy function is restricted to be a polyno-
mial because of the constraint equation (5.19).

5.3.6 Compressible Lodge Materials


Here material models are constructed from Eqs. (5.10), (5.27), and (5.28)
using the invariants from Eq. (5.26).

5.3.6.1 Hydrostatic Pressure


The dilatoric response of a typical Lodge elastic solid is governed by
the strain-energy function Wb = 92 2 such that when substituted into
Eq. (5.10) one gets
 
p = 3 =  ln(dv/dV) = 12  ln det C 1 , (5.44)
which is the same as for the compressible Green solid.

5.3.6.2 First-Order Model


A first-order Lodge elastic solid has an elastic strain-energy function for
shear of Ws = 1 (IN2 13 IN12 ), which satisfies the constraint equation (5.27),
so that
  1
Sx = 2 1 E x
x 1 tr EC C (5.45)
3
where 1 is the first-order shear modulus.

5.3.6.3 Third-Order Model


A third-order Lodge elastic solid has an elastic strain-energy function for
shear of Ws = 1 (IN2 13 IN12 ) + 12 3 (IN22 23 IN12 IN2 + 19 IN14 ), which satisfies the
constraint equation (5.27), so that
    1
Sx = 2 1 + 3 IN2 13 IN12 Ex 1 tr EC
3
x C (5.46)
Explicit Elasticity 127

where 3 is the third-order shear modulus. Like the third-order Green


model, Ws,3 = 0 whenever the strain-energy function is restricted to be a
polynomial because of its constraint equation (5.27).

5.3.6.4 Fifth-Order Model


A fifth-order Lodge elastic solid has an elastic strain-energy function for
shear of Ws = 1 (IN2 13 IN12 ) + 12 3 (IN22 23 IN12 IN2 + 19 IN14 ) + 13 5 (IN23 IN12 IN22 +
1 IN 4 IN 1 IN 6 ), which satisfies the constraint equation (5.27), so that
3 1 2 27 1
   
Sx = 2 1 + 3 IN2 13 IN12 + 5 IN22 23 IN12 IN2 + 19 IN14
  1
 E x 1 tr EC x C (5.47)
3

where 5 is the fifth-order shear modulus. Like the third-order model,


Ws,3 = 0 whenever the strain-energy function is restricted to be a polyno-
mial because of the constraint equation (5.27).

5.3.7 Compressible Material Model of Renown


5.3.7.1 Hookean Solid
The material model that is the classical theory of linear elasticity, when
expressed as an admissible finite-strain theory, combines the dilatoric re-
sponse of Eq. (5.44) with the deviatoric response of Eq. (5.45) according
to the decomposition of stress specified in Eq. (4.40), thereby producing
  1
S =  ln(dv/dV) 23 tr EC x C + 2 E x (5.48)

where  and are the bulk and shear moduli, respectively. This formula re-
duces to the classical theory (Marsden and Hughes 1983) under conditions
of infinitesimal strain whereby kEk 1. Recall that the Lam constant of
classical elasticity is =  23 .

5.4 Numerical Implementation

Even though explicit elastic solids have no dependence upon history,


prudence, nonetheless, requires their BVPs be solved by employing an
incremental, numeric, integration strategy to ensure a stable and conver-
gent outcome, especially in large FE and CFD codes. This is done by
128 Soft Solids

approximating a given constitutive equation with a linear dierential equa-


tion that is homogeneous in strain rate and whose tangent modulus is
acquired through a process of consistent or algorithmic linearization of
the constitutive equation in question; cf. e.g., Belytschko et al. (2000),
Holzapfel (2000), and Simo and Hughes (1998). You are encouraged to
study Appendix C before proceeding, as the various dyadic products that
arise in the description of tangent moduli are defined therein.
Appendix D presents two robust methods for solving these ODEs.
These are third-order methods, one being explicit and the other implicit.
Methods used to solve for stress often need to be more accurate than those
used to compute strain in FEA due to the degree of nonlinearity present in
a typical stress response of soft solids.

5.4.1 Incompressible Materials


Dierentiating the definition for extra stress found in Eq. (4.25) leads
to the governing dierential equation
dS = C 1 d} } dC 1 + d (5.49)
where } is the Lagrange multiplier which enforces an isochoric constraint
and d} is its rate of change. The Lagrange multiplier } is typically an un-
known function in an analysis, except for simple BVPs like those addressed
in this text.
With this objective in mind, it is convenient to introduce a tangent mod-
ulus M = MIJKL eI eJ eK eL such that3
d
d = M W dE or d = W dE (5.50)
dE
where M = d /dE = dIJ/dEKL eI eJ eK eL will depend only upon
the material, while } will depend upon both the material and the BVP to
which it is being applied. Equation (5.49) therefore becomes
 
dS = C 1 d} + M + 2} C 1 C 1 W dE (5.51)
where use has been made of the strain-rate identity (3.46), along with the
fact that dC 1 = 2 dE. The contravariant tensors M and C 1C 1 are of
fourth rank, with the inner dyadic product being defined in Appendix C.
Up to this point, the mathematical properties of M remain unspecified,
3 Green strain rate is chosen because it is the strain rate typically employed in
the solvers of FE codes. The Lodge strain rate could have also been chosen, the
dierence being only cosmetic since dE = C 1  dE  C 1 from Eq. (3.46).
Explicit Elasticity 129

other than its tensorial character which governs how it maps between con-
figurations, viz., all four indices map contravariantly.
Material stability will exist, in the sense of Hill (1968), cf. Eq. (4.36),
whenever
 
tr d dE = dE W M W dE > 0 8 dE 0, (5.52)
i.e., whenever the tangent modulus M is positive definite. This result fol-
lows by contracting Eq. (5.51) with dE , which produces tr(dS dE ) =
tr(C 1 dE ) d} + dE W M W dE + 2} tr(dE dE ), where the constraint
tr(C1 dE ) = 0 follows because of the imposed isochoric assumption.
When this expression is substituted into Hills result (4.36), the outcome
is the inequality stated above. The tangent modulus M in Eq. (5.52) is in
the form of a Hessian. Nicholson (2013) has recently developed a tool that
can probe a region in state space to determine if such a Hessian is positive
definite, which is otherwise a dicult task to verify analytically.
Equations (5.51 and 5.52) provide a general theoretical framework that
can be readily coded into software, after which time the implementation of
a new explicit elastic solid of interest to an engineer would only require the
derivation of its respective tangent modulus. This is eectively how most
existing commercial FE packages operate. They have the capability of
allowing a user to supply their own material model by linking a subroutine
for a users tangent modulus with the vendors software package.

5.4.1.1 Incompressible Green Materials


An incompressible Green elastic solid (5.3) has a consistent tangent mod-
ulus that is quantified via
d2 W (E )
M= (5.53)
dE dE
for any admissible strain-energy function W (E ). A methodology for im-
plementing M into software so that BVPs can be solved is the topic of
Appendix C. Inner and outer dyadic products are defined therein; they
appear in the following tangent moduli.
First-Order Model: Substituting the elastic strain-energy function that
describes the first-order Green elastic solid (5.29) into Eq. (5.53) leads to
a tangent modulus of
M = 2 1 C 1 C 1 . (5.54)
130 Soft Solids

Third-Order Model: Substituting the elastic strain-energy function that


describes the third-order Green elastic solid (5.30) into Eq. (5.53) leads to
a tangent modulus of
 
M = 2 1 + 3 I2 C 1 C 1 + 4 3 C 1EC 1 C 1EC 1 . (5.55)

Fifth-Order Model: Substituting the elastic strain-energy function that


describes the fifth-order Green elastic solid (5.31) into Eq. (5.53) leads
to a tangent modulus of
 
M = 2 1 + 3 I2 + 5,2 I22 C 1 C 1
 
+ 4 3 + 2 5,2 I2 C 1EC 1 C 1EC 1
 
+ 3 5,3 I3 C 1 C 1EC 1 + C 1EC 1 C 1
+ 9 5,3 C 1EC 1EC 1 C 1EC 1EC 1 . (5.56)

5.4.1.2 Incompressible Lodge Materials


An incompressible Lodge elastic solid (5.8) has a consistent tangent mod-
ulus that is described by

  d2 W (E)  1 
M = C 1 C 1 W W C C 1
dE dE !
   
dW (E) dW (E)
2 C 1 C 1 C 1 + C 1 C 1 C 1
dE dE
(5.57)
for any admissible strain-energy function W (E). Application of the chain
rule to arrive at this result requires the relationship dE/dE = C 1 C 1 ,
which comes from the identity dE = C 1  dE  C 1 , i.e., from Eq. (3.46).
The minus two in the second line comes from dC 1 = C 1  dC  C 1 =
2 C 1  dE  C 1 with C 1  dE  C 1 = (C 1 C 1 ) : dE .

First-Order Model: Substituting the elastic strain-energy function that


describes the first-order Lodge elastic solid (5.32) into Eq. (5.57) leads
to a tangent modulus of
 
M = 2 1 C 1 C 1 4 1 C 1 E + E C 1 , (5.58)
which is also the tangent modulus for the neo-Hookean solid.
Explicit Elasticity 131

MooneyRivlin Model: Mixing Eqs. (5.54) and (5.58) leads to the tangent
modulus of a MooneyRivlin (5.37) elastic solid, viz.,
 
M = 2 1 C 1 C 1 2(1 ) C 1 E + E C 1 . (5.59)

Third-Order Model: Substituting the elastic strain-energy function that


describes the third-order Lodge elastic solid (5.33) into Eq. (5.57) leads to
a tangent modulus of
 
M = 2 1 + 3 I2 C 1 C 1 + 4 3 E E
  
4 1 + 3 I2 C 1 E + E C 1 . (5.60)
Fifth-Order Model: Substituting the elastic strain-energy function that
describes the fifth-order Lodge elastic solid (5.34) into Eq. (5.57) leads
to a tangent modulus of
 
M = 2 1 + 3 I2 + 5,2 I22 C 1 C 1
 
+ 4 3 + 2 5,2 I2 E E
 
+ 3 5,3 I3 C 1 E + E C 1
+ 9 5,3 EC E EC E
  
4 1 + 3 I2 + 5,2 I22 C 1 E + E C 1
 
6 5,3 I3 C 1 EC E + EC E C 1 . (5.61)

5.4.2 Compressible Materials


The tangent modulus of a compressible material is considered to be made
up of two independent contributions, specifically
 
dS = V + D : dE where S = p C 1 + Sx (5.62)

with V = d(p C 1 )/dE designating the volumetric tangent modulus and


D = dSx /dE denoting the deviatoric tangent modulus, cf. Freed and Ein-
stein (2013, Corollary 2).

5.4.2.1 Compressible Green Materials


The Green elastic solid (5.20), with invariants defined according to
Eq. (5.18), has a deviatoric tangent modulus acquired by applying the chain
rule to Eq. (4.40) producing
132 Soft Solids

d2 Ws (Ex )  
D= : I I 13 C C 1
x
dE dE x
(5.63)
dSx (Ex )  
= : I I 13 C C 1
dEx

where dEx /dE = I I 13 C C 1 follows from dierentiating Eq. (3.56)


with respect to strain E (cf. Freed and Einstein 2013, Appendix B). Like-
wise, the volumetric tangent modulus becomes

1 d2 Wb ()  1  2 dWb ()  1 


V= C C 1 C C 1
9 d d 3 d
1 dp()  1   
= C C 1 + 2p C 1 C 1 (5.64)
3 d

where the first term comes from expanding C 1 (dp/dE ) : dE and the
second term comes from expanding p (dC 1/dE ) : dE , while using consti-
tutive equation (5.5) to quantify p. Freed and Einstein (2013, Appendix B)
determined dp/dE = 13 (dp/d) C 1 and dC 1 /dE = 2 C 1 C 1 .

Hencky Material: A Hencky volumetric response, as described by


Eq. (5.40), has a tangent modulus of
 
V =  C 1 C 1 + ln det C 1 C 1 C 1 (5.65)

where the identity  = 13 ln det F = 16 ln det C = 16 ln(det C 1 ) has been


used.

First-Order Material: The deviatoric response for the first-order Green


elastic solid (5.41) has a tangent modulus of

D = 2 1 C 1 C 1 13 C 1 C 1 . (5.66)

Equations (5.65 and 5.66), when combined with Eq. (5.62), constitute an
explicit Hookean elastic solid based upon Green strain E that extends the
classical theory of linear elasticity into the domain of finite deformations.

Third-Order Material: The deviatoric response for the third-order Green


elastic solid (5.42) has a tangent modulus of
Explicit Elasticity 133


D = 2 1 + 3 (IN2 1 IN 2 ) C 1 C 1 13 C 1 C 1
3 1

+ 4 3 C 1Ex C 1 C 1Ex C 1
 
13 IN1 C 1 C 1Ex C 1 + C 1Ex C 1 C 1

+ 19 IN12 C 1 C 1 (5.67)

where invariants IN1 and IN2 are defined in Eq. (5.18).


Fifth-Order Material: The deviatoric response for the fifth-order Green
elastic solid (5.43) has a tangent modulus of
   
D = 2 1 + 3 IN2 13 IN12 + 5 IN22 23 IN12 IN2 + 19 IN14

 C 1 C 1 13 C 1 C 1
  1
+ 4 3 + 2 5 IN2 13 IN12 C Ex C 1 C 1Ex C 1
 
13 IN1 C 1 C 1Ex C 1 + C 1Ex C 1 C 1

+ 19 IN12 C 1 C 1 . (5.68)

5.4.2.2 Compressible Lodge Materials


The Lodge elastic solid (5.28), with invariants defined according to
Eq. (5.26), has a deviatoric tangent modulus acquired through the chain
rule that looks like
 x 
 d2 Ws (E) 
D = C 1 C 1 : : C 1 C 1 13 C 1 C 1
dEx dEx
0 ! ! 1
x
dWs (E) 1 x
dWs (E) 1
2 @C 1 C 1 C + C 1 C C 1A
dEx dEx

x 
dSx (E) 
= : C 1 C 1 13 C 1 C 1
dEx
 1 
2 C Sx + Sx C 1 , (5.69)
x
where the identity dE/dE = C 1 C 1 13 C 1 C 1 follows from
x
contracting dE/dE = I I 13 C 1 C with dE/dE = C 1 C 1 arising
from dE/dC 1 = 12 I I and dC 1 /dE = 2C 1 C 1 (cf. Freed
and Einstein 2013, Appendix B). The last line in the above formula, viz.,
134 Soft Solids

2(C 1 Sx + Sx C 1 ), is not present in the deviatoric tangent modulus


of a Green elastic solid (5.63), demonstrating their uniqueness.
In contrast, the volumetric tangent modulus for Lodge elastic solids
1 d2 Wb ()  1  2 dWb ()  1 
V= C C 1 C C 1
9 d d 3 d
1 dp()  1 1   
= C C + 2p C 1 C 1 (5.70)
3 d
is the same tangent modulus for the bulk response as that of the compress-
ible Green elastic solid, viz., Eq. (5.64).

Hencky Material: A Hencky volumetric response, as described by


Eq. (5.44), has a tangent modulus of
 
V =  C 1 C 1 + ln det C 1 C 1 C 1 . (5.71)

First-Order Model: The deviatoric response for the first-order Lodge


elastic solid (5.45) has a tangent modulus of

D = 2 1 C 1 C 1 13 C 1 C 1
 
2 C 1 Ex +Ex C 1 + 2 IN1 C 1 C 1 . (5.72)
3

Equations (5.71 and 5.72), when combined with Eq. (5.62), constitute an
explicit Hookean elastic solid based upon Lodge strain E that extends the
classical theory of linear elasticity into the domain of finite deformations. It
is the compressible neo-Hookean solid and is similar to, yet distinct from,
the formulation based upon Green strain E listed in Eqs. (5.65) and (5.66).

Third-Order Model: The deviatoric response for the third-order Lodge


elastic solid (5.46) has a tangent modulus of
 
D = 2 1 + 3 (IN2 13 IN12 ) C 1 C 1 13 C 1 C 1
 
2 C 1 E x +E x C 1 + 2 IN1 C 1 C 1
3
 
+ 4 3 E x E x 1 IN1 C 1 E x C 1 + 1 IN2 C 1 C 1 . (5.73)
x +E
3 9 1

Fifth-Order Model: The deviatoric response for the fifth-order Lodge


elastic solid (5.47) has a tangent modulus of
Explicit Elasticity 135


D = 2 1 + 3 IN2 1 IN2  + IN 2 2 IN 2 IN + 1 IN 4 
3 1 5 2 3 1 2 9 1

 C 1 C 1 13 C 1 C 1

 
2 C 1 E x C 1 + 2 IN1 C 1 C 1
x +E
3
 
+ 4 3 + 2 5 IN2 13 IN12
 
 E x Ex 1 IN1 C 1 E
x +Ex C 1 + 1 IN2 C 1 C 1 . (5.74)
3 9 1

5.5 Examples

All of the examples considered in this text have a Finger deformation tensor
b that belongs to a class of deformations described by
2 3
b11 b12 0
[b] = 4b21 b22 0 5 (5.75)
0 0 b33
with b21 = b12 from its symmetry. As has already been pointed out on
page 97, the Kirchho stress s and the Finger deformation b commute
for all explicit elastic solids, i.e., [b][s] = [s][b]; cf. Moon and Truesdell
(1974). Therefore, as a direct consequence of the assumed deformations
considered herein, viz., Eq. (5.75), it follows that the Kirchho stress can,
at most, have nonzero components of
2 3
s11 s12 0
[s] = 4s21 s22 0 5 (5.76)
0 0 s33
where s21 = s12 from its symmetry, which follows from the conservation
of angular momentum. From [b][s] = [s][b] one arrives at the constraint
(s11 s22 ) b12 = (b11 b22 ) s12 (5.77)
which is satisfied for all explicit elastic solids whose deformation and stress
are special cases of Eqs. (5.75) and (5.76). Hence, this constraint equa-
tion applies to all motions discussed in this text. It is trivially satisfied for
shear-free motions. Its influence is only felt during shearing motions.
An imposition of plane stress, i.e., s33 = 0, independent of b33 , is an-
other constraint (beyond [s][b] = [b][s]) that will be imposed.
136 Soft Solids

Only BVPs for the incompressible explicit elastic solids of Green and
Lodge are solved below. Solutions for their compressible versions are left
as exercises for you to learn from.

5.5.1 Uniaxial Extension


Components of the Signorini and Almansi strains and the Kirchho stress
are found in Eqs. (3.65), (3.66), and (4.48), respectively. From these
fields, one can quantify the various constitutive responses coming from
the Green and Lodge material models listed in Eqs. (5.295.34) or the
MooneyRivlin material of Eq. (5.38). From this catalog of formul, one
can select a model that best describes the uniaxial behavior of an isotropic
incompressible material of interest, assuming it behaves like an explicit
elastic solid.

5.5.1.1 First-Order Models


Green Material: Under uniaxial extension, this material (5.29) has
a Lagrange multiplier (determined from its 22- or 33-component
equations) of
} = 1 ( 1) (5.78)
and, as such, the component quantifying nominal traction (or engineering
stress) that comes from the 11-component equation is given by
 
T = 1 1 1/ 3 , (5.79)
where T is the traction, is the stretch, and 1 is the shear modulus.
Taking the derivative of T with respect to and evaluating it at = 1
gives dT/d | =1 = 3 1 , which is the infinitesimal elastic (or Youngs)
modulus for the material, i.e., E = 3 1 . This result agrees with the classical
theory of elasticity whenever Poissons ratio is set to a half, i.e., whenever
a condition for incompressibility is imposed. At the other extreme, where
 1, the tangent response tends towards dT/d | 1 = 0, i.e., the mate-
rial response has infinite resilience there.
Lodge Material, a.k.a. the Neo-Hookean Solid: Under uniaxial exten-
sion, this material (5.32) has a Lagrange multiplier of
} = 1 (1 1/ ) (5.80)
and a nominal traction of
 
T = 1 1/ 2 . (5.81)
Explicit Elasticity 137

The derivative of T with respect to at = 1 gives dT/d | =1 = 3 1 , like


the Green solid; however, for  1, one gets dT/d | 1 = 1 , which,
unlike the Green solid, is 1/3 the stiness of the initial material response.
Consequently, a uniaxial experiment is capable of distinguishing between
Green and Lodge elastic solids.
MooneyRivlin Solid: Under uniaxial extension, this material (5.38) has
a Lagrange multiplier of
 
} = 1 (1 ) ( 1)/ +  ( 1) (5.82)
and a nominal traction of
   
T = 1 (1 ) 1/ 2 +  1 1/ 3 . (5.83)
The derivative of T with respect to at = 1 gives dT/d | =1 = 3 1 ,
while for  1, one gets dT/d | 1 = 1 (1 ). Measuring these two
tangent moduli allows one to quantify parameter  from experimental data.

5.5.1.2 Third-Order Models


If the data to be fit are too nonlinear for the first-order models to describe,
then one may need to consider a third-order model.
Green Material: Under uniaxial extension, this material (5.30) has a
Lagrange multiplier of
} = ( 1 + 3 I2 )( 1) (5.84)
and a nominal traction of
 
T = ( 1 + 3 I2 ) 1 1/ 3 , (5.85)
with the second invariant having the value
 
I2 = 14 ( 2 1)2 / 4 + 2(1 )2 . (5.86)
The derivative of T with respect to at = 1 gives dT/d | =1 = 3 1 ,
independent of 3 .
Lodge Material: Under uniaxial extension, this material (5.33) has a
Lagrange multiplier of
} = ( 1 + 3 I2 )(1 1/ ) (5.87)
and a nominal traction of
 
T = ( 1 + 3 I2 ) 1/ 2 , (5.88)
138 Soft Solids

with the second invariant having the value


 
I2 = 14 ( 2 1)2 + 2(1 )2 / 2 . (5.89)
The derivative of T with respect to at = 1 gives dT/d | =1 = 3 1 ,
independent of 3 .

5.5.1.3 Fifth-Order Models


If the data cannot be fit by either a first- or third-order model, then a
fifth-order model may need to be considered.
Green Material: Under uniaxial extension, this material (5.31) has a
Lagrange multiplier of
 
} = 1 + 3 I2 + 5,2 I22 ( 1) + 34 5,3 (1 )2 (5.90)
and a nominal traction of
  
T = 1 + 3 I2 + 5,2 I22 1 1/ 3
 
+ 34 5,3 ( 2 1)2 / 4 (1 )2 / , (5.91)
with the second invariant being given in Eq. (5.86) and the third invariant
having the value
 
I3 = 18 ( 2 1)3 / 6 + 2(1 )3 . (5.92)
The derivative of T with respect to at = 1 gives dT/d | =1 = 3 1 ,
independent of 3 , 5,2 , and 5,3 .
Lodge Material: Under uniaxial extension, this material (5.34) has a
Lagrange multiplier of
 
} = 1 + 3 I2 + 5,2 I22 ( 1)/ + 34 5,3 (1 )2 / 2 (5.93)
and a nominal traction of
  
T = 1 + 3 I2 + 5,2 I22 1/ 2
 
+ 34 5,3 ( 2 1)2 (1 )2 / 2 / , (5.94)
with the second invariant being given in Eq. (5.89) and the third invariant
having the value
 
I3 = 18 ( 2 1)3 + 2(1 )3 / 3 . (5.95)
The derivative of T with respect to at = 1 gives dT/d | =1 = 3 1 ,
independent of 3 , 5,2 , and 5,3 .
The first-, third-, and fifth-order elastic solids predict the same uniaxial
response within a neighborhood surrounding the stress-free state. Their
responses only dier at the larger stretches.
Explicit Elasticity 139

5.5.2 Equi-biaxial Extension


Components of the Signorini and Almansi strains and the Kirchho stress
are found in Eqs. (3.71), (3.72), and (4.57), respectively. From these
fields, one can quantify the various constitutive responses coming from
the Green and Lodge material models listed in Eqs. (5.295.34), or the
MooneyRivlin material of Eq. (5.38). From this catalog of formul,
one can choose a model that best describes the equi-biaxial behavior of
an isotropic incompressible material of interest, given that it behaves like
an explicit elastic solid.

5.5.2.1 First-Order Models


Green Material: Under equi-biaxial extension, this material (5.29) has a
Lagrange multiplier (determined from its 33-component equation) of

} = 1 ( 4 1) (5.96)
and, as such, the component quantifying nominal traction (determined
from the 11- or 22-component equations) is given by
 
T = 1 3 1/ 3 (5.97)
where T is the traction, is the stretch, and 1 is the shear modulus.
Taking the derivative of T with respect to and evaluating it at =
1 gives dT/d | =1 = 6 1 , so equi-biaxial extension produces an initial
response with twice the stiness of uniaxial extension, which agrees with
the classical theory of elasticity. The large stretch response of this material
grows with the cube in stretch.

Lodge Material, a.k.a. the Neo-Hookean Solid: Under equi-biaxial ex-


tension, this material (5.32) has a Lagrange multiplier of

} = 1 (1 1/ 4 ) (5.98)
and a nominal traction of
 
T = 1 1/ 5 . (5.99)
The derivative of T with respect to at = 1 gives dT/d | =1 = 6 1 . This
is a well-established experimental observation (Treloar 1975). At large
stretches  1 the tangent modulus becomes dT/d | 1 = 1 , the same
as in uniaxial extension.
140 Soft Solids

Mooney-Rivlin Solid: Under equi-biaxial extension, this material (5.38)


has a Lagrange multiplier of
 
} = 1 (1 ) ( 4 1)/ 4 +  ( 4 1) , (5.100)
and a nominal traction of
   
T = 1 (1 ) 1/ 5 +  3 1/ 3 . (5.101)

The derivative of T with respect to at = 1 gives dT/d | =1 = 6 1 ,


independent of .

5.5.2.2 Third-Order Models


Green Material: Under equi-biaxial extension, this material (5.30) has a
Lagrange multiplier of

} = ( 1 + 3 I2 )( 4 1) (5.102)
and a nominal traction of
 
T = ( 1 + 3 I2 ) 3 1/ 3 (5.103)
whose second invariant has the value
 
I2 = 14 2( 2 1)2 / 4 + (1 4 )2 . (5.104)
The derivative of T with respect to at = 1 gives dT/d | =1 = 6 1 ,
independent of 3 .

Lodge Material: Under equi-biaxial extension, this material (5.33) has a


Lagrange multiplier of

} = ( 1 + 3 I2 )(1 1/ 4 ) (5.105)
and a nominal traction of
 
T = ( 1 + 3 I2 ) 1/ 5 (5.106)
whose second invariant has the value
 
I2 = 14 2( 2 1)2 + (1 4 )2 / 8 . (5.107)
The derivative of T with respect to at = 1 gives dT/d | =1 = 6 1 ,
independent of 3 .
Explicit Elasticity 141

5.5.2.3 Fifth-Order Models


Green Material: Under equi-biaxial extension, this material (5.31) has a
Lagrange multiplier of
 
} = 1 + 3 I2 + 5,2 I22 ( 4 1) + 34 5,3 (1 4 )2 (5.108)

and a nominal traction of


  
T = 1 + 3 I2 + 5,2 I22 3 1/ 3
 
+ 34 5,3 ( 2 1)2 / 4 (1 4 )2 / (5.109)

whose second invariant is given in Eq. (5.104) and whose third invariant
has the value
 
I3 = 18 2( 2 1)3 / 6 + (1 4 )3 . (5.110)

The derivative of T with respect to at = 1 gives dT/d | =1 = 6 1 ,


independent of 3 , 5,2 and 5,3 .

Lodge Material: Under equi-biaxial extension, this material (5.34) has a


Lagrange multiplier of
 
} = 1 + 3 I2 + 5,2 I22 ( 4 1)/ 4 + 34 5,3 (1 4 )2 / 8 (5.111)

and a nominal traction of


  
T = 1 + 3 I2 + 5,2 I22 1/ 5
 
+ 34 5,3 ( 2 1)2 (1 4 )2 / 8 / (5.112)

whose second invariant is given in Eq. (5.107) and whose third invariant
has the value
 
I3 = 18 2( 2 1)3 + (1 4 )3 / 12 . (5.113)

The derivative of T with respect to at = 1 gives dT/d | =1 = 6 1 ,


independent of 3 , 5,2 and 5,3 .
The first-, third-, and fifth-order elastic solids predict the same
equi-biaxial response within a neighborhood surrounding the stress-free
state. Their responses only dier at the larger stretches.
142 Soft Solids

5.5.3 Simple Shear


The constraint arising from the commuting of Kirchho stress with the
Finger deformation, viz., Eq. (5.77), restricts simple shear in that
s11 s22 =  s12 . (5.114)
This result applies to all constitutive equations derived from the explicit
theory for isotropic elasticity, which includes all models in this chapter.
The left-hand side of Eq. (5.114) is referred to as the first normal-stress
dierence in the rheological literature, cf. e.g., Bird et al. (1987a,b). Lodge
(1984) has argued that admissible viscoelastic constitutive equations ought
to reproduce this elastic result under jump conditions in shear strain, and
those that do not obey it ought to be dismissed from further considera-
tiona means for winnowing out bad constitutive equations. This may
be a bit harsh. Certainly, what it does provide is a litmus test to determine
whether the instantaneous elastic response of a material is governed by an
explicit elastic solid or by some other type of elastic solid, e.g., the implicit
elastic solid of Chap. 6.
The nonzero components of traction that arise during simple shear are
established in Eq. (4.63); they are
      
T1 s11  s12 s12 0 s
{T }e2 = = = 12 (5.115)
T2 s21  s22 s22 1 s22
where T1 = s12 is the shearing traction and T2 = s22 is the normal traction,
both residing on the face whose normal N aligns with the 2-direction in
Fig. 4.4. These components of traction equate with two of the three com-
ponents of the Kirchho stress. The third component, i.e., s11 , is carried
internally through the grips and cannot be measured directly, with s33 = 0
following from the plane-stress assumption. Nevertheless, because of the
constraint equation (5.114), it follows that s11 =  s12 + s22 for the models
of this chapter.

5.5.3.1 First-Order Models


Green Material: From the strain components in Eq. (3.78) and the stress
components of Eq. (4.60), the response of this material (5.29) in simple
shear is
 
s11 s12 0 
= 1 (5.116)
s21 s22   2
Explicit Elasticity 143

and, therefore,
T1 = 1  and T2 = 1  2 =  T1 , (5.117)
which predicts a linear response in shear with a compressive quadratic re-
sponse normal to the shear plane.
Lodge Material, a.k.a. the Neo-Hookean Solid: From the strain compo-
nents in Eq. (3.79) and the stress components of Eq. (4.60), the response
of this material (5.32) in simple shear is
  2
s11 s12  
= 1 (5.118)
s21 s22  0
and, therefore,
T1 = 1  and T2 = 0, (5.119)
which predicts a linear response in shear but no normal response in the
2-direction.
MooneyRivlin Solid: From the strain components in Eqs. (3.78) and
(3.79) and the stress components of Eq. (4.60), the response of this ma-
terial (5.38) in simple shear is
   2 
P11 P12 0  0
= 1  (5.120)
P21 P22  0  3  2
and, therefore,
T1 = 1  and T2 =  T1 . (5.121)

5.5.3.2 Third-Order Models


Green Material: From the strain components in Eq. (3.78) and the stress
components of Eq. (4.60), the response of this material (5.30) in simple
shear is
 
s11 s12   0 
= 1 + 3 I2 (5.122)
s21 s22   2
whose second invariant has the value
 
I2 = 14  2 2 +  2 (5.123)
and, therefore,
 
T1 = 1 + 3 I2  and T2 =  T1 . (5.124)
144 Soft Solids

Lodge Material: From the strain components in Eq. (3.79) and the stress
components of Eq. (4.60), the response of this material (5.33) in simple
shear is
 
s11 s12   2 
= 1 + 3 I2 (5.125)
s21 s22  0
whose second invariant has the value
 
I2 = 14  2 2 +  2 (5.126)
and, therefore,
 
T1 = 1 + 3 I2  and T2 = 0. (5.127)

5.5.3.3 Fifth-Order Models


Green Material: From the strain components in Eq. (3.78) and the stress
components of Eq. (4.60), the response of this material (5.31) in simple
shear is
 
s11 s12  2  0 
= 1 + 3 I2 + 5,2 I2
s21 s22   2
 2
3   3
+ 4 5,3 I3 (5.128)
 3  2 (1 +  2 )
whose second invariant is given in Eq. (5.123) and whose third invariant
has the value
 
I3 = 18  4 3 +  2 (5.129)
and, therefore,
 
T1 = 1 + 3 I2 + 5,2 I22  34 5,3 I3  3 ,
  (5.130)
T2 = 1 + 3 I2 + 5,2 I22  2 + 34 5,3 I3  2 (1 +  2 ).
Lodge Material: From the strain components in Eq. (3.79) and the stress
components of Eq. (4.60), the response of this material (5.34) in simple
shear is
  2
s11 s12  2  
= 1 + 3 I2 + 5,2 I2
s21 s22  0
 2
3  (1 +  2 )  3
+ 4 5,3 I3 (5.131)
3 2
Explicit Elasticity 145

whose second invariant is given in Eq. (5.126) and whose third invariant
has the value
 
I3 = 18  4 3 +  2 (5.132)
and, therefore,
 
T1 = 1 + 3 I2 + 5,2 I22  + 34 5,3 I3  3 ,
(5.133)
T2 = 34 5,3 I3  2
where we now have a Lodge model that will impose a tensile force in the
normal direction.

5.5.4 Homogeneous Planar Membranes


From the constraint (5.77) arising from [b][s] = [s][b], as it applies to the
analysis of a membrane, one determines that
   
1 22 + 2 21 (s11 s22 ) = 21 (1 22 ) 22 (1 12 ) s12 , (5.134)
which applies to all explicit elastic solids, compressible or incompressible.
The third- and fifth-order models are omitted here as their expressions
become rather involved. They can be easily acquired.

5.5.4.1 First-Order Green Material


From the strain components in Eq. (3.89) and the stress components in
Eq. (4.60), a homogeneously deformed membrane of this material (5.29)
will have a constitutive response of
 
s11 s12 1
4 22 22 21 2 21 + 1 22
= 2 (5.135)
s21 s22
2 21 + 1 22
4 21 12 22
where, from Eq. (1.23), one recalls that
= 1 2 (1 1 2 ). The Lagrange
multiplier for this BVP is
 
} = 1
2 1 , (5.136)
which follows from enforcing the plane-stress condition of s33 = 0. With
the components of Kirchho stress now known, tractions T1 and T2 pre-
dicted by the first-order Green material follow from substituting the various
components of Eq. (5.135) into Eqs. (4.80) and (4.81).
146 Soft Solids

5.5.4.2 First-Order Lodge Material, a.k.a. the Neo-Hookean


Solid
From the strain components in Eq. (3.90) and the stress components in
Eq. (4.60), a homogeneously deformed membrane of this material (5.32)
will have a constitutive response of
  2
s11 s12 +  2 2
2 2 21 + 1 22
= 1 1 21 2 2 (5.137)
s21 s22 2 1 + 1 2 22 + 22 21
2
whose Lagrange multiplier for this BVP is described by
} = 1 (1
2 ), (5.138)
which follows from enforcing the plane-stress condition of s33 = 0. With
the components of Kirchho stress now known, tractions T1 and T2 pre-
dicted by the neo-Hookean (first-order Lodge) material follow from sub-
stituting the various components of Eq. (5.137) into Eqs. (4.80) and (4.81).
Response of the MooneyRivlin solid (5.38) is gotten by mixing these
two material models, as has been done throughout this chapter.

5.6 Applications

Two distinct theories of elasticity were derived in the early pages of this
chapter, the Green and Lodge elastic solids, so named because of the strain
measures they employ. These two theories associate with dierent pairs of
thermodynamic conjugate variables. In the pages that follow, first-, third-,
and fifth-order models were constructed for each theory so you can com-
pare their similarities and contrast their dierences. Models were devel-
oped for three types of construction: general constructions, incompress-
ible constructions, and isotropic/deviatoric constructions. With all of these
models, which represent just a sampling, how does one go about selecting
an appropriate model for use? Coming to grips with this question is the
objective of the remaining pages in this chapter.
Other stress/strain pairs are admissible, too, e.g., theories based on the
Biot (1939) and Hencky (1931) strain measures and their thermodynamic
stresses have been derived (Ogden 1984), but are not discussed here. There
is also the excellent and popular elastic model of Ogden (1972), cast in
terms of principal values of stretch and stress, that is not addressed in this
text either. Principal formulations, such as Ogdens, are useful in the ap-
proach of explicit elasticity where the stress response is independent of
Explicit Elasticity 147

Fig. 5.1 A reconstruction of Figs. 5.4 and 5.6 from Treloars (1975) book, which
are experimental plots for natural rubber in uniaxial tension and compression
with the latter being constructed from an equivalent two-dimensional extension
experiment (cf. Fig. 5.5 in his book). Treloars raw data are listed in Table 5.1
at the end of the chapter. Correlations of these data with the rst-order Lodge
(5.32) and Green (5.29) elastic solids for 1 = 355 kPa are displayed as curves

deformation path, but they are probably not very useful for material the-
ories where stress depends upon the path of deformation, which includes
implicit elasticity an viscoelasticity.4 It is for this reason, and in keeping
with the intended scope of the book, that such theories are not addressed
here. The interested reader is referred to the texts of Holzapfel (2000) and
Ogden (1984).
The classic textbook written by Treloar (1975) remains a rich resource
for understanding rubbers and elastomers, both experimentally and theoret-
ically. The experimental data presented in his text have been re-digitized
by numerous authors, including this one; they are cataloged in Table 5.1 at
the end of the chapter.
Investigation begins by comparing the incompressible first-order elastic
solids of Green (5.29) and Lodge (5.32) in Fig. 5.1 against the experimental
4 Private communication with R. W. Ogden, 2013.
148 Soft Solids

data for natural rubber from Treloar, where a shear modulus of 1 =


355 kPa describes the small-deformation response. The Green elastic solid
has a domain of applicability of about 2 ( 3/4 , 1 1/2 ), while the Lodge (or
neo-Hookean) elastic solid has a substantially larger domain of applicabil-
ity of about 2 ( 1/3 , 5). A MooneyRivlin response is bounded above
and below by the Lodge and Green material responses, respectively. When
this option was considered, the author found that  = 0.02 for Treloars
data. In other words, the MooneyRivlin mixture consists of about 98 %
Lodge material and 2 % Green material for natural rubber. For all practi-
cal purposes, natural rubber is a Lodge elastic solid. This complies with
the theory of rubber elasticity derived from statistical mechanics within the
domain of stretching 2 ( 1/3 , 5).
It would be a rare engineering application that would stretch natural
rubber to strains in excess of what Fig. 5.1 shows to be the range of appli-
cability for the neo-Hookean solid. Physically, what is happening in natural
rubber at stretches beginning at about 4 and ending at about 6 is a gradual
transition from an entropic deformation caused by chemical bond rotations
into an energetic deformation where the bond angles and bond lengths are
being stretched.
More complex material models are available, e.g., those of Gent (1996)
and Ogden (1972, 1984), that provide excellent fits over the whole data set
published by Treloar (1975), if such a modeling capability is needed. Here
we shall investigate other models, viz., the higher-order models of a Lodge
elastic solid for describing stretches greater than 5. Fixing 1 at 355 kPa,
the third-order Lodge solid (5.33) was found to have an optimal value of
3 = 390 Pa.5 When graphed in Fig. 5.2, it becomes clear that the non-
linearity of the third-order eect comes on too quickly. Consequently, the
fifth-order Lodge solid (5.34) was considered, turning o the third-order
contribution, resulting in the satisfactory outcome displayed in Fig. 5.2.
No information is revealed in these data to make an informed selection of
5,2 over 5,3 , or vice versa, or for a partitioning between them. Neverthe-
less, selecting 3 = 5,2 = 0, which led to 5,3 = 0.48 Pa, was not entirely
arbitrary. The decision to select 5,3 over 5,2 had to do with the fact that
5,3 is the only parameter in the Lodge models to lead to a tensile normal
stress during simple shear (cf. Eq. (5.133)), a phenomenon that the author
has observed recently in preliminary data acquired in his laboratory.

5 Parameters were secured by using a genetic algorithm (Goldberg 1989, 2002).


Explicit Elasticity 149

Fig. 5.2 Treloars (1975) data for natural rubber from Fig. 5.1 are redisplayed
with curves being plotted for the rst- ( 1 = 355 kPa), third- ( 1 = 355 kPa,
3 = 390 Pa), and fth- ( 1 = 355 kPa, 5,3 = 0.48 Pa, 3 = 5,2 = 0) order elastic
Lodge solids

5.6.1 Guidelines for Selecting an Elastic Material Model


No perfect protocol exists for the purpose of selecting one material model
over another. The following scheme is one that the author has found to be
useful.
For the material in question, begin by examining its elastic response to
uniaxial extension.
(1) If the response curve is approximately linear, select a Hookean solid.
(2) If the response curve softens, then select an explicit elastic solid:
(a) If the response curve softens to an asymptotic response, compute a
ratio of tangent moduli between the asymptotic and initial states:
i. If this ratio is approximately 1/3 , then select the neo-Hookean (or
linear Lodge) elastic solid.
ii. If this ratio is approximately 0, i.e., the material does not exhibit
a strain hardening in its asymptotic regime, then select the linear
Green elastic solid.
150 Soft Solids

iii. If this ratio lies somewhere between 0 and 1/3 , then select the
MooneyRivlin elastic solid.
(b) If the response curve softens and later hardens, the material model
will be of higher order. To select such a model:
i. Compute a ratio of tangent moduli between the plateau of in-
flection and the initial states and apply step (a) to select the
appropriate strain measure to adopt.
ii. Consider a third-order model first. If it describes the data, select
it. If its influence turns on too soon, as was the case for natu-
ral rubber, then select a fifth-order model for a more delayed
response. If it describes the data, select it.
(3) If the response curve stiens, then select an implicit elastic solid from
the next chapter.

Once a preliminary model has been selected, compare its predictions


against other BVPs, if you are fortunate enough to have such data.

5.7 Exercises

5.7.1 Pure Shear


Using the results for stress and strain that you derived in the prior chapters
for pure shear, derive expressions for the Lagrange multiplier } and the
two nominal tractions Ti , i = 1, 2, in the two loading directions (recall that
only T1 can be experimentally measured) for the first- (5.32), third- (5.33),
and fifth- (5.34) order neo-Hookean solids. You will now be able to predict
T2 .
Determine the stiness of these models at the onset of deformation
where = 1, viz., derive dT1 /d | =1 . How do they compare with the ini-
tial stiness predicted during uniaxial and equi-biaxial extensions for these
material models? Determine their stinesses as becomes large. How do
they compare with their uniaxial and equi-biaxial counterparts?

5.7.2 Biaxial Extension


Using the results for stress and strain that you derived in the prior chapters
for biaxial extension, derive expressions for the Lagrange multiplier } and
the two nominal tractions Ti , i = 1, 2, in the two loading directions for
Explicit Elasticity 151

the first- (5.32), third- (5.33), and fifth- (5.34) order neo-Hookean solids.
Show that they reduce to their equi-biaxial cases whenever T1 = T2 .

5.7.3 Extension Followed by Simple Shear


Using the results for stress and strain that you derived in the prior chapters
for simple shear following an axial extension, derive expressions for the
Lagrange multiplier } and the two components of traction Ti , i = 1, 2,
acting on the surface whose normal N aligns with the 2-direction for the
first- (5.32), third- (5.33), and fifth- (5.34) order neo-Hookean solids. You
will now be able to predict the 11-component of stress.

5.7.4 Other Problems


(1) Using the chain rule, show that S = @W (E )/@E is equivalent to the
Green elastic solid P T = @W (F)/@F, as implied by Eqs. (5.2) and
(5.7). Hint: Prove that tr(P T dF) = tr(S dE ).
(2) Show that the three invariants arising from the CayleyHamilton the-
orem, viz., I, II, and III, for the Finger deformation tensor b are equal
to their counterparts for the Green deformation tensor C , i.e., prove
that I(b) = I(C ), II(b) = II(C ), and III(b) = III(C ).
(3) Derive Eq. (5.14) from Eqs. (5.3) and (5.12). Hint: Use index nota-
tion.
(4) Prove that Eq. (5.15) is Eq. (5.14) pushed forward from 0 into .
(5) Derive the component form of the neo-Hookean constitutive equation
for a planar membrane in the material configuration 0 . Note that
Eq. (5.137) is for the spatial configuration .
(6) Using Eqs. (5.12), (A.77), (A.78), and (A.80) and
 
I1 = tr C 1E ,
 
I2 = tr C 1EC 1E ,
 
I3 = tr C 1EC 1EC 1E ,
derive the tangent modulus M for the classic hyperelastic solid, viz.,
derive the functional form for
@2 W (I1 , I2 , I3 )
M= where dS = M : dE ,
@E @E
which in component notation becomes
152 Soft Solids

1 @2 W (I1 , I2 , I3 ) @2 W (I1 , I2 , I3 )
MIJKL = +
4 @EIJ @EKL @EIJ @ELK
!
@2 W (I1 , I2 , I3 ) @2 W (I1 , I2 , I3 )
+ +
@EJI @EKL @EJI @ELK

where M = MIJKL eI eJ eK eL . How does this tangent modulus


compare with the one in Eq. (5.59)?
(7) Solve the compressible Hookean solid of Eq. (5.48) for (i) uniaxial
extension, (ii) equi-biaxial extension, and (iii) simple shear.
(8) Using the material parameters obtained earlier for natural rubber
for the various Lodge models, plot their predicted responses against
Treloars experimental data listed in Table 5.1 for equi-biaxial ex-
tension that are drawn in Fig. 5.3, as described in Sect. 5.5.2. How
good are these parameters at describing these data? What can you
say about their values (too big, too small, about right , etc.) as they
pertain to describing these data?
An investigation of Treloars (1975) equi-biaxial data for natural rub-
ber indicates a first-order shear modulus of 1 = 390 kPa, an increase
of roughly 10 % over the shear modulus acquired from his uniaxial
data; cf. Fig. 5.3. The author has no way of assigning a probable
cause for this discrepancy, so they are treated as independent data
sets.
(9) Using the material parameters obtained earlier for natural rubber for
the various Lodge models and your mathematical representation of
these models for pure shear that you arrived at in Sect. 5.7.1, plot
their predicted responses against Treloars experimental data listed in
Table 5.1 for pure shear that are drawn in Fig. 5.4. How good are
these parameters at describing these data? What can you say about
their values (too big, too small, about right, etc.) as they pertain to
describing these data?
(10) Taking the average of three experiments, estimate the parameters for
the neo-Hookean (5.32) elastic solid for the silicon elastomer PDMS
using the uniaxial data tabulated in Table 5.2 and shown in Fig. 5.5.
Plot your correlation against these data. Discuss the influence that
variability in these data has on the overall fitting process. Does ev-
idence exist to suggest using the MooneyRivlin or a higher-order
Lodge elastic solid over the simpler neo-Hookean solid based upon
this data set alone? Justify your position.
Explicit Elasticity 153

Fig. 5.3 A reconstruction of Fig. 11.1 from Treloars (1975) book, which is an
experimental plot for natural rubber in equi-biaxial extension. These data are
listed in Table 5.1

(11) Estimate the parameters for the neo-Hookean elastic solid for the
biological elastomer known as elastin using the uniaxial fiber data
presented in Table 5.3 and shown in Fig. 5.6. Plot your correlation
against these data. Discuss the influence that variability in these data
has on the overall fitting process. Does evidence exist to suggest using
the Hookean, MooneyRivlin, or a higher-order Lodge elastic solid
over the neo-Hookean solid based upon this data set alone? Justify
your position.
(12) Derive the experimental tractions T1 and T2 for the third- and
fifth-order Lodge elastic membranes.
154 Soft Solids

Treloar's Data for Pure Shear


Natural Rubber
2

1.5
Traction, T1 (MPa)

0.5

0
1 2 3 4 5
Stretch,

Fig. 5.4 A reconstruction of Fig. 5.8 from Treloar (1975), which is an experi-
mental plot for natural rubber in pure shear. These data are listed in Table 5.1
Explicit Elasticity 155

Fig. 5.5 Uniaxial extension data for the silicon elastomer PDMS. These data
are listed in Table 5.2
156 Soft Solids

Elastin Fibers
16

12
Traction, T (MPa)

0
1 1.5 2 2.5 3
Stretch,

Fig. 5.6 Uniaxial extension of swollen elastin bers (58 m in diameter), as


reported in Aaron and Gosline (1981). These data are listed in Table 5.3
Explicit Elasticity 157

Table 5.1 A re-digitization of experimental data for natural rubber taken


from Treloar (1975). These data are displayed in Figs. 5.1, 5.3, and 5.4

Tension Compression Equi-biaxial Pure shear


T (MPa) T (MPa) T (MPa) T1 (MPa)
1.024 0.044 0.941 0.097 1.264 0.393 1.031 0.061
1.117 0.148 0.875 0.179 1.661 0.616 1.109 0.164
1.209 0.232 0.794 0.323 1.925 0.734 1.186 0.242
1.341 0.323 0.766 0.382 2.454 0.943 1.294 0.327
1.503 0.422 0.686 0.562 3.035 1.218 1.417 0.424
1.869 0.530 0.568 0.985 3.458 1.441 1.847 0.582
2.165 0.600 0.481 1.456 3.828 1.703 2.354 0.745
2.410 0.690 0.412 2.100 4.119 1.965 2.938 0.927
3.022 0.864 0.347 3.065 4.330 2.188 3.445 1.103
3.571 1.065 4.515 2.424 3.891 1.279
4.009 1.249 4.306 1.455
4.779 1.605 4.645 1.612
5.311 1.975 4.952 1.782
5.735 2.301
6.126 2.700
6.426 3.047
6.605 3.437
6.853 3.786
7.036 4.141
7.173 4.498
7.273 4.877
7.362 5.248
7.470 5.606
7.605 6.342
158 Soft Solids

Table 5.2 Uniaxial extension data for the


silicon elastomer PDMS.a Data are from
Randall Schmidt of Dow Corning Corpora-
tion, Midland, MI, 2012. For educational
use only. They are displayed in Fig. 5.5

Stretch Traction (MPa)


T, Run 1 T, Run 2 T, Run 3
1.00 0.000 0.000 0.000
1.05 0.030 0.028 0.022
1.10 0.060 0.062 0.052
1.15 0.084 0.083 0.075
1.20 0.107 0.110 0.100
1.25 0.123 0.127 0.120
1.30 0.139 0.140 0.136
1.35 0.152 0.154 0.150
1.40 0.167 0.164 0.159
1.45 0.174 0.176 0.172
1.50 0.185 0.189 0.184
1.55 0.193 0.197 0.193
1.60 0.203 0.203 0.198
1.65 0.214 0.214 0.211
1.70 0.222 0.226 0.220
1.75 0.230 0.231 0.229
1.80 0.239 0.241 0.238
1.85 0.249 0.249 0.248
1.90 0.255 0.255 0.255
1.95 0.266 0.265 0.262
2.00 0.275 0.275 0.272
2.05 0.284 0.284 0.280
2.10 0.295 0.293 0.290
2.15 0.299 0.300 0.298
2.20 0.311 0.311 0.306
2.25 0.317 0.319 0.314
2.30 0.325 0.326 0.321
2.35 0.334 0.334 0.331
2.36 0.336
2.40 0.343 0.344
2.45 0.350 0.352
2.46 0.355
2.48 0.359
a Tensile (ASTM D412) specimens were
die cut (1.0 in. long  0.25 in. wide
gage region) from a cured lm measur-
ing 1.79 mm. in thickness and tested to
failure at ambient temperature and hu-
midity at a stretch rate of 500 mm./min
Explicit Elasticity 159

Table 5.3 Stretch vs. engineering stress


data for the axial extension of swollen
elastin bers (58 m in diameter) extracted
from puried bovine ligamentum nuchae.
These data were re-digitized from Aaron
and Gosline (1981, Fig. 2a). Each da-
tum point represents a separate experi-
ment. They are displayed in Fig. 5.6

T T
kPa kPa
1.00587 59 1.67200 4,588
1.00786 353 1.68839 5,765
1.03007 235 1.70499 4,824
1.04472 588 1.74282 4,824
1.06857 941 1.75980 6,000
1.07750 529 1.79423 5,412
1.09576 824 1.79861 6,471
1.10274 1,235 1.83362 5,706
1.11216 1,000 1.88264 6,824
1.13153 882 1.92793 6,412
1.14139 1,294 1.93710 6,235
1.16676 1,353 1.95543 5,941
1.17196 1,765 1.96926 6,765
1.22884 1,882 1.98786 7,824
1.22884 2,353 2.03474 7,824
1.26919 2,118 2.06322 6,765
1.32087 3,059 2.10144 8,647
1.33979 2,647 2.14005 9,647
1.34939 2,941 2.19378 7,882
1.36893 2,647 2.19866 8,765
1.39890 3,176 2.22819 10,059
1.42622 3,882 2.29299 10,353
1.44712 3,471 2.33829 9,235
1.47564 4,059 2.33319 11,588
1.49008 3,176 2.44524 12,118
1.51951 4,176 2.45043 10,353
1.54963 5,059 2.50204 13,412
1.56486 3,765 2.50204 15,765
1.56874 4,588 2.52282 12,294
1.61161 5,059 2.58012 13,412
1.65569 4,118 2.61161 14,765
1.65972 5,412
Chapter 6

Implicit Elasticity

Explicit elastic solids establish stress through a potential function in strain.


This potential has its origin in the thermodynamics of reversible processes.
The theory produces material models in which the value of stress only
depends upon the current state of strain. The final state of stress has no
dependence upon the path that strain has traversed in order to reach the
current state of stress (Holzapfel 2000; Marsden and Hughes 1983; Ogden
1984; Treloar 1975).
Implicit elastic solids express stress rate as a homogeneous function in
strain rate (Rajagopal 2003; Rajagopal and Srinivasa 2007, 2009). If the
modulus that contracts with strain rate is a function of stress, but not of
strain, the resulting theory is said to be hypoelastic (Bernstein 1960; Noll
1955; Truesdell 1955, 1956). Implicit elasticity, like explicit elasticity, has
its origin in the thermodynamics of reversible processes. However, unlike
explicit elasticity, where the resulting state of stress only depends upon the
imposed state of strain and not upon its history, in implicit elasticity, the
acquired state can also depend upon the path traversed through state space
in order to reach the final state. Alternatively, dierent paths traversed
through state space that reach a common final state, dierent from the ini-
tial state, may require dierent amounts of work to accomplish the same
feat, which is in accordance with recent experimental findings (Criscione
et al. 2003a,b). Each path traveled through state space can have its own
elasticity! Closed cycles, however, must not dissipate energy, derived for
the hypoelastic solid, the proof to a theorem by Noll (1955, p. 35) also ap-
plies to the more general theory of implicit elasticity, as constructed herein.
His theorem states:
Theorem 6.1. If stress rate is a homogeneous function of strain rate, then
for a given initial stress, the stress at a final state depends only on the

A.D. Freed, Soft Solids: A Primer to the Theoretical Mechanics of Materials, 161
Modeling and Simulation in Science, Engineering and Technology,
DOI 10.1007/978-3-319-03551-2_6, Springer International Publishing Switzerland 2014
162 Soft Solids

paths by which the material points reach the final state and not upon the
rate at which they traverse these paths.
Curiously, implicit elasticity, although expressed as a rate theory, has
a stress/strain response that does not depend upon time, yet it can depend
upon the path traveled. This allows us to express our theory in terms of
dierentials instead of derivatives.
Implicit elasticity, like explicit elasticity, provides a collection of elastic
material models. Whether a particular elastic material is better described
with an explicit- or implicit-elastic material model is a matter for experi-
ment to discern.

6.1 Motivation

What motivated the author (Freed 2008, 2009, 2010; Freed and Einstein
2012, 2013; Freed et al. 2010) to consider stress-rate/strain-rate elastic the-
ories over stress/strain elastic theories for describing the passive response
of soft, fibrous, biological tissues is the fact that, ever since Fung (1967)
published his pioneering work, mechanicians have known that the stress/
strain response of tissue is, to a good approximation, exponential. Even
single-molecule proteins like titin, the largest known protein in Nature (the
muscle protein), exhibit force-extension curves with an exponential quality
to them when they are pulled in the stage of an atomic force microscope
(Linke and Grtzner 2008).
Since Fungs original paper, tissue mechanicians have proposed nu-
merous ways to incorporate this exponential quality into their stress/
strain frameworks [see, e.g., the review articles of Humphrey (2002b,
2008), Sacks (2000), Sacks and Sun (2003), Viidik (1973), and Weiss and
Gardiner (2001); the textbooks of Fung (1993) and Humphrey (2002a);
and this authors works (Freed and Diethelm 2007; Freed et al. 2005)]. A
stress-rate/strain-rate theory subsumes this exponential quality through its
(numerical) integration, as the characteristic solution to a linear, first-order,
ordinary, dierential equation (ODE) is an exponential.
In his landmark paper, Fung (1967, Eq. 22) introduced an empirically
based formula for the one-dimensional passive response of soft elastic tis-
sues; it being the linear first-order ODE:
dT( )
= E + T( ) with IC T(1) = 0. (6.1)
d
This formula is known today as the Fungs law. It is a relationship between
traction T and stretch , not between stress and strain (see Eqs. 2.44 and
Implicit Elasticity 163

Y.-C. Fung's 1967 Experimental Data


500

400
Tangent Modulus, dT/d (kPa)

300

200

100

0
0 10 20 30 40 50
Traction, T (Pa)

Fig. 6.1 A reconstruction of Fung (1967, Fig. 8), which is an experimental plot
for rabbit mesentery. A tting of these data to Fungs law, viz., Eq. (6.1), pro-
duces model parameters of E = 11.7 kPa and = 10.9 from a least squares regression
of the data, whose t is drawn as a line

4.43), and it has two material parameters, E and , where E is Youngs


elastic modulus, which has units of stress, while , referred to herein as
Fungs parameter, is dimensionless. Figure 6.1 presents the original ex-
perimental data from which Fung drew his inspiration. Here a linear fit
to the data is shown, viz., Fungs law.1 Fungs law for soft solids is the
analog of Hookes law for hard solids. As we shall see later in this chapter,
Fungs law applies to many synthetic elastomers, as well as to soft biolog-
ical tissues.
Fungs law is usually presented in its integrated hyperelastic or explicit
form of
E "
T(") = e 1 (6.2)

1 In
his paper, Fung (1967) ts a quadratic curve to these data (with no constant
term E) where the contribution arising from the quadratic term was found to
be minimal ( 3 % at maximum load) and, therefore, of negligible consequence.
This allowed Fung to subsequently propose Eq. (6.1) as his empirical model,
which herein has been elevated to the status of being called a material law. My
parameter is his parameter a and my parameter E is his parameter product a.
164 Soft Solids

where " = 1 is the engineering strain; it is the 11-component from both


the Biot and BellEricksen strain tensors defined in Problems 1 and 2 in
Sect. 3.7.4 for uniaxial extension. The biomechanics community at large,
including Fung himself (1967; 1971; 1973; 1993), has been preoccupied
with this version of Fungs empirical law for over a half century in the
hope of acquiring a suitable extrapolation into a full 3D theory.
Alternatively, by multiplying both sides of Eq. (6.1) with P = d /dt,
Fungs law can be rewritten in its hypoelastic rate or implicit form of
 
P = E + T(t) (t)
T(t) P with IC T(t0 ) = 0 (6.3)
where time is now the independent variable, instead of stretch or strain,
even though the response is independent of timea consequence of Nolls
theorem. Here, the stretch of deformation is prescribed/controlled via some
known function of time
P = F (t) with IC (t0 ) = 1.
(t) (6.4)
Like hypoelasticity, implicit elasticity, in its general setting, is described by
a pair of ODEs: one provides the control, the other governs the response.
Version (6.3) of Fungs law is extended from 1D to 3D in this chap-
ter. To the resulting theory, the name Rajagopal elasticity is coined in
honor of his original conceptual thinking (Rajagopal 2003) and his ensu-
ing thermodynamic framework (Rajagopal and Srinivasa 2007, 2009) from
which Freed and Einstein (2013) were able to derive the elastic theory pre-
sented in this chapter. In the uniaxial extension, the Rajagopal elastic solid,
derived in the following pages, reduces to2
dT( ) E/ + T( )
= + T( ) (6.5)
d ln 1 1 (1 2 )
2
with IC T(1) = 0. The denominator on the right-hand side contains a
term that this author calls the Rajagopal eect. It arises out of a coupling
between the state variables for stress and strain present within the thermo-
dynamic framework of a potential function describing the elastic internal

2 The term (E/ + T)/(1 (1 2 )/2) in Eq. (6.5), which arises from thermo-
dynamic considerations, has a numerator that is akin to a Fung (1967) modulus
and a denominator that is akin to a Carton et al. (1962) compliance. The model
of Carton et al. predates Fungs model by 5 years. Their model supposes
 innites-
imal strain to be described by the formula  = max 1 eT/Emax , whose tangent
response is 1 dT/d = Emax /(max ), wherein max is the maximum allowable
strain.
Implicit Elasticity 165

energy. The Rajagopal eect becomes negligible whenever strain is su-


ciently small, under which conditions Eq. (6.5) simplifies to
dT( )
= E/ + (1 + )T( ) (6.6)
d ln
whose appearance is remarkably similar to, but not identical with, Fungs
empirical law stated in Eq. (6.1).
Hookes law is applicable for infinitesimal strains. Fungs law is appli-
cable for moderate strains. The implicit theory of elasticity, or Rajagopal
elasticity, is applicable at finite strains.

6.1.1 Attempts at Capturing Fungs Law Using Explicit Elasticity


A large number of material models have been proposed over the years
whose primary objective, although not specifically stated as such, has been
to approximate Fungs law using the theory of rubber elasticity, referred to
herein as explicit elasticity. The uniaxial stress response of an incompress-
ible, isotropic, rubber-elastic solid is described by the general formula
@W   @W  
T=2 2 + 2 1 3 , (6.7)
@I1 @I2
whose local tangent modulus is given by
dT @W 3 @W  4 
=6 +2 4 1
d @I1 @I2
@2 W  2 2 + 4 @ W 1 3 2 + 1 T (6.8)
2
+4
@I1 2 @I2 2
where I1 = tr(b) = 2 + 2 1 and I2 = tr(b1 ) = 2 + 2 in the simple
extension of an incompressible hyperelastic solid. Modelers of the past
have sought to approximate Fungs law dT/d = E + T using, eectively,
the above formula. Obviously, this would be a daunting task at best, if such
a solution can be attained at all.

6.1.1.1 Extension of neo-Hookean Elasticity


The simplest such model was proposed independently by Blatz et al. (1969)
and Demiray (1972). It has an elastic strain-energy function whose mathe-
matical structure mimics the explicit form of Fungs law (6.3), it being
 (I1 3) 
W= e 1 (6.9)
2
166 Soft Solids

where I1 3 is its strain measure. This energy function linearizes to the


neo-Hookean strain-energy function associated with Eq. (5.32), viz., W =
12 (I1 3), which was a design objective in the development of their model.
This model has a uniaxial stress response of
 
T = e( +2 3) 2 ,
2 1
(6.10)
producing a local tangent modulus of
dT  
= e( +2 3) 3 3 + 2( 2 )2 + 1 T.
2 1
(6.11)
d
Comparing the above foruml with any version of Fungs law, viz.,
Eqs. (6.1)(6.3), it becomes apparent that the above parameter has a dif-
ferent physical interpretation from the parameter in Fungs law. Specif-
ically, the analog for sought between Eqs. (6.2) and (6.9) was lost in its
implementation because of the nonlinearity in model (6.10).

6.1.1.2 Extensions of Mooney-Rivlin Elasticity


Slightly more elaborate models, when translated into our notation, are
those of Veronda and Westmann (1970)
(1 )  (I1 3)    
W= e 1 + I2 3 (6.12)
2 2
and of Vito (1973)
 
(1 ) (I1 3)+ 1

(I2 3)
W= e 1 , (6.13)
2
both of which linearize to W = (1 ) 12 (I1 3) +  12 (I2 3); it being
the Mooney-Rivlin strain-energy function associated with Eqs. (5.37) and
(5.38). Again, achieving a bridge to a well-known rubber elastic solid was
a design objective in their model developments. The Veronda-Westmann
model has a uniaxial stress response of
   
T = (1 )e( +2 3) 2 +  1 3
2 1
(6.14)
with a local tangent modulus of
dT  
= (1 )e( +2 3) 3 3 + 2( 2 )2
2 1

d
 
+  4 4 1 + 1 T, (6.15)
while Vitos model has a uniaxial stress response of
Implicit Elasticity 167

 

( 2 +2 1 3)+ 1 (2 + 2 3)
T = (1 )e
    
 2 + 1 3 (6.16)
1
with a local tangent modulus of
 

dT ( 2 +2 1 3)+ 1 (2 + 2 3)
= (1 )e
d
  4 
 3 3 + 4 1
1
 2  2  2
+ 2 2 + 2 1 3 + 1 T. (6.17)
1
Because such elastic models are, at best, complex approximations to
Fungs otherwise simple law (6.1), an alternative approach for the mod-
eling of soft biological tissues was sought, cf. (Freed 2008, 2009, 2010;
Freed and Einstein 2012, 2013; Freed et al. 2010).
We take Einsteins counsel spoken of in the foreword: Seek simplicity!

6.2 Theory

Several approaches exist for determining the thermodynamic admissibil-


ity of rate-based elastic theories. Conventional approaches, like those of
Ericksen (1958), Leonov (2000), and Bernstein and Rajagopal (2008) are
mathematically complex and often dicult to grasp, at least for this au-
thor. In what follows, a vastly simpler approach is put forward, whose
idea traces back to Rajagopal (2003, Sect. 4.2) and the ensuing papers of
Rajagopal and Srinivasa (2007, 2009), but whose development herein fol-
lows a somewhat dierent vein from what they presented, cf. Freed and
Einstein (2013). Like Albert Einstein, Rajagopal sought clarity and thereby
achieved simplicity by increasing the dimension of state space.
For purely mechanical theories of elastic solids where temperature is
held fixed, which is a reasonable consideration for soft mammalian tissues,
the Clausius-Planck inequality, i.e., Eq. (5.1), is D = tr(S dE ) dU with
D = 0 implying that there is no dissipation. This is what one means by
saying a response is elastic. Consequently, the thermodynamic function U
represents an elastic internal energy per unit volume evaluated in the mate-
rial frame 0 . Instead of assigning an explicit functional dependence for
the strain-energy function (e.g., U = W (E ) as adopted by Green (1841),
168 Soft Solids

or U = W (E) as put forward in Chap. 5), Rajagopal (2003) introduced, in


eect, the idea that U = U(E , S ) so that
D = tr(S dE ) dU(E , S ) = 0 (6.18)
where the elastic internal energy is taken to be an implicit function of state,
i.e., it is assumed to depend upon both stress and strain.
Considering U = U(E , S ) to be a continuous and suciently dieren-
tiable function of state, an expansion of dU via the chain rule allows one
to write
@U(E , S ) @U(E , S )
S : dE : dS : dE = 0. (6.19)
@S @E
This is the elastic theory introduced by Rajagopal (2003). Because the
second Piola-Kirchho stress S is now considered to be an independent
variable, which was not the case in our construction of an explicit theory
for elasticity in Chap. 5, Freed and Einstein (2013) took a derivative with
respect to stress of the above expansion gotten from the chain rule, as-
suming that the internal energy function is suciently dierentiable, and
therefrom obtained3
!
@2 U(E , S ) @2 U(E , S )
: dS = I I : dE (6.20)
@S @S @S @E
that in component notation looks like
!
1 @2 U @2 U @2 U @2 U
+ + + dSKL
4 @SIJ @SKL @SIJ @SLK @SJI @SKL @SJI @SLK

= 12 IK JL + IL JK
!!
1 @2 U @2 U @2 U @2 U
+ + + dEKL (6.21)
4 @SIJ @EKL @SIJ @ELK @SJI @EKL @SJI @ELK
where averaging the partial derivatives on both sides of the equation en-
sures symmetry of the imposed arguments for S and E . The inner dyadic
product I I defined in Eq. (C.39) is a mixed tensor field arising from
dierentiating work tr(S dE ) with respect to stress S returning I I : dE .
3 In a prior paper by the author (Freed and Einstein 2012), he dierentiated
Eq. (6.19) with respect to strain, instead of with respect to stress, as is done here
and in Freed and Einstein (2013). Dierentiating with respect to stress leads to
an elastic structure that belongs to a more general class of implicit elastic solids
introduced by Rajagopal and Srinivasa (2007, 2009).
Implicit Elasticity 169

Formula (6.20) has the structure of a Rajagopal and Srinivasa (2007)


y : dS + B
elastic solid, A y : dE = 0, where A y and B
y are fourth-rank tensors
with appropriate symmetries. This structure allows for an elastic response
of infinite stiness (i.e., dE = 0) whenever there is no compliance (viz.,
whenever B y 1 : A
y = 0). It also allows for an elastic response of infinite
resilience (i.e., dS = 0) whenever there is no modulus (viz., whenever
y 1 : B
A y = 0). Consequently, the range of thermodynamically admissible
elastic solutions is vast within the theoretical structure of Rajagopal and
Srinivasa. In their own words: The solution space is rich.
Rajagopal elasticity, as the terminology is used by this author, is a sub-
class of materials belonging to the Rajagopal and Srinivasa (2007) implicit
elastic solid. Here we shall consider that det(@2 U/@S @S ) 0 so that
(@2 U/@S @S )1 exists. Therefore, solving Eq. (6.20) for stress rate leads
to our class of material models, which belongs to a larger class, that is,
Rajagopals implicit constitutive theory (Rajagopal 2003). A Rajagopal
elastic solid, as defined herein, is described by the constitutive equation
!1 !
@2 U(E , S ) @2 U(E , S )
dS = : I I : dE (6.22)
@S @S @S @E
where stress-rate dS is a homogeneous function of strain-rate dE , as in the
theory of hypoelasticity and, accordingly, Nolls theorem stated on p. 162
applies here, too. But unlike hypoelasticity, here the theory can also depend
upon strain. Equation (6.22) has a mathematical structure like Fungs law
(6.3). This becomes apparent after one rearranges Eq. (6.3) to look like

dT = E 1 + E T d (6.23)

and, therefore, @2 U/@S @E must be first order in stress S in order to com-


ply with Fungs law.
The analogy adopted here is not between Fungs 1D phenomenological
law and an elastic strain-energy function, as was the case for the hyper-
elastic examples cited in Sect. 6.1.1. Rather, it is between Fungs 1D
phenomenological law and a 3D constitutive law that arises from thermo-
dynamics. The fourth-rank tensor @2 U/@S @S is analogous to the com-
pliance tensor from classical elasticity theory. It is the other fourth-rank
tensor, viz., @2 U/@S @E , that is new and interesting. Fungs law provides
guidance here, in that the outcome from this mixed derivative needs to be
first-order in stress. Consequently, @2 U/@S @S must have a strain depen-
dence in any thermodynamically admissible model for soft-solid elasticity
170 Soft Solids

that conforms with Fungs law, as interpreted by the theory just put for-
ward. This higher-order coupling with strain in the resilience @2 U/@S @S
is the thermodynamic origin of what this author calls the Rajagopal eect.
Material models derived from Eq. (6.20) are elastic. They do not dissi-
pate energy. They are not integrable. They are not hyperelastic in the sense
of Green (1841). Nor are they hypoelastic in the sense of Truesdell (1955).
Nevertheless, special cases can and do exist.

6.2.1 Dilatoric/Deviatoric Formulation


In some applications, it will be advantageous to decompose the rate at
which work is being done dW at a material point P into separate isotropic
dWb (bulk) and isochoric dWs (shear) contributions so that
dWb = 3p d,
dW = dWb + dWs with   (6.24)
dWs = tr Sx dEx ,
which follows from Eq. (4.41). Consequently, the dissipation of Eq. (5.1)
can be described alternatively as
Db = 3p d dUb ,
D = Db + Ds with   (6.25)
Ds = tr Sx dEx dUs
so that an elastic response will occur whenever D = 0 of which a sucient
condition is that Db = 0 and Ds = 0 both apply.
Considering that the internal energy caused by the bulk response Ub
can be described by a continuous and suciently dierentiable implicit
function of its state, i.e., Ub = Ub (, p), allows the dierential dUb ap-
pearing in Eq. (6.25) to be expanded via the chain rule so that for an elastic
response
 
@Ub (, p) @Ub (, p)
Db = 3p + d dp = 0 (6.26)
@ @p
whose gradient with respect to pressure p is
!
@2 Ub (, p) @2 Ub (, p)
3+ d dp = 0, (6.27)
@p @ @p2
which can be rearranged so that
!1 !
@2 Ub (, p) 1 @2 Ub (, p)
dp = 3 1+ d, (6.28)
@p2 3 @p @
Implicit Elasticity 171

with d being defined in Eq. (3.51). This general constitutive equation


governs the pressure or isotropic response in our implicit elastic solid.
In like manner, considering that the internal energy caused by the shear
response Us is a continuous and suciently dierentiable implicit function
of its state, i.e., Us = Us (Ex , Sx ), allows the dierential dUs appearing in
Eq. (6.25) to be expanded via the chain rule so that for an elastic response
!
@U s (Ex , Sx ) @Us (Ex , Sx )
Ds = Sx : dEx : dSx = 0, (6.29)
@Ex @Sx

whose gradient with respect to deviatoric stress Sx , taken from the left, is
!
@2 Us (Ex , Sx ) @2 Us (Ex , Sx )
I I : dEx : dSx = 0, (6.30)
@Sx @Ex @Sx @Sx

which can be rearranged so that


!1 !
@2 Us (Ex , Sx ) @2 Us (Ex , Sx )
dSx = : I I : dEx (6.31)
@Sx @Sx @Sx @Ex

with dEx being defined according to Eq. (3.55). This constitutive equation
governs the deviatoric stress response in our implicit elastic solid.
After the implicit constitutive equations for pressure (6.28) and devia-
toric stress (6.31) have been integrated, the total state of stress S can then
be quantified according to Eq. (4.40), viz., S = p C 1 + Sx .

6.3 Modulus and Compliance

Before continuing with our development, it is useful to define four


fourth-order tensors. From Eq. (6.22), one has a classical elastic com-
pliance-like contribution referred to herein as the resilience4
@2 U(E , S ; C )
A= (6.32)
@S @S

4 This is not the modulus of resilience taught in introductory courses on the


mechanics of materials. Resilience is the ability of a substance or object to
spring back into shape of which A is a measure.
172 Soft Solids

with a Fung-like adjustment of5


@2 U(E , S ; C )
B= (6.33)
@S @E
where both have minor symmetry, but only A has major symmetry, in
general.
From these definitions, and from Eq. (6.22), one arrives at an elastic
tangent modulus of
 
M = A 1 : I I B (6.34)
so that
dS = M : dE or dSIJ = MIJKL dEKL (6.35)
or, equivalently, an elastic tangent compliance of
 1
C = I I B : A, (6.36)
so that
dE = C : dS or dEIJ = CIJKL dSKL . (6.37)
It readily follows then that C = M1 or M = C 1 whenever these inverses
exist. If M is a function of stress S , then constitutive equation (6.35) is
implicit; likewise, if C is a function of strain E , then constitutive equation
(6.37) is implicit. Fourth-order tensors C and M are tangents to hyper-
surfaces in state space.
Definition 6.1. Any material model that obeys Eqs. (6.32)(6.37) is said to
be a Rajagopal elastic solid.
The models introduced in this chapter are all Rajagopal elastic solids.
Definition 6.2. Any Rajagopal elastic solid wherein B is independent of
strain and first-order in stress is said to be a Fung elastic solid.
What is defined here to be a Rajagopal elastic solid is, actually, a special
case of his more general implicit elastic theory (Rajagopal 2003).
The existence of tensors C and M tacitly requires the invertibility of
both A and I I B . This is a mathematical restriction, not a (physical)
5B is named after Y. C. Fung, not because Fung introduced this tensor eld,
for he did not. The author named it after Fung because B is the origin of the
eect in his law (6.1) when his law is extrapolated from a 1D phenomenological
model into a thermodynamically admissible 3D elastic theory.
Implicit Elasticity 173

thermodynamic restriction. In certain applications it may be reasonable to


impose a restriction of invertibility so that the resulting material model will
be well behaved in, e.g., FE and CFD implementations of the theory.
In our computational world, a strain (rate) is typically imposed to which
a stress (rate) responds, i.e., Eq. (6.35) applies. This is the opposite of re-
ality, where forces cause displacements, as described in Eq. (6.37). This
inversion between cause and eect is admissible in implicit elasticity be-
cause Eq. (6.36) is the inverse of Eq. (6.34) and vice versa.
Most commercial FE codes implement dS = M : dE as their in-
ternal constitutive equation. This allows them to admit user-definable
material models, where users can supply their own tangent moduli M .
Consequently, our implicit theory of elasticity is ideally suited for
implementation into existing nonlinear FE and CFD codes.
Explicit elastic solids can also be written as dS = M : dE . What
distinguishes an implicit solid from an explicit solid is how their tangent
moduli are derived from their respective elastic energy functions. Recall
that for an explicit, Green, elastic solid M = d2 W (E )/dE dE , Eq. (5.53),
while for an implicit elastic solid, M is given by Eqs. (6.32)(6.34).
How formul like Eqs. (6.34) and (6.36) can actually be solved in ap-
plications is addressed in Appendix C. If you are new to this field, it would
be instructive for you to study Appendix C before continuing with this
chapter, if you have not already done so, as the various dyadic products
that appear in the following pages are defined in that appendix.

6.3.1 Incompressible Materials


The general constitutive structure dS = M : dE for an incompressible
material takes on the form of
d = M : dE (6.38)
where = } C + S is the extra stress of Eq. (4.25) or, equivalently,
1
 
dS = C 1 d} + M + 2} C 1 C 1 : dE (6.39)
where } is the Lagrange multiplier that ensures an isochoric response. This
is the same formula (5.51) that was introduced in the previous chapter for
the numerical implementation of an incompressible, explicit, elastic solid.
What distinguishes these two theories, explicit vs. implicit, is not their
mathematical structure; rather it is how their tangent moduli M are de-
rived from the conservation of energy, viz., their internal energies and the
dierent potential structures with which they associate.
174 Soft Solids

6.3.2 Compressible Materials


A general constitutive structure dS = M : dE is provided for a class of
compressible materials whereby M = V + D according to Eq. (5.62), with
V pertaining to the volumetric or bulk changes, and D pertaining to the
deviatoric or shear changes. This has the same structure that was proposed
earlier for the explicit elastic solid. The explicit and implicit theories dif-
fer in how their tangent moduli V and D are constructed. Derivations of
the tangent moduli that follow can be found in Freed and Einstein (2013,
Appendix B).
Dierentiating the pressure contribution p C 1 in the decomposition
of stress S = p C 1 + Sx with respect to strain E and gathering terms
leads to a volumetric response of
1 dp 1
V = 2p C 1 C 1 C C 1 (6.40)
3 d
with the gradient term coming from Eq. (6.28), viz.,
!1 !
1 dp @2 Ub (, p) 1 @2 Ub (, p)
= 1+ . (6.41)
3 d @p2 3 @p @

In the explicit theory of elasticity , pressure p is specified through its con-


stitutive equation (5.40), while its gradient dp/d is gotten from its dier-
entiation, which leads to Eq. (5.65) for a Hencky response. In the implicit
theory of elasticity , the gradient of pressure 13 dp/d is specified through
its constitutive equation (6.28), while pressure p is gotten from its integra-
tion. The solution process is reversed.
Dierentiating the deviatoric contribution Sx in the decomposition of
stress S = p C 1 + Sx with respect to strain E leads to a deviatoric re-
sponse of
dSx  
D= : I I 13 C C 1 (6.42)
dEx
with the gradient term coming from Eq. (6.31), viz.,
!1 !
dSx @2 Us (Ex , Sx ) @2 Us (Ex , Sx )
= : I I . (6.43)
dEx @Sx @Sx @Sx @Ex
The right-hand side of Eq. (6.43) looks like Eq. (6.34), viz., it looks like
x 1 : (I I B
A x ) where dEx = (I I 1 C C 1 ) : dE .
3
Implicit Elasticity 175

6.3.3 Eulerian Formulations


Because stress S is contravariant, all four indices of tensor A obey a co-
variant map in their transfer of field from a reference configuration 0 into
the current configuration  and, as such,
aijk` = [F1 ]Ii [F1 ]Jj AIJKL [F1 ]Kk [F1 ]L` (6.44)
where a = aijk` e i e j e k e ` is the Eulerian form of A . This means
that A 1 is contravariant.
In contrast, because S is contravariant and E is covariant, the left two
indices of tensors B and I I in Eqs. (6.34) and (6.36) map covariantly,
while their right two indices map contravariantly so that
1 I 1 J KL k `
bk`
ij = [F ] i [F ] j BIJ FK FL (6.45)
where b = bk` ij e e ek e` is the Eulerian form of B . Consequently,
i j

dierences I I B B and I I bb map as mixed fields; the left two indices


(i and j) map according to a covariant law, while the right two indices (k
and `) map according to a contravariant law.
The above two field-transfer operations allow one to push the rate form
of Eqs. (6.34) and (6.35) forward from 0 into  as
M  
s = m : d wherein m = a 1 : I I b (6.46)
and to push the rate form of Eqs. (6.36) and (6.37) forward as
M  1
d = c : s wherein c = I I b : a, (6.47)
which are their Eulerian counterparts. Here we observe that all four indices
of the tangent moduli m and M obey a contravariant field-transfer map in
that
m ijk` = FIi FJ MIJKL FKk FL` ,
j
(6.48)
while all four indices of the tangent compliances c and C obey the covariant
field-transfer map present in Eq. (6.44).

6.3.3.1 Co-Rotational Formulation


Instead of using the upper-convected derivatives of Oldroyd (1950), many
mechanicians prefer to use the objective stress rate of Zaremba (1903) and
Jaumann (1911), also known as the co-rotational derivative of stress, viz.,
Ds
sV = ws + sw (6.49)
Dt
176 Soft Solids

where w is the skew-symmetric vorticity tensor defined in Eq. (2.21).


The co-rotational derivative can always be used in implicit elastic-
ity within the modulus formulation of Eq. (6.46). Following a proof by
Thomas (1955), one can always rewrite (6.46) as follows:
 : d with m
sV = m  =m+sI +I s (6.50)
where m is defined by Eq. (6.46). It turns out that the terms s I + I s
arise from one of the admissible invariants in the isotropic representation
of an elastic energy function U, as stated in Sect. 6.4, so formula (6.50)
is thermodynamically admissible, provided that m  is constructed from a
thermodynamically admissible modulus m .

6.3.3.2 Cauchy Stress Formulation


Cauchy stress is the most common measure of stress in use today, although
it is not the stress measure adopted for use herein. The Cauchy stress T ,
the Kirchho stress s, and the second Piola-Kirchho stress S obey field-
transfer maps of (Holzapfel 2000, p. 127)
1
T = FS F T and s = FS F T (6.51)
det(F)
where the former is the mapping of a relative contravariant tensor field,
while the latter pertains to an absolute contravariant tensor field (cf. e.g.,
Oldroyd (1950), see also Appendix B). Consequently, their objective Lie
derivatives obey the following maps
 1 M
T = F SP F T and s = F SP F T (6.52)
det(F)
from which one arrives at the identity
M 
s = det(F) T , (6.53)
wherein

T = DT /Dt + tr(l)T lT T l T (6.54)
is the Lie derivative of Cauchy stress, which is more commonly referred to
today as Truesdells (1953) stress rate even though its structure appeared
earlier in the seminal paper of Oldroyd (1950). Combining Eqs. (6.46) and
(6.47) with Eq. (6.53) allows one to express an equivalent set of implicit
elastic laws; in particular,
 1
T =m z :d with m z = m,
det(F) (6.55)

d = cQ : T with cQ = det(F) c
m c
wherein and are given by Eqs. (6.46) and (6.47), respectively.
Implicit Elasticity 177

For incompressible materials, where det(F) = 1 and tr(l) = 0, it follows


 M
that T  s and, therefore, m
z  m and cQ  c .

6.3.4 Stability
Material stability, in the sense of Drucker (1959) and Hill (1968), is, in
principle, straightforward to discern; specifically, any implicit elastic solid
will be stable if
8

dE : M : dE > 0 8 dE 0,



<dS : C : dS > 0
8 dS 0,
or (6.56)



d :m :d >0 8 d 0,

:M M M
s :c : s >0 8 s 0,
depending upon which formulation one chooses to use. A like result ap-
plies for the explicit elastic solid, cf. Eq. (5.52). This is stability in a math-
ematical sense. It is not a requirement of any physical law. These inequal-
ities follow from inserting Eqs. (6.34) or (6.36) into Eq. (4.32) or inserting
Eqs. (6.46) or (6.47) into Eq. (4.38), as appropriate.
A material response will be stable whenever its fourth-order tangent-
modulus/tangent-compliance matrices are positive definite. Having said
this, seldom is it a simple endeavor to prove that any given fourth-order
tensor (with minor symmetry but not necessarily with major symmetry)
is, or is not, positive definite. To address this need, Nicholson (2013) just
proposed a tool to check for positive definiteness that ought to prove useful
in future studies.

6.3.5 Plane-Stress Formulation


The plane-stress assumption, which is considered to hold for all seven
of the example BVPs studied in this text, implies that no normal com-
ponent of stress acts in the 3-direction nor are any shear components of
stress associated with the 3-direction. Only stress components within the
1-2 plane can exist; therefore, s33 = s13 = s23 = s31 = s32 = 0 and
sP 33 = sP13 = sP23 = sP31 = sP32 = 0 when specified in the current configura-
tion .
Because of these constraints, one can construct a matrix representation
for the elastic compliance of a reduced size, in particular, of dimension
four, cf. Appendix C. Once this compliance matrix is known, a matrix
178 Soft Solids

representation for the elastic modulus under plane-stress conditions, which


will also be of dimension four, can be obtained by inverting its asso-
ciated reduced compliance matrix. This process would be reversed for
plane-strain considerations, which are not addressed in this text.
In terms of the Kronecker ten operator (Nicholson 2008) of
Appendix C, the tangent compliance is constructed as ten(cc) = ten((I
a) (cf. Eq. 6.59) with
I b )1 ) ten(a
 2
1 0 0 0
3 2
b1111 b1121 b1112 b1122

3 1
60 1/2 1/2 07 6b2111 b2121 b2112 b2122 7
ten(cc) = 6 7 6
40 1/2 1/2 05 4b1211 b1221 b1212 b1222 5
7

0 0 0 1 b2211 b2221 b2212 b2222


2 3
a1111 a1121 a1112 a1122
6a2111 a2121 a2112 a2122 7
 6 7
4a1211 a1221 a1212 a1222 5 (6.57)
a2211 a2221 a2212 a2222
whose tangent modulus ten(m m) = ten(cc1 ) = ten(aa1 ) ten(I I b ) can
then be evaluated as
2 31
c1111 c1121 c1112 c1122
6 c2111 c2121 c2112 c2122 7
m) = 6
ten(m 4c1211 c1221 c1212 c1222 5
7 (6.58)
c2211 c2221 c2212 c2222
such that ten(cc) ten(m
m) = ten(I I). These are general results that apply
to any plane-stress BVP imposed upon a Rajagopal elastic solid.
Remark: This example illustrates a subtle point regarding Kronecker no-
tation; specifically, because stress and strain are symmetric, it follows that
ten(cc ) ten(cc1 ) = ten(cc1 ) ten(cc) = ten(I I) (6.59)
where I I is the identity operator for fourth-order tensors, i.e., it maps a
symmetric second-order tensor into itself, whose Kronecker form for plane
stress has a matrix representation of
2 3
1 0 0 0
60 1/2 1/2 07
ten(I I) = 6 7
40 1/2 1/2 05 . (6.60)
0 0 0 1
Therefore, care needs to be exercised when computing an inverse, e.g.,
ten(cc1 ) = ten1 (cc), when using the Kronecker notation. Classical nu-
merical methods for computing the inverse of a matrix, which assume an
Implicit Elasticity 179

identity matrix that has ones along the diagonal and zeros elsewhere, can-
not be used in this application. Nevertheless, techniques exist that employ
Crout reduction, e.g., which one can use to numerically solve Eqs. (6.57)
and (6.58) (Press et al. 2007).

6.4 Isotropic Materials

As was done in our explicit formulation for elasticity, we once again call
upon invariant theory to replace the tensorial state variables in the internal
energy function, which are now two in number, with an appropriate set of
scalar fields known as invariants. Whenever two symmetric tensor fields
are present, ten separate invariants exist (Rivlin and Smith 1969). Of these
ten, only a few are used. Using the Green C and Cauchy C 1 metrics of
the deformation to construct admissible traces, in accordance with general
tensor analysis, three contributing moment invariants exist for each tensor
argument, viz.,
     
tr C 1E , tr C 1EC 1E , tr C 1EC 1EC 1E (6.61)

tr(SC ), tr(SCSC ), tr(SCSCSC ) (6.62)


and four additional paired invariants exist, i.e.,
tr(SE ), tr(C 1ESE ), tr(SESC ), tr(SESE ) (6.63)
so that U(E , S ) ! U(I1 , I2 , : : : , I10 ) in the most general setting. Not all
ten invariants are useful; in fact, most are not. We seek guidance from other
areas, like experimental evidence, to select reasonable subsets for use.
Objectivity requires W (E , S ) = W (QE QT , QS QT ) for any orthogo-
nal rotation Q, cf. Eq. (5.13). The metric of deformation C and its inverse
C 1 ensure that covariant indices contract with contravariant indices and
vice versa, in accordance with the principles of general tensor analysis
that apply in the Lagrangian frame. They also rotate between frames as
QC QT and QC 1 QT because Q1 = QT . Consequently, any elastic in-
ternal energy function constructed in terms of these ten invariants will be
an objective isotropic function of E and S .

6.4.1 Implicit Hookean Solid


The classic theory of linear elasticity, when written in rate form, is a special
case of a Rajagopal elastic solid, just like it was a special case of explicit
180 Soft Solids

elasticity. Here an elastic energy function is constructed as a quadratic


function in stress (instead of strain) and, as such, is hypoelastic in con-
struction, viz.,
2 1+
U= tr (SC ) + tr(SCSC ) (6.64)
2E 2E
so that B = 0 and, therefore, C = (I I)1 : A = A which implies that
1+
C = C C + C C (6.65)
E E
where E is the elastic modulus and is Poissons ratio, with ! 1/2 from
below for materials whose response approaches incompressibility. The
inner and outer dyadic products are defined in Eqs. (C.39) and (C.33), re-
spectively. The outer dyadic product C C predominantly associates with
the dilational or isotropic response of a material, while the inner dyadic
product C C predominantly associates with the distortional or isochoric
response of a material.
The inverse to this compliance tensor yields a modulus tensor of
E E
M= C 1 C 1 + C 1 C 1
(1 + )(1 2 ) 1+ (6.66)
= C 1 C 1 + 2 C 1 C 1
whose derivation follows from a field-transfer operation (shown later). A
common practice is to replace the coecients E/(1+ )(12 ) and E/2(1+
) with their Lam counterparts of and , respectively.
The compliance matrix in Eq. (6.65) simplifies to its well-known cousin
from classical linear elasticity whenever C ! I, as occurs in its Eulerian
representation introduced below.
Although this is referred to as the implicit Hookean solid, it is not im-
plicit in a mathematical sense in that its compliance C does not depend
upon strain E , nor does its modulus M depend upon stress S . Never-
theless, it was derived from a theoretical framework that establishes the
implicit theory of elasticity.
Note: M becomes singular whenever = 1/2 , a condition of incompress-
ibility, because of the Lam modulus = E/(1 + )(1 2 ), yet C remains
robust. Consequently, from a theoretical viewpoint, dE = C : dS has favor
over dS = M : dE whenever one is to model an incompressible (or nearly
incompressible) Hookean response. However, most FE softwares are not
set up to use dE = C : dS . Therefore, alternative strategies have been
developed that, for example, introduce extra terms into the FE variational
principle to handle incompressibility (Belytschko et al. 2000; Bonet and
Wood 1997; Simo and Hughes 1998).
Implicit Elasticity 181

6.4.1.1 Eulerian Formulation


Because all four indices of the elastic compliance tensor C obey the co-
variant map of Eq. (6.44) in their transfer of field from the reference con-
figuration 0 to the current configuration , the compliance tensor of the
implicit Hookean solid described by Eq. (6.65) pushes forward from 0 to
 as
1+
c = I I + I I (6.67)
E E
whose inverse is well known, cf., e.g., with Marsden and Hughes (1983,
p. 241),6 i.e.,
E E
c 1 = m = I I + I I
(1 + )(1 2 ) 1+ (6.68)
= I I + 2 I I.
These fourth-order fields are the elastic compliance and modulus tensors
belonging to the classical theory of linear elasticity, but here, they are ex-
pressed in a format that remains valid for finite deformations. Because
the indices of c map covariantly, according to Eq. (6.44), the indices of
c1 = m must map contravariantly, according to Eq. (6.48), and, therefore,
the pullback of Eq. (6.68) yields Eq. (6.66).

6.4.1.2 Plane-Stress Formulation


The elastic compliance (6.67) of an isotropic Hookean solid loaded in plane
stress has a Kronecker form (cf. Appendix C) that is described by
2 3 2 3
1001 1 0 0 0
6 7 6 1 1 7
60 0 0 07 + 1 + 60 /2 /2 07
ten(cc) = 4 5
E 0 0 0 0 E 40 1/2 1/2 05
1001 0 0 0 1
2 3 (6.69)
1 0 0
1 6 0 1+ /2 1+ /2 0 7
= 6 7
E 4 0 1+ /2 1+ /2 0 5
0 0 1
whose inverse, the elastic modulus (6.68), has a Kronecker form of

6 What Marsden and Hughes (1983) refer to as c is denoted as c 1 herein.


182 Soft Solids

2 3
1 0 0
E 6 60
1 /2 1 /2 07
7
m) =
ten(m (6.70)
1 2 40 1 /2 1 /2 05
0 0 1

where ten(mm) ten(cc) = ten(I I) because ten(m m) = ten(cc1 ).


Because the above compliance and modulus tensors do not depend upon
any of the state variables, viz., stress or strain, these tangent functions de-
scribe an elastic response for an isotropic Hookean material that is appli-
cable for any plane-stress BVP.

6.4.1.3 Bulk/Shear Formulation


Inserting the expression for the decomposition of stress into hydrostatic
and deviatoric parts, as stated in Eq. (4.40), into the elastic energy function
of Eq. (6.64) allows us to rewrite the internal energy function for a Hookean
solid as
1 2 1
U= p + tr(Sx C Sx C ) (6.71)
2 4
where  = E/3(1 2 ) is the bulk modulus and = E/2(1 + ) is the shear
modulus with =  23 being the Lam modulus. The first term on
the right-hand side is the elastic bulk energy Ub . The second is the elastic
shear energy Us .
For the explicit Hookean solid, the energy function is written in terms
of strain. For the implicit Hookean solid, the energy function is written in
terms of stress.
Placing this elastic energy function into the expression for the volumet-
ric tangent modulus of an implicit elastic solid given in Eqs. (6.40) and
(6.41) leads to the formula
V = 2p C 1 C 1 +  C 1 C 1 with p = 3 (6.72)
because dp = 3 d integrates to p = 3. Equations (5.65) and (6.72)
are the same. They describe the simpler bulk response of two proposed by
Hencky (1931), which is valid to dilatations of several percent. For greater
dilatations, the reader is referred to Henckys second and more accurate
phenomenological model for describing bulk behavior or, alternatively, to
the bulk energy function derived from thermodynamics for materials B and
C below.
Implicit Elasticity 183

Likewise, placing the elastic energy function (6.71) into the expres-
sion for the deviatoric tangent modulus of an implicit elastic solid given in
Eqs. (6.42) and (6.43) leads to the formula
 
D = 2 C 1 C 1 13 C 1 C 1 , (6.73)
which is the same as the deviatoric tangent modulus of the explicit Green
elastic solid given in Eq. (5.66).
By definition, the tangent modulus M is the sum of its volumetric V
and deviatoric D parts implying that
 
M =  23 C 1 C 1 + 2( + p) C 1 C 1 (6.74)
with the hydrostatic pressure being p = 3. This fourth-order tensor
pushes forward from 0 to  to become
 
m =  23 I I + 2( + p) I I, (6.75)
which describes the same material response as does Eqs. (5.65) and (5.66)
for the explicit Hookean solid based on Green strain. It reduces to the clas-
sic linear elastic solid stated of Eq. (6.68) whenever |p| and C 1  I
because =  23 .
The fact that pressure modulates the shear response via 2p I I, in
the Eulerian frame, is a curious result that follows from dierentiating
the hydrostatic/deviatoric decomposition of stress S = p C 1 + Sx giv-
ing SP = Pp C 1 + 2p EP + SxP (cf. Sect. 4.7.1) that when pushed forward
M M
from 0 to  becomes s = Dp/Dt I + 2p d + sN = m : d . The pressure
M
term 2p I I arises from the contribution 2p d to stress rate s. This is
the same source that causes the pressure eect arising in Hills criterion
for material stability stated in Eq. (4.39). This contribution, which makes
physical sense, is missing from the Hookean solid of Eq. (6.68) derived
from an energy function that does not split the overall energy into isotropic
and isochoric constituents.

6.5 Rajagopal Elastic Solids

Three Rajagopal elastic solids are put forth, as established by Definition 6.1
stated on p. 172. The first, Material A, is an elastic solid that builds upon
the implicit Hookean solid of Sect. 6.4.1. The other two material mod-
els are elastic solids constructed within the framework of an isotropic/
deviatoric split, expanding upon the implicit Hookean solid of Sect. 6.4.1.3.
184 Soft Solids

Material B is an elastic solid that predicts final states of stress which do not
depend upon the path traversed through state space. Materials A and C
are elastic solids where the final states do depend upon the path traversed
through state space, in accordance with Nolls theorem stated on page 162
for which the experimental observations of Criscione et al. (2003a,b) are
illustrative. All three models are consistent with Fungs law, viz., they
are 3D representations of his 1D phenomenological law (6.1); specifically,
they comply with Definition 6.2 stated on p. 172.
The implicit Hookean solid is comprised of invariant combinations that
are quadratic in stress, in particular, they are from the invariants listed
in Eq. (6.62). In the spirit of Fungs law, viz., Definition 6.2, additional
invariant combinations are sought from the list (6.616.63) that are both
quadratic in stress and linear in strain and which extrapolate an existing
implicit Hookean solid.

6.5.1 Material A
An internal energy function that incorporates the Hookean elastic solid of
Eq. (6.64) and complies with Fungs law, as put forward in Definition 6.2,
is
2 1+
U= tr (SC ) + tr(SCSC ) tr(SESC ) (6.76)
2E 2E 2E
where the elastic modulus E and Poisson ratio have the same meaning
here as they do in the Hookean solid, while Fungs parameter is new to
this material definition.
B )1 : A ,
Recalling that the elastic compliance is given by C = (I I B
as defined in Eq. (6.36), then for the elastic energy function defining this
soft solid, i.e., Eq. (6.76), the embedded second-order derivatives of the
thermodynamic potential describe a resilience of
@2 U 1+
A= = C C + C C
@S @S E E
1 
C E + E C (6.77)
E 2
with a Fung adjustment of
@2 U  
B= = 12 I CS + CS I (6.78)
@S @E E
where and are separate dyadic operators with definitions found in
Eqs. (C.33), (C.39), and (C.41).
Implicit Elasticity 185

The first two terms on the right-hand side of Eq. (6.77) comprise the
elastic compliance of an isotropic Hookean solid, cf. Eq. (6.65); it is a
first-order eect. The last term in Eq. (6.77) constitutes the Rajagopal ef-
fect, which is an important second-order phenomenon. It is caused by a
coupling between the state variables for stress and strain that originates
within a thermodynamic potential of a Rajagopal and Srinivasa (2007,
2009) elastic solid.
The second thermodynamic potential B is a new elastic eect. Recall
that B = 0 for the implicit Hookean elastic solid. Through this potential,
Fungs eect enters into Rajagopals theory for elastic solids, viz., the
terms in Eqs. (6.1)(6.3).
An analytic construction of the compliance tensor C = (I I B B )1 : A
is not possible when expressed in terms of the two thermodynamic matri-
ces of Eqs. (6.77) and (6.78). This is a minor inconvenience when solving
BVPs such as those addressed in this text, but it presents no real obstacles
when implemented into software, as the theory is readily coded as an im-
plicit system of ODEs that can be solved through numerical methods, like
the implicit solver presented in Appendix D when applied to a Kronecker
representation of the model, as described in Appendix C. In practice, the
integrator of stress, e.g., using a method from Appendix D, often advances
with a finer step size than the integrator for deformation, e.g., those of
Algorithms 2.12.3.

6.5.1.1 Eulerian Formulation


All four indices of the resilience A obey the covariant map of Eq. (6.44)
in their transfer of field. Consequently, Eq. (6.77) pushes forward into the
Eulerian frame as
1+  
a= I I + I I 12 I e + e I (6.79)
E E E
wherein e is the Almansi strain of Eq. (3.10). Again, the first two terms
in the resilience a constitute the elastic compliance of an implicit Hookean
solid, while the last term introduces a Rajagopal eect, whose strain de-
pendence makes this a second-order eect whose influence is felt whenever
strains become finite.
The left two indices of the Fung adjustment B map covariantly, while
the right two indices map contravariantly, according to Eq. (6.45), and, as
such, Eq. (6.78) pushes forward into the Eulerian frame as
186 Soft Solids

1 
b = 2 I s+sI , (6.80)
E
which possesses both minor and major symmetries.

6.5.1.2 Plane-Stress Formulation


In the Eulerian configuration , this Rajagopal elastic solid has a resilience
(6.79) with a plane-stress Kronecker form (cf. Appendix C) of
2 3
1 0 0
1 6 0 1+ /2 1+ /2 0 7
ten(aa) = 6 7
E 4 0 1+ /2 1+ /2 0 5
0 0 1
2 3
2e11 e12 e12 0
6 e12 21 (e11 + e22 ) 21 (e11 + e22 ) e12 7
6 7 (6.81)
E 4 e 1 (e + e ) 1 (e + e ) e 5
12 2 11 22 2 11 22 12
0 e12 e12 2e22
and a Fung adjustment (6.80) of
2 3
2s11 s12 s12 0
6 s12 2 (s11 + s22 )
1 1
2 (s11 + s22 ) s12 7
7
b) = 6
ten(b (6.82)
E 4 s12 21 (s11 + s22 ) 1 (s + s ) s 5
2 11 22 12
0 s12 s12 2s22
where the actual components of Almansi strain e ij (3.10) and Kirchho
stress s ij (4.4) will depend upon the particular BVP being solved, thereby
illustrating the implicit quality of Rajagopals (2003) implicit theory for
elasticity. Once ten(a a) and ten(bb ) are known, the tangent compliance
ten(cc) can be gotten from Eq. (6.57) whose inverse is the tangent modu-
lus ten(mm) = ten(cc1 ), where ten(cc) ten(m
m) = ten(I I).

6.5.2 Material B
The simplest model that one can consider in the material class where bulk
and shear responses are separable has internal energies described by
p2  
Ub = 1 +  , (6.83)
2
1  
Us = tr(Sx C Sx C ) 1 + tr(C 1 Ex ) (6.84)
4
Implicit Elasticity 187

where and are the bulk and shear Fung parameters, respectively, which
are dimensionless, while  and are the bulk and shear moduli from
Hookean elasticity, which have units of stress. Recall that tr(C 1 dEx ) = 0
by definition; nevertheless, tr(C 1 Ex ) 0 with negative values typically
arising in extensions, which is why the sign changes here vs. the presence
of in Materials A and C.
These two energy functions introduce the notion of a limiting state in
strain. For the bulk response, this idea originates with Hencky (1931)
who introduced it in the second of his two, phenomenological, constitutive
equations that he used to describe Bridgmans (1923) high-pressure exper-
iments. In the biological literature, this idea traces back to Carton et al.
(1962), whose model was expressed in terms of a compliance. Later, Fung
(1967) introduced a 1D model with a limiting state, like Carton et al.s, but
that was expressed in terms of a modulus.

6.5.2.1 Bulk Response


When the bulk energy function (6.83) is substituted into its governing con-
stitutive equation (6.41), an evolution equation ensues for describing the
pressure (or the dilatation), viz.,
3 + p 1 + 
dp = d or, equivalently, d = dp (6.85)
1 +  3 + p
depending upon which variable is the independent variable for the BVP of
interest. This first-order dierential equation is separable with the solution
3 p
(3 + p)(1 + ) = 3 so p = or  = (6.86)
1 +  3 + p
whose selection depends upon which variable is being controlled. Pressure
is predicted to grow without bound as  ! 1/ from above. Likewise,
dilatation is predicted to grow without bound as p ! 3/ from above.
(Recall that p = 13 tr T and, therefore, dilation grows without bound as
tr T ! 9/ from below.)
The fact that the bulk constitutive response for a Fungean solid can be
solved in a closed form is a unique and powerful result. It allows one to
write the volumetric tangent modulus strictly in terms of dilatation, viz.,
from Eqs. (6.40), (6.85), and (6.86), it follows that

V = 2p C 1 C 1 + C 1 C 1 (6.87)
(1 + )2
with pressure p being defined according to Eq. (6.86). Formula (6.87) re-
duces to its Hookean response (6.72) whenever = 0.
188 Soft Solids

6.5.2.2 Shear Response


When the shear energy function (6.84) is substituted into its governing
constitutive equation (6.43), one arrives at
 
dSx 2 1 1 x 1
=   C C S C
dEx 1 + tr C 1 Ex 2

where we have made use of the identities (C C )1 = C 1 C 1 and


(C 1 C 1 ) : C Sx C = Sx . Contracting the above expression with I I
1 C C 1 , according to Eq. (6.42), leads to the simple result
3

2
D=   C 1 C 1 13 C 1 C 1 (6.88)
1 + tr C 1 Ex

because Sx C 1 : (I I 13 C C 1 ) = 0. This tangent modulus reduces


to its Hookean response (6.74) whenever = 0.

6.5.2.3 Tangent Response


The remarkable feature of this material model is that it retains the same
tensorial structure as the Hookean elastic solid, replacing the bulk 
and shear moduli with material tangent functions /(1 + )2 and
/(1 + tr(C 1 Ex )), respectively, while Hookean pressure p = 3 is
replaced by a Hencky-like pressure p = 3/(1 + ). Consequently, an
FE or CFD implementation of a Hookean elastic solid can be easily aug-
mented to implement our Rajagopalean elastic solid denoted as Material B;
specifically, its Lagrangian and Eulerian tangent moduli are
   
M = Q 23 Q C 1 C 1 + 2 Q + p C 1 C 1 , (6.89)
   
m = Q 23 Q I I + 2 Q + p I I (6.90)

wherein
 3
Q = , Q = , and p= (6.91)
(1 + )2 1 + tr(Ne) 1 + 

with Q and Q being the respective bulk and shear tangent moduli, while eN
denotes the Almansi distortion whose value is returned by Algorithm 3.1,
where the identity tr(C 1 Ex ) = tr([F 1 ]T [Ex ] [F 1 ]) = tr(Ne) has been used.
Implicit Elasticity 189

6.5.3 Material C
There is another way that one can construct a Fungean material model
utilizing the invariants listed in Eqs. (6.61)(6.63), as they apply to the iso-
choric response, specifically
p2  
Ub = 1 +  , (6.92)
2
1 x x  x xx 
Us = tr S C S C tr S E S C (6.93)
4 4
where this bulk internal energy Ub is the same as that of Material B. It is
the shear internal energy Us that distinguishes Material C from Material B,
it being more like the contribution present in Material A.
Substituting the shear internal energy function (6.93) into its governing
constitutive equation (6.43) produces
 
dSx   1
= 2 C C C Ex + Ex C
dEx 2
 
 x x

: I I + I CS + CS I
4
so that the deviatoric tangent modulus defined in Eq. (6.42) becomes
 
  1
D = 2 C C x
C E +E C x
2

: I I 13 C C 1

1   1 
x x x 1
+ I CS + CS I 3 CSC C (6.94)
2 2
with the Lagrangian tangent modulus M = V + D being the sum of
Eqs. (6.87) and (6.94). Notice that the Fung parameter appears twice
in this expressiona consequence of the thermodynamic coupling arising
from the potential function Us assumed in Eq. (6.93). This is due to the
Rajagopal eect.
This deviatoric tangent modulus is very similar to the tangent mod-
ulus M of Material A with E and S being replaced by their deviatoric
counterparts Ex and Sx with the additional terms on the right-hand side of

the colon, i.e., 13 (C + 2 C Sx C ) C 1 , arising as a consequence of the
term 13 C C 1 in dEx /dE = I I 13 C C 1 needed to come to
dS = M : dE .
190 Soft Solids

6.6 Examples

6.6.1 Uniaxial Extension


The axial extension of an isotropic solid is described by a deformation
gradient F and a first PiolaKirchho stress P with components
2 3 2 3
00 T00
[F] = 4 0 05 , [P ] = 4 0 0 05 (6.95)
00 0 00
where and are the axial and transverse stretches, respectively, while
T is the traction, a.k.a. the engineering stress, with = 1/2 describing
an isochoric response. These variables can be experimentally quantified,
thereby allowing comparisons to be made between theory and experiment
for this BVP.

6.6.1.1 Implicit Hookean Solid


The Hookean solid, whose compliance is given in Eq. (6.65), has a uniaxial
response described by the ODEs
dT d/
= E/ + T and = (6.96)
d ln d /
with IC T(1) = 0. The response will be isochoric whenever = 1/2 , in
accordance with Sect. 2.5.2.
The linear ODE that describes the evolution of traction can be solved
in a closed form,7 producing
 
T = E 12 1 . (6.97)
The strain measure 12 ( 1 ) for the compressible Hookean solid found in
Eq. (6.97) is dierent from the strain measures of 13 (1 3 ) and 13 ( 2 )
found in Eqs. (5.79) and (5.81) for the explicit, incompressible, Green and
neo-Hookean (Lodge) solids, respectively, which is dierent from the clas-
sic strain measure of 1 that arises from the linear theory of elasticity.
7 The linear ODE
dy
= f (x) + g(x) y
dx
has the closed-form solution [see Polyanin and Zaitsev (2003, Sect. 0.1.25)]
R Z R 
y(x) = e g(x) dx e g(x) dx f (x) dx + C .
Implicit Elasticity 191

This illustrates the fact that the implicit Hookean solid is distinct from
the classical theories of linear and hyperelasticity outside the domain of
infinitesimal strain. It turns out that 12 ( 1 ) is a reasonable approxima-
tion for the true strain of Hencky (1928) in simple extension, viz., ln , cf.
Eq. (3.19).

6.6.1.2 Rajagopal Elastic Solid: Material A


Our elastic theory for the soft biological tissues, whose compliance is given
by Eqs. (6.77) and (6.78), has a uniaxial response described by the ODE8
dT E/ + T
= +T (6.98)
d ln 1 1 (1 2 )
2
with IC T(1) = 0, which reduces to the uniaxial response of the implicit
Hookean solid whenever = 0.
The denominator on the right-hand side of Eq. (6.98) contains a term
called the Rajagopal eect that, whenever strain is suciently small, can
be neglected so that
dT
= E/ + (1 + )T
d ln whenever |strain| 1 . (6.99)
d/
=
d /
The second ODE implies that the Poisson response for this material will
be the same as that of a Hookean solid in neighborhoods of small strain.
The first ODE has a mathematical structure that is very similar to Fungs
law (6.1), although the values for their respective Fung parameters dier
by 1. The 1 in the 1 + term of the above ODE arises from the convective
contribution to stress rate in the Lie derivative. It has a physical origin
rather than a constitutive origin.
The first ODE in Eq. (6.99) has a closed-form solution of
1  1+ 1 
T = E 2+ (6.100)
1 ( 1+ 1 ) is a power law. Power
where the governing strain measure 2+
laws were used prior to Fungs introduction of the exponential, e.g., in the
one were to use D = tr(CSC dE) dU(E, S ) = 0 instead of Eq. (6.18), then
8 If

Signorinis strain measure (3.13) of 12 ( 2 1) would appear in the denominator of


Eq. (6.98) instead of Almansis strain measure (3.10) of 12 (1 2 ) which would
cause this equation to saturate much sooner, i.e., at a smaller critical stretch max .
192 Soft Solids

Fig. 6.2 Curves of maximum stretch arising from the Rajagopal eect for uni-
axial and equi-biaxial ( = 1/2 ) extensions as a function of Fungs parameter

early literature on leather, cf. Morgan (1960). The above strain measure
arose from a direct integration of Eq. (6.99) that, because of the occurrence
and placement of the independent variable (viz., the stretch ) in the ODE
being solved, produced a power-law response instead of an exponential
response. This theoretical consequence is somewhat subtle.
The Rajagopal eect implies that this material model will exhibit infi-
nite stiness in a uniaxial extension at some maximum stretch max . For
this mode of deformation, this limiting stretch, in the sense of Carton et al.
(1962), is calculated to be
p
max = /( 2), (6.101)
as shown in Fig. 6.2. Equation (6.101) provides a convenient means for es-
timating the Fung parameter in materials that exhibit a limiting stretch in
their experimental response, as is often the case for soft biological tissues.

6.6.2 Equi-biaxial Extension


Considering the material to be isotropic, the deformation gradient F and
first PiolaKirchho stress P have components for equi-biaxial extension
that are described by
Implicit Elasticity 193

2 3 2 3
00 T00
[F] = 4 0 05 , [P ] = 4 0 T 05 (6.102)
00 0 00
where and are the biaxial and transverse stretches, respectively, while
T is the biaxial traction, with = 2 in an isochoric response. These
variables can be experimentally quantified, thereby allowing comparisons
to be made between theory and experiment for this BVP.

6.6.2.1 Hookean Solid


The Hookean solid, whose compliance is given in Eq. (6.65), has an
equi-biaxial response described by the ODE
dT E/
= +T (6.103)
d ln 1
with IC T(1) = 0. This linear ODE has a solution of
E 1 
T= 1 , (6.104)
1 2

which is the same response derived for a uniaxial extension, but at 1/(1 )
times the stiness, which is 2 for the incompressible case where = 1/2
and, therefore, is in agreement with the same result predicted by explicit
theories of elasticity.

6.6.2.2 Rajagopal Elastic Solid: Material A


This more general implicit elastic solid, whose compliance is given by
Eqs. (6.77) and (6.78), has an equi-biaxial response described by the ODE
dT E/ + T
= +T (6.105)
d ln 1 1 (1 2 )
2
with IC T(1) = 0, which reduces to the biaxial response of the Hookean
solid whenever = 0.
As with uniaxial loading, the Rajagopal eect limits the stretch to max
that in the biaxial case depends upon the Poisson ratio , too, specifically
s

max = (6.106)
2(1 )
which is plotted in Fig. 6.2 for the incompressible material, viz., = 1/2 .
194 Soft Solids

6.6.3 Simple Shear


The simple shearing of a compressible material requires a slightly more
general description for its deformation gradient than is required of an in-
compressible material to ensure a plane-stress response; specifically,

2 3 2 3
1 0
1 1  0
4
[F] = 0 1 05 so that F 4
= 0 1 0 5 (6.107)
00 0 0 1

while the Kirchho stress s, because of the plane-stress assumption, has


components

2 3 2 3
s11 s12 0 s11  s12 s12 0
[s] = 4s21 s22 05 so [P ] = 4s21  s22 s22 05 (6.108)
0 0 0 0 0 0

wherein  is the magnitude of shearing and is the responding stretch


across the thickness. For the case of incompressible materials, = 1 and
Eq. (6.107) becomes Eq. (2.52). The normal and shear tractions, as speci-
fied in Fig. 4.4, are, therefore, given by T2 = s22 and T1 = s12 , respectively.
This is an experiment of fundamental importance, yet it is seldom done
on solids. It is important because the eigenvectors of deformation rotate
in the body throughout the history of the deformation (Lodge 1964, pp. 66
& 281283). Published simple-shear experiments done on soft solids are
almost nonexistent. In the study of biological tissues, the passive responses
of rat septum and porcine ventricular myocardium done by Dokos et al.
(2000, 2002) are as excellent as they are rare.
Simple shear is also important because it provides a means to distin-
guish between explicit and implicit elastic materials. Explicit elastic solids
obey s11 s22 =  s12 during simple shear, which follows from the fact that
[b] [s] = [s] [b] for this material class. The challenge is acquiring s11 from
experimental data.
O-diagonal terms in the deformation and stress fields make seeking
closed-form solutions for this Rajagopal elastic solid challenging. For this
reason, a strategy best suited for numerical analysis is created; specifically,
a plane-stress solution is constructed in a Kronecker form according to
Appendix C. In the Eulerian frame , Eq. (6.47) becomes
d 2
Implicit Elasticity

3

M
s 11
195

11 c1111 c1121 c1112 c1122


M
d21 6c2111 c2121 c2112 c2122 7 s 21
=6
4c1211
7 (6.109)
d12 c1221 c1212 c1222 5 M
s 12
d22 c2211 c2221 c2212 c2222 M
s 22


that when inverted into Eq. (6.46) becomes a coupled system of ODEs
2 3
sP11 2T1 P m1111 m1121 m1112 m1122 0
TP1 T2 P 6m2111 m2121 m2112 m2122 7 P /2
6 7 (6.110)
TP 1 = T2 P + 4m1211 m1221 m1212 m1222 5 P /2
TP2 0 m2211 m2221 m2212 m2222 0
M
where the components for s came from Eq. (4.69), while the components
for d came from Eq. (2.60). Recall that the stress component s11 cannot be
gotten experimentally, but it can be predicted theoretically.

6.6.3.1 Hookean Elastic Solid


m) for the Hookean elastic solid is listed in
The tangent modulus ten(m
Eq. (6.70).

6.6.3.2 Rajagopal Elastic Solid: Material A


The resilience a of this Rajagopal elastic solid in the Eulerian frame, viz.,
Eq. (6.79), has a Kronecker form (6.81) for simple shear of
2 3 2 3
1 0 0 0 1 1 0
1 60 1+/2 1+/2 07 61  /2  /2 17
ten(a ) = 4 
05 4  /2 1 5
1+/2 1+/2 (6.111)
E 0 4E 1  /2
0 0 1 0 1 1 2
where the contribution from the second matrix constitutes the Rajagopal
eect. Likewise, the Fung adjustment b given by Eq. (6.80) has a Kro-
necker form (6.82) for simple shear of
2 3
2s11 T1 T1 0
6 T1 21 (s11 + T2 ) 21 (s11 + T2 ) T1 7
b) = 6
ten(b 7, (6.112)
E 4 T1 21 (s11 + T2 ) 12 (s11 + T2 ) T1 5
0 T1 T1 2T2
196 Soft Solids

which introduces the Fung eect into the material response. From these
two Kronecker matrices one can construct a tangent compliance according
to Eq. (6.57) whose inverse, via Eq. (6.59), establishes its tangent modulus,
as stated in Eq. (6.58).
This example further illustrates the implicit quality of a Rajagopal elas-
tic solid.

6.6.4 Homogeneous Planar Membranes


The deformation gradient F and its inverse F 1 for a compressible mem-
brane undergoing homogeneous deformation are described by
" # " #
1  1 2 0
1 1 2 1 2 0
[F] = 2 1 2 0 and F = 2 1 1 0 (6.113)
0 0
0 0
/

where is the stretch across the thickness and


is the areal stretch (1.23),
with the Kirchho stress s having components
2 3
s11 s12 0
[s] = 4s21 s22 05 , (6.114)
0 0 0

from which all other fields can be determined.


For incompressible materials, a Lagrange multiplier is introduced via
the extra stress, with coming from the isochoric constraint, i.e., =
1 .

6.6.4.1 Implicit Hookean Solid


m) for the Hookean elastic solid is listed in
The tangent modulus ten(m
Eq. (6.70).

6.6.4.2 Rajagopal Elastic Solid: Material A


This BVP is best solved numerically. The resilience ten(a a) is defined
by Eq. (6.81) where the components for Almansi strain are taken from
Eq. (3.89). The Fung adjustment ten(b b ) is defined by Eq. (6.82) where the
components for Kirchho stress are taken from Eq. (4.78). A tangent mod-
ulus is then constructed in the manner of Eqs. (6.57) and (6.58) wherein
ten(cc1 ) is calculated according to Eq. (6.59).
Implicit Elasticity 197

6.7 Applications

Figure 6.3 presents data that establish, and leave little to no doubt, that
pericardium behaves as an implicit elastic solid in the sense of Rajagopal.
The curves drawn for T vs. 2 dT/d were constructed from data taken
from uniaxial and equi-biaxial extension experiments (Freed et al. 2010).
These curves have the same initial slope, which is governed by , with
an apparent but small oset in their intercepts, which is due to a Pois-
son eect via , cf. Eqs. (6.98) and (6.105). Unlike Fungs original data,9
which are plotted in Fig. 6.1, where a straight line provides an excellent fit,
here, in Fig. 6.3, one observes a degree of curvature in these data. This is
the Rajagopal eect, which enters into the denominator of Eqs. (6.98) and
(6.105), driving them to singularity at some critical threshold in stretch.
The experimental stress/stretch plots that associate with the tangent
moduli plots in Fig. 6.3 are presented in Fig. 6.4, where the extent of the
materials nonlinearity becomes immediately apparent and the existence of
a limiting stretch is strongly suggested. Furthermore, in accordance with
the trends of Fig. 6.2, this threshold in stretch arises earlier in equi-biaxial
stretching than in uniaxial stretching. The raw data that comprise Figs. 6.3
and 6.4 are tabulated in Tables 6.16.3, which are placed at the end of the
chapter. The data of Freed et al. (2010) have been truncated here at a stress
level of 500 kPa because of the substantial experimental noise (as reported
in Freed et al.) present in their tangent data that lie between the stresses of
500 kPa and 1 MPa.
The fact that the response curves in Figs. 6.3 and 6.4 are virtually the
same in the 1- and 2-directions for the biaxial data is an excellent in-
dicator of material isotropy, in a mechanical sense, not in a physiologic
sense. The stark dierences between the stretch/stress curves from uniax-
ial and biaxial extensions observed in Fig. 6.4 are caused by nonlinearity,
not anisotropy. Nonlinear eects are often misinterpreted in the literature
as being anisotropic in origin. Such errors often stem from a persons in-
correct extrapolation from their bias/experience with linear mechanics into
an application of nonlinear mechanics. This is yet another example of why
this text forgoes any detailed discussion of classic linear elasticity, which
you can study at a later time, if you are so inclined.
9 What is not known about Fungs experimental data are the stretches that
coincide with the data that are presented in Fig. 6.1. If the strains were moderate
(larger than innitesimal, but not large in a nite sense), as this author suspects,
then the Rajagopal eect would not yet be felt and the T vs. 2 dT/d response
would be eectively linear, as observed.
198 Soft Solids

Paracardium
40

Uniaxial
Tangent Modulus, 2dT/d (MPa)

Biaxial: 1 Direction
30 Biaxial: 2 Direction

20

10

0
0 100 200 300 400 500
True Stress, T (kPa)

Fig. 6.3 A partial reconstruction of Fig. 6 from Freed et al. (2010). The stress-
stretch curves that associate with these Fung plots are displayed in Fig. 6.4

Paracardium
500

Uniaxial
400 Biaxial: 1 Direction
Biaxial: 2 Direction
True Stress, T (kPa)

300

200

100

0
1 1.05 1.1 1.15
Stretch,

Fig. 6.4 A partial reconstruction of Fig. 7 from Freed et al. (2010) showing the
uniaxial and equi-biaxial stretch/stress response of a bovine pericardium
Implicit Elasticity 199

6.8 Exercises

6.8.1 Pure Shear


Using the results you acquired from the exercises of prior chapters, derive
the ODEs that govern the response of the implicit Hookean and Rajagopal
elastic solids of Sects. 6.4.1 and 6.5, i.e., derive dT1 /d and dT2 /d for
these four material models. For the Hookean solid, solve the ODEs to get
their closed-form solutions.

6.8.2 Biaxial
Using the results you acquired from the exercises of prior chapters, derive
the ODEs that govern the response of the implicit Hookean and Rajagopal
elastic solids of Sects. 6.4.1 and 6.5, i.e., derive dT1 /d 1 and dT2 /d 2 for
these four material models.

6.8.3 Extension Followed by Simple Shear


Using the results you acquired from the exercises of prior chapters, derive
the ODEs that govern a simple-shear response imposed after an axial pre-
stretch for the implicit Hookean and Rajagopal elastic solids of Sects. 6.4.1
and 6.5. Which of these matrix equations can be solved in closed form, if
any?

6.8.4 Other Problems


(1) Perform the integrations required to get the solutions listed in
Eqs. (6.97), (6.100), or (6.104), as selected by your instructor.
(2) Estimate the model parameters for the implicit Hookean and Ra-
jagopal elastic solids using the uniaxial and equi-biaxial data listed
in Tables 6.16.3 that associate with Figs. 6.3 and 6.4. Plot your re-
sults.
(3) Using finite dierences, construct a T vs. 2 dT/d plot for the cal-
caneal fat pad data in Table 6.4. Discuss what you find.
(4) Estimate the model parameters for the implicit Hookean and Ra-
jagopal elastic solids using the uniaxial compression data for the cal-
caneal fat pad that are presented in Table 6.4. Plot your correlations
against these data and discuss your findings (Fig. 6.5).
(5) Estimate the model parameters for the implicit Hookean and
Rajagopal elastic solids using the uniaxial extension data for
200 Soft Solids

Calcaneal Fat Pad


0

-2
Traction, T (kPa)

-4

-6 N = 10

-8
0.5 0.6 0.7 0.8 0.9 1
Stretch,

Fig. 6.5 Experimental data averaged over ten compression experiments done on
the calcaneal fat pad, as reported in Miller-Young et al. (2002), and whose values
are tabulated in Table 6.4

resin-reinforced PDMS silicon elastomer that are presented in


Table 6.5 and graphed in Fig. 6.6. Plot your correlations against these
data and discuss your findings.
(6) Estimate the model parameters for the implicit Hookean and Ra-
jagopal elastic solids using the uniaxial extension data for filler-re-
inforced PDMS silicon elastomer that are presented in Table 6.6 and
graphed in Fig. 6.7. Plot your correlations against these data and dis-
cuss your findings. How do these parameters compare with those for
the resin-reinforced PDMS material of question 5?
(7) Revisit the uniaxial elastin data presented in Fig. 5.6 whose values
are tabulated in Table 5.3. Estimate the model parameters for the
implicit Hookean solid against these data. How does its ability to
fit these data compare with the capabilities of the neo-Hookean and
MooneyRivlin models of Chap. 5?
Implicit Elasticity 201

Resin Reinforced PDMS Silicon Elastomer


10

Run 1
8 Run 2
Run 3
Traction, T (MPa)

0
1 1.2 1.4 1.6 1.8 2
Stretch,

Fig. 6.6 Uniaxial data from three experiments done on a resin-reinforced PDMS
silicon elastomer. Data were provided by Randall Schmidt of Dow Corning Cor-
poration, Midland, MI, 2012, and are for educational use only. Their values are
reported in Table 6.5
202 Soft Solids

Filler Reinforced PDMS Silicon Elastomer


10

Run 1
8 Run 2
Run 3
Traction, T (MPa)

0
1 2 3 4 5 6 7 8
Stretch,

Fig. 6.7 Uniaxial data from three experiments done on a ller-reinforced PDMS
silicon elastomer. Data were provided by Randall Schmidt of Dow Corning Cor-
poration, Midland, MI, 2012, and are for educational use only. Their values are
reported in Table 6.6
Implicit Elasticity 203

Table 6.1 The raw uniaxial data of Freed et al. (2010)


for bovine paracardium displayed in Figs. 6.3 and 6.4

T 2 dT/d T 2 dT/d
kPa MPa kPa MPa
1.000 1.1 0.27 1.107 137.9 4.69
1.002 2.0 0.38 1.109 146.2 5.06
1.005 3.0 0.35 1.111 156.1 5.88
1.008 4.2 0.42 1.113 165.8 6.12
1.012 6.0 0.50 1.115 175.0 7.31
1.016 8.2 0.42 1.116 186.5 9.05
1.021 9.8 0.41 1.117 198.6 9.56
1.026 12.1 0.48 1.119 209.8 10.57
1.031 14.6 0.45 1.120 222.9 11.67
1.036 16.7 0.50 1.121 234.8 11.53
1.041 19.7 0.66 1.122 245.9 12.13
1.046 23.3 0.70 1.123 257.2 13.97
1.051 26.5 0.81 1.124 269.4 15.22
1.055 30.8 0.96 1.125 280.9 17.23
1.060 35.0 0.96 1.126 294.7 20.14
1.064 39.0 1.04 1.126 309.2 19.70
1.068 43.5 1.26 1.127 321.9 18.79
1.072 48.8 1.33 1.128 335.3 19.91
1.076 53.4 1.49 1.129 349.5 18.66
1.079 59.4 1.90 1.129 360.5 19.08
1.083 66.3 2.04 1.130 374.3 22.42
1.086 72.6 2.14 1.131 388.3 23.40
1.089 79.6 2.51 1.131 401.9 26.22
1.092 87.4 2.50 1.132 417.7 29.46
1.095 93.5 2.69 1.133 433.0 28.00
1.098 101.6 3.33 1.133 445.3 29.60
1.100 110.2 3.51 1.134 460.4 32.04
1.103 118.2 4.04 1.134 473.6 31.09
1.105 128.0 4.77 1.135 486.8 35.33
204 Soft Solids

Table 6.2 The raw equi-biaxial data in the 1-direction


of Freed et al. (2010) for bovine paracardium displayed
in Figs. 6.3 and 6.4

T 2 dT/d T 2 dT/d
kPa MPa kPa MPa
1.000 1.1 0.15 1.086 133.8 5.51
1.003 1.7 0.25 1.087 140.9 6.11
1.006 2.5 0.35 1.088 148.7 6.37
1.009 3.9 0.45 1.090 155.8 6.51
1.012 5.3 0.50 1.091 163.2 7.04
1.015 7.0 0.54 1.092 170.8 7.30
1.018 8.7 0.55 1.093 178.0 8.15
1.021 10.4 0.62 1.094 186.5 9.18
1.024 12.6 0.71 1.095 194.9 9.33
1.028 14.8 0.75 1.096 202.9 10.51
1.031 17.2 0.85 1.097 212.6 11.61
1.034 20.0 0.90 1.098 221.7 10.71
1.037 22.6 0.94 1.099 238.5 12.25
1.040 25.6 1.07 1.100 247.1 11.66
1.043 28.9 1.09 1.101 254.8 12.09
1.046 32.0 1.12 1.101 263.4 13.63
1.049 35.3 1.32 1.102 272.5 14.11
1.051 39.3 1.44 1.103 281.0 15.27
1.054 43.1 1.54 1.104 300.4 17.32
1.057 47.3 1.75 1.105 309.9 18.43
1.059 51.9 1.94 1.106 329.2 17.56
1.062 56.8 2.05 1.107 347.9 20.09
1.064 61.6 2.23 1.108 364.9 20.78
1.066 66.8 2.36 1.109 384.8 22.91
1.068 71.8 2.39 1.109 394.1 24.86
1.071 76.7 2.66 1.110 405.2 27.80
1.073 82.3 2.89 1.110 426.5 26.36
1.075 87.7 3.01 1.111 436.6 26.79
1.076 93.2 3.37 1.111 446.7 26.64
1.078 99.2 3.83 1.112 456.2 27.11
1.080 105.7 4.10 1.112 466.3 28.18
1.081 112.0 4.51 1.112 476.2 27.82
1.083 119.1 5.25 1.113 485.6 29.87
1.084 126.8 5.48 1.113 496.7 33.74
Implicit Elasticity 205

Table 6.3 The raw equi-biaxial data in the 2-direction


of Freed et al. (2010) for bovine paracardium displayed
in Figs. 6.3 and 6.4

T 2 dT/d T 2 dT/d
kPa MPa kPa MPa
1.003 1.5 0.32 1.086 127.8 5.31
1.006 2.5 0.33 1.088 134.2 5.66
1.008 3.4 0.37 1.089 141.9 6.52
1.011 4.6 0.43 1.090 149.6 6.34
1.014 5.9 0.48 1.092 156.2 6.92
1.017 7.3 0.46 1.093 164.5 7.87
1.020 8.6 0.52 1.094 172.4 7.61
1.023 10.3 0.60 1.095 179.5 8.55
1.026 12.0 0.63 1.096 188.5 9.61
1.028 13.9 0.69 1.097 196.9 9.40
1.031 16.0 0.74 1.098 204.9 10.21
1.034 18.1 0.76 1.099 222.4 10.86
1.037 20.3 0.81 1.100 230.8 11.36
1.040 22.6 0.92 1.101 239.4 12.16
1.042 25.4 0.95 1.102 248.2 12.62
1.045 27.8 1.03 1.103 265.7 15.17
1.048 30.9 1.22 1.104 276.1 15.40
1.051 34.1 1.21 1.105 294.2 17.41
1.053 37.1 1.38 1.106 304.2 16.75
1.056 41.0 1.56 1.107 322.9 20.73
1.058 44.7 1.56 1.108 341.4 21.18
1.060 48.4 1.82 1.108 352.4 23.74
1.063 53.0 2.02 1.109 362.2 22.76
1.065 57.3 2.00 1.109 371.6 25.43
1.067 61.5 2.31 1.110 382.8 27.22
1.069 66.9 2.63 1.110 393.0 26.97
1.071 71.9 2.67 1.111 403.5 29.41
1.073 77.1 2.94 1.111 415.0 31.28
1.075 82.8 3.24 1.111 426.4 30.57
1.077 88.5 3.40 1.112 436.9 29.79
1.079 94.3 3.67 1.112 447.4 31.70
1.080 100.5 4.17 1.113 458.8 31.55
1.082 107.3 4.33 1.113 468.8 30.45
1.084 113.4 4.56 1.113 479.3 34.25
1.085 120.3 5.28 1.114 491.5 34.43
206 Soft Solids

Table 6.4 Stretch vs. engineering


stress data for the quasi-static ax-
ial compression of the human, cal-
caneal, fat pad. Data are from
Miller-Young et al. (2002, Fig. 3).a
They are displayed in Fig. 6.5
T T
kPa kPa
1.00 0.0 0.75 1.27
0.99 0.0127 0.74 1.38
0.98 0.0289 0.73 1.50
0.97 0.0483 0.72 1.63
0.96 0.0708 0.71 1.76
0.95 0.0963 0.70 1.90
0.94 0.124 0.69 2.04
0.93 0.155 0.68 2.19
0.92 0.188 0.67 2.36
0.91 0.223 0.66 2.54
0.90 0.261 0.65 2.73
0.89 0.305 0.64 2.95
0.88 0.357 0.63 3.18
0.87 0.421 0.62 3.43
0.86 0.421 0.61 3.70
0.85 0.493 0.60 3.99
0.84 0.566 0.59 4.31
0.83 0.693 0.58 4.65
0.82 0.743 0.57 5.03
0.81 0.793 0.56 5.44
0.80 0.852 0.55 5.89
0.79 0.921 0.54 6.37
0.78 0.998 0.53 6.91
0.77 1.08 0.52 7.49
0.76 1.17
a Each data point represents the
mean from ten specimens taken
from multiple cadaver feet. Re-
ported standard deviation for the
experimental variability around
the mean stress for these ten ex-
periments was about 10 % of the
value at the point of maximum
compression
Implicit Elasticity 207

Table 6.5 Uniaxial extension data for a resin-re-


inforced PDMS silicon elastomer.a Data are from
Randall Schmidt of Dow Corning Corporation,
Midland, MI, 2012, and are for educational use
only. They are displayed in Fig. 6.6

Stretch Traction (MPa)


T, Run 1 T, Run 2 T, Run 3
1.00 0.000 0.000 0.000
1.05 0.200 0.200 0.210
1.10 0.335 0.335 0.340
1.15 0.460 0.460 0.470
1.20 0.585 0.575 0.580
1.25 0.740 0.720 0.735
1.30 0.900 0.975 0.890
1.35 1.095 1.065 1.080
1.40 1.315 1.285 1.305
1.45 1.605 1.570 1.590
1.50 2.050 1.950 1.990
1.55 2.550 2.470 2.520
1.60 3.240 3.110 3.190
1.65 4.280 4.100 4.220
1.70 5.400 5.190 5.300
1.75 6.650 6.400 6.530
1.76 6.680
1.80 7.700 7.420
1.82 8.095
a Tensile (ASTM D412) specimens were die cut
(1.0 in. long  0.25 in. wide gage region) from a
cured lm measuring 2.08 mm. in thickness and
tested to failure at ambient temperature and hu-
midity at a stretch rate of 500 mm./min
208 Soft Solids

Table 6.6 Uniaxial extension data for a ller-re-


inforced PDMS silicon elastomer.a Data are from
Randall Schmidt of Dow Corning Corporation,
Midland, MI, 2012, and are for educational use
only. They are displayed in Fig. 6.7

Stretch Traction (MPa)


T, Run 1 T, Run 2 T, Run 3
1.00 0.000 0.000 0.000
1.25 0.250 0.250 0.235
1.50 0.380 0.380 0.370
1.75 0.545 0.545 0.530
2.00 0.735 0.750 0.720
2.25 0.940 0.960 0.910
2.50 1.100 1.170 1.100
2.75 1.305 1.370 1.305
3.00 1.530 1.610 1.540
3.25 1.800 1.860 1.770
3.50 2.050 2.100 2.020
3.75 2.320 2.380 2.300
4.00 2.620 2.700 2.600
4.25 2.900 2.990 2.900
4.50 3.220 3.330 3.240
4.75 3.530 3.660 3.565
5.00 3.870 4.030 3.930
5.25 4.230 4.420 4.310
5.50 4.620 4.830 4.730
5.75 5.020 5.270 5.150
6.00 5.440 5.750 5.620
6.25 5.870 6.280 6.090
6.50 6.340 6.760 6.630
6.75 6.780 7.240 7.080
7.00 7.200 7.650 7.500
7.25 7.620 8.010 7.920
7.30 7.990
7.41 7.890
7.50 8.500
7.79 9.000
a Tensile (ASTM D412) specimens were die cut
(1.0 in. long  0.25 in. wide gage region) from
a cured lm measuring 1.98 mm. in thickness
and tested to failure at ambient temperature and
humidity at a stretch rate of 500 mm./min
Chapter 7

Viscoelasticity

Viscoelastic materials exhibit both elastic and viscous behaviors through


their simultaneous storage and dissipation of mechanical energies. Boltz-
mann (1874) formulated his viscoelastic theory in terms of a convolu-
tion integral with a hereditary kernel, what we now refer to as a Volterra
(1930) integral equation of the second kind. Coleman and Noll (1961)
have summarized the mathematical and physical considerations necessary
to construct a linear theory of viscoelasticity in their review paper of some
50 years ago. The textbooks by Christensen (1971) and Tschoegl (1989)
provide excellent presentations of the linear theory of viscoelasticity, while
the text by Mainardi (2010) furthers this development through an appli-
cation of the fractional calculus, to which he has assembled a substantial
bibliography. To this literature, Wineman (2009) has written a review arti-
cle whose focus is on the topic of nonlinear viscoelasticity.
The range of viscoelastic models and theories that can be found in the
literature is vast: from simple to complex and from phenomenological to
theoretical. Constructions have been developed from the physics of contin-
uous media at one end of the spectrum to the physics of molecular dynam-
ics and statistical mechanics at the other end of the spectrum. Viscoelastic
models have been cast in terms of differential and integral equations. In this
text, integral equations are chosen over differential equations, as they have
the capability of being more flexible in applications. Numerous textbooks
address this topic, to which the interested reader is referred, e.g., Bird et al.
(1987a), Bird et al. (1987b), Christensen (1971), Ferry (1980), Holzapfel
(2000), Lakes (1998), Lodge (1964, 1974), Mainardi (2010), Phan-Thien
(2002), Pipkin (1972), Simo and Hughes (1998), Truesdell and Noll (2004),
Tschoegl (1989), and Wineman and Rajagopal (2000), while the anelas-
tic (i.e., viscoelastic) effects exhibited by minerals and metals have been

A.D. Freed, Soft Solids: A Primer to the Theoretical Mechanics of Materials, 209
Modeling and Simulation in Science, Engineering and Technology,
DOI 10.1007/978-3-319-03551-2_7, Springer International Publishing Switzerland 2014
210 Soft Solids

documented by, e.g., Gittus (1975), Nowick and Berry (1972), and Zener
(1948). The text by Phan-Thien (2002) provides photographs and short bio-
sketches of some of the prominent contributors to the field over the early
years, providing a historical perspective.
Polymeric materials exhibit viscoelastic behavior at temperatures that
exceed what is known as the glass transition temperature, or Tg , which lies
below the materials melting temperature Tm . At temperatures below Tg ,
their material behavior changes to, typically, an elasticplastic solid, which
constitutes a class of material models that is not discussed in this text. (The
interested reader is referred to, e.g., Havner (1992), Kachanov (1971), and
Stouer and Dame (1996).) Most components made from polymers are
processed at temperatures between Tg and Tm .
As this is an introductory text, the approach adopted here is intended to
be more straightforward. The idea is to provide insight and understanding
by developing useful viscoelastic theories as integral extensions that are
analytic continuations of the elastic theories presented in the previous two
chapters.

7.1 1D Viscoelastic Solid

Before constructing 3D viscoelastic theories, it is instructive to investigate


the properties and behavior of a 1D linear viscoelastic material whose shear
response is described by a Boltzmann (1874) convolution or hereditary
integral of the Stieltjes type (Mainardi 2010, pp. 2527)
Z t
 (t) = G(t t0 ) d (t0 ), (7.1)
0
where  is the shear stress associating with the 12-component for stress in
Eq. (4.68), while 12  (0, t0 ) is the shear strain associating with the 12-com-
ponent for strain in Eq. (3.79), as measured between configurations 0
and 0 , associated with times t = 0 and t = t0 , and whose rate of change
is 12 d (t0 ) at time t0 in state 0 . The materials shear modulus G(t t0 ) has
units of stress. In viscoelasticity, the shear modulus appears as a hereditary
kernel, i.e., as a function in time, which contrasts with elastic theory where
it has a fixed value.
When used to describe a solid, the shear modulus G introduced above,
a.k.a. the relaxation modulus, can be expressed as
 
G(t) = 1 + 0 1 G (t), (7.2)
Viscoelasticity 211

SLOW INTERMEDIATE FAST

Fig. 7.1 Behavior of a linear viscoelastic solid exposed to cyclic loading. The
response appears to be elastic at both suciently slow and adequately fast cyclic
frequencies, while over a range in frequency between these two extremes the
stress/strain response is elliptic (for a linear response) with a clockwise rotation

where
G (t) 2 [0, 1] 8 t  0. (7.3)
Here G (t) is called the reduced relaxation function or relaxation kernel,
which is dimensionless. It is normalized so G (0) = 1 and G (1) = 0;
therefore, G(0) = 0 and G(1) = 1 , with 0 being the glassy (or in-
stantaneous) shear modulus and 1 being the rubbery (or equilibrium)
shear modulus. They order as 0 > 1  0, with 1 > 0 implying that
the material is a solid ( 1 = 0 for fluids); see Fig. 7.1.
One of the fundamental assumptions pertaining to viscoelastic kernel
functions, born out of experimental observation and thermodynamic con-
siderations (Coleman and Mizel 1968), is that the viscoelastic kernel func-
tion must be a nonincreasing (monotonically decreasing) function in time,
viz.,
G(t1 )  G(t2 )  0 8 t2 > t1  0 ) G (t1 )  G (t2 )  0 (7.4)
with G (1) = 0, which is a mathematical way of saying the material has a
fading memory.
Inserting the expression for the viscoelastic relaxation modulus G of
Eq. (7.2) into Boltzmanns viscoelastic integral (7.1), noting that G (0) = 1
and G (1) = 0, brings about an alternative 1D constitutive description for
the Boltzmann viscoelastic solid, i.e.,
Z
  t
 (t) = 1  (t) + 0 1 G (t t0 ) d (t0 ), (7.5)
0
212 Soft Solids

where an initial condition of  (0) = 0 is imposedRso that  (0) = 0 follows.1


Here a transient (viscous) response ( 0 1 ) 0t G (t t0 ) d (t0 ) is being
added to the rubbery (elastic) response of 1  .

7.1.1 Clocks
Viscoelastic materials often respond according to their own internal clock,
whose speed can vary with respect to laboratory time in the presence of
environmental factors like temperature (Ferry 1980; Williams et al. 1955),
radiation or chemical reactions (Tobolsky 1960), or biological processes
like osmosis (Lillie and Gosline 1996), or agents like hormones (Rajagopal
and Wineman 2010). They may also vary with respect to mechanical fac-
tors like the state of stress (Bernstein and Shokooh 1980; Ferry 1980) or
the state of strain (Duenwald et al. 2010). The internal or material time
that comes from such a clock replaces the actual time as the independent
variable in a models viscoelastic kernel, so the concept is fairly straight-
forward to implement, viz., only the kernel is aected. These are often
important considerations that need to be taken into account in applications,
but their modeling lies outside the intended scope of this introductory text.

7.1.2 Stress Relaxation Experiment


Many of the deformation histories that are used to characterize the visco-
elastic behavior of man-made materials [e.g., Bird et al. (1987a), Ferry
(1980), Lakes (1998), Lodge (1964), Lodge (1974), Phan-Thien (2002),
and Tschoegl (1989)] are too imposing for living tissues, eectively
killing them over the duration of the experiment. Another challenge that
arises when testing living matter is that their physiologic dimensions are
often too small for the testing apparatuses ordinarily used by engineers to
determine the viscoelastic properties of man-made materials. For example,
Dokos et al. (2000) designed and built a unique miniature testing apparatus
so that they could perform simple-shear experiments on myocardial heart
muscle excised from rats.
Stress relaxation is a viable experiment for soft materials of all kinds,
be they made by human or by Nature. This BVP has even been used to
describe seismic afterslip around a fault caused by a viscoelastic mantle
1 Jump osets at time 0, as occur in an idealized creep or relaxation experiment,
are not addressed here. The interested reader is referred to, e.g., Christensen
(1971), Mainardi (2010), or another intermediate-level text in viscoelasticity.
Viscoelasticity 213

/ 0 / 0

0
TIME

Fig. 7.2 Behavior of a viscoelastic material during stress relaxation. Use of log
linear or loglog axes will often transform the data into presentations that can
be a useful aid when inferring which relaxation kernel to select

after a subduction earthquake (Wang et al. 2012). This experiment has an


idealized strain history described by
1 t  0+ ,
 (t) = 0 H(t), where H(t) = 1/2 t = 0, (7.6)
0 t
0 .

Here H(t) is the Heaviside step function with 0 =  (0 , 0+ ) being an im-


posed jump in strain occurring virtually instantaneously over [0 , 0+ ].
For this experimental BVP, Boltzmanns viscoelastic equation (7.1) in-
tegrates to an expression that allows the functional form of the relaxation
modulus to be determined experimentally, viz.,
 (t)  (t)/0 1
G(t) = or G (t) = , (7.7)
0 0 1
where  (t) is the measured stress response from the experiment. This is the
reason why G is called the relaxation modulus. The measured stress  (t)
is often normalized by the initial stress 0 =  (0+ ) = 0 0 when plotting,
as illustrated in Fig. 7.2, in which case one has
G(t)  (t)
= . (7.8)
G(0) 0
The shape of a relaxation curve is an experimental outcome designating
what a relaxation kernel must be able to correlate.
214 Soft Solids

This analysis applies for the linear theory of viscoelasticity, not to non-
linear theories of viscoelasticity. In the case of nonlinear viscoelasticity,
eects arising from material nonlinearity (e.g., via a variable tangent mod-
ulus) blur with eects arising from the relaxation modulus, and the process
of material characterization becomes more challenging.
The relaxation modulus G is one of several viscoelastic material func-
tions commonly used. It is the only one that will be addressed in this text.
Another is the creep compliance, which is the inverse (in the Laplace trans-
form domain) to the relaxation modulus in linear viscoelasticity (Mainardi
2010; Pipkin 1972). A set of dynamic functions also exists that most visco-
elastic testing apparatuses are designed to measure. These are the storage
and loss moduli and their ratio, the loss tangent, a.k.a. tan (Lakes 1998).

7.1.3 Causal Deformations


All viscoelastic deformations are considered to be causal in this text. Math-
ematically, this means that  (a, 0) = 0 for all a < 0. This is why the lower
limit of integration in our presentation of Boltzmanns integral equation
(7.1) is taken to be 0 instead of 1, as is commonly encountered in the
viscoelastic literature.
Because causality is impossible to verify (it is almost always violated
in that it is almost a certainty that  (a, 0) 0 for some a 0), what
one means by imposing the causal deformation assumption, in a practical
sense, is that the material to be tested is initially unstressed and in a state
that is near to or in static equilibrium at time t = 0. Said dierently, any
memory that a material may have had from prior deformations has been
essentially forgotten by the material before the onset of the experiment.
Rule of Thumb: Causality can be considered to exist whenever  (a, 0)=0
over some recent interval in time, say, 5 . a < 0, where  is a charac-
teristic time for the materialan idea that will be given meaning later in
Sect. 7.2.

7.1.3.1 Preconditioning
The biomechanics literature speaks of preconditioning specimens before
testing them; cf. Fung (1993, p. 270). Sometimes the preconditioning pro-
tocol is reported; sometimes it is not. The intended purpose of precondi-
tioning is to take a sample from its static equilibrium state and place it into a
stable dynamic orbit (a repeatable response) at the onset of the experiment,
Viscoelasticity 215

the idea being that this dynamic state is supposed to better represent the
actual, physiologic, in vivo (i.e., in a living organism) resting state for the
particular tissue of interest. In practice, preconditioning is often more than
a mechanical consideration; it is a biological consideration, too, e.g., as a
mechanical stimulus for osmotic rehydration of an excised tissue back to
its natural physiologic levels before testing.
To describe the response of a preconditioned specimen requires full
knowledge of the deformation history traversed since leaving a causal state
because
Z t
 (t) = G(t t0 ) d (t0 )
Z0 t 0 Z t
= 0
G(t t ) d (t ) +0
G(t t0 ) d (t0 ), (7.9)
0 0 t

preconditioning experiment

where preconditioning occurs over some interval [0, t0 ), with the experi-
ment then proceeding over interval [t0 , t]. Of significance is the fact that the
preconditioned contribution to stress varies with time for all t0 , t0
t0
t,
due to current time t being present in G(t t0 ) with t denoting the time at the
end of the actual experiment, i.e., time t is at the upper limit of integration
in the second integral. Consequently, memory of a preconditioning from 0
to t0 will continue to fade over the lifetime of the experiment, viz., from t0
to t.
Because preconditioned deformation histories are, in general, only
causal before time t = 0 and do not return to a causal state by time t = t0 ,
any viscoelastic modeling of a history that contains a preconditioning can
only hope to be correct if the entire preconditioning portion of the defor-
mation history is taken into account in ones experimental analysis (Freed
and Doehring 2005).

7.1.4 Memory Kernel Formulation


An integration by parts of Eq. (7.1), as specified in Eq. (B.87), allows the
rate in the integrand to be taken from the forcing function, i.e., the strain
rate, and moved to the kernel function, causing Boltzmanns viscoelastic
model to be recast in an alternative form where
216 Soft Solids

Z t
 (t) = 0  (t) M(t t0 )  (t0 ) dt0
0
Z t
 
= 0  (t) 0 1 M(t t0 )  (t0 ) dt0 (7.10)
0
with
dG (t t0 )
M(t) = ( 0 1 )M(t) and M(t t0 ) = , (7.11)
dt0
where the forcing function in this integrand is now strain  , not strain-rate
d /dt. In contrast with Eq. (7.5), where a viscous overstress is added to
an equilibrium or rubbery elastic response, here the integral diminishes the
dynamic or glassy elastic response.
Lodge calls the viscoelastic material function M(t) the memory func-
tion, which has units of stress rate, while M is referred to as the memory
kernel. Unlike the relaxation kernel G , which is dimensionless, the mem-
ory kernel M has units of reciprocal time, i.e., it is a rate. Lodge (1956)
introduced his memory function into a model that he would later call the
rubberlike liquid (Lodge 1964, pp. 101104). Memory functions entered
into the biomechanics literature some 25 years later in a paper by Zhu et al.
(1991). The terminology of a memory function is a synonym for the
rate-of-relaxation function.
Both viscoelastic kernels, i.e., relaxation G (t) and memory M(t), must
monotonically decay to 0 as time t goes from 0 to 1 in order to be thermo-
dynamically viable (Coleman and Mizel 1968). The relaxation kernel G (t)
is bound to the interval [0, 1] for physical reasons, but no such restriction
applies to the memory kernel M(t). In fact, it is common for M(0) = 1
with the implication being that the material has a perfect recollection of
the present. If relaxation kernels G (t) were allowed to become unbounded
at time t = 0, like memory kernels often are, then (predicted) wave fronts
would travel through solids with infinite velocity (Caputo and Mainardi
1971b,a; Mainardi 2010) which is not physical. It is because of this physi-
cal characteristic that the restriction of G (0) = 1 is justified.

7.1.5 Additive Strain Formulation


Strain is a two-state field, and it is important in some viscoelastic formula-
tions that one distinguish what two states a particular strain measure asso-
ciates with. When expressed as  (a, b), the first argument, in this case a,
associates with configuration a at time t = a, while the second argument,
Viscoelasticity 217

taken as b, associates with configuration b at time t = b, such that this


strain is a measure of the stretch that has occurred between configurations
a and b , where a and b are ordered, i.e., 0
a
b, but are otherwise
arbitrary. Whenever strain is written with only a single argument in time,
say  (b), then it is understood that strain is being measured against the
reference state 0 , viz.,  (b)   (0, b).
Some, but not all, strain measures are additive in the sense that
 (0, t) =  (0, t0 ) +  (t0 , t) 8 t0 2 [0, t]. (7.12)
Restricting our analysis to employ only those strain measures that are ad-
ditive allows Eq. (7.10) to be rewritten as
Z t
 (t) = G(t)  (t) + M(t t0 )  (t0 , t) dt0
0

= 1  (t) + ( 0 1 ) G (t)  (t)


Z t
+ ( 0 1 ) M(t t0 )  (t0 , t) dt0 , (7.13)
0
Rt R  
where through the identity 0 M(t t0 ) dt0 = 0t dG (t t0 )/dt0 dt0 =
R G (0)
G (t) dG = G (0) G (t) = 1 G (t), with G(t) = 1 + ( 0 1 )G (t)
and M(t) = ( 0 1 )M(t), one arrives at the above result. This is the first
of two versions of Boltzmanns viscoelastic theory used herein.
The reference configuration for strain in the integrand of Eq. (7.13)
is t0 ; it is the initial configuration of the viscoelastic response. This is
consistent with the physical notion that interval [t0 , t] constitutes that part of
the deformation history which the material recollects, while the preceding
interval [0, t0 ) represents that part of the history which the material has
forgotten. In this regard, Eq. (7.13) is preferred over Eq. (7.10) because it
better reflects the underlying physics. Equation (7.13) is in a format that
is amenable for extending explicit (stress/strain) elasticity into the visco-
elastic domain.
The first term on the right-hand side of the first line in Eq. (7.13) has
two constituents, which are written out in the second line. One accounts
for an equilibrium or elastic contribution to stress, i.e., 1  (t). The other
accounts for a transient or viscous contribution to stress that will fade away
over time, viz., ( 0 1 )G (t)  (t). The combined response from this latter
term, along with the integral term displayed in the third line, describes a
218 Soft Solids

nonequilibrium contribution to the overall state of stress that adjusts itself


over the strain history. In this version of Boltzmanns viscoelastic theory,
like in Eq. (7.5), transient contributions to stress are being added to the
quasi-static or rubbery state of stress associated with equilibrium.
Preconditioning a specimen eectively removes ( 0 1 )G (t)  (t) as
a contributor to the transient response so that Eq. (7.13) reduces to
Z t
 (t)  1  (t) + ( 0 1 ) M(t t0 )  (t0 , t) dt0 . (7.14)
0
This is most readily seen during cyclic loading histories that stabilize to a
repeating or saturated cycle after several transient cycles.
Many numerical algorithms have been specifically designed for ap-
proximating the integration of hereditary integrals like those that arise in
viscoelastic material models with either discrete (Simo and Hughes 1998)
or continuous (Bra 1977; Brunner 2004) relaxation spectra, including a
method developed by Diethelm and Freed (2006) presented in Appendix E.
Our algorithm is very capable in this regard. It can handle weak singular-
ities that often occur at the upper limit of integration, i.e., kernels where
M(0) = 1 are allowed. It can also handle two-state strain measures, as
seen in Eqs. (7.13) and (7.14), in a natural and ecient way. Plus, it has
a p-adic memory management scheme that enhances its performance from
one of O(N 2 ) to O(N log N), where N is the number of integration steps
taken.

7.1.6 Quasi-Linear Viscoelasticity


Up to this point, our 1D viscoelastic formulation has been linear. To de-
scribe the viscoelastic response of soft tissues, whose constitutive response
is nonlinear, Fung (1971) introduced, in eect, a dimensionless tangent
modulus T ( ), i.e., T ( ) = 1 d&( )/d where  = &( ), into Boltzmanns
integral equation (7.1) so that
Z t
 (t) = G(t t0 ) T ( ,  ; t0 ) d (t0 ), (7.15)
0
where here we have introduced a tangent modulus that can depend upon
both stress and strain, viz., T ( ,  ) = 1 @&( ,  )/@ . Fung called
his theory quasi-linear viscoelasticity (QLV), the juxtaposition of Boltz-
manns linear theory of viscoelasticity with a nonlinear theory of elasticity.
Function T = 1 returns Eq. (7.1); function T = T ( ) supplies an explicit
Viscoelasticity 219

tangent modulus, like Fung used in his original theory and like the tangent
moduli derived in Chap. 5; while function T = T ( ,  ) supplies an implicit
tangent modulus of the type introduced in Chap. 6. The tangent modulus
is to be normalized in the sense that T (0, 0) = 1, i.e., Eq. (7.15) reduces to
a linear viscoelastic solid Eq. (7.1) in a neighborhood around the reference
state 0 ; otherwise, it can become nonlinear. The tangent modulus intro-
duced here, i.e., T , diers from the tangent moduli of Chaps. 5 and 6, viz.,
M, in that T = 1 M = 1 @&/@ is taken to be dimensionless.
A diversion from the developmental path traveled by Fung (1971, 1993)
is taken here. A somewhat dierent path is followed leading to a form
similar to Eq. (7.13); specifically, consider a change in variable

d:( ,  ; t) = T ( ,  ; t) d (t) (7.16)

that is a homogeneous function in strain rate that when integrated becomes


Z t
:( ,  ; 0, t) = T ( ,  ; t0 ) d (t0 ), (7.17)
0
where :(0, t)  :( ,  ; 0, t) is a nonlinear measure of strain that asso-
ciates with the beginning and ending configurations 0 and . We call :
a Guth strain after Guth et al. (1946), who used the neo-Hookean strain
of Eq. (5.81) as a nonlinear strain measure in their 1D viscoelastic model.
A Guth strain, as the terminology is used herein, is any stress response
(normalized to become a strain) that arises from an admissible nonlinear
elastic theory that can be put into the form of dS = M : dE .
Because Guth strain is described by an integral equation, it follows
straightaway that it is additive in the sense that :(0, t) = :(0, t0 ) + :(t0 , t) for
all t0 2 [0, t]. Furthermore, according to a theorem of Nolls (1955) recited
on p. 162, Guth strain may depend upon the path traversed between states
0 and , but it will never have a direct dependence upon time. Time is
only the dummy variable of integration.
Consequently, Eqs. (7.15) and (7.16) can be collectively rewritten as
Z t
 (t) = G(t t0 ) d:(t0 ), (7.18)
0
where d:(t0 ) d:( ,  ; t0 ), and
we retain the form of Boltzmanns (1874)
linear viscoelastic hereditary integral, Eq. (7.1), but now with a nonlinear
strain measure :, hence the terminology, QLV.
220 Soft Solids

Following the same line of reasoning that was used to arrive at


Eq. (7.13) from Eq. (7.1), Eq. (7.18) can be transformed into an expression
that apparently originated with Guth et al. (1946) (25 years before Fungs
development); it being
Z t
 (t) = G(t) :(t) + M(t t0 ) :(t0 , t) dt0
0

= 1 :(t) + ( 0 1 ) G (t) :(t)


Z t
+ ( 0 1 ) M(t t0 ) :(t0 , t) dt0 (7.19)
0
wherein :(t0 , t) = :(t) :(t0 ) which, in turn, is coupled with an integral
equation to establish its Guth strain, viz.,
Z t
:(t) = T ( ,  ; t0 ) d (t0 ). (7.20)
0
These two formul comprise the second viscoelastic representation of
Boltzmanns theory used in this text.
QLV, as it is presented here, viz., Eqs. (7.19) and (7.20), is described
by a pair of integral equations. Whenever the tangent modulus depends
upon stress, i.e., whenever T = T ( ), these equations become coupled
and, therefore, some type of implicit or iterative solution strategy will be
required. Otherwise these formul are uncoupled, thereby allowing for a
sequential analysis scheme that would first solve for strain : and then solve
for stress  .

7.2 Viscoelastic Kernels

Many kernel functions have been proposed in the literature. Most have
theoretical underpinnings whose physics lie beyond the scope of this intro-
ductory text. Some of the more important and commonly used kernels are
presented below. A kernel is every bit as much a material characteristic as
are, say, its rubbery and glassy moduli. The ability to select an appropriate
kernel function is an important experimental outcome.
Numerous viscoelastic kernels have been derived from mechanical
models of springs and dashpots, whose discussions pervade the literature.
They are commonly used for instruction in introductory textbooks on the
subject, e.g., Mainardi (2010), Pipkin (1972), Tschoegl (1989), Wineman
and Rajagopal (2000), and Zener (1948). Mechanical models described
Viscoelasticity 221

this way present the student with a double-edged sword. As a positive,


they oer the student a mental image of how dissipation can be mimicked
via a mechanical analog. As a negative, they can easily cloud the physics
that are actually taking place, often becoming an all-consuming exercise
that leads to kernels whose parameters have little or no physical meaning
left to them. Viscoelastic kernels are a direct consequence of molecular/
statistical mechanics, not of springs, dashpots, and the like, i.e., they are
functions derived from modern theoretical physics, which is the approach
taken here for introducing them to the student, but at a higher level, not at
the detailed level of a statistical mechanics analysis.

7.2.1 IOV Kernel


The simplest viscoelastic material model, when written as an ODE, that
can reproduce the basic time-dependent behavior of a solid in a physically
meaningful way was developed by Zener (1948, p. 43) for describing the
physics of a thermal relaxation caused by solute diusion in a thermoelastic
solid [cf. Mainardi (2010, pp. 7172)]. Zener called this model the standard
linear solid. Here, it is called the integer-order viscoelastic (IOV) model,
as Zeners solid can be cast as a first-order dierential equation whose in-
tegral formulation becomes an instance of Boltzmanns (1874) viscoelastic
theory containing a MaxwellDebye relaxation kernel described by a de-
caying exponential (Maxwell 1867), i.e.,
t
G (t) = exp ,  > 0, (7.21)

whose associated memory kernel is another decaying exponential, viz.,
1 t
M(t) = exp , (7.22)
 
where parameter  is called the characteristic time, which has units of time.
It is the time required to relax away about 63 % of the transient contribution
to stress, sometimes referred to as an overstress.
Diusion is the physical cause of viscoelastic eects (Zener 1948). The
IOV relaxation kernel arises when modeling molecular diusion via statis-
tical mechanics where just the simplest set of physical assumptions are
imposed. A Brownian random walker under the control of the diusion
equation, sampled at uniform intervals in time and space, provides the un-
derlying theoretical basis for this kernel (Metzler and Klafter 2002). Such
222 Soft Solids

Markovian physics describe systems that are in thermal equilibrium. This


idea can be traced back to Maxwells [1867] theory for an ideal gas.

7.2.2 FOV Kernel


The kinetics of a fractional-order viscoelastic (FOV) model can be de-
scribed by a specific fractional-order dierential equation (Caputo and
Mainardi 1971b,a; Doehring et al. 2005). The Laplace transform tech-
nique can be applied to solve linear fractional-order dierential equations
(Mainardi 2010), thereby allowing a whole class of viscoelastic material
models to be recast as Volterra (1930) integral equations whose solutions
are otherwise described in terms of fractional-order dierential equations
[cf. e.g., with the review paper of Mainardi and Gorenflo (2007) and the
recent book by Mainardi (2010)]. Consequently, you need not fully appre-
ciate the intricacies of the fractional calculus in order to be able to eec-
tively use viscoelastic theories derived using the fractional calculus. Text-
books on the fractional calculus have been written by Baleanu et al. (2012),
Miller and Ross (1993), Diethelm (2010), Mainardi (2010), Meerschaert
and Sikorskii (2012), Oldham and Spanier (1974), Herrmann (2011), Pod-
lubny (1999), and Samko et al. (1993).
Mainardis FOV kernel has been well studied. It was formally intro-
duced by Caputo and Mainardi (1971a,b) where the first-order derivatives
in Zeners IOV solid were replaced (analytically continued) with deriva-
tives of fractional order of the Caputo (1967) type. Bagley (1991) calls
this the thermorheologically complex material as a metaphoric play on
words regarding Christensens (1971, p. 220) thermorheologically simple
material. In actuality, they both have the same mathematical complexity.
They are both linear Volterra integral equations of the second kind. They
just employ dierent kernels of integration.
Bagley and Torvik (1983) were among the first to link viscoelastic mod-
els based on the fractional calculus with models derived from molecular
physics, where predictions of fractional orders of 1/2 and 2/3 were shown
to associate with the molecular theories of Rouse (1953) and Zimm (1956),
respectively. The RouseZimm theories were derived to describe the eect
that small amounts of a polymer, when suspended in a solvent, can have on
the overall behavior of a fluid [cf. Bird et al. (1987b)].
By applying the Laplace transform, Caputo and Mainardi (1971a,b)
were able to solve their fractional-order dierential equation, acquiring
a Volterra (1930) integral equation of the second kind in the form of a
Viscoelasticity 223

Boltzmann (1874) viscoelastic material whose relaxation kernel is


 
G (t) = E,1 (t/) ,  > 0, 0<
1 (7.23)

with an aliated memory kernel (Freed and Diethelm 2006; Gross 1947)
 
E,0 (t/)
M(t) = , (7.24)
t

where signifies the fractional order governing the diusion kinetics,


which is a dimensionless parameter, while  is its characteristic time,
which has units of time. The parametric constraints in Eq. (7.23) ensure
that the viscoelastic kernels of the FOV solid will decay monotonically
with G (t) 2 [0, 1] such that G (0) = 1 and G (1) = 0.
The Mittag-Leer (1904) function E, (x) is defined by the infinite
series (Mainardi 2010; Podlubny 1999)

X
1
xk
E, (x) = , > 0, (7.25)
 ( + k)
k=0

which reduces to the exponential function whenever = = 1, i.e.,


E1,1 (t/) = et/ because  (1 + k) = k! whenever k = 0, 1, 2, : : : . For
the case of = 1/2 and = 1, the Mittag-Leer function equates with
E 1/2 ,1 ((t/) 1/2 ) = et/ erfc((t/) 1/2 ), where erfc designates the complemen-
tary error function, i.e., erfc(x) = 1 erf(x) within which the error func-
R 2
tion erf is defined via erf(x) = p2 0x et dt. A 1/2 order arises naturally
when describing the temporal contribution to diusion processes (Oldham
and Spanier 1974, Chap. 11). Algorithms for the numerical evaluation of
E, (x) can be found in papers by Gorenflo et al. (2002) and Hilfer and
Seybold (2006), the former being replicated in a review paper by Diethelm
et al. (2005). This algorithm, presented in Appendix F, is simplified for
the case of real-valued arguments with 2 (0, 1] and 2 {0, 1}, which is
sucient for computing G and M in Eqs. (7.23) and (7.24).
Understanding the significance of the Mittag-Leer function is aided
by drawing an analogy with the exponential function. It is well known
that the exponential is the eigenfunction of the dierential operator, i.e.,
given that y(x) = ecx , then dy(x)/dx = cy(x) for all x  0. It turns out that
given y(x) = E,1(cx ), then d y(x)/dx = cy(x) for all > 0 and x  0
224 Soft Solids

(Diethelm 2010) where d/dx denotes a fractional derivative of order in


the sense of Caputo (1967). The exponential function is the characteristic
solution to an ODE of integer order. The Mittag-Leer function is the
characteristic solution to an ODE of fractional order.
The FOV kernel first appeared in a paper written by Gross (1947), a
culmination of his earlier works (Gross 1937, 1938) wherein he derived
(without realizing it) a fractional-order dierential equation for describing
an electrical circuit. His kernel appeared in the form of a memory func-
tion, viz., dE ((t/) )/dt where E (x)  E,1(x) is the one-parameter
Mittag-Leer function.2 The Mittag-Leer function was rigorously ali-
ated with the fractional calculus some years later in a pair of papers written
by Caputo and Mainardi (1971a), Caputo and Mainardi (1971b), with
being the fractional order of a certain class of dierential equations.3
Relaxation functions described in terms of the Mittag-Leer func-
tion arise naturally out of a renewal theory for condensed-matter relax-
ation (Douglas 2000). They also follow from a random walker under
the control of a fractional FokkerPlanck drift/diusion equation, whose
walker spreads less eciently than a Brownian random walker (Metzler
and Klafter 2002) and, therefore, describes diusional systems that are
close to thermal equilibrium. In their study of the fractional diusion equa-
tion, Meerschaert and Sikorskii (2012) found that fractional derivatives in
space model anomalous super-diusion processes, while fractional deriva-
tives in time model anomalous sub-diusion processes. Fractional-order
dierential equations also arise from an analytic description of diusion
processes associated with fractal boundaries (Douglas 2000), whereas in-
teger-order dierential equations arise from diusion processes associated
with smooth boundaries.
Like the exponential function, the Mittag-Leer function, as a visco-
elastic relaxation kernel, has a rigorous connection with theoretical
physics. The exponential function describes diusion processes in systems
that are in thermal equilibrium, while the Mittag-Leer function describes
diusion processes in systems that are close to thermal equilibrium.

2 Gross (1947) did not make use of the two-parameter Mittag-Leer function
E, (x).
3 Advancements were also made to fractional-order viscoelasticity in the Russian
literature around this same time, as noted in the text by Mainardi (2010).
Viscoelasticity 225

0
0.2
0.4
0.6
0.8
1
1
0.8

G (t) 0.6
0.4
0.2
0
0
1
2
t 3

Fig. 7.3 A 3D plot of the FOV relaxation function dened in Eq. (7.23), i.e.,
G (t) = E,1 ((t/) ) with  = 1, and where t 2 [0, 3]

7.2.2.1 Behavior of the FOV Kernels


When compared with the exponential relaxation kernel from IOV (where
= = 1), the FOV relaxation kernel (with 0 <
1, = 1) ex-
hibits a much faster rate of decay for smaller arguments (super-diusion:
its derivative tends to 1 as time moves towards 0 from above, in com-
parison with the IOV kernel where the derivative is 1 at time 0), but with
a much slower rate of decay for larger arguments (sub-diusion: algebraic
decay in comparison with exponential decay). This long-term behavior
gives rise to what has been referred to in the literature as an ultra-slow re-
laxation process (Gorenflo and Rutman 1995). These characteristics can
be visualized in Fig. 7.3, where a surface plot of the FOV relaxation ker-
nel is drawn using calculations acquired from the algorithm presented in
Appendix F.
When the fractional order is very small, but still greater than zero,
the memory kernel M(t) = E,0 ((t/) )/t, plotted in Fig. 7.4 for  = 1,
has a response that behaves like an impulse function. This indicates that
the material has perfect knowledge of its current state. Also, its integrated
response over time, shown in Fig. 7.3, behaves like the Heaviside step func-
tion (7.6) at = 0, whose value is 1/2 . In contrast, especially whenever is
small, the memory kernel has virtually no recollection of even the most re-
cent of past states. The memory function continues to maintain its perfect
226 Soft Solids

0
0.2
0.4
0.6
0.8
1
2

1.5

M (t) 1

0.5

0
0
1
2
t 3

Fig. 7.4 A 3D plot of the FOV memory function dened in Eq. (7.23), i.e.,
M(t) = E,0 ((t/) )/t with  = 1, and where t 2 [0.001, 3]

knowledge of the current state as approaches unity (i.e., M is infinite at


t = 0, except at = 1, with the strength of this singularity diminishing as
! 1). To this complete remembrance of the current state, the function
then adds an increasing recollection of past events with increasing , albeit
this being a memory that fades away rapidly over the passage of time.
A visual inspection of Figs. 7.3 and 7.4 indicates that a significant
numerical advantage should exist when employing the memory kernel
defined in Eq. (7.24) over its corresponding relaxation kernel given in
Eq. (7.23) for the kernel of viscoelastic convolution, provided that the sin-
gularity at the upper limit of integration can be handled eectively and
eciently.
Further inspection of Fig. 7.4 may mislead one to draw a false conclu-
sion that the FOV memory kernel fades faster than the exponential belong-
ing to the IOV memory kernel, which is refuted in Eq. (7.26) below and
is beautifully illustrated in Fig. 7.5. For 2 (0, 1) and an argument t/
that is less than about three, the FOV memory kernel does indeed collapse
faster than exponential decay, except in a neighborhood around the origin.
However, as the argument exceeds approximately three, this trend begins
to reverse itself, and the FOV memory kernel starts to exhibit its true char-
acter of possessing an asymptote that approaches zero algebraically as time
approaches infinity. Memory decay is exponential only when = 1.
Viscoelasticity 227

0
0.2
0.4
0.6
0.8
1
3

2
M(t) / et

0
0
1
2
3
4
t 5

Fig. 7.5 A 3D plot of M(t)/et = E,0 ((t/) )/(t et/ ) with  = 1 and where t 2
[0.001, 5], demonstrating that the FOV memory function in Eq. (7.23) asymptotes
algebraically to zero as t ! 1 for all 2 (0, 1)

The asymptotic behavior of the FOV relaxation kernel Eq. (7.23) is de-


scribed by the formul (Diethelm et al. 2005)


1 (t/)
(1+) / exp (t/) for t ,
G (t) (/t) (7.26)
1 for t  .
 (1) / 1+(t/)

The FOV kernel interpolates between two well-known relaxation kernels


at its extremities, i.e., whenever t  or t  , specifically the KWW
or stretched-exponential kernel (for t ) and the CCM or power-law
kernel (for t  ) (Metzler and Klafter 2002). Both the KWW and CCM
relaxation kernels are widely used in practice and are discussed below.

7.2.2.2 KWW Kernel


The kernel of Kohlrausch (1847) and Williams and Watts (1970) (KWW)
considers a stretched exponential for its reduced relaxation function, viz.,
 
G (t) = exp (t/) ,  > 0, 0 <
1, (7.27)

whose aliated memory kernel is


t exp(t/) 
M(t) = , (7.28)
 t
228 Soft Solids

where  and are its material constants. The constraints imposed on


these parameters ensure that these viscoelastic kernels monotonically de-
cay. Like the FOV kernel, the KWW kernel reduces to the IOV kernel at its
boundary of = 1. It also has been derived from statistical physics when
describing condensed-matter relaxation (Douglas 2000).

7.2.2.3 CCM and MPL Kernels


The ColeCole (1941; 1942) model (CCM) was introduced to describe
the dispersion and absorption properties of dielectrics. It has a relaxation
kernel of
1
G (t) = ,  > 0, > 0, (7.29)
1 + (t/)
whose aliated memory kernel is
(t/)
M(t) =  2 , (7.30)
t 1 + (t/)

where  and are its material parameters. This variant of a power-law


model is used herein.4
An alternative form for a power-law model is the modified power law
(MPL) of Williams (1964), who took the power-law creep compliance of
Nutting (1921) and expressed it as a normalized relaxation function so that
G (0) = 1. It has a relaxation function of
1
G (t) = ,  > 0, > 0,
(1 + t/)

which is distinct from the ColeCole kernel above. Williams considered


his power law to be a broadband approximation to the IOV kernel. This
kernel first appeared in the Russian literature in a 1937 paper by Kobeko,
Kuvshinskij, and Gurevitch.5 The CMM and MPL kernels have the same
asymptotic response for t  .

4 The fractional model of Scott Blair (1944), where G(t) / t , is similar to the
CCM model; CCM is a regularization of the Scott Blair model. The Scott Blair
kernel is not normalized in that G (0) = 1 and, hence, his kernel propagates waves
with innite velocity.
5 This citation came from Tschoegl (1989, p. 320). The author has been unable
to secure this Russian document to corroborate his citation.
Viscoelasticity 229

Bagley (1987) has shown that Caputos 1967 fractional Voigt model for
viscoelasticity and the power-law kernels have the same response at large
times, while Freed and Diethelm (2006) suggest that power-law kernels
are a kind of regularization of the Abel kernel that appears in the fractional
dierential operator studied by Caputo (1967).

7.2.3 BOX Kernel


This relaxation model, developed by Neubert (1963) and made popular
by Fung (1971), has become the de facto standard for characterizing soft-
tissue viscoelasticity in the biomechanics literature of today, where it is
often referred to as the QLV kernel. It is described by the relaxation kernel6

E1 (t/2 ) E1 (t/1 )
G (t) =
ln(2 /1 )
Z 2 t/
1 e
= d, 2 > 1 > 0, (7.31)
ln(2 /1 ) 1 
whose aliated memory kernel is
exp(t/2 ) exp(t/1 )
M(t) = . (7.32)
t ln(2 /1 )
These viscoelastic functions have two characteristic times, 1 and 2 ,
whose stated constraint in Eq. (7.31) ensures a fading memory. Function
Z 1 
e
E1 (t) = d, t > 0 (7.33)
t 
is one of several equivalent definitions for the exponential integral that can
be found in the literature (Abramowitz and Stegun 1964; Mainardi 2010).
The asymptotic behavior of the BOX relaxation kernel Eq. (7.31) is de-
scribed by (Neubert 1963):
6 To compute the BOX relaxation kernel in Eq. (7.31), it is best to numerically
R
solve the integral 12 et/ / d, because taking the dierence between the two
exponential integrals E1 (t/2 ) E1 (t/1 ) will lead to measurable numerical error
whenever t is in the vicinity of either 1 or 2 given that standard algorithms are
used to solve the exponential integral, as presented in, e.g., Press et al. (2007).
The BOX relaxation kernel, as written down in Eq. (7.31), does not usually
appear in the QLV literature, e.g., Fung (1971, Sect. 7.6). Rather, a parameter
c > 0 appears that relates the glassy modulus 0 to the rubbery modulus 1 via
1 = 0 /(1 + c ln(2 /1 )) where c represents the height of a rectangular relaxation
spectrum that begins at 1 and ends at 2 .
230 Soft Solids

( t/1 t/2
1 ln( for t ,
G (t) 2 /1 ) (7.34)
2 exp(t/2 )1 exp(t/1 )
t ln(2 /1 ) for t  .
When compared with the exponential relaxation kernel of the IOV,
Eq. (7.21), the BOX relaxation kernel exhibits a flexible initial rate of
decay, i.e., the BOX kernel has a derivative at t = 0 that is given by
(2 1 )/1 2 ln(2 /1 ), while the derivative of the IOV kernel is 1 at
t = 0. However, at large times, the BOX and IOV kernels exhibit similar
behaviors.
This kernel was derived from the phenomenological consideration
of a constant relaxation spectrum defined over a finite interval in time/
frequency, viz., a box (hence its name). This kernel has no contribution
arising from the spectrum for times less than 1 nor for times greater than
2 while, over the interval [1 , 2 ], the relaxation spectrum is held constant,
normalized by ln(2 /1 ) so that G (0) = 1 (Neubert 1963; Tschoegl 1989).
The author is not aware of any theory from statistical mechanics
whereby this model has been shown to be derived from a more fundamen-
tal analysis of the physics, i.e., it is phenomenological; nevertheless, it is
very popular.

7.2.4 Implementing a Physical Kernel: The MCM Kernel


The
R t level0 of 0numerical eort required to integrate a convolution integral
0 2
0 k(t t ) f (t ) dt is on the order of O(N ) where k is the kernel, f is the
forcing function, and N is the number of integration steps used to parti-
tion the interval of integration [0, t]. The algorithm of Diethelm and Freed
(2006), presented in Appendix E, significantly reduces this level of work
to O(N log N). It even applies to weakly singular kernels with continuous
relaxation spectra.
Whenever the kernel of convolution associates with a discrete spectrum,
i.e., it is constructed from decaying exponentials, an integration algorithm
can be constructed whose level of eort can be reduced down to O(nN)
where n is the number of exponentials in a Prony series used to approx-
imate the kernel (Simo and Hughes 1998, pp. 353355). Using a Prony
series representation for a physical kernel makes sense, from a computa-
tional perspective, only when log N  n. Such an approach is discussed
below, and is widely used in CFD and FE implementations of viscoelastic
models.
Viscoelasticity 231

A materials characterization and implementation will often employ


two dierent viscoelastic kernels. In practice, the Maxwell chain model
(MCM) kernel introduced below is used as an approximation function
for a physically motivated kernel selected for material characterization.
Common practice in computer science is to use approximation functions
to eciently and accurately compute estimates for mathematical functions
(Hart et al. 1968). Here a Prony series is used as the approximating func-
tion. As in approximation theory, coecients in an approximating function
cannot be uniquely determined for any given function beyond the most triv-
ial cases. This is most certainly true here, too. Consequently, parameters
in a Prony series lack physical interpretation.
The viscoelastic kernels just discussed have all been physically moti-
vated to one extent or another. The MCM, often referred to as the gener-
alized Maxwell model, is not so much a physical kernel as it is a useful
kernel, although some would argue otherwise. The MCM is composed of a
finite sum of K discrete Maxwell elements, i.e., the viscoelastic relaxation
and memory kernels are described by Prony series

X
K   X
K  
t ck t
G (t) = ck exp and M(t) = exp ,
k  k
k=1 k=1 k
X
K
ck = 1, ck > 0 8 k, 0 < 1 < 2 <    < K , (7.35)
k=1
where each term in each sum can be thought of as being associated with
a discrete integral or a separate internal-state variable (Simo and Hughes
1998, Chap. 10), each obeying the physics of a Maxwell relation. The sum
of all ck s equaling 1 ensures that G (0) = 1, while G (1) = 0 follows if
k > 0 for all k, and a monotonic decay will happen if ck > 0 for all k. The
MCM kernel obeys the principle of fading memory under these conditions.
The sheer number of material parameters that this kernel can employ
makes it a nearly impossible task to try to gain any physical insight into a
material via these parameters. This is why this author does not advocate
selecting the MCM kernel for the purpose of material characterization. It
is, however, a reasonable kernel to select for numerical implementation of
a viscoelastic model into a computationally intensive software application
like CFD or FE codes, for reasons that are explained below. For example,
Puso and Weiss (1998) adopted this approach to approximate the contin-
232 Soft Solids

uous spectrum of a BOX kernel with the discrete spectrum of a Prony se-
ries over its range [1 , 2 ] to implement their model for tendon into an FE
code.7
It is the recursion property of the exponential function, viz., et+t =
t t
e e , that makes exponential kernels so useful and powerful in applica-
tions, i.e., that provides a mechanism whereby an O(nN) algorithm can
be constructed. In particular, whenever a Prony series, or MCM solid, is
used as the kernel in a Volterra integral equation, that kernel will allow the
integral to be decomposed according to the scheme
Z t+t
0
et+tt f (t0 ) dt0
0
Z t Z t+t
0 0
= et+tt f (t0 ) dt0 + et+tt f (t0 ) dt0
0 t
Z t Z t+t (7.36)
t tt0 0 0 t+tt0 0 0
=e e f (t ) dt + e f (t ) dt
0 t
R 0
where 0t ett f (t0 ) dt0 is known from the previous integration step, it being
adjusted in the current step by a simple scale factor of et . An integration
R 0
of tt+t et+tt f (t0 ) dt0 is all that needs to be solved, and it is over an
interval of incremental length t so it can be evaluated quite economically
by standard numerical techniques. The IOV and MCM kernels, and the
BOX memory kernel, can exploit this recursive property to their advantage,
whereas the advantages of the CCM, FOV, KWW, and MPL kernels lie in
the physical insight that one can acquire by contrasting values for their
parameters across a class of materials.
A relaxation kernel is completely monotonic if (1)k dk G (t)/dtk  0
for all k = 0, 1, 2, : : : and, therefore, possesses a continuous relaxation
spectrum H (t) defined as a forcing function in the convolution integral
R 0
G (t) = 01 et/t H (t0 ) dt0 . The BOX, FOV, KWW, and the power-law ker-
nels all have relaxation functions that are completely monotonic. Because
they possess this property, the continuous relaxation spectrum of each can
be approximated by a sum of discrete relaxation spectra over any range in
time or frequency yielding nonunique MCM kernels.

7 The memory function of the BOX kernel is the dierence between two expo-
nentials; cf. Eq. (7.32). Approximating a BOX kernel with a Prony series for
integrating its viscoelastic convolution integral is not necessary provided that the
viscoelastic model is built to accept a memory function, as in Eqs. (7.13) or (7.19).
Viscoelasticity 233

Rule of Thumb: One to two Maxwell chains are needed for each decade
in time or frequency response that separates the glassy 0 and rubbery 1
plateaus with characteristic times k evenly spaced in logarithmic time over
the frequency range separating the rubbery and glassy plateaus.
For synthetic polymers, the number of Maxwell chains typically
exceeds seven.8 For biologic tissues, the number of terms in a Prony series
is usually around three or four, e.g., Miller-Young et al. (2002) and Puso
and Weiss (1998). There are two parameters per link of chain, i.e., per
exponential term in the Prony series. Yikes! The MCM kernel can easily
become a parameter estimation nightmare.
Tschoegl (1989, Sect. 3.6.2) describes a collocation method for fitting
the parameters of an MCM kernel, but it can lead to nonphysical, negative,
spectral lines.
Fulchiron et al. (1993) and Simhambhatla and Leonov (1993) propose
using an automated PadLaplace technique to obtain optimum MCM pa-
rameters. Here, a Pad expansion of chosen order is used to fit the data
(in our case, a characterized physical viscoelastic kernel) in the Laplace
domain where the problem becomes well posed. The results are then trans-
formed back into the time domain for use.
Another application that uses the Laplace transform was developed by
Park and Schapery (1999) and Schapery and Park (1999) to map the Prony
parameters between the various viscoelastic functions that exist in the lit-
erature, e.g., between the relaxation modulus and the creep compliance.
Another scheme has been proposed by Stuebner and Haider (2010),
where Gauss-Lagrange quadrature was applied to a relaxation kernel G (t),
in their case the BOX kernel, to quantify an MCM kernel using integration
nodes that are logarithmically distributed over time.
This authors opinion is that, whenever possible, the numerical inte-
gration of a convolution integral whose kernel derives from a continuous
relaxation spectrum, as do the BOX, CCM, FOV, KWW, and MPL ker-
nels, ought to be done with an algorithm that systematically samples its
history over the whole of its integration, as is the case with the algorithm
presented in Appendix E. Only in applications where the computational
eort will be excessive, e.g., in CFD and FEA codes where log(N)  n
(with N being the number of integration steps, while n is the number of
8 Forexample, Park and Schapery (1999) use ten Prony elements for polyisobuty-
lene and eleven for polymethyl methacrylate. The author has even witnessed the
use of a Prony series with eighteen elements (thirty-six parameters) in a confer-
ence presentationyou can t an elephant with thirty parameters (Wei 1975).
234 Soft Solids

terms in a Prony series) does it make sense to approximate a physical


kernel with a nonphysical MCM kernel for use in an integration scheme
based on Eq. (7.36). This urgency is further exasperated by the fact that an
integration must take place at each Gauss point within a computational
mesh which, at present day, often number in the tens to hundreds of mil-
lions per geometric model.
One must also bear in mind that Prony approximations have the poten-
tial to produce erroneous results whenever an approximation is inaccurate
or whenever the domain of application travels outside the range where the
approximating MCM kernel was fit.

7.3 Additive Strain Fields

The derivation of the 1D viscoelastic models in Eqs. (7.13) and (7.19),


which will be extended into 3D viscoelastic models in Sects. 7.4 and 7.5,
requires that the strain measure be additive. Not all finite-strain measures
are additive; in fact, many are not. Fortunately, the four predominant strain
measures used in this book can all be generalized so that they become
additive.
Strains, as defined in Chap. 3, are measures of stretch between two con-
figurations, typically between 0 and . The Lagrangian strain measures
E and E of Green (1841) and Lodge (1964) employ 0 as their refer-
ence state, while the Eulerian strain measures e and e of Almansi (1911)
and Signorini (1930) employ  as their reference state. Strains e and
E map between themselves as covariant fields, while strains e and E map
between themselves as contravariant fields.
A generalized version of strain follows. A third intermediate configu-
ration, denoted as 0 and aliated with time t0 , 0
t0
t, is incorporated
in a manner so that strains become additive; specifically, it is desired that
9
E (0, t) = E (0, t0 ) + E (t0 , t)>
>
>
E(0, t) = E(0, t0 ) + E(t0 , t) =
8 t0 , 0
t0
t. (7.37)
e(0, t) = e(0, t0 ) + e(t0 , t) >>
>
;
e(0, t) = e(0, t0 ) + e(t0 , t)
The left-hand sides in these formul are known. They are defined in
Eqs. (3.10), (3.11), (3.13), and (3.14). The strain measures on the right-
hand side of these formul are constructed below.
Viscoelasticity 235

7.3.1 Lagrangian Strains


The strain fields defined over 0 in Eqs. (3.11) and (3.14) can be con-
structed in such a way so as to become additive whenever Green strain
(3.11) is reinterpreted as
 
E (0, t) = 12 C (0, t) C (0, 0)
 
= 12 F T(0, t) F(0, t) F T(0, 0) F(0, 0) ,
 
E (0, t0 ) = 12 C (0, t0 ) C (0, 0)
 
= 12 F T(0, t0 ) F(0, t0 ) F T(0, 0) F(0, 0) ,
 
E (t0 , t) = 12 C (0, t) C (0, t0 )
 
= 12 F T(0, t) F(0, t) F T(0, t0 ) F(0, t0 ) ,
so that E (0, t) = E (0, t0 ) + E (t0 , t) (7.38)
and whenever Lodge strain (3.14) is reinterpreted as
 
E(0, t) = 12 C 1 (0, 0) C 1 (0, t)
 
= 12 F 1 (0, 0) F T(0, 0) F 1 (0, t) F T(0, t) ,
 
E(0, t0 ) = 12 C 1 (0, 0) C 1 (0, t0 )
 
= 12 F 1 (0, 0) F T(0, 0) F 1 (0, t0 ) F T (0, t0 ) ,
 
E(t0 , t) = 12 C 1 (0, t0 ) C 1 (0, t)
 
= 12 F 1 (0, t0 ) F T(0, t0 ) F 1 (0, t) F T(0, t) ,
so that E(0, t) = E(0, t0 ) + E(t0 , t), (7.39)
where F(0, 0) = I by its definition Eq. (2.2), and given that 0
t0
t.
Notice that the first temporal argument in either C or C 1 belongs to time
0, which follows because all of these strain measures are defined over the
initial configuration 0 , i.e., they are Lagrangian fields.
Contained within these definitions is a generalized deformation gradi-
ent that is defined by
@(X , b) @(X , a)
F(a, b) = with F 1 (a, b) = , (7.40)
@(X , a) @(X , b)
where
@(X , c)
F(c, c) = F 1 (c, c) = = I, (7.41)
@(X , c)
236 Soft Solids

so that from the chain rule

F(0, t) = F(t0 , t) F(0, t0 ) (7.42)

wherein  is the motion map established in Eq. (1.1) with 0


a
b
t
where instants a and b associate with configurations a and b . Instead
of the notation F(a, b) where a
b, Noll (1958) introduced the nota-
tion Fb (a), which he defined via F(a) = Fb (a) F(b) so that his Fb (a) is
equivalent to our F 1 (a, b). The author finds the notation introduced in
Eq. (7.40) to be more intuitive then that of Nolls. Nolls notation, how-
ever, has widespread application throughout the rheology community.
The Lagrangian deformation tensors C and C 1 of Green (1841) and
Cauchy (1827) generalize as
C (0, b) = F T(0, b) F(0, b),
, (7.43)
C 1 (0, b) = F 1 (0, b) F T(0, b)
where the first temporal argument in the argument lists on the right-hand
sides of Eqs. (7.38) and (7.39) aliates with the initial time of t = 0. This
is because these deformation measures are Lagrangian and, as such, are
defined over an initial configuration 0 .

7.3.2 Eulerian Strains


The strain fields defined over  in Eqs. (3.10) and (3.13) can be constructed
in such a way so as to become additive, too; specifically, whenever Almansi
strain (3.10) is reinterpreted as
 
e(0, t) = 12 b1 (t, t) b1 (0, t)
 
= 12 F T(t, t) F 1 (t, t) F T(0, t) F 1 (0, t) ,
 
e(0, t0 ) = 12 b1 (t0 , t) b1 (0, t)
 
= 12 F T(t0 , t) F 1 (t0 , t) F T(0, t) F 1 (0, t) ,
 
e(t0 , t) = 12 b1 (t, t) b1 (t0 , t)
 
= 12 F T(t, t) F 1 (t, t) F T(t0 , t) F 1 (t0 , t) ,
so that e(0, t) = e(0, t0 ) + e(t0 , t) (7.44)
Viscoelasticity 237

and whenever Signorini strain (3.13) is reinterpreted as


 
e(0, t) = 12 b(0, t) b(t, t)
 
= 12 F(0, t) F T(0, t) F(t, t) F T(t, t) ,
 
e(0, t0 ) = 12 b(0, t) b(t0 , t)
 
= 12 F(0, t) F T(0, t) F(t0 , t) F T(t0 , t) ,
 
e(t0 , t) = 12 b(t0 , t) b(t, t)
 
= 12 F(t0 , t) F T(t0 , t) F(t, t) F T(t, t) ,
so that e(0, t) = e(0, t0 ) + e(t0 , t), (7.45)
then these strains become additive, too, given 0
t0
t. From the defini-
tion of the generalized deformation gradient found in Eq. (7.40), it follows
that b(t, t) = I because F(t, t) = I.
The Eulerian deformation tensors b and b1 of Finger (1894) and Piola
(1833) generalize as
b(a, t) = F(a, t) F T(a, t),
(7.46)
b1 (a, t) = F T(a, t) F 1 (a, t),
where the second temporal argument is at current time t in the right-hand
side arguments of Eqs. (7.44) and (7.45), which follows because these de-
formation measures are Eulerian and, therefore, are defined over the cur-
rent configuration .

7.3.3 Field Transfer


The covariant strain fields defined in Eqs. (7.38) and (7.44) obey the map-
pings
[E (0, t)] = [F(0, t)]T [e(0, t)] [F(0, t)],
[E (0, t0 )] = [F(0, t)]T [e(0, t0 )] [F(0, t)], (7.47)
[E (t0 , t)] = [F(0, t)]T [e(t0 , t)] [F(0, t)]
or, reversing these maps, they obey
[e(0, t)] = [F 1 (0, t)]T [E (0, t)] [F 1 (0, t)],
[e(0, t0 )] = [F 1 (0, t)]T [E (0, t0 )] [F 1 (0, t)], (7.48)
[e(t0 , t)] = [F 1 (0, t)]T [E (t0 , t)] [F 1 (0, t)],
238 Soft Solids

where the Eulerian fields found in Eq. (7.44) pull back from  into 0 ac-
cording to the covariant map Eq. (B.20), thereby producing the Lagrangian
fields found in Eq. (7.38) or vice versa via the map Eq. (B.19).
Similarly, the contravariant strain fields defined in Eqs. (7.39) and
(7.45) obey the mappings

[E(0, t)] = [F 1 (0, t)] [e(0, t)] [F 1 (0, t)]T ,


[E(0, t0 )] = [F 1 (0, t)] [e(0, t0 )] [F 1 (0, t)]T , (7.49)
[E(t0 , t)] = [F 1 (0, t)] [e(t0 , t)] [F 1 (0, t)]T
or, reversing these maps, they obey

[e(0, t)] = [F(0, t)] [E(0, t)] [F(0, t)]T ,


[e(0, t0 )] = [F(0, t)] [E(0, t0 )] [F(0, t)]T , (7.50)
0 0 T
[e(t , t)] = [F(0, t)] [E(t , t)] [F(0, t)] ,
where the Eulerian fields found in Eq. (7.45) pull back from  into 0
according to the contravariant map Eq. (B.16), thereby producing the La-
grangian fields found in Eq. (7.39) or, vice versa, via the map Eq. (B.15).

7.4 K-BKZ Viscoelasticity

The conjecture of Kaye (1962) and Bernstein, Kearsley and Zapas (1963)
(K-BKZ) is adopted here as a means for analytically extending the explicit
theory of finite elasticity developed in Chap. 5 into a viscoelastic theory.
Other techniques also exist, e.g., the nonlinear theory of Pipkin and Rogers
(1968). The K-BKZ technique takes the 1D formulation of Eq. (7.13)
and extrapolates it into a 3D construction that can be used to model soft
isotropic materials. Their theory was created for viscoelastic fluids, but
here, their hypothesis is applied to solids. The conjecture put forward by
Bernstein et al. (1963) states:

Conjecture 7.1. For the Coleman and Noll (1964) fluid, the stress at time
t depends upon the history of the relative deformation between the config-
uration at time t and all configurations at times prior to t. To this idea we
add the following notions: (1) The eect of the configuration at time t0 < t
on the stress at time t is equivalent to the eect of stored elastic energy
with the configuration at time t0 as the preferred configuration. The eect
Viscoelasticity 239

depends on t t0 , the amount of time elapsed between time t0 and time t.


(2) The stress at time t is the sum (integral) of all the contributions from all
t0 < t. : : :
In eect, we are taking the concept of a strain-energy function associ-
ated with the theory of finite elastic deformations, which is formulated in
terms of a preferred configuration, and incorporating it in a fluid theory
of the Coleman-Noll type by treating all past configurations as preferred
configurations.
This hypothesis takes the potential structure of an explicit elastic solid,
arising from thermostatics, and analytically continues it into neighboring
states of thermodynamic irreversibility where viscoelastic phenomena ex-
ist. The thermodynamic admissibility of this hypothesis was discussed in a
follow-up paper by Bernstein et al. (1964), with an alternative formulation
being provided by Rao and Rajagopal (2007) in terms of multiple natural
configurations.
Point 1 in their conjecture states that configuration 0 , aliated with
time t0 , is to be the viscoelastic reference state from which strain is mea-
sured. Point 2 implies that its eect, when summed over all past configura-
tions via a hereditary integral, as implicated by Point 1, describes the state
of stress. Consequently, Eq. (7.13) is expressed precisely in the form of the
K-BKZ hypothesis, viz.,
Z t
 (t) = G(t)  (t) + M(t t0 )  (t0 , t) dt0 , (7.13)
0
as it applies to the 1D case. The derivation of this formula from Boltz-
manns viscoelastic theory Eq. (7.1) requires that  (t0 , t) =  (t)  (t0 ), i.e.,
that the strain measure be additive.
The K-BKZ hypothesis was created for viscoelastic fluids, in which
case G(t) = 0 G (t) and M(t) = 0 M(t). Here, it is applied to solids where
G(t) and M(t) are given by Eqs. (7.2) and (7.11).

7.4.1 Viscoelastic Lodge Solid


With the prior definitions for additive strain measures in hand, one can in-
corporate the 3D isotropic solid of Lodge stated in Eq. (5.8) into the 1D
viscoelastic structure found in Eq. (7.13). In accordance with the guide-
lines put forth in the K-BKZ hypothesis, one can write down, straightaway,
an analytic continuation of this explicit elastic solid, it being
240 Soft Solids

 
0 1 dW (E; t) 1
S (t) = 1 + G (t) C 1 (t) C (t)
1 dE
Z
0 1 t dW (E; t0 , t) 1 0
+ M(t t0 ) C 1 (t0 , t) C (t , t) dt0 , (7.51)
1 0 dE
which describes the K-BKZ viscoelastic Lodge solid expressed in terms
of Lagrangian fields. A subtle restriction arising from the derivation of
Eq. (7.13) is that the resulting forcing function within the integrand must
be an additive strain field; therefore, only choices for W (E) that lead to
additive strain measures are admissible for analytic continuation via the
K-BKZ hypothesis.
For incompressible materials, S is replaced by .
This viscoelastic material model reduces to Lodge elasticity in the
quasi-static limit. This fact illustrates a fundamental premise of the K-BKZ
hypothesis: The tensorial structure of a viscoelastic material is the tenso-
rial structure of its limiting elastic response. This provides a great simplifi-
cation to the overall process of constructing a viscoelastic material model.
This material model has two elastic parameters: the glassy 0 and rub-
bery 1 shear moduli, plus whatever parameters are required to quantify
the reduced relaxation G and memory M kernels and whatever parame-
ters are needed to describe a strain-energy function W . Because the elastic
strain energy has dimensions of stress, its gradients have been normalized
by the rubbery shear modulus in the above expression.
Any Lodge elastic solid can be analytically continued into an admissi-
ble viscoelastic solid via Eq. (7.51), provided that its resulting strain field
is additive, which is actually a very strong restriction. For example, the
first-order, incompressible, Lodge, elastic solid of Eq. (5.32) becomes
 
(t) = 2 1 + ( 0 1 ) G (t) E(t)
Z t
+ 2( 0 1 ) M(t t0 ) E(t0 , t) dt0 (7.52)
0
or, equivalently, when pushed forward into the Eulerian frame according to
Eqs. (3.17), (4.27), (7.50), and (B.15), it becomes
 
(t) = 2 1 + ( 0 1 ) G (t) e(t)
Z t
+ 2( 0 1 ) M(t t0 ) e(t0 , t) dt0 . (7.53)
0
There are representations of the K-BKZ viscoelastic neo-Hookean solid
written in terms of Lagrangian or Eulerian fields, respectively. These are
Viscoelasticity 241

admissible K-BKZ models in the sense that the Lodge E and Signorini e
strains are additive; cf. Eqs. (7.39) and (7.45).
In the limiting case of a fluid, which is obtained by setting 1 = 0,
Eq. (7.52) reduces to the Lodge (1956, 1958) rubberlike liquid that, re-
markably,
has been derived from two dierent molecular theories: the bead-
-spring theory of Rouse and Zimm for very dilute solutions of de-
formable long molecules in an incompressible Newtonian solvent: : :,
and the network theory of Green and Tobolsky, Yamamoto, and Lodge
which is developed for concentrated polymer solutions and undiluted
or molten polymers. (Lodge et al. 1978)

So, like the neo-Hookean elastic solid, the neo-Hookean viscoelastic solid
has a sound foundation in theoretical physics.

7.4.2 Viscoelastic Green Solid


In like manner, the Green elastic solid of Eq. (5.3) can also be analytically
continued into the viscoelastic domain via the K-BKZ hypothesis; how-
ever, the constraint of additivity of strain requires rewriting S = dW /E as
CSC = C (dW /dE ) C thereby producing
 
0 1 dW (E ; t)
C (t) S (t) C (t) = 1 + G (t) C (t) C (t)
1 dE
Z
0 1 t dW (E ; t0 , t)
+ M(t t0 ) C (t0 , t) C (t0 , t) dt0 , (7.54)
1 0 dE
which describes the K-BKZ viscoelastic Green solid expressed in terms of
Lagrangian fields.
Any Green elastic solid can be analytically continued into an admissible
viscoelastic solid via Eq. (7.54) provided that its resulting strain measure
is additive, which is why Eq. (7.54) was recast as a covariant equation. For
example, the first-order, incompressible, Green, elastic solid of Eq. (5.29),
when written as a covariant equation, becomes
 
C (t) (t) C (t) = 2 1 + ( 0 1 ) G (t) E (t)
Z t
+ 2( 0 1 ) M(t t0 ) E (t0 , t) dt0 (7.55)
0
242 Soft Solids

that, when pushed forward into the Eulerian frame in accordance with
Eqs. (3.1), (3.16), (4.27), (7.48), and (B.19), becomes
 
(t) = 2 1 + ( 0 1 ) G(t) e(t)
Z t
+ ( 0 1 ) M(t t0 ) e(t0 , t) dt0 . (7.56)
0
These are admissible viscoelastic material models in the sense of the
K-BKZ hypothesis, because the Green E and Almansi e strain fields are
additive; cf. Eqs. (7.38) and (7.44).

7.4.3 Viscoelastic MooneyRivlin Solid


Like the MooneyRivlin elastic solid of Eq. (5.37), which is a mixture of
the first-order Green and Lodge elastic solids of Eqs. (5.29) and (5.32) with
a mixing strength of , 0

1, one can also construct a mixture of
their viscoelastic formulations (7.52) and (7.55) by writing
 Z t 
0 0 0
(t) = 2(1 ) G(t) E(t) + M(t t ) E(t , t) dt
0

+ 2 C 1 (t) G(t) E (t)
Z t 
+ 0 0 0
M(t t ) E (t , t) dt C 1 (t) (7.57)
0
that when pushed forward into the Eulerian frame in accordance with
Eqs. (3.4), (3.16), (3.17), (4.27), (7.48), (7.50), (B.15), and (B.19) becomes
 Z t 
0 0 0
(t) = 2(1 ) G(t) e(t) + M(t t ) e(t , t) dt
0
 Z t 
+ 2 G(t) e(t) + M(t t0 ) e(t0 , t) dt0 . (7.58)
0

These represent the K-BKZ viscoelastic MooneyRivlin solid. This model


has three material parameters: the glassy 0 and rubbery 1 shear moduli
and the scaling parameter , 0

1, plus whatever parameters are
needed to characterize the viscoelastic functions G and M.
While Green strain E is additive, strain C 1 EC 1 = 12 (C 1 C 2 ) is
not. This is why the pre- and post-multipliers of C 1 are pulled out in front
and pushed behind the covariant terms that are being summed in Eq. (7.57).
Viscoelasticity 243

This is consistent with Eq. (7.55) and the fact that C 1 is used as a met-
ric of deformation in 0 in this context, which is required to convert the
covariant strain fields that lie between these two contravariant metrics into
its respective contravariant field so that it becomes compatible with stress,
which maps contravariantly. All sense of tensorial structure gets lost in its
Eulerian formulation, Eq. (7.58).

7.5 Quasi-Linear Viscoelasticity

One is pressed to construct a viscoelastic model from the K-BKZ approach


outlined in the prior section for those cases where the limiting elastic re-
sponse is described by a nonlinear strain measure, because such strain mea-
sures are, typically, not additive. Constructing a scheme to overcome this
limitation is the primary focus of this section.
Up to this point, attention has been focused on four topics: a 1D
overview of Boltzmanns (1874) viscoelastic theory, the kernel functions
that arise in viscoelasticity, how to construct suitable strain measures for
use in viscoelastic models, and how to analytically continue explicit elastic
models into viscoelastic models. A process is proposed below whereby
any elastic model characterized by a tangent modulus whose dierential
equation is homogeneous in strain rate, which includes all of the models in
Chaps. 5 and 6, can be analytically continued into a theoretically admissi-
ble viscoelastic model via the K-BKZ conjecture.
The construction of Eq. (7.19) puts elastic material models that admit
a tangent modulus into a mathematical framework whereby the K-BKZ
hypothesis can be adopted. Recall that in the 1D case
Z t
 (t) = G(t) :(t) + M(t t0 ) :(t0 , t) dt0 , (7.19)
0
where
Z t
1 @&( ,  )
:(t) = T ( ,  ; t0 ) d (t0 ) with T ( ,  ) = (7.20)
0 1 @
and, because
Z t
:(0, t) = T ( ,  ; t00 ) d (t00 )
0
Z t0 Z t
(7.59)
= T ( ,  ; t00 ) d (t00 ) + T ( ,  ; t00 ) d (t00 )
0 t0
0 0
= :(0, t ) + :(t , t),
244 Soft Solids

the constraint of additivity of strain becomes a consequence of integration,


i.e., :(t0 , t) = :(t) :(t0 ), assuming that your particular integrand for T is
integrable in strain  . This scheme establishes a strain measure in terms
of an arbitrary tangent modulus T whose functional form derives from an
elastic constitutive assumption. Guth et al. (1946) implemented this 1D
model using the neo-Hookean strain of uniaxial extension, viz., 11 (
2 ), for : in Eq. (7.19).
Adaptation of the K-BKZ conjecture allows the above 1D formulation
to be extended into a 3D formulation straightaway, resulting in
Z t
S (t) = 2G(t) Z (t) + 2 M(t t0 ) Z (t0 , t) dt0
0
 
= 2 1 + ( 0 1 ) G (t) Z (t)
Z t
+ 2( 0 1 ) M(t t0 ) Z (t0 , t) dt0 , (7.60)
0

where Z (t0 , t) = Z (t) Z (t0 ) for all t0 such that 0


t0
t. Equivalently,
when pushed forward into the Eulerian frame according to Eqs. (4.18),
(B.15), and (B.83), it becomes
Z t
s(t) = 2G(t) z(t) + 2 M(t t0 ) z(t0 , t) dt0
0
 
= 2 1 + 2( 0 1 ) G (t) z(t)
Z t
+ 2( 0 1 ) M(t t0 ) z(t0 , t) dt0 , (7.61)
0

where z(t0 , t) = z(t) z(t0 ).


For incompressible material models, S will be replaced by and s
by , according to Eq. (4.25).
Equations (7.60) and (7.61) have the exact same mathematical structure
as Eqs. (7.52) and (7.53), except that the Lodge E and Signorini e strains
have been replaced with Guth strains Z and z, respectively, where
Z t
Z (t) = T (S , E , E; t0 ) : dE (t0 )
1
Z0 t with T = M (7.62)
0 0 0 2 1
= T (S , E , E; t ) : E (t ) dt
P
0
Viscoelasticity 245

which pushes forward into the Eulerian frame according to Eqs. (3.37),
(7.42), and (B.83), resulting in
Z t
 
z(t) = F(t0 , t) t (s, e, e; t0 ) : d (t0 ) F T (t0 , t) dt0 (7.63)
0
or, in component form, as
Z t
j
ij
z = FIi (t0 , t) tIJKL (s, e, e; t0 ) dKL (t0 ) FJ (t0 , t) dt0 (7.64)
0
which, in practice, is evaluated more simply via
Z t
 
z(t) = F(t) F 1 (t0 ) t (s, e, e; t0 ) : d (t0 ) F T (t0 ) dt0 F T (t), (7.65)
0
where Z (t0 , t) = Z (t) Z (t0 ) and z(t0 , t) = z(t) z(t0 ) follow from the addi-
tive property of integrals. The viscoelastic strains Z and z of Guth9 obey
the contravariant maps of Eqs. (B.15) and (B.16), so a simpler pushforward
operation is
[z] = [F] [Z ] [F]T , (7.66)
implying that Guth strain is most easily integrated in the Lagrangian
configuration 0 . The viscoelastic tangent moduli T and t map as fourth-
order contravariant tensors according to Eq. (6.48).
To be able to write the Stieltjes integral, the first line in Eq. (7.62), as a
Riemann integral, the second line in (7.62), requires that the tangent mod-
ulus T be a continuous function and that the strain rate EP be integrable.
Equation (7.60) applies as an analytic continuation for explicit and im-
plicit elastic solids alike via the K-BKZ hypothesis. They dier in how
one quantifies their respective Guth strains Z . In essence, Guth strain can
be thought of as a dimensionless solution to any R t admissible elastic solid
1 M
that can be expressed in the form of Z = 2 1 0 (t ) : dE (t0 ) wherein
0

M = dS /dE .

7.5.1 Guth Strains for Explicit Elastic Solids


A Green elastic solid Eq. (5.3) with tangent modulus Eq. (5.53) has a Guth
strain
Z t 2
1 d W (E ; t0 )
Z = : dE (t0 ) (7.67)
2 1 0 dE dE
9 Guth et al. (1946) constructed a viscoelastic strain measure based upon the
strain resulting from a neo-Hookean elastic solid.
246 Soft Solids

whose elastic strain energy W (E ) is expressed as an explicit function of


Green strain E = 12 (C I). Dividing through by 2 1 makes the strain Z
dimensionless, as the strain-energy function W has units of stress.
A Lodge elastic solid Eq. (5.8) with tangent modulus Eq. (5.57) has an
associated Guth strain of
Z t
1  1  d2 W (E; t0 )  1 
Z= C C 1 : : C C 1
2 1 0 dE dE

dW (E; t0 ) 1
2 C 1 C 1 C
dE

dW (E; t0 ) 1
+ C 1 C C 1 : dE (t0 ) (7.68)
dE
whose elastic strain energy W (E) is expressed as an explicit function of
Lodge strain E = 12 (I C 1 ).
What we observe is that Guth strains are the stress responses of finite-
strain elastic theories normalized by their rubbery moduli so as to become
dimensionless. Specification of a functional form for either W (E ) or W (E)
establishes a material model.

7.5.2 Guth Strains for Implicit Elastic Solids


An implicit elastic solid has a tangent modulus described in Eq. (6.22) that
results in a Guth strain of
Z t !1
1 @2 U(E , S ; t0 )
Z=
2 1 0 @S @S
!
@2 U(E , S ; t0 )
: I I : dE (t0 ) (7.69)
@S @E
whose elastic internal energy U(E , S ) is an implicit function of strain E
and stress S . The dependence of this integral upon stress will make the
Volterra integral in Eq. (7.60) of implicit construction that will likely aect
ones choice of a numerical scheme for acquiring solutions.

7.5.3 Bulk/Shear Split


Equation (5.62) states that stress S can be split into separate hydrostatic
and deviatoric parts, viz., S = p C 1 + Sx , where p is the hydrostatic
Viscoelasticity 247

pressure and Sx is the deviatoric part to the second PiolaKirchho stress,


which is traceless, viz., tr(Sx C ) = 0. Explicit and implicit elastic solids
in Chaps. 5 and 6 have been constructed so that they can be put into a
constitutive expression of the form dS = (V V + D ) : dE wherein V is the
bulk (volumetric) tangent modulus and D is the shear (deviatoric) tangent
modulus so that
dSb = V : dE dS = dSb + dSs
and, therefore, (7.70)
dSs = D : dE S = Sb + Ss .

These two constituents of stress have separate viscoelastic characteristics.

7.5.3.1 Bulk Response


A viscoelastic bulk response is described by its glassy 0 and rubbery 1
bulk moduli, which are the analogs to 0 and 1 for the shear response.
In a typical application, the bulk response will be taken to be elastic.
Two reasons exist for this. First, 0 /1 is usually less than 10 (Ferry 1980,
Chap. 18), i.e., the eect is small compared to what arises from the shear
response where 0 / 1 typically falls within the range of 103 107 . And
second, volumetric relaxations typically occur much faster than their shear-
ing counterparts because b s , i.e., the characteristic time for bulk
relaxation is usually much smaller than the characteristic time for shear
relaxation. Consequently, the shear response will be the rate-controlling
response in most applications.
Nevertheless, applications exist where the bulk viscoelastic properties
of a material are important and need consideration, e.g., injection mold-
ing and capillary flows of polymers during processing (Leonov 1996)
and the viscoelastic eect that surfactant has on the alveolar response of
lung parenchyma (Smith and Stamenovic 1986; Stamenovic and Smith
1986a,b).
When bulk relaxations are a matter of importance, then an application
of the K-BKZ hypothesis to the volumetric response of Eqs. (5.64), (5.70),
(6.40), and (7.70) leads to a Volterra integral equation for the bulk stress of
 
Sb = 1 + (0 1 ) K(t) Y (t)
Z t
+ (0 1 ) K(t t0 ) Y (t0 , t) dt0 , (7.71)
0
248 Soft Solids

where K is the reduced bulk relaxation kernel, akin to G for the shear
response, and Y (t0 , t) = Y (t) Y (t0 ) with Y establishing the volumetric
strain of Guth defined by
Z t
1
Y (t) = V (p, ; t0 ) : dE (t0 )
1 0
Z t
1
= 2 p(t0 ) C 1 (t0 ) C 1 (t0 )
1 0

1 dp 0 1 0
(t ) C (t ) C (t ) : dE (t0 ) (7.72)
1 0
3 d
which simplifies to
Z t
1
Y (t) = 2 p(t0 ) E(t
P 0)
1 0

1 dp 0  P 0  1 0
(t ) tr C E; t C (t ) dt0 , (7.73)
3 d
where the hydrostatic pressure p and its gradient dp/d are described by an
appropriate elastic constitutive equation that, in this case, may be explicit
or implicit in origin. In the explicit case, p is given by a thermodynamic
potential from which dp/d is gotten by dierentiation, while in the im-
plicit case, dp/d is given by a thermodynamic potential from which p is
gotten through integration.

7.5.3.2 Deviatoric Response


The distortional response has a K-BKZ form similar to Eq. (7.60) in that
the shear stress response is

 
Ss (t) = 2 1 + ( 0 1 ) G (t) Z (t)
Z t
+ 2( 0 1 ) M(t t0 ) Z (t0 , t) dt0 . (7.74)
0
Whenever Green strain E is the independent strain measure, then its Guth
strain extends Eqs. (5.63) and (6.42) such that
Z t x
1 dS 0  
Z (t) = (t ) : I I 13 C (t0 ) C 1 (t0 ) : dE (t0 ) (7.75)
2 1 0 dEx
Viscoelasticity 249

or equivalently
Z t x
1 dS 0
Z (t) = (t ) : dEx (t0 ) (7.76)
2 1 0 dEx
where Eq. (3.55) defines the deviatoric strain rate dEx . Whenever Lodge
strain E is the independent strain measure, then Guth strain will extend
Eq. (5.69), producing
Z t
1 dSx 0  1 
Z (t) = (t ) : C C 1 13 C 1 C 1 (t0 )
2 1 0 dE x

 1  0
2 C S + S C (t ) : dE (t0 ) (7.77)
x x 1

or, equivalently,
Z t x
1 dS 0 x 0)
Z (t) = (t ) : dE(t
x
2 1 0 dE
Z t
1  
P Sx + Sx C EP (t0 ) dt0 , (7.78)
EC
1 0
x Gradients dSx /dEx and dSx /dE
where Eq. (3.61) defines dE. x follow from
their appropriate thermodynamic potentials, which can be of explicit or
implicit origin.

7.5.4 Tangent Moduli


The various elastic material models of Chaps. 5 and 6 can all be extended
into viscoelastic material models via the theoretical constructs put forward
in Eqs. (7.60)(7.63). All one needs to know is what their respective tan-
gent moduli are. The Guth strain for some models can be solved for analyt-
ically; others must be solved for numerically. A few models are cataloged
below.

7.5.4.1 Incompressible Materials


From Eq. (5.29), the first-order Green elastic solid has a strain energy of
W = 1 tr(C 1EC 1E ) so that, when substituted into Eq. (7.67), one gets
a Guth strain of
Z t Z t
1 0 1 0 0
Z (t) = C (t ) C (t ) : dE (t ) = dE(t0 ) = E(t) (7.79)
0 0
250 Soft Solids

which, when substituted into Eq. (7.60), becomes the classic viscoelastic
neo-Hookean solid of Eq. (7.52), where the strain-rate identity found in
Eq. (3.46) has been made use of.
From Eq. (5.32), the first-order Lodge elastic solid has a strain energy
of W = 1 tr(EC EC ) so that, when substituted into Eq. (7.68), one gets
a Guth strain of
Z t
 1 
Z (t) = C C 1 (t0 )
0
 1
1  0
2 C E + E C (t ) : dE (t0 )
Z t
 
= E(t) 2 P E + EC EP (t0 ) dt0
EC (7.80)
0

wherein identity (C 1 C 1 ) : (C C ) : (C 1 C 1 ) = C 1 C 1 has


been used.
An implicit elastic solid has a tangent modulus defined according to
Eq. (6.34) whose compliance is specified in Eq. (6.36) with an associated
viscoelastic compliance of
 1
2 1C = I I B : A y, (7.81)

where A y is a dimensionless version of A. Specifically, for the Rajagopal


elastic solid denoted as Material A, B remains as defined in Eq. (6.78) in
that
1  
B= 2 I CS + CS I , (7.82)
E1
where is Fungs parameter and E1 = 2 1 (1 + ) is the rubbery Youngs
modulus, while, from Eq. (6.77), one has
 
y = C C + C C 1 C E + E C
A (7.83)
1+ 1+ 2

which reduces to the viscoelastic Hookean solid whenever = 0.


Because
1  
T = y 1 : I I B ,
C 1 = A (7.84)
2 1
it necessarily follows that the integrals for integrating stress S and Guth
strain Z are coupled, as defined in Eqs. (7.60) and (7.62), for the analytic
Viscoelasticity 251

continuation of a Rajagopal elastic solid into the viscoelastic domain via


the K-BKZ conjecture. For plane-stress problems, like those considered
herein, one should quantify T by taking the inverse of C whenever C de-
pends upon the stress S .

7.5.4.2 Compressible Material


The implicit Hookean solid, whose elastic strain energy is described by
Eq. (6.71), has a tangent modulus for its bulk response described by
Eq. (6.72), whose associated Guth strain in Eq. (7.73) becomes
Z t
   
Y (t) = P t0 C 1 (t0 ) + ln det C 1 (t0 ) E(t
tr C E; P 0 ) dt0 (7.85)
0
and it has a tangent modulus for its shear response described by Eq. (6.74),
whose associated Guth strain becomes
x
Z (t) = E(t) (7.86)
x being defined by Eq. (3.62). When sub-
with the distortional strain E
stituted into Eqs. (7.71) and (7.74), respectively, these results produce a
viscoelastic model for the Hookean solid that is applicable for finite-strain
analysis.

7.6 Examples

To be able to write expressions for any of the various viscoelastic mod-


els, as they apply to any of the BVPs studied in this text, it is instruc-
tive to first determine the components of the relative deformation gradient
F(t0 , t) = F(0, t) F 1 (0, t0 ) and its inverse F 1 (t0 , t) for each particular BVP,
with these operators being defined in Eq. (7.40).
To present formul in as compact a notation as possible, the visco-
elastic kernels G (t) and M(t) are reverted back to their associated functions
G(t) and M(t), recalling that
G(t) = 1 + ( 0 1 ) G (t) & M(t) = ( 0 1 ) M(t) (7.87)
establish these relationships where 0 > 1  0, with 1 > 0 implying
a solid.
So as to keep the presentation size manageable, only the K-BKZ
viscoelastic neo-Hookean solid of Eq. (7.52) is analyzed in this examples
section.
252 Soft Solids

7.6.1 Uniaxial Extension


Components for the relative deformation gradient F(t0 , t) that describe the
uniaxial extension of an isotropic material (cf. Eq. 6.95) are given by
2 3
(t)/ (t0 ) 0 0
[F(t0 , t)] = 4 0 (t)/(t0 ) 0 5, (7.88)
0 0 0
(t)/(t )

while its inverse has components


2 0 3

1 0 (t )/ (t) 0 0
F (t , t) = 4 0 (t0 )/(t) 0 5, (7.89)
0 0 0
(t )/(t)

where and are the axial and transverse stretches, respectively, with =
1/2 for an isochoric response. These matrices can, in turn, be used in any
of the various definitions for relative strain found in the integrands of the
various viscoelastic models. In the above formul and those to follow, the
one-argument stretches (t) and (t0 ) are shorthand notations for (0, t) and
(0, t0 ), respectively. In all cases that follow, stretches are measured against
the initial frame associating with time 0, i.e., a Lagrangian viewpoint is
adopted.
The dependent variable in these viscoelastic models, viz., the traction
or engineering stress T, can be experimentally quantified via Eq. (4.43),
thereby allowing the component of stress in Eq. (6.95) to be determined
from which comparisons between theory and experiment can follow.
The viscoelastic neo-Hookean solid Eq. (7.52) has a Lagrange multi-
plier for uniaxial extension of
  Z t  
1 0 (t0 )
} = G(t) 1 M(t t ) 1 dt0 , (7.90)
(t) 0 (t)
leading to an axial traction of
  Z t  
1 0 (t) (t0 )
T(t) = G(t) (t) 2 + M(t t ) dt0 , (7.91)
(t) 0 2 (t0 ) 2 (t)

where Eqs. (3.66), (4.25), (4.48), (7.39), and (7.52) have been used. This
is not the viscoelastic model proposed by Guth et al. (1946), viz., the strain
measure in the integrand is dierent from (t0 ) 1/ 2 (t0 ), as they supposed
it to be.
Viscoelasticity 253

7.6.2 Equi-biaxial Extension


The material is assumed to be isotropic, with a deformation being described
by Eq. (2.48), so that the engineering stresses T1 and T2 established in
Eqs. (4.51) and (4.52) are approximately equal along the two loading di-
rections for this BVP. If this is the case, then the following BVP applies.
Components for the relative deformation gradient F(t0 , t) that describe
an isochoric equi-biaxial extension (cf. Eq. 2.48) are given by
2 3
(t)/ (t0 ) 0 0
[F(t0 , t)] = 4 0 (t)/ (t0 ) 0 5, (7.92)
0 0 0
(t )/ (t)
2 2

whose inverse has components


2 0 3

1 0 (t )/ (t) 0 0
F (t , t) = 4 0 (t0 )/ (t) 0 5, (7.93)
0 0 2 0
(t)/ (t )
2

from which the various definitions for relative strain measures can be de-
termined.
Substituting Eq. (7.92) for F(t0 , t) into the definition for the relative
Lodge strain E(t0 , t) given in Eq. (7.39), while using Eq. (4.57) to quantify
stress and Eq. (3.72) to quantify strain, the Lagrange multiplier for the
viscoelastic neo-Hookean solid Eq. (7.52) obtained from the 33 component
is given by

  Z t !
1 4 (t0 )
}(t) = G(t) 1 4 M(t t0 ) 1 4 dt0 , (7.94)
(t) 0 (t)

so that the engineering stress T = T1 = T2 obeys

  Z t !
1 (t) 4 (t0 )
T(t) = G(t) (t) 5 + M(t t0 ) dt0 , (7.95)
(t) 0 2 (t0 ) 5 (t)

whose experimental counterparts can be determined via Eqs. (4.51) and


(4.52).
254 Soft Solids

7.6.3 Simple Shear


The relative deformation gradient F(t0 , t) = F(0, t) F 1 (0, t0 ) has compo-
nents that describe simple shear (cf. Eq. 2.52) of
2 3
1  (t)  (t0 ) 0
[F(t0 , t)] = 40 1 05 , (7.96)
0 0 1

while its inverse F 1 (t0 , t) = F(0, t0 ) F 1 (0, t) is described by


2 3

1 0 1  (t0 )  (t) 0
F (t , t) = 40 1 05 , (7.97)
0 0 1
from which the various definitions for relative strain measures can be de-
termined. In the above formul, the one-argument shears  (t) and  (t0 ) are
shorthand notations for  (0, t) and  (0, t0 ), respectively. The magnitude
of shear is measured against an initial frame associated with time 0, viz.,
 (0, 0) = 0.
The Lagrange multiplier for motions of rectilinear simple shear will be
zero, assuming that a condition of plane stress applies, as is the case here,
because simple shear is a planar motion, i.e., P 3 = 0.
From the definition for the relative strain of Lodge E(t0 , t) given in
Eq. (7.39), using Eq. (3.77) to quantify the Cauchy deformation C 1 and
recalling that the Lagrange multiplier for the viscoelastic hyper-Hookean
solid Eq. (7.53) is zero for simple shear, assuming plane stress, it follows
then that the second PiolaKirchho stress S has components
  2
S11 (t) S12 (t)  (t)  (t)
= G(t)
S21(t) S22 (t)  (t) 0
Z t   2 
0  (t)  2 (t0 )  (t)  (t0 )
+ M(t t ) dt0 . (7.98)
0  (t)  (t0 ) 0
Each component is described by a Volterra integral equation of the second
kind. Equivalently, components of the first PiolaKirchho stress P are
described by
Z t
 
P11 (t) = M(t t0 )  (t0 )  (t)  (t0 ) dt0 ,
0
Z t (7.99)
0  0  0
P12 (t) = G(t)  (t) + M(t t )  (t)  (t ) dt ,
0
Viscoelasticity 255

where P22 = 0 and P21 = P12 , in this case. Components P12 and P22
can both be experimentally measured, but P11 cannot, as documented in
Eqs. (4.63) and (4.68). The negative state of stress arising for P11 may
manifest itself as a wrinkling of the surface whenever thin membranes are
sheared. Notice that  (t)  (t0 ) is the shear strain between states 0 and
, in accordance with the K-BKZ hypothesis.

7.6.4 Planar Membranes


Components for the relative deformation gradient F(t0 , t) that describe an
isochoric planar membrane (cf. Eq. 2.61) are given by

0 1 1 (t) 2 (t0 ) 1 (t)2 (t0 ) 1 (t0 )1 (t) 1 (t)1 (t0 )
[F(t , t)] = ,

(t0 ) 2 (t0 )2 (t) 2 (t)2 (t0 ) 1 (t0 ) 2 (t) 1 (t0 )2 (t)
(7.100)

while its inverse has components



1 1 (t0 ) 2 (t) 1 (t0 )2 (t) 1 (t)1 (t0 ) 1 (t0 )1 (t)
[F 1 (t0 , t)] = ,

(t) 2 (t)2 (t0 ) 2 (t0 )2 (t) 1 (t) 2 (t0 ) 1 (t)2 (t0 )
(7.101)

which follow from F(t0 , t) = F(0, t) F 1 (0, t0 ) and whose associated areal
stretch is

(t)

(t0 , t) = , (7.102)

(t0) 
as derived from
(t0 , t) = det F(t0 , t) = det F(0, t) F 1 (0, t0 ) = det F(0, t)
det F 1 (0, t0 ) = det F(0, t)/ det F(0, t0 ) =
(0, t)/
(0, t0 ), where use has
been made of properties (A.74) and (A.75). These results can be used
in any of the various definitions for relative strain found in the integrands
of our various viscoelastic models. In the above formul, the one-argu-
ment variables are shorthand notations, e.g., 1 (t) and 1 (t0 ) are shorthand
notations for 1 (0, t) and 1 (0, t0 ), respectively.
The viscoelastic material models considered in this text present them-
selves as 2  2 matrix equations that are best solved numerically in their
Lagrangian constructions. To be able to solve the incompressible models
requires that one know how to calculate their Lagrange multipliers, which
need to be solved prior to solving the constitutive equation itself. For this
BVP, this has been done by assuming that the membranes are planar and
in a state of plane stress, from which the Lagrange multiplier can then be
derived from the 33 component straightaway.
256 Soft Solids

The viscoelastic neo-Hookean solid Eq. (7.52) is

S (t) = }(t) C 1 (0, t)


Z t
+ 2G(t) E(0, t) + 2 M(t t0 ) E(t0 , t) dt0 , (7.103)
0
whose Lagrange multiplier }, for this BVP, is described by the Volterra
integral equation
  Z t !
1
2 (t0 )
}(t) = G(t) 1 2 M(t t0 ) 1 2 dt0 , (7.104)

(t) 0
(t)

where
is the areal stretch defined in Eq. (1.23). Unlike the prior BVPs
considered, where the Lagrange multiplier has been assimilated into the
equation governing the stress components, here it is best left as a separate
equation to be sequentially solved with the constitutive equation.

7.7 Applications

The process of characterizing a viscoelastic material model can be broken


down into a two-step process. First, one characterizes its viscous behav-
ior, viz., a viscoelastic kernel is selected, and the associated rubbery and
glassy shear moduli are quantified. Second, one characterizes its limiting
elastic behavior, viz., a constitutive structure is selected, and the remaining
parameters are determined, if any exist.

7.7.1 Selecting a Kernel


To be able to select a kernel, one needs to first have an idea of how the
various kernels behave. Of particular importance is to take notice of the
short- and long-time relaxation behaviors of the material to be character-
ized, as this is where the greatest distinctions between the various kernel
responses reside. To illustrate the short-time behavior of the various ker-
nels, linear plots in time vs. reduced relaxation are presented in Figs. 7.6,
7.8, 7.10, and 7.12. To illustrate the long-time behavior of the various ker-
nels, loglog plots in time vs. reduced relaxation are presented in Figs. 7.7,
7.9, 7.11, and 7.13. Applications of man-made materials often place im-
portance on the long-term response, while the very nature of living tissue
typically emphasizes the short-term response in their applications.
Viscoelasticity 257

FOV Kernel
1
Reduced Relaxation Modulus, G(t)

= 1.0
0.8
= 0.8
= 0.6
= 0.4
0.6 = 0.2

0.4

0.2

0
0 1 2 3
t/

Fig. 7.6 Plot of normalized time t/ vs. the reduced relaxation modulus G (t) for
the fractional-order viscoelastic model where = 1 associates with the integer-
order viscoelastic model of Zener. Emphasis is on the short-time response

The FOV relaxation kernel of Eq. (7.23) is drawn in Figs. 7.6 and 7.7.
Of note in the short-time response is that the FOV kernel behaves as the
IOV kernel at = 1. Stress relaxation quickens with decreasing values
of , until about half the stress has been recovered. Beyond that point,
the rate or relaxation slows way down, relative to exponential decay, and
may even appear to plateau out. An examination of this kernels long-term
relaxation response, displayed in Fig. 7.7, demonstrates that this apparent
plateau is a fictional artifact of the material becoming power law in its
response, i.e., it is likely that no such plateau exists, and one must be careful
not to falsely associate it with the rubbery modulus. So, what one looks for
in an experimental data set that suggests considering the FOV kernel is an
extremely rapid stress recovery in the early stages of relaxation followed
by an ultraslow recovery at large time.
The KWW relaxation kernel of Eq. (7.27) is drawn in Figs. 7.8 and
7.9. Its short-time response has a similar character to that of the FOV,
with a few notable distinctions. For 0
t , the FOV and KWW
kernels are eectively equivalent, as noted in Eq. (7.26). As time gets
larger, these kernels predict vastly dierent responses. The KWW kernel
has a point in common for all values of , i.e., at t =  the stress will have
258 Soft Solids

FOV Kernel
100
Reduced Rlaxation Modulus, G(t)

10-1

= 1.0
= 0.8
10-2 = 0.6
= 0.4
= 0.2

10-3

10-4
10-4 10-3 10-2 10-1 100 101 102 103 104
t/

Fig. 7.7 Loglog plot of normalized time t/ vs. the reduced relaxation modulus
G (t) for the fractional-order viscoelastic model where = 1 associates with the
integer-order viscoelastic model of Zener. Emphasis is on the long-time response

relaxed about 63 % regardless of the value of . It is at large times that the


stretched exponential and the Mittag-Leer function have widely diering
responses, viz., whenever t  . Here, the KWW kernel behaves as an
exponential, while the FOV kernel behaves as a power law. So, what one
looks for in an experimental data set that suggests considering the KWW
kernel is an extremely rapid stress recovery in the early stages of relaxation
that, over time, gives way to an exponential decay at large times.
The CCM relaxation kernel of Eq. (7.29) is drawn in Figs. 7.10 and
7.11. Unlike the FOV and KWW kernels, the CCM kernel does not con-
tain the IOV kernel as a special case. What one sees when looking at the
relaxation curves for the CCM kernel is that they are very similar in shape
to those of the FOV kernel, both in the short- and long-time asymptotes.
In fact, they become equivalent for t  , as noted in Eq. (7.26). What is
predominantly dierent between them is how one interprets the time con-
stant . In the CCM kernel,  represents the half-life of relaxation where
half the transient stress has relaxed away and half remains, independent of
the value of . In this regard, the CCM power-law kernel of Cole and Cole
(1941, 1942) is an excellent approximating function for the physically mo-
tivated Mittag-Leer kernel belonging to the fractional-order Zener model
Viscoelasticity 259

Fig. 7.8 Plot of normalized time t/ vs. the reduced relaxation modulus G (t) for
the Kohlrausch and Williams and Watts model where = 1 associates with the
integer-order viscoelastic model of Zener. Emphasis is on the short-time response

FOV of Caputo and Mainardi (1971b,a) (see also Doehring et al. (2005)),
with the advantage being that the CCM kernels are much easier and quicker
to compute.
The BOX relaxation kernel of Eq. (7.31) is drawn in Figs. 7.12 and
7.13. In these figures, one observes that the BOX kernel can mimic the
short-time behaviors of the other kernels. When time t has exceeded 2 ,
stress will relax with an exponential rate of decay, like the KWW kernel,
whereas for times that lie between the characteristic times 1 and 2 , the
BOX kernel behaves more like a power law. For these reasons, the BOX
kernel has found application in biological tissues. Because their relaxation-
like histories do not reach large times, the characterized response behaves
like the FOV kernel since the exponential tail at long times is not reached
in practice.
The BOX, CCM, FOV, and KWW viscoelastic kernels can all be used
to account for rapid short-time relaxation behavior that is common among
soft solids, relative to the IOV kernel. So, if short-time relaxation is
all that is to be modeled, as is usually the case when modeling living
260 Soft Solids

Fig. 7.9 Loglog plot of normalized time t/ vs. the reduced relaxation modulus
G (t) for the Kohlrausch and Williams and Watts model where = 1 associates
with the integer-order viscoelastic model of Zener. Emphasis is on the long-time
response

biological tissues, then little phenomenological evidence exists to support


selecting one over the others. If long-time relaxation data are available,
then sucient phenomenological evidence will exist to warrant a selection
of one or two models over the others for the purpose of material parame-
terization.

7.7.1.1 Polyisobutylene
To illustrate this process, consider what is likely to be the most thoroughly
characterized viscoelastic material: the polyisobutylene prepared by the
National Bureau of Standards in the early days of polymer research when
testing techniques and standards were being developed, which necessitated
a uniformity in material being tested by the various university and indus-
trial laboratories involved. The experimental data presented in Figs. 7.14
and 7.15 were published by Catsi and Tobolsky (1955) where a time/
temperature translation of the data has taken place, being corrected to
25 C.
Viscoelasticity 261

CCM Kernel
1

IOV kernel
Reduced Relaxation Modulus, G(t)

0.8 = 1.0
= 0.8
= 0.6
= 0.4
0.6 = 0.2

0.4

0.2

0
0 1 2 3
t/

Fig. 7.10 Plot of normalized time t/ vs. the reduced relaxation modulus G (t)
for the ColeCole model contrasted against the integer-order viscoelastic model
of Zener. Emphasis is on the short-time response

The linearlinear plot in time vs. reduced relaxation in Fig. 7.14 shows
the short-time response of polyisobutylene with the stress relaxing at a
much faster rate than an exponential would suggest, while the loglog plot
in Fig. 7.15 shows the response becoming power law in character at large
times. Hence, the exponential decay properties of the BOX, KWW, and
IOV kernels eliminate them from further consideration for this particular
material, leaving the CCM and FOV kernels as candidate models.
Fitting the CCM and FOV kernels to the experimental data of
Catsi and Tobolsky (1955) leads to the following parameterizations,
with 95 % confidence intervals being reported. For the CCM kernel,
= 0.6850 0.0025 and  = (4.05 0.11)  109 s with a coecient
of determination of R2 = 0.999. For the FOV kernel, = 0.6491 0.0028
and  = (9.79 0.40)  109 s with a coecient of determination
of R2 = 0.998. These parameters were fit using a genetic algorithm
(Goldberg 1989, 2002) for parameter estimation in the loglog space of
Fig. 7.15 (not in the linearlinear space of Fig. 7.14).
When viewed in the space in which they were fit, i.e., Fig. 7.15, the
CCM kernel is seen to do an excellent job of describing the data over the
entire range of the data, whereas, although the FOV kernel does a very
262 Soft Solids

CCM Kernel
100
Reduced Rlaxation Modulus, G(t)

10-1

IOV kernel
= 1.0
10-2 = 0.8
= 0.6
= 0.4
= 0.2

10-3

10-4
10-4 10-3 10-2 10-1 100 101 102 103 104
t/

Fig. 7.11 Loglog plot of normalized time t/ vs. the reduced relaxation modulus
G (t) for the ColeCole model contrasted against the integer-order viscoelastic
model of Zener. Emphasis is on the long-time response

good job, it does not do as well as the CCM kernel. The superior quality of
fit of the CCM kernel over the FOV kernel for this material is immediately
apparent in the short-time response data of Fig. 7.14. For polyisobutylene,
the CCM kernel of Cole and Cole (1941, 1942) is the best kernel of the
viscoelastic kernels considered in this text at describing the time-dependent
viscoelastic characteristics of polyisobutylene.

7.7.1.2 Natural Rubber


The creation of chemical bonds or crosslinks between individual poly-
mer chains produces a polymer network, i.e., a rubber or an elastomer.
This loosely pinned network is why these materials can exhibit large elas-
tic deformations. Sulfur reacts with natural rubber to bring about cross-
linkinga process called vulcanization. Viscoelastic attributes can arise
in elastomeric materials whenever crosslinks break under loada process
called scission (Tobolsky 1960; Wineman and Min 2003).
To study the viscoelastic eect due to scission, Tobolsky and Mercurio
(1959) cured natural rubber of high purity by bombarding it with an elec-
tron beam, so that cross-linking could take place without the aid of any
Viscoelasticity 263

BOX Kernel
1
IOV kernel
1 = 5x10-1, 2 = 5x100
Reduced Relaxation Modulus, G(t)

0.8
1 = 5x10-2, 2 = 5x101
1 = 5x10-3, 2 = 5x102
0.6
1 = 5x10-4, 2 = 5x103

0.4

0.2

0
0 1 2 3
t

Fig. 7.12 Plot of time t vs. the reduced relaxation modulus G (t) for the BOX
model of Neubert contrasted against the integer-order viscoelastic model of Zener.
Emphasis is on the short-time response

chemical agents. Chemical activity can have a strong eect on relaxation


behavior (Tobolsky 1960). Their experiments separate these concerns, al-
lowing just scission eects to be quantified. Their stress-relaxation data
are presented in Fig. 7.16.
Of the kernels considered in this text, the experimental data in Fig. 7.16
are best represented by the exponential IOV kernel where  = 13, 500
250 s with a coecient of determination of R2 = 0.9. This is not an ex-
ceptionally good fit. No kernel considered herein captures the physics of
scission correctly. These data show a faster than exponential behavior in
the short-term response, which the BOX, CCM, FOV, and KWW kernels
are all capable of delivering on. These data also show a faster than expo-
nential behavior for the long-term response, which no model considered
herein can deliver on. When a genetic algorithm (Goldberg 1989, 2002)
was applied to fit the FOV and KWW kernels to these data, it returned
estimates for of 1 for both kernels, which is the IOV limit in these ker-
nels. Even for the BOX kernel, the genetic algorithm returned an expo-
nential response in that 1  2 , i.e., the optimizer reduced it to the limit-
ing case of an IOV kernel, too. Other mechanisms are at work that cause
264 Soft Solids

BOX Kernel
100
Reduced Rlaxation Modulus, G(t)

10-1

10-2 IOV kernel


1 = 5x10-1, 2 = 5x100
1 = 5x10-2, 2 = 5x101
10-3
1 = 5x10-3, 2 = 5x102
1 = 5x10-4, 2 = 5x103

10-4
10-4 10-3 10-2 10-1 100 101 102 103 104
t

Fig. 7.13 Loglog plot of time t vs. the reduced relaxation modulus G (t) for the
BOX model of Neubert contrasted against the integer-order viscoelastic model of
Zener. Emphasis is on the long-time response

scission, mechanisms that these kernels are not capable of representing; cf.
Wineman and Min (2003).

7.7.2 Selecting a Constitutive Equation


Selecting an appropriate constitutive equation, in the spirit of a K-BKZ
viscoelastic theory, boils down to determining what the tensorial character
of a viable model is. Since all of our viscoelastic models are analytic con-
tinuations of an elastic model, the selection process reduces to determining
what the limiting elastic response of the material is. This can be done using
any of the BVPs studied herein, for example, by performing the experiment
suciently fast enough so that the material responds according to its glassy
behavior or suciently slow enough so that the material responds accord-
ing to its rubbery behavior. This is easier said than done. Once these data
are in hand one can employ the selection guidelines outlined in Sect. 5.6.1.
If this is not possible, an alternative optimization strategy will likely need
to be developed, which is not an ideal strategy, but it is the strategy most
likely encountered in practice.
Viscoelasticity 265

Reduced Relaxation Modulus, G(t) Experimental Data


CCM kernel
0.8 FOV kernel

0.6

0.4

0.2

0
0 2x10-9 4x10-9 6x10-9 8x10-9 10-8
Time, t (s)
Fig. 7.14 Linear plot of time t vs. the reduced relaxation modulus G (t) for the
polyisobutylene data of Catsi and Tobolsky (1955) corrected for 25 C, zooming
in on the short-time response. The glassy and rubbery shear moduli are 0 =
1.0 GPa and 1 = 250 kPa (Tobolsky 1956). The time axis represents about 2
for the CCM kernel and about 1 for the FOV kernel

7.8 Exercises

7.8.1 Pure Shear


Using the results you acquired from the exercises of prior chapters, derive
the formul that govern the response of a viscoelastic neo-Hookean solid
Eq. (7.52) subjected to pure shear, i.e., derive the Lagrange multiplier }
and the two tractions T1 and T2 .

7.8.2 Biaxial Extension


Using the results you acquired from the exercises of prior chapters, derive
the formul that govern the response of a viscoelastic neo-Hookean solid
Eq. (7.52) subjected to biaxial extension, i.e., derive the Lagrange multi-
plier } and the two tractions T1 and T2 .
266 Soft Solids

100

Reduced Relaxation Modulus, G(t)

10-1

10-2
Experimental Data
CCM kernel
FOV kernel

10-3

10-4
10-11 10-10 10-09 10-08 10-07 10-06 10-05 10-4 10-3 10-2
Time, t (s)

Fig. 7.15 Loglog plot of time t vs. the reduced relaxation modulus G (t) for
the polyisobutylene data of Catsi and Tobolsky (1955) corrected for 25 C. The
glassy and rubbery shear moduli are 0 = 1.0 GPa and 1 = 250 kPa (Tobolsky
1956). Emphasis is on the long-time response

7.8.3 Extension Followed by Simple Shear


Using the results you acquired from the exercises of prior chapters, derive
the formul that govern the response of a viscoelastic neo-Hookean solid
Eq. (7.52) subjected to a simple shear following an axial extension, i.e.,
derive the Lagrange multiplier }, the two tractions T1 and T2 acting on the
face whose normal is in the 2-direction, and the 11-component of stress.

7.8.4 Other Problems


1. In 1952, McLoughlin and Tobolsky (1952) determined the master
stress-relaxation curve for polymethyl methacrylate (PMMA), a.k.a.
acrylic. They normalized their data for 40 C by shifting or translating
stress-relaxation curves achieved at dierent temperatures to the curve
at 40 C. This technique of constructing a master curve is now standard
practice in polymer rheology, but the details of its implementation lie
beyond the scope of this book. Given their time-shifted master curve,
whose raw data are reproduced in Table 7.1 and displayed in Fig. 7.17,
select an appropriate relaxation kernel, justifying your decision.
Viscoelasticity 267

100

Reduced Relaxation Modulus, G(t)

10-1
Experimental Data
IOV kernel

10-2
102 103 104 105
Time, t (s)

Fig. 7.16 Loglog plot of time t vs. the reduced relaxation modulus G (t) for the
radiation-cured natural rubber data of Tobolsky and Mercurio (1959) at 130 C.
The glassy and rubbery shear moduli are 0 = 850 MPa (Tobolsky 1956) and
1 = 350 kPa, cf. p. 147

Then determine its parameters and draw your fitted curve against their
data. Discuss the results.
2. An alternative power-law kernel to the CCM kernel that one finds in
the literature is the MPL kernel G (t) = 1/(1 t/) . What is its memory
kernel? Construct a graph that shows the similarities and dierences
between the CCM and MPL kernels. Discuss what you find.
3. The data presented in Tables 7.2, 7.3, and 7.4 and drawn in Figs. 7.18,
7.19, and 7.20 were obtained from simple-shear experiments done on
a sample of porcine myocardial heart tissue by Dokos et al. (2002).
These are precious data; they are rare. Shear was imposed on three
orthogonal planes. The fact that the responses are dierent in dierent
material directions is an indication of material anisotropy, a topic not
addressed in this text. The hysteretic area is proportional to the energy
lost per cycle due to the inherent viscoelastic behavior of muscle. The
extreme nonlinearity is indicative of a tensorial character suggested by
the implicit elastic solid of Chap. 6 that was analytically continued into
the viscoelastic domain in Sect. 7.5. Based on this, your instructor may
ask you to use these data in any number of dierent ways to illustrate
anyone of these issues, which are topics that lie beyond the intended
scope of this text.
268 Soft Solids

Fig. 7.17 Experimental data for the reduced relaxation modulus G (t) are from
McLoughlin and Tobolsky (1952) for polymethyl methacrylate (PMMA) corrected
for 40 C, as recorded in Table 7.1. The glassy and rubbery shear moduli are
0 = 750 MPa and 1 = 750 kPa (Tobolsky 1956)
Viscoelasticity 269

Stabilized Passive Shear Response of Porcine Myocardium


Muscle Fibers Lie Normal to the Shear Plane
16

12
FN
FS
8
Shear Stress, T (kPa)

-4

-8

-12

-16
-0.6 -0.4 -0.2 0 0.2 0.4 0.6
Shear,

Fig. 7.18 Dynamic response of porcine myocardium subjected to simple shear


over a range of 1/2

1/2 with a 30 s period. The muscle bers are orthogonal
to the plane of shearing. Experiment FN shears the sample in the direction of the
normal to the muscle sheet. Experiment FS shears the sample along the muscle
sheet. The experimental data are from Dokos et al. (2002) with values being
reported in Table 7.2
270 Soft Solids

Stabilized Passive Shear Response of Porcine Myocardium


Fiber Sheets are Parallel with the Shear Plane
2

NF
NS
1
Shear Stress, T (kPa)

-1

-2
-0.6 -0.4 -0.2 0 0.2 0.4 0.6
Shear,

Fig. 7.19 Dynamic response of porcine myocardium subjected to simple shear


over a range of 1/2

1/2 with a 30 s period. The sheets of muscle ber lie
parallel to the plane of shearing. Experiment NF shears the sample in the direction
of the muscle bers. Experiment NS shears the sample along the muscle sheet.
The experimental data are from Dokos et al. (2002) with values being reported
in Table 7.3
Viscoelasticity 271

Stabilized Passive Shear Response of Porcine Myocardium


Fiber Sheets are Orthogonal to the Shear Plane
6

4 SF
SN
Shear Stress, T (kPa)

-2

-4

-6
-0.6 -0.4 -0.2 0 0.2 0.4 0.6
Shear,

Fig. 7.20 Dynamic response of porcine myocardium subjected to simple shear


over a range of 1/2

1/2 with a 30 s period. The sheets of muscle ber are
orthogonal to the plane of shearing. Experiment SF shears the sample in the
direction of the muscle bers. Experiment SN shears the sample in a direction
orthogonal to the muscle sheet. The experimental data are from Dokos et al.
(2002) with values being reported in Table 7.4
272 Soft Solids

Table 7.1 Experimental data for the


reduced relaxation modulus G (t) for
shear, computed from the timetemper-
ature-shifted elastic moduli E(t) data
of McLoughlin and Tobolsky (1952,
Table II) for polymethyl methacrylate
(PMMA) corrected for 40 C where
G(t) = E(t)/3 has been assumeda

log10 (t) G (t) log10 (t) G (t)


hr. hr.
3 0.989 5 0.183
2 0.955 6 0.0692
1 0.902 7 0.0193
0 0.803 8 0.00494
1 0.708 9 0.00158
2 0.609 10 0.00043
3 0.495 11 0.00008
4 0.355 12 0.00002
a These data are drawn in Fig. 7.17
Viscoelasticity 273

Table 7.2 Passive saturated shear response of porcine myocardiuma

FN FS
Shear Traction Shear Traction Shear Traction Shear Traction
 T1 (kPa)  T1 (kPa)  T1 (kPa)  T1 (kPa)
0.495 11.220 0.494 15.531 0.493 15.360 0.494 12.091
0.495 10.894 0.487 13.780 0.494 14.933 0.495 11.910
0.490 10.199 0.477 11.330 0.489 14.092 0.491 11.148
0.479 8.231 0.461 8.462 0.479 11.613 0.481 9.434
0.461 6.028 0.437 5.307 0.464 8.603 0.467 7.598
0.437 4.005 0.408 3.212 0.444 5.772 0.448 5.578
0.407 2.421 0.372 1.737 0.420 3.694 0.417 3.433
0.371 1.351 0.338 0.975 0.389 2.043 0.387 2.148
0.331 0.768 0.296 0.468 0.352 1.138 0.352 1.232
0.289 0.412 0.267 0.338 0.318 0.690 0.311 0.709
0.243 0.213 0.225 0.162 0.278 0.449 0.266 0.400
0.194 0.133 0.176 0.075 0.232 0.240 0.215 0.250
0.141 0.061 0.124 0.031 0.182 0.149 0.166 0.156
0.088 0.019 0.074 0.056 0.133 0.087 0.138 0.060
0.036 0.063 0.018 0.069 0.082 0.057 0.084 0.061
0.023 0.106 0.039 0.134 0.035 0.025 0.036 0.001
0.078 0.112 0.091 0.177 0.018 0.031 0.014 0.023
0.134 0.233 0.137 0.253 0.075 0.093 0.062 0.085
0.183 0.377 0.182 0.376 0.124 0.156 0.113 0.179
0.235 0.682 0.228 0.572 0.173 0.306 0.161 0.240
0.281 1.219 0.274 0.885 0.224 0.496 0.207 0.416
0.324 1.975 0.318 1.402 0.266 0.796 0.249 0.687
0.363 3.168 0.355 2.122 0.304 1.291 0.292 1.105
0.393 4.588 0.389 3.239 0.343 2.057 0.330 1.771
0.424 6.837 0.418 4.592 0.381 3.247 0.368 2.913
0.450 9.628 0.443 6.235 0.412 4.776 0.406 4.994
0.469 12.190 0.464 8.159 0.440 6.643 0.438 7.794
0.482 14.222 0.479 9.668 0.460 8.479 0.464 10.869
0.490 15.519 0.490 10.827 0.477 10.188 0.481 13.669
0.494 15.745 0.495 11.220 0.488 11.537 0.493 15.360

Data are from Dokos et al. (2002, Fig. 6) whose raw values were supplied to the
author with permission granted for use by Prof. Ian J. LeGrice, University of
Auckland, Auckland, New Zealand
a Samples 3  3  3 mm in size were subjected to simple shearing over a range in
shear of  = 0.5 under a sinusoidal waveform with a 30 s period. Their frame
of reference is : F aligns with the direction of the muscle bers, S is transverse
to F and lies within the sheet of muscle bers, and N is normal to this sheet.
Six shearing orientations were exercised on the same tissue sample, designated by
pairs, e.g., FS. The rst character, F in this case, denes the direction normal to
the shearing plane (i.e., direction 2 in Fig. 1.5). The second character, S in this
case, denes the direction of shearing (i.e., direction 1 in Fig. 1.5)
274 Soft Solids

Table 7.3 Passive saturated shear response of porcine myocardiuma

NF NS
Shear Traction Shear Traction Shear Traction Shear Traction
 T1 (kPa)  T1 (kPa)  T1 (kPa)  T1 (kPa)
0.499 1.839 0.499 1.669 0.499 1.990 0.499 1.596
0.497 1.804 0.495 1.598 0.497 1.920 0.494 1.502
0.489 1.665 0.485 1.441 0.488 1.754 0.482 1.305
0.477 1.437 0.469 1.196 0.474 1.478 0.467 1.055
0.460 1.160 0.450 0.984 0.457 1.242 0.448 0.859
0.440 0.919 0.424 0.758 0.434 0.961 0.421 0.637
0.415 0.710 0.389 0.545 0.405 0.705 0.392 0.461
0.386 0.522 0.348 0.369 0.371 0.499 0.359 0.334
0.351 0.348 0.306 0.246 0.333 0.361 0.315 0.212
0.312 0.245 0.251 0.123 0.289 0.220 0.284 0.162
0.273 0.138 0.223 0.087 0.247 0.125 0.243 0.109
0.231 -yy0.103 0.177 0.086 0.202 0.085 0.198 0.064
0.186 0.053 0.122 0.036 0.149 0.036 0.144 0.035
0.141 0.035 0.070 0.019 0.102 0.014 0.092 0.010
0.090 0.000 0.017 0.018 0.055 0.011 0.043 0.011
0.043 0.000 0.042 0.017 0.004 0.035 0.010 0.036
0.005 0.017 0.092 0.071 0.046 0.060 0.064 0.088
0.053 0.035 0.150 0.104 0.099 0.060 0.112 0.109
0.101 0.069 0.197 0.156 0.157 0.129 0.163 0.151
0.148 0.104 0.246 0.227 0.213 0.158 0.208 0.221
0.194 0.140 0.289 0.313 0.255 0.232 0.252 0.339
0.239 0.227 0.326 0.434 0.294 0.294 0.292 0.431
0.283 0.335 0.361 0.572 0.337 0.454 0.339 0.617
0.327 0.458 0.398 0.799 0.372 0.580 0.379 0.878
0.362 0.615 0.428 1.058 0.403 0.777 0.413 1.104
0.396 0.792 0.451 1.282 0.432 0.974 0.441 1.361
0.424 1.001 0.472 1.508 0.457 1.174 0.462 1.597
0.450 1.195 0.487 1.700 0.474 1.351 0.481 1.808
0.474 1.403 0.496 1.804 0.490 1.548 0.493 1.944
0.489 1.579 0.499 1.839 0.497 1.621 0.499 1.990

Data are from Dokos et al. (2002, Fig. 6) whose raw values were supplied to the
author with permission granted for use by Prof. Ian J. LeGrice, University of
Auckland, Auckland, New Zealand
a Samples 3  3  3 mm in size were subjected to simple shearing over a range in
shear of  = 0.5 under a sinusoidal waveform with a 30 s period. Their frame
of reference is: F aligns with the direction of the muscle bers, S is transverse
to F and lies within the sheet of muscle bers, and N is normal to this sheet.
Six shearing orientations were exercised on the same tissue sample, designated by
pairs, e.g., FS. The rst character, F in this case, denes the direction normal to
the shearing plane (i.e., direction 2 in Fig. 1.5). The second character, S in this
case, denes the direction of shearing (i.e., direction 1 in Fig. 1.5)
Viscoelasticity 275

Table 7.4 Passive saturated shear response of porcine myocardiuma

SF SN
Shear Traction Shear Traction Shear Traction Shear Traction
 T1 (kPa)  T1 (kPa)  T1 (kPa)  T1 (kPa)
0.498 4.287 0.497 5.719 0.498 4.810 0.497 2.927
0.494 4.086 0.492 5.285 0.494 4.668 0.489 2.576
0.483 3.557 0.482 4.533 0.485 4.245 0.474 2.092
0.467 2.909 0.467 3.628 0.467 3.432 0.455 1.564
0.444 2.168 0.447 2.710 0.448 2.764 0.430 1.110
0.420 1.626 0.424 1.928 0.424 2.124 0.402 0.758
0.392 1.098 0.395 1.277 0.395 1.515 0.369 0.508
0.356 0.708 0.362 0.787 0.363 1.037 0.329 0.299
0.315 0.378 0.324 0.448 0.322 0.646 0.290 0.196
0.271 0.227 0.283 0.239 0.284 0.429 0.239 0.139
0.227 0.115 0.240 0.151 0.242 0.269 0.193 0.094
0.182 0.056 0.196 0.062 0.199 0.153 0.147 0.064
0.129 0.014 0.148 0.027 0.148 0.108 0.090 0.064
0.078 0.006 0.112 0.003 0.100 0.053 0.062 0.024
0.022 0.041 0.053 0.017 0.052 0.023 0.002 0.008
0.030 0.086 0.006 0.059 0.001 0.007 0.054 0.036
0.086 0.106 0.056 0.100 0.050 0.022 0.112 0.051
0.134 0.175 0.108 0.139 0.105 0.034 0.166 0.122
0.183 0.257 0.155 0.197 0.159 0.080 0.216 0.253
0.227 0.384 0.204 0.278 0.210 0.125 0.262 0.414
0.270 0.548 0.252 0.410 0.260 0.197 0.309 0.732
0.309 0.795 0.298 0.646 0.306 0.345 0.350 1.139
0.350 1.157 0.337 0.959 0.350 0.549 0.389 1.706
0.383 1.683 0.372 1.369 0.388 0.874 0.420 2.358
0.415 2.375 0.411 2.011 0.421 1.255 0.444 2.967
0.442 3.234 0.440 2.651 0.449 1.768 0.465 3.665
0.462 4.029 0.463 3.277 0.470 2.223 0.480 4.186
0.480 4.890 0.481 3.819 0.486 2.661 0.491 4.594
0.490 5.462 0.492 4.210 0.495 2.926 0.496 4.840
0.496 5.795 0.498 4.287 0.499 3.001 0.498 4.810

Data are from Dokos et al. (2002, Fig. 6) whose raw values were supplied to the
author with permission granted for use by Prof. Ian J. LeGrice, University of
Auckland, Auckland, New Zealand
a Samples 3  3  3 mm in size were subjected to simple shearing over a range in
shear of  = 0.5 under a sinusoidal waveform with a 30 s period. Their frame
of reference is: F aligns with the direction of the muscle bers, S is transverse
to F and lies within the sheet of muscle bers, and N is normal to this sheet.
Six shearing orientations were exercised on the same tissue sample, designated by
pairs, e.g., FS. The rst character, F in this case, denes the direction normal to
the shearing plane (i.e., direction 2 in Fig. 1.5). The second character, S in this
case, denes the direction of shearing (i.e., direction 1 in Fig. 1.5)
Appendix A

Linear Algebra

This appendix provides an overview of linear algebra. Coverage of this


topic is only sufficient to comprehend and understand the contents of this
textbook. It is not meant to be a substitute for a course on the topic nor a
good handbook.

A.1 Arrays and Vectors

Vectors are arrays with physical units that obey a linear transformation rule
between coordinate frames, like the pushforward and pull-back operators
discussed in Eqs. (B.7)(B.14) of Appendix B.
Several vector notations are used throughout this text, some being
illustrated in this appendix. To help keep things as compact as possi-
ble, two-dimensional (2D) vectors are considered, with their extension to
three dimensions (3D) being straightforward. A Cartesian metric tensor
g = ij e i e j is used throughout this text, wherein ij denotes the Kro-
necker delta: 1 if i = j, 0 otherwise.
Consider two vectors U and V defined as
(  
Ui ei , U1
U = {U } = , (A.1)
Ui e i , U2

and
(  
V i ei , V1
V = {V } = (A.2)
Vi e i , V2

A.D. Freed, Soft Solids: A Primer to the Theoretical Mechanics of Materials, 277
Modeling and Simulation in Science, Engineering and Technology,
DOI 10.1007/978-3-319-03551-2, Springer International Publishing Switzerland 2014
278 Soft Solids

where ei and e i are the unit base vectors in the ith coordinate direction with
i = 1, 2 whose components are
   
1 0
e1  e 1 = and e2  e 2 = , (A.3)
0 1
where ei  e i in Cartesian tensor analysis; they dier in general tensor
analysis. Vector components Ui and Ui , say, possess dierent physical
properties, which is the topic of Appendix B. These two vectors add ac-
cording to
(  
(Ui + V i ) ei , U1 + V1
U +V = or {U + V } = , (A.4)
(Ui + Vi ) e i , U2 + V2

and they subtract according to


(  
(Ui V i ) ei , U1 V1
U V = or {U V } = , (A.5)
(Ui Vi ) e i , U2 V2

where it is noted, e.g., that Ui Vi is not defined. Vector addition obeys


U +V = V +U commutative, (A.6)
U + (V + W ) = (U + V ) + W associative. (A.7)
Multiplication by a scalar c implies that
(    
cUi ei , U1 cU1
cU = or c{U } = c = (A.8)
cUi e i , U2 cU2

and, as such, scalar multiplication is


c(U + V ) = cU + cV distributive. (A.9)
A zero vector exists and is denoted as
(  
0 i ei , 0
0= or {0} = . (A.10)
0i e i, 0
This completes the definitions needed to establish a vector algebra.
In the above notation, observe that indexers, e.g., the i in Ui ei , are
given Latin characters located in either a subscript or superscript position.
They appear in pairs that are summed over in accordance with Einsteins
P
summation convention, e.g., Ui ei means 3i=1 Ui ei = U1 e1 + U2 e2 + U3 e3
(in 3-space), with one Latin index being a superscript and the other Latin
Linear Algebra 279

index being a subscript. On the other hand, elements of a vector, e.g., U1 ,


are given numerical characters that are always positioned as subscripts.
Two vectors contract to a scalar by a process called the dot product
U  V = Ui Vi = Ui V i = U1 V1 + U2 V2 , (A.11)
which also goes by the terminology of the inner product between two vec-
tors. The Euclidean norm, or magnitude, of a vector is defined as
p q q q
kU k = U  U = U ij U = Ui U j = U21 + U22
i j ij
(A.12)
and is equivalent to the Pythagorean description for distance. Here ij and
ij arise as the metrics of Cartesian space. Properties (A.4)(A.10) and
(A.12) constitute a normed vector space, a.k.a. a Hilbert space.
Two vectors can be multiplied via a dyadic product (cf. Holzapfel 2000,
p. 10) so as to produce a matrix valued quantity called a tensor; specifically,
8 i j

U V ei e j ,

<Ui V e e j , 
j i U1 V1 U1 V2
U V = where [U V ] = . (A.13)

Ui V e e j ,
j i U2 V1 U2 V2
:
Ui V j e i e j ,
This goes by the terminology of the outer (dyadic or tensor) product be-
tween two vectors, where ei e j forms a basis for a tensor wherein the
first base vector associates with the row and the second base vector asso-
ciates with the column.

A.2 Matrices and Tensors

Tensors are matrices with physical units that obey a linear transformation
rule between coordinate frames, like the pushforward and pull-back oper-
ators discussed in Eqs. (B.15)(B.26) of Appendix B.
Consider two tensors M and N defined by1
8

<M ei e j ,
ij 
M11 M12
M = Mij ei e j , with [M ] = (A.14)
: M21 M22
Mij e i e j ,
j
1 There are two mixed tensors in general tensor analysis, e.g., Mij and Mi (cf.
Truesdell and Noll 2004). However, in Cartesian tensor analysis, one variant for
a mixed tensor suces. It is being denoted as simply Mij .
280 Soft Solids

and
8

<N ei e j ,
ij 
N11 N12
N = Nij ei e j , with [N ] = (A.15)
: N21 N22
Nij e i e j ,
where the i index associates with the row and the j index associates with the
column. Again, 2D fields are used to help keep things concise. No issues
arise when extending the following concepts into 3D. The above tensors
add according to
8
<(M + N ) ei e j ,
ij ij

M + N = (Mij + Nij ) ei e j , (A.16)


:
(Mij + Nij ) e e ,
i j

with

M11 + N11 M12 + N12
[M + N ] = , (A.17)
M21 + N21 M22 + N22
and they subtract according to
8
<(M N ) ei e j ,
ij ij

M N = (Mij Nij ) ei e j , (A.18)


:
(Mij Nij ) e i e j ,
with

M11 N11 M12 N12
[M N ] = , (A.19)
M21 N21 M22 N22
where, e.g., Mij Nij is not defined. Matrix addition obeys
M +N =N +M commutative, (A.20)
L + (M + N ) = (L + M ) + N associative. (A.21)
Multiplication by a scalar c implies that
8

<cM ei e j ,
ij

cM = cMij ei e j , (A.22)
:
cMij e i e j ,
with
 
M11 M12 cM11 cM12
c[M ] = c = (A.23)
M21 M22 cM21 cM22
Linear Algebra 281

and, as such, scalar multiplication is


c(M + N ) = cM + cN distributive. (A.24)
A zero tensor exists and is denoted as
8

<0 ei e j ,
ij 
00
0 = 0 ij ei e j , or [0] = . (A.25)
: 00
0 ij e i e j ,
This completes the definitions needed to establish a tensor algebra.
The transpose of a tensor is achieved by swapping its o-diagonal ele-
ments so that
8

<M ei e j ,
ji 
T `i

T M11 M21
M = jk M` ei e , or M =
k j , (A.26)
: M12 M22
Mji e e ,
i j

where component notation for the transpose of a mixed tensor is not


straightforward.2 This allows a tensor to be decomposed into a symmetric
part
8 
1 Mij + Mji  e e ,
  < 2 i
 j
sym(M ) = 12 M + M T = 12 Mij + jk Mk` `i ei e j , (A.27)
: 1   i
2 Mij + Mji e e ,
j

with
" #
M11 1 (M + M )
2 12 21
[sym(M )] = 1 (M + M ) (A.28)
2 21 12 M 22

and a skew (or antisymmetric) part


8 
1 Mij Mji  e e ,
  < 2 i
 j
skew(M ) = 12 M M T = 12 Mij jk Mk` `i ei e j , (A.29)
: 1   i
2 M ij Mji e e j,

2 The transpose of a mixed tensor, e.g., M T = jk Mk` `i ei e j , has component


indices that are pulled up and pushed down by the metric of the space (cf. Marsden
and Hughes 1983, p. 48). The metric tensor belonging to a Cartesian coordinate
frame has elements described by the Kronecker delta. The transpose of a matrix,
denoted as [M ]T , is distinct from the transpose of a tensor, denoted as M T ,
especially for mixed tensors, in which case [M T ] [M ]T , in general.
282 Soft Solids

with
" #
0 1 (M M )
2 12 21
[skew(M )] = 1 (M M ) , (A.30)
2 21 12 0
so that
M = sym(M ) + skew(M ) (A.31)
where
sym(M T ) = sym(M ) and skew(M T ) = skew(M ). (A.32)
In three-dimensional analysis it is at times advantageous to express a skew
tensor, say M , as a vector, say m, via
2 3
0 m3 m2 Mij =  ijk mk ,
[M ] = 4m3 0 m1 5 with, e.g., (A.33)
m2 m1 0 Mij = ijk mk ,

where ijk and  ijk denote the permutation symbol whose values are : 1 if
indices ijk are unique and cyclically ordered, i.e., 123, 231, or 312; 1 if
they are unique with a reverse cyclic ordering, viz., 132, 213, or 321; or 0
whenever two or three indices have the same indical value.
A tensor can contract with a vector, producing another vector. This is a
kind of linear mapping, i.e., a transformation matrix, and is defined by
8

<M V j ei ,
ij

U = M V = Mij V j ei , (A.34)
:
Mij V e ,
j i

with
 
M11 V1 + M12 V2
{U } = [M ]{V } = . (A.35)
M21 V1 + M22 V2
A dierent contraction occurs when
8

<M V j ei ,
ji

U = M T V = ij Mkj k` V` ei , (A.36)
:
Mji V j e i ,
with
 

M11 V1 + M21 V2
{U } = M T {V } = , (A.37)
M12 V1 + M22 V2
Linear Algebra 283

which is sometimes written as V T M in the literature, although that nota-


tion is not used here. The vector outcomes are dierent for these two prod-
ucts. These contractions can take on several forms in component notation
depending upon whether a particular index is a subscript or a superscript,
whose origin is the topic of Appendix B.
A special tensor maps a vector back into itself. It is the identity tensor
and is described by, e.g.,
(
ij V j ei ,
V = IV = j (A.38)
i V j ei,
where

10
I= ij ei e j = ei e i
or [I] = (A.39)
01
where components of the Cartesian metric I have values assigned by the
Kronecker delta ij , which are 1 if i = j and 0 otherwise.
Two tensors can be multiplied together in a variety of ways, too, and
a variety of notations exist in the literature for this purpose. As above, a
simple notation has been adopted here. The dot or inner product between
two tensors is expressed as
(
Mik Nkj ,
T = MN with, e.g., Tij = (A.40)
Mik Nkj ,
whose matrix notation is

M11 N11 + M12 N21 M11 N12 + M12 N22
[T ] = [M ][N ] = . (A.41)
M21 N11 + M22 N21 M21 N12 + M22 N22
An alternative, but distinct, inner product occurs for
(
Mki Nkj ,
T = M T N with, e.g., Tij = ik ` (A.42)
Mk `m Nmj ,
whose matrix notation is


M11 N11 + M21 N21 M11 N12 + M21 N22
[T ] = M T [N ] = , (A.43)
M12 N11 + M22 N21 M12 N12 + M22 N22
while another form occurs for
(
Mik Njk ,
T = MN T with, e.g., Tij = (A.44)
Mik k` Nm` mj ,
284 Soft Solids

whose matrix notation is




T M11 N11 + M12 N12 M11 N21 + M12 N22
[T ] = [M ] N = (A.45)
M21N11 + M22 N12 M21 N21 + M22 N22
where the tensor outcomes are dierent in all three of these products. These
are all contractions over a single index. In Eqs. (A.40), (A.42), and (A.44),
tensor T = Tij ei e j is taken to be of mixed character. Like expressions
follow whenever T = Tij ei e j or T = Tij e i e j .
Matrix multiplication obeys
L(MN ) = (LM )N associative, (A.46)
L(M + N ) = LM + LN distributive, (A.47)
but, in general, matrix multiplication does not commute, viz.,
MN NM . (A.48)
The fact that matrix multiplication does not commute has major conse-
quences and implications in applications. Some matrices do commute, and
this special property makes them useful in certain applications, e.g., the
Kirchho stress s and the Finger deformation b tensors commute within
the mathematical structure of the explicit theory of elasticity; cf. Chap. 5.
Commuting matrices were also used in the derivation of the integrator pre-
sented in Appendix C (Butcher 2008). Nevertheless, commuting matrices
are a rare occurrence in mechanics.
Contractions can also occur over two indices, with such summations
reducing to scalar fields. When just one tensor field is involved, this oper-
ation is called the trace and is given by
tr M = Mii = M11 + M22 , (A.49)
which is one of two invariants of M . (Three invariants exist for a tensor
field in 3-space.) The trace is defined for tensor arguments of mixed char-
acter. Whenever there is a single argument in a trace, e.g., tr(M ), it will
often be written simply as tr M . A double contraction can exist whenever
two tensors are involved, specifically
8
ji
<M Nji ,
M : N = tr(M N ) = ik M`k `j Nij ,
T
(A.50)
:
Mji Nji ,
whose matrix notation is
M : N = M11 N11 + M21 N21 + M12 N12 + M22 N22 , (A.51)
Linear Algebra 285

where
p q
kM k = M : M = M211 + M221 + M212 + M222 (A.52)
is the Frobenius norm, which is the tensor analog to the Euclidean norm
defined in Eq. (A.12) for vectors.
The inner and outer dyadic products between two matrices, e.g., M N
and M N , are defined in Appendix C.
The determinant is the second invariant of tensor M in 2-space and is
given by

 i  M11 M12
det M = det M j = = M11 M22 M21 M12 , (A.53)
M21 M22
while for a tensor in 3-space, say T , it is given by

T22 T23 T21 T23 T21 T22
det T = T11
T12
+ T13 , (A.54)
T32 T33 T31 T33 T31 T32
wherein the 2  2 determinants are cofactors Cij of det T given by (1)i+j
times its associated minor determinant (obtained by removing the ith row
and jth column from T ). The determinant is defined for tensor arguments
of mixed character (Truesdell and Noll 2004, p. 15). A tensor is said to be
singular if its determinant is zero; otherwise, it is non-singular.
The determinant and its cofactors play an important role in quantifying
the inverse of a tensor field. For a tensor in 2D, its inverse is given by
 
1 1 C11 C21 1 M22 M12
[M ] = = , (A.55)
det M C12 C22 det M M21 M11
while for a tensor in 3D, its inverse is given by
2 3
C11 C21 C31
1 4C12 C22 C32 5
[T 1 ] =
det T C C C
13 23 33
2 3
T T T32 T23 T32 T13 T12 T33 T12 T23 T22 T13
1 4 22 33
= T31 T23 T21 T33 T11 T33 T31 T13 T21 T13 T11 T23 5 , (A.56)
det T T T T T T T T T T T T T
21 32 31 22 31 12 11 32 11 22 21 12

which will exist if their respective matrices are not singular. A more con-
densed way to express the tensor inverse is via the expression
adj M
M 1 = provided that det M 0 (A.57)
det M
286 Soft Solids

where adj M is the adjoint of M , which is the transpose of the matrix of


cofactors.
Orthogonal matrices (cf. Appendix B) have a special property, viz.,
QQT = I or, in other words, Q1 = QT . (A.58)
A direct consequence of this requirement is that
det Q = 1, (A.59)
with the condition that det Q = 1 being referred to as a proper orthogonal
tensor. Orthogonal tensors are rotation maps that, in 2D, have components

cos( ) sin( )
[Q] = (A.60)
sin( ) cos( )
for some angle of rotation  , with a positive rotation occurring counter-
clockwise, i.e., from the positive 1-axis towards the positive 2-axis about
the positive 3-axis. A proper orthogonal tensor maps a right-handed co-
ordinate frame into another right-handed frame, whereas a non-proper or-
thogonal tensor (i.e., its determinant is 1) maps a right-handed frame into
a left-handed one, or vice versa.
Some useful properties of tensor fields, irrespective of dimension, in-
clude:
MM 1 = M 1 M = I, (A.61)
(M 1 )1 = M, (A.62)
(MN )1 = N 1 M 1 , (A.63)
1 1
(cM )1 = M , (A.64)
c
M 2 = M 1 M 1 , (A.65)
T T
(M ) = M , (A.66)
(MN )T = N T M T , (A.67)
1 T T 1 T
(M ) = (M ) =M , (A.68)
tr(M T ) = tr M , (A.69)
tr(MN ) = tr(NM ), (A.70)
tr(M + N ) = tr(M ) + tr(N ), (A.71)
tr(cM ) = c tr M , (A.72)
det(M T ) = det M , (A.73)
Linear Algebra 287

det(MN ) = det(M ) det(N ), (A.74)


1
det(M 1 ) = , (A.75)
det M
det(cM ) = cd det M (A.76)
where d is the dimension of the space, which will be either 2 or 3 in our
applications.
Some useful gradients that arise during constitutive construction are:
@ tr M
= I, (A.77)
@M
@ tr(M 2 ) @ tr(MM ) @M TW M
= = = 2M T , (A.78)
@M @M @M
@ tr(MM T ) @ tr(M T M ) @M W M
= = = 2M , (A.79)
@M @M @M
@ det M
= det(M ) M T , (A.80)
 @M 
@ ln det(M )
= M T , (A.81)
@M
@ tr(M 1 )
= M T M T , (A.82)
@M
d det M
= det(M ) tr(M 1 dM /dt), (A.83)
 dt 
d ln det(M )
= tr(M 1 dM /dt). (A.84)
dt
Specifically, these gradients can occur because of the thermodynamic po-
tential functions, out of which constitutive equations are derived.
The CayleyHamilton theorem states: Any square matrix satisfies its
own characteristic equation. When applied to a 2D tensor field, say M ,
this implies that
M 2 IM + III = 0 (A.85)
where
I(M ) = tr M and II(M ) = det M (A.86)
are the two invariants of the 2D tensor M .
The same CayleyHamilton theorem, when applied to a 3D tensor field,
say T , requires that
T 3 IT 2 + IIT IIII = 0 (A.87)
288 Soft Solids

where
I(T ) = tr T ,
 
II(T ) = 12 tr(T )2 tr(T 2 ) , (A.88)
III(T ) = det T
so now there are three invariants. This theorem is useful for the purpose
of expressing a higher-order power of a tensor field in terms of a polyno-
mial whose order is no greater than the dimension of the matrix and whose
coecients are restricted to be scalar functions of its invariants.
To illustrate how one can use the CayleyHamilton theorem, consider
the exponential of a tensor defined by its Taylors series as
exp T = I + T + 2! 1 T2 + 1 T3 + 1 T4 + 1 T5 +:::
3! 4! 5!
that, because of the CayleyHamilton theorem, can be approximated as

exp T  1 + 3! 1 III + 1 I III + 1 III I 2 II  I
4! 5!

+ 1 3!1 II 1 I II III  1 II(I 2 II) I III  T

4! 5!

+ 2! 1 + 1 I + 1 I 2 II  + 1 I 3 2I II + III  T 2 ,

3! 4! 5!
which has been truncated at the sixth term in the expansion.
Appendix B

Covariant and Contravariant Issues:


Conguration Physics

The reasons for selecting one tensor field over another, when multiple
choices exist, can be rationally explained, but it requires more mathe-
matical capability than the typical reader of this text would be expected
to have. That is why the authors reasoning is outlined here, in an ap-
pendix. Like the body of this text, full mathematical rigor is not pre-
sented here; rather, continuity in the flow of ideas and thought processes is
sought. You can get a deeper understanding of the topics covered in this
appendix from textbooks that address general tensor analysis, e.g., Chad-
wick (1976), Holzapfel (2000), Lodge (1964, 1974), Marsden and Hughes
(1983), Sokolniko (1964), Truesdell and Noll (2004), and Truesdell and
Toupin (1960), some of which address applications in the general theory
of relativity, too.
Vectors can map from one configuration into another in one of two
ways: as a covariant vector or as a contravariant vector. Normals to sur-
faces are examples of covariant vectors, e.g., Eq. (2.11), while tangents to
curves are examples of contravariant vectors, e.g., Eq. (2.2). The need to
distinguish between these two vector types arises in two important con-
siderations when using Cartesian tensor analysis: in the transfer of field
from one configuration into another and when taking time derivatives or
integrals in the spatial configuration. Their importance in general tensor
analysis is much more profound.
It is absolutely essential that the reader becomes familiar with the con-
cepts and ideas presented in this appendix. The physics described here
prepare the canvas upon which a theory for the mechanics of materials can
then be painted.

A.D. Freed, Soft Solids: A Primer to the Theoretical Mechanics of Materials, 289
Modeling and Simulation in Science, Engineering and Technology,
DOI 10.1007/978-3-319-03551-2, Springer International Publishing Switzerland 2014
290 Soft Solids

{v}

A B
[Q]
e3 e2
e1
e2 e3

[Q1]
e1

Fig. B.1 Invariance of an observer: observer A views event v as a vector {v} in a


frame of reference (e1 , e2 , e3 ) with components {v1 v2 v3 }T , while observer B views
the same event v from a dierent frame of reference, i.e., (e10 , e20 , e30 ), producing
a dierent set of components, viz., {v10 v20 v30 }T , describing vector {v0 }. These
two sets of components relate to one another via the formula v 0 i = Q j0 i v j or,
equivalently, via the formula v j = [Q1 ] 0 i v 0 i where matrix [Q1 ] 0 i is the inverse to
j j

matrix Q j0 i . These formul are expressed more simply in their matrix notations
{v0 } = [Q] {v} and {v} = [Q1 ] {v0 }

B.1 Invariant Observer

A physical field must not depend upon the frame in which an observer
observes an event and in which the field is to be quantified. Said dierently,
consider a scenario where observer A observes physical event v from a
coordinate frame of (e1 , e2 , e3 ), while observer B observes the same event
but from a dierent perspective, viz., from some other frame (e10 , e20 , e30 ),
as illustrated in Fig. B.1. The numerical values they assigned to this event
will likely be dierent, yet the event was the same. How can this be? If the
same event was observed by the two observers, then it stands to reason that
the physical fields being measured ought to be the same. Consequently, the
cause of this quandary must be in how their separate values associate with
their individual coordinate frames of reference.
Investigating further, one finds that these two observers coordinate
frames relate to one another via the linear transformations
{e10 } = [Q] {e1 }, {e20 } = [Q] {e2 }, {e30 } = [Q] {e3 },
(B.1)
{e1 } = [Q1 ] {e10 }, {e2 } = [Q1 ] {e20 }, {e3 } = [Q1 ] {e30 }
where
j
[Q] = Q 0j i ei0 e j and [Q1 ] = [Q1 ] 0 i e j e 0 i (B.2)
Covariant and Contravariant Issues: Configuration Physics 291

with [Q] [Q1 ] having components Q 0ki [Q1 ] k0j = 00ji and [Q1 ] [Q]
having components [Q1 ] i0 k Q 0j k = ij and, therefore, [Q] [Q1 ] =
[Q1 ] [Q] = [I]. For the matrix array Q 0j i , index 0 i is the row indexer
j
and index j is the column indexer. For the matrix array [Q1 ] 0 i , index j is
the row indexer and index 0 i is the column indexer. Matrix [Q] is orthog-
onal because its inverse [Q1 ] = [Q]1 equals its transpose [Q]T , while
its determinant is det Q = 1. The determinant will be +1 whenever both
bases have the same handedness and 1 whenever one basis is left handed
and the other is right handed.
A physical vector, say v, is described by an array {v} whose elements
v i associate with a basis, say (e1 , e2 , e3 ), that when described in another
basis, say (e10 , e20 , e30 ), has elements v 0 i that obey a linear mapping, i.e.,
v = v i ei = v 0 i ei0 (B.3)
has elements that obey
v 0 i = Q 0j i v j , {v0 } = [Q] {v},
or (B.4)
j
v j = [Q1 ] 0 i v 0 i , {v} = [Q]T {v0 },
because [Q1 ] = [Q]T . Here notation assists us in our understanding that
it is not the physical field v that is dierent, as measured by our two ob-
servers; rather, it is their frame of reference that diers. Furthermore, if
one knows how to map between these two frames of observation, then ei-
ther observer can transform their measurements into the others frame of
reference, at which point their results should agree, within experimental
error.
A physical tensor, say T , is described by a matrix [T ] whose elements
T ij associate with a basis, say (e1 , e2 , e3 ), that when described in another
basis, say (e10 , e20 , e30 ), has elements T 0 ij that obey a linear mapping, i.e.,
T = T ij ei e j = T 0 ij ei0 e j0 (B.5)
with
0j
T 0 ij = Q 0ki T k` Q ` , [T 0 ] = [Q] [T ] [Q]T ,
or (B.6)
j
T ij = [Q1 ] i0 k T 0 k` [Q1 ] 0 ` , [T ] = [Q]T [T 0 ] [Q],
because [Q1 ]T = [Q]. Here again the notation assists us in our under-
standing that a tensor and its components are distinct ideas, although both
are representations of the same physical entity.
292 Soft Solids

B.2 Field Transfer

Field transfer is an operation by which a vector or tensor field is mapped


from one configuration into another (Lodge 1964, 1974). In the formul
that follow, field transfer takes place between a reference configuration
0 where the fields depend upon material coordinates X = XI eI and a
current configuration  where the fields depend upon spatial coordinates
x = x i ei . Both configurations are taken to associate with the same triad of
base vectors (e1 , e2 , e3 ) in this text, thereby keeping the ideas of the prior
section separate from those in this section.
Field transfer is a process whereby a tensor field from one configuration
is mapped into another configuration. Tensor equations, on the other hand,
are defined over a single configuration. Therefore, formul that describe
field transfer are not tensor equations. It is for this reason that field-transfer
operations are written as matrix equations, e.g., {dx} = [F] {dX }, instead
of as tensor equations, in this case dx = F dX .
Each and every configuration is aliated with its own normed vector
space, i.e., a Hilbert space, and, therefore, an algebra is available for the
physical fields belonging to each configuration, but no algebra is avail-
able that spans the physical spaces of all configurations. Consequently,
arithmetic, algebra, and the calculus are all available to the mechanician
whenever operations are done within a single configuration, but not when
multiple configurations are involved, at least not without special consid-
erations. This important yet subtle consequence comes from the Hilbert
spaces upon which mechanics is built (Freed 2010; Lodge 1964, 1974).
For example, one is not permitted to directly add a physical vector from
0 with another physical vector from . One must first map one of these
two vectors into the others configuration, at which point both will belong
to same configuration and, therefore, to the same Hilbert space where addi-
tion is a defined operation. All mappings between configurations are done
in ambient space (e1 , e2 , e3 ), i.e., in our world or universe, within which
all physical configurations are embedded.
Contravariant vectors in  and 0 of g(x, t) = g i (x, t) ei and
G (X , t) = GI (X , t) eI transfer their fields according to the mappings
@x i I
g i (x, t) = G (X , t), (B.7)
@XI
@XI i
GI (X , t) = g (x, t) (B.8)
@x i
Covariant and Contravariant Issues: Configuration Physics 293

[F]

@ t0 @t

0
[F1]

Fig. B.2 Field-transfer operations for a contravariant vector eld, e.g., a tangent
to a curve; cf. Eqs. (2.2) and (2.3). Matrix [F] is the pushforward map; matrix
[F 1 ] is the pull-back map

that when expressed in matrix notation become


{g(x, t)} = [F] {G (X , t)} pushforward, (B.9)
{G (X , t)} = [F 1 ] {g(x, t)} pullback, (B.10)
as shown schematically in Fig. B.2, and where [F 1 ] = [F]1 .
Covariant vectors in  and 0 of g(x, t) = g i (x, t) e i and G (X , t) =
GI (X , t) e I transfer their fields according to the mappings
@XI
g i (x, t) = GI (X , t)
, (B.11)
@x i
@x i
GI (X , t) = g i (x, t) I (B.12)
@X
that when expressed in matrix notation become
{g(x, t)} = [F 1 ]T {G (X , t)} pushforward, (B.13)
T
{G (X , t)} = [F] {g(x, t)} pullback, (B.14)
as shown schematically in Fig. B.3.
Vectors are expressed in a slanted blackboard bold font. Vectors de-
fined over 0 are expressed in uppercase, while vectors defined over  are
expressed in lowercase. Contravariant vectors have a component indexer in
the superscript location, while covariant vectors have a component indexer
in the subscript location. The deformation gradient F has a component
representation of FIi where i points to the row and I points to the column in
its matrix representation. Its inverse F 1 has a component representation
of [F1 ]Ii where now I points to the row and i points to the column.
294 Soft Solids

[F1]T

@ t0 @t

0 [F]T

Fig. B.3 Field-transfer operations for a covariant vector eld, e.g., a normal to
a surface; cf. Eq. (2.11). Matrix [F 1 ]T is the pushforward map; matrix [F]T is
the pull-back map

Note: [F]T {g} F Tg; they are distinct. From Eq. (B.14), [F]T {g}
has components FIi g i , whereas, from Eq. (A.36), F Tg has a component
j
representation of IJ FJ ji g i . The indical locations of index i on vector
g are dierent. The former is a matrix operation; the latter is a tensor
operation. The transfer of field from one configuration into another is a
matrix operation. It is a mapping between two Hilbert spaces. Tensor
operations are only defined within a single Hilbert space. This is why these
j
two notations have been introduced. In general tensor analysis, IJ FJ ji
j
would become [G1 ]IJ FJ g ji wherein [G1 ]IJ denotes the components of
the contravariant metric G 1 of 0 , while g ij denotes the components of
the covariant metric g of . This distinction arises whenever curvilinear
coordinates are employed.
Contravariant tensors in  and 0 of g(x, t) = g ij (x, t) ei e j and
G (X , t) = GIJ (X , t) eI eJ transfer their fields according to the mappings
@x i @x j IJ
g ij (x, t) = G (X , t), (B.15)
@XI @XJ
@XI @XJ ij
GIJ (X , t) = g (x, t) (B.16)
@x i @x j
that in matrix notation become
[g(x, t)] = [F] [G (X , t)] [F]T pushforward, (B.17)
[G (X , t)] = [F 1 ] [g(x, t)] [F 1 ]T pullback. (B.18)
Covariant and Contravariant Issues: Configuration Physics 295

Matrix symmetry of a contravariant tensor, if present, is preserved when-


ever its field is transferred from one configuration into another. However,
the inverse of a contravariant tensor, if it exists, maps as a covariant tensor.
Covariant tensors in  and 0 of g(x, t) = g ij (x, t) e i e j and
G (X , t) = GIJ (X , t) e I e J transfer their fields according to the mappings
@XI @XJ
g ij (x, t) = GIJ (X , t) , (B.19)
@x i @x j
@x i @x j
GIJ (X , t) = g ij (x, t) I (B.20)
@X @XJ
that in matrix notation become
[g(x, t)] = [F 1 ]T [G (X , t)] [F 1 ] pushforward, (B.21)
[G (X , t)] = [F]T [g(x, t)] [F] pullback. (B.22)
Matrix symmetry of a covariant tensor, if present, is preserved whenever
its field is transferred from one configuration into another. However, the
inverse of a covariant tensor, if it exists, maps as a contravariant tensor.
Mixed tensors in  and 0 have one contravariant index and one
covariant index1 with dyadic notations g(x, t) = g ji (x, t) ei e j and
G (X , t) = GIJ (X , t) eI e J whose fields transfer according to the map-
pings
@x i I @XJ
g ji (x, t) =G (X , t) , (B.23)
@XI J @x j
@XI i @x j
GIJ (X , t) = g (x, t) (B.24)
@x i j @XJ
that in matrix notation become
[g(x, t)] = [F] [G (X , t)] [F 1 ] pushforward, (B.25)
[G (X , t)] = [F 1 ] [g(x, t)] [F] pullback. (B.26)
The inverse of a mixed tensor, if it exists, is preserved whenever its field is
transferred from one configuration into another, i.e., the inverse of a mixed
tensor maps as a mixed tensor. In contrast, matrix symmetry of a mixed

1 What is referred to as a mixed tensor here is often referred to as a right-


covariant mixed tensor in the literature. That is because they introduce a left-
covariant mixed tensor that we do not. There is no need for such elds in Carte-
sian tensor analysis within the discipline of continuum mechanics.
296 Soft Solids

Table B.1 Field-transfer mappings between the material 0 and spatial


 congurations for some of the more common tensor elds used herein

Type Material Spatial Field Dened in


of eld tensor tensor maps as equation(s)
Deformation I b Contravariant Eq. (3.6)
Deformation I b1 Covariant Eq. (3.8)
Deformation C I Covariant Eq. (3.1)
Deformation C 1 I Contravariant Eq. (3.4)
Stretch U v Mixed Eq. (2.13)
Velocity gradient L l Mixed Eqs. (2.19) and (2.24)
Strain E e Covariant Eqs. (3.10) and (3.11)
Strain E e Contravariant Eqs. (3.13) and (3.14)
O
Strain rate EP ed Covariant Eqs. (3.36) and (3.37)
M
Strain rate EP ed Contravariant Eqs. (3.39) and (3.40)
Stress S s Contravariant Eqs. (4.4) and (4.18)
M
Stress rate SP s Contravariant Eq. (4.21)
Extra stress  Contravariant Eq. (4.25)
M
Extra-stress rate P  Contravariant Eqs. (4.28) and (4.29)

tensor, if present, is not preserved by its field-transfer law. For example,


the transpose of matrix equation (B.25) is [g]T = [F 1 ]T [G ]T [F]T , which
is the field-transfer map of a left-covariant mixed tensor (our mixed tensors
are right covariant). Matrix [g] is said to by symmetric whenever [g]T =
[g]. Tensor g is said to be symmetric whenever Eq. (A.27) applies.
In the case of tensors, the appropriate pushforward and pull-back op-
erations of Figs. B.2 and B.3 are applied twice, once to each index of the
field being transferred. One-state tensor fields are expressed in a slanted
bold font. Like vectors, tensors defined over 0 are expressed in upper-
case, while tensors defined over  are expressed in lowercase. Two-state
tensor fields that serve as maps between configurations 0 and , like F,
are expressed in an upright bold font and usually in uppercase.
By pushforward one means that the field of interest is being pushed
from 0 forward into , and by pullback, one means that the field of inter-
est is being pulled from  back into 0 . A collection of the more common
field-transfer maps between the material 0 and spatial  configurations
of tensor fields used throughout this text is presented in Table B.1.
From the polar decomposition of F, given in Eq. (2.13), the field trans-
fer operator F has two parts: one that is due to a rotation between the two
configurations, akin to the discussion of the prior section, and another that
Covariant and Contravariant Issues: Configuration Physics 297

is due to a deformation or shape change that takes place over the history or
path that has been traversed between these two configurations.

B.3 Material Derivatives

The process is straightforward for taking the time derivative of a field at-
tached to a particle, say at P in body B, whenever that field is expressed
in terms of material coordinates X (i.e., it is a Lagrangian field). In this
case, the coordinates associate with the fixed particle P of interest, so the
time rate of change of a material field is just its partial derivative taken with
respect to time. For material vector fields, one has
@G (X , t)
GP (X , t) = , (B.27)
@t
while for material tensor fields, one has
@G (X , t)
GP (X , t) = , (B.28)
@t
with there being no need to distinguish between contravariant, covariant,
or mixed fields for these operations.
These rates are obvious. It turns out, though, that physically admissi-
ble time derivatives are not so intuitive whenever spatial coordinates are
selected for use.
To take the time derivative of a field attached to particle P originally
located at coordinates X but whose field of interest is expressed in terms
of spatial coordinates x = (X , t), not material coordinates X (i.e., it is an
Eulerian field), requires that one pay more attention to the details. The time
derivative of a spatial field taken at a fixed particle is called its material
derivative and is denoted by D/Dt. The material derivative of a spatial
vector field is given by
Dg(x, t) @g(x, t) @g(x, t) @x
= + 
Dt @t @x @t (B.29)
= g(x,
P t) + grad g(x, t)  v(x, t)
of which the acceleration vector in Eq. (1.9) is an example, while the ma-
terial derivative for a spatial tensor field is given by
Dg(x, t) @g(x, t) @g(x, t) @x
= + 
Dt @t @x @t (B.30)
P
= g(x, t) + grad g(x, t)  v(x, t)
298 Soft Solids

with no distinction being required between contravariant, covariant, or


mixed fields for either type of operator. The gradient of a vector field
yields a tensor field that, when contracted with another vector field, pro-
duces a vector field, e.g., grad(g)  v = (@g i /@x j ) v j ei , while, whenever g
is a tensor, grad(g)  v = (@g ij /@x k ) v k ei e j produces a tensor.
The material derivative of a scalar field is objective, i.e., invariant of
frame. However, except for the velocity and acceleration fields, the mate-
rial derivative of a vector field is not an objective rate nor is the material
derivative of a tensor field an objective rate. Taking the material derivative
of such fields does not provide a complete description for these fields time
rates of change because the material derivative is not a frame-invariant op-
erator for any type of field other than (absolute) scalars. Specifically, non-
physical rotations can enter into a solution. To correct this deficiency, one
must also take into account the motion of a bodys configuration as it moves
through space. After all, the current configuration is translating, rotating,
and stretching, and all these rates of change need to be accounted for.

B.4 Lie Derivatives

The Lie derivative taken along the path of motion of some particle P , often
referred to as an Oldroyd (1950, 1970) derivative, is an objective measure
for the time rate of change of spatial fields. An objective measure is one
that does not introduce any unwanted rotational eects into its outcome.
Objective rates are physically admissible measures for quantifying time
rates of change of a field in the spatial configuration .
A recipe for computing the Lie derivative of any spatial tensor field is
fairly straightforward (cf. Holzapfel 2000, p. 106):

1. Pull back a spatial field of interest in  into its corresponding material


field in 0 by using its appropriate mapping, e.g., Eqs. (B.8), (B.12),
(B.16), (B.20), or (B.24).
2. Take the partial derivative of its mapped material field with respect to
time at the fixed particle P in 0 .
3. Push forward the derivative of the material field achieved in 0 into
its corresponding spatial field in  by using its appropriate mapping,
e.g., Eqs. (B.7), (B.11), (B.15), (B.19), or (B.23).
Covariant and Contravariant Issues: Configuration Physics 299

With this methodology in place, objective derivatives of contravariant


M M
P I (X , t) eI transfer their fields
vectors g(x, t) = g i (x, t) ei and GP (X , t) = G
according to the mappings
Mi @x i P I
g (x, t) = G (X , t), (B.31)
@XI
P I (X , t) = @X M
I
G g i (x, t) (B.32)
@x i
that when written in matrix notation become
M
{g(x, t)} = [F] {GP (X , t)} pushforward, (B.33)
M
{GP (X , t)} = [F 1 ] {g(x, t)} pullback. (B.34)
O O
Objective derivatives of covariant vectors g(x, t) = g i (x, t) e i and
P I (X , t) e I transfer their fields according to the mappings
GP (X , t) = G
O @XI
g i (x, t) = GP I (X , t)
, (B.35)
@x i
P I (X , t) = O @x i
G g i (x, t) I (B.36)
@X
that when written in matrix notation become
O
{g(x, t)} = [F 1 ]T {GP (X , t)} pushforward, (B.37)
O
{GP (X , t)} = [F]T {g(x, t)} pullback. (B.38)
M M
Objective derivatives of contravariant tensors g(x, t) = g ij (x, t) ei e j and
P IJ (X , t) eI eJ transfer their fields according to the mappings
GP (X , t) = G
Mij @x i @x j P IJ (X , t),
g (x, t) = G (B.39)
@XI @XJ
P IJ (X , t) = @X @XJ
I Mij
G g (x, t) (B.40)
@x i @x j
that in matrix notation become
M
[g(x, t)] = [F] [GP (X , t)] [F]T pushforward, (B.41)
M
[GP (X , t)] = [F 1 ] [g(x, t)] [F 1 ]T pullback. (B.42)
300 Soft Solids

O O
Objective derivatives of covariant tensors g(x, t) = g ij (x, t) e i e j and
P IJ (X , t) e I e J transfer their fields according to the mappings
GP (X , t) = G

P IJ (X , t) @X @X ,
I J
O
g ij (x, t) = G (B.43)
@x i @x j
P IJ (X , t) = O @x i @x j
G g ij (x, t) I (B.44)
@X @XJ

that in matrix notation become

O
[g(x, t)] = [F 1 ]T [GP (X , t)] [F 1 ] pushforward, (B.45)
O
[GP (X , t)] = [F]T [g(x, t)] [F] pullback. (B.46)


And objective derivatives of mixed tensors g(x, t) = g ij (x, t) ei e j and
P I (X , t) eI e J transfer their fields according to the mappings
GP (X , t) = GJ

i @x i @XJ
g j (x, t) = GP IJ (X , t) j , (B.47)
@XI @x
P I (X , t) = @X @x j
I i
G J g j (x, t) J (B.48)
@x i @X

that in matrix notation become


[g(x, t)] = [F] [GP (X , t)] [F 1 ] pushforward, (B.49)

[GP (X , t)] = [F 1 ] [g(x, t)] [F] pullback. (B.50)

Here, M denotes an upper-convected derivative and O denotes a lower-con-


vected derivative in the terminology of Oldroyd (1950). The former applies
to contravariant fields, while the latter applies to covariant fields. Oldroyd
did not consider mixed tensor fields in his applications, so he did not spec-
ify the mixed Lie derivative denoted here as .
A convected derivative expressed in the spatial configuration  in the
sense of Oldroyd (a.k.a. the Lie derivative taken along the path of trajectory
followed by particle P during its motion from 0 to ) represents the
Covariant and Contravariant Issues: Configuration Physics 301

same physical time rate of change of an Eulerian field as does the simple
partial derivative taken with respect to time of its equivalent Lagrangian
field when expressed in the material or reference configuration 0 .
Lie derivatives are quantified via the following formul. For contravari-
ant vectors g(x, t) = g i (x, t) ei , the Lie derivative is computed as2
Mi @g i @g i j
g (x, t) = + j v ` ij g j ,
@t @x (B.51)
M
{g(x, t)} = {Dg/Dt} [l] {g},
where l = ` ij ei e j is the velocity gradient with ` ij = @v i (x, t)/@x j . For
covariant vectors g(x, t) = g i (x, t) e i , the Lie derivative is computed as
O @g i @g i j j
g i (x, t) = + j v + `i g j ,
@t @x (B.52)
O
{g(x, t)} = {Dg/Dt} + [l]T {g}.
For contravariant tensors g(x, t) = g ij (x, t) ei e j , the Lie derivative is
computed as
M ij @g ij @g ij k j
g (x, t) = + k v `ki g kj g ik ` k ,
@t @x (B.53)
M
[g(x, t)] = [Dg/Dt] [l] [g] [g] [l]T .
For covariant tensors g(x, t) = g ij (x, t) e i e j , the Lie derivative is com-
puted as
O @g ij @g ij k
g ij (x, t) = + k v + ` ki g kj + g ik ` kj ,
@t @x (B.54)
O
[g(x, t)] = [Dg/Dt] + [l]T [g] + [g] [l].

2 These formul come from dierentiating their associated eld-transfer laws,


@ @
 1  P
e.g., for a contravariant vector eld @t {G (X , t)} = @t [F] {g(x, t)} = [F]1 {g} +
    M M
[F]1 {g}P + [grad(g)] {v} = [F]1 {Dg/Dt} [l] {g} = [F]1 {g}, yielding {g} =
P
{Dg/Dt} [l] {g} where the identity F 1 = F 1 FF P 1 = F 1 l from Eq. (3.35) has
been used.
302 Soft Solids

And for mixed tensors g(x, t) = g ji (x, t) ei e j , the Lie derivative is com-
puted as
i @g ij @g ji
g j (x, t) = + v k ` ik g jk + g ki ` kj ,
@t @x k (B.55)

[g(x, t)] = [Dg/Dt] [l] [g] + [g] [l].
These are referred to as the Oldroyd (1950, 1970) convected derivatives,
as he was the first to derive and apply them in the literature of continuous
media mechanics. It is the opinion of this author that Oldroyds 1950 paper
was the single-most important paper published in the twentieth century in
the field of continuum mechanics.

B.4.1 Integration
The inverse of dierentiation is integration, and a like recipe exists for
integrating Lie derivatives:
1. Pull back the Lie derivative of a spatial field of interest given in  into
its material equivalent in 0 by applying, e.g., Eqs. (B.32), (B.36),
(B.40), (B.44), or (B.48).
2. Integrate this rate over time, which is now defined in the reference
configuration 0 , i.e., it is defined over a single Hilbert space where
an algebra exists, thereby permitting the summing of integration to
take place (Freed 2010).
3. Push forward the result obtained in 0 back into the spatial config-
uration  by using the appropriate mapping, e.g., Eqs. (B.7), (B.11),
(B.15), (B.19), or (B.23).
With this methodology in place, objective integrations of objective
derivatives of contravariant vectors g(x, t) = g i (x, t) ei and G (X , t) =
GI (X , t) eI are given by the integral equations
t Z
Mi 0 0 0 @x i t P I
g (x , t ) dt = G (X , t0 ) dt0 , (B.56)
0 @X I
0
Z t I tM
PGI (X , t0 ) dt0 = @X g i (x 0 , t0 ) dt0 (B.57)
0 @x i
0
that in matrix notation become
t Z t
M 0 0
{g(x , t )} dt = [F] {GP (X , t0 )} dt0
0 pushforward, (B.58)
0 0
Covariant and Contravariant Issues: Configuration Physics 303

Z t t
M
{GP (X , t0 )} dt0 = [F 1 ] {g(x 0 , t0 )} dt0 pullback (B.59)
0 0
wherein x 0 = (X , t0 ).
Objective integrations of objective derivatives of covariant vectors
g(x, t) = g i (x, t) e i and G (X , t) = GI (X , t) e I transfer their fields accord-
ing to the mappings
t Z t
O @XI
g i (x 0 , t0 ) dt0 = GP I (X , t0 ) dt0 , (B.60)
0 0 @x i
Z t t
P I (X , t0 ) dt0 = O @x i
G g i (x 0 , t0 ) dt0 I (B.61)
0 0 @X
that in matrix notation become
t Z t
O
{g(x 0 , t0 )} dt0 = [F 1 ]T {GP (X , t0 )} dt0 pushforward, (B.62)
Z 0t t0
O
{GP (X , t0 )} dt0 = [F]T {g(x 0 , t0 )} dt0 pullback. (B.63)
0 0
Objective integrals of objective derivatives of contravariant tensors
g(x, t) = g ij (x, t) ei e j and G (X , t) = GIJ (X , t) eI eJ transfer their
fields according to the mappings
t Z
Mij 0 0 @x i @x j t P IJ
g (x , t ) dt0 = G (X , t0 ) dt0 , (B.64)
0 @X I @XJ
0
Z t I @XJ t M
0 0 @X
G IJ
P (X , t ) dt = g ij (x 0 , t0 ) dt0 (B.65)
0 @x i @x j
0
that in matrix notation become
t Z t
M 0 0
[g(x , t )] dt = [F] [GP (X , t0 )] dt0 [F]T
0 pushforward, (B.66)
0 0
Z t t
M
[G (X , t )] dt = [F ] [g(x 0 , t0 )] dt0 [F 1 ]T
P 0 0 1 pullback. (B.67)
0 0
Objective integrals of objective derivatives of covariant tensors g(x, t) =
g ij (x, t) e i e j and G (X , t) = GIJ (X , t) e I e J transfer their fields ac-
cording to the mappings
t Z t
P IJ (X , t0 ) dt0 @X @X ,
O I J
0 0 0
g ij (x , t ) dt = G (B.68)
0 0 @x i @x j
Z t t
PGIJ (X , t0 ) dt0 = O @x i @x j
g ij (x 0 , t0 ) dt0 I (B.69)
0 0 @X @XJ
304 Soft Solids

that in matrix notation become


t Z t
O 0 0 0 1 T
[g(x , t )] dt = [F ] [GP (X , t0 )] dt0 [F 1 ] pushforward,
0 0
(B.70)
Z t t
0 0 O
[GP (X , t )] dt = [F]T [g(x 0 , t0 )] dt0 [F] pullback.
0 0
(B.71)
And objective integrals of objective derivatives of mixed tensors g(x, t) =
g ji (x, t) ei e j and G (X , t) = GIJ (X , t) eI e J transfer their fields accord-
ing to the mappings
t Z
i 0 0 @x i t P I 0 ) dt0 @X ,
J
g j (x , t ) dt0 = G J (X , t (B.72)
0 @XI 0 @x j
Z t I t
P I (X , t0 ) dt0 = @X 0 @x
j
i 0 0
G J g (x , t ) dt (B.73)
0 @x i 0 j @XJ
that in matrix notation become
t Z t

[g(x 0 , t0 )] dt0 = [F ] [GP (X , t0 )] dt0 [F 1 ] pushforward, (B.74)
Z 0t 0
t

[GP (X , t0 )] dt0 = [F 1 ] [g(x 0 , t0 )] dt0 [F] pullback. (B.75)
0 0

Here, 0t  dt0 denotes a convected integration over [0, t] (the inverse op-
R
erator of a convected dierentiation), while 0t  dt0 denotes a standard
integration, in the sense of Riemann, which is the inverse operator to the
derivative. The derivations and terminologies of convected dierentiation
and convected integration were introduced by Oldroyd (1950).
Convected integration, in the sense of Oldroyd, represents the same
summation over time of a spatial field expressed in the Eulerian configura-
tion  as does the classic Riemannian sum of its associated material field
expressed in the Lagrangian configuration 0 .
From formul (B.7) and (B.57), convected integration of a contravari-
ant vector g(x, t) = gi (x, t) ei is described by
t Z
Mi 0 0 0 @x i t @XI M j 0 0 0
g (x , t ) dt = g (x , t ) dt ,
@XI 0 @x 0j
t0 Z t (B.76)
M 0 0 0 1 0 M 0 0 0
{g(x , t )} dt = [F] [F (t )] {g(x , t )} dt
0 0
Covariant and Contravariant Issues: Configuration Physics 305

wherein F(t0 ) = F 0I i e i0 e I with F 0I i = @ i (X , t0 )/@XI . For covariant


vectors g(x, t) = gi (x, t) e i , convected integration of their Lie derivatives
is given by
t Z t
O 0 0 0 O @x 0j @XI
g i (x , t ) dt = g j (x 0 , t0 ) I dt0 i ,
@X @x
0t 0
Z t (B.77)
O 0 0 0 1 T 0 T O 0 0 0
{g(x , t )} dt = [F ] [F(t )] {g(x , t )} dt .
0 0
It is important to distinguish the time dependence of the dierent deforma-
tion gradients present, which pertain to their respective configurations.
For contravariant tensors g(x, t) = g ij (x, t) ei e j , convected integra-
tion of their Lie derivatives is given by
t Z
Mij 0 0 0 @x i @x j t @XI @XJ Mk` 0 0 0
g (x , t ) dt = g (x , t ) dt ,
0 @XI @XJ 0 @x 0 k @x 0 `
t Z t (B.78)
M 0 0 0 1 0 M 0 0 1 0 T 0 T
[g(x , t )] dt = [F] [F (t )] [g(x , t )] [F (t )] dt [F] .
0 0
For covariant tensors g(x, t) = g ij (x, t) e i e j , convected integration of
their Lie derivatives proceeds as
t Z t
O 0 0 0 O @x 0 k @x 0 ` 0 @XI @XJ
g ij (x , t ) dt = g k` (x 0 , t0 ) dt ,
0 0 @XI @XJ @x i @x j
t Z t (B.79)
O 0 0 0 1 T 0 T O 0 0 0 0 1
[g(x , t )] dt = [F ] [F(t )] [g(x , t )] [F(t )] dt [F ].
0 0
And for mixed tensors g(x, t) = g ji (x, t) ei e j , convected integration of
their Lie derivatives proceeds as
t Z
i 0 0 0 @x i t @XI k 0 0 @x 0 ` 0 @XJ
g j (x , t ) dt = g (x , t ) dt ,
0 @XI 0 @x 0 k ` @XJ @x j
t Z t (B.80)
0 0 0 1 0 0 0 0 0 1
[ g(x , t )] dt = [F] [F (t )] [g(x , t )] [F(t )] dt [F ].
0 0
Oldroyd (1950) introduced these integrals into the mechanics literature in
the same paper where he introduced his convected derivatives.
Besides writing his formul in terms of components of convected co-
ordinates, he also brought the deformation gradients that reside outside the
integrals in the above formul (B.76)(B.80), moving them inside their in-
tegrands, thereby producing generalized deformation gradients, as defined
306 Soft Solids

in Eq. (7.40). Oldroyd did this nearly a decade before Noll (1958) intro-
duced this idea in his body of work. Using the generalized deformation
gradient of Eq. (7.40), convected integration of a contravariant vector be-
comes
t Z t
Mi 0 0 0 @x i M j 0 0 0
g (x , t ) dt = g (x , t ) dt ,
0 @x
0j
0
t Z t (B.81)
M 0 0 0 0 M 0 0 0
{g(x , t )} dt = [F(t , t)] {g(x , t )} dt .
0 0
Convected integration of a covariant vector becomes
t Z t
O O @x 0j
g i (x 0 , t0 ) dt0 = g j (x 0 , t0 ) i dt0 ,
@x
0t Z0 t (B.82)
O O
{g(x 0 , t0 )} dt0 = [F 1 (t0 , t)]T {g(x 0 , t0 )} dt0 .
0 0
Convected integration of a contravariant tensor becomes
t Z t
Mij 0 0 0 @x i @x j Mk` 0 0 0
g (x , t ) dt = g (x , t ) dt ,
@x 0 k @x 0 `
0t Z0 t (B.83)
M M
[g(x 0 , t0 )] dt0 = [F(t0 , t)] [g(x 0 , t0 )] [F(t0 , t)]T dt0 .
0 0
Convected integration of a covariant tensor becomes
t Z t
O O @x 0 k @x 0 ` 0
g ij (x 0 , t0 ) dt0 = g k` (x 0 , t0 ) dt ,
0 0 @x i @x j
t Z t (B.84)
O 0 0 0 O
[g(x , t )] dt = [F 1 (t0 , t)]T [g(x 0 , t0 )] [F 1 (t0 , t)] dt0 .
0 0
And convected integration of a mixed tensor becomes
t Z t
i 0 0 0 @x i k 0 0 @x 0 ` 0
g j (x , t ) dt = 0k `
g (x , t ) dt ,
0 0 @x @x j
t Z t (B.85)

[g(x 0 , t0 )] dt0 = [F(t0 , t)] [g(x 0 , t0 )] [F 1 (t0 , t)] dt0 .
0 0
Formul (B.81)(B.85) are often preferred when working with convected
integrals in a mathematical setting; however, it is often more convenient to
use Eqs. (B.76)(B.80) whenever numerical solutions are sought.
Covariant and Contravariant Issues: Configuration Physics 307

B.4.1.1 Liebnitzs Rule for Dierentiating an Integral

Z Z b
d b @F (x, y) db da
F (x, y) dy = dy + F (x, b) F (x, a) (B.86)
dx a a @x dx dx

B.4.1.2 Integration by Parts

Z Z
u dv = uv v du
Z Z  Z Z 
dv
or uv dx = u dx v u dx dx (B.87)
dx

B.5 Weighted Field Transfer

Because the focus of this textbook is on isochoric materials, the topic of


weight (i.e., the Jacobian entering into a field-transfer operation) can be
somewhat avoided. When deformations are comprised of both distortions
and dilations, the weight of a field can become important. Most physi-
cal fields are without weight, in which case they are said to be absolute
fields, whereas relative fields have weight. For example, Cauchy stress T
is a relative contravariant tensor with a weight of 1, cf. Eq. (4.20), while
Truesdell strain Ey is a relative contravariant tensor with a weight of 2. Be-
cause of the importance of Cauchy stress in all of mechanics, the author
feels that a brief introduction to weighted fields is warranted.
Recall that a contravariant index maps as a tangent vector, while a co-
variant index maps as a normal to a surface. In contrast, weight does not
associate with an index. A scalar field can have weight. In fact, a mapping
of an elemental volume dV from 0 into dv in  via dv = det(F) dV (which
follows from the conservation of mass, as shown in Fig. B.4) establishes a
mapping with weight +1.
Nansons formula (4.12) maps the normal to an element of area as a
relative covariant vector with weight +1, while the Cauchy stress T (with
units of force per unit current area ) pulls back into the second PiolaKirch-
ho stress S (with units of force per unit reference area ) as a relative
contravariant tensor with weight 1. The weight belonging to this trans-
fer of field changes the measure of area between these two stress tensors.
308 Soft Solids

det(F )

B @ t0 dA B@t
dV

0 det( F1 )

Fig. B.4 Field-transfer operations for the weighted aspect of a eld, e.g., a vol-
ume element; cf. Sect. 2.2.1. For positive weighted elds, as drawn here, det F is
the pushforward map, and det F 1 = 1/ det F is the pull-back map. For negative
weighted elds, det F 1 is the pushforward map, and det F is the pull-back map

Truesdells stress rate (see p. 176) is the convected or Lie derivative


associated with Cauchy stress, wherein the term containing tr l maps its
weight.
Because det F = 1 for most materials of interest in this textbook, weight
is of little importance for us. This is why we use the Kirchho s and second
PiolaKirchho S stresses in our studies, where [s] = [F] [S ] [F]T , i.e.,
these stresses obey an absolute, contravariant, field-transfer mapping.
Appendix C

Kronecker Products
A Tensor-to-Array Mapping Scheme

The explicit and implicit elastic materials introduced in Chaps. 5 and 6 are
described in terms of fourth-order tensors that are tangents to response sur-
faces, these being the tangent compliance C and tangent modulus M ten-
sors. In the case where an imposed stress rate causes a strain rate to occur,
the governing constitutive equation is typically written in the Lagrangian
frame as
dE = C W dS dEIJ = CIJKL dSKL
or
where the tangent compliance C = CIJKL e Ie Je Ke L is a fourth-order
covariant tensor with minor symmetries, i.e., CIJKL = CIJLK = CJIKL =
CJILK , which follow from the symmetries required of stress SIJ = SJI and
strain EIJ = EJI . The elastic tangent compliance may or may not have
major symmetry. It is said to possess major symmetry whenever CIJKL =
CKLIJ .
The compliance tensor derived from the implicit theory of elasticity
presented in Chap. 6 has the form
 1
C = I I B :A
where A and B are fourth-order tensors derived from a thermodynamic
potential function that describes the internal elastic energy. Tensor A has
both minor and major symmetries, while tensor B need only possess minor
symmetry.
Equations like this are common in the study and application of consti-
tutive equations, but it is not obvious how to handle the four indices be-
longing to tensors A , B , and C . A number of schemes have been devised
over the years to deal with this dilemma that basically map second-order
matrices into vectors of higher dimension and fourth-order matrices into
second-order matrices of higher dimension, most being somewhat ad hoc
A.D. Freed, Soft Solids: A Primer to the Theoretical Mechanics of Materials, 309
Modeling and Simulation in Science, Engineering and Technology,
DOI 10.1007/978-3-319-03551-2, Springer International Publishing Switzerland 2014
310 Soft Solids

in their construction. Voigt notations (Belytschko et al. 2000; Nadeau and


Ferrari 1998) are commonly employed, but they are restricted to working
with symmetric fields. A contraction between two symmetric tensors need
not produce a symmetric tensor; furthermore, many fields arising in con-
stitutive constructions, like the velocity gradient, are not symmetric. For
these and other reasons, a more robust methodology for mapping matri-
ces into vectors is required. The Kronecker product approach to linear
algebra (Graham 1981) has been adopted for use here, which Nicholson
and Lin (1998, 1999) introduced to the mechanics community. Applica-
tions in this text are restricted to 2D plane-stress analysis. The interested
reader is referred to Nicholsons textbook (2008) for a 3D implementation
of Kronecker products and their applications in continuum mechanics, in
particular finite elements.
Plane-stress conditions imply that all o-diagonal components in stress
that relate to the 3-direction are zero, viz., S13 = S23 = S31 = S32 = 0 plus,
in the absence of surface pressure, the normal component S33 will be
zero, too. Only stress components in the 12-plane can be nonzero. For
plane-stress analysis, like the BVPs addressed in this book, the problem
space can be cast in R2 instead of R3 . The E33 component of strain is not
captured in this or any other Voigt-like scheme for plane stress. This issue
is left to be addressed on a case-by-case basis, as needed.

C.1 The vec Operator

In Kronecker product notation, second-order tensor fields defined in R2


produce 4  1 arrays; for example, for an arbitrary contravariant tensor
A = AIJ eI eJ , one denotes
A
11
A 21 
vec(A) = and vecT(A) = A11 A21 A12 A22 . (C.1)
A 12
A 22

Two, fundamental, 4  4, mapping matrices exist; they are the identity map
2 3
1000
60 1 0 07
vec(A) = I vec(A) where I = 6 40 0 1 05
7 (C.2)
0001
Kronecker Products A Tensor-to-Array Mapping Scheme 311

and the permutation map


2 3
1000
  60 0 1 07
vec A T = J vec(A) where J = 6 7
40 1 0 05 (C.3)
0001
with J 2 = I and J = J T = J 1 . From these definitions, it follows that
 
vec(sym(A)) = 12 I + J vec(A), (C.4)
 
vec(skew(A)) = 2 I J vec(A)
1 (C.5)
and, therefore, vec(sym(A)) + vec(skew(A)) = vec(A). Other useful prop-
erties of the vec operator include
vec(A) = vec(A), (C.6)
vec(A B) = vec(A) vec(B), (C.7)
tr(AB) = vecT(A) J vec(B) (C.8)
for an arbitrary scalar and for arbitrary 2  2 tensors A and B.
The dyadic (or tensor) product between two arbitrary vectors a and b,
as defined in Eq. (A.13), associates with the vec operator accordingly
 
vec(a b) = vec abT (C.9)
because   

T a1  a 1 b1 a 1 b2
ab = b1 b2 = , (C.10)
a2 a 2 b1 a 2 b2
which agrees with Eq. (A.13).

C.2 Kronecker Product

The Kronecker product is a dyadic product1 between two matrices that pro-
duces a single matrix at twice the dimension of its two originators; specif-
ically, it is defined as (Graham 1981)
2 3
 A11 B11 A11 B12 A12 B11 A12 B12
A B A12 B 6A B A B A B A B 7
A ~ B = 11 = 4 11 21 11 22 12 21 12 22 5 , (C.11)
A BA B
21 22 A B A B A B A B
21 11 21 12 22 11 22 12
A21 B21 A21 B22 A22 B21 A22 B22
1 Inthe literature, the Kronecker product is denoted as (Graham 1981; Nichol-
son 2008), but here, it is denoted as ~ because the operator is used to denote
the outer dyadic product between two vectors or two tensors, as established in
Eqs. (A.13) and (C.33)(C.35). Denitions for and ~ dier in their constructs.
312 Soft Solids

which is distinct from the inner and outer dyadic products and used
in this text. Establishing how these dyadic operators relate to one another
is the main objective of this appendix.
The Kronecker sum and dierence are defined by (cf. Nicholson
2008, p. 45)
A B = A ~ I + I ~ B, (C.12)
A B = A ~ I I ~ B. (C.13)
The Kronecker product, sum, and dierence have eigenvalues that are
products, sums, and dierences between their constituents eigenvalues,
hence their names.
Many relationships exist for the Kronecker product (Graham 1981):
vec(A) = I ~ A vec(I), (C.14)
vec(AB) = I ~ A vec(B) = B T ~ I vec(A), (C.15)
T
vec(ABC ) = C ~ A vec(B), (C.16)
 
A ~ B = J B ~ A J, (C.17)
 T
A ~ B = AT ~ B T , (C.18)
 1
A~B = A 1 ~ B 1 , (C.19)
  
A ~ I I ~ B = A ~ B, (C.20)
  
I ~ A I ~ B = I ~ AB, (C.21)
  
A ~ B C ~ D = AC ~ BD (C.22)
where A, B, C , and D are arbitrary tensors, while I is the identity tensor.

C.3 The ten Operator

Nicholson and Lin (1998) handled tensor equations like dE = C : dS , or


dEIJ = CIJKL dSKL , by surmising the existence of a Kronecker operation of
the form
C ) vec(dS ),
vec(dE ) = ten(C (C.23)
wherein the ten operator possesses properties (Nicholson and Lin 1999)
ten(CC ) = ten(C
C ), (C.24)
B C ) = ten(B
ten(B B) ten(CC ), (C.25)
ten(BB : C ) = ten(B
B) ten(C
C) (C.26)
Kronecker Products A Tensor-to-Array Mapping Scheme 313

where is a scalar and B and C are fourth-order tensors constructed from


dyadic products between two second-order tensors, say A and B.
The inverse of a fourth-order tensor without symmetry is defined in
Kronecker notation by
A) ten(A
ten(A A1 ) ten(A
A1 ) = ten(A A) = I , (C.27)
with
ten(A A),
A1 ) = ten1 (A (C.28)
while the inverse of a fourth-order tensor with minor symmetry is defined
in Kronecker notation by
A) ten(A
ten(A A1 ) ten(A
A1 ) = ten(A A) = ten(I I), (C.29)
as presented in Eq. (6.59), where I I is the fourth-order identity tensor
whose Kronecker form for plane-stress analysis is
2 3
1 0 0 0
60 1/2 1/2 07
I + J) = 6
ten(I I) = 12 (I 7
40 1/2 1/2 05 . (C.30)
0 0 0 1
A fourth-order tensor C has both minor and major symmetries if and only
if (Nicholson 2008)
ten(C C ) and
C ) = tenT(C C ) = J ten(C
ten(C C) J. (C.31)
It immediately follows that C 1 will possess both minor and major sym-
metries if C possesses both minor and major symmetries.
Twenty-four possible index orientations exist that one must consider
when assigning a ten operator, which lead to twenty-four distinct Kro-
necker forms (Nicholson and Lin 1998). The Kronecker product between
two tensors, say A and B, associates with components
[A ~ B]IJKL = BIK AJL . (C.32)
In the following, the two dyadic products that arise in this textbook are
expressed in terms of their representative Kronecker products as a vehicle
for implementing our theories into software.
The outer dyadic product usually appears in one of three formats
[A A]IJKL = AIJ AKL , (C.33)
[A B]IJKL = AIJ BKL , (C.34)

1  IJKL 1  IJ KL KL BIJ .
2 A B + B A = 2 A B + A (C.35)
314 Soft Solids

These three outer dyadic operators have Kronecker forms of


ten(A A) = vec(A) vecT(A), (C.36)
ten(A B) = vec(A) vecT(B), (C.37)
   
ten 12 (A B + B A) = 12 ten(A B) + ten(B A) . (C.38)
Symmetry of A and B, although commonplace, is not a requirement of
these operators.
In like manner, the inner dyadic product also appears in three formats2
 
[A A]IJKL = 12 AIK AJL + AIL AJK , (C.39)
 
[A B]IJKL = 12 AIK BJL + AIL BJK , (C.40)

1 IJKL 1  IK JL
2 (A B + B A) = 4 A B + AIL BJK

+ AJL BIK + AJK BIL . (C.41)
These three inner dyadic operators have Kronecker forms of
I + J ),
ten(A A) = 12 (A ~ A)(I (C.42)
ten(A B) = 1 (B
2 I + J ),
~ A)(I (C.43)
1 
ten 2 (A B + B A) = 1 (B
4 I + J ),
~ A + A ~ B)(I (C.44)
Again, symmetry of A and B is not a requirement of these operators, al-
though that is the typical situation.

C.4 Coordinate Transformations

For a change in basis from {eI } to {eI0 } where components of tensors A =


0 0
AIJ eIeJ and A = AIJKL eIeJeKeL rotate via Q into A = AI J eI0
0 0 0 0 0 0 0 0
eJ0 and A = AI J K L eI0 eJ0 eK0 eL0 with AI J = Q IK Q JL AKL and
0 0 0 0 0 0 0 L 0 MNOP given that QQT = I and det Q = 1
AI J K L = Q IM Q JN Q K O QP A

2 The author (Freed et al. 2005) introduced a notation of A  B, which is equiva-


lent to the notation 12 (A B +B A). The latter notation is selected here because
it better reects how major symmetry is achieved at the component level in this
dyadic product, i.e., it is akin to the notation used to express the symmetric part
of some tensor T , viz., sym(T ) = 12 (T + T T ).
Kronecker Products A Tensor-to-Array Mapping Scheme 315

(cf. Sect. B.1), then such a change in basis carries over into the vec and ten
Kronecker forms as (Nicholson and Lin 1999)
vec0 (A) = Q ~ Q vec(A), (C.45)
ten0 (A
A) = Q ~ Q ten(A
A) (Q ~ Q)T (C.46)
0
with the rotation matrix Q = Q IJ eI0 e J residing in physical space.
Appendix D

Solver for a First-Order ODE

Numerical solutions to elastic material models described by a system of


ODEs of the form SP = M (E , S ) : EP , as presented in Chaps. 5 and 6, can
be cast into the general form of a nonlinear ODE
y 0 = f (x, y) with IC y(x0 ) = y0 where x 2 [x0 , X], (D.1)
wherein y 0 = dy/dx. The methods and algorithms presented in this ap-
pendix solve this class of problems.
Inherent RungeKutta stable (IRKS) methods are a subclass to a
larger class of numerical methods known as general linear (GL) meth-
ods. GL methods seek to capture the best attributes from the multistage
RungeKutta (RK) methods and combine these with the best attributes
from the multistep AdamsBashforth/Moulton methods. The mathemat-
ics underlying these methods are sophisticated, even daunting at times, the
details of which you can study in a text by Butcher (2008) with Butcher
and Podhaisky (2006) assessing their error. Only an overview is presented
herein, sucient for understanding how these integrators work, conceptu-
ally, with algorithms being provided to illustrate their implementation.
The explicit method presented in this appendix can be used to solve
explicit elastic solids, as presented in Sect. 5.4, and the implicit Hookean
solid (since it is an explicit system of ODEs). The implicit method pre-
sented below is L-stable and can be used to solve implicit elastic solids,
since they are comprised of implicit systems of ODEs. IRKS methods are
not self-starting, in general, so a special starter method is also provided,
which happens to be an implicit IRKS integrator, too.
In GL methods, r quantities y1[n1] , y2[n1] , : : :, yr[n1] (e.g., they are the
history or multistep values acquired over previous integration steps, viz.,
y [n1] , y [n2] , : : :, y [nr] ) are updated via integration from step n1 to step
n, thereby producing a subsequent set of new estimates y1[n] , y2[n] , : : :, yr[n] ,
A.D. Freed, Soft Solids: A Primer to the Theoretical Mechanics of Materials, 317
Modeling and Simulation in Science, Engineering and Technology,
DOI 10.1007/978-3-319-03551-2, Springer International Publishing Switzerland 2014
318 Soft Solids

each being an array of dimension N, the number of simultaneous equations


being solved. The IRKS methods presented here assemble these arrays into
a super-array with Nordsieck scaling, i.e., the values over the past r steps

 
going back into the history are brought forward into the current step as a
truncated Taylor series expansion, viz.,

y1[n] y(xn )
y2[n] hy 0 (xn )
[n]
Y [n] = y3 = h2 y 00 (xn ) (D.2)
.. ..
. .
yr[n] hr1 y (r1) (xn )

where y 0 = dy/dx, y 00 = d2 y/dx2 , etc., with the super-array Y [n] having a


length of rN, while h = xn xn1 is the step size of integration (x being the
independent variable of integration, which is usually time). This truncated
Taylor series can be used to interpolate between neighboring pairs of nodal
points, if needed.
To secure a solution in the form of Eq. (D.2) using a multistage integra-
tion from step n1 to step n, s stage values Y1 , Y2 , : : :, Ys are computed
over an interval [xn1 , xn ] along with their s stage derivatives F1 , F2 , : : :,
Fs , in RK fashion, where Yi approximates the solution y(xn1 + ci h) and


Fi approximates its rate f (xn1 + ci h, Yi ). The stage derivative arrays are
assembled into a super-array

F1 (Y1 )
F2 (Y2 )
F = F3 (Y3 ) (D.3)
..
.
Fs (Ys )

where F is of length s  N. For the three, third-order, IRKS integrators


presented in this appendix, r = s = 4.
Lack of knowledge of a Taylor expansion at the start of an integra-
tion, viz., Y [0] , is why these methods are, generally, not self-starting. The
start-up method of Table D.1 is designed to accept an input or IC of
T
Y [0] = y0 0 0 : : : 0 (D.4)

while returning an output for Y [1] in accordance with Eq. (D.2).


Solver for a First-Order ODE 319

GL methods are comprised of four matrices; they are A = [Aij ]ss ,


U = [Uij ]sr , B = [Bij ]rs , and V = [Vij ]rr . They appear in a special
linear system of equations used to update the state; specifically,
X
s X
r
Yi = h Aik Fk (Yk ) + Uik yk[n1] , i = 1, 2, : : : , s, (D.5)
k=1 k=1

[n]
X
s X
r
[n1]
yi =h Bik Fk (Yk ) + Vik yk , i = 1, 2, : : : , r, (D.6)
k=1 k=1
where matrix A contains stage coecients, akin to RungeKutta methods;
matrix B contains step coecients, akin to AdamsBashforth/Moulton
methods; while matrices U and V couple these two processes together
into a unified integration algorithm.
The sum on index k over the RungeKutta-like Aik coecients termi-
nates at i1 for explicit methods, at i for diagonally implicit methods, or at
s for fully implicit methods (which are not discussed here).

D.1 Estimate Error

A Butcher tableau for the GL method is a partitioned (s + r)  (s + r) matrix


of the form
" #
A U
. (D.7)
B V
Property F, a.k.a. a first-same-as-last method in the RK literature, is a
desirable property for an IRKS method to have in that it enables one to
compute an asymptotically accurate estimate of the local truncation er-
ror (Butcher and Podhaisky 2006). It is readily apparent by examining
a methods Butcher tableau whether it does or does not possess this prop-
erty. If it does, then the last row in matrix [A|U ] will be identical to the
first row in matrix [B|V ], and the second row in matrix [B|V ] will be all
zeros, except for the last column in matrix B where there will be a one.
To obtain an estimate for the local truncation error of a solution for an
IRKS method with property F, one first constructs the dierence vector
!
X
s
[n] = h 0 y 0 [n1] i Fi , (D.8)
i=1
320 Soft Solids

Table D.1 The partitioned matrix for a


start-up method for use with third-order IRKS
T
methods where c = 1/4 , 1/2 , 3/4 , 1 a
1/4 0 0 0 1 0 0 0
1/4 1/4 0 0 1 0 0 0
5/4 7/4 1/4 0 1 0 0 0
5/12 5/12 1/12 1/4 1 0 0 0
5/12 5/12 1/12 1/4 1 0 0 0
0 0 0 1 0 0 0 0
29/10 67/10 7/10 31/10 0 0 0 0
56/5 88/5 8/5 24/5 0 0 0 0
a This method was created for the author by
Professor John Butcher so that the author
could start Butchers integrators listed in Ta-
bles D.2 and D.3 with a method that would re-
turn a compatible Nordsieck array of the Tay-
lor series, which can in turn be passed to the
integrators of Tables D.2 and D.3 for continued
integration

from which one can then compute a normalized estimate for the error via
k[n] k
errorn =   (D.9)
max 1, ky [n1] k
where 0 and  = {1 , 2 , : : : , s }T are parameters of the integrator.
Note that y [n1] and hy 0 [n1] are stored as the first two sub-arrays in the
super-array Y [n1] .

D.2 IRKS Methods

Tables D.2 and D.3 provide Butcher tableau for third-order explicit and
implicit IRKS methods, respectively, each with property F. These two
methods are not self-starting, so a start-up method must be called upon to
take the first step of integration. A compatible, third-order, start-up method
for these integrators is listed in Table D.1.
Typical implementations impose a fixed step size h. If a need arises
to adjust the step size to, say, keep the error under control, then, at the
point of interruption, the next step taken with the new step size needs to be
done with the starter method of Table D.1, after which point integration can
proceed using the regular integrator as before, but now with this new step
Table D.2 Partitioned matrix for an explicit IRKS method of third order with prop-
erty F where c = { 1/4 , 1/2 , 3/4 , 1}T with error-estimation coecients of 0 = 98.267807 and
 = {329.071228, 397.606843, 201.071228, 34.267807}T a

0 0 0 0 1 0.25 0.0625 0.015625


0.845232 0 0 0 1 0.345232 0.172616 0.033481
0.0449691 0.882754 0 0 1 0.087785 0.29777 0.231759
1.03061 0.483638 0.157846 0 1 0.672094 0.235712 0.177668
1.03061 0.483638 0.157846 0 1 0.672094 0.235712 0.177668
0 0 0 1 0 0 0 0
0.712022 0.63459 0.0665772 1.06962 0 0.343573 0.0487644 0.512949
Solver for a First-Order ODE

0.707945 0.57053 0.210071 0.622123 0 0.866424 0.00463586 0.0487644


a Parameters are for method e3a extracted from the Atlas of general linear methods with inherent
RungeKutta stability, http://www.math.auckland.ac.nz/hpod/atlas
321
322

Table D.3 The partitioned matrix for an L-stable implicit IRKS method of third order that
T
possesses the property F where c = 1/4 , 1/2 , 3/4 , 1 and whose coecients for error estimation
are 0 = 43.700369 and  = {110.801474, 70.202212, 17.198526, 20.299631}T a

0.225 0 0 0 1 0.25 0.05 0.0265625


0.211287 0.225 0 0 1 0.063713 0.0806435 0.0833663
0.946338 0.342943 0.225 0 1 0.0783954 0.0947737 0.121956
0.52149 0.662474 0.490476 0.225 1 0.425507 0.216014 0.103603
0.52149 0.662474 0.490476 0.225 1 0.425507 0.216014 0.103603
Soft Solids

0 0 0 1 0 0 0 0
0.0423385 0.695379 0.784079 1.0116 0 0.880558 0.521284 0.774748
0.077564 0.246379 0.321806 0.274145 0 0.276282 0.350743 0.521284
a Parameters
are for method i3a extracted from the Atlas of general linear methods with inherent
RungeKutta stability, http://www.math.auckland.ac.nz/hpod/atlas
Solver for a First-Order ODE 323

size. This is necessary because the Taylor series retained in the Nordsieck
vector associates with a single step size.
Explicit integrators, like the one in Table D.2, are appropriate for
non-sti problems, while implicit integrators, like the one in Table D.3,
are needed whenever a system of ODEs exhibits stiness, as they often do.
Here we advocate using an implicit integrator to solve our implicit system
of ODEs. Our implementation of an implicit integrator calls upon New-
tons iterative method for finding roots; hence, the cost per step is greater
than that of an explicit method; however, the allowable step size is often
much greater, which often outweighs their per-step cost in overall perfor-
mance.
In FE codes, the step size of the global solver, e.g., the step size pertain-
ing to Algorithms 2.12.3, and the step size for the local solver that updates
a constitutive equation at the Gauss points are, typically, dierent. A finer
step size is often required of the constitutive solver in order to obtain con-
vergent solutions, usually due to material nonlinearity, and quite often by
an inherent stiness present in the system of ODEs to be solved.
IRKS methods, like those belonging to Tables D.1, D.2, and D.3, can be
implemented via Algorithm D.2, from which implicit IRKS methods call
Algorithm D.1. Algorithm D.1 uses Newtons method to iteratively solve
two linear equations simultaneously; they are
Yi h Aii Fi (Yi ) YQ i = 0, (D.10)
hFi hf (xn1 + ci h, Yi ) = 0, (D.11)
Pi1 Pr
wherein YQ i = h k=1 Aik Fk (Yk ) + k=1 Uik yk[n1] is the value for Yi sup-
plied by Algorithm D.2. Equation (D.11) is multiplied through by the step
size h so that the dimensionality of the two equations is the same. Solving
for hFi (instead of Fi outright) has the advantage of placing ones along
the diagonal of the Jacobian (cf. Algorithm D.1). Procedure Solve in Al-
gorithm D.2 receives values for xn1 and Y [n1] belonging to step n1 and
returns values for xn and Y [n] belonging to step n along with an error esti-
mate for the returned solution y over [xn1 , xn ].
324 Soft Solids

Algorithm D.1 Newtons method for implicit IRKS


var F variables and functions supplied by Algorithm D.2
integer N F number of equations in the ODEs
real h F step size
matrix A F appropriate RK Butcher tableau
function f (x, y) = {f j (x, y)} F system of ODEs f = dy/dx
function J(x, y) = [Jjk (x, y)] F their Jacobian J = df /dy

function NewtonF (i, x, YQ , Y , F )


for j 1, N do
gj Y j  Aii hF j YQ j  F solve for Yi
g j+N h F j f j (x, Y ) F solve for hFi
end for
return g
end function

function NewtonJ (i, x, Y )


J 0
for j 1, 2N do
Jjj 1
end for
for j 1, N do
Jj,j+N Aii
end for
for j 1, N do
for k 1, N do
Jj+N,k hJjk (x, Y )
end for
end for
return J
end function
Solver for a First-Order ODE 325

Algorithm D.1 continuation


procedure Newton (i, x, var Y , var F )
YQ Y
repeat F Newtons algorithm
G NewtonF (i, x, YQ , Y , F )
J NewtonJ (i, x, Y )
dg J 1 G
repeat
for j 1, N do
Yj G j dg j
Fj (Gj+N dg j+N )/h
end for
g NewtonF (i, x, YQ , Y , F )
dg dg/2
until kgk < kG k
until kgk < m
for j 1, N do F assign vectors Y and F for return
Yj gj
Fj g j+N /h
end for
end procedure
326 Soft Solids

Algorithm D.2 IRKS ODE solver


var F variables and functions defining the problem
boolean firstStep F flag to start or restart the integrator
integer N F number of equations in the ODEs
function f (x, y) = {f j (x, y)} F system of ODEs f = dy/dx
function J(x, y) = [Jjk (x, y)] F their Jacobian J = df /dy

procedure Initialize (h, method)


F step size h, integrator method 2 {explicit, implicit}
firstStep true
end procedure

procedure Solve (var x, var Y , var error)


if firstStep then
Get parameters from Table D.1
else if method = explicit then
Get parameters from Table D.2
else if method = implicit then
Get parameters from Table D.3
end if
Y [n1] Y
for i 1, s do F solve Eq. (D.5)
! x + ci h
Y 0 F denotes Yi
for k 1, r do
for j 1, N do
[n1]
Yj Y j + Uik Yj+(k1)N
end for
end for
for k 1, i 1 do
for j 1, N do
Yj Y j + h Aik Fj+(k1)N
end for
end for
Solver for a First-Order ODE 327

Algorithm D.2 first continuation


if method = explict then
for j 1, N do
Fj f j (!, Y ) F assign Fi = f (xn1 + ci h, Yi )
end for
else F estimate Fi = f (xn1 + ci h, Yi )
if i = 1 then
for j 1, N do
[n1]
Fj Yj+N /h F these are property F methods
end for
else
for j 1, N do
Fj Fj+(i2)N F use prior stage for estimate
end for
end if
Newton (i, !, Y , F ) F solve for stage Yi and derivative Fi
end if
for j 1, N do
Fj+(i1)N Fj F populate super-array F with vector Fi
end for
end for
for i 1, r do F solve Eq. (D.6)
Y 0 F denotes yi[n]
for k 1, r do
for j 1, N do
[n1]
Yj Y j + Vik Yj+(k1)N
end for
end for
for k 1, s do
for j 1, N do
Yj Y j + h Bik Fj+(k1)N
end for
end for
for j 1, N do
Yj+(i1)N Yj F update Nordsieck solution Y [n]
end for
end for
x x+h F update independent variable xn
328 Soft Solids

Algorithm D.2 second continuation


for j 1, N do F estimate local truncation error
[n1]
yj Yj
[n1]
j 0 Yj+N
end for
for i 1, s do
for j 1, N do
j  j i hFj+(i1)N
end for
end for
error kk/max(1, kyk)
firstStep false F update stored variable
end procedure
Appendix E

Solver for Convolution Integrals

Numerical solutions for viscoelastic material models of the Boltzmann


type, as presented in Chap. 7, entail approximating a convolution integral
Z x
 
K(x y) F g(y), g(x) dy, x 2 [0, X], (E.1)
0
containing a univariate kernel function K and a bivariate forcing function
F with arguments g that are themselves univariate functions, e.g., F(t),
with x denoting time in the case of viscoelasticity. A suitable methodology
for approximating Eq. (E.1) is presented in this appendix. It is the third
method derived by Diethelm and Freed (2006) and is comprised of four
algorithms.
A straightforward numerical integration of Eq. (E.1) is an O(N 2 ) pro-
cess, with there being N integration steps taken over the interval [0, X]. The
algorithm presented here is a more ecient O(N log N) process. This in-
creased eciency is possible because viscoelastic kernels K monotonically
decay, asymptotically approaching zero with increasing argument. Conse-
quently, numerical sampling and summing can occur at a logarithm gait.
Algorithm E.1 uses a p-adic scheme to determine the number of integra-
tion steps that will be incurred over each subinterval of integration for any
x 2 [0, X]. The idea is to break up the integral of Eq. (E.1) into a sequence
of integrals
Z QT Z Q2 T Z Q T !
 
+ + + K(x y) F g(y), g(x) dy (E.2)
0 QT Q1 T

where x = Q T with T > 0 being a characteristic time for the kernel K,


e.g.,  in the viscoelastic kernels of Sect. 7.2, and Q 2 {3, 5, 7, : : :} is a
quality parameter that sets the number of characteristic times that reside in
A.D. Freed, Soft Solids: A Primer to the Theoretical Mechanics of Materials, 329
Modeling and Simulation in Science, Engineering and Technology,
DOI 10.1007/978-3-319-03551-2, Springer International Publishing Switzerland 2014
330 Soft Solids

each subinterval of integration (the larger Q is, the more accurate the result
will be, but at an increased cost of integration) with = dlogQ (X/T)e
designating the number of subintervals assigned.
Numerical integration occurs with a mesh size of h = T/S over the first
subinterval [0, QT] in Eq. (E.2) (where S  4 is the number of integration
steps over [0, T] in the first subinterval of integration), a mesh size of hQ
over the second interval [QT, Q2T], a mesh size of hQ2 over the third in-
terval [Q2 T, Q3 T], etc. Mesh size increases with a logarithmic gait, an idea
proposed by Ford and Simpson (2001).
As a solution to Eq. (E.1) advances by steps of size h over an interval
[0, X] to its final step N = dX/he, history variables belonging to the integral
will need to be stored for reuse in future steps not yet taken. The challenge,
and the elegance of Algorithm E.2, is to decimate only those history data
that will no longer be requiredretaining only those data with future re-
quirements. This is not a trivial specification. It was the cornerstone upon
which this numerical method of Diethelm and Freed (2006) was designed.
Retaining only those history data that will be needed in future integration
steps, recalling that the gait of sampling increases logarithmically the fur-
ther back into its history one goes, minimizes the storage requirements of
this numerical method. Algorithm E.21 manages the history of field vari-
ables g(xn ), n = 0, 1, : : : , N, that are the arguments of ones forcing function
F in Eq. (E.1).
Not all kernel functions K are simple and ecient to calculate, e.g., the
MittagLeer function E, of Appendix F. For this reason, kernel func-
tion calls are handled by Algorithm E.3, where the actual kernel function
K of interest is tabulated once, prior to integration, assembling an array of
instances with logarithmic gait over the interval [0, X] of integration. Eval-
uations are made at the exact nodes of integration over the first subinterval
of integration [0, QT] in Eq. (E.2) where a typical kernel has its greatest
variation. A fourth-order NevilleAitken interpolation is used for estimat-
ing the kernel in the remaining subintervals where a typical kernel exhibits
little variation.
1 There is a typo at the end of Algorithm 4, as stated in Diethelm and Freed
(2006), which is Algorithm E.2 herein. In Diethelm and Freed (2006) it reads:
FOR i 1 TO Ln DO
I I + Pn [i].
It should read:
FOR i 1 TO Ln1 DO
I I + Pn1 [i].
This error has been xed in Algorithm E.2 by adjusting the ow of the algorithm.
Solver for Convolution Integrals 331

Algorithm E.4 solves Eq. (E.2), which in turn calls Algorithms E.1E.3.
Each call to this algorithm advances the solution by one incremental step
of size h along its path [0, X], requiring field g(xn ) as input at each step
n 2 [1, N]. To allow kernels that may be singular at the upper limit of in-
tegration, which often occur in viscoelastic models, a useful quadrature
method must not have a node located at the upper limit of integration.
Diethelm and Freed (2006) selected a midpoint rule with a Laplace end
correction, specifically (cf. Bra 1977, Sect. V.8)
8
Z b <d(ba)/he
X
f (x) dx  h f a + 2j1 h
a : 2
j=1
703 h i
+ f a + 12 h + f b 12 h
5760
463 h i
(E.3)
f a + 32 h + f b 32 h
1920
101 h i
+ f a + 52 h + f b 52 h
640 )
223 h 7

7
i
f a+ 2 h +f b 2h ,
5760

where d(b a)/he  4 and, as such, this method is not self-starting. The
four-point MacLaurin rule is used to integrate steps n = 1, 2, 3, viz.,
Z xn
nh n h 7 1
k(xn y) f (y, xn ) dy  13 k 8 xn f 8 xn , xn
0 48
i
+ k 18 xn f 78 xn , xn
h (E.4)
5 3
+ 11 k 8 xn f 8 xn , xn
io
+ k 38 xn f 58 xn , xn ,

which also avoids the upper limit of integration as a quadrature node. Both
integrators converge as O(h5 ) to the actual solution provided that the inte-
grand is a C5 function. The midpoint rule with a Laplace end correction
works with our memory management scheme, whereas the MacLaurin rule
does not.
332 Soft Solids

Algorithm E.1 O(log n) partitioning of integration nodes


procedure InitPAdic (N, Q, S)
` blogQ (N/S)c + 1 F Pn = {Pn[1], : : : , Pn [`]}T
end procedure.

procedure PAdic (n, var Ln , var Pn )


m n div S F p-adic representation of m
if m = 0 then
Ln 1
Pn [1] 0
else
Ln 0
while m > 0 do
Ln Ln + 1
Pn [Ln ] m mod Q
m m div Q
end while
for i Ln + 1, ` do
Pn [i] 0
end for
for i 1, Ln 1 do F put in non-standard p-adic form
if Pn [i] < 1 then F ensures Pn [i] 0 8 i 2 [1, Ln ]
Pn [i] Pn [i] + Q
Pn [i + 1] Pn [i + 1] 1
end if
end for
if Pn [Ln ] = 0 then
Ln Ln 1
end if
end if
for i 1, Ln do F convert to # of integration steps
Pn [i] SPn [i]
end for
Pn [1] Pn [1] + n mod S
end procedure.
Solver for Convolution Integrals 333

Algorithm E.2 Manage history variables


var integers I, n F managed history variables
field g(xn1 )
integer array Pn1 of length Ln1

procedure InitHistory (N, Q, S, g(x0 ))


` S(1 + Q(1 + dlogQ (N/S)e)) F G = {G[0], : : : , G[`]}T
PAdic (0, Ln1 , Pn1 )
PAdic (1, Ln , Pn )
g(xn1 ) g(x0 )
I 1
n 1
end procedure.

procedure History (g(xn ), var G ) F note that G[n] g(xn )


if I = 1 then
G[0] g(xn1 )
end if
G[I] (g(xn1 ) + g(xn ))/2
if (n 1 > S) and ((n S) mod (QS) = 0) then
for j 1, Ln do F garbage collection
if Pn [j] < Pn1 [j] then
a 0
for i j + 1, Ln do
a a + Pn1 [i]
end for
b S
for i 1, j 1 do
b b + Pn [i]
end for
for i 1, S do
G[a + i] G[a + iQ (Q 1) div 2]
end for
for i 1, b do
G[a + S + i] = G[a + i + QS]
end for
end if
end for
end if
334 Soft Solids

Algorithm E.2 continued


I 0 F update managed history variables
for i 1, Ln do
I I + Pn [i]
end for
n n+1
g(xn1 ) g(xn )
Ln1 Ln
for i 1, Ln do
Pn1 [i] Pn [i]
end for
PAdic (n, Ln , Pn )
end procedure.
Solver for Convolution Integrals 335

Algorithm E.3 Manage kernel function calls


procedure InitKernel (N, Q, S, h, K)
L dlogQ (N/S)e
` LS(Q 1) + S 1 F locations with logarithmic gait
m 0
for i 1, SQ 1 do
m m+1
X[m] h(i 1/2)
end for
y h(SQ 1)
for i 2, L do
for j 1, S(Q 1) do
m m+1
X[m] y + h(j 1/2)Qi1
end for
y y + hS(Q 1)Qi1
end for
for i 1, ` do F create array of kernel values
[i] K(X[i]) F K is a univariate function
end for
end procedure.

function Kernel (x)


lo 1
hi `
repeat F locate x in X so X[lo] < x
X[hi]
mid (lo + hi) div 2
if x > X[mid] then
lo mid
else
hi mid
end if
until lo = hi 1
if lo < 3 then F set indexer for interpolation
m 1
else if lo < ` 2 then
m lo 1
else
m `3
end if
336 Soft Solids

Algorithm E.3 continued: NevilleAitken interpolation


(xX[m+1]) [m](xX[m]) [m+1]
K12 X[m]X[m+1]
(xX[m+2]) [m+1](xX[m+1]) [m+2]
K23 X[m+1]X[m+2]
(xX[m+3]) [m+2](xX[m+2]) [m+3]
K34 X[m+2]X[m+3]
(xX[m+2]) K12 (xX[m]) K23
K123 X[m]X[m+2]
(xX[m+3]) K23 (xX[m+1]) K34
K234 X[m+1]X[m+3]
(xX[m+3]) K123 (xX[m]) K234
K X[m]X[m+3]
return K
end function.
Solver for Convolution Integrals 337

Algorithm E.4 Fast convolution quadrature


procedure Initialize (Q, S, T, X, g(x0 ), K, F )
F Q 2 {3, 5, 7, : : :}, S > 3, T > 0, X > 0, univariate K, bivariate F
h T/S
N dX/he
for j 4, 7 do F Laplaces weights of quadrature
for i 1, j do
`i,j 1
end for
`1,j `1,j + 703/5760
`2,j `2,j 463/1920
`3,j `3,j + 101/640
`4,j `4,j 223/5760
`j,j `j,j + 703/5760
`j1,j `j1,j 463/1920
`j2,j `j2,j + 101/640
`j3,j `j3,j 223/5760
end for
Q1 1 + 703/5760
Q2 1 463/1920
Q3 1 + 101/640
Q4 1 223/5760
InitPAdic (N, Q, S)
InitHistory (N, Q, S, g(x0 ))
InitKernel (N, Q, S, h, K)
end procedure.

function ConvQuad (g(xn )) R


F solves for qn  0xn K(xn y) F (g(y), g(xn )) dy
History (g(xn ), G ) F n is from Algorithm E.2
if n < 4 then F MacLaurin approximation for qn
if n = 1 then F g(xn1 ) is from Algorithm E.2
g1 g(xn1 ) + (g(xn ) g(xn1 ))/8;
g2 g(xn1 ) + 3(g(xn ) g(xn1 ))/8;
g3 g(xn1 ) + 5(g(xn ) g(xn1 ))/8;
g4 g(xn1 ) + 7(g(xn ) g(xn1 ))/8;
else if n = 2 then
g1 G[0] + (G[1] G[0])/2;
g2 G[1] + (g(xn1 ) G[1])/2;
g3 g(xn1 ) + (g(xn ) g(xn1 ))/4;
g4 g(xn1 ) + 3(g(xn ) g(xn1 ))/4;
338 Soft Solids

Algorithm E.4 first continuation


else
g1 G[0] + 3(G[1] G[0])/4;
g2 G[1] + 5(G[2] G[1])/8;
g3 G[2] + 3(g(xn1 ) G[2])/4;
g4 g(xn1 ) + 5(g(xn ) g(xn1 ))/8;
end if
qn nh{13[K(7nh/8) F (g1 , g(xn )) + K(nh/8) F (g4 , g(xn ))]
+11[K(5nh/8) F (g2 , g(xn )) + K(3nh/8) F (g3 , g(xn ))]}/48
else F Laplace approximation for qn
m I F I is from Algorithm E.2
if Pn1 [1] < 7 then F Pn1 is from Algorithm E.2
qn h`1,Pn1 [1]+1
 Kernel(h/2) F ((g(xn1 ) + g(xn ))/2, g(xn ))
else
qn h`Q1 Kernel(h/2) F ((g(xn1 ) + g(xn ))/2, g(xn ))
end if
if Pn1 [1] < 7 then F integrate with step size h
for i 2, Pn1 [1] + 1 do
qn qn + h`i,Pn1 [1]+1
 Kernel(h(i 1/2)) F (G[m], g(xn ))
m m1
end for
else
for i 2, 4 do
qn qn + h`Qi Kernel(h(i 1/2)) F (G[m], g(xn ))
m m1
end for
for i 5, Pn1 [1] 3 do
qn qn + h Kernel(h(i 1/2)) F (G[m], g(xn ))
m m1
end for
for i Pn1 [1] 2, Pn1 [1] + 1 do
qn qn + h`QPn1 [1]+2i
 Kernel(h(i 1/2)) F (G[m], g(xn ));
m m1
end for
end if
Solver for Convolution Integrals 339

Algorithm E.4 second continuation


y h(1 + Pn1 [1]) F integrate with step size hQj1
for j 2, Ln1 do F Ln1 is from Algorithm E.2
if Pn1 [j] < 8 then
for i 1, Pn1 [j] do
qn qn + hQj1 `i,Pn1 [j]
 Kernel(y + h(i 1/2)Qj1 ) F (G[m], g(xn ))
m m1
end for
else
for i 1, 4 do
qn qn + hQj1 Qi
 Kernel(y + h(i 1/2)Qj1 ) F (G[m], g(xn ))
m m1
end for
for i 5, Pn1 [j] 4 do
qn qn + hQj1
 Kernel(y + h(i 1/2)Qj1 ) F (G[m], g(xn ))
m m1
end for
for i Pn1 [j] 3, Pn1 [j] do
qn qn + hQj1 QPn1 [j]+1i
 Kernel(y + h(i 1/2)Qj1 ) F (G[m], g(xn ))
m m1
end for
end if
y y + hPn1 [j]Qj1
end for
end if
return qn
end function.
Appendix F

Solver for the MittagLeer Function

The two-parameter MittagLeer function E, (z) is an entire function


(in z 2 C) of order 1/ defined by the power series (Erdlyi et al. 1955,
Sect. 18.1)
X
1
zk
E, (z) = , 2 <+ , 2 <, z 2 C , (F.1)
 ( + k)
k=0
where  is the gamma function, < is the set of real numbers, <+ is the
set of positive real numbers, and C is the set of complex numbers. The
one-parameter MittagLeer function E (x) = E,1 (x) is the original func-
tion studied by Mittag-Leer (1904). The MittagLeer function plays
the same role in dierential equations of fractional order that the exponen-
tial function ez plays in ordinary dierential equations; in fact, ez = E1,1 (z).
A general numerical algorithm for computing the MittagLeer func-
tion defined in Eq. (F.1) has been derived by Luchko (Gorenflo et al. 2002)
and is reproduced in a review article written by Diethelm et al. (2005). Our
needs are more modest here, viz.,
X
1
xk
E, (x) = , 0 <
1, 2 {0, 1}, x 2 < (F.2)
 ( + k)
k=0
covers the domain of the MittagLeer function used in viscoelasticity.
The MittagLeer function of restricted domain specified in Eq. (F.2) can
be solved for using Algorithm F.1.R This algorithm requires the ability to ex-
!
ecute a numerical integration of 0 0 K(, , !, x) d! to machine precision
m for which the author uses a GaussKronrod quadrature formula readily
available in many numerical packages.

A.D. Freed, Soft Solids: A Primer to the Theoretical Mechanics of Materials, 341
Modeling and Simulation in Science, Engineering and Technology,
DOI 10.1007/978-3-319-03551-2, Springer International Publishing Switzerland 2014
342 Soft Solids

Algorithm F.1 MittagLeffler function


function K (, , !, x) (
1 x exp(!1/ ) !1/ sin((1 + )), =0
K  !2 2!x cos()+x2
sin(), =1
return K
end function.

function MittagLeffler (, , x) F 0 <


1, 2 {0, 1}, x 2 <
if x = 0 then (
0, = 0
E, (x)
1, = 1
else if = 1 then (
x exp(x), = 0
E, (x)
exp(x), =1
else if |x| <  then F  is adjustable,  0.6 for 64-bit reals
k0 dln(m (1 |x|))/ ln(|x|)e F m is machine precision
if = 0 then
k0 max(k0 , d1/e)
end if Pk0 k
E, (x) k=0 x / ( + k)
else if |x| < b10+ 5c then 
!0 max 1, 2|x|, ( ln(m /6))
R !0
E, (x) 0 K(, , !, x) d!
if x > 0 then
E, (x) E, (x) + 1 x(1)/ exp(x1/ )
end if
else
k0 b ln(m )/ ln(|x|)c
Pk0 k
E, (x) k=1 x / ( k)
if x > 0 then
E, (x) E, (x) + 1 x(1)/ exp(x1/ )
end if
end if
return E, (x)
end function.
Bibliography

Aaron, B. B., & Gosline, J. M. (1981). Elastin as a random-network elastomer:


A mechanical and optical analysis of single elastin fibers. Biopolymers, 20,
12471260.
Abramowitz, M., & Stegun, I. A. (Eds.). (1964). Handbook of mathematical
functions: With formulas, graphs, and mathematical tables. Washington, DC:
National Bureau of Standards. Republished by New York, NY: Dover
Publications.
Almansi, E. (1911). Sulle deformazioni finite dei solidi elastici isotropi. Rendi-
conti della Reale Accademia dei Lincei: Classe di scienze fisiche, matematiche
e naturali (Vol. 20, pp. 705714). Roma: LAccademia.
Atluri, S. N., & Cazzani, A. (1995). Rotations in computational solid mechanics.
Archives of Computational Methods in Engineering, 2, 49138.
Bagley, R. L. (1987). Power law and fractional calculus model of viscoelasticity.
AIAA Journal, 27(10), 14121417.
Bagley, R. L. (1991). The thermorheologically complex material. International
Journal of Engineering Science, 29, 797806.
Bagley, R. L., & Torvik, P. J. (1983). A theoretical basis for the application of
fractional calculus to viscoelasticity. Journal of Rheology, 27, 201210.
Baleanu, D., Diethelm, K., Scalas, E., & Trujillo, J. J. (2012). Fractional calculus
models and numerical methods. Series on complexity, nonlinearity and chaos
(Vol. 3). Singapore: World Scientific.
Bell, J. F. (1983). Continuum plasticity at finite strain for stress paths of arbitrary
composition and direction. Archive for Rational Mechanics and Analysis, 84,
139170.
Belytschko, T., Liu, W. K., & Moran, B. (2000). Nonlinear finite elements for
continua and structures. Chichester: Wiley.
Bernstein, B. (1960). Hypo-elasticity and elasticity. Archive for Rational Mechan-
ics and Analysis, 6, 90104.

A.D. Freed, Soft Solids: A Primer to the Theoretical Mechanics of Materials, 343
Modeling and Simulation in Science, Engineering and Technology,
DOI 10.1007/978-3-319-03551-2, Springer International Publishing Switzerland 2014
344 Soft Solids

Bernstein, B., & Rajagopal, K. R. (2008). Thermodynamics of hypoelasticity.


Zeitschrift fr angewandte Mathematik und Physik, 59, 537553.
Bernstein, B., & Shokooh, A. (1980). The stress clock function in viscoelasticity.
Journal of Rheology, 24, 189211.
Bernstein, B., Kearsley, E. A., & Zapas, L. J. (1963). A study of stress relaxation
with finite strain. Transactions of the Society of Rheology, 7, 391410.
Bernstein, B., Kearsley, E. A., & Zapas, L. J. (1964). Thermodynamics of per-
fect elastic fluids. Journal of Research of the National Bureau of StandardsB.
Mathematics and Mathematical Physics, 68B, 103113.
Biot, M. A. (1939). Non-linear theory of elasticity and the linearized case for a
body under initial stress. London, Edinburgh and Dublin Philosophical Maga-
zine and Journal of Science, 27, 468489.
Bird, R. B., Armstrong, R. C., & Hassager, O. (1987a). Dynamics of polymeric
liquids: Fluid mechanics (2nd ed., Vol. 1). New York: Wiley.
Bird, R. B., Curtiss, C. F., Armstrong, R. C., & Hassager, O. (1987b). Dynamics
of polymeric liquids: Kinetic theory (2nd ed., Vol. 2). New York: Wiley.
Blatz, P. J., Chu, B. M., & Wayland, H. (1969). On the mechanical behavior of
elastic animal tissue. Transactions of the Society of Rheology, 13(1), 83102.
Boltzmann, L. (1874). Zur Theorie der elastischen Nachwirkung. Sitzungsberichte
der Mathematisch-Naturwissenschaftlichen Classe der Kaiserlichen Akademie
der Wissenschaften, Wien, 70(2), pp. 275300).
Bonet, J., & Wood, R. D. (1997). Nonlinear continuum mechanics for finite ele-
ment analysis. Cambridge: Cambridge University Press.
Bowen, R. M. (1989). Introduction to continuum mechanics for engineers. Math-
ematical concepts and methods in science and engineering (Vol. 39). New
York: Plenum Press. (Republished by Mineola, NY: Dover Publications, re-
vised, 2007, 2009)
Bra, H. (1977). Quadraturverfahren. Studia mathematica (Vol. 3). Gttingen:
Vandenhoeck & Ruprecht.
Bridgman, P. W. (1923). The compressibility of thirty metals as a function of
pressure and temperature. Proceedings of the American Academy of Arts and
Sciences, 58, 165242.
Brunner, H. (2004). Collocation methods for Volterra integral and related func-
tional equations. Cambridge monographs on applied and computational math-
ematics (Vol. 15). Cambridge: Cambridge University Press.
Butcher, J. C. (2008). Numerical methods for ordinary dierential equations (2nd
ed.). Chichester: Wiley.
Butcher, J. C., & Podhaisky, H. (2006). On error estimation in general linear meth-
ods for sti ODEs. Applied Numerical Mathematics, 56, 345357.
Caputo, M. (1967). Linear models of dissipation whose Q is almost frequency
independent II. Geophysical Journal of the Royal Astronomical Society, 13,
529539.
Bibliography 345

Caputo, M., & Mainardi, F. (1971b) A new dissipation model based on memory
mechanism. Pure and Applied Geophysics, 91, 134147.
Caputo, M., & Mainardi, F. (1971a). Linear models of dissipation in anelastic
solids. Rivista del Nuovo Cimento, 1, 161198.
Carton, R. W., Dainauskas, J., & Clark, J. W. (1962). Elastic properties of single
elastic fibers. Journal of Applied Physiology, 17(3), 547551.
Catsi, E., & Tobolsky, A. V. (1955). Stress-relaxation of polyisobutylene in the
transition region (1,2). Journal of Colloid and Interface Science, 10, 375392.
Cauchy, A. -L. (1827) Exercices de mathmatiques (Vol. 2). Paris: de Bure Frres.
Chadwick, P. (1976), Continuum Mechanics: Concise theory and problems.
London: George Allen & Unwin. (Republished by Mineola, NY: Dover Publi-
cations, 2nd ed., 1999)
Cheng, H., & Gupta, K. C. (1989). An historical note on finite rotations. Journal
of Applied Mechanics, 56, 139145.
Christensen, R. M. (1971). Theory of viscoelasticity: An introduction. New York:
Academic.
Cole, K. S., & Cole, R. H. (1941). Dispersion and absorption in dielectrics: I.
Alternating current characteristics. Journal of Chemical Physics, 9, 341351.
Cole, K. S., & Cole, R. H. (1942). Dispersion and absorption in dielectrics: II.
Direct current characteristics. Journal of Chemical Physics, 10, 98105.
Coleman, B. D., & Mizel, V. J. (1968). On the general theory of fading memory.
Archive for Rational Mechanics and Analysis, 29, 1831.
Coleman, B. D., & Noll, W. (1961). Foundations of linear viscoelasticity. Reviews
of Modern Physics, 33(2), 239249.
Coleman, B. D., & Noll, W. (1964). Simple fluids with fading memory. In
M. Reiner & D. Abir (Eds.), Second-order eects in elasticity, plasticity and
fluid dynamics (pp. 530551). New York: Pergamon Press.
Criscione, J. C., Sacks, M. S., & Hunter, W. C. (2003a). Experimentally tractable,
pseudo-elastic constitutive law for biomembranes: I. theory. Journal of Biome-
chanical Engineering, 125, 9499.
Criscione, J. C., Sacks, M. S., & Hunter, W. C. (2003b) Experimentally tractable,
pseudo-elastic constitutive law for biomembranes: II application. Journal of
Biomechanical Engineering, 125, 100105.
Demiray, H. (1972). A note on the elasticity of soft biological tissues. Journal of
Biomechanics, 5, 309311.
Dienes, J. K. (1979). On the analysis of rotation and stress rate in deforming bod-
ies. Acta Mechanica, 32, 217232.
Dienes, J. K. (2003). Finite deformation of materials with an ensemble of de-
fects. Technical report LA13994MS. Los Alamos: Los Alamos National
Laboratory.
346 Soft Solids

Diethelm, K. (2010). The analysis of fractional dierential equations: An appli-


cation-oriented exposition using dierential operators of Caputo type. Lecture
notes in mathematics (Vol. 2004). Heidelberg: Springer.
Diethelm, K., & Freed, A. D. (2006). An ecient algorithm for the evaluation
of convolution integrals. Computers and Mathematics with Applications, 51,
5172.
Diethelm, K., Ford, N. J., Freed, A. D., & Luchko, Y. (2005). Algorithms for the
fractional calculus: A selection of numerical methods. Computer Methods in
Applied Mechanics and Engineering, 194, 743773.
Doehring, T. C., Freed, A. D., Carew, E. O., & Vesely, I. (2005). Fractional order
viscoelasticity of the aortic valve cusp: an alternative to quasilinear viscoelas-
ticity. Journal of Biomechanical Engineering, 127, 700708.
Dokos, S., LeGrice, I. J., Smaill, B. H., Kar, J., & Young, A. A. (2000). A triaxi-
al-measurement shear-test device for soft biological tissues. Journal of Biome-
chanical Engineering, 122, 471478.
Dokos, S., Smaill, B. H., Young, A. A., & LeGrice, I. J. (2002). Shear properties
of passive ventricular myocardium. American Journal of PhysiologyHeart and
Circulatory Physiology, 283, H2650H2659.
Douglas, J. F. (2000). Polymer science applications of path-integration, integral
equations, and fractional calculus. In R. Hilfer (Ed.), Applications of fractional
calculus in physics (pp. 241330). Singapore: World Scientific.
Drucker, D. C. (1959). A definition of stable inelastic material. Journal of Applied
Mechanics, 27, 101106.
Duenwald, S. E., Vanderby, R., Jr., & Lakes, R. S. (2010). Stress relaxation and
recovery in tendon and ligament: Experiment and modeling. Biorheology, 47,
114.
Einstein, A. (1933). On the method of theoretical physics. New York: Oxford
University Press. (The Herbert Spencer lecture delivered at Oxford, 10 June
1933)
Erdlyi, A., Magnus, W., Oberhettinger, F., & Tricomi, F. G. (1955). Higher tran-
scendental functions. Bateman manuscript project (Vol. 2). New York: Mc-
Graw-Hill.
Ericksen, J. L. (1958). Hypo-elastic potentials. Quarterly Journal of Mechanics
and Applied Mathematics, 11, 6772.
Ferry, J. D. (1980). Viscoelastic properties of polymers (3rd ed.). New York:
Wiley.
Finger, J. (1894). ber die allgemeinsten beziehungen zwischen endlichen de-
formationen und den zugehrigen spannungen in aeolotropen und isotropen
substanzen. Sitzungsberichte der Akademie der Wissenschaften, Wien, 103,
10731100.
Fitzgerald, J. E. (1980). A tensorial Hencky measure of strain and strain rate for
finite deformations. Journal of Applied Physics, 51, 51115115.
Bibliography 347

Flanagan, D. P., & Taylor, L. M. (1987). An accurate numerical algorithm for


stress integration with finite rotations. Computer Methods in Applied Mechan-
ics and Engineering, 62, 305320.
Ford, N. J., & Simpson, A. C. (2001). The numerical solution of fractional dier-
ential equations: Speed versus accuracy. Numerical Algorithms, 26, 333346.
Freed, A. D. (1995). Natural strain. Journal of Engineering Materials and Tech-
nology, 117, 379385.
Freed, A. D. (2008). Anisotropy in hypoelastic soft-tissue mechanics, I: Theory.
Journal of Mechanics of Materials and Structures, 3(5), 911928.
Freed, A. D. (2009). Anisotropy in hypoelastic soft-tissue mechanics, II: Simple
extensional experiments. Journal of Mechanics of Materials and Structures,
4(6), 10051025.
Freed, A. D. (2010). Hypoelastic soft tissues, part I: Theory. Acta Mechanica, 213,
189204.
Freed, A. D., & Diethelm, K. (2006). Fractional calculus in biomechanics: A 3D
viscoelastic model using regularized fractional derivative kernels with applica-
tion to the human calcaneal fat pad. Biomechanics and Modeling in Mechanobi-
ology, 5, 203215.
Freed, A. D., & Diethelm, K. (2007). Caputo derivatives in viscoelasticity: A non-
linear finite-deformation theory for tissue. Fractional Calculus and Applied
Analysis, 10(3), 219248.
Freed, A. D., & Doehring, T. C. (2005). Elastic model for crimped collagen fibrils.
Journal of Biomechanical Engineering, 127, 587593.
Freed, A. D., & Einstein, D. R. (2012). Hypo-elastic model for lung parenchyma.
Biomechanics and Modeling in Mechanobiology, 11, 557573.
Freed, A. D., & Einstein, D. R. (2013). An implicit elastic theory for lung
parenchyma. International Journal of Engineering Science, 62, 3147.
Freed, A. D., Einstein, D. R., & Sacks, M. S. (2010). Hypoelastic soft tissues, part
II: In-plane biaxial experiments. Acta Mechanica, 213, 205222.
Freed, A. D., Einstein, D. R., & Vesely, I. (2005). Invariant formulation for dis-
persed transverse isotropy in aortic heart valves: An ecient means for model-
ing fiber splay. Biomechanics and Modeling in Mechanobiology, 4, 100117.
Fulchiron, R., Verney, V., Cassagnau, P., Michael, A., Levoir, P., & Aubard, J.
(1993). Deconvolution of polymer melt stress relaxation by the Pad-Laplace
method. Journal of Rheology, 37, 1734.
Fung, Y. C. (1967). Elasticity of soft tissues in simple elongation. American Jour-
nal of Physiology, 28, 15321544.
Fung, Y. -C. (1971). Stressstrain-history relations of soft tissues in simple
elongation. In Y. -C. Fung, N. Perrone, & M. Anliker (Eds.), Biomechan-
ics: Its foundations and objectives, chap. 7 (pp. 181208). Englewood Clis:
Prentice-Hall.
Fung, Y. -C. (1973). Biorheology of soft tissues. Biorheology, 10, 139155.
348 Soft Solids

Fung, Y. C. (1993). Biomechanics: Mechanical properties of living tissues (2nd


ed.). New York: Springer.
Gent, A. N. (1996). A new constitutive relation for rubber. Rubber Chemsitry and
Technology, 69, 5961.
Gittus, J. (1975). Creep, viscoelasticity and creep rupture in solids. New York:
Halsted Press.
Goldberg, D. E. (1989). Genetic algorithms in search, optimization, and machine
learning. Boston: Addison-Wesley.
Goldberg, D. E. (2002). The design of innovation: Lessons learned from and for
competent genetic algorithms. Genetic algorithms and evolutionary computa-
tion (Vol. 7). Boston: Kluwer.
Gorenflo, R., & Rutman, R. (1995). On ultraslow and intermediate processes. In
P. Rusev, I. Dimovski, & V. Kiryakova (Eds.) Transform methods and special
functions, sofia 1994 (pp. 6181). Singapore: Science Culture Technology Pub-
lishing.
Gorenflo, R., Loutchko, I., & Luchko, Y. (2002). Computation of the Mittag-L-
eer function E_, (z) and its derivatives. Fractional Calculus and Applied
Analysis, 5, 491518. [Erratum: 6, 111112 (2003)]
Graham, A. (1981). Kronecker products and matrix calculus: With applications.
Ellis Horwood series in mathematics and its applications. Chichester: Ellis
Horwood Limited.
Green, G. (1841). On the propagation of light in crystallized media. Transactions
of the Cambridge Philosophical Society, 7, 121140.
Gross, B. (1937). ber die anomalien der festen dielektrika. Zeitschrift fr Physik,
107, 217234.
Gross, B. (1938). Zum verlauf des einsatzstromes im anomalen dielektrikum.
Zeitschrift fr Physik, 108, 598608.
Gross, B. (1947). On creep and relaxation. Journal of Applied Physics, 18,
212221.
Gurtin, M. E. (1981). An introduction to continuum mechanics. Mathematics in
science and engineering (Vol. 158). New York: Academic.
Gurtin, M. E., Fried, E., & Anand, L. (2010). The mechanics and thermodynamics
of continua. Cambridge: Cambridge University Press.
Guth, E., Wack, P. E., & Anthony, R. L. (1946). Significance of the equation of
state for rubber. Journal of Applied Physics, 17, 347351.
Hart, J. F., Cheney, E. W., Lawson, C. L., Maehly, H. J., Mesztenyi, C. K., Rice,
J. R., et al. (1968). Computer approximations. The SIAM series in applied
mathematics. New York: Wiley.
Havner, K. S. (1992). Finite plastic deformation of crystalline solids. Cambridge
monographs on mechanics and applied mathematics. Cambridge: Cambridge
University Press.
Bibliography 349

Hencky, H. (1928). ber die Form des Elastizittsgesetzes bei ideal elastischen
Stoen. Zeitschrift fr technische Physik, 9, 215220. (Translated from German
to English in NASA TT-21602, Washington DC, 1994)
Hencky, H. (1931). The law of elasticity for isotropic and quasi-isotropic sub-
stances by finite deformations. Journal of Rheology, 2, 169176.
Herrmann, R. (2011). Fractional calculus: An introduction for physicsts.
Singapore: World Scientific.
Hilfer, R., & Seybold, H. J. (2006). Computation of the generalized Mittag-Leer
function and its inverse in the complex plane. Integral Transforms and Special
Functions, 17, 637652.
Hill, R. (1957). On uniqueness and stability in the theory of finite elastic strain.
Journal of the Mechanics and Physics of Solids, 5, 229241.
Hill, R. (1958). A general theory of uniqueness and stability in elastic-plastic
solids. Journal of the Mechanics and Physics of Solids, 6, 236249.
Hill, R. (1968). On constitutive inequalites for simple materials I. Journal of the
Mechanics and Physics of Solids, 16, 229242.
Hoger, A. (1986). The material time derivative of logarithmic strain. International
Journal of Solids and Structures, 22, 10191032.
Holzapfel, G. A. (2000). Nonlinear solid mechanics: A continuum approach for
engineering. Chichester: Wiley.
Humphrey, J. D. (2002a) Cardiovascular solid mechanics; cells, tissues, and
organs. New York: Springer.
Humphrey, J. D. (2002b) Continuum biomechanics of soft biological tissues. Pro-
ceedings of the Royal Society, London A, 459, 346.
Humphrey, J. D. (2008). Biological soft tissues. In W. N. J. Sharpe (Ed.), Springer
handbook of experimental solid mechanics (pp. 169185). New York: Springer.
James, H. M., & Guth, E. (1943). Theory of the elastic properties of rubber. The
Journal of Chemical Physics, 11, 455481.
James, H. M., & Guth, E. (1944). Theory of the elasticity of rubber. Journal of
Applied Physics, 15, 294303.
James, H. M., & Guth, E. (1947). Theory of the increase in rigidity of rubber
during cure. The Journal of Chemical Physics, 15, 669683.
Jaumann, G. (1911). Geschlossenes system physikalischer und chemischer dier-
entialgesetze. Sitzungsberichte der Kaiserlichen Akademie der Wissenschaften:
Mathematisch-naturwissenschaftliche Klasse, 120, 385530.
Kachanov, L. M. (1971). Foundations of the theory of plasticity. North-Holland se-
ries in applied mathematics and mechanics (Vol. 12). Amsterdam: North-Hol-
land Publishing.
Kaye, A. (1962). A non-Newtonian flow in incompressible fluids. Technical report
134. Cranfield: The College of Aeronautics.
350 Soft Solids

Kirchho, G. (1852). ber die Gleichungen des Gleichgewichtes eines elastischen


Krpers bei nicht unendlich kleinen Verschiebungen seiner Theile. Sitzungs-
berichte der Akademie der Wissenschaften, Wien, 9, 763773.
Kohlrausch, R. (1847). Ueber das Dellmannsche Elektrometer. Annalen der
Physik und Chemie, 72(11), 353405.
Lai, W. M., Rubin, D., & Krempl, E. (1974). Introduction to continuum mechan-
ics. Pergamon Unified Engineering Series. New York: Pergamon Press.
Lakes, R. S. (1998). Viscoelastic solids. CRC Mechanical Engineering Series.
Boca Raton: CRC Press.
Leonov, A. I. (1996). On the constitutive equations for nonisothermal bulk relax-
ation. Macromolecules, 29, 83838386.
Leonov, A. (2000). On the conditions of potentiality in finite elasticity and
hypo-elasticity. International Journal of Solids and Structures, 37, 25652576.
Lillie, M. A., & Gosline, J. M. (1996). Swelling and viscoelastic properties of
osmotically stressed elastin. Biopolymers, 39, 641652.
Linke, W. A., & Grtzner, A. (2008). Pulling single molecules of titin by
AFMrecent advances and physiological implications. Pflgers Archiv Eu-
ropean Journal of Physiology, 456, 101115.
Lodge, A. S. (1956). A network theory of flow birefringence and stress in concen-
trated polymer solutions. Transactions of the Faraday Society, 52, 120130.
Lodge, A. S. (1958). A network theory of constrained elastic recovery in concen-
trated polymer solutions. Rheologica Acta, 1, 158163.
Lodge, A. S. (1964). Elastic liquids: An introductory vector treatment of finite-s-
train polymer rheology. London: Academic.
Lodge, A. S. (1974). Body tensor fields in continuum mechanics: With applica-
tions to polymer rheology. New York: Academic.
Lodge, A. S. (1984). A classification of constitutive equations based on stress
relaxation predictions for the single-jump shear strain experiment. Journal of
Non-Newtonian Fluid Mechanics, 14, 6783.
Lodge, A. S. (1999). An introduction to elastomer molecular network theory.
Madison: Bannatek Press.
Lodge, A. S., McLeod, J. B., & Nohel, J. A. (1978). A nonlinear singularly
perturbed Volterra integrodierential equation occurring in polymer rheology.
Proceedings of the Royal Society of Edinburgh, 80A, 99137.
Mainardi, F. (2010). Fractional calculus and waves in linear viscoelasticity. Lon-
don: Imperial College Press.
Mainardi, F., & Gorenflo, R. (2007). Time-fractional derivatives in relaxation
processes: A tutorial survey. Fractional Calculus and Applied Analysis, 10,
269308.
Malvern, L. E. (1969). Introduction to the mechanics of a continuous medium.
Prentice-Hall series in engineering of the physical sciences. Englewood Clis:
Prentice-Hall.
Bibliography 351

Marsden, J. E., & Hughes, T. J. R. (1983). Mathematic foundations of elasticity.


Englewood Clis: Prentice-Hall. Republished by Mineola, NY: Dover Publi-
cations, 1994.
Maxwell, J. C. (1867). On the dynamical theory of gases. Philosophical Transac-
tions of the Royal Society, London, 157, 4988.
McLoughlin, J. R., & Tobolsky, A. V. (1952). The viscoelastic behavior of poly-
methyl methacrylate. Journal of Colloid and Interface Science, 7, 555568.
Meerschaert, M. M., & Sikorskii, A. (2012). Stochastic models for fractional cal-
culus. De Gruyter studies in mathematics (Vol. 43). Berlin: De Gruyter.
Metzler, R., & Klafter, J. (2002). From stretched exponential to inverse power-law:
Fractional dynamics, Cole-Cole relaxation processes, and beyond. Journal of
Non-crystalline Solids, 305, 8187.
Miller, K. S., & Ross, B. (1993). An introduction to the fractional calculus and
fractional dierential equations. New York: John Wiley.
Miller-Young, J. E., Duncan, N. A., & Baroud, G. (2002). Material properties of
the human calcaneal fat pad in compression: Experiment and theory. Journal
of Biomechanics, 35, 15231531.
Mittag-Leer, G. (1904). Sur la reprsentation analytique dune branche uni-
forme dune fonction monogne. Acta Mathematica, 29, 101168.
Moon, H., & Truesdell, C. (1974). Interpretations of adscititious inequalities
through the eects pure shear stress produces upon an isotropic elastic solid.
Archive for Rational Mechanics and Analysis, 55, 117.
Mooney, M. (1940). A theory of large elastic deformations. Journal of Applied
Physics, 11, 582592.
Morgan, F. R. (1960). The mechanical properties of collagen fibres: Stress-strain
curves. Journal of the International Society of Leather Trades Chemists, 44,
170182.
Moyer, A. E. (1977). Robert Hookes ambiguous presentation of Hookes law.
Isis, 68(242), 266275.
Nadeau, J. C., & Ferrari, M. (1998). Invariant tensor-to-matrix mappings for eval-
uation of tensorial expressions. Journal of Elasticity, 52, 4361.
Neubert, H. K. P. (1963). A simple model representing internal damping in solid
matrials. The Aeronautical Quarterly, 14, 187210.
Nicholson, D. W. (2008). Finite element analysis: Thermomechanics of solids
(2nd ed.). Boca Raton: CRC Press.
Nicholson, D. W. (2013). An analysis of invariant convexity in hyperelasticity.
Submitted to International Journal of Engineering Science.
Nicholson, D. W., & Lin, B. (1998). On the tangent modulus tensor in hyperelas-
ticity. ACTA Mechanica, 131, 121131.
Nicholson, D. W., & Lin, B. (1999). Extensions of Kronecker product algebra
with applications in continuum and computational mechanics. ACTA Mechan-
ica, 136, 223241.
352 Soft Solids

Noll, W. (1955). On the continuity of the solid and fluid states. Journal of Rational
Mechanics and Analysis, 4, 381.
Noll, W. (1958). A mathematical theory of the mechanical behavior of continuous
media. Archive for Rational Mechanics and Analysis, 2, 197226.
Noll, W. (1972). A new mathematical theory of simple materials. Archive for Ra-
tional Mechanics and Analysis, 48, 150.
Nowick, A. S., & Berry, B. S. (1972). Anelastic relaxation in crystalline solids.
Materials science series. New York: Academic.
Nutting, P. G. (1921). A new general law of deformation. Journal of the Franklin
Institute, 191, 679685.
Ogden, R. W. (1972). Large deformation isotropic elasticityon the correlation of
theory and experiment for incompressible rubberlike solids. Proceedings of the
Royal Society, London A, 326, 565584.
Ogden, R. W. (1984). Non-linear elastic deformations. New York: John Wiley.
(Republished by Mineola, NY: Dover Publications, 1997)
Oldham, K. B., & Spanier, J. (1974). The fractional calculus: Theory and appli-
cations of dierentiation and integration to arbitrary order. New York: Aca-
demic. (Republished by Mineola, NY: Dover Publications, revised, 2006)
Oldroyd, J. G. (1950). On the formulation of rheological equations of state. Pro-
ceedings of the Royal Society, London A, 200, 523541.
Oldroyd, J. G. (1970). Equations of state of continuous matter in general relativity.
Proceedings of the Royal Society, London A, 316, 128.
Park, S. W., & Schapery, R. A. (1999). Methods of interconversion between linear
viscoelastic material functions, part I: A numerical method based on Prony
series. International Journal of Solids and Structures, 36, 16531675.
Phan-Thien, N. (2002). Understanding viscoelasticity: Basics of rheologoy.
Berlin: Springer.
Piola, G. (1833). La meccanica d corpi naturalmente estesi: Trattata col calcolo
delle variazioni. Opuscoli Matematici e Fisici di Diversi Autori, 1, 201236.
Pipkin, A. C. (1972). Lectures on viscoelasticity theory. Applied mathematical
sciences (Vol. 7). New York: Springer.
Pipkin, A. C., & Rogers, T. G. (1968). A non-linear integral representation for
viscoelastic behaviour. Journal of the Mechanics and Physics of Solids, 16,
5972.
Podlubny, I. (1999). Fractional dierential equations: An introduction to frac-
tional derivatives, fractional dierential equations, to methods of their solu-
tion and some of their applications. Mathematics in science and engineering
(Vol. 198). San Diego: Academic.
Polyanin, A. D., & Zaitsev, V. F. (2003). Handbook of exact solutions for ordinary
dierential equations (2nd ed.). Boca Raton: Chapman & Hall/CRC.
Bibliography 353

Press, W. H., Teukolsky, S. A., Vetterling, W. T., & Flannery, B. P. (2007). Numer-
ical recipies: The art of scientific computing (3rd ed.). Cambridge: Cambridge
University Press.
Puso, M. A., & Weiss, J. A. (1998). Finite element implementation of anisotropic
quasi-linear viscoelasticity using a discrete spectrum approximation. Journal
of Biomechanical Engineering, 120, 6270.
Rajagopal, K. R. (2003). On implicit constitutive theories. Applications of Math-
ematics, 48(4), 279319.
Rajagopal, K. R. (2011a) Conspectus of concepts of elasticity. Mathematics and
Mechanics of Solids, 16, 536562.
Rajagopal, K. R. (2011b) On the cavalier attitude towards referencing. Interna-
tional Journal of Engineering Science. In press.
Rajagopal, K. R., & Srinivasa, A. R. (2007). On the response of non-dissipative
solids. Proceedings of the Royal Society, London A, 463, 357367.
Rajagopal, K. R., & Srinivasa, A. R. (2009). On a class of non-dissipative materi-
als that are not hyperelastic. Proceedings of the Royal Society, London A, 465,
493500.
Rajagopal, K. R., & Wineman, A. (2010). Applications of viscoelastic clock mod-
els in biomechanics. Acta Mechanica, 213, 255266.
Rao, I. J., & Rajagopal, K. R. (2007). The status of the K-BKZ model within
the framework of materials with multiple natural configurations. Journal of
Non-Newtonian Fluid Mechanics, 141, 7984.
Rivlin, R. S., & Saunders, D. W. (1951). Large elastic deformations of isotropic
materials vii. experiments on the deformation of rubber. Philosophical Trans-
actions of the Royal Society, London, A 243, 251288.
Rivlin, R. S., & Smith, G. F. (1969). Orthogonal integrity basis for N symmetric
matrices. In D. Abir (Ed.), Contributions to mechanics (pp. 121141). New
York: Pergamon Press.
Rouse, P. E. (1953). A theory of the linear viscoelastic properties of dilute solu-
tions of coiling polymers. Journal of Chemical Physics, 21, 12721280.
Sacks, M. S. (2000). Biaxial mechanical evaluation of planar biological materials.
Journal of Elasticity, 61, 199246.
Sacks, M. S., & Sun, W. (2003). Multiaxial mechanical behavior of biological
materials. Annual Review of Biomedical Engineering, 5, 251284.
Samko, S. G., Kilbas, A., & Marichev, O. I. (1993). Fractional integrals and
derivatives: Theory and applications. Yverdon: Gordon and Breach.
Schapery, R. A., & Park, S. W. (1999). Methods of interconversion between lin-
ear viscoelastic material functions, part II: An approximate analytical method.
International Journal of Solids and Structures, 36, 16771699.
Scott Blair, G. W. (1944). Analytical and integrative aspects of the stress-strain-
time problem. Journal of Scientific Instruments, 21, 8084.
354 Soft Solids

Signorini, A. (1930). Sulle deformazioni thermoelastiche finite. In C. W. Oseen, &


W. Weibull (Eds.), Proceedings of the 3rd International Congress for Applied
Mechanics (Vol. 2, pp. 8089). Stockholm: Ab. Sveriges Litografiska Tryck-
erier.
Simhambhatla, M., & Leonov, A. (1993). The extended Pad-Laplace method
for ecient discretization of linear viscoelastic spectra. Rheologica Acta, 32,
589600.
Simo, J. C., & Hughes, T. J. R. (1998) Computational inelasticity. Interdisci-
plinary applied mathematics (Vol. 7). New York: Springer.
Smith, J. C., & Stamenovic, D. (1986). Surface forces in lungs. I. Alveolar sur-
face tension-lung volume relationships. Journal of Applied Physiology, 60(4),
13411350.
Sokolniko, I. S. (1964). Tensor analysis: Theory and applications to geometry
and mechanics of continua (2nd ed.). Applied Mathematics Series. New York:
Wiley.
Spencer, A. J. M. (1972). Deformations in fibre-reinforced materials. Oxford sci-
ence research papers. Oxford: Clarendon Press.
Stamenovic, D., & Smith, J. C. (1986a). Surface forces in lungs. II. Microstruc-
tural mechanics and lung stability. Journal of Applied Physiology, 60(4),
13511357.
Stamenovic, D., & Smith, J. C. (1986b). Surface forces in lungs. III. Alveolar
surface tension and elastic properties of lung parenchyma. Journal of Applied
Physiology, 60(4), 13581362.
Stouer, D. C. & Dame, L. T. (1996). Inelastic deformation of metals: Models,
mechanical properties, and metallurgy. New York: Wiley
Stuebner, M. & Haider, M. A. (2010). A fast quadrature-based numerical method
for the continuous spectrum biphasic poroviscoelastic model of articular carti-
lage. Journal of Biomechanics, 43, 18351839.
Thomas, T. Y. (1955). On the structure of the stress-strain relations. Proceed-
ings of the National Academy of Sciences of the United States of America, 41,
716719.
Tobolsky, A. V. (1956). Stress relaxation studies of the viscoelastic properties of
polymers. Journal of Applied Physics, 27, 673685.
Tobolsky, A. V. (1960). Properties and structure of polymers. New York: Wiley
Tobolsky, A. V. & Mercurio, A. (1959). Oxidative degradation of polydiene vul-
canizates. Journal of Applied Polymer Science, 2, 186188.
Treloar, L. R. G. (1975). The physics of rubber elasticity (3rd ed.). Oxford: Claren-
don Press.
Truesdell, C. (1953). The mechanical foundations of elasticity and fluid dynamics.
Journal of Rational Mechanics and Analysis, 2, 593616.
Truesdell, C. (1955). Hypoelasticity. Journal of Rational Mechanics and Analysis,
4, 83133.
Bibliography 355

Truesdell, C. (1956). Hypo-elastic shear. Journal of Applied Physics, 27, 441447.


Truesdell, C. (1958). Geometric interpretation for the reciprocal deformation ten-
sors. Quarterly of Applied Mathematics, 15, 434435.
Truesdell, C. (1961). Stages in the development of the concept of stress. Problems
of continuum mechanics (Muskhelisvili anniversary volume) (pp. 556564).
Philadelphia: Society of Industrial and Applied Mathematics.
Truesdell, C. & Noll, W. (2004). The non-linear field theories of mechanics (3rd
ed.). Berlin: Springer.
Truesdell, C. & Toupin, R. (1960). The classical field theories. In S. Flgge (Ed.),
Encyclopedia of physics. Principles of classical mechanics and field theory
(Vol. III/1, pp. 226793). Berlin: Springer.
Tschoegl, N. W. (1989). The phenomenological theory of linear viscoelastic be-
havior: An introduction. Berlin: Springer.
Veronda, D. R. & Westmann, R. A. (1970) Mechanical characterization of skin:
Finite deformations. Journal of Biomechanics, 3, 111124.
Viidik, A. (1973). Functional properties of collagenous tissues. International Re-
view of Connective Tissue Research, 6, 127215.
Vito, R. (1973). A note on arterial elasticity. Journal of Biomechanics, 6, 561564.
Volterra, V. (1930). Theory of functionals and of integral and integro-dierential
equations. Glasgow: Blackie and Son. Republished by Mineola, NY: Dover
Publications.
Wang, M. C. & Guth, E. (1952). Statistical theory of networks of non-Gaussian
flexible chains. The Journal of Chemical Physics, 20, 11441157.
Wang, K., Hu, Y. & He, J. (2012). Deformation cycles of subduction earthquakes
in a viscoelastic Earth. Nature, 484, 327332.
Wei, J. (1975). Least square fitting of an elephant. Chemtech, 5, 128129.
Weiss, J. A. & Gardiner, J. C. (2001). Computational modeling of ligament me-
chanics. Critical Reviews in Biomedical Engineering, 29, 303371.
Williams, M. L. (1964). Structural analysis of viscoelastic materials. AIAA Jour-
nal, 2(5), 785808.
Williams, G. & Watts, D. C. (1970). Non-symmetrical dielectric relaxation be-
haviour arising from a simple empirical decay function. Transactions of the
Faraday Society, 66, 8085.
Williams, M. L., Landel, R. F. & Ferry, J. D. (1955). The temperature dependence
of relaxation mechanisms in amorphous polymers and other glass-forming liq-
uids. Journal of the American Chemical Society, 77, 37013707.
Wineman, A. (2009). Nonlinear viscoelastic solidsa review. Mathematics and
Mechanics of Solids, 14, 300366.
Wineman, A. & Min, J. -H. (2003). Time dependent scission and cross-linking in
an elastomeric cylinder undergoing circular shear and heat conduction. Inter-
national Journal of Non-Linear Mechanics, 38, 969983.
356 Soft Solids

Wineman, A. S. & Rajagopal, K. R. (2000). Mechanical response of polymers, an


introduction. Cambridge: Cambridge University Press.
Zaremba, S. (1903). Sur une forme perfectionne de la thorie de la relaxation.
Bulletin de lAcadmie de Cracovie, 8, pp. 594614.
Zener, C. (1948). Elasticity and anelasticity of metals. Chicago: University of
Chicago Press.
Zhu, W., Lai, W. M. & Mow, V. C. (1991). The density and strength of proteogly-
can-proteoglycan interaction sites in concentrated solutions. Journal of Biome-
chanics, 24, 10071018.
Zimm, B. H. (1956). Dynamics of polymer molecules in dilute solution: vis-
coelasticity, flow birefringence and dielectric loss. Journal of Chemical
Physics, 24, 269278.
Index

Aaron and Gosline (1981), 156, 159, Bonet and Wood (1997), 61, 88, 180,
343 344
Abramowitz and Stegun (1964), 229, Bowen (1989), ix, 344
343 Bra (1977), 218, 331, 344
Almansi (1911), 50, 234, 343 Bridgman (1923), 61, 125, 187, 344
Atluri and Cazzani (1995), 34, 35, 343 Brunner (2004), 218, 344
Bagley and Torvik (1983), 222, 343 Butcher and Podhaisky (2006), 317,
Bagley (1987), 228, 343 319, 344
Bagley (1991), 222, 343 Butcher (2008), 284, 317, 344
Baleanu et al. (2012), 222, 343 Caputo and Mainardi (1971a), 216,
222, 224, 259, 345
Bell (1983), 73, 343
Caputo and Mainardi (1971b), 216,
Belytschko et al. (2000), 33, 61, 88,
222, 224, 259, 345
111, 128, 180, 310, 343
Caputo (1967), 222, 224, 229, 344
Bernstein and Rajagopal (2008), 167,
Carton et al. (1962), 164, 187, 192,
344
345
Bernstein and Shokooh (1980), 212,
Catsiff and Tobolsky (1955), 260, 261,
344
265, 266, 345
Bernstein et al. (1963), 238, 344 Cauchy (1827), 49, 71, 77, 79, 91,
Bernstein et al. (1964), 239, 344 236, 345
Bernstein (1960), 161, 343 Chadwick (1976), ix, 289, 345
Biot (1939), 73, 146, 344 Cheng and Gupta (1989), 35, 345
Bird et al. (1987a), 9, 97, 142, 209, Christensen (1971), 209, 212, 222, 345
212, 344 Cole and Cole (1941), 228, 258, 262,
Bird et al. (1987b), 9, 142, 209, 222, 345
344 Cole and Cole (1942), 228, 258, 262,
Blatz et al. (1969), 165, 344 345
Boltzmann (1874), 209, 210, 219, 221, Coleman and Mizel (1968), 211, 216,
223, 243, 344 345

A.D. Freed, Soft Solids: A Primer to the Theoretical Mechanics of Materials, 357
Modeling and Simulation in Science, Engineering and Technology,
DOI 10.1007/978-3-319-03551-2, Springer International Publishing Switzerland 2014
358 Soft Solids

Coleman and Noll (1961), 209, 345 Freed (2010), 58, 60, 162, 167, 292,
Coleman and Noll (1964), 238, 345 302, 347
Criscione et al. (2003a), 161, 184, 345 Fulchiron et al. (1993), 233, 347
Criscione et al. (2003b), 161, 184, 345 Fung (1967), 162164, 187, 347
Demiray (1972), 165, 345 Fung (1971), 164, 218, 219, 229, 347
Dienes (1979), 33, 345 Fung (1973), 164, 347
Dienes (2003), 35, 345 Fung (1993), 162, 164, 214, 219, 347
Diethelm and Freed (2006), 218, 230, Gent (1996), 148, 348
329331, 346 Gittus (1975), 210, 348
Diethelm et al. (2005), 223, 227, 341, Goldberg (1989), 148, 261, 263, 348
346 Goldberg (2002), 148, 261, 263, 348
Diethelm (2010), 222, 224, 346 Gorenflo and Rutman (1995), 225,
Doehring et al. (2005), 222, 259, 346 348
Dokos et al. (2000), 194, 212, 346 Gorenflo et al. (2002), 223, 341, 348
Dokos et al. (2002), 194, 267, Graham (1981), 310312, 348
269271, 273275, 346 Green (1841), 49, 50, 53, 110, 167,
Douglas (2000), 224, 228, 346 170, 234, 236, 348
Drucker (1959), 86, 177, 346 Gross (1937), 224, 348
Duenwald et al. (2010), 212, 346 Gross (1938), 224, 348
Einstein (1933), xxix, 346 Gross (1947), 223, 224, 348
Erdlyi et al. (1955), 341, 346 Gurtin et al. (2010), ix, 348
Ericksen (1958), 167, 346 Gurtin (1981), ix, 348
Ferry (1980), 97, 209, 212, 247, 346 Guth et al. (1946), 219, 220, 244, 245,
Finger (1894), 49, 237, 346 252, 348
Fitzgerald (1980), 51, 346 Hart et al. (1968), 231, 348
Flanagan and Taylor (1987), 34, 346 Havner (1992), 210, 348
Ford and Simpson (2001), 330, 347 Hencky (1928), 18, 51, 61, 124, 191,
Freed and Diethelm (2006), 223, 229, 349
347 Hencky (1931), 51, 61, 124, 125, 146,
Freed and Diethelm (2007), 162, 347 182, 187, 349
Freed and Doehring (2005), 215, 347 Herrmann (2011), 222, 349
Freed and Einstein (2012), 162, 167, Hilfer and Seybold (2006), 223, 349
168, 347 Hill (1957), 86, 349
Freed and Einstein (2013), 61, 125, Hill (1958), 86, 349
131133, 162, 164, 167, 168, 174, Hill (1968), 47, 86, 88, 121, 124, 129,
347 177, 349
Freed et al. (2005), 162, 314, 347 Hoger (1986), 51, 349
Freed et al. (2010), 14, 42, 162, 167, Holzapfel (2000), ix, xxxi, xxxii, 7,
197, 198, 203205, 347 25, 30, 49, 81, 109, 110, 122, 128,
Freed (1995), 52, 347 147, 161, 176, 209, 279, 289, 298,
Freed (2008), 162, 167, 347 349
Freed (2009), 162, 167, 347 Humphrey (2002a), 14, 162, 349
Index 359

Humphrey (2002b), 162, 349 Miller-Young et al. (2002), 200, 206,


Humphrey (2008), 162, 349 233, 351
James and Guth (1943), 122, 123, 349 Mittag-Leffler (1904), 223, 341, 351
James and Guth (1944), 122, 123, 349 Moon and Truesdell (1974), 119, 135,
James and Guth (1947), 122, 349 351
Jaumann (1911), 175, 349 Mooney (1940), 123, 351
Kachanov (1971), 210, 349 Morgan (1960), 192, 351
Kaye (1962), 238, 349 Moyer (1977), 74, 351
Kirchhoff (1852), 79, 82, 83, 91, 350 Nadeau and Ferrari (1998), 310, 351
Kohlrausch (1847), 227, 350 Neubert (1963), 229, 230, 351
Lai et al. (1974), ix, 350 Nicholson and Lin (1998), 310, 312,
Lakes (1998), xxxv, 209, 212, 214, 313, 351
350 Nicholson and Lin (1999), 310, 312,
Leonov (1996), 247, 350 315, 351
Leonov (2000), 167, 350 Nicholson (2008), 9, 178, 310313,
Lillie and Gosline (1996), 212, 350 351
Linke and Grtzner (2008), 162, 350 Nicholson (2013), 86, 129, 177, 351
Lodge et al. (1978), 241, 350 Noll (1955), 161, 219, 351
Lodge (1956), 51, 216, 241, 350 Noll (1958), 28, 86, 236, 306, 352
Lodge (1958), 51, 241, 350 Noll (1972), 86, 352
Lodge (1964), ix, 13, 4951, 54, 55,
Nowick and Berry (1972), 210, 352
95, 194, 209, 212, 216, 234, 289,
Nutting (1921), 228, 352
292, 350
Ogden (1972), 146, 148, 352
Lodge (1974), 911, 13, 27, 50, 51, 55,
Ogden (1984), ix, 30, 81, 86, 146148,
82, 97, 119, 209, 212, 289, 292, 350
161, 352
Lodge (1984), 142, 350
Lodge (1999), 55, 110, 122, 350 Oldham and Spanier (1974), 222, 223,
352
Mainardi and Gorenflo (2007), 222,
350 Oldroyd (1950), 82, 84, 175, 176, 298,
Mainardi (2010), xxxv, 209, 210, 212, 300, 302, 304, 305, 352
214, 216, 220224, 229, 350 Oldroyd (1970), 84, 298, 302, 352
Malvern (1969), 9, 350 Park and Schapery (1999), 233, 352
Marsden and Hughes (1983), 27, 28, Phan-Thien (2002), 209, 210, 212, 352
127, 161, 181, 281, 289, 350 Piola (1833), 49, 82, 83, 237, 352
Maxwell (1867), 221, 351 Pipkin and Rogers (1968), 238, 352
McLoughlin and Tobolsky (1952), Pipkin (1972), 209, 214, 220, 352
266, 268, 272, 351 Podlubny (1999), 222, 223, 352
Meerschaert and Sikorskii (2012), Polyanin and Zaitsev (2003), 190, 352
222, 224, 351 Press et al. (2007), 179, 229, 353
Metzler and Klafter (2002), 221, 224, Puso and Weiss (1998), 231, 233, 353
227, 351 Rajagopal and Srinivasa (2007), 161,
Miller and Ross (1993), 222, 351 164, 167169, 185, 353
360 Soft Solids

Rajagopal and Srinivasa (2009), 161, Truesdell and Toupin (1960), ix, xxxi,
164, 167, 168, 185, 353 9, 289, 355
Rajagopal and Wineman (2010), 212, Truesdell (1953), 176, 354
353 Truesdell (1955), 161, 170, 354
Rajagopal (2003), 161, 164, 167169, Truesdell (1956), 161, 354
172, 186, 353 Truesdell (1958), 56, 355
Rajagopal (2011a), 110, 353 Truesdell (1961), 80, 355
Rajagopal (2011b), 6, 353 Tschoegl (1989), 209, 212, 220, 228,
Rao and Rajagopal (2007), 239, 353 230, 233, 355
Rivlin and Saunders (1951), 123, 353
Veronda and Westmann (1970), 166,
Rivlin and Smith (1969), 114, 179, 353
355
Rouse (1953), 222, 353
Viidik (1973), 162, 355
Sacks and Sun (2003), 162, 353
Vito (1973), 166, 355
Sacks (2000), 14, 162, 353
Samko et al. (1993), 222, 353 Volterra (1930), 209, 222, 355
Schapery and Park (1999), 233, 353 Wang and Guth (1952), 123, 355
Scott Blair (1944), 228, 353 Wang et al. (2012), 213, 355
Signorini (1930), 50, 58, 234, 354 Weiss and Gardiner (2001), 162, 355
Simhambhatla and Leonov (1993), Wei (1975), 233, 355
233, 354 Williams and Watts (1970), 227, 355
Simo and Hughes (1998), 30, 37, 61, Williams et al. (1955), 212, 355
128, 180, 209, 218, 230, 231, 354 Williams (1964), 228, 355
Smith and Stamenovic (1986), 247, Wineman and Min (2003), 262, 264,
354 355
Sokolnikoff (1964), ix, xxxii, 24, 27, Wineman and Rajagopal (2000),
28, 48, 53, 82, 289, 354 xxxv, 209, 220, 356
Spencer (1972), 114, 354 Wineman (2009), 209, 355
Stamenovic and Smith (1986a), 247, Zaremba (1903), 175, 356
354 Zener (1948), 210, 220, 221, 356
Stamenovic and Smith (1986b), 247, Zhu et al. (1991), 216, 356
354
Zimm (1956), 222, 356
Stouffer and Dame (1996), 210, 354
Stuebner and Haider (2010), 233, 354
acceleration, 8
Thomas (1955), 176, 354
algorithm
Tobolsky and Mercurio (1959), 262,
267, 354 angular velocity, 34
Tobolsky (1956), 265268, 354 convolution integral, 337
Tobolsky (1960), 212, 262, 263, 354 distortion
Treloar (1975), 110, 122, 123, 139, Cauchy, 66
147149, 152154, 157, 161, 354 Green, 63
Truesdell and Noll (2004), ix, xxxi, 9, IRKS ODE solver, 326
78, 110, 209, 279, 285, 289, 355 Mittag-Leffler function, 342
Index 361

rotation, 36 generalized, 235


stretch, 38 Green, 48
angular velocity, 33 generalized, 236
homogeneous, 25, 67
Berra, Yogi, 107 isochoric, 26, 32
body, 5 mapping, 27
Piola, 48
Cauchys generalized, 237
postulate, 77 planar, 42
theorem, 78 pure shear, 44
Cayley-Hamilton theorem, 287 simple shear, 40
configuration, 5 uniaxial, 38
conservation law derivative
angular momentum, 80 co-rotational, 175
energy, 109 convected, 300
linear momentum, 80 Lie, 67, 298
mass, 25 material, 8, 67, 297
Newtons 2nd law, 79 Oldroyd, 67, 300
constitutive equation, 84, 109 Truesdell, 176
continuum, 5 dilatation, 61
law of, 6, 23 dilation, 61
coordinate system, 5 displacement, 23
coordinates, 5 distortion
Cauchy, 65
data Green, 62
Aaron & Gosline, 160
Dokos, 276 Einstein
Dow Corning, 160, 209 notation, xxxii
McLoughlin & Tobolsky, 276 Einstein, Albert, xxix, 107
Miller-Young, 209 elastic
Sacks, 209 biaxial, 150, 199
Treloar, 160 equi-biaxial, 139, 192
deformation, 23 extension & shear, 151, 199
biaxial, 44 planar, 145, 196
Cauchy, 48 pure shear, 150, 199
generalized, 236 simple shear, 142, 194
causal, 214 uniaxial, 136, 190
equi-biaxial, 39 elasticity
extension & shear, 45 explicit, 109
Finger, 48 Green solid, 110
generalized, 237 Hookean, 127
gradient, 23 implicit, 161
362 Soft Solids

linear, xxxvi kernel


Lodge solid, 112 Cole-Cole, 227
models, 119 exponential, 221
Mooney-Rivlin, 123 exponential integral, 229
neo-Hookean, 122 fractional, 222
Rajagopal, 169 Kohlrausch, Williams & Watt, 227
strain-energy function, 114 Maxwell chain model, 230
Eulerian, 6 memory, 215
Mittag-Leffler, 222
experiment
relaxation, 210
biaxial, 19
kinematics, 5
equi-biaxial, 11
Kronecker analysis, 309
extension & shear, 20 coordinate transformation, 314
membrane, 16 inner-dyadic product, 314
planar, 14 Kronecker product, 311
pure shear, 18 outer-dyadic product, 313
simple shear, 12 ten operator, 312
stress relaxation, 212 vec operator, 310
thin-walled tube, 19
uniaxial, 10 Lagrange multiplier, 85
Lagrangian, 6
fading memory, 211 Lie
field transfer, 292 derivative, 298
weighted, 307 integral, 302
functions Liebnitz rule, 307
exponential integral, 229 Lodge, Arthur, 3
Mittag-Leffler, 223
material
Fungs law, 162
elastin, 160
fat, 209
gage length, 10
myocardium, 267
PDMS, 160
Hooke, Robert, 75
fiber reinforced, 209
resin reinforced, 209
incompressible, 9 pericardium, 197
infinitesimal strain, 47 PMMA, 267
integration polyisobutylene, 260
by parts, 307 polymethyl methacrylate, 276
Lie, 302 rubber, 147, 262
invariants, 114, 179, 287 membrane, 97
isochoric, 10 Mohrs circle, 18
isotropic, 9 motion, 6
Index 363

homogeneous, 7 linear, 52
planar, 8 Lodge, 50, 55
shear free, 10 generalized, 235
planar, 71
Nansons formula, 81 pure shear, 73
Nolls theorem, 161 rate, 57
Signorini, 50
ODE solver, 317 generalized, 236
overstress, 221 simple shear, 69
uniaxial, 67
particle, 5 stress, 77
Pasteur, Louis, vii 1st Piola-Kirchhoff, 81
plane stress, 97, 177 2nd Piola-Kirchhoff, 83
polar decomposition, 28, 40 biaxial, 101
precondition, 214 Biot, 103
pressure, 88 Cauchy, 78
deviatoric, 88
Rajagopal effect, 184, 189 equi-biaxial, 92
rotation, 28 extension & shear, 102
extra, 85
shear magnitude, 13 Kirchhoff, 79
stability, 86, 177 planar, 97
strain, 47 pure shear, 101
additive, 216, 234 rates, 83
Almansi, 50 simple shear, 95
generalized, 236 uniaxial, 90
Bell-Ericksen, 74 stress equations of motion, 80
biaxial, 73 stretch
Biot, 74 areal, 14
deviatoric, 61 axial, 10
dilatation, 61 tensor, 28
distortional volumetric, 61
Green, 63 stretching, 31
Lodge, 65 surface, 27, 54
dual, 51
equi-biaxial, 68 tangent compliance, 172
extension & shear, 73 tangent modulus, 128, 172
geometry, 53 deviatoric, 131, 174
Green, 50, 53 volumetric, 131, 174
generalized, 235 tensor, 279
Guth, 219, 244 components, 279
Hencky, 51 contravariant, 49, 294
364 Soft Solids

covariant, 49, 294 physical, 291


determinant, 285 tangent, 23
inverse, 285 velocity, 8
mixed, 31 gradient, 30
norm, 284 viscoelastic
orthogonal, 286 equi-biaxial, 253
physical, 291 extension & shear, 266
trace, 284 planar, 255
traction, 77 pure shear, 265
simple shear, 254
vector, 277 uniaxial, 252
base, 5 viscoelasticity, 209
components, 277 K-BKZ, 238
contravariant, 27, 292 quasi-linear, 218, 243
covariant, 27, 292 vorticity, 31
norm, 279
normal, 27, 54, 81 work, 89

You might also like