pdf

208
GEOMETRY AND MECHANICS OF GROWING, NONLINEARLY ELASTIC PLATES AND MEMBRANES by Joseph Brian McMahon Copyright c Joseph Brian McMahon 2009 A Dissertation Submitted to the Faculty of the GRADUATE INTERDISCIPLINARY PROGRAM IN APPLIED MATHEMATICS In Partial Fulfillment of the Requirements For the Degree of DOCTOR OF PHILOSOPHY In the Graduate College THE UNIVERSITY OF ARIZONA 2009

Upload: w-danizz-swarie

Post on 26-Nov-2015

12 views

Category:

Documents


0 download

TRANSCRIPT

GEOMETRY AND MECHANICS OF GROWING,NONLINEARLY ELASTIC PLATES AND MEMBRANES

by

Joseph Brian McMahon

Copyright c© Joseph Brian McMahon 2009

A Dissertation Submitted to the Faculty of the

GRADUATE INTERDISCIPLINARY PROGRAM IN APPLIEDMATHEMATICS

In Partial Fulfillment of the RequirementsFor the Degree of

DOCTOR OF PHILOSOPHY

In the Graduate College

THE UNIVERSITY OF ARIZONA

2009

UMI Number: 3387379

All rights reserved

INFORMATION TO ALL USERS The quality of this reproduction is dependent upon the quality of the copy submitted.

In the unlikely event that the author did not send a complete manuscript

and there are missing pages, these will be noted. Also, if material had to be removed, a note will indicate the deletion.

UMI 3387379

Copyright 2010 by ProQuest LLC. All rights reserved. This edition of the work is protected against

unauthorized copying under Title 17, United States Code.

ProQuest LLC 789 East Eisenhower Parkway

P.O. Box 1346 Ann Arbor, MI 48106-1346

2

THE UNIVERSITY OF ARIZONAGRADUATE COLLEGE

As members of the Dissertation Committee, we certify that we have read the dis-sertation prepared by Joseph Brian McMahonentitled Geometry and Mechanics of Growing, Nonlinearly Elastic Plates and Mem-branesand recommend that it be accepted as fulfilling the dissertation requirement for theDegree of Doctor of Philosophy.

Date: 1 December 2009Alain Goriely

Date: 1 December 2009Michael Tabor

Date: 1 December 2009Shankar Venkataramani

Date: 1 December 2009

Date: 1 December 2009

Final approval and acceptance of this dissertation is contingent upon the candidate’ssubmission of the final copies of the dissertation to the Graduate College.I hereby certify that I have read this dissertation prepared under my direction andrecommend that it be accepted as fulfilling the dissertation requirement.

Date: 1 December 2009Dissertation Director: Alain GorielyDissertation Co-Director: Michael Tabor

3

STATEMENT BY AUTHOR

This dissertation has been submitted in partial fulfillment of requirements for anadvanced degree at the University of Arizona and is deposited in the UniversityLibrary to be made available to borrowers under rules of the Library.

Brief quotations from this dissertation are allowable without special permission,provided that accurate acknowledgment of source is made. Requests for permissionfor extended quotation from or reproduction of this manuscript in whole or in partmay be granted by the copyright holder.

SIGNED: Joseph Brian McMahon

4

ACKNOWLEDGEMENTS

I would like to thank The University of Arizona’s Graduate College and the Pro-gram in Applied Mathematics for their generous financial support during my firstyear in Arizona. I am indebted to the Department of Mathematics’ committeefor reviewing applications for funding through the National Science Foundation’sVertical InteGration of Research and Education (VIGRE) program, without whichmy research would have proceeded much more slowly during the last two years.I would also like to express my gratitude for support under Army Research OfficeMURI Award 50342-PH-MUR (Co-PI: Ildar Gabitov) and NSF grant DMS-0604704(Co-PIs: Alain Goriely and Michael Tabor).

I was able to focus on classwork, seminars, colloquia, and research instead ofon forms, signatures, and deadlines thanks to Linda Silverman, Anne Keyl, andespecially Stacey Wiley, all of the Program in Applied Mathematics. Stacey hasoffered an unending ball of twine through the labyrinth of University regulations.

I have grown as a student and researcher thanks to the faculty of the Program inApplied Mathematics. No one has had greater impact than my advisors, ProfessorMichael Tabor and Professor Alain Goriely, both of the Department of Mathematicsand the Program in Applied Mathematics. I could not have asked for a bettercombination of enthusiasm tempered with caution and an eye for detail matched byan eye locked on the “big picture”.

I sincerely thank you all.

5

DEDICATION

This dissertation is dedicated to Dr. Herbert Wood and Mrs. Olive Lewis of St.

Anselm’s Abbey School, my early inspirations in physical science and mathematics;

to my parents, Bernadette and Jim, who encourage and support me in everything

I try; and most of all to Bridget, who stands beside me not just as wife and best

friend, but as fellow mathematician.

6

TABLE OF CONTENTS

LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

ABSTRACT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

CHAPTER 1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . 15

CHAPTER 2 EUCLIDEAN GEOMETRY WITH CURVILINEAR COORDI-NATES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222.1 Euclidean Point Space . . . . . . . . . . . . . . . . . . . . . . . . . . 222.2 Curvilinear Coordinates on a Body . . . . . . . . . . . . . . . . . . . 23

2.2.1 Example: Cylindrical Coordinates . . . . . . . . . . . . . . . . 242.3 Contravariant and Covariant Vectors . . . . . . . . . . . . . . . . . . 272.4 Metric Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312.5 Christoffel Symbols of the Second Kind . . . . . . . . . . . . . . . . . 322.6 Christoffel Symbols of the First Kind . . . . . . . . . . . . . . . . . . 342.7 Ricci’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 352.8 Further Properties of the Covariant Derivative . . . . . . . . . . . . . 36

2.8.1 Covariant Derivative with Respect to a Vector Field . . . . . . 362.8.2 Covariant Derivative as a Tensor Field . . . . . . . . . . . . . 38

2.9 Curvature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 382.9.1 Example: Cartesian Coordinates . . . . . . . . . . . . . . . . 40

2.10 Abstract Manifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . 402.10.1 Coordinates and Tangent Spaces . . . . . . . . . . . . . . . . 402.10.2 Metric Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . 422.10.3 Linear Connection . . . . . . . . . . . . . . . . . . . . . . . . 432.10.4 Riemannian Geometry . . . . . . . . . . . . . . . . . . . . . . 44

2.11 Isometric Immersion and Embedding of a Riemannian Manifold . . . 452.11.1 Example: Immersion without Embedding . . . . . . . . . . . . 46

CHAPTER 3 KINEMATICS OF DEFORMATION . . . . . . . . . . . . . . 493.1 Deformation versus Change of Coordinates . . . . . . . . . . . . . . . 493.2 Deformation Gradient . . . . . . . . . . . . . . . . . . . . . . . . . . 52

3.2.1 Example: Cartesian Coordinates . . . . . . . . . . . . . . . . 533.2.2 Example: Cylindrical Deformation . . . . . . . . . . . . . . . 54

TABLE OF CONTENTS – Continued

7

3.3 Local Deformation of Euclidean Lines . . . . . . . . . . . . . . . . . . 553.3.1 Example: Cylindrical Deformation . . . . . . . . . . . . . . . 56

3.4 Local Deformation of Volumes . . . . . . . . . . . . . . . . . . . . . . 563.5 Local Deformation of Areas . . . . . . . . . . . . . . . . . . . . . . . 573.6 Differential and Pullback . . . . . . . . . . . . . . . . . . . . . . . . . 59

3.6.1 Example in Cylindrical Coordinates . . . . . . . . . . . . . . . 62

CHAPTER 4 INCOMPATIBLE GROWTH . . . . . . . . . . . . . . . . . . 644.1 Multiplicative Decomposition of Deformation Gradient . . . . . . . . 664.2 Change of Riemannian Metric . . . . . . . . . . . . . . . . . . . . . . 674.3 Example: Immersion without Embedding . . . . . . . . . . . . . . . . 694.4 Example: Lack of Immersion . . . . . . . . . . . . . . . . . . . . . . . 704.5 Non-Metric Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

CHAPTER 5 ELEMENTARY MECHANICS OF SOLIDS . . . . . . . . . . 765.1 Material vs. Spatial Formulation; Conservation of Mass . . . . . . . . 765.2 Other Integral Balance Laws . . . . . . . . . . . . . . . . . . . . . . . 785.3 Stress Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 795.4 Differential Balance Laws . . . . . . . . . . . . . . . . . . . . . . . . 80

5.4.1 Linear Momentum . . . . . . . . . . . . . . . . . . . . . . . . 805.4.2 Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . 82

CHAPTER 6 HYPERELASTICITY . . . . . . . . . . . . . . . . . . . . . . 836.1 Constitutive Relations . . . . . . . . . . . . . . . . . . . . . . . . . . 836.2 Simplifications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 846.3 Stress Power and Hyperelasticity . . . . . . . . . . . . . . . . . . . . 846.4 Objectivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 866.5 Isotropic Hyperelasticity . . . . . . . . . . . . . . . . . . . . . . . . . 86

6.5.1 Isotropic Functions . . . . . . . . . . . . . . . . . . . . . . . . 866.5.2 Form of the Nominal Stress . . . . . . . . . . . . . . . . . . . 87

6.6 Hyperelasticity with Growth . . . . . . . . . . . . . . . . . . . . . . . 886.6.1 Units without Growth . . . . . . . . . . . . . . . . . . . . . . 886.6.2 Units with Growth . . . . . . . . . . . . . . . . . . . . . . . . 89

6.7 Example: Compressible Neo-Hookean . . . . . . . . . . . . . . . . . . 92

CHAPTER 7 CAVITATION IN THE HAUGHTON-OGDEN MEMBRANE 957.0.1 Kinematics of an Axisymmetric Membrane with Growth . . . 967.0.2 Thickness-Averaged Hyperelasticity . . . . . . . . . . . . . . . 977.0.3 Equilibrium Equation . . . . . . . . . . . . . . . . . . . . . . . 977.0.4 Growth-Induced Cavitation . . . . . . . . . . . . . . . . . . . 98

TABLE OF CONTENTS – Continued

8

7.1 Absence of Cavitation in Neo-Hookean Material . . . . . . . . . . . . 997.1.1 λ2 À λ1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1007.1.2 λ1

À∼ λ2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1017.2 Varga Material . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

7.2.1 An Equivalent ODE System . . . . . . . . . . . . . . . . . . . 1047.2.2 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . 1057.2.3 A Necessary Condition . . . . . . . . . . . . . . . . . . . . . . 1087.2.4 Constructing an Appropriate g′(J) . . . . . . . . . . . . . . . 1087.2.5 Converting to an Autonomous System . . . . . . . . . . . . . 1097.2.6 Converting to a Lotka-Volterra System . . . . . . . . . . . . . 1097.2.7 Asymptotics of the Lotka-Volterra System . . . . . . . . . . . 1107.2.8 Boundary Condition at Outer Radius . . . . . . . . . . . . . . 1147.2.9 Numerical Construction of Solutions . . . . . . . . . . . . . . 116

7.3 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

CHAPTER 8 GLOBAL CONSTRAINTS AND THE INDUCED THEORYOF SHELLS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1198.1 Kirchhoff’s Classical Theory of Plates . . . . . . . . . . . . . . . . . . 1198.2 Foppl-von Karman Equations . . . . . . . . . . . . . . . . . . . . . . 1218.3 Plate Theory Induced from Three-Dimensional Elasticity . . . . . . . 122

8.3.1 Global Constraints . . . . . . . . . . . . . . . . . . . . . . . . 1238.3.2 Stress Decomposition . . . . . . . . . . . . . . . . . . . . . . . 1248.3.3 Weak Formulation of Balance Laws . . . . . . . . . . . . . . . 124

CHAPTER 9 THE FLAT KIRCHHOFF PLATE . . . . . . . . . . . . . . . 1269.1 Kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126

9.1.1 Deformation Gradient . . . . . . . . . . . . . . . . . . . . . . 1269.1.2 Incompatible Growth . . . . . . . . . . . . . . . . . . . . . . . 1269.1.3 Elastic Response . . . . . . . . . . . . . . . . . . . . . . . . . 1299.1.4 Virtual Displacements . . . . . . . . . . . . . . . . . . . . . . 129

9.2 Compressible Neo-Hookean Constitutive Relation . . . . . . . . . . . 1329.2.1 A Restriction on γ1 and γ2 at R = 0 . . . . . . . . . . . . . . 134

9.3 Conversion to a Dynamical System . . . . . . . . . . . . . . . . . . . 1349.4 γ1(R) ≡ 1, γ′2(R) > 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . 136

9.4.1 Fixed Points and Linearization . . . . . . . . . . . . . . . . . 1389.4.2 Existence of Solution . . . . . . . . . . . . . . . . . . . . . . . 139

9.5 Numerical Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . 1449.6 γ1(R) ≡ 1, γ′2(R) < 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . 145

TABLE OF CONTENTS – Continued

9

CHAPTER 10 AXISYMMETRIC BUCKLING OF THE KIRCHHOFF PLATE15410.1 Kirchhoff Constraints . . . . . . . . . . . . . . . . . . . . . . . . . . . 15410.2 Kinematics of an Axisymmetric Kirchhoff Plate . . . . . . . . . . . . 15510.3 Radial Virtual Displacement . . . . . . . . . . . . . . . . . . . . . . . 15910.4 Director Virtual Displacement . . . . . . . . . . . . . . . . . . . . . . 16210.5 Addressing Lagrange Multipliers . . . . . . . . . . . . . . . . . . . . . 16310.6 Problem As a Pair of ODEs . . . . . . . . . . . . . . . . . . . . . . . 16710.7 Applying the Constitutive Relation . . . . . . . . . . . . . . . . . . . 17110.8 The Flat Plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17410.9 Differential Equations for Bifurcation . . . . . . . . . . . . . . . . . . 17510.10Linear Boundary Conditions for Bifurcation . . . . . . . . . . . . . . 17910.11Evidence of Bifurcation . . . . . . . . . . . . . . . . . . . . . . . . . . 18010.12Numerical Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18210.13Numerical Example: Buckling Near a Bifurcation Point . . . . . . . . 18310.14Numerical Example: Buckling Due to Immersion-Preventing Growth . 18610.15Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187

CHAPTER 11 CONCLUSION . . . . . . . . . . . . . . . . . . . . . . . . . . 190

APPENDIX A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192

APPENDIX B . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196

APPENDIX C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198

APPENDIX D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200

REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202

10

LIST OF FIGURES

1.1 Polymerization-induced curvature in gel discs and sheets . . . . . . . 17

2.1 Cartesian vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222.2 Labeling a Euclidean point by a vector . . . . . . . . . . . . . . . . . 222.3 A differentiable manifold . . . . . . . . . . . . . . . . . . . . . . . . . 242.4 A cylindrical coordinates chart . . . . . . . . . . . . . . . . . . . . . . 252.5 Another cylindrical coordinates chart . . . . . . . . . . . . . . . . . . 252.6 Tangent vectors in E3, R3, and B . . . . . . . . . . . . . . . . . . . . 412.7 An isometric immersion that is not an embedding . . . . . . . . . . . 47

3.1 Change of coordinates on a body in E3 . . . . . . . . . . . . . . . . . 503.2 Deformation of a body in E3 . . . . . . . . . . . . . . . . . . . . . . . 51

4.1 Local azimuthal expansion . . . . . . . . . . . . . . . . . . . . . . . . 644.2 Stress-free “virtual configuration” . . . . . . . . . . . . . . . . . . . . 654.3 Stress-free “virtual configuration” . . . . . . . . . . . . . . . . . . . . 664.4 Isometric immersion that is not an embedding . . . . . . . . . . . . . 75

5.1 Stress tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

7.1 Phase portrait . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1157.2 Cavity radius as function of γ2 . . . . . . . . . . . . . . . . . . . . . . 117

8.1 A slice of a Cosserat shell . . . . . . . . . . . . . . . . . . . . . . . . 123

9.1 Isoclines in the (τ , N)-plane . . . . . . . . . . . . . . . . . . . . . . . 1409.2 Vector field of the dynamical system run in reverse . . . . . . . . . . 1429.3 Curve formed by unstable manifold orbits . . . . . . . . . . . . . . . 1439.4 Stereographic projection of curve of initial conditions . . . . . . . . . 1449.5 Radial stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1469.6 τ = r/R . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1469.7 Impact of κ on radial stress profiles . . . . . . . . . . . . . . . . . . . 1479.8 Impact of κ on τ = r/R . . . . . . . . . . . . . . . . . . . . . . . . . 1479.9 Impact of rate of change of γ2 on radial stress profiles . . . . . . . . . 1489.10 Impact of rate of change of γ2 on τ = r/R . . . . . . . . . . . . . . . 1489.11 Vector field of dynamical system run in reverse . . . . . . . . . . . . . 1499.12 Radial and azimuthal stresses . . . . . . . . . . . . . . . . . . . . . . 1509.13 τ = r/R . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151

LIST OF FIGURES – Continued

11

9.14 Impact of κ on stresses . . . . . . . . . . . . . . . . . . . . . . . . . . 1519.15 Impact of κ on τ = r/R . . . . . . . . . . . . . . . . . . . . . . . . . 1529.16 Doubling γ2 − 1 approximately doubles the stress . . . . . . . . . . . 1529.17 Doubling γ2 − 1 approximately doubles the expansion . . . . . . . . . 153

10.1 Kinematics of axisymmetric buckled Kirchhoff plate . . . . . . . . . . 15610.2 Stress perturbation at R = R1 . . . . . . . . . . . . . . . . . . . . . 18210.3 Moment perturbation at R = R1 . . . . . . . . . . . . . . . . . . . . 18310.4 Bifurcation value of γ1 . . . . . . . . . . . . . . . . . . . . . . . . . . 18310.5 First three critical values of γ1 . . . . . . . . . . . . . . . . . . . . . . 18410.6 Buckling angle φ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18510.7 Middle surface height ζ . . . . . . . . . . . . . . . . . . . . . . . . . . 18510.8 Integrated radial stress . . . . . . . . . . . . . . . . . . . . . . . . . . 18610.9 Integrated radial moment . . . . . . . . . . . . . . . . . . . . . . . . 18610.10Integrated stresses . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18710.11Buckling angle φ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18810.12Middle surface height ζ . . . . . . . . . . . . . . . . . . . . . . . . . . 18810.13Top view of middle surface . . . . . . . . . . . . . . . . . . . . . . . . 189

12

LIST OF TABLES

2.1 Equivalent contravariant vectors in E3 and R3 . . . . . . . . . . . . . 422.2 Equivalent covariant vectors in E3 and R3 . . . . . . . . . . . . . . . 42

3.1 Body manifold with cylindrical coordinates . . . . . . . . . . . . . . . 623.2 Secondary manifold with cylindrical coordinates . . . . . . . . . . . . 62

7.1 Fixed points in the in (z1, z2)-plane . . . . . . . . . . . . . . . . . . . 111

13

ABSTRACT

Until the twentieth century, theories of elastic rods and shells arose from collections

of geometric and mechanical assumptions and approximations. These theories often

lacked internal consistency and were appropriate for highly proscribed and some-

times unknown geometries and deformation sizes. The pioneering work of Trues-

dell, Antman, and others converted mechanical intuition into rigorous mathematical

statements about the physics and mechanics of rods and shells. The result is the

modern, geometrically exact theory of finite deformations of rods and shells.

In the latter half of the twentieth century, biomechanics became a major focus

of both experimental and theoretical mechanics. The genesis of residual stress by

non-elastic growth has significant impact on the shape and mechanical properties

of soft tissues. Inspired by the geometry of blood vessels and adopting a formalism

found in elasto-plasticity, mechanicians have produced rigorous and applied results

on the effect of growth on finite elastic deformations of columns and hollow tubes.

Less attention has been paid to shells.

A theory of growing elastic plates has been constructed in the context of linear

elasticity. It harnessed many results in the theory of Riemann surfaces and has

produced solutions that are surprisingly similar to experimental observations. Our

intention is to provide a finite-deformation alternative by combining growth with the

geometrically exact theory of shells. Such a theory has a clearer and more rigorous

foundation, and it is applicable to thicker structures than is the case in the current

theory of growing plates.

This work presents the basic mathematical tools required to construct this alter-

native theory of finite elasticity of a shell in the presence of growth. We make clear

that classical elasticity can be viewed in terms of three-dimensional Riemannian ge-

ometry, and that finite elasticity in the presence of growth must be considered in this

14

way. We present several examples that demonstrate the viability and tractability of

this approach.

15

CHAPTER 1

INTRODUCTION

The theory of elastic shells, which includes plates and membranes, can be traced

back to Euler (1766), who derived a model of what would today be called an elastic

membrane. While French and German researchers dominated the construction of

the theory through much of the nineteenth century, around the turn of the century

Russians scientists made great contributions while applying elastic models to prob-

lems in naval architecture. The aerospace industry and modern numerical computing

have provided another major impetus for theoretical refinements and computational

solutions; see Todhunter and Pearson (1960), Ventsel and Krauthammer (2001), and

Szilard (2004). More recently, the theory of elastic solids has provided the founda-

tion for much of modern biomechanics. This should be recognized as a return of

mechanics to biology, as some nineteenth-century botanists used mechanical theory

to explain experimental observations; see Strasburger (1895).

Mechanics of soft tissues has become a focus of much experimental and theoret-

ical research. Two of the most-studied phenomena in soft tissues are the generation

of residual stress — stress that remains in a body after body forces and applied

tractions have been removed — and its role in morphogenesis and changes in me-

chanical properties, especially in cardiovascular tissues. See, for example, Rachev

et al. (1998), Lin and Taber (1995), Taber (1998a), Taber (1998b), Taber and Pe-

rucchio (2000), Humphrey (2006), and Chapter 9 of Humphrey (2002).

Soft tissues undergo significant deformations, so much research has been set in

the context of finite elasticity. Non-elastic growth has been incorporated into this

theory by considering deformations consisting of two steps. In the first, the body

undergoes non-elastic volumetric growth. This growth is described locally by a

growth tensor G that determines how infinitesimal material lines are distorted by

the growth. If the local values of the growth tensor do not form the gradient of a

16

deformation of the body, the growth is called incompatible. In the presence of incom-

patible growth, the body undergoes an instantaneous elastic response. This elastic

response is also described locally by a tensor A, which determines how “grown”

infinitesimal material lines are distorted. The product A · G is the gradient of a

true deformation of the body. The grown body is assumed to be stress-free, if not

necessarily a continuous body in Euclidean space. The elastic response distorts the

grown body so that it “fits” into Euclidean space as a solid continuum once again,

at the cost of carrying residual stress.

While much work has been devoted to importing the decomposition into soft

tissue mechanics (Klisch et al. (2003), Lubarda and Hoger (2002)), creating a the-

ory for initially residually stressed bodies (Hoger (1997), Klarbring et al. (2007)),

deriving evolution laws for the growth tensor (Rodriguez et al. (1994), DiCarlo and

Quiligotti (2002)), and clarifying the foundations of the decomposition independent

of its origin in elasto-plasticity theory (Chen and Hoger (2000)), most applications of

the theory have avoided analysis of buckling due to growth. Exceptions include the

analysis of spherical shells in Ben Amar and Goriely (2005), Goriely and Ben Amar

(2005), Goriely et al. (2006), and of cylindrical columns in Goriely et al. (2008),

Vandiver and Goriely (2008).

A theory of grown plates has been constructed and has produced solutions

that share many qualitative features of experimental results (Efrati et al. (2007a),

Efrati et al. (2007c), Efrati and Kupferman (2009), Efrati et al. (2009)). In

the experiments, thin discs of gel are formed by mixing water, monomers of N-

isopropylacrylamide (NIPA), bisacrylamide (a “cross-linker”), and polymerization

agents. The concentration of NIPA varies with radius in the disc. Due to the non-

uniformity of NIPA distribution, above a critical temperature the disc is described

as possessing a non-Euclidean metric. This is meant to indicate that intrinsic dis-

tances in the middle surface of the (three-dimensional) disc are now assigned by a

(two-dimensional) metric that precludes the middle surface from residing in two-

dimensional Euclidean space. See Figure 1.1. See Efrati et al. (2007b) for a brief

video of the buckling.

17

Figure 1.1: Discs and cylindrical sheets of gel with polymerization-induced curva-ture. Different shapes were produced by varying the thickness and distribution ofN-isopropylacrylamide, which polymerizes. From Efrati et al. (2007a). Used withthe kind permission of Eran Sharon, Racah Institute of Physics, The Hebrew Uni-versity of Jerusalem.

The mechanics in Efrati et al. (2007a) and related work is derived from linear

elasticity. In particular, the work adapts the Foppl-von Karman equations for large

deformations of plates. The energy density in this theory is a linear combination of

a bending energy quadratic in the curvatures of the middle surface and a stretching

energy quadratic in the difference between the curvatures of the middle surfaces of

the non-Euclidean plate and of a “nearby” Euclidean plate. Each of these com-

18

ponents is scaled by an appropriate power of the plate’s thickness. The choice of

Foppl-von Karman equations limits the variety of constitutive relations that can

be employed. Further, this work has considered only non-Euclidean plates that are

not even immersible (locally embeddable), neglecting the simpler case of immersible

but not globally embeddable plates. Nevertheless, this theory has made good use

of results on Riemann surfaces, and its results are qualitatively similar to a wide

variety of observed phenomena.

Recently, another example of polymerization-induced growth and elastic re-

sponse has been constructed in the laboratory of Shuji Hashimoto, of Waseda Univer-

sity’s Department of Applied Physics. The physical dimensions of the “gel caterpil-

lar”, in which the oscillatory Belousov-Zhabotinsky reaction causes polymer growth

and shrinkage, put it outside the realm of membranes, where Foppl-von Karman

is applicable, but not outside the geometrically exact theory of shells; see Simonite

(2009).

Our goal is to provide an alternative to the linearized theory of incompatibly

grown shells. We combine the growth/elastic multiplicative decomposition with the

geometrically exact theory of elastic shells. The result is not only free of Foppl-von

Karman’s limits of applicability, but also makes explicit the form of the Euclidean-

to-non-Euclidean transformation, which is obscured in a linearized approach.

In Chapter 2 we introduce the machinery of curvilinear coordinates in three-

dimensional Euclidean space. We introduce the relationship between coordinate sys-

tems and tangent vectors, covariant differentiation of vector fields, and the Riemann-

Christoffel curvature tensor. We then see that Euclidean space is an example of a

Riemannian manifold. Incompatibly grown solid bodies will be portrayed as Rie-

mannian manifolds.

Chapter 3 presents the kinematics of finite deformations of solid bodies, with a

special emphasis on the distinction between coordinate points and Euclidean points.

We emphasize that a deformation of a body in three-dimensional Euclidean space

should be viewed as the composition of a map from Euclidean points to coordinate

points, followed by a map from coordinate points to new coordinates, and completed

19

by a map from the new coordinates back into Euclidean space. This explicit division

of stages is obscured in standard introductions such as Gurtin (1981), Ogden (1984),

Antman (1995), and Liu (2002). We introduce the form and role of the deforma-

tion gradient. Once this classical material is covered, we present some elementary

notions of differential geometry and show their relation to the kinematics of finite

deformations. In particular, we see that a deformation of a solid body changes the

metric tensor on the Riemannian manifold that represents the body, so that intrinsic

arclengths of curves are different after a deformation.

In Chapter 4 we use differential geometry to show that incompatible growth,

when described via a growth tensor, causes a change of metric tensor in the same

fashion that a true deformation does. In the case of incompatible growth, the new

metric precludes the body from residing in three-dimensional Euclidean space. We

consider simple examples of incompatible growth in cylindrical coordinates.

In Chapter 5 we introduce the fundamental balance laws of solid mechanics. We

adopt the material of Lagrangian formulation of continuum mechanics and show

how differential equations are derived from integral balance laws.

Chapter 6 is an introduction to constitutive relations. In particular, we explore

hyperelastic materials and how growth is incorporated via the multiplicative de-

composition A · G. The detail devoted to the geometric roles of the deformation

gradient and the stress tensors make clear in this section why the nominal stress,

rather than the otherwise more fundamental first Piola-Kirchhoff stress, is of central

importance in the computation of stress power.

In Chapter 7 we introduce growth to a simple (and not geometrically exact)

axisymmetric model of a membrane and demonstrate that, unlike in the growthless

case, it supports solutions with cavities.

Before tackling geometrically exact models of plates, we open Chapter 8 with a

brief look at the original Kirchhoff theory of plates, the Foppl-von Karman equa-

tions, and rational mechanical objections to the latter. As an alternative, we intro-

duce the Principle of Virtual Power for plates and review the theory by which we

derive the differential equations of equilibrium for plates.

20

In Chapter 9 we examine the genesis of residual stress by incompatible growth

in a flat Kirchhoff plate. The type of growth considered precludes the grown body

from being put into three-dimensional Euclidean space, even locally, and yet we find

in our examples that the elastic response is capable of bringing the plate back into

Euclidean space, without the assistance of applied tractions.

Chapter 10 is devoted to incompatible growth-induced axisymmetric buckling of

the Kirchhoff plate. We find that incompatible growth of a type simpler than that

found in Efrati et al. (2007a), etc., leads to the existence of buckled configurations

after the elastic response to the growth, without applied tractions. We examine the

onset of buckling’s dependence on physical parameters such as thickness.

It is hoped that this work will serve not just as a first step in the geometrically

exact theory of incompatibly grown shells, but also as an introduction to finite

elasticity as an application of three-dimensional Riemannian geometry. Multiple

pages are devoted to fundamental notation and machinery because they are not

included in standard introductions, while more advanced treatises, such as Noll

et al. (1968) and Bloom (1979), lack basic examples and leapfrog directly from

Euclidean geometry to non-Riemannian geometry.

Comments on notation

Many of the quantities in this work are tensor-valued. The tensor products we

consider will be outer products of two vectors, although tensor products of other

objects can be defined. For example, the products a⊗ b can be viewed as

a⊗ b = abT =

a1

a2

a3

( b1 b2 b3 )

Given this structure, one can take a dot-product from the left or the right:

(a⊗ b) · c = (b · c) a,

d · (a⊗ b) = (d · a)b,

where b · c and d · a are scalar constants or scalar-valued functions.

21

Since constant tensors and tensor-valued functions can be viewed as linear com-

binations of such tensor products, we will also use the dot-product notation for the

action of tensors on vectors, such as F · v instead of Fv, and for inner products of

tensors, such as A ·B instead of AB.

We use the Einstein summation convention unless otherwise noted. If an “upper”

index and a “lower” index in an expression are the same, then the expression is to

be understood as the sum of that same expression over that index. This is true for

each matched pair of indices. For example,

Eij Ei ⊗Ej = E11 E1 ⊗E1 + E12 E1 ⊗E2 + E13 E1 ⊗E3

+ E21 E2 ⊗E1 + E22 E2 ⊗E2 + E23 E2 ⊗E3

+ E31 E3 ⊗E1 + E32 E3 ⊗E2 + E33 E3 ⊗E3,

as there is one upper and one lower i and one upper and one lower j. With this

convention, the Kronecker delta will have one upper index and one lower index.

If the same index label is used more than once for an upper or a lower index, a

mistake has been made, and the author will be grateful for notification.

22

CHAPTER 2

EUCLIDEAN GEOMETRY WITH CURVILINEAR COORDINATES

2.1 Euclidean Point Space

The setting for continuum mechanics is three-dimensional Euclidean point space,

denoted by E3. E3 consists of points, which may be denoted with respect to a fixed

orthogonal system of Cartesian coordinates (X,Y, Z). Each point in E3 is equipped

with a three-dimensional real inner-product space whose elements are called tangent

vectors. This vector space is called the tangent space at that point. The Cartesian

tangent vectors i, j, and k provide an orthonormal basis for the tangent space; see

Figure 2.1. All tangent vectors can be expressed as linear combinations of i, j, and

k. Tensors, to be discussed later, can be expressed in terms of the tensor products

i` ⊗ im, i` ⊗ im ⊗ in, and so on, where i1 = i, i2 = j, and i3 = k.

Z j

i

k

XY

Figure 2.1: i, j, and k are unit vectorsanchored at a point in E3, that point alongthe X-, Y -, and Z-axes, respectively.

H1,1,1L

i

j

k

X=i+ j+k

E3

Figure 2.2: We will refer to a point in E3

by the tangent vector that points from theorigin to the point.

This endows E3 with a metric: if we have two points with Cartesian coordinates

23

(x0, y0, z0) and (x1, y1, z1), then the distance vector between them is

v = (x1 − x0)i + (y1 − y0)j + (z1 − z0)k, (2.1)

where the Cartesian vectors i, j, and k are anchored at (x0, y0, z0) and point along

the x-, y-, and z-axes, respectively. This vector can be viewed as an arrow anchored

at (x0, y0, z0) and pointing to (x1, y1, z1). The distance between the two points is

the Euclidean norm of this vector:

dist ((x0, y0, z0), (x1, y1, z1)) =√

v · v=

√(x1 − x0)2 + (y1 − y0)2 + (z1 − z0)2. (2.2)

We will most often refer to a point in E3 not by its Cartesian coordinates, but

by the difference vector between the point and the origin; see Figure 2.2.

2.2 Curvilinear Coordinates on a Body

We will describe a solid body by a differentiable manifold B. B consists of material

points that can be assigned coordinates in an invertible fashion, at least locally. We

call the map Ψ : B → R3 a coordinate system on B. Each material point p ∈ B is

given a triple of real coordinates

Ψ(p) =(ξ1(p), ξ2(p), ξ3(p)

). (2.3)

Some manifolds may require different coordinate systems for different subsets. See

Figures 2.4 and 2.5 for an example in cylindrical coordinates.

The coordinate points Ψ(p) are mapped to spatial points in E3 by a position

function

X : Ψ(B) → E3 :(ξ1, ξ2, ξ3

) 7→ X(ξ1, ξ2, ξ3

). (2.4)

The map from material points to coordinate points to spatial points is illustrated

in Figure 2.3.

At each spatial point (X Ψ)(p) the tangent vector-valued partial derivatives

Ei =∂X

∂ξi(2.5)

24

B

Y YHBL

R3

A BR

Π

2 Π

Q

X`

HX`ëYLHBL

E3

Figure 2.3: The material points p ∈ B are mapped to coordinate points Ψ(p) ∈ R3 bythe coordinate system Ψ. The coordinate points Ψ(p) are mapped to spatial points

in E3 by the position function X. See Figures 2.4 and 2.5 for a caveat regardingcoordinate systems.

provide a basis for the tangent space at that point. The basis formed by the vectors

Ei is called the coordinate basis (Marsden and Hughes (1983), page 29) for the

tangent space at that point. There is a corresponding dual basis of vectors Ej

defined so that

Ej ·Ei = δji . (2.6)

The existence of the dual basis is guaranteed, even in the case of a non-orthogonal

coordinate basis E1,E2, E3, by the fact that if a, b, and c are linearly indepen-

dent, then the triple product det (abc) = a · (b× c) is non-zero. We set

Ei =Ej ×Ek

Ei · (Ej ×Ek), (2.7)

where (ijk) is a cyclic permutation of (123), so that

Ei ·Ej = Ei ·Ek = 0 and Ei ·Ei =Ei · (Ej ×Ek)

Ei · (Ej ×Ek)= 1. (2.8)

2.2.1 Example: Cylindrical Coordinates

First we demonstrate that we may need to choose different charts, or local patches

of coordinates, to put an open subset of a manifold into R3. In Figure 2.4, we use

a chart that assigns Θ-coordinates between 0 and 2π. Bo, the open interior of B, is

mapped into the open subset of R3 shown in the middle of Figure 2.4.

25

BoY YHBoL R3

0.25 0.75 R

Π

2 Π

3 ΠQ

X`

HX`ëYLHBoL

E3

Figure 2.4: One chart for cylindrical coordinates assigns Θ-coordinates between 0and 2π. The position function X for this chart sends this portion of R3 into E3.

Bo

YYHBoL

R3

0.25 0.75 R

Π

2 Π

3 ΠQ

X`

HX`ëYLHBoL

E3

Figure 2.5: Some subsets of Bo (the open interior of B) may straddle the line thatcorresponds to Θ = 0 = 2π in the chart in Figure 2.4. To avoid a disconnectedimage of the subset in R3, we choose a different chart, which assigns Θ-coordinatesbetween π and 3π. The position function X for this chart sends this portion of R3

into E3.

The case shown in Figure 2.5, though, requires a different chart. The subset

pictured would straddle the line marking Θ = 0 = 2π, and using the chart shown

in Figure 2.4 would leave the image of the subset disconnected in R3. We use a

different chart, which amounts to choosing a different branch of arctangent. In this

chart, the Θ-coordinates assigned to points in B are between π and 3π.

In the case of cylindrical coordinates (R, Θ, Z) the standard position function is

X(R, Θ, Z) = R cos Θ i + R sin Θ j + Z k, (2.9)

which can be used for both charts shown in Figures 2.4 and 2.5. The corresponding

26

coordinate basis consists of

ER =∂X

∂R= cos Θ i + sin Θ j, (2.10)

EΘ =∂X

∂Θ= −R sin Θ i + R cos Θ j, (2.11)

EZ =∂X

∂Z= k, (2.12)

and the dual basis vectors are

ER = cos Θ i + sin Θ j, (2.13)

EΘ = − 1

Rsin Θ i +

1

Rcos Θ j, (2.14)

EZ = k, (2.15)

so that Ei ·Ej = δji , the Kronecker delta.

Note that the cylindrical coordinate basis vectors do not all have the same lengths

or even the same units. To compute the arclength of a curve parametrized by

cylindrical coordinates, these tangent vectors must be weighted differently. Consider

a differentiable curve γ : s 7→ (R(s), Θ(s), Z(s)) in Ψ(B), defined for some interval

of s-values. Via the position function X, this corresponds to a curve in E3:

(X γ)(s) = R(s) cos Θ(s) i + R(s) sin Θ(s) j + Z(s) k. (2.16)

The velocity of a point tracing out this Euclidean curve is now easily computed:

(X γ)′(s) = (R′(s) cos Θ(s)−R(s)Θ′(s) sin Θ(s)) i

+ (R′(s) sin Θ(s) + R(s)Θ′(s) cos Θ(s)) j

+ Z ′(s) k, (2.17)

where ()′ indicates differentiation with respect to the parameter s. The size of this

velocity is the Euclidean norm of the right-hand side:

√(X γ

)′·(X γ

)′=

√(R′)2 + (RΘ′)2 + (Z ′)2. (2.18)

27

We see now that the parameter s does not correspond to Euclidean arclength, and

that the derivatives R′, Θ′, and Z ′ must be properly weighted to find the Euclidean

velocity.

Note the equivalent expression for the velocity of the Euclidean curve:

(X γ

)′= ERR′ + EΘΘ′ + EZZ ′ = Ei

dξi

ds, (2.19)

where (ξ1, ξ2, ξ3) = (R, Θ, Z), and where we have employed Einstein’s summation

notation. Thus the absolute velocity is

√(X γ

)′·(X γ

)′=

√(Ei

dξi

ds

)·(

Ejdξj

ds

)

=

√Ei ·Ej

dξi

ds

dξj

ds. (2.20)

When computing Euclidean speed from curvilinear velocities, the dot products Eij =

Ei ·Ej provide the proper weights for the curvilinear velocities. The 3 × 3 matrix

with entries Eij is symmetric and positive-definite, so it provides a quadratic form

for converting a triple of velocities into the corresponding total Euclidean velocity

squared: (dξ1

ds,dξ2

ds,dξ3

ds

)7→ Eij

dξi

ds

dξj

ds. (2.21)

We will explore more properties of the matrix [Eij] in the following sections.

2.3 Contravariant and Covariant Vectors

When a position function X exists, there need not be any distinction between the

spaces to which the basis vectors Ei and the dual basis vectors Ej belong. We will

eventually see, however, the distinct roles they play in more abstract geometries.1

For the moment we treat them as equivalent.

Given a point in (X Ψ)(B), we have two bases for the tangent space at that

point: E1,E2, E3 andE1,E2, E3

. Hence, each tangent vector v can be ex-

1For those familiar with Dirac’s bra-ket notation in quantum mechanics it is worth noting that

Ei and Ei are related in exactly the same fashion that |φ〉 and 〈φ| are related.

28

pressed with respect to either basis:

v = viEi = viEi. (2.22)

We will show that under a change of variables, the coefficients vi and vi change in

distinct ways.

Suppose Φ = (η1, η2, η3) is another coordinate system on B and that the position

function for this coordinate system is X. Then (η1, η2, η3) can be viewed as functions

of (ξ1, ξ2, ξ3) such that

X(η1

(ξ1, ξ2, ξ3

), η2

(ξ1, ξ2, ξ3

), η3

(ξ1, ξ2, ξ3

))= X

(ξ1, ξ2, ξ3

). (2.23)

This immediately provides the relationship between the coordinate basis generated

by (ξ1, ξ2, ξ3) and X and that generated by (η1, η2, η3) and X:

Ei =∂

∂ξiX

(ξ1, ξ2, ξ3

)=

∂ξiX

(η1, η2, η3

)

=∂ηj

∂ξi

∂X

∂ηj=

∂ηj

∂ξiEj. (2.24)

Inherent in this formula is the assumption that each coordinate ηi is treated as a

differentiable function of (ξ1, ξ2, ξ3). If, on the other hand, each ξi is a differentiable

function of (η1, η2, η3), then we have

Ei =∂

∂ηiX

(η1, η2, η3

)=

∂ηiX

(ξ1, ξ2, ξ3

)

=∂ξj

∂ηi

∂X

∂ξj=

∂ξj

∂ηiEj. (2.25)

We define the dual basis vectors Ejso that E

j · Ei = δji . Hence, the matrix that

has Eias columns satisfies

(E

1E

2E

3)

=(E1 E2 E3

)−T

, (2.26)

where −T indicates the transpose of the inverse. Since

(E1 E2 E3

)= (E1 E2 E3)

∂ξ1

∂η1∂ξ1

∂η2∂ξ1

∂η3

∂ξ2

∂η1∂ξ2

∂η2∂ξ2

∂η3

∂ξ3

∂η1∂ξ3

∂η2∂ξ3

∂η3

, (2.27)

29

(E

1E

2E

3)

=(E1 E2 E3

)−T

=

(E1 E2 E3)

∂ξ1

∂η1∂ξ1

∂η2∂ξ1

∂η3

∂ξ2

∂η1∂ξ2

∂η2∂ξ2

∂η3

∂ξ3

∂η1∂ξ3

∂η2∂ξ3

∂η3

−T

= (E1 E2 E3)−T

∂ξ1

∂η1∂ξ1

∂η2∂ξ1

∂η3

∂ξ2

∂η1∂ξ2

∂η2∂ξ2

∂η3

∂ξ3

∂η1∂ξ3

∂η2∂ξ3

∂η3

−T

=(E1 E2 E3

)

∂η1

∂ξ1∂η2

∂ξ1∂η3

∂ξ1

∂η1

∂ξ2∂η2

∂ξ2∂η2

∂ξ2

∂η1

∂ξ3∂η3

∂ξ3∂η3

∂ξ3

, (2.28)

where in the second line we consider the ξi to be functions of ηj and in the last line

we consider the ηj to be functions of ξi. The conclusion is that

Ei=

∂ηi

∂ξjEj and Ei =

∂ξi

∂ηjE

j. (2.29)

Now consider a vector v expressed with respect to the various bases arising from

the various coordinate systems:

v = viEi = vjEj = vj ∂ηi

∂ξjEi, (2.30)

v = viEi= vjE

j = vj∂ξj

∂ηiE

i. (2.31)

We conclude that there are two distinct linear transformation rules for the coeffi-

cients:

vi = vj ∂ηi

∂ξjand vi = vj

∂ξj

∂ηi. (2.32)

The coefficients vi are called the contravariant components of the vector v, while

the coefficients vi are called the covariant components. A vector of the form viEi is

often called a contravariant vector, while viEi is called a covariant vector.

A multi-indexed object that is contravariant or covariant in each index is called

a tensor. This ensures that, regardless of coordinate system chosen, when a tensor

30

value is re-written in terms of the Cartesian tangent vectors i, j, and k, the result

is the same.

An object of this type with m covariant and n contravariant indices will be called

a tensor of type(

mn

).2 For example, the metric tensor of a coordinate system, which

we will explore in the next section, has the form

Eij Ei ⊗Ej, (2.33)

where Eij = Ei ·Ej. We will show that this is a tensor of type(02

)by considering

a change of variables. The coefficients transform as

Eij = Ei ·Ej =

(∂ηk

∂ξiEk

)·(

∂η`

∂ξjE`

)= Ek`

∂ηk

∂ξi

∂η`

∂ξj, (2.34)

where Ek` = Ek · E`. There is one summation for each index. The full object

transforms as

Eij Ei ⊗Ej =

(Ek`

∂ηk

∂ξi

∂η`

∂ξj

) (∂ξi

∂ηmE

m)⊗

(∂ξj

∂ηnE

n)

=

(∂ξi

∂ηm

∂ηk

∂ξi

∂η`

∂ξj

∂ξj

∂ηn

)Ek` E

m ⊗ En

= δkmδ`

n Ek` Em ⊗ E

n

= Emn Em ⊗ E

n. (2.35)

This ensures that the metric tensor has the same formula, in terms of basis vectors,

regardless of the coordinate system chosen.

To show how tensors with upper indices transform, consider the inverse of the

metric tensor:

Eij Ei ⊗Ej. (2.36)

This is a tensor of type(20

), and its coefficients transform as

Eij = Ei ·Ej =

(∂ξi

∂ηkE

k)·(

∂ξj

∂η`E

`)

= Ek` ∂ξi

∂ηk

∂ξj

∂η`.

2This convention is adopted from page 65 of Marsden and Hughes (1983), which is the opposite

of the convention defined on page 12 of Lee (1997).

31

2.4 Metric Tensor

Eij ei ⊗ Ej is a tensor of type(02

)and is called the metric tensor induced by the

position function X. We have seen an example in which the metric tensor for a

coordinate system provides a method to convert a triple of curvilinear velocities

into the square of the corresponding Euclidean speed. In fact, the metric tensor

provides a method for assigning a norm to any contravariant vector. If v = viEi is

such a vector, then

(viEi

) · (Ek` Ek ⊗E`) · (vjEj

)= Ek`v

ivj(Ei ·Ek

) (E` ·Ej

)

= Eklvivjδk

i δ`j

= Eijvivj. (2.37)

The metric tensor assigns size to vectors when it is used as a quadratic form. It

can also be used as a bilinear form, providing an inner product of pairs of vectors.

Let u = uiEi and v = viEi. Their inner product is

(uiEi

) · (Ek` Ek ⊗E`) · (vjEj

)= Ek`u

ivj(Ei ·Ek

) (E` ·Ej

)

= Ekluivjδk

i δ`j

= Eijuivj. (2.38)

We continue to treat the contravariant (viEi) and covariant (viEi) expressions

for a vector v as alternative descriptions of the same object. The metric tensor

provides a method for converting between such expressions. First note that

viEi = vjEj

(Ek`E

k ⊗E`) · (viEi

)=

(Ek`E

k ⊗E`) · (vjE

j)

Ek`viδ`

i Ek = Ek`vjE`jEk

Ek`v`Ek = Ek`E

`jvjEk. (2.39)

It is shown in Appendix A that the matrix [Eij] is the inverse of the matrix [Eij]

formed by the dot-products Eij. Hence the last line above can be re-written as

Ek`v`Ek = δj

kvjEk = vkE

k, (2.40)

32

from which we conclude

vk = Ek`v`. (2.41)

Similar manipulations show the inverse transformation,

vj = Ejivi. (2.42)

Even after we abandon the equality of viEi and viEi, there will remain a one-

to-one correspondence between contravariant and covariant vectors, and Eqs. (2.41)

and (2.42) will remain the formulas relating the coefficients.

2.5 Christoffel Symbols of the Second Kind

If v is a vector field on (X Ψ)(B), then its rate of change with position is expressed

in terms of the rates of change of its coefficients vi and also of the basis vectors Ei:

∂v

∂ξj=

∂vi

∂ξjEi + vi ∂Ei

∂ξj.

∂Ei/∂ξj is vector-valued, so it can be expressed as a linear combination of the basis

vectors. One way to define the Christoffel symbols of the second kind (for a given

coordinate system), denoted by Γkij, is by their role in the expression

∂Ei

∂ξj= Γk

jiEk. (2.43)

They can also be defined as follows. Since Ej ·Ei = δji ,

∂ξj

(Ek ·Ei

)= 0

∂Ek

∂ξj·Ei + Ek · ∂Ei

∂ξj= 0

Ek · ∂Ei

∂ξj

︸ ︷︷ ︸Γk

ji

= −Ei · ∂Ek

∂ξj. (2.44)

33

The derivative of a vector field can now be written as

∂v

∂ξj=

∂vi

∂ξjEi + viΓk

jiEk

=∂vi

∂ξjEi + vkΓi

jkEi

=

(∂vi

∂ξj+ Γi

jkvk

)Ei. (2.45)

The coefficients in this expression are denoted by the shorthand ∇jvi or vi

||j:

∇jvi = vi

||j =∂vi

∂ξj+ Γi

jkvk. (2.46)

The tensor formed by ∇jvi is called the covariant derivative of the vector field v.

Note that ∇jvi has one upper and one lower index.

We can also derive the form of the covariant derivative of the corresponding

covariant expression:

∂ξj

(viE

i)

=∂vi

∂ξjEi + vi

∂Ei

∂ξj

=∂vi

∂ξjEi + vk

∂Ek

∂ξj

=∂vi

∂ξjEi + vk

(Ei ⊗Ei

) · ∂Ek

∂ξj

=∂vi

∂ξjEi + vk

(Ei · ∂Ek

∂ξj

)Ei

=∂vi

∂ξjEi + vk

(−Γkji

)Ei

=

(∂vi

∂ξj− Γk

jivk

)Ei, (2.47)

where we have used the fact that Ei ⊗ Ei is the identity tensor among the con-

travariant vectors. The shorthand for the coefficients of this derivative is

∇jvi = vi||j =∂vi

∂ξj− Γk

jivk. (2.48)

34

The Christoffel symbols of the second kind are symmetric in their lower indices:

Γkji = Ek · ∂Ei

∂ξj= Ek · ∂2X

∂ξj∂ξi

= Ek · ∂2X

∂ξi∂ξj= Ek · ∂Ej

∂ξi

= Γkij. (2.49)

The Christoffel symbols of the first kind do not form a tensor, as their transfor-

mation under a change of variables, proved in Appendix A, lacks the proper form:

Γkji =

∂ξk

∂η`

∂ηm

∂ξj

∂ηn

∂ξiΓ`

mn +∂ξk

∂η`

∂2η`

∂ξj∂ξi. (2.50)

However, the difference Γkji − Γk

ij does transform properly:

Γkji − Γk

ij =∂ξk

∂η`

∂ηm

∂ξj

∂ηn

∂ξi

(Γ`

mn − Γ`nm

), (2.51)

so the difference Γkji − Γk

ij is a tensor of type(12

). It is called the torsion tensor (of

the connection provided by the Christoffel symbols Γkji; this will be discussed later).

Although this is the zero tensor here, it need not be so in more abstract geometries.

2.6 Christoffel Symbols of the First Kind

The Christoffel symbols of the first kind are produced by “lowering” the lone “upper”

index in Γkij:

Γji` = E`kΓkji = E`kE

k · ∂Ei

∂ξj(2.52)

These symbols are symmetric in their first two indices. The symbols alone do not

form a tensor, but the difference Γji` − Γij` is a tensor of type(03

):

Γjik − Γijk =∂η`

∂ξk

∂ηm

∂ξj

∂ηn

∂ξi

(Γmn` − Γnm`

). (2.53)

We will again exploit the equality of contravariant and covariant representations

of a vector:

Γji` = E`kΓkji = E`kE

k · ∂Ei

∂ξj= E` · ∂Ei

∂ξj. (2.54)

35

This equality and the symmetry in the first two indices allows us to express Γjik

completely in terms of the metric tensor and its derivatives:

Γjik =1

2

(Ek · ∂Ei

∂ξj+ Ek · ∂Ej

∂ξi

)

=1

2

(∂

∂ξj(Ek ·Ei)−Ei · ∂Ek

∂ξj+

∂ξi(Ek ·Ej)−Ej · ∂Ek

∂ξi

)

=1

2

(∂Eki

∂ξj+

∂Ekj

∂ξi−Ei · ∂2X

∂ξj∂ξk−Ej · ∂2X

∂ξi∂ξk

)

=1

2

(∂Eki

∂ξj+

∂Ekj

∂ξi−Ei · ∂Ej

∂ξk−Ej · ∂Ei

∂ξk

)

=1

2

(∂Eki

∂ξj+

∂Ekj

∂ξi− ∂Eij

∂ξk

). (2.55)

Hence the Christoffel symbols of the second kind can be expressed completely in

terms of metric tensor coefficients, their derivatives, and inverse metric tensor coef-

ficients via

Γkji = Ek`Γji` =

1

2Ek`

(∂E`i

∂ξj+

∂E`j

∂ξi− ∂Eij

∂ξ`

). (2.56)

2.7 Ricci’s Theorem

Proposition 2.7.1 (Ricci’s Theorem) The covariant derivative of the metric tensor

is zero.

Proof

∂ξk

(EijE

i ⊗Ej)

=∂Eij

∂ξkEi ⊗Ej + Eij

∂Ei

∂ξk⊗Ej + EijE

i ⊗ ∂Ej

∂ξk

=∂Eij

∂ξkEi ⊗Ej − EijΓ

ik`E

` ⊗Ej − EijΓjk`E

i ⊗E`

=

(∂Eij

∂ξk− EmjΓ

mki − EimΓm

kj

)Ei ⊗Ej

=

(∂Eij

∂ξk− Γkij − Γkji

)Ei ⊗Ej. (2.57)

However,

Γkij + Γkji =1

2

(∂Eji

∂ξk+

∂Ejk

∂ξi− ∂Eki

∂ξj

)+

1

2

(∂Eik

∂ξj+

∂Eij

∂ξk− ∂Ejk

∂ξi

)

=∂Eij

∂ξk. (2.58)

36

This can be expressed as

∇kEij = Eij||k =∂Eij

∂ξk− EmjΓ

mki − EimΓm

kj = 0, (2.59)

which illustrates the fact that the covariant derivative of a tensor with two lower

indices includes one sum with Christoffel symbols for each index, and that these

sums are subtracted.

Each coefficient in the covariant derivative of the metric tensor is zero, so

∂ξk

(EijE

i ⊗Ej)

= O, (2.60)

the zero tensor.

Ricci’s Theorem ensures that the rate of change of the “size” of a vector field

is due entirely to the covariant derivatives of the components. For example, the

squared norm of a vector v is Eijvivj, and the derivative of this quantity is

∂ξk

(Eijv

ivj)

=∂

∂ξk

((v`E`

) · (EijEi ⊗Ej

) · (vmEm))

=∂v`E`

∂ξk· (EijE

i ⊗Ej) · (vmEm)

+(v`E`

) ·(

∂ξk

(EijE

i ⊗Ej)) · (vmEm)

+(v`E`

) · (EijEi ⊗Ej

) · ∂vmEm

∂ξk

=(v`||kE`

) · (EijEi ⊗Ej

) · (vmEm)

+(v`E`

) · (EijEi ⊗Ej

) · (vm||kEm

)

= Eij

(vi||kv

j + vivj||k

). (2.61)

2.8 Further Properties of the Covariant Derivative

2.8.1 Covariant Derivative with Respect to a Vector Field

We have already seen the definition and proper method for computing the total

derivative of a vector field with respect to coordinates ξi. This definition provides a

37

method for computing a derivative of a vector field with respect to another vector

field.

Suppose that γ : s 7→ γ(s) = (ξ1(s), ξ2(s), ξ3(s)) is a curve in Ψ(B). Then X γ

is its image in E3. Let v = viEi be a vector field on (XΨ)(B). Though v is a vector

field on the Euclidean body, it can still be expressed as a function of coordinates.

Its rate of change along X γ is

d

ds(v γ) (s) =

d

ds

(vi γ

)(s) (Ei γ) (s)

=(vi γ

)′(s) (Ei γ) (s) +

(vi γ

)(s) (Ei γ)′ (s), (2.62)

where

(vi γ

)′(s) = lim

ε→0

vi (γ(s + ε))− vi (γ(s))

ε

=∂vi

∂ξj

∣∣∣∣γ(s)

dγj

ds

∣∣∣∣γ(s)

, (2.63)

and

(Ei γ)′ (s) = limε→0

Ei (γ(s + ε))−Ei (γ(s))

ε

=∂Ei

∂ξj

∣∣∣∣γ(s)

dγj

ds

∣∣∣∣γ(s)

= ΓkjiEk

∣∣γ(s)

dγj

ds

∣∣∣∣γ(s)

. (2.64)

The full derivative is then

d

ds(v γ) (s) =

(∂vi

∂ξj+ Γi

jkvk

)dγj

dsEi = vi

||jdγj

dsEi. (2.65)

To compute the rate of change of v along this curve, the components vi||j of the

covariant derivative are weighted by components dγj/ds of the tangent vector to the

curve.

A continuously differentiable non-zero vector field u on (X Ψ)(B) and an initial

point in (XΨ)(B) can be used to define a curve γ that passes through that point; u

serves as the tangent vector to this curve ( dds

(X γ) = u). Based on the observation

38

above, we can define the covariant derivative of a vector field v with respect to the

vector field u:

∇uv =

(∂vi

∂ξj+ Γi

jkvk

)ujEi = vi

||jujEi. (2.66)

2.8.2 Covariant Derivative as a Tensor Field

If v = viEi is a vector field, then the coefficients vi||j form a tensor of type

(11

). The

full object is

vi||jEi ⊗Ej, (2.67)

which yields a contravariant vector only after contraction in the lower index:

(vi||jEi ⊗Ej

) ·Ek =(vi||jEi

)δjk = vi

||kEi. (2.68)

In other words, the covariant derivative of a contravariant vector field is a tensor

field of type(11

)that can be “fed” a contravariant vector to yield a contravariant

vector field. If the vector that it is fed is u = ukEk, then the result is

(vi||jEi ⊗Ej

) · (ukEk

)= vi

||jukδj

kEi = vi||ju

jEi = ∇uv, (2.69)

the covariant derivative of v with respect to u.

The proof that the coefficients vi||j transform properly can be found in Appendix

A.

2.9 Curvature

Consider the impact of repeated covariant differentiation on a vector field:

∇w∇uv = ∇w

((∂vi

∂ξj+ Γi

jkvk

)ujEi

)

=

(∂

∂ξ`

[(∂vi

∂ξj+ Γi

jkvk

)uj

]+ Γi

m`

(∂vm

∂ξj+ Γm

jkvk

)uj

)w`Ei

=

∂2vi

∂ξ`∂ξjuj +

∂vi

∂ξj

∂uj

∂ξ`+

∂Γijk

∂ξ`vkuj + Γi

jk

∂vk

∂ξ`uj + Γi

jkvk ∂uj

∂ξ`

+ Γim`

(∂vm

∂ξj+ Γm

jkvk

)uj

w`Ei. (2.70)

39

Computing repeated covariant differentiation in the opposite order and finding the

difference yields

∇w∇uv −∇u∇wv =

∂vi

∂ξj

(∂uj

∂ξ`w` − ∂wj

∂ξ`u`

)+ Γi

jkvk

(∂uj

∂ξ`w` − ∂wj

∂ξ`u`

)

+

(∂Γi

jk

∂ξ`− ∂Γi

`k

∂ξj+ Γi

m`Γmjk − Γi

mjΓm`k

)vkujw`

Ei. (2.71)

Next we define the Lie bracket of two vector fields:

[w,u] =[w`E`, u

jEj

]=

(wk ∂uj

∂ξk− uk ∂wj

∂ξk

)Ej, (2.72)

which is a contravariant vector field. The covariant derivative with respect to the

Lie bracket is then

∇[w,u]v = ∇(wk ∂uj

∂ξk−uk ∂wj

∂ξk

)Ej

v

= ∇wk ∂uj

∂ξk Ejv −∇

uk ∂wj

∂ξk Ejv

=

(∂vi

∂ξj+ Γi

j`v`

)wk ∂uj

∂ξkEi −

(∂vi

∂ξj+ Γi

j`v`

)uk ∂wj

∂ξkEi

=

∂vi

∂ξj

(∂uj

∂ξkwk − ∂wj

∂ξkuk

)+ Γi

j`v`

(∂uj

∂ξkwk − ∂wj

∂ξkuk

)Ei.(2.73)

The difference between Eq. (2.71) and Eq. (2.73) is

∇w∇uv −∇u∇wv −∇[w,u]v

=

(∂Γi

jk

∂ξ`− ∂Γi

`k

∂ξj+ Γi

m`Γmjk − Γi

mjΓm`k

)vkujw`Ei (2.74)

Note that the inputs are three contravariant vector fields (u, w, and v) and the

output is a contravariant vector field. There are three lower indices and one upper

index. We set

R(w,u)v = ∇w∇uv −∇u∇wv −∇[w,u]v, (2.75)

Rikj` =

∂Γijk

∂ξ`− ∂Γi

`k

∂ξj+ Γi

m`Γmjk − Γi

mjΓm`k. (2.76)

This defines a tensor field of type(13

)called the Riemann-Christoffel curvature ten-

sor. Before we explore the information carried in R, we consider an especially simple

case.

40

2.9.1 Example: Cartesian Coordinates

In Cartesian coordinates (X, Y, Z) with position function X i + Y j + Z k, each

Christoffel symbol is identically zero. Hence each coefficient of the Riemann-

Christoffel curvature tensor is also identically zero:

Rikj` ≡ 0, so R ≡ O. (2.77)

Since R is a tensor field, however, its value is invariant under a change of coordinates.

Hence R is identically zero in all coordinate systems on E3.

2.10 Abstract Manifolds

We have derived numerous objects directly from the position function X. But this is

just one example of a three-dimensional differentiable manifold. Many possibilities

arise when there is no position function.

2.10.1 Coordinates and Tangent Spaces

A three-dimensional differentiable manifold B is a set such that for each point p ∈ Bthere is an open neighborhood U(p) of p and a diffeomorphism Ψ between U(p)

and an open subset of R3. The image Ψ(U(p)) ⊂ R3 consists of coordinate points,

and the diffeomorphism Ψ is called a coordinate system for that neighborhood. If

p ∈ B is a point in a manifold and lies in the domain of two coordinate systems

Ψ = (ξ1, ξ2, ξ3) and Φ = (η1, η2, η3), then the change-of-coordinates maps Ψ Φ−1

and Φ Ψ−1 must be C∞.3

Each point in a differentiable manifold has a vector space of tangent vectors

attached to it. If p ∈ B, its tangent space is denoted TpB. For example, in the

presence of a position function, for each tangent vector at a point in E3 there is

a corresponding tangent vector in the space of coordinates, as well as an abstract

tangent vector at p. See Figure 2.6.

3In some applications, including solid mechanics, the requirement of infinite differentiability

relaxed is relaxed for these functions and for others, such as metric tensors. See the definitions

41

E3

ER

R3

¶R

A BR

Π

2 Π

Q

v

Figure 2.6: When a position function exists, each tangent vector at (XΨ)(p) in theEuclidean body corresponds to a tangent vector at Ψ(p) and to an abstract tangentvector at p ∈ B.

When there is no position function X available, vectors may not be expressed

in terms of the partial derivatives ∂X/∂ξi. Instead, once a coordinate system is

assigned to a neighborhood, the vectors are expressed as linear combinations of the

partial derivative operators ∂/∂ξi. These vectors are contravariant, since under a

change of variables they transform as

∂ηi=

∂ξj

∂ηi

∂ξj. (2.78)

The linear functionals on the tangent space TpB also form a vector space, denoted

T ∗pB. If ∂/∂ξi, i = 1, 2, 3 form a basis for TpB, then a basis for T ∗

pB is provided by

the linear functionals dξj, j = 1, 2, 3, defined by

dξj

[∂

∂ξi

]= δj

i , (2.79)

where we use square brackets to indicate that the tangent vector ∂/∂ξi is the argu-

ment of dξj. The functionals dξj are called differential 1-forms, and T ∗pB is called a

cotangent space. The dξi are the covariant vectors on an abstract manifold, as

dηi =∂ηi

∂ξjdξj. (2.80)

In the case of cylindrical coordinates with position function X, the correspon-

dence between vectors is summarized in Tables 2.1 and 2.2.

in Ciarlet (2005).

42

Contravariant vector at X(R, Θ, Z) Tangent vector at (R, Θ, Z)

ER = ∂X∂R

∂∂R

EΘ = ∂X∂Θ

∂∂Θ

EZ = ∂X∂Z

∂∂Z

Table 2.1: Equivalent tangent vectors at X(R, Θ, Z) and at (R, Θ, Z)

Covariant vector at X(R, Θ, Z) 1-form at (R, Θ, Z)

ER dREΘ dΘEZ dZ

Table 2.2: Equivalent covariant vectors at X(R, Θ, Z) and 1-forms at (R, Θ, Z)

2.10.2 Metric Tensor

When a space of coordinates Ψ(B) has a position function X, there is a metric

tensor induced on Ψ(B). If M = Eij Ei⊗Ej is the metric tensor on the Euclidean

body, the corresponding metric tensor on Ψ(B) is

M = Eij dξi ⊗ dξj, (2.81)

where the Eij are now simply scalar functions of the coordinates (ξ1, ξ2, ξ3). Since

M is induced by X, the arclength of a differentiable curve γ in Ψ(B) computed

with M agrees with the Euclidean arclength of its image X γ. We will find that

deformations and growth change the metric tensor on Ψ(B) in such a way that the

arclengths computed with the new metric tensor do not necessarily agree with the

Euclidean arclengths of curves in (X Ψ)(B).

Without a position function X, a differentiable manifold does not automatically

come equipped with a metric tensor. A priori, each tangent space is a real vector

space only, not an inner-product space. Assigning a metric tensor field to a manifold

provides an inner product. A metric tensor has the form

g = gij dξi ⊗ dξj, gij = g

(∂

∂ξi,

∂ξj

)=

⟨∂

∂ξi,

∂ξj

⟩, (2.82)

where 〈 , 〉 indicates the inner product defined by g.

43

Definition 2.10.1 A(02

)tensor field is a Riemannian metric if

(i) it is C∞ (this degree of differentiability is relaxed in solid mechanics);

(ii) it is symmetric, i.e.

g(v,u) = g(u,v); (2.83)

(iii) and it is positive-definite, i.e.

g(v,v) ≥ 0 for all vectors v, and g(v,v) = 0 iff v = 0. (2.84)

2.10.3 Linear Connection

In the absence of a position function, there is no obvious coordinate-invariant

method for computing the derivative of a vector field. More concretely, there is

no method for computing the acceleration of a curve t 7→ γ(t) = (ξ1(t), ξ2(t), ξ3(t))

in the space of coordinates. When there is a position function X, one can compute

both the velocity and the acceleration of the curve’s Euclidean image X γ:

position: (X γ)(t) = X(ξ1(t), ξ2, ξ3(t)

),

velocity:d

dt(X γ)(t) =

∂X

∂ξi

∣∣∣∣∣γ(t)

(ξi

)′(t) = Ei|γ(t)

(ξi

)′(t),

acceleration:d2

dt2(X γ)(t) =

∂Ei

∂ξj

∣∣∣∣γ(t)

(ξj

)′(t)

(ξi

)′(t)

+ Ei|γ(t)

(ξi

)′′(t), (2.85)

where ′ indicates differentiation with respect to the parameter t. The tangent vec-

tors Ei are functions of coordinates (ξ1, ξ2, ξ3), so the derivatives ∂Ei/∂ξj can be

computed.

In the absence of a position function X, however, there is no Euclidean image

of the curve γ, and the tangent vector to the curve is

(ξi

)′(t)

∂ξi

∣∣∣∣γ(t)

, (2.86)

44

whose derivative with respect to t is not defined without more information. We

must assign a method for computing derivatives of the vectors ∂/∂ξi.

A linear connection provides a definition for differentiation of a vector field. In

terms of differentiation of coordinate basis vectors with respect to other coordinate

basis vectors, a linear connection is a set of 33 = 27 smooth functions Γkji, known as

connection coefficients, with which we define the covariant derivative

∇ ∂

∂ξj

∂ξi= Γk

ji

∂ξk. (2.87)

Definition 2.10.2 A linear connection ∇ is a bilinear map (u,v) 7→ ∇uv on vector

fields that satisfies

(i) linearity over smooth real-valued functions in the first argument:

∇fu+hwv = f∇uv + h∇wv for f, h ∈ C∞(B); (2.88)

(ii) linearity over real numbers in the second argument:

∇u (av + bw) = a∇uv + b∇uw for a, b ∈ R; (2.89)

(iii) the following product rule in the second argument:

∇u (fv) = f∇uv +

(ui ∂f

∂ξi

)v for f ∈ C∞(B). (2.90)

The Christoffel symbols of the second kind, as derived in Eq. (2.56), are connec-

tion coefficients for a special case of a linear connection. A linear connection need

not be derived from the metric tensor. On an abstract manifold, the metric tensor

and connection may be assigned independently of each other. In fact, for a manifold

that can be completely covered by one coordinate chart, any collection of 27 smooth

functions Γkji can define a linear connection; see Lemma 4.4 in Lee (1997).

2.10.4 Riemannian Geometry

When connection coefficients are derived directly from a Riemannian metric via

Eq. (2.56), the resulting connection is called the Riemannian connection or the

45

Levi-Civita connection of the manifold. This connection is the unique connection

with two properties found in the connection derived from the position function X.

Theorem 2.10.3 “Fundamental Lemma of Riemannian Geometry” (Theorem 5.4

in Lee (1997)) The Riemannian connection on a manifold B with Riemannian metric

g is the sole linear connection that satisfies

(i) symmetry: the connection coefficients for a coordinate basis4 satisfy Γkij = Γk

ji,

or ∇uv −∇vu = [u,v] for all vector fields u, v; and

(ii) metricity: the covariant derivative (defined by the connection) of the metric

tensor is zero: ∇g = O.

See Lemma 5.2 and Theorem 5.4 in Lee (1997), or Theorem 4.24 in Marsden and

Hughes (1983).

While a three-dimensional Riemannian manifold may or may not “fit” into E3

(see below), it does have analogues of length, area, and volume, so lineal, areal, and

volumetric quantities can be defined on it.

2.11 Isometric Immersion and Embedding of a Riemannian Manifold

The position function X maps Ψ(B) into E3 in a special way. It is a diffeomorphism

from the manifold into E3, and for each curve γ in Ψ(B), the arclength of γ and the

arclength of its Euclidean image X γ agree. These properties earn X the status

of an isometric embedding of Ψ(B) into E3.

We have seen that starting with an isometric embedding X, computing from

it a Riemannian metric, and deriving the Riemannian connection from the metric,

produce a Riemann-Christoffel curvature tensor that is identically zero:

existence of isometric embedding into E3 =⇒ R ≡ O. (2.91)

4The connection coefficients of a symmetric connection are symmetric in the lower indices when

the coefficients are computed with respect to the coordinate basis ∂/∂ξi, i = 1, 2, 3. This need not

be true of coefficients for other bases. See the warning on the top of page 246 of Frankel (1997).

46

The three-dimensional version of a result dubbed the “Fundamental Theorem of

Riemannian Geometry” in Ciarlet (2005) gives a partial converse.

Definition 2.11.1 Let Ω be an open subset of R3. A mapping Y : Ω → E3 is

an immersion at (ξ1, ξ2, ξ3) ∈ Ω if it is differentiable at (ξ1, ξ2, ξ3) and the partial

derivatives ∂Y /∂ξi, i = 1, 2, 3, at (ξ1, ξ2, ξ3) are linearly independent.

Theorem 2.11.2 (Theorem 1.6-1 in Ciarlet (2005)) Let Ω be a simply connected set

in R3 and let C = [Eij] ∈ C2(Ω,S3>) be a symmetric positive-definite matrix-valued

function on Ω that satisfies

R ≡ O in Ω.

Then there exists an immersion Y ∈ C3(Ω,E3) such that

Eij =∂Y

∂ξi· ∂Y

∂ξjin Ω.

The isometric immersion promised by the theorem is a local position function.

It is a map from simply connected open subsets of the manifold into E3 that is

locally diffeomorphic and such that metric tensor-computed arclengths of curves in

coordinates and arclengths of their Euclidean images agree. An isometric immersion

may fail to be an embedding if it is not globally invertible.

2.11.1 Example: Immersion without Embedding

Consider the manifold B with coordinate system Ψ = (R, Θ, Z) satisfying

A ≤ R ≤ B, 0 ≤ Θ ≤ 3π

2, H1 ≤ Z ≤ H2, (2.92)

See Figure 2.7. Now consider the map

Y (R, Θ, Z) = R cos(2Θ) i + R sin(2Θ) j + Z k. (2.93)

47

The coefficients of the corresponding metric tensor are

ERR ERΘ ERZ

EΘR EΘΘ EΘZ

EZR EZΘ EZZ

=

1 0 0

0 4R2 0

0 0 1

. (2.94)

Computing the corresponding Christoffel symbols and Riemann-Christoffel curva-

ture tensor via Eq. (2.56) yields an identically zero curvature tensor.

While Y might seem to be a good candidate for a position function, Figure 2.7

shows that the Euclidean images of two disjoint simply connected open subsets of

Ψ(B) will have a nonempty intersection.

W1

R3

A B R

Π

2

Π

3 Π2

2 ΠQ

Y`HW1L

E3

-2 -1 1 2

-2

-1

1

2

W2

R3

A B R

Π

2

Π

3 Π2

2 ΠQ

Y`HW2L

E3

-2 -1 1 2

-2

-1

1

2

Figure 2.7: An isometric immersion that is not an embedding: the images Y (Ω1)

and Y (Ω2) overlap.

We have seen that Y is an isometric immersion of this manifold with this metric

into E3, but Y is not an embedding. Suppose we discard the would-be position

48

function Y and equip Ψ(B) with the metric tensor

dR⊗ dR + 4R2 dΘ⊗ dΘ + dZ ⊗ dZ, (2.95)

whose associated (through the Riemannian connection) Riemann-Christoffel curva-

ture tensor is identically zero. Is there an isometric embedding of the (open, simply

connected) interior of Ψ(B) with this metric into E3? A theorem on the “rigidity”

of isometric immersions shows that the answer is negative.

Theorem 2.11.3 Let Ω be an open connected subset of R3 and let Y ∈ C1(Ω,E3)

and Y ∈ C1(Ω,E3) be two isometric immersions such that their associated coordi-

nate bases satisfy

∂Y

∂ξi· ∂Y

∂ξj=

∂Y

∂ξi· ∂Y

∂ξjin Ω.

Then there are a vector c ∈ E3 and a proper rotation Q such that

Y (ξ1, ξ2, ξ3) = Q · Y (ξ1, ξ2, ξ3) + c

for each (ξ1, ξ2, ξ3) ∈ Ω.

See Theorem 1.7-1 in Ciarlet (2005).

A uniform rotation and translation of Y will not alleviate its global failure of

injectivity on the interior of Ψ(B). Since there is no isometric embedding of the

partial annulus, any complete annulus equipped with the same metric also lacks an

isometric embedding.

49

CHAPTER 3

KINEMATICS OF DEFORMATION

3.1 Deformation versus Change of Coordinates

A change of coordinates consists of two sets of coordinates and a position function

for each set. As shown in Figure 3.1, each occupied spatial point is mapped to a

coordinate point (η1, η2, η3) by the inverse position function X−1

; in Figure 3.1,

these are the Cartesian coordinates (X, Y, Z). Each coordinate point is mapped to

a coordinate point (ξ1, ξ2, ξ3) in a new coordinate system ((R, Θ, Z) in Figure 3.1)

in a continuously differentiable and invertible manner, at least locally. Each new

coordinate point is then mapped back into E3 by a new position function X. The

final results is that

X(η1, η2, η3

)= X = X

(ξ1

(η1, η2, η3

), ξ2

(η1, η2, η3

), ξ3

(η1, η2, η3

)). (3.1)

The same material points occupy the same spatial points. Only the coordinates

labeling the points have changed.

The transformation shown in Figure 3.2 is different. There is a diffeomorphism

f between the triples (R, Θ, Z) ∈ Ψ(B) and the triples (r, θ, z) ∈ (f Ψ)(B), but the

position function x sends (r, θ, z) to a point in E3 different from the point where X

sends the corresponding (R, Θ, Z). If this transformation satisfies certain conditions

to be discussed below, then it is called a deformation of the body.

Given coordinate system Ψ on B and position function X on Ψ(B), a deformation

of the body is a continuously differentiable and invertible function χ : (X Ψ)(B) →E3 : X 7→ x that has the form

x(ζ1, ζ1, ζ3) = (x f)(ξ1, ξ2, ξ3)

= (x f X−1

)(X)

= χ(X). (3.2)

50

FHBL

R3

0 1

0

12

1

X

Y

YHBL

R3

0 2

0

Π

4

Π

2

R

Q

X -1

X`

E3

HXëFLHBL = HX

`ëYLHBL

Figure 3.1: In a change from Cartesian to cylindrical coordinates, X−1

maps pointsin (X Φ)(B) ⊂ E3 to the coordinate points (X,Y, Z) ∈ Φ(B). Each Cartesiancoordinate point (X,Y, Z) is mapped to a cylindrical coordinate point (R, Θ, Z) ∈Ψ(B), which is then mapped, via X, to the point X(R, Θ, Z) = X(Z, Y, Z).

See Figure 3.2. We will refer to (X Ψ)(B) as the reference configuration and to

(χ X)(B) as the deformed configuration.

Note that computing a deformation requires three steps. First, the point

X ∈ (XΨ)(B) is mapped to the corresponding coordinate point (ξ1, ξ2, ξ3) ∈ Ψ(B)

by the inverse position function X−1

. Second, the coordinate point (ξ1, ξ2, ξ3) is

mapped to a coordinate point (ζ1, ζ2, ζ3) = f(ξ1, ξ2, ξ3). Finally, x maps the coor-

dinate point (ζ1, ζ2, ζ3) to a spatial point x(ζ1, ζ2, ζ3).

Just as X induces a metric tensor on Ψ(B), x induces a metric tensor on the

space of (ζ1, ζ2, ζ3)-coordinates. In the case of cylindrical coordinates (r, θ, z), we

51

R3

0

0

Π

2 Π

R

Q

f

R3

0

0

Π

2 Π

r

Θ

X` -1 x`

E3

E3

Figure 3.2: In a deformation, a point in the body may be assigned new coordinatesand be sent to a new point in E3. Further, the distances between material pointsmay change.

will use the position function

x(r, θ, z) = r cos(θ) i + r sin(θ) j + z k. (3.3)

The corresponding Euclidean metric tensor is

m = eij ei ⊗ ej

= er ⊗ er + r2 eθ ⊗ eθ + ez ⊗ ez, (3.4)

52

where eij = ei · ej and

er =∂x

∂r= cos(θ) i + sin(θ) j,

eθ =∂x

∂θ= r(− sin(θ) i + cos(θ) j),

ez =∂x

∂z= k. (3.5)

The metric tensor m corresponds to the metric tensor

m = dr ⊗ dr + r2 dθ ⊗ dθ + dz ⊗ dz (3.6)

on the space of (r, θ, z)-coordinates.

In a change of variables, the intrinsic arclengths of curves remain the same.

Deformation, on the other hand, changes the intrinsic arclengths. Local changes of

length and of direction are described via the deformation gradient.

3.2 Deformation Gradient

A local description of a deformation is provided by considering its impact on in-

finitesimal material lines. Consider the points X(ξ1, ξ2, ξ3) and X(ξ1 + c1ε, ξ2 +

c2ε, ξ3 + c3ε) in E3 and consider the difference quotient of χ evaluated at these two

points:

limε→0(χ X)(ξ1 + εc1, ξ2 + εc2, ξ3 + εc3)− (χ X)(ξ1, ξ2, ξ3)

ε

=∂x

∂ζj

∣∣∣∣X(ξ1,ξ2,ξ3)

ci ∂ζj

∂ξi

∣∣∣∣(ξ1,ξ2,ξ3)

, (3.7)

where (ζ1, ζ2, ζ3) = f(ξ1, ξ2, ξ3). The following is an equivalent but more illustrative

expression of this directional derivative:

∂x

∂ζjci ∂ζj

∂ξi=

∂ζj

∂ξi

∂x

∂ζj

(ckδi

k

)

=∂ζj

∂ξi

∂x

∂ζj

(ckEi ·Ek

)

=

(∂ζj

∂ξi

∂x

∂ζj⊗Ei

)

︸ ︷︷ ︸deformation gradient

· (ckEk

). (3.8)

53

The tensor in the last line maps a contravariant vector at X(ξ1, ξ2, ξ3) to a con-

travariant vector at x(ζ1, ζ2, ζ3). Because it maps in this way, this tensor, called

the deformation gradient, is called a two-point tensor.

The deformation gradient is usually expressed as follows:

F =∂ζj

∂ξiej ⊗Ei, (3.9)

where ej = ∂x/∂ζj is a tangent vector at the point x(ζ1, ζ2, ζ3). F is a tensor

field whose domain is the reference configuration (X Ψ)(B). Ei is evaluated at

(ξ1, ξ2, ξ3) and anchored at X(ξ1, ξ2, ξ3), while ej is evaluated at (η1, η2, η3) and

anchored at x(ζ1, ζ2, ζ3).

3.2.1 Example: Cartesian Coordinates

We consider a reference configuration with Cartesian coordinates (X, Y, Z) and po-

sition function

X(X,Y, Z) = X i + Y j + Z k. (3.10)

Note that

EX =∂X

∂X= i, EY =

∂X

∂Y= j, EZ =

∂X

∂Z= k. (3.11)

The deformed configuration is assigned Cartesian coordinates (x, y, z) and posi-

tion function

x(x, y, z) = x i + y j + z k, (3.12)

so that

ex =∂x

∂x= i, ey =

∂x

∂y= j, ez =

∂x

∂z= k. (3.13)

In the deformation, we consider x, y, and z to be functions of X, Y , and Z. The

54

deformation gradient is

F =∂x

∂Xex ⊗EX +

∂x

∂Yex ⊗EY +

∂x

∂Zex ⊗EZ

+∂y

∂Xey ⊗EX +

∂y

∂Yey ⊗EY +

∂y

∂Zey ⊗EZ

+∂z

∂Xez ⊗EX +

∂z

∂Yez ⊗EY +

∂z

∂Zez ⊗EZ

=∂x

∂Xi⊗ i +

∂x

∂Yi⊗ j +

∂x

∂Zi⊗ k

+∂y

∂Xj ⊗ i +

∂y

∂Yj ⊗ j +

∂y

∂Zj ⊗ k

+∂z

∂Xk ⊗ i +

∂z

∂Yk ⊗ j +

∂z

∂Zk ⊗ k. (3.14)

The vectors on the left-hand sides of the tensor products are anchored at the de-

formed point x(x, y, z), and the vectors on the right-hand sides are anchored at the

original point X(X,Y, Z) in the reference configuration.

Note that if we use the notation ∇x for the gradient of the function x, we can

write F as

F = i⊗ (∇x) + j ⊗ (∇y) + k ⊗ (∇z) . (3.15)

This gives an idea of the origin of the notation F = ∇⊗χ, which is used in Ogden

(1984). A clearer but still vague alternative notation would be F = (∇⊗ χ)T .

3.2.2 Example: Cylindrical Deformation

A cylindrical deformation of a body is expressed, in cylindrical coordinates (R, Θ, Z)

for the reference configuration and (r, θ, z) for the deformed configuration, by

(χ X)(R, Θ, Z) = x(r(R), θ(Θ), z(Z))

= r(R) (cos(θ(Θ)) i + sin(θ(Θ)) j) + z(Z) k, (3.16)

with θ(Θ) = Θ and z(Z) = Z. In this particular case, the deformation gradient has

only three terms:

F = r′er ⊗ER + eθ ⊗EΘ + ez ⊗EZ , (3.17)

where each vector on the right of a tensor product is anchored to the pre-deformation

point X(R, Θ, Z), and each vector on the left of a tensor product is anchored at

55

the post-deformation point x(r, θ, z) = (χ X)(R, Θ, Z).

Note the action of the deformation gradient on tangent vectors in the pre-

deformation body:

F ·ER = r′ er, F ·EΘ = eθ, F ·EZ = ez. (3.18)

3.3 Local Deformation of Euclidean Lines

The formula in Eq. (3.9) is equivalent to

F · ∂X

∂ξi=

∂ζj

∂ξi

∂x

∂ζj. (3.19)

If we allow ourselves the use of infinitesimals dξi and dζj, which we are not treating

as 1-forms, this can be written as

F · ∂X

∂ξidξi =

∂x

∂ζjdζj. (3.20)

Eq. (3.20) presents the deformation gradient as a linear map of “infinitesimal

material lines” in (X Ψ)(B) to their images in (χ X Ψ)(B). Infinitesimal

material lines are useful devices for describing the local strain of a deformation.

Consider an infinitesimal material line of the form

∂X

∂ξidξi = Ei dξi (3.21)

whose length is √(Ei dξi) · (Ej dξj) =

√Eijdξidξj. (3.22)

The image of this material line under the deformation is

F · (Ei dξi)

= F ·Ei dξi, (3.23)

and the length of this image is

√(F ·Ei dξi) · (F ·Ei dξi) =

√(F ·Ei)

T (F ·Ej) dξidξj

=√

Ei ·(F T · F ) ·Ejdξidξj. (3.24)

56

Note that the expression in Eq. (3.24) describes the length of the deformed infinites-

imal line in terms of the infinitesimals dξi on (X Ψ)(B).

While F is a linear map from the tangent space at X = X(ξ1, ξ2, ξ3) to the

tangent space at x = x(ζ1, ζ2, ζ3), F T ·F , known as the right Cauchy-Green tensor,

is a linear map from the tangent space at a point X = X(ξ1, ξ2, ξ3) in the reference

configuration to that same tangent space:

F : tangent space at X → tangent space at x, (3.25)

F T : tangent space at x → tangent space at X, (3.26)

F T · F : tangent space at X → tangent space at X. (3.27)

Such a tensor is called a Lagrangian tensor.

3.3.1 Example: Cylindrical Deformation

For the cylindrical deformation example, the length of an un-deformed infinitesimal

material line is √Eijdξidξj =

√(dR)2 + R2 (dΘ)2 + (dZ)2, (3.28)

while the length of its deformed image is

√Ei ·

(F T · F ) ·Ejdξidξj =

√(r′)2 (dR)2 + r2 (dΘ)2 + (dZ)2. (3.29)

The right Cauchy-Green tensor is

F T · F =(r′ ER ⊗ er + EΘ ⊗ eθ + EZ ⊗ ez

) · (r′ er ⊗ER + eθ ⊗EΘ + ez ⊗EZ)

= (r′)2err ER ⊗ER + eθθ EΘ ⊗EΘ + ezz EZ ⊗EZ

= (r′)2ER ⊗ER + r2 EΘ ⊗EΘ + EZ ⊗EZ . (3.30)

3.4 Local Deformation of Volumes

It is assumed that each subset of (X Ψ)(B) with positive volume is mapped to a

subset of (χ X Ψ)(B) with positive volume. This precludes deformations that

57

feature infinite mass density or inversion of the handedness of a neighborhood of

matter.

Consider an infinitesimal parallelepiped in the reference configuration, bounded

by infinitesimal material lines Eidξi, i = 1, 2, 3. The (infinitesimal) volume is

det(E1dξ1 E2dξ2 E3dξ3

)= det (E1E2E3) dξ1dξ2dξ3. (3.31)

The image of this parallelepiped will be bounded by infinitesimal material lines

F ·Eidξi, so the volume of the image will be

det(F ·E1dξ1 F ·E2dξ2 F ·E3dξ3

)

= det (F · (E1E2E3)) dξ1dξ2dξ3

= det (F ) det (E1E2E3) dξ1dξ2dξ3. (3.32)

Hence the deformation gradient must have positive determinant: det F > 0. This

will preclude not just the collapse of volumes to areas, but also reversal of orientation.

3.5 Local Deformation of Areas

An infinitesimal area is described by the cross-product of two noncollinear infinites-

imal lines. Consider an infinitesimal directed area of the form Ejdξj ×Ekdξk, and

let c be an arbitrary tangent vector in the same tangent space from which Ej and

Ek are drawn (so that it is acted upon by the same F ). Then the triple product

c · (Ej ×Ek) = det (c EjEk) , (3.33)

is transformed to

(F · c) · ((F ·Ej)× (F ·Ek)) = det ((F · c) (F ·Ej) (F ·Ek))

= det (F ) det (c EjEk)

= det (F ) c · (Ej ×Ek) . (3.34)

By the definition of tensor transpose F T ,

(F · c) · ((F ·Ej)× (F ·Ek)) = c · F T · ((F ·Ej)× (F ·Ek)) , (3.35)

58

so that

c · F T · ((F ·Ej)× (F ·Ek)) = det (F ) c · (Ej ×Ek) . (3.36)

Since c is arbitrary,

F T · ((F ·Ej)× (F ·Ek)) = det (F ) (Ej ×Ek) , (3.37)

so

(F ·Ej)× (F ·Ek) = det (F ) F−T · (Ej ×Ek) , (3.38)

where F−T is the inverse of the transpose F T . Hence the deformed area, expressed

in the coordinates ξi, is

det (F ) F−T · (Ej ×Ek) dξjdξk. (3.39)

The expression in Eq. (3.39) holds for each infinitesimal directed area, as Ej ×Ekdξjdξk, i, j = 1, 2, 3, i 6= j, form a basis for such areas. This transformation of

area, called Nanson’s formula, is sometimes summarized as

n ds = (det F ) F−T ·N dS, (3.40)

where N and n are unit vectors normal to the surface in the reference and deformed

configurations, respectively; and dS and ds are the infinitesimal scalar areas (or

area measures) in the reference and deformed configurations, respectively.

Eq. (3.40) is not technically correct as stated, however. The object on the left-

hand side of Eq. (3.40) is a vector-valued area measure on surfaces in the deformed

configuration, while the object on the right-hand side is an area measure on surfaces

in the reference configuration. The area measure on the right-hand side allows one to

compute surface integrals in the deformed configuration while still using coordinates

on the reference configuration:∮

∂(χX)(Ω)

n ds =

∂X(Ω)

(det F ) F−T ·N dS, (3.41)

for every measurable subset Ω of Ψ (B) with well-behaved boundary ∂Ω.

59

Since Eq. (3.41) demonstrates equivalence of measures, it can be extended to

equality of integrals. For continuously differentiable scalar-valued functions ϕ,

∂(χX)(Ω)

ϕ(χ−1(x)

)n(x) ds =

∂X(Ω)

ϕ(X) (det F (X)) F (X)−T ·N(X) dS,

(3.42)

where x = χ(X). Let A be a continuously differentiable tensor field function whose

value at x ∈ (χX Ψ)(B) maps from the tangent space at x to that same tangent

space. Such a tensor is called an Eulerian tensor, and

∂(χX)(Ω)

A (x) · n(x) ds =

∂X(Ω)

(det F (X)) A(χ(X)) · F (X)−T ·N(X) dS.

(3.43)

3.6 Differential and Pullback

Before we explore the difference between growth and deformation, we delve into the

machinery derived from coordinate-to-coordinate maps between manifolds equipped

with positions functions.

Just as the metric tensor M on Ψ(B) is the analogue of M on (X Ψ)(B),

there are also analogues of F and F T · F . The primary object of study will be the

invertible map f that performs the change-of-coordinate step in a deformation, as

pictured in Figure 3.2.

Consider our original body manifold B with coordinate system Ψ = (ξ1, ξ2, ξ3)

and position function X on Ψ(B), and another three-dimensional manifold M with

coordinate system ψ = (ζ1, ζ2, ζ3) and position function x on ψ(M).

Definition 3.6.1 For a diffeomorphism f : Ψ(B) → (f Ψ)(B) ⊂ ψ(M) defined by

f(ξ1, ξ2, ξ3) =(ζ1(ξ1, ξ2, ξ3), ζ2(ξ1, ξ2, ξ3), ζ3(ξ1, ξ2, ξ3)

), (3.44)

there is a corresponding linear map f∗ : Tp(B) → Tf(p)M between tangent spaces

that is expressed in coordinates by

f∗ =∂ζ i

∂ξj

∂ζ i⊗ dξi. (3.45)

60

f∗ is called the differential of f .1

The map f∗ sends tangent vectors ∂/∂ξj at (ξ1, ξ2, ξ3) ∈ Ψ(B) to tangent vectors

∂/∂ζ i at (ζ1, ζ2, ζ3) = f(ξ1, ξ2, ξ3) ∈ (f Ψ)(B) = ψ(M):

f∗

[vk ∂

∂ξk

]=

(∂ζ i

∂ξj

∂ζ i⊗ dξj

)[vk ∂

∂ξk

]

= dξj

[vk ∂

∂ξk

]∂ζ i

∂ξj

∂ζ i

= vk δjk

∂ζ i

∂ξj

∂ζ i

= vj ∂ζ i

∂ξj

∂ζ i, (3.46)

where vj and ∂ζ i/∂ξj are evaluated at (ξ1, ξ2, ξ3) = f−1(ζ1, ζ2, ζ3).

Just as the differential linearly maps tangent vectors at (ξ1, ξ2, ξ3) to tangent

vectors at f(ξ1, ξ2, ξ3), there is a linear map of cotangent vectors at (ζ1, ζ2, ζ3) to

cotangent vectors at f−1(ζ1, ζ2, ζ3).

Definition 3.6.2 The pull-back f ∗ of the diffeomorphism f : Ψ(B) → (f Ψ)(B) ⊂ψ(M) is the linear map satisfying, for each 1-form β ∈ T ∗

f(p)M and each vector

v ∈ TpB,

(f ∗β) [v] = β [f∗v] , (3.47)

where f∗v ∈ Tf(p)M.

1On page 38 of Marsden and Hughes (1983), the equivalent object is called a tangent map and

is written Tf .

61

In coordinates, Eq. (3.47) is written as

(f ∗βidζ i

) [vj ∂

∂ξj

]= βidζ i

[f∗vj ∂

∂ξj

]

= βidζ i

[vj ∂ζk

∂ξj

∂ζk

]

= βi vj ∂ηk

∂ξjdζ i

[∂

∂ζk

]

= βi vj ∂ηk

∂ξjδik

= βk vj ∂ηk

∂ξj. (3.48)

See Section 2.3 of Frankel (1997).

The pull-back f ∗ can be applied to covariant tensors, as well. In particular, the

metric tensor m on ψ(M) can be pulled back, via f ∗, to a metric tensor on Ψ(B).

If

m = eij dζ i ⊗ dζj, (3.49)

then the pull-back of m is found by applying f ∗ to each covariant vector in m:

f ∗m = eij

(f ∗dζ i

)⊗ (f ∗dζj

)

= eij

(∂ζ i

∂ξkdξk

)⊗

(∂ζj

∂ξ`dξ`

)

= eij∂ζ i

∂ξk

∂ζj

∂ξ`dξk ⊗ dξ`, (3.50)

where the eij are evaluated at (ζ1, ζ2, ζ3) and the partial derivatives are evaluated

at (ξ1, ξ2, ξ3) = f−1(ζ1, ζ2, ζ3).

f ∗m is a metric tensor on Ψ(B). It is used to compute the intrinsic distances

between material points in B following a deformation. If the position function x

on ψ(M) maps (f Ψ)(B) invertibly into E3, then (x f Ψ)(B) is the deformed

Euclidean body. In this case, if γ is a curve in Ψ(B), then its arclength computed

with f ∗m matches the Euclidean arclength of its Euclidean image, (x f γ).

62

3.6.1 Example in Cylindrical Coordinates

We consider two manifolds with cylindrical coordinates, standard Euclidean metric

tensor for cylindrical coordinates, and standard position function for cylindrical

coordinates. See Tables 3.1 and 3.2 for details.

Manifold BCoordinates Ψ = (R, Θ, Z)Metric tensor M = dR⊗ dR + R2 dΘ⊗ dΘ + dZ ⊗ dZ

Position function X(R, Θ, Z) = R cos(Θ) i + R sin(Θ) j + Z k

Table 3.1: Body manifold with cylindrical coordinates

Manifold MCoordinates ψ = (r, θ, z)Metric tensor m = dr ⊗ dr + r2 dθ ⊗ dθ + dz ⊗ dzPosition function x(r, θ, z) = r cos(θ) i + r sin(θ) j + z k

Table 3.2: Secondary manifold with cylindrical coordinates

If f has the form

f(R, Θ, Z) = (r(R), θ(R, Θ), Z), (3.51)

then f∗ has the form

f∗ = r′∂

∂r⊗ dR +

∂θ

∂R

∂θ⊗ dR +

∂θ

∂Θ

∂θ⊗ dΘ +

∂z⊗ dZ, (3.52)

where each ∂/∂ζ i is anchored at f(R, Θ, Z) ∈ ψ(M) and each dξj is anchored at

(R, Θ, Z) ∈ Ψ(B). The Euclidean analogue, under the position functions X and x,

is

F = r′ er ⊗ER +∂θ

∂Reθ ⊗EΘ +

∂θ

∂Θeθ ⊗EΘ + ez ⊗EZ . (3.53)

The pull-back of the metric tensor m is

f ∗m =

((r′)2

+ r2

(∂θ

∂R

)2)

dR⊗ dR + r2 ∂θ

∂R

∂θ

∂Θ(dR⊗ dΘ + dΘ⊗ dR)

+ r2

(∂θ

∂Θ

)2

dΘ⊗ dΘ + dZ ⊗ dZ. (3.54)

63

The Riemann-Christoffel curvature tensor associated with this metric is identically

zero. See the Mathematica code included in Appendix D to perform this computa-

tion.

The Euclidean analogue of f ∗m, using only the position function X, is

F T · F =

((r′)2

+ r2

(∂θ

∂R

)2)

ER ⊗ER + r2 ∂θ

∂R

∂θ

∂Θ

(ER ⊗EΘ + EΘ ⊗ eR

)

+ r2

(∂θ

∂Θ

)2

EΘ ⊗EΘ + EZ ⊗EZ . (3.55)

Just as the right Cauchy-Green tensor F T ·F gives a measure of the differential

arclength in the deformed Euclidean configuration, as a function of X ∈ (X Ψ)(B), the pull-back f ∗m gives a measure of the differential deformed arclength, as

a function of (R, Θ, Z) ∈ Ψ(B).

64

CHAPTER 4

INCOMPATIBLE GROWTH

While any two-point tensor with positive determinant at a fixed point X ∈ (XΨ)B)

can be the value of a deformation gradient, not every two-point tensor field with

positive determinant is equal globally to a deformation gradient. Further, some

two-point tensor fields are global deformation gradients on some bodies but not on

others.

For example, recall the (X Ψ)(B)-to-E3 map in Eq. (2.93):

(χ X)(R, Θ, Z) = x(r, θ, z)

= x(R, 2Θ, Z)

= R cos(2Θ) i + R sin(2Θ) j + Z k. (4.1)

Locally this performs an azimuthal expansion by a factor of 2, with no radial or

vertical deformation.

Figure 4.1: The map in Eq. (4.1) locally performs an azimuthal expansion. Thisneighborhood may also be translated and rotated by the deformation of the rest ofthe body.

If the body on the left-hand side of Figure 4.1 were the entire reference con-

figuration, then this would be a true deformation of the body. We have seen in

Figure 2.7, however, that this map is not globally injective on all Euclidean bodies.

In particular, an annulus whose pieces undergo such a transformation cannot remain

a solid annulus in E3.

65

If G is a continuously differentiable two-point tensor field on (X Ψ)(B) with

positive determinant but is not equal to the gradient of a deformation of (X Ψ)(B),

then G is called an incompatible growth tensor (field). In some treatments of

incompatible growth, the grown state is pictured as a disconnected collection of

stress-free neighborhoods (Rodriguez et al. (1994); Skalak et al. (1996); Taber and

Humphrey (2001); Klarbring et al. (2007)). When the body is considered this way,

incompatible growth is interpreted to describe how local neighborhoods of matter

would grow if they were not constrained by the presence of neighboring matter. See

Figure 4.2.

incompatible

growth

elasticresponse

HaL HbL HcL

Figure 4.2: The stress-free “virtual configuration” (b) is often drawn as a collectionof unconnected sets of matter.

However, we will show that incompatible growth amounts to a change of metric

on a Riemannian manifold, so we find it convenient to view such growth as chang-

ing the shape of the body manifold B in such a way that it does not “fit” into

three-dimensional Euclidean point space.1 The connectedness of the manifold is

unchanged by the growth. See Figure 4.3.

1We must specify that it is three-dimensional Euclidean point space that cannot accommo-

date the incompatibly grown body, as the first embedding theorem of Nash ensures that an n-

dimensional Riemannian manifold has a C1 isometric embedding into E2n+1. See Chapter 1 of

Han (2006) for a brief history of the subject.

66

HaL HbL HcL

Figure 4.3: We view incompatible growth as changing the metric of a Riemannianmanifold in such a way that the body cannot be isometrically embedded in E3. Allthe connectedness properties of the manifold remain unchanged by the growth.

4.1 Multiplicative Decomposition of Deformation Gradient

Regardless of the view taken of the intermediate, “grown” state, a solid body resides

in E3. It is assumed that there is an elastic response to the incompatible growth and

that this response acts on the disconnected neighborhoods or on the Riemannian

manifold in such a way that the body again occupies a closed subset of E3. The

elastic response is described locally by a two-point tensor A, which is analogous to

a deformation gradient. See Figures 4.2 and 4.3.

The result of the incompatible growth and the elastic response is a true defor-

mation of the body. The gradient of this deformation has the form

F = A ·G. (4.2)

If G represents incompatible growth, then G is not a deformation gradient. In

this case, A is also not a deformation gradient, and the elastic response is not a

deformation.

Multiplicative decompositions of the form in Eq. (4.2) arose after the introduc-

tion of Riemannian geometry into continuum mechanics in Eckart (1948). The form

F = F elastic · F plastic (4.3)

is sometimes labeled the Kroner-Lee decomposition, after Kroner (1959) and Lee

and Liu (1967), though the decomposition appears only in the latter. Kroner (1959)

67

is an exploration of linear connections other than the Levi-Civita connection, for

use in the theory of dislocations in crystalline solids. Lee and Liu introduced the

multiplicative decomposition after observing that

“[t]he combination of elastic and plastic strains, both finite, calls for a

more careful study of the kinematics than the usual assumption that the

total strain components are simply the sum of the elastic and plastic

components, as for infinitesimal strain theory.”

Rodriguez et al. (1994) is frequently cited as the introduction into bioelasticity

of the multiplicative decomposition in Eq. (4.2). It has since been used in an array

of models of soft tissue, but according to Chen and Hoger (2000),

“neither Rodriguez et al. (1994), nor Hoger (1999) provided a theoreti-

cally sound foundation for the two central ideas in their theory of growth:

the decomposition of the gradient of the total deformation; and the re-

lated assertion that the response of the material should depend only on

the elastic portion of the decomposed gradient.”

Chen and Hoger (2000) provided a foundation while considering a reference

configuration-free approach to elasticity with incompatible growth.

4.2 Change of Riemannian Metric

As in the introduction to deformation of a Euclidean body, the reference configu-

ration is described by a body manifold B with a coordinate system Ψ = (ξ1, ξ2, ξ3)

and an associated position function X, which induces a metric tensor M on Ψ(B)

as follows:

Ei =∂X

∂ξi, Eij = Ei ·Ej, M = Eij dξi ⊗ dξj. (4.4)

We will need a second three-dimensional manifold M with coordinate system

ψ = (ζ1, ζ2, ζ3) and associated position function x, which induces a metric tensor

68

m in an analogous fashion:

ei =∂x

∂ζ i, eij = ei · ej, m = eij dζ i ⊗ dζj. (4.5)

Each continuously differentiable2 two-point tensor field G with positive determi-

nant on the reference configuration (X Ψ)(B), can be used to induce a change in

the metric on Ψ(B), which amounts to a change in metric on B itself. If a two-point

tensor field of the form

G = Gij ei ⊗Ej, (4.6)

is completely specified, then the E3-point where ei is anchored, is specified. This

means that there is a local diffeomorphism f : Ψ(B) → ψ(M) such that the left-hand

side of each tensor product in Eq. (4.6) is the vector value

ei(ζ1, ζ2, ζ3) = ei(f(ξ1, ξ2, ξ3)) =

∂x

∂ζ i

∣∣∣∣(ζ1,ζ2,ζ3)=f(ξ1,ξ2,ξ3)

, (4.7)

and it is anchored at (x f)(ξ1, ξ2, ξ3).

This makes G a two-point tensor that maps vectors at X(ξ1, ξ2, ξ3) to vec-

tors at (x f)(ξ1, ξ2, ξ3). The corresponding object that maps tangent vectors at

(ξ1, ξ2, ξ3) ∈ Ψ(B) to tangent vectors at (ζ1, ζ2, ζ3) = f(ξ1, ξ2, ξ3) ∈ ψ(M) is

G = Gij

∂ζ i⊗ dξj, (4.8)

We are assuming that the coefficients Gij form a matrix [Gi

j] with positive de-

terminant, so we can compute something akin to the pull-back of the metric tensor

m:

eij Gik Gj

` dξk ⊗ dξ`. (4.9)

Since [eij] is a symmetric, positive-definite matrix and [Gik] is a matrix with posi-

tive determinant, [eijGikG

j`] is another symmetric, positive-definite matrix, and the

object in Eq. (4.9) is Riemannian metric.

2We are relaxing the requirement that metric tensors be infinitely differentiable.

69

We ask, “Is there an embedding of Ψ(B) into E3 that has G as its gradient

and that is isometric with this new metric?” If the answer is affirmative, then that

embedding is a deformation of the body, and G is a deformation gradient. The map

from Ψ(B) has the form x f , where f(ξ1, ξ2, ξ3) = (ζ1, ζ2, ζ3) and x is the position

function from (ζ1, ζ2, ζ3)-space to E3.

There are several ways in which such a tensor field can fail to be a deformation

gradient. We consider simple examples in cylindrical coordinates. We will use the

notation introduced in Tables 3.1 and 3.2.

4.3 Example: Immersion without Embedding

We have already considered an example of two-point tensor field of the form

G = γ1 er ⊗ER +γ2

γ1

eθ ⊗EΘ + ez ⊗EZ , (4.10)

where γ1 and γ2 are constants. If G is a local deformation gradient, then it must

have the form

er ⊗GRAD r + eθ ⊗GRAD θ + ez ⊗GRAD z

= er ⊗(

∂r

∂RER +

∂r

∂ΘEΘ +

∂r

∂ZEZ

)

+ eθ ⊗(

∂θ

∂RER +

∂θ

∂ΘEΘ +

∂θ

∂ZEZ

)

+ ez

(∂z

∂RER +

∂z

∂ΘEΘ +

∂z

∂ZEZ

). (4.11)

If G is of this form, then

GRAD r = γ1 ER, GRAD θ = γ2 EΘ, GRAD z = EZ . (4.12)

To test whether these are actual gradients, we compute the curls of each. For

example,

CURL(γ1 ER

)= REΘ∂γ1

∂Z− EZ

R

∂γ1

∂Θ= 0. (4.13)

The curls of γ2 EΘ and EZ are also identically zero. This means that G is locally

a deformation gradient. We will see below that G induces an immersible metric

70

tensor on Ψ(B). See Appendix F in Marion and Thornton (1988) for gradient and

curl in cylindrical coordinates.

In the spaces of coordinates, G corresponds to

G = γ1∂

∂r⊗ dR +

γ2

γ1

∂Θ⊗ dΘ +

∂z⊗ dZ. (4.14)

This is the tangent map for

(r, θ, z) = f(R, Θ, Z) =

(γ1R,

γ2

γ1

Θ, Z

). (4.15)

The new metric tensor on Ψ(B) is then

eij Gik Gj

` dξk ⊗ dξ` = γ21 dR⊗ dR +

(γ2

γ1

)2

r2 dΘ⊗ dΘ + dZ ⊗ dZ

= γ21 dR⊗ dR + γ2

2 dΘ⊗ dΘ + dZ ⊗ dZ. (4.16)

The Riemann-Christoffel curvature tensor associated with this metric tensor is iden-

tically zero, which guarantees that Ψ(B) can be mapped locally into E3 in a way

that agrees with the new metric.

The immersion of Ψ(B) isometric with the new metric is

(x f)(R, Θ, Z) = γ1R cos

(γ2

γ1

Θ

)i + γ1R sin

(γ2

γ1

Θ

)i + Z k. (4.17)

Figure 4.4 shows the images of Ψ(B) under x f if γ1 6= γ2. While (x f) may

send Ψ(B) into E3 in a fashion isometric with the new metric, the result is not

an embedding. Theorem 2.11.3 and the failure of (x f) ensure that there is no

embedding of Ψ(B) into E3 that agrees with the new metric.

4.4 Example: Lack of Immersion

Consider a one-to-one map of the form

f(R, Θ, Z) = (r(R), θ(R, Θ), Z). (4.18)

The gradient of (x f X−1

) is

F = r′ er ⊗ER +∂θ

∂Reθ ⊗ER +

∂θ

∂Θeθ ⊗EΘ + ez ⊗EZ , (4.19)

71

may not be the deformation of a globally one-to-one (X Ψ)(B)-to-E3 map. This

gradient is a two-point tensor that corresponds to

f∗ = r′∂

∂r⊗ dR +

∂θ

∂R

∂θ⊗ dR +

∂θ

∂Θ

∂θ⊗ dΘ +

∂z⊗ dZ. (4.20)

The pull-back of m is then

f ∗m =

((r′)2

+ r2

(∂θ

∂R

)2)

dR⊗ dR + r2 ∂θ

∂R

∂θ

∂Θ(dR⊗ dΘ + dΘ⊗ dR)

+ r2

(∂θ

∂Θ

)2

dΘ⊗ dΘ + dZ ⊗ dZ. (4.21)

Applying Eqs. (2.44) and (2.76) yields an identically zero Riemann-Christoffel curva-

ture tensor. See the Mathematica code in Appendix D to perform this computation.

By Theorem 2.11.2, Ψ(B) can be immersed in E3 isometrically with f ∗m.

We may create a metric tensor with non-zero Riemann-Christoffel curvature by

dropping some terms in f ∗m. For example, let M be the “diagonal” portion of f ∗m:

M =

((r′)2

+ r2

(∂θ

∂R

)2)

dR⊗ dR + r2

(∂θ

∂Θ

)2

dΘ⊗ dΘ + dZ ⊗ dZ (4.22)

Computing the Riemann-Christoffel curvature tensor associated with this metric

tensor has some rather messy nonzero terms (modify the Mathematica commands

in Appendix D), so we simplify yet again:

˜M = (r′)2

dR⊗ dR + r2

(∂θ

∂Θ

)2

dΘ⊗ dΘ + dZ ⊗ dZ. (4.23)

The Riemann-Christoffel curvature tensor associated with˜M (modify the Math-

ematica code in Appendix D yet again) has four terms that may be nonzero:

RΘRRΘ = −RΘRΘR = RRΘΘR = −RRΘRΘ

=rθΘ

r′

(2 (r′)2

θRΘ − rr′′θRΘ + rr′θRRΘ

). (4.24)

Since we assume r is a function of R alone, these terms are not identically zero

unlessr2θRΘ

r′= constant in R. (4.25)

72

4.5 Non-Metric Problems

Consider the identity deformation, in which each material point remains in the same

position. The deformation gradient is

F = er ⊗ER + eθ ⊗EΘ + ez ⊗EZ , (4.26)

where r = R, θ = Θ, and z = Z. The corresponding right Cauchy-Green tensor is

F T · F = err eR ⊗ER + eθθ EΘ ⊗EΘ + ezz EZ ⊗EZ

= eR ⊗ER + r2 EΘ ⊗EΘ + EZ ⊗EZ

= eR ⊗ER + R2 EΘ ⊗EΘ + EZ ⊗EZ (4.27)

because r = R.

Now consider

G = (cos φ er − sin φ ez)⊗ER + (− sin φ er + sin φ ez)⊗EZ

+ eθ ⊗EΘ, (4.28)

again with r = R, θ = Θ, and z = Z. φ is an R-dependent angle. This corresponds

to a rotation of the vectors er and ez in the θ = constant plane. The analogue of

the right Cauchy-Green tensor is

GT ·G =(err cos2 φ + ezz sin2 φ

)ER ⊗ER + eθθ EΘ ⊗EΘ

+(err sin2 φ + ezz cos2 φ

)EZ ⊗EZ

= ER ⊗ER + R2 EΘ ⊗EΘ + EZ ⊗EZ , (4.29)

which is the same as the Cauchy-Green tensor for the identity deformation. We

know that the associated Riemann-Christoffel curvature tensor will be identically

zero.

For a deformation with coordinates (R, Θ, Z) and (r, θ, z), the portion of the

deformation gradient corresponding to r is

∂r

∂Rer ⊗ER +

∂r

∂Θer ⊗EΘ +

∂r

∂Zer ⊗EZ

= er ⊗(

∂r

∂RER +

∂r

∂ΘEΘ +

∂r

∂ZEZ

)

︸ ︷︷ ︸gradient of r

. (4.30)

73

If G is the gradient of a deformation — even a local deformation — then the gradient

of r is

GRAD r = cos φ ER − sin φ EZ . (4.31)

We take the curl of this vector field. The components of the curl are

1

R

∂Θ(− sin φ) = 0, (4.32)

∂Zcos φ− ∂

∂R(− sin φ) = −φ′ cos φ, (4.33)

− 1

R

∂Θcos φ = 0. (4.34)

At a spot where φ′ cos φ 6= 0, the expression in Eq. (4.31) is not the gradient of a

function, and G is not even a local deformation, even though the Cauchy-Green

tensor corresponds to a metric tensor with identically zero Riemann-Christoffel cur-

vature tensor. G fails to be even a local deformation gradient because it describes

a re-orientation of material fibers without any bending, stretching, or compression

of the body. If the material fibers were re-oriented in this fashion in a deformation,

the result of this microscopic re-orientation would be a macroscopic change in the

shape of the body.

The problem with G cannot be detected with the Riemann-Christoffel curvature

tensor because the problem with G is not metric in nature. The only difference

between G and F in Eq. (4.26) is a rotation: er and ez in F are rotated about eθ

to create G. GT · G, the associated metric tensor, and the associated Riemann-

Christoffel curvature tensor cannot detect the rotation. Note that if F = R ·U is

the polar decomposition of a deformation gradient, then

F T · F = (R ·U )T · (R ·U ) = UT ·U . (4.35)

This shows that all information about orientation is absent from the right Cauchy-

Green tensor. It contains information about local arclengths only. Hence F T ·F , GT · G, and the associated metric tensors on Ψ(B) determine only that local

arclengths in the body manifold are compatible, at least locally, with a Euclidean

image of the manifold. These metric tensors cannot be used to detect changes in

orientation.

74

Deeper study of a growth tensor such as G in Eq. (4.28) requires consideration

of linear connections other than the Levi-Civita connection and will not be pursued

here. We will consider the types of incompatible growths seen in the previous

sections: (1) those that are local deformations but not global deformations, and (2)

those that are not isometrically immersible and have non-zero Riemann-Christoffel

curvature tensor.

75

R3

0 A B

0

Π

2 Π

R

Q

fR

3

0 Γ1A Γ1B

0

Π

Γ2

Γ12 Π

r

Θ

X` -1 x`

E3

E3

R3

0 A B

0

Π

2 Π

R

Q

f

R3

0 Γ1A Γ1B

0

Π

2 ΠΓ2

Γ12 Π

r

Θ

X` -1 x`

E3

E3

Figure 4.4: Top: If γ1 < γ2, then the image of the annulus under the isometricimmersion double-covers part of E3. Bottom: If γ1 > γ2, the image of the annulusdoes not close.

76

CHAPTER 5

ELEMENTARY MECHANICS OF SOLIDS

5.1 Material vs. Spatial Formulation; Conservation of Mass

To discuss continuum mechanics we assume the existence of a one-parameter family

of deformations χ(t, ·) : (X Ψ)(B) → E3, where χ(0, ·) is the identity map on (X Ψ)(B). Hence, a deformed configuration is often called not the final configuration,

but the current configuration (at time t).1

All physical quantities can be expressed as functions of three types of arguments:

the curvilinear coordinates (ξ1, ξ2, ξ3) of material points, the Euclidean locations

X(ξ1, ξ2, ξ3) of the material points in the reference configuration, or the Euclidean

locations x(ζ1, ζ2, ζ3) of the current configuration. The notations used will be dif-

ferent.

For example, if there are no sources or fluxes of mass in or on the body, then the

mass can be expressed in three ways as the integral of a mass density. First, one can

integrate a mass density over B, the set of curvilinear coordinates corresponding to

material points:

m(B) =

Ψ(B)

ρ(ξ1, ξ2, ξ3

)det (E1E2E3) dξ1dξ2dξ3. (5.1)

The mass is also equal to the integral of a density over (X Ψ)(B) ⊂ E3, the portion

of Euclidean space occupied by the body before any deformation:

m(B) =

(XΨ)(B)

ρ0(X) dV, (5.2)

1Later on we will consider statics, and we will consider deformations that consist of a growth

step followed by an elastic step. The configuration formed after both steps are complete will be

called the final state.

77

where dV is volume measure on (X Ψ)(B). Finally, the mass can be considered as

an integral of a density over the current configuration:

m(B) =

χ(t,(XΨ)(B))ρ(t, x) dv

=

(XΨ)(B)

ρ(t, χ(t, X)) det F (t, X) dV, (5.3)

where dv is volume measure on χ(t, (X Ψ)(B)), F (t, ·) is the gradient of χ(t, ·),and dV is volume measure on (X Ψ)(B). Since the position x = χ(t, X) ∈ E3 is

time-dependent, the density ρ(t, ·) : E3 → [0,∞) must also change with time.

Since these expression are also equivalent for all measurable subsets Ω of Ψ(B),

we can write

X(Ω)

ρ0(X) dV = m(Ω) =

χ(t,X(Ω))ρ(t,x) dv

=

X(Ω)

ρ(t,χ(t,X)) det F (t, X) dV, (5.4)

from which we conclude the pointwise equality

ρ0(X) = ρ(t,χ(t, X)) det F (t, X) (5.5)

for (almost) every X ∈ (X Ψ)(B). This is an expression of the conservation of

mass.

Since we will assume that there are no sources or fluxes of mass in or on the

body, we will express all quantities as functions whose domain is the reference con-

figuration, (X Ψ)(B). That is, the original spatial position (X Ψ)(p) of the

material point p ∈ B is the independent non-temporal variable. In some sense, the

coordinate point Ψ(p) = (ξ1, ξ2, ξ3) may be considered the independent variable.

This approach is known as the material or Lagrangian formulation of continuum

mechanics, as opposed to the spatial or Eulerian formulation, in which a stationary

spatial point x ∈ E3 is the independent spatial variable.

78

5.2 Other Integral Balance Laws

The balance laws for continuous bodies with no exchange of mass or momentum are

generalizations of those for collections of particles. It should be noted that these

are axioms of solid mechanics, as they cannot be derived from analogous results

for Newtonian mechanics of collections of point particles. See page 402 of Antman

(1995).

First, the rate of change of linear momentum is equal to the total force on the

body:

d

dt

(XΨ)(B)

∂χ

∂t(t, X)ρ0(X) dV

=

(XΨ)(B)

f(t, X) dV +

∂(XΨ)(B)

t(t, X, ∂(X Ψ)(B)) dS, (5.6)

or, since ρ0 and the spatial variable X ∈ (X Ψ)(B) are time-independent,

(XΨ)(B)

∂2χ

∂t2(t, X)ρ0(X) dV

=

(XΨ)(B)

f(t, X) dV +

∂(XΨ)(B)

t(t, X, ∂(X Ψ)(B)) dS. (5.7)

Here f is the body force per unit reference volume, which includes outside forces

such as gravity; and t is the traction applied to the body’s surface. The notation

indicates that t is a function of position and of the boundary ∂(X Ψ)(B).

The rate of change of angular momentum is equal to the applied torque:

d

dt

(XΨ)(B)

χ(t, X)× ∂χ

∂t(t, X)ρ0(X) dV

=

(XΨ)(B)

χ(t, X)× f(t, X) dV +

∂(XΨ)(B)

χ(t, X)× t(t, X) dS. (5.8)

Note that

d

dt

(χ× ∂χ

∂t

)=

∂χ

∂t× ∂χ

∂t+ χ× ∂2χ

∂t2

= χ× ∂2χ

∂t2. (5.9)

79

Balance of angular momentum can now be written as∮

∂(XΨ)(B)

χ× t dS +

(XΨ)(B)

(χ× f − ρ0χ× χtt) dV = 0. (5.10)

5.3 Stress Tensors

Local versions of the integral balance laws above can be found if we adopt Cauchy’s

postulate that the traction t depends on the surface ∂(X Ψ)(B) only through the

unit normal vector N at X ∈ ∂(X Ψ)(B); we drop the previous notation and write

t (t, X,N) instead. In such a case, Cauchy’s Stress Theorem ensures a relatively

simple expression for the traction.

Theorem 5.3.1 (Cauchy’s Stress Theorem) If t(t, ·, N(·)) and ρ0(·)χtt(t, ·) −f(t, ·) are continuous on (X Ψ)(B), then there exists a tensor field P (t, ·) on

(X Ψ)(B) such that

t (t, X,N ) = P (t, X) ·N .

See Section XII.7.14 in Antman (1995).

P is called the first Piola-Kirchhoff stress tensor. Note that it maps a vector in

the reference configuration (N) to a vector in the deformed configuration (t); see

Figure 5.1. Hence, P is a two-point tensor. The traction on the surface of the body

can now be expressed as ∮

∂(XΨ)(B)

P (t, X) ·N (X)dS. (5.11)

Recall Eq. (3.43), which described how to compute, in reference configuration coor-

dinates, a tensorial flux in the deformed configuration: for each measurable subset

Ω of Ψ(B) with well-behaved boundary ∂Ω,∮

∂(χX)(Ω)

A (x) · n(x) ds =

∂X(Ω)

(det F (X)) A(χ(X)) · F−T (X)︸ ︷︷ ︸P (X)

·N(X) dS,

(5.12)

where A is an Eulerian tensor. If we let P be the tensor integrand on the right-hand

side, then we define another tensor by

T = (det F )−1 P · F T . (5.13)

80

T is called the Cauchy stress tensor. It maps a unit normal vector n in the de-

formed configuration to the stress vector on the surface that has n as its normal;

see Figure 5.1. Since T maps from the tangent space at x into the same tangent

space, it is an Eulerian tensor.

N

X`

HWL

P×N = t t

HΧëX`

LHWL

T×n = t n

HΧëX`

LHWL

Figure 5.1: Left: The first Piola-Kirchhoff stress tensor maps unit normal vectors inthe reference configuration to traction vectors in the deformed configuration. Right:The Cauchy stress tensor maps unit normal vectors in the deformed configurationto traction vectors in the deformed configuration.

5.4 Differential Balance Laws

5.4.1 Linear Momentum

The entire body (X Ψ)(B) is not special with respect to the balance laws, so

Eqs. (5.7) and (5.10) should hold for each measurable subset Ω of Ψ(B) with well-

behaved boundary ∂Ω. The balance of linear momentum, combined with Cauchy’s

stress theorem, yields

X(Ω)

(χttρ0 − f) dV =

∂X(Ω)

P ·N dS. (5.14)

The first Piola-Kirchhoff stress tensor can be written as

P = P iJ ei ⊗EJ = P iJ ei ⊗EJ , (5.15)

where we use an upper-case index for reference configuration and a lower-case index

for deformed configuration. We have also taken advantage of the fact that in E3

covariant and contravariant vectors are both traditional vectors of the same type.

81

Note that for a given deformation χ, the deformed coordinate vector ei is, ultimately,

a function of X ∈ (X Ψ)(B). We cannot simply apply the Divergence Theorem

to P iJEJ and ignore the change in ei as X varies. There is an applicable extension

of the Divergence Theorem for this situation.

Theorem 5.4.1 Divergence Theorem for a Two-Point Tensor

∂X(Ω)

P ·N dS =

∂X(Ω)

(P iJ ei ⊗EJ

) ·N dS

=

X(Ω)

P iJ||J ei dV, (5.16)

where the covariant derivative for a two-point tensor is

P iJ||J =

∂P iJ

∂ξJ+ ΓJ

LJP iL + PmJF `Jγi

m`, (5.17)

where γim` are Christoffel symbols of the second kind for the coordinate system on

the deformed configuration, anchored at x = χ(X).

See Proposition 4.29 and Theorem 2.6 in Marsden and Hughes (1983).

The balance of linear momentum can thus be written in differential form as

ρ0χtt = f + DIVP , (5.18)

where

DIVP = P iJ||J ei. (5.19)

Note that all quantities in Eq. (5.18) have spatial argument X ∈ (X Ψ)(B) but

take vector values anchored in the current configuration.

The expression equivalent to Eq. (5.18) in the spatial formulation is

ρ(t, x)d

dtv(t, x) = f(t,x) + div T (t, x), (5.20)

where we now consider f to have spatial argument x instead of X, div is divergence

with respect to coordinates in the current configuration, and

v(t, x) = χt(t, χ−1(t, x)). (5.21)

82

5.4.2 Angular Momentum

The analogous local expression of balance of angular momentum is

P · F T = F · P T . (5.22)

A derivation is presented in Appendix B.

Eq. (5.22) implies that the Cauchy stress tensor T is symmetric:

T T =((det F )−1 P · F T

)T

= (det F )−1 F · P T

= (det F )−1 P · F T = T . (5.23)

This is a local expression of balance of angular momentum in the spatial description.

83

CHAPTER 6

HYPERELASTICITY

6.1 Constitutive Relations

Consider for a moment Eq. (5.18) expressed with Cartesian coordinates (X, Y, Z)

for the reference configuration and Cartesian coordinates (x, y, z) for the current

configuration. Then the first Piola-Kirchhoff stress tensor is expressed as

P = P `M i` ⊗ iM , where i1 = i, i2 = j, i3 = k, (6.1)

and its divergence is

DIV P =∂P `M

∂XMi` (6.2)

The remaining quantities in Eq. (5.18) are

χ = χ` i`, f = f ` i`. (6.3)

Eq. (5.18) amounts to three scalar equations:

ρ0χ`tt = f ` +

∂P `M

∂XM, ` = 1, 2, 3. (6.4)

In the absence of body forces, which we shall assume, f ` ≡ 0. Eq. (5.22) represents

three linear constraints on P `M and the spatial partial derivatives of the χ`, so

we have six equations for our three unknowns χ` and nine unknowns P `M . The

remaining six equations come in the form of a constitutive relation that describes the

six independent entries in the stress tensor P as functions of the strains ∂χ`/∂XM ,

the position X, and perhaps other quantities, such as temperature.

It is through the constitutive relation that the specific physics of a material

is encoded into the differential equations of solid mechanics. For example, while

a block of rubber and a block of cement can be described as solid continua that

satisfy conservation of mass, of linear momentum, and of angular momentum, their

responses to applied tractions are expected to be quite different.

84

6.2 Simplifications

Numerous simplifications are applied in the search for constitutive relations. A

simple material is one in which the stress tensor at a point X ∈ (X Ψ)(B) in

the reference configuration depends on time t, position X, and the history of the

deformation gradient at that point (as opposed to the history of the deformation

gradient in a neighborhood of X):

P (t, X) = P (t, X,F t(·,X)), (6.5)

where the superscript indicates that this is the history of F up to the time t. P is

called the response function of the material.

A special case of a simple material is an elastic material, in which the only value

of F that plays a role in determining the stress is the instantaneous value at time

t, and t is not explicitly an argument of the response function:

P (t,X) = P (X,F (t, X)). (6.6)

In a Cauchy elastic material, the point X is also not an explicit argument:

P (t, X) = P (F (t, X)). (6.7)

6.3 Stress Power and Hyperelasticity

The power spent in the deformation of a solid is found through the stress and the

rate of strain. In the material formulation, the rate of strain is measured by the time

derivative of the deformation gradient. Recall the form of the deformation gradient:

F =∂ζ i

∂ξjej ⊗Ei. (6.8)

Since we are considering a fixed point in the reference configuration, the time-

dependence is in the deformed coordinates ζ i, and the time derivative of F has the

form

F t =∂ζ i,t∂ξj

ei ⊗Ej. (6.9)

85

The first Piola-Kirchhoff stress tensor is almost the appropriate stress to combine

with this strain rate. But both F t and P map from tangent spaces in the reference

configuration to tangent spaces in the current configuration. We need to use the

transpose of the first Piola-Kirchhoff stress tensor, denoted by S and often called

the nominal stress tensor:

S =(P k` ek ⊗E`

)T= P k` E` ⊗ ek. (6.10)

S · F t = ekiPk` ∂ζ i,t

∂ξjE` ⊗Ej. (6.11)

The trace of S ·F t is equal to the total inner product S : F t and is the stress power

of the deformation:

S : F t = tr (S · F t) . (6.12)

In a Cauchy elastic material, S will be determined by the local, instantaneous

value of F :

S = S(F ) =(P (F )

)T

, (6.13)

and the stress power density can be expressed as

tr(S(F ) · F t

). (6.14)

A Green elastic or hyperelastic material is a Cauchy elastic material for which

there exists a strain-energy density function W whose value at reference point X at

time t is determined by F (t, X). In such a case, the stress power density satisfies

d

dtW (F (t, X)) = tr

(S(F (t, X)) · F t(t,X)

). (6.15)

For a scalar function of a time-dependent tensor argument,

d

dtW (F (t, X)) =

∂W

∂F

∣∣∣∣F (t,X)

: F t(t, X); (6.16)

see Section XI.2 of Antman (1995). We conclude that for a hyperelastic material

the nominal stress tensor has the form

S(t, X) =∂W

∂F

∣∣∣∣F (t,X)

. (6.17)

86

6.4 Objectivity

One of the properties that all constitutive relations must have is objectivity or frame-

indifference. If the current configuration is viewed through a time-dependent set of

coordinates, the motion is expressed in the form

χ(t− a, X) = Q(t) · χ(t, X) + c(t), (6.18)

where Q(t) is a time-dependent rotation, c(t) is a time-dependent spatial transla-

tion, and a is a fixed translation in time. The first Piola-Kirchhoff stress tensor from

the second point of view must have form

P (t− a, X) = Q(t) · P (t, X). (6.19)

In Cauchy elastic materials, this requirement proscribes the types of dependence P

can have on F .

Theorem 6.4.1 A Cauchy elastic constitutive relation is invariant under rigid mo-

tions and time shifts if and only if it has the form

P (F ) = R · P (U ), (6.20)

where F = R ·U is the polar decomposition of F .

See Theorem XII.11.13 in Antman (1995) or Sections 4.2.1 and 4.2.2 in Ogden

(1984). See Appendix D for a discussion of the standard proof of the existence of

the polar decomposition and its relation to Theorem 6.4.1.

6.5 Isotropic Hyperelasticity

6.5.1 Isotropic Functions

We simplify yet again and assume that the value of the strain-energy density W

is unchanged if the direction of a stretch or compression is changed. This isotropy

87

indicates that there is no preferred direction in the material, i.e. that the mechanical

properties of the material are the same in each direction. In this case,

W (Q · F ) = W (F ) (6.21)

for each (Eulerian) rotation Q. In particular, if F = R·U is the polar decomposition

of the deformation gradient, then

W (F ) = W (R ·U ) = W (RT ·R ·U ) = W (U ). (6.22)

The strain-energy density is thus equal to some function of the stretch tensor U . Due

to the complications mentioned in Appendix D, however, we prefer to use a strain-

energy density that is a function the right Cauchy-Green tensor F T · F = UT ·U :

W (F ) = W (F T · F ). (6.23)

Real-valued isotropic functions of symmetric tensor arguments have a special

form.

Theorem 6.5.1 ϕ : Sym → R is isotropic if and only if there exists some function

ϕ : (0,∞)3 → R such that for each symmetric C,

ϕ(C) = ϕ(I1, I2, I3),

where I1, I2, I3 are the principal invariants of the symmetric tensor C:

I1 = tr (C) , I2 = det C tr(C−1

), I3 = det C.

See Section 37 in Gurtin (1981) for a proof.

6.5.2 Form of the Nominal Stress

For a isotropic hyperelastic material, the nominal stress has the form

S(F ) =∂

∂FW (F T · F ) = 2

∂W

∂C

∣∣∣∣∣F T ·F

· F T , (6.24)

88

since∂

∂F

(F T · F )

= 2F T . (6.25)

We use C as the symbol of the Lagrangian symmetric tensor argument of W . Note

the direction of the tangent space-to-tangent space mapping:

F T : tangent space at x −→ tangent space at X, (6.26)

and since the argument of W is Lagrangian,

∂W

∂C

∣∣∣∣∣F T ·F

: tangent space at X −→ tangent space at X. (6.27)

It should also be noted that, by Theorem 6.5.1, there is a function˜W such that

W (F T · F ) =˜W (I1, I2, I3), (6.28)

where I1, I2, I3 are the principal invariants of F T · F . The derivative of the strain-

energy with respect to its symmetric tensor argument then has the form

∂W

∂C

∣∣∣∣∣F T ·F

=∂˜W

∂I1

∂I1

∂C

∣∣∣∣∣∣F T ·F

+∂˜W

∂I1

∂I2

∂C

∣∣∣∣∣∣F T ·F

+∂˜W

∂I1

∂I3

∂C

∣∣∣∣∣∣F T ·F

. (6.29)

6.6 Hyperelasticity with Growth

6.6.1 Units without Growth

In hyperelasticity without growth, the strain-energy density has units

[W (F )] =elastic energy

reference volume, (6.30)

so the nominal stress tensor has units

[S] =[W (F )]

[F ]=

elastic energy / reference volume

deformed length / reference length

=elastic energy / deformed length

reference area. (6.31)

If we recognize elastic energy per final length as elastic force in the deformed con-

figuration, then the units of S reflect the role of P , its transpose, as a linear map

89

converting area in the reference configuration into stress in the deformed configura-

tion.

The Cauchy stress tensor T is related to S by

T = (det F )−1 F · S, (6.32)

so the units of T are

[T ] = [det F ]−1 [F ] [S]

=

(deformed volume

reference volume

)−1deformed length

reference length· elastic energy / deformed length

reference area

=elastic energy / deformed length

deformed area. (6.33)

These units reflect T ’s role as a linear map converting area in the deformed config-

uration to force in the deformed configuration.

6.6.2 Units with Growth

In a body with growth included via the multiplicative decomposition in Eq. (4.2),

we distinguish three states of the body. The pre-growth state is the traditional

reference configuration; the post-growth, pre-elastic response state (described by a

non-embeddable Riemannian manifold) is called intermediate; and the(Euclidean)

configuration found after the elastic response to the growth is called the final con-

figuration. The lengths, areas, and volumes attached to these three states will be

considered distinct.

In this model, the elastic strain-energy density’s argument is A, the sole (first-

order) descriptor of the elastic response. Just as F maps from a tangent space

in the reference configuration to a tangent space in the deformed configuration in

traditional hyperelasticity, A maps from a tangent space in the intermediate state

to a tangent space in the final configuration. Considering the units of W (F ) in

Eq. (6.30), we conclude that the units of W (A) are

[W (A)] =elastic energy

intermediate volume. (6.34)

90

The analogue of the nominal stress has units

[∂

∂AW (A)

]=

elastic energy / intermediate volume

final length / intermediate length

=elastic energy / final length

intermediate area, (6.35)

while the analogue of the Cauchy stress has units

[(det A)−1 A · ∂

∂AW (A)

]

=

(final volume

intermediate volume

)−1

· final length

intermediate length· elastic energy / final length

intermediate area

=elastic energy / final length

final area. (6.36)

A value of this analogue of the Cauchy stress is a linear map that coverts area in

the final configuration to stress in the final configuration. Since the elastic strain-

energy depends only on the elastic response A, this is the true Cauchy stress for a

body in which the deformation gradient is decomposed as F = A ·G.

The formula above provides the form of the (Eulerian) Cauchy stress tensor for

a hyperelastic body with growth, but we will need the first Piola-Kirchhoff stress

tensor P to express balance laws. In hyperelasticity with growth, P remains a

linear map from tangent spaces in the reference configuration to tangent spaces in

the final (post-growth, post-elastic response) configuration, and it is related to the

Cauchy stress tensor in the traditional manner:

P = (det F ) T · F−T

=

(det F

det A

)A ·

(∂

∂AW (A)

)· F−T

= (det G) A ·(

∂AW (A)

)· F−T . (6.37)

While pursuing a formulation of constitutive relations for bodies undergoing

incompatible growth, Chen and Hoger (2000) made the argument above rigorous

and applicable beyond hyperlasticity. They showed that the Cauchy stress in a

body that has undergone growth as described by a growth tensor G, should be

91

given by

T (t, x) = T (F (t, χ−1(t, x)) ·G−1(χ−1(t, x)))

= T (A(t, χ−1(t, x))), (6.38)

where T is a traditional Cauchy response function.

The relationship between Cauchy stress and first Piola-Kirchhoff stress is un-

changed by growth, so in this case the first Piola-Kirchhoff stress tensor has the

form

P = (det F ) T · F−T = (det F ) T (A) · F−T .

In the hyperelastic case, we find a traditional Cauchy response function by first

finding the nominal stress and then its transpose, the first Piola-Kirchhoff stress:

S =∂W

∂F, P =

(∂W

∂F

)T

. (6.39)

Then we employ Eq. (5.13):

T = (det F )−1 P · F T = (det F )−1

(∂W

∂F

)T

· F T . (6.40)

The Cauchy response function is then

T (A) = (det A)−1

(∂W

∂A

)T

·AT , (6.41)

where the partial derivative ∂W/∂A is evaluated at A.

The first Piola-Kirchhoff stress is then

P = (det F ) (det A)−1

(∂W

∂A

)T

·AT · F−T = (det G)

(∂W

∂A

)T

·G−T . (6.42)

This is expressed in Ben Amar and Goriely (2005) by

S = JG−1 ·WA, (J = det G) (6.43)

for the nominal stress, the transpose of the first Piola-Kirchhoff stress.

92

6.7 Example: Compressible Neo-Hookean

We consider here an isotropic strain-energy density of neo-Hookean type, expressed

as a function of eigenvalues:

W (λ1, λ2, λ3) = µ

(1

2

(λ2

1 + λ22 + λ2

3 − 3)− ln J

)+ λ (J − 1− ln J) , (6.44)

where λ21, λ2

2, λ23 are the eigenvalues of F T ·F (λ1, λ2, and λ3 are eigenvalues of F ),

and J = λ1λ2λ3. The eigenvalues and principal invariants of a symmetric matrix

are related by

I1 = λ21 + λ2

2 + λ23, (6.45)

I2 = λ21λ

22 + λ2

2λ23 + λ2

3λ21, (6.46)

I3 = λ21λ

22λ

23, (6.47)

where it must be remembered that λ21, λ2

1, and λ21 are the eigenvalues of F T ·F , not

the squares of the eigenvalues. See Chapter 1 of Gurtin (1981).

The given strain-energy density can be expressed as a function of the principal

invariants of F T · F :

˜W (I1, I2, I3) = µ

(1

2(I1 − 3)− ln

√I3

)+ λ

(√I3 − 1− ln

√I3

)

= µ

(1

2(I1 − 3)− 1

2ln I3

)+ λ

(√I3 − 1− 1

2ln I3

). (6.48)

The nonzero partial derivatives of˜W are

∂˜W

∂I1

2, (6.49)

∂˜W

∂I3

=1

2

(λI

−1/23 − (λ + µ) I−1

3

)(6.50)

The derivatives of the principal invariants are

∂I1

∂C= I, (6.51)

∂I3

∂C= (det C) C−1, (6.52)

93

where I is the Lagrangian identity tensor. See Section 3.5 of Marsden and Hughes

(1983). The full derivative of˜W is

∂˜W

∂C=

µ

2I +

1

2

(λI

−1/23 − (λ + µ) I−1

3

)(det C) C−1

2I +

1

2

(λI

1/23 − (λ + µ)

)C−1 (6.53)

because I3 = det C.

Evaluating this derivative at C = F T · F , we have

∂˜W

∂C

∣∣∣∣∣∣F T ·F

2I +

1

2(λJ − (λ + µ)) F−1 · F−T , (6.54)

where J = λ1λ2λ3 = I1/23 . The nominal stress tensor is

S = 2∂˜W

∂C

∣∣∣∣∣∣F T ·F

· F T = µF T + (λJ − (λ + µ)) F−1. (6.55)

The corresponding Cauchy stress tensor is

T (F ) = (det F )−1 F · S=

µ

JF · F T +

(λ− λ + µ

J

)I, (6.56)

where I is now the Eulerian identity tensor.

Now we consider the case with growth. The Cauchy stress determined by A is

T (A) =µ

JA

A ·AT +

(λ− λ + µ

JA

)I,

=1

JA

µ

(λ2

1 a1 ⊗ a1 + λ22 a2 ⊗ a2 + λ2

3 a3 ⊗ a3

)

+ (λJA − (λ + µ)) I , (6.57)

where JA = det A and we have expressed the symmetric Eulerian tensor A ·AT as

its spectral resolution:

A ·AT = λ21 a1 ⊗ a1 + λ2

2 a2 ⊗ a2 + λ23 a3 ⊗ a3, (6.58)

94

where the λ2i are the eigenvalues of A·AT and the ai = ai are the unit eigenvectors of

A ·AT , which can be chosen to form an orthonormal basis due to the real symmetry

of A ·AT .

The corresponding first Piola-Kirchhoff stress is

P = (det F ) T (A) · F−T

= JA (det G) T (A) ·A−T ·G−T (6.59)

= (det G)µA ·G−T + (λJA − (λ + µ)) F−T

. (6.60)

95

CHAPTER 7

CAVITATION IN THE HAUGHTON-OGDEN MEMBRANE

Elastic cavitation is the mechanism by which an empty cavity forms inside an elastic

solid under tension. The modern study of cavitation in elastic solids began with Gent

and Lindley’s experimental observation and theoretical consideration of rupture in

rubber cylinders Gent and Lindley (1959). The subject became a staple of nonlinear

elasticity after Ball (1982) put the subject on a solid mathematical foundation in

the context of finite hyperelasticity. In particular, Ball transformed the analysis

into a bifurcation problem and showed that a neo-Hookean sphere under hydrostatic

load supports both trivial (with no inner cavity) and cavitated solutions for a load

exceeding 5E/6, where E is Young’s modulus. Following Ball’s work, cavitation,

bifurcations, and solutions have been studied for a variety of specific strain energy

densities, geometries, loads, and constraints (Horgan and Polignone (1993), Horgan

and Polignone (1995), Sivaloganathan (1999), Horgan (2001), Sivaloganathan and

Spector (2002), Pericak-Spector et al. (2002)). Further experimental and numerical

works have validated the theoretical approaches of some of these theories (Tvergaard

(1996), Fond (2001), Dollhofer et al. (2004)). However, the role of elastic cavitation

as a mechanism for fracture initiation remains a controversial topic.

Elastic cavitation has also been studied for membranes. Interestingly, depend-

ing on the modeling assumptions on the membrane, cavitation may not be possible

Haughton (2001). More precisely, in a model of an elastic membrane with Varga

strain-energy density and with thickness negligible compared to the smallest curva-

ture of the “middle surface” of the membrane, no cavitated solution exists (Haughton

(1986)). However, if the membrane is considered to be a true two-dimensional object

with no thickness, cavitation is possible (Steigmann (1992)). It is therefore partic-

ularly interesting to study the case of a membrane subject to growth and see how

it allows for the existence of cavitated solutions, a drastic change of basic material

96

behavior.

7.0.1 Kinematics of an Axisymmetric Membrane with Growth

We consider an axisymmetric membrane, so we assign cylindrical coordinates to

both reference and deformed configurations. We assume r is a function of R alone,

θ = Θ, and z is a function of R and Z. The deformation gradient has the form

F =

(dr

dRer +

∂z

∂Rez

)⊗ER + eθ ⊗EΘ +

dz

dZez ⊗EZ

=

(dr

dRh1 +

∂z

∂Rk

)⊗ h1 +

r

Rh2 ⊗ h2 +

dz

dZk ⊗ k. (7.1)

We consider a growth tensor that is locally a deformation gradient but is not the

gradient of an embedding of the membrane:

G = γ1 h1 ⊗ h1 + γ2 h2 ⊗ h2 + k ⊗ k, (7.2)

where γ1 and γ2 are constants. The corresponding A is then

A = γ−11

(dr

dRh1 +

∂z

∂Rk

)⊗ h1 +

r

γ2Rh2 ⊗ h2 +

dz

dZk ⊗ k. (7.3)

The elastic strain quantity used in this model is the average of A over the

thickness of the membrane:

A =1

2H

∫ H

−H

A dZ

= γ−11

(dr

dRh1 +

1

2H

∫ H

−H

∂z

∂RdZ k

)⊗ h1 +

r

γ2Rh2 ⊗ h2

+1

2H

∫ H

−H

dz

dZdZ k ⊗ k (7.4)

A is simplified with the assumption of uniformity in the Z-direction of the membrane

and the assumption that the thickness is negligible compared to the smallest radius

of curvature found in the middle surface:

A = γ−11

dr

dRh1 ⊗ h1 +

r

γ2Rh2 ⊗ h2 +

h

Hk ⊗ k, (7.5)

97

where the R-dependent function h is the half-thickness of the deformed membrane.

See Haughton and Ogden (1978).

The eigenvalues of A are

λ1 = γ−11

dr

dR, λ2 =

r

γ2R, λ3 =

h

H. (7.6)

7.0.2 Thickness-Averaged Hyperelasticity

Haughton and Ogden (1978) showed that the mechanics of a hyperelastic membrane

can be approximated by treating the strain-energy density as a function of the

thickness-averaged strain tensor. We consider a traditional strain-energy density

but use A as its argument. Based on Eq. (6.42) for the first Piola-Kirchhoff stress

tensor, which is the transpose of the nominal stress tensor, the nominal stress arising

from this energy is

S = (det G) G−1 · ∂W

∂A

∣∣∣∣A

(7.7)

The corresponding Cauchy stress tensor is found via the thickness-averaged defor-

mation gradient:

T =(det F

)−1F · S

=(det A

)−1(det G)−1 (det G) A ·G ·G−1 · ∂W

∂A

∣∣∣∣A

=(det A

)−1A · ∂W

∂A

∣∣∣∣A

. (7.8)

If the strain-energy density is isotropic, then there is some real-valued W such

that W (A) =˜W (I1, I2, I3), where I1, I2, I3 are the principal invariants of of A

T ·A.

7.0.3 Equilibrium Equation

The assumptions of axial symmetry and of uniformity in the Z-direction reduce

the Eulerian equation of equilibrium (Eq. (5.20)) to a single ordinary differential

equation. The two-dimensional divergence of the radial force in the membrane is

the integral of T over the thickness of the membrane. In the deformed configuration,

98

this is 2hT , and the equation of equilibrium is

d

dr

(2ht11

)+

2h(t11 − t22

)

r= 0, (7.9)

to which a “plane stress” condition is added:

t33 = 0. (7.10)

Here, t11 is the radial Cauchy stress, t22 is the azimuthal Cauchy stress, and t33 is

the Cauchy stress normal to the middle surface of the sheet.

If we assume constant half-thickness H of the reference configuration, then since

h = Hλ3, Eq. (7.9) simplifies to

1

λ3

d

dr

(λ3t11

)+

t11 − t22

r= 0. (7.11)

Since this model assumes minimal curvature of the membrane’s middle surface,

r = r(R) is a monotonically increasing function. Thus Eq. (7.11) can be re-written

as

1

λ3

(dr

dR

)−1d

dR

(λ3t11

)+

t11 − t22

r= 0, (7.12)

or,

1

λ3λ1

d

dR

(λ3t11

)+

t11 − t22

r= 0, (7.13)

where all quantities are now considered to be functions of R.

7.0.4 Growth-Induced Cavitation

We consider the possibility of growth-induced cavitation. That is, we prescribe

growth parameters and zero applied loads and then seek an axisymmetric deformed

configuration such that r(0) > 0. Since the total radial traction at the cavity’s sur-

face is 2πrh t11

∣∣R=0

= 2πrλ3H t11

∣∣R=0

and since r(0) > 0, we impose the restriction

λ3 t11

∣∣R=0

= 0, (7.14)

with the same boundary condition at R = Rmax. In the following sections we

consider two types of elastic strain energy densities.

99

7.1 Absence of Cavitation in Neo-Hookean Material

The compressible neo-Hookean strain energy density has the form

W (A) = W(AT ·A)

=˜W (I1, I2, I3)

= µ

(1

2I1 − g(

√I3)

), (7.15)

where µ is constant, I1 = tr(AT ·A)

, and I3 = det(AT ·A)

= (det A)2.

The derivative of W is then

∂W

∂A= 2

∂W

∂C·AT

= 2

˜W

∂I1

∂I1

∂C+

∂˜W

∂I3

∂I3

∂C

·AT

= 2µ

(1

2I − 1

2I−1/23 g′

(√I3

)(det C) C−1

)·AT

= 2µ

(1

2I − 1

2I−1/23 g′

(√I3

)(det A)2 A−1 ·A−T

)·AT

= µ(AT − g′

(√I3

)(det A) A−1

). (7.16)

As a function of A, T is

T =(det A

)−1A · ∂W

∂A

∣∣∣∣A

= µ(det A

)−1A ·

(A

T − g′(√

I3

) (det A

)A−1

)

= µ(det A

)−1(A ·AT − (

det A)g′

(det A

)I)

, (7.17)

where I is the Eulerian identity tensor.

The radial component of the Cauchy stress is

t11 = h1 · T · h1 = µJ−1(λ2

1 − Jg′(J))

= µ

(λ1

λ2λ3

− g′(J)

), (7.18)

where λ1, λ2, λ3 are the eigenvalues of A, and J = det A = λ1λ2λ3.

100

The azimuthal Cauchy stress is

t22 = h2 · T · h2 = µJ−1(λ2

2 − Jg′(J))

= µ

(λ2

λ1λ3

− g′(J)

). (7.19)

The third principal Cauchy stress is

t33 = k · T · k = µJ−1(λ2

3 − Jg′(J))

= µ

(λ3

λ1λ2

− g′(J)

). (7.20)

The plane stress condition (Eq. (7.10)) can now be expressed as

λ23 = Jg′(J). (7.21)

The equilibrium equation takes the form

1

λ3γ1λ1

d

dR

(λ3 t11

)+

µ

r

(λ1

λ2λ3

− λ2

λ1λ3

)= 0, (7.22)

so thatd

dR

(λ3 t11

)=

µγ1

r

(λ2 − λ2

1

λ2

). (7.23)

We consider two types of behavior as R → 0: λ2 À λ1 and otherwise. The latter

condition will be denoted λ1À∼ λ2, which indicates that, asymptotically, λ1 grows

at a rate greater than or equivalent to the rate of growth of λ2.

7.1.1 λ2 À λ1

We assume λ1 = ε λ2, where ε is continuous and ε(R) → 0 as R → 0. The

equilibrium equation, Eq. (7.23), can be re-written as

d

dR

(λ3 t11

)=

µγ1

rλ2

(1− ε2

)

=µγ1

r

r

γ2R

(1− ε2

)

=µγ1

γ2

(1− ε2

)R−1. (7.24)

Given δ ∈ (0, 1), for small-enough R > 0,

d

dR

(λ3 t11

)>

µγ1

γ2

(1− δ2

)R−1, (7.25)

101

so that for sufficiently small R2 > R1 > 0,

λ3 t11

∣∣R=R2

− λ3 t11

∣∣R=R1

>µγ1

γ2

(1− δ2

)ln

(R2

R1

). (7.26)

If we fix R2 and let R1 → 0, we see that the total force on the surface at r(R1)

diverges. In this case a cavity cannot be sustained, even with (finite) nonzero radial

traction at the inner radius.

7.1.2 λ1À∼ λ2

In this case we let λ2 = ελ1, where ε is continuous and

limR→0

ε(R) =

0 if λ1 À λ2,

c > 0 if λ1 and λ2 are on the same scale.

The definitions in Eq. (7.6) and the relation ε λ1 = λ2 imply

ε(R)

γ1

dr

dR=

r

γ2 R, (7.27)

so that, given any δ > 0, for sufficiently small R2 > R1 > 0,

∫ R2

R1

dr

r=

γ1

γ2

∫ R2

R1

dR

ε(R) R>

γ1

γ2

∫ R2

R1

dR

(c + δ)R. (7.28)

Here c = 0 if λ1 À λ2. Integration yields

ln

(r(R2)

r(R1)

)>

γ1

γ2(c + δ)ln

(R2

R1

), (7.29)

or

Rγ1/γ2(c+δ)1 r(R2) > R

γ1/γ2(c+δ)2 r(R1). (7.30)

If we fix R2 and let R1 → 0, we see that r(R1) → 0. Cavitation is precluded in this

scenario, as well.

7.2 Varga Material

Haughton (2001) proved that elastic materials with the Varga strain energy density

(without volumetric growth) do not have cavitated configurations without radial

102

traction at the cavity’s surface. We show that a particular class of Varga strain en-

ergy densities permits cavitation in certain cases of homogeneous anisotropic growth.

The compressible Varga strain energy density has the form

W (A) = W (λ1, λ2, λ3) = µ (λ1 + λ2 + λ3 − g(J)) , (7.31)

where µ is constant; λ1, λ2, λ3 are the eigenvalues of A; and J = det A = λ1λ2λ3.

This strain-energy density is an isotropic function of A, but it cannot be ex-

pressed simply in terms of the principal invariants of AT ·A. We appeal to more

advanced results to find the Cauchy stress.

If W is an isotropic function of A and A = U ·R is the polar decomposition of

A, then there is a function W such that W (A) = W (U ), and

∂W

∂A=

∂W

∂U·RT . (7.32)

See Section 4.2 of Ogden (1984).

The sum λ1 +λ2 +λ3 is the first principal invariant (I1) of U , and J is the third

principal invariant (I3) of U . Their derivatives with respect to U are

∂I1

∂U= I,

∂I3

∂U= I3 U−1, (7.33)

where I is the Lagrangian identity tensor. See Section 4.2 of Ogden (1984).

The derivative of the strain-energy density is

∂W

∂A=

∂W

∂U·RT

= µ

(∂I1

∂U− g′(I3)

∂I3

∂U

)·RT

= µ(I − g′(J)JU−1

) ·R−1

= µ(RT − Jg′(J)A−1

), (7.34)

where RT = R−1 because R is orthogonal.

103

The Cauchy stress T is

T =(det A

)−1A · ∂W

∂A

∣∣∣∣A

= JA · µ(R

T − Jg′(J)A−1

)

= µJ−1(A ·RT − Jg′(J) I

)

= µJ−1(R ·U ·RT − Jg′(J) I

), (7.35)

where A = R · U is a polar decomposition of A. The other polar decomposition

of A has the form A = V ·R, where V is a symmetric, positive-definite Eulerian

tensor in the same sense that U is a symmetric, positive-definite Lagrangian tensor;

see Appendix C.

The conclusion is that

T = µ(J−1V − g′(J)I

)

= µ(J−1 (λ1 h1 ⊗ h1 + λ2 h2 ⊗ h2 + λ3 k ⊗ k)− g′(J)I

). (7.36)

The principal Cauchy stresses are

t11 = h1 · T · h1 = µ

(1

λ2λ3

− g′(J)

),

t22 = h2 · T · h2 = µ

(1

λ1λ3

− g′(J)

),

t33 = k · T · k = µ

(1

λ1λ2

− g′(J)

), (7.37)

and the plane-stress condition implies

1

λ1λ2

= g′(J), or λ3 = Jg′(J). (7.38)

We impose three restrictions on the function g found in the strain energy density.

First, since there should be no elastic energy in the absence of elastic deformation,

we require W (1, 1, 1) = 0, which implies g(1) = 3. Second, there should be no

stress in the elastically unstrained state, so each partial derivative of W must be

zero-valued when λ1 = λ2 = λ3 = 1; this implies g′(1) = 1. Finally, the requirement

104

of a positive bulk modulus in the unstrained state implies (see Scott (2007))

λ

3

d

dλtii(λ, λ, λ)

∣∣∣∣λ=1

> 0. (7.39)

For the compressible Varga material, we have

λ

(− 2

λ3− 3λ3g′′(λ3)

)∣∣∣∣λ=1

= − µ

3(2g′′(1) + 3) > 0. (7.40)

In response, we require that g′′(1) < −2/3.

7.2.1 An Equivalent ODE System

The equation of equilibrium and the plane stress condition can be converted into a

pair of ordinary differential equations with independent variable R and dependent

variables r and J . The first step is to combine Eqs. (7.13) and (7.37):

µ

γ1λ1λ3

d

dR

(λ3

(1

λ2λ3

− g′(J)

))+

µ

r

(1

λ2λ3

− 1

λ1λ3

)= 0,

which simplifies to

1

γ1λ1λ3

d

dR

(1

λ2

− λ3g′(J)

)+

1

r

(1

λ2λ3

− 1

λ1λ3

)= 0. (7.41)

Note that

d

dRλ−1

2 =d

dR

(γ2R

r

)= γ2

(1

r− Rr′

r2

)=

γ2

r

(1− γ1λ1

γ2λ2

).

When this is plugged into Eq. (7.41) we find

1

γ1λ1λ3

(γ2

r

(1− γ1λ1

γ2λ2

)− d

dR(λ3g

′(J))

)+

1

r

(1

λ2λ3

− 1

λ1λ3

)= 0.

d

dR(λ3g

′(J)) =γ2

r

(1− γ1λ1

γ2λ2

)+

γ1λ1λ3

r

(1

λ2λ3

− 1

λ1λ3

)

=γ2

r− γ1λ1

λ2r+

γ1λ1

λ2r− γ1

r

=γ2 − γ1

r

105

The plane stress condition, Eq. (7.38), allows us to eliminate λ3:

d

dR

(J (g′(J))

2)

=γ2 − γ1

r. (7.42)

If J ′ is the derivative of J with respect to R, then we have

J ′ (g′(J))2+ 2J g′(J) g′′(J) J ′ =

γ2 − γ1

r,

or

J ′ =γ2 − γ1

r g′(J) (g′(J) + 2J g′′(J)),

which is our first ordinary differential equation.

To derive our second ordinary differential equation, we adapt the definition of

λ1 (see Eq. (7.6)) with the definition J = λ1λ2λ3.

dr

dR= γ1λ1 = γ1

J

λ2λ3

= γ1J

(r/γ2R)Jg′(J)=

γ1γ2R

rg′(J)(7.43)

In summary, we have a pair of ordinary differential equations:

J ′ =γ2 − γ1

r g′(J) (g′(J) + 2Jg′′(J)), (7.44)

r′ =γ1γ2 R

r g′(J). (7.45)

In light of these equations, we impose more restrictions on g. We want both J ′

and r′ to remain finite within the material, so we require g′(J) and g′(J) + 2Jg′′(J)

to remain nonzero. As such, both of these expressions have fixed sign within the

material. Recall that g′(1) = 1 and g′′(1) < −2/3; based on this, we assume that

g′(J) is positive for J > 0 and g′(J) + 2Jg′′(J) is negative for J > 0.

7.2.2 Boundary Conditions

We seek growth-induced cavitation, so we assume zero radial force at both the inner

and outer radii of the cavitated configuration. With the Varga strain energy density,

the left-hand side of Eq. (7.14) takes the form

λ3 t11 = λ3 µ

(1

λ2λ3

− g′(J)

)= µ

(1

λ2

− J (g′(J))2

), (7.46)

106

where the last equality holds because of the plane-stress condition, Eq. (7.38). In

the presence of a cavity, there is no reason to assume that the principal stretches

are well-behaved near the surface of the cavity, which is reached in the R → 0 limit.

We consider the possible behaviors of J in this limit.

limR→0 J(R) ∈ (0,∞)

In the case of a nonzero finite limit limR→0 J(R) = J0, we have

limR→0

λ3 t11 = limR→0

µ

(1

λ2

− J (g′(J))2

)= − µJ0 (g′(J0))

2, (7.47)

where λ−12 = γ2 R/r → 0 as R → 0. Since we assume that g′(J) is positive for J > 0,

this leaves us with a nonzero force at the inner radius of the deformed configuration,

and this case is ruled out.

limR→0 J(R) = 0

We divide this case into three distinct sub-cases, depending on the behavior of g′(J)

as J → 0.

Sub-Case I: g′(0) = 0 Suppose first that g′(0) = 0. Since we assume g to

be twice continuously differentiable for J > 0, we have

g′(J) =

∫ J

0

g′′(x) dx < 0 (7.48)

because g′′(J) < 0 for J > 0. But this contradicts our assumption that g′(J) > 0

for J > 0.

Sub-Case II: g′(0) ∈ (0,∞) Now suppose that g′(0) ∈ (0,∞). Then J → 0

and J (g′(J))2 → 0 as R → 0. Recall that

J ′ =γ2 − γ1

r g′(J) (g′(J) + 2Jg′′(J)), while

d

dRJ (g′(J))

2=

γ2 − γ1

r.

Since J(0) = 0 and since J(R) > 0 for R > 0, J ′ must be nonnegative at R = 0.

Since g′(J) > 0 and g′(J) + 2Jg′′(J) < 0 for J > 0 (and thus for R > 0), we have

sign (J ′) =sign (γ2 − γ1)

sign (r g′(J)) sign (g′(J) + 2Jg′′(J))= sign (γ1 − γ2) . (7.49)

107

Thus we require γ1 > γ2.

But J (g′(J))2 is also zero at R = 0 and must also be positive for R > 0. Now

note that

sign

(d

dRJ (g′(J))

2

)=

sign (γ2 − γ1)

sign(r)= sign (γ2 − γ1) . (7.50)

For positivity of J (g′(J))2, we require γ2 > γ1, which contradicts the last result.

Thus we eliminate the case in which g′(0) = 0.

Sub-Case III: limJ→0 g′(J) = ∞ Finally, we consider limJ→0 g′(J) = ∞.

Since

d

dRJ (g′(J))

2=

γ2 − γ1

r

and since the right-hand side of this equation converges to a finite value as R → 0,

J (g′(J))2 also converges as R → 0. As argued above, since J(0) = 0 and since

J(R) must be positive for R > 0, we require γ1 > γ2. As a result, the derivative of

J (g′(J))2 is negative at R = 0, so J (g′(J))2, must be positive at R = 0. Otherwise,

J (g′(J))2, would be negative for small positive values of R.

Let limR→0 J (g′(J))2 = β > 0. Then

J(R) (g′(J(R)))2 − β =

∫ R

0

γ2 − γ1

r(s)ds < 0. (7.51)

This implies

0 < β − J(R) (g′(J(R)))2

= (γ1 − γ2)

∫ R

0

ds

r(s)<

(γ1 − γ2)R

r(0), (7.52)

as r(0) < r(R) for R > 0. Hence J (g′(J))2 = β +O(R) as R → 0. The radial force

at r(0) is then proportional to

limR→0

λ3 t11 = limR→0

µ

(1

λ2

− J (g′(J))2

)= − µβ.

Because of this nonzero radial force at R = 0, we rule out this case.

108

limR→0 J(R) = ∞

The only possibility that remains is the divergence of J as R → 0. We consider two

distinct limiting behaviors of g′(J) as J →∞.

Sub-Case I: limJ→∞ g′(J) > 0 In this case, proceeding as above shows that

limR→0

λ3 t11 = µ

(limR→0

1

λ2

− limR→0

J (g′(J))2

)= −∞. (7.53)

This precludes the possibility that limJ→∞ g′(J) > 0.

Sub-Case II: limJ→∞ g′(J) = 0 In this category we have found an example

of growth-induced cavitation that is sustained with zero radial traction on the cav-

ity’s surface. This example is presented in the next section.

7.2.3 A Necessary Condition

Before we explore this very specific example, we note a physically meaningful con-

dition that is necessary for growth-induced cavitation. Since J →∞ as R → 0 and

since J must be finite within the body, we must have J ′ → −∞ as R → 0. However,

the form of J ′ (Eq. (7.44)) and the restrictions on g′(J) and g′(J) + 2Jg′′(J) ensure

that J ′ is negative throughout the body. And since

sign(J ′) = sign (γ1 − γ2) ,

cavitation with zero radial traction is possible only if γ2 > γ1.

7.2.4 Constructing an Appropriate g′(J)

The following class of functions satisfies all the requirements so far imposed on g(J):

g(J) = A

(− 1

αJ−α + 3 +

1

α

)+ (1− A)

(2J1/2 + 1

),

where α > 0 and A > (6α + 3)−1.

However, even this small class must be reduced before we have a physically

acceptable g. Since the thickness of the membrane is proportional to λ3, λ3 must

109

not diverge as R → 0. The plane-stress condition requires λ3 = Jg′(J) (Eq. (7.38)),

which in this case amounts to

λ3 = AJ−α + (1− A)J1/2.

Since we now assume J →∞ as R → 0, this would imply λ3 →∞ as R → 0. Hence

we set A = 1:

g(J) = − 1

αJ−α + 3 +

1

α, (7.54)

g′(J) = J−α−1, (7.55)

g′(J) + 2Jg′′(J) = −(2α + 1)J−α−1. (7.56)

Eqs. (7.44)-(7.45) now take the form

J ′ =(γ1 − γ2) J2α+2

(2α + 1) r, (7.57)

r′ =γ1γ2 R Jα+1

r. (7.58)

7.2.5 Converting to an Autonomous System

Thanks to the explicit presence of R in Eq. (7.58), Eqs. (7.57)-(7.58) do not define

an autonomous system. To produce an equivalent autonomous system, set τ =

− ln(R/B) for some B > 0, and set q = r/R.

dJ

dτ=

(γ2 − γ1

2α + 1

)J2α+2 q, (7.59)

dq

dτ= q − γ1γ2 Jα+1 q−1, (7.60)

and the behavior as R → 0 corresponds now to the behavior of the two dependent

variables as the new “time” τ →∞

7.2.6 Converting to a Lotka-Volterra System

Demonstrating the existence of a solution with the desired asymptotic behavior is

easier if the autonomous system above is converted yet again, to a Lotka-Volterra

110

system. Set

Z1 = J2α+1 q−1, (7.61)

Z2 = Jα+1 q−2, (7.62)

so that

dZ1

dτ= −Z1 + (γ2 − γ1) Z2

1 + γ1γ2 Z1Z2, (7.63)

dZ2

dτ= −2Z2 + 2γ1γ2 Z2

2 +(γ2 − γ1)(α + 1)

2α + 1Z1Z2 (7.64)

7.2.7 Asymptotics of the Lotka-Volterra System

We have established that in a configuration with a cavity, both J and q diverge as

R → 0. Hence as τ →∞,

Z2

Z1

=Jα+1q−2

J2α+1q−1= J−αq−1 → 0. (7.65)

Any orbit of the Lotka-Volterra system for which Z2/Z1 diverges or converges to a

positive constant, does not correspond to a cavitated configuration.

We introduce one more useful fact before addressing specific orbits. The defini-

tion of q and Eq. (7.60) reveal that

dr

dτ=

d

(q B e−τ

)= q′ B e−τ − q B e−τ = − γ1γ2 Jα+1q−1 B e−τ . (7.66)

Combining this with the definition q = r/R and Eq. (7.62) shows that

d

dτln r = −γ1γ2 Z2. (7.67)

Now we may consider the growth or decay of Z2 to determine whether ln r converges

as R → 0 (as τ →∞).

Unbounded Orbits

We consider only those unbounded orbits on which Z1 À Z2 asymptotically. It is

evident from Eq. (7.64) that if the orbit does not converge to a fixed point and

Z1 À Z2, then Z2 does not decay rapidly to zero as τ →∞. Hence,

limτ→∞

ln r = −∞, (7.68)

111

Coordinates (z1, z2) Linearization Df(z1, z2) Eigenvalues

(0, 0)

( −1 00 −2

)−1, −2

(0, 1

γ1γ2

) (0 0

(γ2−γ1)(α+1)γ1γ2(2α+1)

2

)0, 2

(1

γ2−γ1, 0

) (1 γ1γ2

γ2−γ1

0 − (3α+12α+1

))

1, − (3α+12α+1

)

Table 7.1: Fixed points in the in (z1, z2)-plane

and limR→0 r = 0 on this orbit. No unbounded orbit corresponds to a cavitated

configuration.

The system defined by Eqs. (7.63)-(7.64) has three fixed points. The coordinates,

linearizations of the equations, and eigenvalues of these fixed points are listed in the

table below.

Fixed Point (0, 0)

The eigenvalues of Df(0, 0) indicate that (0, 0) is a sink. We consider orbits attracted

to (0, 0). Note that by Eqs. (7.61)-(7.62),

J3α+1 =Z2

1

Z2

(7.69)

so that by Eqs. (7.63)-(7.64),

d

dτJ3α+1 =

(γ2 − γ1)(3α + 1)

2α + 1Z1 J3α+1, (7.70)

ord

dτln

(J3α+1

)=

(γ2 − γ1)(3α + 1)

2α + 1Z1, (7.71)

112

or, even more directly,d

dτln J =

(γ2 − γ1

2α + 1

)Z1. (7.72)

By Eq. (7.63), given any δ ∈ (0, 1), for (Z1, Z2) in the first quadrant and close

enough to (0, 0),dZ1

dτ< (−1 + δ) Z1. (7.73)

For large-enough τ0 < τ , then,

Z1(τ) < Z1(τ0) e(−1+δ)(τ−τ0). (7.74)

Integrating Eq. (7.72) yields

ln

(J(τ)

J(τ0)

)=

(γ2 − γ1

2α + 1

) ∫ τ

τ0

Z1(s)ds

<

(γ2 − γ1

2α + 1

)Z1(τ0)

∫ τ

τ0

e(−1+δ)(s−τ0)ds

=

(γ2 − γ1

2α + 1

)Z1(τ0)

1− δ

(1− e(−1+δ)(τ−τ0)

)(7.75)

Taking the τ →∞ limit of Eq. (7.75) shows

limτ→∞

ln

(J(τ)

J(τ0)

)≤

(γ2 − γ1

2α + 1

)Z1(τ0)

1− δ< ∞. (7.76)

Eq. (7.76) reveals that on an orbit attracted to the fixed point (0, 0), J does not

diverge as τ → ∞. Such orbits correspond to configurations with nonzero radial

traction at the surface of the cavity.

Fixed Point(0, γ−1

1 γ−12

)

The presence of one positive eigenvalue and one zero eigenvalue of Df(γ−11 γ−1

2 , 0)

ensures the existence of a one-dimensional center manifold for this fixed point; see,

e.g., Perko (1996). At the fixed point, the center manifold is tangent to the 0-

eigenspace of Df(γ−11 γ−1

2 , 0). If any orbit approaches this fixed point , it lies on the

center manifold.

Suppose that there is an orbit that approaches this fixed point as τ →∞. Since

Z2 converges to a positive constant, Jα+1 and q2 must have the same asymptotic

growth rates. Since Jα+1 diverges, q2 grows at the same rate. But then

113

Z1 =J2α+1

q=

(Jα+1)(2α+1)/(α+1)

(q2)1/2>

Jα+1

q2→ (γ1γ2)

−1 .

However, Z1 → 0 as the orbit approaches this fixed point, contradicting the above

inequality. Hence the orbit approaching this fixed point (along the center manifold)

as τ →∞ does not correspond to a cavitated configuration.

Fixed Point((γ2 − γ1)

−1 , 0)

The linearization at this fixed point has one positive and one negative eigenvalue,

so this fixed point has a one-dimensional stable manifold; see Perko (1996). We

consider the orbit that approaches the fixed point along the stable manifold from

the first quadrant. Z1 converges to (γ2 − γ2)−1, while Z2 converges to 0. We will

now show that the stable manifold corresponds to a function r(R) that converges

to a positive constant as R → 0 (i.e. as τ →∞).

Recall from Eq. (7.67) that

d

dτln r = −γ1γ2 Z2. (7.77)

Z2 → 0 on the stable manifold, and we will show that Z2 → 0 fast enough that

limτ→∞ ln r > −∞.

Since Z2 → 0 and Z1 → (γ2 − γ1)−1 on the stable manifold, for small δ > 0 and

(Z1, Z2) close enough to((γ2 − γ1)

−1 , 0),

dZ2

dτ<

(−2 + δ +

(γ2 − γ1)(α + 1)

2α + 1

(1

γ2 − γ1

+ δ

))Z2

=

(−

(3α + 1

2α + 1

)+

(1 +

(γ2 − γ1)(α + 1)

2α + 1

)Z2,

so that for τ > τ0,

Z2(τ)

Z2(τ0)< exp

(−

(3α + 1

2α + 1

)+

(1 +

(γ2 − γ1)(α + 1)

2α + 1

(τ − τ0)

). (7.78)

If δ is chosen small enough and τ0 is chosen large enough, then from Eqs. (7.77) and

(7.78) for τ > τ0 we have

− γ1γ2 Z2(τ0) e−η(τ−τ0) <d

dτln r < 0, (7.79)

114

where

η =

(3α + 1

2α + 1

)−

(1 +

(γ2 − γ1)(α + 1)

2α + 1

)δ > 0.

Integration yields

γ1γ2 Z2(τ0)

η

(e−ητ − e−ητ0

)< ln

(r(τ)

r(τ0)

)< 0, (7.80)

so that

limτ→∞

ln r(τ) ≥ ln r(τ0)− γ1γ2 Z2(τ0)

ηe−ητ0 > −∞. (7.81)

Hence r(R) converges to a positive constant as R → 0.

It can be shown that each orbit attracted to the fixed point (0, 0) also has a

radius converging to a positive number as τ → ∞ (R → 0). However, such orbits

have been discounted because J does not diverge as τ →∞. We now establish that

J does diverge on the stable manifold of((γ2 − γ1)

−1 , 0).

It can be shown from Eqs. (7.61)-(7.62) that

J3α+1 =Z2

1

Z2

. (7.82)

On the stable manifold of((γ2 − γ1)

−1 , 0), Z1 → (γ2 − γ1)

−1 > 0 and Z2 → 0 as

τ → ∞. Hence J diverges as τ → ∞, and the radial traction at the surface of the

cavity is zero.

We have established the existence of a solution R 7→ r(R) that converges to a

positive constant as R → 0 and has zero radial stress in the same limit. Now we

must demonstrate that this solution has zero radial stress at some positive value

of R. The corresponding value of r is the outer radius of a cavitated configuration

with zero radial stress at the inner and outer radii.

7.2.8 Boundary Condition at Outer Radius

From Eqs. (7.46) and (7.55) we see that

λ3 t11 = µ

(γ2

q− J−(2α+1))

). (7.83)

115

1 1.5 2 2.5 3 3.5 4 4.5 5 5.5

Z1

Z1 = g

2−1

Figure 7.1: Phase portrait for γ1 = 1.0, γ2 = 1.2, α = 1: Each fixed point is markedby a black square. The verical line is Z1 = γ−1

2 . The solid lines are two orbits of thesystem. Between the slanted dashed lines, dZ1/dτ > 0 and dZ2/dτ < 0. The arrowemanating from the fixed point

((γ2 − γ1)

−1 , 0)

= (5, 0) points along the stableeigenspace at this point. The stable manifold at this fixed point is asymptoticallytangent to this line.

For R > 0 both J and q are positive and finite, so the zero-radial force condition

implies q/γ2 = J2α+1, or

Z1 =J2α+1

q=

1

γ2

. (7.84)

If the stable manifold of the fixed point ((γ2− γ1)−1, 0) intersects the line Z1 = γ−1

2

in the (Z1, Z2)-plane, then the portion of the orbit from the fixed point to this line

corresponds to a cavitated configuration that has zero radial traction applied at its

inner and outer radii.

dZ1

dτ= 0 for Z2 = −

(γ2 − γ1

γ2γ1

)Z1 +

1

γ1γ2

, (7.85)

dZ1

dτ= 0 for Z2 = −

(α + 1

4α + 2

)(γ2 − γ1

γ2γ1

)Z1 +

1

γ1γ2

. (7.86)

116

The slope of line in Eq. (7.85) is more severe than that of the line in Eq. (7.86), and

these lines intersect at the fixed point (0, γ−11 γ−1

2 ). It can be seen from Eqs. (7.63)-

(7.64) that dZ1/dτ > 0 and dZ2/dτ < 0 between these lines in the first quadrant

of the plane. Hence any orbit that is in this region for some τ0 is trapped in this

region for τ ≤ τ0 and approaches the fixed point (0, γ−11 , γ−1

2 ) as τ → −∞.

The stable manifold of the fixed point((γ2 − γ1)

−1 , 0)

is asymptotically tangent

to the “stable eigenspace” at this fixed point. The linearization of Eqs. (7.63)-(7.64)

at this fixed point reveals that

−(2α + 1)γ2γ2

(5α + 1)(γ2 − γ1)

(7.87)

is a representative vector from this eigenspace. When attached to the fixed point((γ2 − γ1)

−1 , 0), this vector points into the region between the lines described in

Eqs. (7.85)-(7.86); see Fig. 7.1. The stable manifold lies between the lines for all τ

and thus intersects the line Z1 = γ−12 . The orbit of interest to us approaches the

fixed point((γ2 − γ1)

−1 , 0)

as τ →∞ and the fixed point (0, γ−11 γ−1

2 ) as τ → −∞.

7.2.9 Numerical Construction of Solutions

To compute the radius of a cavity numerically, a point on the stable eigenspace

attached to the fixed point((γ2 − γ1)

−1 , 0)

was chosen very close to the fixed point.

With this point as initial condition, the system Eqs. (7.63)-(7.64) was integrated

numerically backward in τ until the numerical orbit reached the line Z1 = γ−12 . This

enforced a condition of zero radial traction at the maximum value of R, i.e. at the

outer surface of the membrane. The elasped τ was then used to compute

r = qR =

(Zα+1

1

Z2α+12

)1/(3α+1)

Be−τ . (7.88)

The coordinates of the initial condition provided the values of Z1 and Z2, and B was

set to 1. The results showing limR→0 r(R) as a function of γ2 are shown for various

values of γ1 in Fig. 7.2.

117

The same numerical integration was used to compute the outer radius of the

membrane after deformation. In Eq. (7.88), τ is set to 0, B is set to 1, Z1 is set to

γ−12 , and the value of Z2 is that found by numerical integration to the point where

the orbit crosses the line Z1 = γ−12 .

3.0

1.5

0.25

1.5

1.75

4.0

0.75

0.5 2.01.0

1.0

3.5

0.0

2.5

1.25

0.5

2.50.5 4.0

3.0

1.5 3.0

2.0

2.0

3.5

1.0

0.5

1.5

3.5

1.0

2.5

b=r(B)a=r(A)

γ2

γ1=0.5

γ 1=0.5 γ

1=1.0

γ 1=1.0

γ1=1.5

γ 1=1.5

γ 1=2.0

γ1=2.0

γ1=2.5

γ 1=2.5

γ1=3

γ1=3 γ

2

Figure 7.2: Cavity radius a = r(0) and outer radius b = r(B) as functions of γ2 forvarious values of γ1

On each curve the radius of the cavity increases with γ2 − γ1. But the curves

describing the cavity’s radius are not mere translates of one another; the growth

with γ2− γ1 is slower for larger values of γ1. On the other hand, the curves showing

the radius of the membrane do appear to be parallel.

7.3 Conclusion

We have added homogeneous anisoptropic growth to a model of a three-dimensional

elastic membrane. We have demonstrated that a membrane whose elastic behavior

is neo-Hookean cannot sustain a cavity without radial traction on the cavity. While

growthless elastically compressible Varga membranes do not permit cavity forma-

tion without radial traction on the cavity, we have found a class of growing Varga

membranes that do permit such solutions.

In this special class of growing Varga membranes, cavity formation with zero ra-

dial traction on the cavity occurs as long as the azimuthal growth factor γ2 exceeds

118

the radial growth factor γ1. Cavity formation is independent of the sizes of these

factors; in particular, cavitation is found regardless of whether the growth factors

indicate shrinkage or expansion in the growth stage. Numerical results show, how-

ever, that the cavity’s radius is a function of both γ1 and γ2 and not of γ2 − γ1

alone.

119

CHAPTER 8

GLOBAL CONSTRAINTS AND THE INDUCED THEORY OF SHELLS

The equations for the Haughton-Ogden membrane were derived via Taylor expansion

of various geometric and stress quantities and keeping finite numbers of terms. While

this approach is popular and has a long history, the assumptions at the foundation of

such a model are not necessarily consistent or physically realistic for large deflections.

Geometrically exact theories, in which explicit functional forms of deformations are

specified from the start, have none of the geometric inconsistencies found in older

theories. Further, geometrically exact theories allow the use of wide varieties of

constitutive relations for the same geometries.

We review briefly the assumptions of Kirchhoff’s classical theory of plates, con-

sider some objections to the popular Foppl-von Karman equations for deflections of

plates, and introduce the geometrically exact theory of plates induced by constrained

three-dimensional finite elasticity.

8.1 Kirchhoff’s Classical Theory of Plates

The mechanical theory of plates begins in 1766 with Euler’s model of elastic mem-

branes (Euler (1766)). Jakob Bernoulli (1759-1789), a student of Euler, extended

Euler’s theory to plates (Bernoulli (1789)). After the German physicist E. F. F.

Chladni’s presentation of vibration patterns on plates in Paris in 1809, the French

Academy of Sciences, per Napoleon’s suggestion, offered a prize for a mathematical

essay of plate vibrations that could predict Chladni’s experimental results. Soon

after, theoretical developments in the theory of plates were dominated by French

researchers, among them Sophie Germain, Navier, Poisson, de Saint Venant, and

Lagrange; see Szilard (2004). But it was Kirchhoff, a German, who first produced

a theory that included both bending and stretching (Kirchhoff (1876)). Kirchhoff

120

also introduced the use of virtual displacements into the theory.

Kirchhoff’s theory began with assumptions on geometry and mechanical behav-

ior, enumerated as follows in Szilard (2004):

1. The material is homogeneous, isotropic, and linearly elastic. That is, it obeys

Hooke’s law.

2. The plate is initially flat.

3. The middle surface of the plate remains unstrained during bending.

4. The constant thickness of the plate is small compared to its other dimensions.

That is, the smallest lateral dimension of the plate is at least ten times the

thickness.

5. The transverse deflections are small compared to the plate thickness. A max-

imum deflection of one tenth of the thickness is considered the limit of the

small-deflection theory.

6. Slopes of the deflected middle surface are small compared to unity.

7. Sections taken normal to the middle surface before deformation remain plane

and normal to the deflected middle surface. Consequently, shear deformations

are neglected.

8. The normal stress in the direction transverse to the plate surface can be ne-

glected.

The modern, geometrically exact theory that includes Kirchhoff plates preserves

some but not all of these assumptions. In particular, a class of allowed deformations

is specified by prescribing explicit functional forms. All assumptions and restrictions

are included in a consistent manner, which cannot be said for the classical theory

of plates. See, for example, Sections X.1 and XIV.15 of Antman (1995).

121

8.2 Foppl-von Karman Equations

A widely-used model for both static and dynamic large deflections of thin plates

began with the German August Foppl (1907) and was extended by the Hungarian

Theodore von Karman (1910). Their equilibrium partial differential equations for

transverse deflection w(x, y) and stress Φ(x, y) are

D

h∆2w = w2

xxΦ2yy + w2

yyΦ2xx − 2wxyΦxy +

pz

h, (8.1)

1

E∆2Φ = −1

2

(w2

xxw2yy + w2

yyw2xx − 2wxywxy

), (8.2)

where h is thickness, E is Young’s modulus, pz(x, y) is lateral load, and

D =Eh3

12(1− ν2)(8.3)

is the flexural rigidity of the plate, where ν is Poisson’s ratio. See Chapter 11 of

Szilard (2004) and Section XIV.14 of Antman (1995).

While the Foppl-von Karman equations (8.1)-(8.2) have found favor among both

applied scientists and mathematicians, numerous objections have arisen from the

school of rational mechanics. In the epilogue of Truesdell (1978), Truesdell com-

plained about an apparent lack of rigor in the derivation of the Foppl-von Karman

equations. He opined, however,

“[t]hat is not to say that theory is bad. Rather, whether it is good or bad

is assessable only after its predictions shall have been compared with a

rational theory.”

Truesdell subsequently asked Stuart Antman, an originator of many of the mod-

ern foundations and implementations of rational mechanics, to specify his objections.

According to Antman, the derivation of thr Foppl-von Karman equations relies upon

1. “ ‘approximate geometry’, the validity of which is assessable only

in terms of some other theory”.

2. “assumptions about the way the stress varies over a cross-section,

assumptions that could be justified only in terms of some other

theory”.

122

3. “commitment to some specific linear constitutive relation — linear,

that is, in some special measure of strain, while such approximate

linearity should be the outcome, not the basis, of a theory”.

4. “neglect of some components of strain — again, something that

should be proved mathematically from an overriding, self-consistent

theory”.

5. “an apparent confusion of the referential and spatial descriptions —

a confusion that is easily justified for classical linearized elasticity

but here is carried over unquestioned, in contrast with all recent

studies of the elasticity of finite deformations”.

Since Truesdell (1978), much rigorous analysis has been devoted to the Foppl-

von Karman equations, in particular by Ciarlet (2005). The equations still arise

only from asymptotics in the small-thickness limit, and the formal validity of the

equations depends on careful scalings of the loads with respect to the thickness.

Different types of scalings have been found to yield a hierarchy of plate models that

includes Foppl-von Karman (Friesecke et al. (2006)).

The end of Section XIV.14 of Antman (1995) points out that the Foppl-von

Karman equations have the advantage of semilinearity, so their analysis is simpler

than some geometrically exact approaches. However, for Cosserat shells, examples

of which we will study, the linearized equations are also semilinear, so the bifurcation

analysis is of the same complexity. For this and other reasons, we avoid the Foppl-

von Karman theory that has had such success in Efrati et al. (2007a), Efrati et al.

(2007c), Efrati and Kupferman (2009), and Efrati et al. (2009).

8.3 Plate Theory Induced from Three-Dimensional Elasticity

In intrinsic theories of shells and plates, the body is treated as intrinsically two-

dimensional, and balance laws are formulated for stress and torque in terms of

integrals along curves in the body. Such a theory requires the introduction of an ob-

jective constitutive relation for the two-dimensional body. We consider here induced

123

theories, which can be viewed as either approximations of or constrained versions of

full three-dimensional theory. In particular, we consider the special Cosserat theory

of shells, in which a shell consists of a middle surface from which material fibers

sprout, to form a three-dimensional body. See Figure 8.1. A plate is a shell that

has no curvature in its middle surface in its reference configuration.

Figure 8.1: A slice of a Cosserat shell, with a finite number of material fibers shown.The material fibers form a continuum in the model.

8.3.1 Global Constraints

The body is assumed to occupy only “plate-like” configurations. This requires spec-

ifying a specific class of allowed configurations. In the case of Kirchhoff constraints

on a plate, the deformed configuration must have the form

(x f)(R, Θ, Z) = r(R, Θ) + Z d(R, Θ), (8.4)

where the image of r is the middle surface of the deformed configuration, and the

director d is a unit vector anchored at r(R, Θ) and normal to the deformed middle

surface:

d =∂r

∂R× ∂r

∂Θ

/wwww∂r

∂R× ∂r

∂Θ

wwww . (8.5)

Eqs. (8.4) and (8.5) allow us to define tangent vectors to the set of allowed

configurations. If we view (xf) as a function of r and its partial derivatives alone,

124

then the tangent vector has the form

Mx = lim

ε→0

1

ε

(x f)

[r + ε

Mr,

∂r

∂R+ ε

∂Mr

∂R,∂r

∂Θ+ ε

∂Mr

∂Θ

]− (x f)

[r,

∂r

∂R,∂r

∂Θ

]

= limε→0

1

ε

(r + ε

Mr + Z

(r,R +εMr,R )× (r,Θ +ε

Mr,Θ )

‖(r,R +εMr,R )× (r,Θ +ε

Mr,Θ )‖

)−

(r + Z

r,R×r,Θ‖r,R×r,Θ ‖

)

=Mr + Z

Mr,R×r,Θ +r,R×M

r,Θ‖r,R×r,Θ ‖ . (8.6)

8.3.2 Stress Decomposition

This approach to shell theory includes a constitutive assumption. The first Piola-

Kirchhoff stress tensor is decomposed as

P = P act + P lat, (8.7)

where P lat satisfies ∫

(XΨ)(B)

P lat(X) :∂

Mx

∂XdV = 0 (8.8)

This is analogous to decomposing the Cauchy stress as T = T (F ) − pI in the

case of incompressibility, where T (F ) is a constitutive function and pI contributes

nothing to the stress power. However, the Kirchhoff constraints cannot be expressed

as pointwise constraints, so the condition in Eq. (8.8) is not a pointwise condition.1

P lat is called the latent or reactive or even Lagrange stress.

8.3.3 Weak Formulation of Balance Laws

The balance of linear momentum can be expressed as the Impulse-Momentum Law:∫ τ

0

∂X(Ω)\S3

P (t, X) · ν(X) dS dt +

∫ τ

0

S3∩∂X(Ω)

τ (t, X) dS dt

+

∫ τ

0

X(Ω)

f(t, X) dV =

X(Ω)

ρ(X) (χt(τ, X)− χt(0, X)) dV, (8.9)

1In Section XII.12 of Antman (1995), Antman does not prove that the Kirchhoff constraints

cannot be expressed as pointwise constraints on the deformation gradient or the right Cauchy-

Green tensor. Rather, he shows that the Kirchhoff constraints do not satisfy the criteria of the

theorems by which constraints such as incompressibility and inextensibility are shown to be local.

125

for well-behaved subsets Ω of Ψ(B), where S3 is the subset of the reference config-

uration on which traction τ is prescribed. In the case of Kirchhoff plates, that will

be the surface formed by the boundary of the middle surface and the material fibers

attached to that boundary.

Antman and Osborn (1979) proved the equivalence of the Impulse-Momentum

Law and the Principle of Virtual Power:

∫ ∞

0

(XΨ)(B)

(P :

∂Mx

∂X− f · M

x− ρχt ·Mxt

)dV dt

−∫ ∞

0

∂(XΨ)(B)

τ (t, X) · Mx(t, X) dS dt

−∫

(XΨ)(B)

ρ(X)χt(0,X) · Mx(0, X) dV = 0 (8.10)

for all well-behavedMx with compact support in X(B)× [0,∞) that satisfy bound-

ary conditions specified by the constraints. For an introduction and examples, see

Sections XII.8 and XII.9 of Antman (1995).

For statics of a Kirchhoff plate free of body forces, the Principle of Virtual Power

gives ∫

(XΨ)(B)

(P act + P lat) :∂

Mx

∂XdV =

∂(XΨ)(B)

τ · Mx dS. (8.11)

By Eq. (8.8), Eq. (8.10) can be re-written as

(XΨ)(B)

P act :∂

Mx

∂XdV =

∂(XΨ)(B)

τ · Mx dS. (8.12)

This will be the basis for our differential equations of equilibrium in the following

chapters. It will be sufficient for deriving equations representing balance of linear

momentum and of angular momentum.

126

CHAPTER 9

THE FLAT KIRCHHOFF PLATE

9.1 Kinematics

9.1.1 Deformation Gradient

In this chapter we will be considering statics of cylindrical deformations of axisym-

metric plates. The reference configuration has the form

X(R, Θ, Z) = R cos(Θ) i + R sin(Θ) j + Z k, (9.1)

while the final configuration has the form

x(r, θ, z) = r cos(θ) i + r sin(θ) j + z k, (9.2)

where r is a function of R alone, θ(Θ) = Θ, and z(Z) = Z.

The deformation gradient is

F = r′ er ⊗ER + eθ ⊗EΘ + ez ⊗EZ

= r′ h1 ⊗ h1 +r

Rh2 ⊗ h2 + k ⊗ k. (9.3)

where the right vector in each tensor product is to be understood as a tangent vector

anchored at a point in the reference configuration, while the left vector is a tangent

vector anchored at a point in the final configuration.

We will need the following expression later:

F−T =1

r′h1 ⊗ h1 +

R

rh2 ⊗ h2 + k ⊗ k. (9.4)

9.1.2 Incompatible Growth

Inspired by the growth tensors used in Goriely and Ben Amar (2005) and Ben Amar

and Goriely (2005) for spherical shells, we consider growth tensors whose expressions

127

in “physical components” take the form

G(R, Θ, Z) = γ1(R) h1(θ)⊗ h1(Θ) + γ2(R) h2(θ)⊗ h2(Θ) + k ⊗ k. (9.5)

The possible (R, Θ)-dependence of the coefficients indicates that such G may not

even be the gradient of an isometric immersion, so it is unclear where the vectors

on the left-hand side of each tensor product are anchored. We seek to remove, or at

least clarify, this ambiguity.

If we have a map (x f) from Ψ(B) into E3, where

f(R, Θ, Z) = (r(R), θ(R, Θ), Z), (9.6)

then the gradient of χ = (x f X−1

) has the form

∂χ

∂X= r′ er ⊗ER +

∂θ

∂Reθ ⊗ER +

∂θ

∂Θeθ ⊗EΘ + ez ⊗EZ

= r′ h1 ⊗ h1 +∂θ

∂Rr h2 ⊗ h1 +

∂θ

∂Θ

r

Rh2 ⊗ h2 + k ⊗ k, (9.7)

where the vectors on the left-hand sides of the tensor products are anchored at

(x f)(R, Θ, Z) = r(R) cos(θ(R, Θ)) i + r(R) sin(θ(R, Θ)) j + Z k

= r(R) h2(θ(R, Θ)) + Z k, (9.8)

where r and θ are assigned by the function f .

The tensor field G in Eq. (9.5) may be produced from a gradient of the kind found

in Eq. (9.7) by dropping the “off-diagonal” term, i.e. the one involving eθ ⊗ ER.

Comparing Eqs. (9.5) and (9.7), we set

r′ = γ1(R) so r(R) = r(0) +

∫ R

0

γ1(s)ds, (9.9)

and∂θ

∂Θ=

R

r(R)γ2(R), so θ(R, Θ) =

R

r(R)γ2(R)Θ + θ(R, 0), (9.10)

where the function θ(·, 0) remains unknown.

Since the given G has no term corresponding to ∂θ/∂R, we lack the information

needed to specify θ(·, 0). even if that off-diagonal term were included, we could

128

specify θ(·, 0) only up to an additive constant, just as for r. This differs from the

case of constant γ1 and γ2 only in that the θ-value for Θ = 0 may be R-dependent.

If we choose an additive constant r(0) and a well-behaved function R 7→ θ(R, 0),

we have a completely specified map f(R, Θ, Z) = (r, θ, z) and the corresponding

differential f∗ and pull-back f ∗. To produce something analogous to G in Eq. (9.5),

we drop the off-diagonal term:

G = γ1∂

∂r⊗ dR +

R

rγ2

∂θ⊗ dΘ +

∂z⊗ dZ. (9.11)

This G, which is analogous to a pull-back but will be equal to a pull-back only if γ1

and γ2 satisfy certain differential equations, can be used to produce a new metric on

Ψ(B) by computing the analogue of the pull-back of m, the standard metric tensor

on the space of (r, θ, z)-points:

m = dr ⊗ dr + r2 dθ ⊗ dθ + dz ⊗ dz. (9.12)

Since this is not a true pull-back of a metric tensor, there is no standard notation

for this operation. We appeal to intuition to describe the operation. First we let m

act on the “left sides” of the tensor products in G:

m [G] = dr

[γ1

∂r

]dr ⊗ dR + r2 dθ

[R

rγ2

∂θ

]dθ ⊗ dΘ + dz

[∂

∂z

]dz ⊗ dZ

= γ1 dr ⊗ dR + rRγ2dθ ⊗ dΘ + dz ⊗ dZ. (9.13)

Then we let the transpose of the two-point tensor m[G] act on the left sides in G:

(m [G])T [G] = dr

[γ1

∂r

]γ1 dR⊗ dRdθ

[R

rγ2

∂θ

]Rrγ2 dΘ⊗ dΘ

+ dz

[∂

∂z

]dZ ⊗ dZ

= γ21 dR⊗ dR + R2γ2

2 dΘ⊗ dΘ + dZ ⊗ dZ. (9.14)

The Mathematica code in Appendix C can be modified to demonstrate that the

Riemann-Christoffel curvature tensor associated with this metric has four terms that

may be nonzero:

RRΘRΘ = −RRΘΘR = −RΘRRΘ = RΘRΘR

=Rγ2

γ1

(γ2γ′1 + Rγ′1γ

′2 − γ1 (2γ′2 + Rγ′′2 )) . (9.15)

129

Since we need a Riemannian metric (so that the grown image of set with positive

volume, has positive volume), we require γ1 > 0 and γ2 > 0. Hence the terms in the

Riemann-Christoffel curvature tensor are identically zero if and only if

γ2γ′1 + Rγ′1γ

′2 − γ1 (2γ′2 + Rγ′′2 ) = 0, (9.16)

orγ′1γ1

=(γ2 + Rγ′2)

γ2 + Rγ′2, (9.17)

which impliesγ2 + Rγ′2

γ1

= constant. (9.18)

In all other cases, the Riemann-Christoffel curvature tensor is not identically zero,

and the growth described by G precludes even isometric immersion.

9.1.3 Elastic Response

The local linear description of the body’s elastic response to incompatible growth

has the form

A = F ·G−1

=r′

γ1

h1 ⊗ h1 +r

γ2Rh2 ⊗ h2 + k ⊗ k, (9.19)

where the vectors on the left-hand sides of the tensor products are anchored at the

point in the final configuration, and the vectors on the right-hand side are anchored

at the position in Eq. (9.8).

The eigenvalues of√

F T · F are called stretches, but since A is not a deformation

gradient, we cannot use that label for the eigenvalues of√

AT ·A. We choose the

label pseudostretches. In this case they are

λ1 =r′

γ1

, λ2 =r

γ2R, λ3 = 1. (9.20)

9.1.4 Virtual Displacements

The Kirchhoff plate consists of a two-dimensional middle surface from which material

fibers extend perpendicularly. The deformation in Eq. (9.2) leaves the middle surface

130

flat and the material fibers unstretched and perpendicular to the middle surface.

The only possible virtual displacements are axially symmetric perturbations of the

middle surface, so in this caseMx has the form

Mx(R, Θ) =

Mr(R) h1(Θ), (9.21)

whereMr is any as-regular-as-needed real-valued function. The Principle of Virtual

Power for statics, Eq. (8.12), takes the form

plate

((h1 · P act · h1)

Mr′+ (h2 · P act · h2)

Mr

R

)dV =

∂(plate)

τ ·(Mrh1

)dS (9.22)

because all other projections of the gradient ∂Mx/∂X are zero.

It should be noted that

(h1 · P act · h1)Mr′+ (h2 · P act · h2)

Mr

R

=(Mr′h1

)· (P act · h1) +

( Mr

Rh2

)· (P act · h2)

=

(∂

Mx

∂R

)· (P act · h1) +

(1

R

∂Mx

∂Θ

)· (P act · h2) +

(∂

Mx

∂Z

)· (P act · k)

=1

R

(∂

∂RR

Mx

)· (P act · h1)− M

x · P act · h1

+

(1

R

∂Mx

∂Θ

)· (P act · h2) +

(∂

Mx

∂Z

)· (P act · k)

= Div(Mx · P act

)− M

x ·(

1

R

∂R(RP act · h1) +

1

R

∂Θ(P act · h2)

), (9.23)

where Div indicates divergence with respect to cylindrical coordinates. We have

assumed that since there is no extension or shear through the thickness of the plate,

(P act · k) is vertical, so that ∂ (P act · k) /∂Z is also vertical and is orthogonal toMx.

By the Divergence Theorem applied to the vector fieldMx · P act,

plate

Div(Mx · P act

)dV =

∂(plate)

Mx · P act · ν dS

=

∂(plate)

Mx · τ dS (9.24)

131

where ∂ (plate) is the boundary of the plate and ν is the outward-pointing unit

normal on this boundary. After cancellation of the surface traction terms, the weak

equation of equilibrium is∫

plate

Mr h1 ·

(1

R

∂R(RP act · h1) +

1

R

∂Θ(P act · h2)

)dV = 0, (9.25)

which must hold for all admissibleMr.

Due to the constraints imposed on the body, we have no equations to determine

the rate of change of stress in the Z-direction. However, in an induced theory

such as this, we consider not the pointwise values of stress in the three-dimensional

body; rather, we consider the stress integrated across the thickness of the plate. The

integrated stress is a function of the planar coordinates R and Θ only.

We take the weak equation of equilibrium and write the integral over the plate

as an iterated integral, first in Z and then in the planar coordinates:∫

slice

Mr h1 ·

[∫ H2

H1

(1

R

∂R(RP act · h1) +

1

R

∂Θ(P act · h2)

)dZ

]R dR dΘ = 0

(9.26)

We change the order of differentiation and Z-integration:∫

slice

Mr h1 ·

(1

R

∂R

∫ H2

H1

(RP act · h1) dZ +1

R

∂Θ

∫ H2

H1

(P act · h2) dZ

)R dR dΘ = 0

(9.27)

By our assumptions of symmetry, P act ·h2 must be proportional to h2, so P act ·h2 = (h2 · P act · h2) h2. And by symmetry, the coefficient (h2 · P act · h2) is Θ-

independent, so

∂Θ

∫ H2

H1

(P act · h2) dZ =∂

∂Θ

∫ H2

H1

(h2 · P act · h2) h2 dZ

=∂h2

∂Θ

∫ H2

H1

(h2 · P act · h2) dZ

= −h1

∫ H2

H1

(h2 · P act · h2) dZ. (9.28)

Our weak equation of equilibrium is now∫

slice

Mr

(∂

∂RR

∫ H2

H1

(h1 · P act · h1) dZ −∫ H2

H1

(h2 · P act · h2) dZ

)dR dΘ = 0.

(9.29)

132

Our symmetry assumption imply that the remaining mechanical quantities are Θ-

independent, so we can integrate in Θ automatically and view the functions in the

result as functions of R alone:

slice

Mr

(d

dRR

∫ H2

H1

(h1 · P act · h1) dZ −∫ H2

H1

(h2 · P act · h2) dZ

)dR = 0. (9.30)

If this is to hold for all admissibleMr, then the other factor of the R-integrand must

be identically zero:

d

dR

(R

∫ H2

H1

(h1 · P act · h1) dZ

)−

∫ H2

H1

(h2 · P act · h2) dZ = 0. (9.31)

This is the equation of equilibrium for our constrained axisymmetric plate. Closure

of the system of differential equations requires a constitutive relation.

9.2 Compressible Neo-Hookean Constitutive Relation

We consider a compressible neo-Hookean constitutive relation with strain energy

density introduced in Chapter 6:

W (λ1, λ2, λ3) = µ

(1

2

(λ2

1 + λ22 + λ2

3 − 3)− ln J

)+ λ (J − 1− ln J) ; (9.32)

In growthless elasticity, λ21, λ2

2, and λ23 are the eigenvalues of the Cauchy-Green

tensors F · F T and F T · F , and J = λ1λ2λ3 = det F .

Following the procedure seen in Section 6.7, we find Eq. (6.57) for the active

Cauchy stress as a function of A = F ·G−1:

T act(A) =1

JA

µ

(λ2

1 a1 ⊗ a1 + λ22 a2 ⊗ a2 + λ2

3 a3 ⊗ a3

)

+ (λJA − (λ + µ)) I , (9.33)

where ai = ai are eigenvectors of A ·AT that form an orthonormal basis. In this

case, the appropriate eigenvalues and eigenvectors are

a1 = h1, λ1 = r′/γ1, (9.34)

a2 = h2, λ2 = r/γ2R, (9.35)

a3 = k, λ3 = 1. (9.36)

133

The active portion of the first Piola-Kirchhoff stress tensor is found via Eq. (6.60):

P act = (det F ) T act(A) · F−T

Only the projections h1 · P act · h1 and h2 · P act · h2 of P act are needed for our

calculations.

h1 · P act · h1 = (det F ) h1 · T act · F−T · h1

=(det A) (det G)

J

(µλ2

1 + λJ − (λ + µ))h1 ·

(1

r′h1

)

= (det G)

(r′

γ1

)2

+ λJ − (λ + µ)

)1

r′

= γ1γ2

r′

γ21

+ λr/R

γ1γ2

− λ + µ

r′

)

=µγ2

γ1

r′ + λr

R− (λ + µ) γ1γ2

r′. (9.37)

A similar calculation shows that

h2 · P act · h2 =µγ1

γ2

r

R+ λr′ − (λ + µ) γ1γ2

r/R. (9.38)

The flatness of the middle surface and the lack of stretching or shear through the

thickness have ensured a complete lack of Z-dependence, so the thickness-integrated

stresses are∫ H2

H1

(h1 · P act · h1) dZ = (H2 −H1)

(µγ2

γ1

r′ + λr

R− (λ + µ) γ1γ2

r′

)(9.39)

∫ H2

H1

(h2 · P act · h2) dZ = (H2 −H1)

(µγ1

γ2

r

R+ λr′ − (λ + µ) γ1γ2

r/R

).(9.40)

This leaves us the ordinary differential equation

d

dR

(R

(µγ2

γ1

r′ + λr

R− (λ + µ) γ1γ2

r′

))=

µγ1

γ2

r

R+ λr′ − (λ + µ) γ1γ2

r/R. (9.41)

After differentiation and cancellation of λr′ from each side, we have(

µγ2

γ1

)′Rr′ +

µγ2

γ1

r′ +µγ2

γ1

Rr′′ − (λ + µ) γ1γ2

r′

− (λ + µ) R (γ1γ2)′

r′+

(λ + µ) Rγ1γ2

(r′)2 r′′

=µγ1

γ2

r

R− (λ + µ) Rγ1γ2

r/R. (9.42)

134

9.2.1 A Restriction on γ1 and γ2 at R = 0

We assume that r is at least twice continuously differentiable and that r′ is bounded

away from zero. By the boundedness of derivatives near R = 0 and the fact that

r(R)/R → r′(0) as R → 0, the limiting form of the equation above is, after cancel-

lation,γ2(0)

γ1(0)r′(0) =

γ1(0)

γ2(0)r′(0). (9.43)

The conclusion is that we need either r′(0) = 0 or γ1(0) = γ2(0) for a configuration

without a cavity.

We consider the first possibility. Suppose that near R = 0, r(R) = A(R)Rα,

where A is a slowly-varying function bounded above zero near R = 0. Since we

want r to be twice continuously differentiable, we assume α = 1 or α ≥ 2. If α = 1,

then r′(0) 6= 0, and we assume γ1(0) = γ2(0). If α ≥ 2, then

r′ = αARα−1 + A′Rα, r′′ = (α− 1)αARα−2 + 2αA′Rα−1 + A′′Rα, (9.44)

and the most singular terms in Eq. (9.42) are

− (λ + µ) (γ1γ2)′

r′/R≈ −(λ + µ) (γ1γ2)

αA

∣∣∣∣R=0

R2−α, (9.45)

−(λ + µ)γ1γ2

r/R2≈ −(λ + µ)γ1γ2

A

∣∣∣∣R=0

R2−α. (9.46)

Requiring that the coefficients agree at R = 0 would put undue restrictions on

the types of growth functions γi we may consider. We opt for the less restrictive

alternative: the growth functions we will employ will satisfy γ1(0) = γ2(0).

9.3 Conversion to a Dynamical System

We perform a change of variables found in Antman and Negron-Marrero (1987) and

in Section X.3 of Antman (1995). Let n be the radial stress,

n =µγ2

γ1

r′ + λr

R− (λ + µ) γ1γ2

r′, (9.47)

and let

τ =r

R. (9.48)

135

Finding r′ as a function of n yields two solutions, only one of which is positive. The

flatness of the middle surface requires that r′ be positive, so we choose the positive

solution:

r′ =γ1

2µγ2

(n− λ

r

R+

√(n− λ

r

R

)2

+ 4µ(λ + µ)γ22

)

=γ1

2µγ2

(n− λτ +

√(n− λτ)2 + 4µ(λ + µ)γ2

2

). (9.49)

The differential equation can now be stated in terms of n and τ :

d

dR(Rn) =

µγ1

γ2

r

R+ λr′ − (λ + µ)γ1γ2

r/R

=µγ1

γ2

τ +λγ1

2µγ2

(n− λτ +

√(n− λτ)2 + 4µ(λ + µ)γ2

2

)

︸ ︷︷ ︸λr′

− (λ + µ)γ1γ2

τ. (9.50)

Note also that

d

dR(Rτ) = r′ =

γ1

2µγ2

(n− λτ +

√(n− λτ)2 + 4µ(λ + µ)γ2

2

). (9.51)

We define a new independent variable s by

R = Rmaxes−1, or s = ln

(R

Rmax

)+ 1. (9.52)

If we set n(s) = n(R) and τ(s) = τ(R), then the R- and s-derivatives are related

by

d

dR(Rn) =

d

dR(Rn) = n + R

ds

dR

dn

ds= n +

dn

ds, (9.53)

d

dR(Rτ) =

d

dR(Rτ) = τ + R

ds

dR

ds= τ +

ds, (9.54)

where the prime on n and τ indicates differentiation with respect to s. The differ-

ential equations in n and τ are

dn

ds=

µγ1

γ2

τ +λγ1

2µγ2

(n− λτ +

√(n− λτ)2 + 4µ(λ + µ)γ2

2

)

− (λ + µ)γ1γ2

τ− n, (9.55)

136

ds=

γ1

2µγ2

(n− λτ +

√(n− λτ)2 + 4µ(λ + µ)γ2

2

)− τ , (9.56)

If we set N = n/µ and κ = λ/µ, then we can simplify these equations some:

dN

ds=

γ1

γ2

τ +κγ1

2γ2

(N − κτ +

√(N − κτ)2 + 4(1 + κ)γ2

2

)

− (1 + κ)γ1γ2

τ−N, (9.57)

ds=

γ1

2γ2

(N − κτ +

√(N − κτ)2 + 4(1 + κ)γ2

2

)− τ , (9.58)

where γi(s) = γi(R).

Were it not for the s-dependence of the γi, these equations would be autonomous,

and they could be analyzed from the point-of-view of dynamical systems. If the γi

satisfied some autonomous differential equations, then adding those equations to the

pair above would form an autonomous system. Below we consider several examples

of γi that satisfy autonomous equations.

To the differential equations in Eqs. (9.57) and (9.58) we add two boundary

conditions:

N(1) = 0, (zero radial stress at R = Rmax) (9.59)

lims→−∞

Rmaxes−1τ(s) = 0. (r(0) = 0) (9.60)

Now we consider examples with specific growth functions γ1 and γ2.

9.4 γ1(R) ≡ 1, γ′2(R) > 0

Suppose that γ1(R) ≡ 1 and γ2(R) = a+ b (R/Rmax)2. Since we want γ1(0) = γ2(0),

we set a = 1. Then γ2 satisfies

d

dR(Rγ2(R)) =

d

dR

(R +

b

R2max

R3

)= 1 + 3b

(R

Rmax

)2

= 1 + 3 (γ2(R)− 1) ,

(9.61)

so γ2 satisfies

γ2

ds+ γ2 = 1 + 3 (γ2 − 1) , or

γ2

ds= 2 (γ2 − 1) . (9.62)

137

We have the system of autonomous equations

ds=

1

2γ2

(N − κτ +

√(N − κτ)2 + 4(1 + κ)γ2

2

)− τ , (9.63)

dN

ds=

τ

γ2

2γ2

(N − κτ +

√(N − κτ)2 + 4(1 + κ)γ2

2

)

− (1 + κ)γ2

τ−N, (9.64)

dγ2

ds= 2 (γ2 − 1) . (9.65)

We will often express this system of differential equations by

dy

ds= f (y) , (9.66)

where y = (τ , N, γ2)T . Points y ∈ R3 will be described via τ -, N -, and γ2-

coordinates.

We will need some notation and basic results from the theory of ordinary differ-

ential equations and dynamical systems.

Definition 9.4.1 The domain of f is

D (f) =(τ , N, γ2) ∈ R3 : f (τ , N, γ2) is defined

. (9.67)

In the case at hand, D(f) consists of all of R3 except for the planes τ = 0 and γ2 = 0.

Definition 9.4.2 Each point y0 ∈ D(f) can be considered as the initial condition

of some solution of Eqs. (9.63)-(9.65):

dy

ds= f(y), y(0) = y0. (9.68)

The maximal interval of existence I(y0) is the largest interval of s-values (both

positive and negative) for which the solution of Eq. (9.68) exists and is unique.

Theorem 9.4.3 If f and the τ -, N-, and γ2-derivatives of f are continuous on the

“box” y ∈ R3 : ‖y − y0‖ < b, then Eq. (9.68) has at most one solution.

138

For a proof, see e.g. Section 3.3 of Brauer and Nohel (1989). Note that for

Eqs. (9.63)-(9.65), f is continuously differentiable in any finite box bounded away

from τ = 0 and γ2 = 0.

Definition 9.4.4 If s 7→ y(s) is a solution of Eq. (9.68), then the curve

y(s) ∈ R3 : s ∈ I(y0)

(9.69)

is called an orbit or trajectory of Eq. (9.68). We will often say that the solution

s 7→ y(s) “traces out” its orbit.

One of the most important properties of orbits of autonomous systems of dif-

ferential equations is that in regions of D(f) where f is continuously differentiable,

distinct orbits do not intersect. Hence, if s 7→ y(s) and s 7→ z(s) are solutions of

Eq. (9.68) with different initial conditions and y(s1) = z(s2) for some s1 6= s1, then

z(s) = y(s + s1 − s2), i.e. s 7→ z(s) is essentially s 7→ y(s) with a constant shift or

argument.

Definition 9.4.5 Much of the analysis will focus on the coordinates (τ , N). If

s 7→ (τ(s), N(s), γ2(s)) is a solution of Eqs. (9.63)-(9.65), then we will refer to the

function s 7→ (τ(s), N(s)) as the solution’s projection into the (τ , N)-plane. We will

also refer to the set of points

(τ(s), N(s)) : s ∈ maximal interval of existence , (9.70)

as the projection of the orbit.

9.4.1 Fixed Points and Linearization

The fixed points of Eqs. (9.63)-(9.65) form curves whose coordinates satisfy

γ2 = 1, N = (1 + κ)(τ + τ−1

). (9.71)

The coordinates of the fixed points with τ > 0 satisfy

γ2 = 1, τ =N +

√N2 + 4 (1 + κ)2

2 (1 + κ). (9.72)

139

At each fixed point, the derivative of the right-hand sides of Eqs. (9.63)-(9.65)

has the form

Df(τ , (1 + κ)

(τ + τ−1

), 1

)(9.73)

=

− (1+κ)(1+τ2)1+κ+τ2

τ2

1+κ+τ2

τ(1+κ−τ2)1+κ+τ2

(1+κ)(1+τ2)(1+κ−(−1+κ)τ2)τ2(1+κ+τ2)

−1−κ+(−1+κ)τ2

1+κ+τ2 −1+κτ − τ − κτ + 2κ(1+κ)τ

1+κ+τ2

0 0 2

.

This matrix has eigenvalues −2, 0, and 2, regardless of the values of κ and τ . This

indicates the presence of a stable manifold, a center manifold, and an unstable

manifold attached to each fixed point; see Perko (1996).

Each of these manifolds is traced out by a solution of Eqs. (9.63)-(9.65). The

unstable manifold is “unstable” in the sense that the solution that traces it out

approaches the fixed point at an exponential rate as s → −∞, i.e. as the dy-

namical system is run backwards. The eigenvector corresponding to the eigenvalue

2 describes the direction from which this solution approaches the fixed point as

s → −∞.

For a solution corresponding to a configuration of the plate, γ2(s) is non-constant

and γ2(s) ↓ 1 as s → −∞. The only eigenvector of the derivative matrix that has

nonzero γ2-component is the one associated with the eigenvalue 2:

(−−τ − κτ + 2τ 3

4 (1 + κ + τ 2), − 1 + 2κ + κ2 + 3τ 2 + 2κτ 2 − κ2τ 2 + 2τ 4 + 2κτ 4

4τ (1 + κ + τ 2), 1

)T

.

(9.74)

This means that the only orbit attached to a fixed point that could possibly cor-

respond to a configuration of the body is an unstable manifold. We will show in

the next section that there is an unstable manifold that corresponds to a plate with

zero radial stress at its periphery.

9.4.2 Existence of Solution

We will demonstrate that there is one solution that traces out an unstable manifold

and whose N -coordinate satisfies N(1) = 0.

140

For each s > −∞, we define the horizontal isocline

Γh(s) =

(τ , N) : τ =

N +√

N2 + 4(1 + κ) (κ + γ2(s)) γ2(s)

2 (κ + γ2(s))

(9.75)

and the vertical isocline

Γv(s) = (τ , N) : τ > 0 and (9.76)

1

2γ2(s)

(N − κτ +

√(N − κτ)2 + 4(1 + κ)γ2(s)2

)− τ = 0

.

Γh(s) consists of the points in the (τ , N)-plane with τ > 0 where the vector field has

zero τ -component at s. Γv(s) consists of the points in the same half-plane where

the vector field has zero N -component at s. If κ > 1, then Γv(s) has two branches,

one above Γh(s) and one below. See Figure 9.1.

GvHsL

GvHsL

GhHsL

GH-¥L

0 1 2 3 4

-5

0

5

10

15

Τ

Figure 9.1: Isoclines in the (τ , N)-plane at some s > −∞. The presence of the upperbranch of Γv(s) indicates that κ > 1 in this example. The line segments attachedto Γ(−∞) show the directions of the unstable eigenvectors.

141

Proposition 9.4.6 At s = −∞, Γh(s) and Γv(s) coincide with Γ(−∞).

Proof In the defining equation of the horizontal isocline Γh(s), set γ2 = 1. This

coincides with the equation for the curve of fixed points. Plug this formula for τ

into the defining equation of the vertical isocline (with γ2 set to 1) and the equation

becomes a tautology.

Proposition 9.4.7 At s = 1, the projection of each solution that traces out an

unstable manifold lies in the region between Γ(−∞) and the lower branch of Γv(1).

Proof First we note that the N -component of the “unstable eigenvector” in

Eq. (9.74),

− (1 + κ) (2τ 4 + (3− κ2) τ 2 + (1 + κ))

4τ (1 + κ + τ 2)(9.77)

is negative for all real τ . Hence each fixed point is asymptotically approached from

below (as s → −∞), which means each fixed point is approached asymptotically

from the region between Γ(−∞) and the lower branch of Γv(s).

Figure 9.2 shows the vector field for the dynamical system run in reverse. If

(τ(s), N(s)) is to the left of Γ(−∞) for some s > −∞, then τ(s) → 0 as s → −∞,

and the orbit does not approach a fixed point. If (τ(s), N(s)) is to the right of thew

lower branch of Γv(s) for some s > −∞, then τ(s) →∞ as s → −∞, and the orbit

does not approach a fixed point.

In particular, each solution whose orbit is an unstable manifold, lies in the region

between Γ(−∞) and the lower branch of Γv(1) at s = 1.

Definition 9.4.8 If we view each point y0 ∈ D(f) as the initial condition of the

problemdy

ds= f(y), y(0) = y0, (9.78)

then the flow of that dynamical system is a function defined as

Φ(y0, s) = y(s), s ∈ I(y0). (9.79)

142

GvHsL

GvHsL

GhHsL

GH-¥L1 2 3 4 Τ

-5

5

10

15

N = n Μ

Figure 9.2: The vector field of the dynamical system run in reverse

Let E be an open subset of D(f) such that f ∈ C1 (E). Define

Ω = (s,y0) ∈ R× E : s ∈ I(y0) . (9.80)

Then Φ ∈ C1(Ω), i.e., Φ is continuously differentiable in both its s and y0 arguments.

See Section 2.5 of Perko (1996).

Proposition 9.4.9 At s = 1, the positions of the solutions tracing out the unstable

manifolds form a continuous curve in the region between Γ(−∞) and the lower

branch of Γv(1).

Proof Fix τ0 > 0. Define E by

E = (τ , N) : (τ , N) is between Γ(−∞)

and the lower branch of Γv(1), and τ > τ0 > 0 . (9.81)

Then f ∈ C1(E), which implies that the flow Φ of the dynamical system is also

continuously differentiable. The Center Manifold Theorem (see Section 2.12 of Perko

(1996)) guarantees the unstable manifolds can be locally (near the fixed points)

143

continuously parametrized by s and an arclength coordinate for the curve of fixed

points. Hence for some s0 > −∞ with |s0| >> 1 there is a continuous curve formed

by the positions of the solutions that trace out unstable manifolds. See Figure 9.3.

Since f ∈ C1(E), the maximal interval of existence for each unstable manifold

includes the half-infinite interval (−∞, 1]. Hence the continuous curve formed by

the solutions at s = s0 is mapped continuously via Φ to its image at s = 1. By

continuity of the s = s0 curve and of Φ, the positions of the solutions at s = 1 also

form a continuous curve.

Note that E is bounded below by τ0 > 0. τ0 can be re-defined to allow E to

extend to any proximity to the τ = 0 line, and the argument still holds. On the

curve of fixed points, N → −∞ as τ → 0, so the continuous curve of positions at

s = 1 is infinite in extent. This argument leaves open the possibility that the map

from the fixed points to this curve may not be uniformly continuous.

GvH1L

GH-¥Lcurve formedby solutions

at s = s0 << 0

curve formedby solutions

at s = 1

0.5 1.0 1.5 2.0

-1

0

1

2

3

4

5

6

Τ

Figure 9.3: The continuity of the curve of fixed points is maintained in the curvesformed by the unstable manifold orbits for s > −∞.

Proposition 9.4.10 The continuous curve of positions of solutions at s = 1 inter-

sects the line N = 0.

Proof (Sketch) This can be proved by considering the one-point compactification

of the (τ , N)-plane, which is a 2-sphere, and examining the stereographic projections

onto the sphere of Γ(−∞), the lower branch of Γv(1), and the curve of solution

144

Figure 9.4: The stereographic projections of the lower branch of Γv(1) (outer curve)and Γ(−∞) (inner curve) intersect at the north pole of the sphere. The dashedcurve that they both intersect is the projection of the N = 0 line. The curve of lociat s = 1 is between these two curves and evolves into Γ(−∞) as s → −∞.

positions at s = 1. Each of these continuous images of R is projected to a closed

loop on the sphere; see Figure 9.4. Each “end” of the projections of Γ(−∞) and the

lower branch of Γv(1) lie at the north pole of the sphere. The closed loop of solution

positions at s = 1 lies between these projections. If the loop of positions is to evolve

into Γ(−∞) as s → −∞, the “ends” of this loop must also lie at the north pole.

Hence the loop of positions must intersect the N = 0 line on the sphere, shown in

Figure 9.4.

9.5 Numerical Procedure

We use the form of the fixed points of Eq. (9.63)-(9.65) to compute numerical solu-

tions of the equations. The method is as follows.

1. Choose τ0 > 0.

2. Set N0 = (1 + κ)(τ0 − τ−1

0

), the N -coordinate of the fixed point with τ -

coordinate equal to τ0.

3. Compute the unstable eigenvector at the fixed point (τ0, N0, 1).

145

4. For some small ε > 0, perturb (τ0, N0, 1) by ε times the unstable eigenvector:

(τ1

N1

(γ2)1

)=

(τ0

N0

1

)+ ε

( ((1 + κ)τ0 − 2τ3

0

)/4

(1 + κ + τ2

0

)

−(1 + κ)(2τ4

0 + (3− κ2)τ20 + (1 + κ)

)/4

(1 + κ + τ2

0

)

1

). (9.82)

5. Set s1 = 1+ 12ln (ε/b2). The formula in the step above gives an approximation

for the position, at s = s1, of the solution tracing out the unstable manifold

attached to the fixed point (τ0, N0, 1).

6. Integrate the equations numerically (for increasing s) until N = 0.

If N reaches zero at some s-value other than 1, adjust τ0 and repeat the above steps

until |N(1)| is near zero to within some tolerance.

Results for Rmax = 1 are pictured in Figures 9.5-9.10. In each case, the radial

stress is zero at the periphery and positive at the center, as expected from the form

of the vector field run in reverse; see Figure 9.2. The azimuthal stress, which agrees

with the radial stress at R = 0, is negative at the periphery, indicating that the

outer edge is under azimuthal compression.

Recall that for κ > 1, Γv(s) has two branches. The impact of the change in the

vector field can be seen by comparing Figures 9.6 and 9.8. As R → 0 (s → −∞), the

terminal τ -value is approached from below in Figure 9.6, indicating that the fixed

point is approached from the left. In Figure 9.8, on the other hand, the terminal

τ -value is approached from larger values of τ , indicating that the fixed point is

approached from the right.

9.6 γ1(R) ≡ 1, γ′2(R) < 0

Now we consider a growth pattern that allows us to use much of the previous section’s

analysis but produces quite different results. We again let the radial growth γ1

be constant, but we assign it a value γ0 greater than one. Further, we assume

γ2(R) ≤ γ0, with equality only at R = 0.

146

0.2 0.4 0.6 0.8 1.0 R

-1.5

-1.0

-0.5

0.5

stressΜ

Γ1HRL º 1, Γ2HRL = 1 + HRRmaxL2, Κ = 1

Figure 9.5: The radial stress (solid curve) drops to zero at the periphery, while theazimuthal stress (dashed curve) is negative at the periphery.

0.2 0.4 0.6 0.8 1.0 R

1.20

1.22

1.23

1.24

1.25

Τ = rRΓ1HRL º 1, Γ2HRL = 1 + HRRmaxL

2, Κ = 1

Figure 9.6: τ = r/R decreases monotonically in this configuration.

If γ1(R) ≡ γ0 > 1 and γ2(R) = γ0 − b (R/Rmax)2, the corresponding differential

147

0.2 0.4 0.6 0.8 1.0 R

-2.0

-1.5

-1.0

-0.5

0.5

1.0

stressΜ

Γ1HRL º 1, Γ2HRL = 1 + HRRmaxL2, Κ = 3

Figure 9.7: Changing κ from 1 to 3 produces little change in the radial (solid) andazimuthal (dashed) stress profiles.

0.2 0.4 0.6 0.8 1.0 R

1.17

1.18

1.19

1.20

Τ = rRΓ1HRL º 1, Γ2HRL = 1 + HRRmaxL

2, Κ = 3

Figure 9.8: Changing κ from 1 to 3 changes the τ = r/R profile dramatically.

equations are

dN

ds=

γ0

γ2

τ +κγ0

2γ2

(N − κτ +

√(N − κτ)2 + 4(1 + κ)γ2

2

)

− (1 + κ)γ0γ2

τ−N, (9.83)

ds=

γ0

2γ2

(N − κτ +

√(N − κτ)2 + 4(1 + κ)γ2

2

)− τ , (9.84)

dγ2

ds= 2 (γ2 − γ0) . (9.85)

148

0.2 0.4 0.6 0.8 1.0 R

-2

-1

1

stressΜ

Γ1HRL º 1, Γ2HRL = 1 + 2HRRmaxL2, Κ = 1

Figure 9.9: Changing b from 1 to 2 produces little change in the stress profiles.

0.2 0.4 0.6 0.8 1.0 R

1.40

1.45

1.50

Τ = rRΓ1HRL º 1, Γ2HRL = 1 + 2HRRmaxL

2, Κ = 1

Figure 9.10: Changing b from 1 to 2 produces little change in the τ = r/R profile.

The fixed points of this system form curves whose coordinates satisfy

γ2 = γ0, τ =N ±

√N2 + 4 (1 + κ)2 γ2

0

2 (1 + κ). (9.86)

τ > 0 on one curve; τ < 0 on the other. We focus on positive τ . At each such fixed

149

point, the linearization Df is

−(1+κ)

(τ2+γ2

0

)

τ2+(1+κ)γ20

τ2

τ2+(1+κ)γ20

τ(−τ2+(1+κ)γ2

0

)

γ0(

τ2+(1+κ)γ20

)

(1+κ)(

τ2+γ20

)(−(−1+κ)τ2+(1+κ)γ2

0

)

τ2(

τ2+(1+κ)γ20

) (−1+κ)τ2−(1+κ)γ20

τ2+(1+κ)γ20

−(1+κ)τ4+(−2−κ+κ2

)τ2γ2

0−(1+κ)2γ40

τγ0(

τ2+(1+κ)γ20

)

0 0 2

, (9.87)

which has eigenvalues 2, 0, and −2, regardless of κ, γ0, and τ . As in the previous

case, the only eigenvector with nonzero γ2-component is the unstable eigenvector,

which in this case is(

τ (−2τ 2 + (1 + κ)γ20)

4γ0 (τ 2 + (1 + κ)γ20)

,−2(1 + κ)τ 4 + (−3− 2κ + κ2) τ 2γ2

0 − (1 + κ)2γ40

4τ γ0 (τ 2 + (1 + κ)γ20)

, 1

)T

.

(9.88)

In this example, γ2(s) ↑ γ0 as s → −∞, which indicates that the unstable mani-

fold approaches the fixed point from the direction opposite to this eigenvector; see

Figure 9.11.

GvH1L

GhH1L

GH-¥L0.5 1.0 1.5 Τ

-1

1

2

3

4

N = nΜ

Figure 9.11: The vector field of the dynamical system run in reverse. In this case,the solutions tracing out the unstable manifolds approach the fixed points from theleft as s → −∞.

As in the previous section, we define horizontal and vertical isoclines Γh(s) and

Γv(s), respectively, and observe that they coincide with Γ(−∞), the curve of fixed

points, as s → −∞. Arguments similar to those in the last section show the existence

of an unstable manifold that intersects the line N = 0 at s = 1. In this case, however,

150

it is possible for a solution tracing out an unstable manifold to satisfy N(1) = 0

as well as lims→−∞ N(s) < 0, which corresponds to compressive radial stress at the

center of the disc.

Using a numerical approach nearly identical to that of the last section (with

negative ε in step 4), we found results with negative radial stress throughout the

body and azimuthal stress that changes sign within the body. In these examples,

γ0 ≥ γ2(R) ≥ 1, with γ2(0) = γ0 and γ2(Rmax) = 1.

0.2 0.4 0.6 0.8 1.0 R

-0.5

0.5

1.0

stressΜ

Γ1HRL º 1.5, Γ2HRL = 1.5 - 0.5HRRmaxL2, Κ = 1

Figure 9.12: The radial stress (solid line) is negative throughout the body, while theazimuthal stress (dashed line) changes sign, leaving it positive at the periphery.

This type of growth tensor induces negative radial stress throughout the body;

it is largest at R = 0 and decays monotonically until it is zero at the periphery. At

R = 0, the azimuthal stress agrees with the radial stress. It grows monotonically

in R and is positive at the periphery. The presence of negative stresses through so

much of the body suggests that these configurations may be prone to buckling.

151

0.2 0.4 0.6 0.8 1.0 R

1.362

1.364

1.366

1.368

1.370

1.372

Τ

Γ1HRL º 1.5, Γ2HRL = 1.5 - 0.5HRRmaxL2, Κ = 1

Figure 9.13: τ = r/R is monotone decreasing with R, but no change in concavity ofthe graph is apparent.

0.4 0.6 0.8 1.0 R

-0.2

-0.1

0.1

0.2

0.3

0.4

0.5

stressΜ

Γ1HRL º 1.5, Γ2HRL = 1.5 - 0.5HRRmaxL2, Κ = 3

Figure 9.14: Changing κ from 1 to 3 changes the magnitudes of the radial (solid)and azimuthal (dashed) stresses.

152

0.4 0.6 0.8 1.0 R

1.155

1.160

1.165

1.170

Τ

Γ1HRL º 1.5, Γ2HRL = 1.5 - 0.5HRRmaxL2, Κ = 3

Figure 9.15: Unlike what was found in the γ′2 > 0 case, changing κ from 1 to 3 haslittle impact on the form of τ = r/R.

0.2 0.4 0.6 0.8 1.0 R

-1.0

-0.5

0.5

1.0

1.5

2.0

2.5

stressΜ

Γ1HRL º 2, Γ2HRL = 2 - HRRmaxL2, Κ = 1

Figure 9.16: Keeping κ = 1 but doubling the growth increment approximatelydoubles the radial (solid) and azimuthal (dashed) stresses.

153

0.2 0.4 0.6 0.8 1.0 R

1.71

1.72

1.73

1.74

Τ

Γ1HRL º 2, Γ2HRL = 2 - HRRmaxL2, Κ = 1

Figure 9.17: Keeping κ = 1 but doubling the growth increment approximatelydoubles the expansion.

154

CHAPTER 10

AXISYMMETRIC BUCKLING OF THE KIRCHHOFF PLATE

10.1 Kirchhoff Constraints

In a Kirchhoff plate, the final configuration is assumed to have the form

(x f)(R, Θ, Z) = r(R, Θ) + Z d(R, Θ), (10.1)

where the function r describes a two-dimensional “middle surface” of the plate, and

the director field d is a unit vector-valued function that satisfies

d =∂r

∂R× ∂r

∂Θ

/wwww∂r

∂R× ∂r

∂Θ

wwww . (10.2)

Thus d can be viewed as a function of r and its partial derivatives. The three-

dimensional Kirchhoff plate consists of a two-dimensional surface from which one-

dimensional material fibers sprout in a perpendicular direction. See Figure 10.1.

The tangent vector to this class of deformations was found in Eq. (8.6):

Mx =

Mr + Z

Mr,R×r,Θ +r,R×M

r,Θ‖r,R×r,Θ ‖ , (10.3)

whereMr is an as-regular-as-needed function of R and Θ.

We could derive the vector equation of equilibrium by considering the weak form

of the Method of Virtual Work with this virtual displacement, but there is an easier

but less direct method. The director field must satisfy the (local) constraints

d · d = 1 and d · r,α = 0, α = R, Θ. (10.4)

Encoding this into the expression P act : ∂Mx/∂X is complicated, even in the axisym-

metric case. If we introduce Lagrange multipliers λ, βR, and βΘ, we can consider

155

instead the integral of the following:

τ k ·(

Mr + Z

Md

),k − λd ·

Md + βα

(Md · r,α + d · M

r,α

)

= τ k · Mr,k + βαd · r,α + τ k ·

(Z

Md

),k + (βαr,α − λd) ·

Md

= (τα + βαd) · Mr,α + τα · Z

Md,α +

(τZ + βαr,α − λd

) ·Md, (10.5)

where k = R, Θ, Z and α = R, Θ. Thanks to the Lagrange multipliers, we can

treatMr and

Md as though they were independent virtual displacements. We will be

considering a plate with no body forces and no applied tractions, so the weak form

of the equations of equilibrium will be

(XΨ)(B)

(τα + βαd) · M

r,α + τα · ZMd,α +

(τZ + βαr,α − λd

) ·Md

dV = 0

(10.6)

for all admissible virtual displacementsMr.

Note that

τR ·(

Mr + Z

Md

),R =

(P act ·ER

) · ∂

∂R

(Mr + Z

Md

)

= (P act · h1) · ∂

∂R

(Mr + Z

Md

)(10.7)

τΘ ·(

Mr + Z

Md

),Θ =

(P act ·EΘ

) · ∂

∂Θ

(Mr + Z

Md

)

=

(P act · h2

R

)· ∂

∂Θ

(Mr + Z

Md

). (10.8)

10.2 Kinematics of an Axisymmetric Kirchhoff Plate

We consider axisymmetric deformations of an axisymmetric Kirchhoff plate. The

reference configuration has the form

X(R, Θ, Z) = R cos(Θ) i + R sin(Θ) j + Z k = R h1(Θ) + Z k, (10.9)

where

h1(Θ) = cos(Θ) i + sin(Θ) j. (10.10)

156

h1

d

k

a

r

Ζ

Φ

Φ

Figure 10.1: A radial slice of an axisymmetric Kirchhoff plate. ζ is the height of apoint of the middle surface, and a material point’s height above the middle surfaceis measured along the unit vector d, which is perpendicular to the middle surface.

As a function of (R, Θ, Z), the final configuration will have the form

(x f)(R, Θ, Z) = r(R) h1(Θ) + ζ(R) k︸ ︷︷ ︸r(R,Θ)

+Z d(R, Θ). (10.11)

Here r is the cylindrical radius of a point in the middle surface, and ζ is the height

of a point in the middle surface. See Figure 10.1.

An outward-pointing radial tangent vector to the middle surface is

d

dR(r(R) h1(Θ) + ζ(R) k) = r′(R) h1(Θ) + ζ ′(R) k. (10.12)

We define a to be the corresponding unit vector:

a =r′ h1 + ζ ′ k√(r′)2 + (ζ ′)2

= cos φ h1 + sin φ k, (10.13)

where φ is the angle formed by a and the radial unit vector h1:

φ = arctan

(ζ ′

r′

). (10.14)

157

The director d is the “upward”-pointing unit vector orthogonal to a:

d = − sin φ h1 + cos φ k. (10.15)

The azimuthal unit vector h2 is the third member of the local orthonormal basis.

To construct the deformation gradient, we appeal to the relation

∂ξi(χ X)(ξ1, ξ2, ξ3) = F ·Ei, (10.16)

where F and the Ei are evaluated at X(ξ1, ξ2, ξ3).

∂R(r(R) h1(Θ) + ζ(R) k + Z d(R, Θ)) = r′(R) h1(Θ) + ζ ′(R) k + Z

∂d

∂R(R, Θ)

= r′ h1 + ζ ′ k − Zφ′ (cos φ h1 + sin φ k)

= r′ h1 + ζ ′ k − Zφ′ a. (10.17)

∂Θ(r(R) h1(Θ) + ζ(R) k + Z d(R, Θ)) = r(R) h2(Θ) + Z

∂d

∂Θ(R, Θ)

= r h2 − Z sin φ∂h1

∂Θ

= (r − Z sin φ) h2. (10.18)

∂Z(r(R) h1(Θ) + ζ(R) k + Z d(R, Θ)) = d(R, Θ) (10.19)

These results show that

F = (r′ h1 + ζ ′ k − Zφ′ a)⊗ER + (r − Z sin φ) h2 ⊗EΘ + d⊗EZ

= (r′ h1 + ζ ′ k − Zφ′ a)⊗ h1 +

(r − Z sin φ

R

)h2 ⊗ h2 + d⊗ k, (10.20)

where it must be understood that the vectors on the right-hand sides of the tensor

products are anchored at the reference point, while those on the left-hand sides are

anchored at the deformed point.

We will find it convenient to use the orthonormal basis a, h2,d in the deformed

configuration, so we will re-write the left-hand side of the first tensor product in F .

158

Note that a

d

=

cos φ sin φ

− sin φ cos φ

h1

k

=⇒

h1

k

=

cos φ − sin φ

sin φ cos φ

a

d

.

(10.21)

As a result,

r′ h1 + ζ ′ k = r′ (cos φ a− sin φ d) + ζ ′ (sin φ a + cos φ d)

= (r′ cos φ + ζ ′ sin φ) a + (ζ ′ cos φ− r′ sin φ) d. (10.22)

By definition of φ, though, ζ ′ = r′ tan φ, so

r′ cos φ + ζ ′ sin φ = r′ (cos φ + tan φ sin φ)

= r′(

cos φ +sin2 φ

cos φ

)

= r′ sec φ, (10.23)

ζ ′ cos φ− r′ sin φ = r′ (tan φ cos φ− sin φ) = 0. (10.24)

The deformation gradient can thus be written as

F = (r′ sec φ− Zφ′) a⊗ h1 +

(r − Z sin φ

R

)h2 ⊗ h2 + d⊗ k. (10.25)

We will need the following expression for later computations:

F−T = (r′ sec φ− Zφ′)−1a⊗ h1 +

(r − Z sin φ

R

)−1

h2 ⊗ h2 + d⊗ k. (10.26)

We will use incompatible growth tensors of the form presented in Eq. (4.6):

G = γ1 Er ⊗ER +γ2

γ1

Eθ ⊗EΘ + Ez ⊗EZ

= γ1 h1 ⊗ h1 +γ2

γ1

rh2 ⊗ h2

R+ k ⊗ k

= γ1 h1 ⊗ h1 + γ2 h2 ⊗ h2 + k ⊗ k, (r = γ1R) (10.27)

with γ1 and γ2 constant. It should be noted that in Eq. (10.27) the vectors on

the right-hand side of each tensor product are anchored at a point in the reference

159

configuration and each vector on the left-hand side of a tensor product is anchored

at the image under the incompatible growth discussed above.

With F and G so defined, the tensor describing the elastic response to growth

is

A = F ·G−1

=

(r′ sec φ− Zφ′

γ1

)a⊗ h1 +

(r − Z sin φ

γ2R

)h2 ⊗ h2 + d⊗ k. (10.28)

The vector on the right-hand side of each tensor product is anchored at a point

in the incompatibly grown state, and each vector on the left-hand side of a tensor

product is anchored at a point in the final configuration.

The most important tensor for the constitutive relation will be the Eulerian

tensor

A ·AT =

(r′ sec φ− Zφ′

γ1

)2

a⊗ a +

(r − Z sin φ

γ2R

)2

h2 ⊗ h2 + d⊗ d. (10.29)

It might be clearer that G maps from the tangent space at a point in a body in

E3 to the tangent space at a point in a body that cannot “fit” into E3, if G were

written as

G = γ1∂

∂r⊗ER +

γ2

γ1

∂θ⊗EΘ +

∂z⊗EZ , (10.30)

where ∂/∂r, ∂/∂θ, and ∂/∂z are tangent vectors at (r, θ, z) = f(R, Θ, Z) in (f Ψ)(B) ⊂ R3. Computing A = F ·G−1 would be just as simple, but computing A·AT

with this notation would require a detour into the form of invariant expressions for

transposes of tensors. Eqs. (10.27), (10.28), and (10.29) are sufficient for our goals.

10.3 Radial Virtual Displacement

First we consider∫

plate

(τα + βαd) · Mr,α dV

=

plate

(P act · h1 + βRd

) · ∂Mr

∂R+

(P act · h2

R+ βΘd

)· ∂

Mr

∂Θ

dV

=

plate

Mr

∂R· (P act · h1 + βRd

)+

∂Mr

∂Θ·(

P act · h2

R+ βΘd

)dV.

160

It will be helpful to consider βR and βΘ as components of a vector β:

β = βRER + βΘEΘ = βR h1 + βΘR h2. (10.31)

With β so defined, we can consider the two-point tensor field P act + d⊗ β.

Consider the planar portion ofMr · (P act + d⊗ β):

Mr · (P act + d⊗ β)planar =

(Mr · P act · h1 +

(Mr · d

)(β · h1)

)h1

+(Mr · P act · h2 +

(Mr · d

)(β · h2)

)h2. (10.32)

In polar coordinates, the divergence of a planar vector field is

Div (f h1 + g h2) =1

R

∂R(Rf) +

1

R

∂g

∂Θ, (10.33)

so the divergence ofMr · (P act + d⊗ β)planar is

Div(Mr · (P act + d⊗ β)planar

)

=1

R

∂R

(R

Mr · P act · h1 + R

(Mr · d

)(β · h1)

)

+1

R

∂Θ

(Mr · P act · h2 +

(Mr · d

)(β · h2)

)

=∂

Mr

∂R· (P act · h1 + βRd

)+

Mr · 1

R

∂R

(R P act · h1 + R βRd

)

+1

R

∂Mr

∂Θ· (P act · h2 + (β · h2) d) +

Mr · 1

R

∂Θ(P act · h2 + (β · h2) d)

=∂

Mr

∂R· (P act · h1 + βRd

)+

Mr · 1

R

∂R

(R P act · h1 + R βRd

)

+1

R

∂Mr

∂Θ· (P act · h2 + βΘR d

)+

Mr · 1

R

∂Θ

(P act · h2 + βΘR d

)

=∂

Mr

∂R· (P act · h1 + βRd

)+

Mr · 1

R

∂R

(R P act · h1 + R βRd

)

+∂

Mr

∂Θ·(

P act · h2

R+ βΘ d

)+

Mr · ∂

∂Θ

(P act · h2

R+ βΘ d

). (10.34)

161

TheMr-portion of Eq. (10.6) can be written as

plate

Mr

∂R· (P act · h1 + βRd

)+

∂Mr

∂Θ·(

P act · h2

R+ βΘd

)dV

=

plate

Div(Mr · (P act + d⊗ β)planar

)dV

−∫

plate

Mr ·

1

R

∂R

(R P act · h1 + R βRd

)

+∂

∂Θ

(P act · h2

R+ βΘd

)dV. (10.35)

By the Divergence Theorem in the plane, the integral of the first integrand can be

re-written

plate

Div(Mr · (P act + d⊗ β)planar

)dv

=

∫ H2

H1

slice

Div(Mr · (P act + d⊗ β)planar

)R dR dΘ

dZ

=

∫ H2

H1

∂(slice)

(Mr · (P act + d⊗ β)planar

)· ν d`

dZ

=

∂(slice)

Mr ·

∫ H2

H1

(P act + d⊗ β)planar · ν dZ

d`, (10.36)

where the “slice” mentioned is a planar slice of the reference configuration of the

plate, ∂(slice) is the curve that forms the planar boundary of the slice, ν is the

outward-pointing unit normal on ∂(slice), and d` is arclength measure on ∂(slice).

In order to balance the boundary term in the weak form of the Principle of

Virtual Work, we must specify the value of

∫ H2

H1

(P act + d⊗ β)planar · ν dZ =

∫ H2

H1

(P act + d⊗ β) · h1 dZ (10.37)

on the boundary of the middle surface.

Now that boundary terms are balanced, we have

plate

Mr ·

1

R

∂R

(R P act · h1 + R βRd

)+

∂Θ

(P act · h2

R+ βΘd

)dV = 0

(10.38)

162

for all as-smooth-as-needed functionMr of R and Θ. Since

plate

Mr ·

1

R

∂R

(R P act · h1 + R βRd

)+

∂Θ

(P act · h2

R+ βΘd

)dv

=

slice

∫ H2

H1

1

R

∂R

(R P act · h1 + R βRd

)

+∂

∂Θ

(P act · h2

R+ βΘd

)dZ · M

r R dR dΘ

= 0, (10.39)

for such a large class ofMr, we conclude that the integral in Z is identically zero:

∫ H2

H1

1

R

∂R

(R P act · h1 + R βRd

)+

∂Θ

(P act · h2

R+ βΘd

)dZ = 0.

(10.40)

10.4 Director Virtual Displacement

We apply the same procedure to the variation inMd. The result will be a weak

version of the balance of angular momentum in the constrained plate with zero

applied moments.

plate

τα · Z

Md,α +

(τZ + βαr,α−λd

) ·Md

dV

=

plate

τR · Z ∂

Md

∂R+ τΘ · Z ∂

Md

∂Θ+

(τZ + βαr,α−λd

) ·Md

dV

=

plate

Z Div

(Md · P act

)

planar

dV

+

plate

Md ·

(τZ + βαr,α−λd

)− Z1

R

∂R(R P act · h1)− Z

∂Θ

(P act · h2

R

)dV

=

∂(slice)

Md ·

(∫ H2

H1

Z P act dZ

)· ν d`

+

slice

Md ·

∫ H2

H1

τZ + βαr,α−λd− Z

1

R

∂R(R P act · h1)

− Z∂

∂Θ

(P act · h2

R

)dZ R dR dθ (10.41)

163

As before, ν = h1 is the outward-pointing unit normal to the planar boundary

∂(slice) of the slice in the reference configuration, and d` is arclength measure on

∂(slice).

To balance the boundary terms, we must set values of the moment∫ H2

H1

Z (P act · h1) dZ (10.42)

on the planar boundary of the middle surface.

We have found that∫

slice

Md ·

∫ H2

H1

τZ + βαr,α−λd− Z

1

R

∂R(R P act · h1)

− Z∂

∂Θ

(P act · h2

R

)dZ R dR dθ = 0 (10.43)

for all as-smooth-as-needed functionsMd of R and Θ. We conclude that

∫ H2

H1

τZ + βαr,α−λd− Z

1

R

∂R(R P act · h1)− Z

∂Θ

(P act · h2

R

)dZ = 0.

(10.44)

10.5 Addressing Lagrange Multipliers

We are not yet ready to introduce the constitutive relation to close the system of

equations. The quantities λ, βR, and βΘ are not constitutively defined. However,

there is enough information in the equations above to get around this complication.

First we claim that if we take the cross-product with d, the term λd will vanish.

By the form of G we are assuming and the form of F−T found above, we can be

sure that F−T ·k = d. If we use an isotropic hyperelastic constitutive relation, then

the form of A = F ·G−1 in such a relation will guarantee that T · d points in the

direction d. As a result,∫ H2

H1

τZdZ =

∫ H2

H1

(P act · k) dZ

=

∫ H2

H1

(det F )(T · F−T · k)

dZ

=

∫ H2

H1

(det F ) (T · d) dZ ∝ d, (10.45)

164

which implies

d×∫ H2

H1

τZdZ = 0, (10.46)

the elimination we sought.

When we take the cross-product of d with Eq. (10.44), we have

d× 1

R

∂R

(R

∫ H2

H1

Z (P act · h1) dZ

)

+ d× ∂

∂Θ

∫ H2

H1

Z

(P act · h2

R

)dZ − d× r,α

∫ H2

H1

βαdZ = 0. (10.47)

We will consider all these terms in detail to find expressions for βR and βΘ in terms

of kinematic and constitutively-defined quantities.

By the symmetry of the deformation, the vector field P act ·h1 has zero projection

onto h2. It is determined solely by its projections onto h1 and k:

(P act · h1) = (h1 · P act · h1) h1 + (k · P act · h1) k (10.48)

Since h1 and k are R-independent,

1

R

∂RR (P act · h1) = h1

1

R

∂RR (h1 · P act · h1) + k

1

R

∂RR (k · P act · h1) .

(10.49)

In particular,

h1 · 1

R

∂RR (P act · h1) =

1

R

∂RR (h1 · P act · h1) , (10.50)

k · 1

R

∂RR (P act · h1) =

1

R

∂RR (k · P act · h1) . (10.51)

165

We have

d× 1

R

∂R

(R

∫ H2

H1

Z (P act · h1) dZ

)

= (cos φ k − sin φ h1)

×

h11

R

∂R

(R

∫ H2

H1

Z (h1 · P act · h1) dZ

)

+ k1

R

∂R

(R

∫ H2

H1

Z (k · P act · h1) dZ

)

= h2

cos φ

R

∂R

(R

∫ H2

H1

Z (h1 · P act · h1) dZ

)

+sin φ

R

∂R

(R

∫ H2

H1

Z (k · P act · h1) dZ

). (10.52)

The symmetry of the deformation also requires that P act · h2 point along h2:

(P act · h2) = (h2 · P act · h2) h2. (10.53)

Further, the coefficient (h2 · P act · h2) is Θ-independent, so the Θ-derivative of

(P act · h2) is

∂Θ(P act · h2) =

∂Θ(h2 · P act · h2) h2

= (h2 · P act · h2)∂h2

∂Θ

= − (h2 · P act · h2) h1. (10.54)

The cross-product with d is

d× ∂

∂Θ

∫ H2

H1

Z

(P act · h2

R

)dZ

= (cos φ k − sin φ h1)×(−h1

∫ H2

H1

Z

(h2 · P act · h2

R

)dZ

)

= −h2cos φ

R

∫ H2

H1

Z (h2 · P act · h2) dZ. (10.55)

166

The remaining terms are

d× r,α

∫ H2

H1

βαdZ

= d×

∂r

∂R

∫ H2

H1

βRdZ +∂r

∂Θ

∫ H2

H1

βΘdZ

= (cos φ k − sin φ h1)×

(r′ h1 + ζ ′ k)

∫ H2

H1

βRdZ + r h2

∫ H2

H1

βΘdZ

=

(r′ cos φ

∫ H2

H1

βRdZ

)h2 +

(ζ ′ sin φ

∫ H2

H1

βRdZ

)h2

−(

r cos φ

∫ H2

H1

βΘ

)h1 −

(r sin φ

∫ H2

H1

βΘ

)k. (10.56)

What we have shown is

cos φ

R

∂R

(R

∫ H2

H1

Z (h1 · P act · h1) dZ

)

+sin φ

R

∂R

(R

∫ H2

H1

Z (k · P act · h1) dZ

)

− cos φ

R

∫ H2

H1

Z (h2 · P act · h2) dZ

h2

= (r′ cos φ + ζ ′ sin φ)

(∫ H2

H1

βRdZ

)h2

− r (cos φ h1 − sin φ k)

∫ H2

H1

βΘdZ. (10.57)

Note that the left-hand side is a scalar multiple of h2, and that∫ H2

H1βΘdZ on the

right-hand side is the coefficient of a linear combination of h1 and k, both of which

are orthogonal to h2. Since this holds for each (R, Θ), we see that∫ H2

H1βΘdZ = 0.

Computing the h2-projection reveals

cos φ

R

∂R

(R

∫ H2

H1

Z (h1 · P act · h1) dZ

)

+sin φ

R

∂R

(R

∫ H2

H1

Z (k · P act · h1) dZ

)

− cos φ

R

∫ H2

H1

Z (h2 · P act · h2) dZ

= (r′ cos φ + ζ ′ sin φ)

(∫ H2

H1

βRdZ

). (10.58)

167

If we multiply by R and divide by cos φ, we have

∂R

(R

∫ H2

H1

Z (h1 · P act · h1) dZ

)

+ tan φ∂

∂R

(R

∫ H2

H1

Z (k · P act · h1) dZ

)

−∫ H2

H1

Z (h2 · P act · h2) dZ

= R (r′ + ζ ′ tan φ)

(∫ H2

H1

βRdZ

)

= R(r′ + r′ tan2 φ

) (∫ H2

H1

βRdZ

)

= Rr′ sec2 φ

(∫ H2

H1

βRdZ

). (10.59)

In the next section we will find another expression for the integral of βR,

and this equation will become a differential equation in terms of kinematical and

constitutively-defined quantities.

10.6 Problem As a Pair of ODEs

Since we have eliminated all differentiation with respect to Θ in the differential

equations, we can view them as a system of ordinary differential equations with

independent variable R.

Thanks to the discoveries of the last section, Eq. (10.40) can now be written as

1

R

d

dR

h1 R

∫ H2

H1

(h1 · P act · h1) dZ

+ k R

∫ H2

H1

(k · P act · h1) dZ + R d

∫ H2

H1

βRdZ

= h1

∫ H2

H1

(h2 · P act · h2

R

)dZ. (10.60)

The projection of this equation onto k is

1

R

d

dR

R

∫ H2

H1

(k · P act · h1) dZ + R cos φ

∫ H2

H1

βRdZ

= 0, (10.61)

168

so that

R

∫ H2

H1

(k · P act · h1) dZ + R cos φ

∫ H2

H1

βRdZ = constant. (10.62)

Recall that we must assign the value of the stress

∫ H2

H1

(P act · h1 + βRd

)dZ (10.63)

on the boundary (R = Rmax) of the middle surface. If this stress is zero, then in

particular its k-projection is zero. This implies that the constant on the right-hand

side of Eq. (10.62) is zero, and

∫ H2

H1

βRdZ = − sec φ

∫ H2

H1

(k · P act · h1) dZ. (10.64)

We will see later that the isotropic hyperelastic constitutive relation we choose

ensures that P act · h1 points in the direction a = cos φ h1 + sin φ k. As a result of

the direction of P act · h1,

P act · h1 = (a · P act · h1) a, (10.65)

so (h1 · P act · h1) and (k · P act · h1) satisfy

k · P act · h1 = (a · P act · h1) k · a = sin φ (a · P act · h1) , (10.66)

h1 · P act · h1 = (a · P act · h1) h1 · a = cos φ (a · P act · h1) , (10.67)

and

(k · P act · h1) = tan φ (h1 · P act · h1) . (10.68)

As a result,

∫ H2

H1

βRdZ = − sec φ tan φ

∫ H2

H1

(h1 · P act · h1) dZ. (10.69)

The h1-projection of Eq. (10.60) becomes

1

R

d

dR

R

∫ H2

H1

(h1 · P act · h1) dZ −R sin φ

∫ H2

H1

βRdZ

=

∫ H2

H1

(h2 · P act · h2

R

)dZ, (10.70)

169

or

d

dR

R sec2 φ

∫ H2

H1

(h1 · P act · h1) dZ

=

∫ H2

H1

(h2 · P act · h2) dZ (10.71)

We have one more scalar ordinary differential equation from the previous section:

d

dR

(R

∫ H2

H1

Z (h1 · P act · h1) dZ

)

+ tan φd

dR

(R

∫ H2

H1

Z (k · P act · h1) dZ

)

−∫ H2

H1

Z (h2 · P act · h2) dZ

= Rr′ sec2 φ

(∫ H2

H1

βRdZ

)

= −Rr′ tan φ sec3 φ

∫ H2

H1

(h1 · P act · h1) dZ, (10.72)

where we have used Eq. (10.69) to re-write the right-hand side. Note that

tan φd

dR

(R

∫ H2

H1

Z (k · P act · h1) dZ

)

= tan φd

dR

(R tan φ

∫ H2

H1

Z (h1 · P act · h1) dZ

)

= φ′ tan φ sec2 φR

∫ H2

H1

Z (h1 · P act · h1) dZ

+ tan2 φd

dR

(R

∫ H2

H1

Z (h1 · P act · h1) dZ

). (10.73)

Eq. (10.72) can be re-written as

(1 + tan2 φ)d

dR

(R

∫ H2

H1

Z (h1 · P act · h1) dZ

)

+ φ′ tan φ sec2 φR

∫ H2

H1

Z (h1 · P act · h1) dZ

=

∫ H2

H1

Z (h2 · P act · h2) dZ

− Rr′ tan φ sec3 φ

∫ H2

H1

(h1 · P act · h1) dZ. (10.74)

170

Since 1 + tan2 φ = sec2 φ = (cos φ)−2, we can re-write this as

d

dR

(R

∫ H2

H1

Z (h1 · P act · h1) dZ

)

= cos2 φ

∫ H2

H1

Z (h2 · P act · h2) dZ

− (Rφ′ + Rr′ sec φ) tan φ

∫ H2

H1

Z (h1 · P act · h1) dZ. (10.75)

In summary, the balance laws for linear momentum and angular momentum for

the plate are

d

dR

(R sec2 φ

∫ H2

H1

(h1 · P act · h1) dZ

)=

∫ H2

H1

(h2 · P act · h2) dZ, (10.76)

and

d

dR

(R

∫ H2

H1

Z (h1 · P act · h1) dZ

)

= cos2 φ

∫ H2

H1

Z (h2 · P act · h2) dZ

− (Rφ′ + Rr′ sec φ) tan φ

∫ H2

H1

Z (h1 · P act · h1) dZ. (10.77)

We have already ensured that∫ H2

H1

(k · P act · h1) dZ + cos φ

∫ H2

H1

βRdZ

= k ·∫ H2

H1

(P act · h1) dZ + d

∫ H2

H1

βRdZ

(10.78)

is identically zero. One of the boundary conditions is that the analogous h1-

projection is zero at the boundary of the middle surface:

h1 ·∫ H2

H1

(P act · h1) dZ + d

∫ H2

H1

βRdZ

=

∫ H2

H1

(h1 · P act · h1) dZ − sin φ

∫ H2

H1

βRdZ

=

∫ H2

H1

(h1 · P act · h1) dZ + sin φ sec φ

∫ H2

H1

(k · P act · h1) dZ

=(1 + tan2 φ

) ∫ H2

H1

(h1 · P act · h1) dZ

= sec2 φ

∫ H2

H1

(h1 · P act · h1) dZ. (10.79)

171

Hence the boundary condition for stress is

sec2 φ

∫ H2

H1

(h1 · P act · h1) dZ

∣∣∣∣R∈Rmin,Rmax

= 0, (10.80)

where we consider Rmax in every case and Rmin only in the case of an annulus.

The boundary condition for moment is

∫ H2

H1

Z (P act · h1) dZ

∣∣∣∣R∈Rmin,Rmax

= 0. (10.81)

We will find a scalar equation that ensures that the h1- and k-projections are both

zero.

10.7 Applying the Constitutive Relation

We use the isotropic hyperelastic constitutive relation seen in Section 6.7 and Chap-

ter 9:

W (λ1, λ2, λ3) = µ

(1

2

(λ2

1 + λ22 + λ3

3 − 3)− ln J

)+ λ (J − 1− ln J) . (10.82)

Following the method outlined in Section 6.7, we find that the active portion of the

Cauchy stress tensor in the presence of growth is

T act =1

J

(λ2

1 a⊗ a + λ22 h2 ⊗ h2 + λ2

3 d⊗ d)

+ (λJ − (λ + µ)) I), (10.83)

where the λi are the “pseudo-stretches”

λ1 =r′ sec φ− Zφ′

γ1

, λ2 =r − Z sin φ

γ2R, λ3 = 1, (10.84)

and J = det A = λ1λ2λ3.

The differential equations Eqs. (10.76) and (10.77) are written in terms of P act =

(det F ) T act · F−T , but we need only certain projections of this tensor. Note that

F−T · h1 = (r′ sec φ− Zφ′)−1a, (10.85)

F−T · h2 =

(r − Z sin φ

R

)−1

h2. (10.86)

172

We have

P act · h1 = (det F ) T act · F−T · h1

= Jγ1γ2 T act · (r′ sec φ− Zφ′)−1a

= γ1γ2

γ21

(r′ sec φ− Zφ′) +λ

γ1γ2

(r′ sec φ− Zφ′)

− λ + µ

r′ sec φ− Zφ′

)a, (10.87)

where

det F = (det A) (det G) = (J) (γ1γ2) . (10.88)

A similar calculation shows that

P act · h2 (10.89)

= γ1γ2

γ22

(r − Z sin φ

R

)+

λ

γ1γ2γ3

(r − Z sin φ

R

)− λ + µ

(r − Z sin φ) /R

)h2.

The projections pertinent for the differential equations are

h1 · P act · h1 (10.90)

= cos φ

(µγ2

γ1

(r′ sec φ− φ′Z) + λ

(r − Z sin φ

R

)− (λ + µ) γ1γ2

r′ sec φ− Zφ′

),

h2 · P act · h2 (10.91)

=µγ1

γ2

(r − Z sin φ

R

)+ λ (r′ sec φ− Zφ′)− (λ + µ) γ1γ2

(r − Z sin φ) /R.

We also assume that the middle surface occupies the true middle of the plate, so

that H1 = −H and H2 = H for some H > 0. The Z-integrals then have the forms∫ H

−H

(h1 · P act · h1) dZ (10.92)

= cos φ

2H

(µγ2

γ1

r′ sec φ + λr

R

)− 2 (λ + µ) γ1γ2

φ′arctanh

(Hφ′

r′ sec φ

)

∫ H

−H

(h2 · P act · h2) dZ (10.93)

= 2H

(µγ1

γ2

r

R+ λr′ sec φ

)− 2 (λ + µ) γ1γ2R

sin φarctanh

(H sin φ

r

)

173

∫ H

−H

Z (h1 · P act · h1) dZ (10.94)

= − cos φ

2

3H3

(µγ2

γ1

φ′ + λsin φ

R

)

+2 (λ + µ) γ1γ2

(φ′)2

(r′ sec φ arctanh

(Hφ′

r′ sec φ

)−Hφ′

)

∫ H

−H

Z (h2 · P act · h2) dZ (10.95)

= −2

3H3

(µγ1

γ2

sin φ

R+ λφ′

)

− 2 (λ + µ) γ1γ2R

sin2 φ

(r arctanh

(H sin φ

r

)−H sin φ

)

We get a further simplification if we divide by µ:

1

µ

∫ H

−H

(h1 · P act · h1) dZ (10.96)

= cos φ

2H

(γ2

γ1

r′ sec φ + κr

R

)− 2 (1 + κ) γ1γ2

φ′arctanh

(Hφ′

r′ sec φ

)

1

µ

∫ H

−H

(h2 · P act · h2) dZ (10.97)

= 2H

(γ1

γ2

r

R+ κr′ sec φ

)− 2 (1 + κ) γ1γ2R

sin φarctanh

(H sin φ

r

)

1

µ

∫ H

−H

Z (h1 · P act · h1) dZ (10.98)

= − cos φ

2

3H3

(γ2

γ1

φ′ + κsin φ

R

)

+2 (1 + κ) γ1γ2

(φ′)2

(r′ sec φ arctanh

(Hφ′

r′ sec φ

)−Hφ′

)

174

1

µ

∫ H

−H

Z (h2 · P act · h2) dZ (10.99)

= −2

3H3

(γ1

γ2

sin φ

R+ κφ′

)

− 2 (1 + κ) γ1γ2R

sin2 φ

(r arctanh

(H sin φ

r

)−H sin φ

),

where κ = λ/µ. The differential equations have the same form if each stress and

moment is replaced by the same quantity divided by µ.

10.8 The Flat Plate

If the plate is un-buckled, then there is no moment, P act ·h1 has no Z-dependence,

and the radial Piola-Kirchhoff stress is radial only, i.e. it has no vertical component.

In the case of constant γ1 and γ2, the equations for the un-buckled plate can be

converted into a pair of autonomous ordinary differential equations, as detailed in

Antman and Negron-Marrero (1987). If we set

R = Rmaxes−1, (10.100)

τ =r

R, (10.101)

N =γ2

γ1

r′ + κr

R− (1 + κ)γ1γ2

r′. (10.102)

When τ and N are viewed as functions of the independent variable s, they satisfy

ds=

γ1

2γ2

(N − κτ +

√(N − κτ)2 + 4(1 + κ)γ2

2

)− τ, (10.103)

dN

ds=

γ1

γ2

τ +κγ1

2γ2

(N − κτ +

√(N − κτ)2 + 4(1 + κ)γ2

2

)

−(1 + κ)γ1γ2

τ−N. (10.104)

We consider an annulus that has zero radial stress at its inner and outer faces.

We can find numerically an initial value for τ such that the boundary conditions

n(s0) = 0 and n(s1) = 0. The functions n(s) and τ(s) correspond to functions N(R)

175

and T (R) via

N(R) = n

(1 + ln

(R

Rmax

)), T (R) = τ

(1 + ln

(R

Rmax

)). (10.105)

This correspondence also gives us a function r(R):

r(R) = RT (R) = Rτ

(1 + ln

(R

Rmax

)), (10.106)

which also provides a function r′(R).

10.9 Differential Equations for Bifurcation

Finding numerical solutions of the boundary-value problem for the buckled plate

has proved a daunting task. To demonstrate the existence of buckled solutions, we

consider perturbation about the flat configuration.

Eqs. (10.76) and (10.77) for the buckled plate, i.e. with stresses and moments

given by Eqs. (10.93)-(10.96), can be expressed in the form

1 0 0 0

0 1 0 0

0 0 a33 a34

0 0 a43 a44

︸ ︷︷ ︸A(R,r,φ,r′,φ′)

d

dR

r

φ

r′

φ′

=

r′

φ′

f3

f4

︸ ︷︷ ︸f(R,r,φ,r′,φ′)

, (10.107)

with

a33 = 2HR

(γ2 sec2 φ

γ1

+γ1γ2(1 + κ)

(r′)2 −H2 (φ′)2 cos2 φ

), (10.108)

a34 = 2R

γ1γ2(1 + κ) sec φ

(φ′)2 arctanh

(Hφ′ cos φ

r′

)

− Hγ1γ2(1 + κ)

r′φ′(1− H2(φ′)2 cos2 φ

(r′)2

) (10.109)

176

a43 = 2R

−γ1γ2(1 + κ)

(φ′)2 arctanh

(Hφ′ cos φ

r′

)

+Hγ1γ2(1 + κ) cos φ

r′φ′(1− H2(φ′)2 cos2 φ

(r′)2

) (10.110)

a44 = 2R

−H3γ2 cos φ

3γ1

+Hγ1γ2(1 + κ) cos φ

(φ′)2

+2γ1γ2(1 + κ) cos φ

(arctanh

(Hφ′ cos φ

r′

)r′ sec φ−Hφ′

)

(φ′)3

− Hγ1γ2(1 + κ) cos φ

(φ′)2(1− H2(φ′)2 cos2 φ

(r′)2

) , (10.111)

f3 = 2Rγ1γ2(1 + κ)arctanh

(H sin φ

r

)csc φ− 2Hκr sec φ

R

+ 2Hκr′ sec φ− 2H

(γ1r

Rγ2

+ κr′ sec φ

)

+ sec φ

2H

(κr

R+

γ2r′ sec φ

γ1

)−

2γ1γ2(1 + κ)arctanh(

Hφ′ cos φr′

)

φ

+2HRγ2r

′φ′ sec2 φ tan φ

γ1

+ R sec φ tan φ

2H

(κr

R+

γ2r′ sec φ

γ1

)−

2γ1γ2(1 + κ)arctanh(

Hφ′ cos φr′

)

φ′

φ′

+2HRγ1γ2(1 + κ)φ′ tan φ

r′(1− H2(φ′)2 cos2 φ

(r′)2

) (10.112)

177

f4 =2H3κ cos φ sin φ

3R

+ Rr′ tan φ

2H

(κr

R+

γ2r′ sec φ

γ1

)−

2γ1γ2(1 + κ)arctanh(

Hφ′ cos φr′

)

φ′

−2Rγ1γ2(1 + κ)arctanh

(Hφ′ cos φ

r′

)r′ tan φ

φ′

− 2

3H3κφ′ cos2 φ +

2HRγ1γ2(1 + κ) sin φ

1− H2(φ′)2 cos2 φ

(r′)2

− cos φ

2γ1γ2(1 + κ)(arctanh

(Hφ′ cos φ

r′

)r′ sec φ−Hφ′

)

(φ′)2

+2

3H3

(κ sin φ

R+

γ2φ′

γ1

)

− cos φ

−2Rγ1γ2(1 + κ) csc2 φ

(rarctanh

(H sin φ

r

)−H sin φ

)

− 2

3H3

(γ1 sin φ

Rγ2

+ κφ′)

(10.113)

The matrix A is invertible symbolically, so the equations can be expressed in

quasilinear form:

d

dR

r

φ

r′

φ′

= A−1f . (10.114)

We consider expressions r = r0 +εr1 and φ = εφ1, where r0 is the radius function

in Eq. (10.106), i.e. the solution of the boundary-value problem for the un-buckled

178

plate. Inserting these into Eq. (10.114), we have

d

dR

r0

0

r′0

0

+ εd

dR

r1

φ1

r′1

φ′1

= A−1f∣∣

r = r0

r′ = r′0φ = 0

φ′ = 0

+ ε∂A−1f

∂(r, φ, r′, φ)

∣∣∣∣ r = r0

r′ = r′0φ = 0

φ′ = 0

·

r1

φ1

r′1

φ′1

+O (ε2

). (10.115)

Equating terms with equal powers of ε, we are left with a system of linear equations

for (r1, φ1, r′1, φ

′1):

d

dR

r1

φ1

r′1

φ′1

=∂A−1f

∂(r, φ, r′, φ)

∣∣∣∣ r = r0

r′ = r′0φ = 0

φ′ = 0

·

r1

φ1

r′1

φ′1

. (10.116)

179

10.10 Linear Boundary Conditions for Bifurcation

We employ first-order expansions of the scalar coefficient of the radial stress in

Eq. (10.93):

1

µ

∫ H

−H

(a · P act · h1) dZ

=1

µ

∫ H

−H

(a · P act · h1) dZ

∣∣∣∣r = r0, r′ = r′0φ = 0, φ′ = 0

+∂

(1µ

∫ H

−H(a · P act · h1) dZ

)

∂(r, φ, r′, φ′)

∣∣∣∣∣∣ r = r0, r′ = r′0φ = 0, φ′ = 0

· ε

r1

φ1

r′1

φ′1

+O (ε2

)

= 2H

(γ2

γ1

r′0 + κr0

R− (1 + κ)γ1γ2

r′0

)

+ ε

(2Hκ

Rr1 +

(2Hγ2

γ1

+2H(1 + κ)γ1γ2

(r′0)2

)r′1

)+O (

ε2). (10.117)

At R = R0 = Rmin and R = R1 = Rmax, the radial stress of the flat annulus is zero,

so for i = 0, 1,

1

µ

∫ H

−H

(a · P act · h1) dZ

∣∣∣∣R=Ri

= ε

(2Hκ

Rr1 +

(2Hγ2

γ1

+2H(1 + κ)γ1γ2

(r′0)2

)r′1

)∣∣∣∣R=Ri

+O (ε2

). (10.118)

This provides one linear boundary condition for R = R0 and one for R = R1:

κ

Rr1 +

(γ2

γ1

+(1 + κ)γ1γ2

(r′0)2

)r′1

∣∣∣∣R=Ri

= 0. (10.119)

180

We perform the same expansion procedure for the moment in Eq. (10.95):

1

µ

∫ H

−H

(a · P act · h1) Z dZ

=1

µ

∫ H

−H

(a · P act · h1) Z dZ

∣∣∣∣ r = r0

r′ = r′0φ = 0

φ′ = 0

+∂

(1µ

∫ H

−H(a · P act · h1) Z dZ

)

∂(r, φ, r′, φ′)

∣∣∣∣∣∣ r = r0

r′ = r′0φ = 0

φ′ = 0

· ε

r1

φ1

r′1

φ′1

+O (ε2

)

= 0 +2H3ε

3

Rφ1 +

(γ2

γ1

+(1 + κ)γ1γ2

(r′0)2

)φ′1

)+O (

ε2). (10.120)

Recall that the un-buckled plate has identically zero moment. This provides linear

boundary conditions at R = R0 and R = R1:

κ

Rφ1 +

(γ2

γ1

+(1 + κ)γ1γ2

(r′0)2

)φ′1

∣∣∣∣R=Ri

= 0. (10.121)

10.11 Evidence of Bifurcation

We now have a system of four linear ordinary differential equations and four lin-

ear boundary conditions. The linear boundary conditions at R = R0 have a two-

dimensional vector space of solutions. Considering that the boundary conditions at

R = R0 can be expressed as

κR

γ2

γ1+ (1+κ)γ1γ2

(r′0)20 0

0 0 κR

γ2

γ1+ (1+κ)γ1γ2

(r′0)2

r1

r′1

φ1

φ′1

∣∣∣∣∣∣∣∣∣∣∣R=R0

=

0

0

, (10.122)

181

we see that the two-dimensional solution space at R = R0 is spanned by, for example,

r1

r′1

φ1

φ′1

∣∣∣∣∣∣∣∣∣∣∣R=R0

=

γ2

γ1+ (1+κ)γ1γ2

(r′0)2

− κR

0

0

∣∣∣∣∣∣∣∣∣∣∣R=R0

(10.123)

and

r1

r′1

φ1

φ′1

∣∣∣∣∣∣∣∣∣∣∣R=R0

=

0

0

γ2

γ1+ (1+κ)γ1γ2

(r′0)2

− κR

∣∣∣∣∣∣∣∣∣∣∣R=R0

(10.124)

Let u and v be solutions of Eq. (10.116) with the initial (R = R0) conditions

above. Then u and v have the forms

u1

u2

u3

u4

=

r1

r′1

0

0

and

v1

v2

v3

v4

=

0

0

φ1

φ′1

. (10.125)

Every linear combination αu + βv is a solution of the differential equations that

satisfies the linear boundary conditions Eq. (10.119) and (10.121) at R = R0.

We seek a linear combination of this form that also satisfies the linear boundary

conditions at R = R1. That is, we want

κR

γ2

γ1+ (1+κ)γ1γ2

(r′0)20 0

0 0 κR

γ2

γ1+ (1+κ)γ1γ2

(r′0)2

αu1

αu2

βv3

βv4

∣∣∣∣∣∣∣∣∣∣∣R=R1

=

0

0

. (10.126)

We can seek the existence of such a linear combination by re-writing this condition

as

κRu1 +

(γ2

γ1+ (1+κ)γ1γ2

(r′0)2

)u2 0

0 κRv1 +

(γ2

γ1+ (1+κ)γ1γ2

(r′0)2

)v2

∣∣∣∣∣∣R=R1

α

β

=

0

0

(10.127)

182

If the matrix has zero determinant, then there is a non-trivial solution (α, β)T of

this pair of linear equations at R = R1.

We know that if γ1 = γ2, then the growth tensor is a true deformation gradient,

and there is no residual stress in the un-buckled annulus. We fix a value for γ2 and

let γ1 vary until the determinant of the matrix in Eq. (10.127) is zero. In truth, we

seek the first γ1-value at which either of the non-zero entries in the matrix is zero.

10.12 Numerical Results

In all numerical results found so far, the first term to drop to zero is the (2, 2)-entry

of the matrix in Eq. (10.127). This means that we can choose a perturbation of

the moment alone. Further, the (2, 2)-entry reaches zero at some γ1 > γ2, which

indicates buckled solutions for the case in which incompatible growth corresponds

to azimuthal contraction.

In the following example, we set γ2 = 1.01, H = 0.01, R0 = 0.01, R1 = 1, and

κ = 1. As seen in Figures 10.2 and 10.3, as functions of γ1, the stress perturbation

(dependent on r1 alone) at R = R1 has no zeros, while the moment perturbation

(dependent on φ1 alone) at R = R1 has multiple zeros, all greater than γ2. The first

four zeros are found at γ1 = 1.01086, 1.01348, 1.01793, and 1.0242.

1.012 1.014 1.016 1.018 1.020 1.022 1.024Γ1

405

410

415

420

Κ r1HR1L

R1+Γ2

Γ1+HΚ + 1L Γ1 Γ2

r0HR1L2

r1¢HR1L

Figure 10.2: The stress perturbation (dependent on r1) at R = R1 has no zeros.

183

1.012 1.014 1.016 1.018 1.020 1.022 1.024Γ1

-50

50

100

150

Κ Φ1HR1L

R1+Γ2

Γ1+HΚ + 1L Γ1 Γ2

r0¢HR1L

2Φ1¢HR1L

Figure 10.3: The moment perturbation (dependent on φ1) at R = R1 has multiplezeros.

0.02 0.04 0.06 0.08 0.10H

1.02

1.04

1.06

1.08

1.10

HΓ1Lfirst critical value

Figure 10.4: The first bifurcation value of γ1, as a function of half-thickness H

10.13 Numerical Example: Buckling Near a Bifurcation Point

Iterating the numerical scheme described in Section 9.5 of Chapter 9, we compute

solutions of Eqs. (9.57)-(9.58) for the flat annulus until a flat solution satisfying the

boundary conditions is found. We call this solution r0. We use r0 in the matrix on

the right-hand side of Eq. (10.116). We also use r0(Rmin) and r′0(Rmin) to choose

values at Rmin for the perturbations r1 and r′1. Since we want r1 and r′1 to satisfy

184

0.02 0.04 0.06 0.08 0.10H

1.1

1.2

1.3

1.4

1.5

1.6

1.7HΓ1Lcritical value

Figure 10.5: The first three critical values of γ1, as functions of H

the boundary condition in Eq. (10.119), we set

r1(Rmin) =γ2

γ1

+(1 + κ)γ1γ2

(r′0(Rmin))2 , r′1(Rmin) = − κ

Rmin

. (10.128)

We also set

φ1(Rmin) = 0, φ′1(Rmin) = 0. (10.129)

With these as initial conditions, we integrate Eq. (10.116).

Fixing r0(Rmin) and r′0(Rmin), we seek numerically for values for φ and φ′ so that

the nonlinear boundary conditions

2H

(γ2

γ1

r′ sec φ + κr

Rmin

)− 2(1 + κ)γ1γ2

φ′arctanh

(Hφ′

r′ sec φ

)= 0 (10.130)

and

2H3

3

(γ2

γ1

φ′ + κsin φ

Rmin

)− 2(1 + κ)γ1γ2

(φ′)2

(r′ sec φarctanh

(Hφ′

r′ sec φ

)−Hφ′

)= 0

(10.131)

are satisfied with r = r0(Rmin), r′ = r′0(Rmin), φ = φ and φ′ = φ′. With these

initial conditions, we integrate Eqs. (10.76)-(10.77) numerically. We find very good

numerical satisfaction of the boundary condition at R = Rmax.

The numerical results for the two methods are compared. The numerical results

for the linear perturbation are scaled to agree as much as possible with the results of

185

integrating Eqs. (10.76)-(10.77). The numerical results of both methods are shown

together below.

In this example, we use parameters

γ2 = 1.01, γ2 =56026090

55424431≈ 1.01086, H = 0.01, Rmin = 0.01, Rmax = 1, κ = 1.

(10.132)

0.2 0.4 0.6 0.8 1.0 R

1.´10-6

1.5´10-6

2.´10-6

2.5´10-6

Φ

Figure 10.6: The buckling angle φ found from the two methods. Solid: nonlinear.Dashed: linear perturbation.

0.2 0.4 0.6 0.8 1.0 R

5.´10-7

1.´10-6

1.5´10-6

2.´10-6

Ζ

Figure 10.7: The middle surface height ζ found from the two methods. Solid:nonlinear. Dashed: linear perturbation.

186

0.2 0.4 0.6 0.8 1.0 R

-0.00006

-0.00005

-0.00004

-0.00003

-0.00002

-0.00001

1Μà-H

H Ha×Pact×h1L âZ

Figure 10.8: Thickness-integrated radial stress found from the two methods. Solid:nonlinear. Dashed: linear perturbation.

0.2 0.4 0.6 0.8 1.0 R

5.´10-12

1.´10-11

1.5´10-11

2.´10-11

1Μà-H

H Ha×Pact×h1LZ âZ

Figure 10.9: Thickness-integrated radial moment found from the two methods.Solid: nonlinear. Dashed: linear perturbation.

By the nature of the bifurcation analysis, the buckling found at each bifurcation

value is small. We present in the next section an example of much larger buckling.

10.14 Numerical Example: Buckling Due to Immersion-Preventing Growth

We employ the growth tensor

G = 1.9 h1 ⊗ h1 +(1.9− 0.9R2

)h2 ⊗ h2 + k ⊗ k. (10.133)

187

This leads to a new metric tensor on Ψ(B) whose associated Riemann-Christoffel

curvature tensor is nonzero, so this growth precludes isometric immersion of the

grown body.

We fix the parameter values

Rmin = 0.2, Rmax = 0.709, H = 0.01, κ = 1, (10.134)

and solve Eqs. (10.76)-(10.77) numerically. We search for initial conditions for which

the boundary conditions at R = Rmin and R = Rmax are satisfied simultaneously.

The results are shown in Figures 10.10-10.13.

0.3 0.4 0.5 0.6 0.7 R

-0.014

-0.012

-0.010

-0.008

-0.006

-0.004

-0.002

Hintegrated stressLΜ

Figure 10.10: Thickness-integrated radial (solid) and azimuthal (dashed) stresses

ζ(0.709) ≈ 0.064, which is 9% of the radius of the reference configuration of the

annulus. This buckling is visible in a plot of the middle surface. See Figure 10.13.

10.15 Conclusion

We have derived geometrically exact equations for an axisymmetric flat and buckled

Kirchhoff plates that has undergone axisymmetric incompatible growth. Through

perturbation about the final (grown and elastically distorted) flat configuration, we

have shown that the equations for the buckled configuration have solutions free of

applied tractions. Further, we have found in the model that an increase of the

plate’s thickness makes the plate more resistant to buckling. We have demonstrated

188

0.3 0.4 0.5 0.6 0.7 R

6.0

6.2

6.4

6.6

6.8

7.0

Φ

Figure 10.11: Buckling angle φ

0.3 0.4 0.5 0.6 0.7 R

0.01

0.02

0.03

0.04

0.05

0.06

Ζ

Figure 10.12: Middle surface height ζ

that incompatible growth can induce buckling with an amplitude several times the

thickness of the plate.

189

Figure 10.13: Top view of middle surface

190

CHAPTER 11

CONCLUSION

Our presentation of finite elasticity has a special emphasis on the functional form

of deformation in order to show that the standard kinematical conventions of fi-

nite deformation can be adapted to describe the geometric impact of incompatible

growth. We have demonstrated that importing growth into finite hyperelasticity via

the multiplicative decomposition F = A · G permits the use of the geometrically

exact theory of shells with just a change in the form of the first Piola-Kirchhoff

stress tensor. We have found examples in which the elastic response to incompat-

ible growth distorts a non-immersible grown body into a Euclidean body without

the assistance of applied tractions. In another example, the elastic response to

incompatible growth induced buckling of a plate free of applied tractions.

The next step in this enterprise should be numerical solution of Eqs. (10.76) and

(10.77) in the regions between critical values of γ1. Though we have demonstrated

the presence of buckled solutions, the size such buckling can reach is of interest.

To test the value of this approach against that of Efrati et al. (2007a), etc., this

theory should be applied Kirchhoff discs that are allowed to buckle azimuthally as

well as radially. The kinematics of such a project will be more complicated, but still

within the realm presented in chapter of Antman (1995). The greater difficulty will

be posed by the partial differential equations that will arise.

The examples presented here have not included a shell, which differs from a

plate only in that its middle surface is curved in the reference configuration. The

locomotive “gel caterpillar” mentioned in chapter 1 (Simonite (2009)), for example,

looks like a bent beam and would require a shell model. Except in cases of special

symmetry, models of shells will produce partial differential equations.

Even more advanced models for shell-like bodies have been suggested. For ex-

ample, DiCarlo et al. (2001) proposed a method for introducing thickness distension

191

into shells. Compared to the complexities coming from such extensions of the theory,

including incompatible growth in the fashion discussed here is hardly a hardship at

all.

Various theories for geometry-independent dynamics of growth have been pro-

posed (Epstein and Maugin (2000), DiCarlo and Quiligotti (2002), Lubarda and

Hoger (2002)). Thin-walled tubes dominate the explicit models of dynamical

growth (Lin and Taber (1995), Taber and Eggers (1996), Taber (1998a), Taber

and Humphrey (2001), Vandiver (2009)). One of the foundational difficulties of

the subject is the derivation of balance laws in residually stressed bodies, including

those that have already undergone incompatible growth and the subsequent elastic

response (Hoger (1997), Chen and Hoger (2000), Klarbring et al. (2007)).

We hope that our combination of incompatible growth and the geometrically

exact theory of shells provides a guide for the construction of models and for the

understanding of the interplay between three-dimensional Riemannian geometry and

elasticity.

192

APPENDIX A

PROOFS OF SOME RESULTS IN CHAPTER 2

Proposition A.0.1 The matrix [Eij] of dot-products Eij = Ei · Ej is the inverse

of the matrix [Eij] of dot-products Eij = Ei ·Ej.

Proof Recall that

det (abc) = a · (b× c) = c · (a× b) = b · (c× a) , (A.1)

which is non-zero if a, b, and c are linearly independent. In particular, the denom-

inator in the definition of Ei is independent of i. Let (ijk) and (`mn) be cyclic

permutations of (123). Then

Ei ·E` =Ej ×Ek

det (E1E2E3)· Em ×En

det (E1E2E3)

=(Ej ×Ek) · (Em ×En)

det [E1E2E3]2

=(Ej ·Em) (Ek ·En)− (Ej ·En) (Ek ·Em)

det (E1E2E3)2

=EjmEkn − EjnEkm

det (E1E2E3)2

=1

det (E1E2E3)2

∣∣∣∣∣∣Ejm Ejn

Ekm Ekn

∣∣∣∣∣∣. (A.2)

The 2×2 determinant in the last line is the i, j-entry of the adjugate matrix of [Eij].

The determinant det (E1E2E3) is related to the determinant of the matrix [Eij] as

193

follows.

det

E11 E12 E13

E21 E22 E23

E31 E32 E33

= det

ET1

ET2

ET3

(E1 E2 E3)

= det

ET1

ET2

ET3

det

(E1 E2 E3

)

= det(E1E2E3)2. (A.3)

We now have

Ei` =(adj[Ejm])i`

det[Ejm], (A.4)

where adj[Ejm] is the adjugate of the matrix [Ejm]. By Cramer’s Rule, this shows

that [Eij] is the inverse of the matrix [Eij] (Strang (1980), page 170).

Proposition A.0.2 Christoffel symbols of the second kind transform as

Γkji =

∂ξk

∂η`

∂ηm

∂ξj

∂ηn

∂ξiΓ`

mn +∂ξk

∂η`

∂2η`

∂ξj∂ξi. (A.5)

Proof

Γkji = Ek · ∂Ei

∂ξj

=

(∂ξk

∂η`E

`)· ∂ηm

∂ξj

∂ηm

(∂ηn

∂ξiEn

)

=∂ξk

∂η`

∂ηm

∂ξj

(∂ηn

∂ξiE

` · ∂En

∂ηm+ E

` · En∂

∂ηm

(∂ηn

∂ξi

))

=∂ξk

∂η`

∂ηm

∂ξj

∂ηn

∂ξiΓ`

mn +∂ξk

∂η`

∂ηm

∂ξjδ`n

∂ξp

∂ηm

∂ξp

(∂ηn

∂ξi

)

=∂ξk

∂η`

∂ηm

∂ξj

∂ηn

∂ξiΓ`

mn +∂ξk

∂η`δ`nδ

pj

∂2ηn

∂ξp∂ξi

=∂ξk

∂η`

∂ηm

∂ξj

∂ηn

∂ξiΓ`

mn +∂ξk

∂η`

∂2η`

∂ξj∂ξi(A.6)

194

Proposition A.0.3 The coefficients vi||j transform as the coefficients of a tensor

of type(11

).

Proof Recall that

vi||j =

∂vi

∂ξj+ Γi

jkvk. (A.7)

The first term transforms as

∂vi

∂ξj=

∂ηm

∂ξj

∂ηm

(vn ∂ξi

∂ηn

)

=∂ηm

∂ξj

∂vn

∂ηm

∂ξi

∂ηn+

∂ηm

∂ξjvn ∂2ξi

∂ηm∂ηn, (A.8)

while the Christoffel symbol of the second kind transforms as

Γijk =

∂ξi

∂η`

∂ηm

∂ξj

∂ηn

∂ξkΓ`

mn +∂ξi

∂η`

∂2η`

∂ξj∂ξk. (A.9)

Γijkv

k =

(∂ξi

∂η`

∂ηm

∂ξj

∂ηn

∂ξkΓ`

mn +∂ξi

∂η`

∂2η`

∂ξj∂ξk

)(∂ξk

∂ηpvp

)

=∂ξi

∂η`

∂ηm

∂ξj

∂ηn

∂ξk

∂ξk

∂ηpΓ`

mn vp +∂ξi

∂η`

∂2η`

∂ξj∂ξk

∂ξk

∂ηpvp

=∂ξi

∂η`

∂ηm

∂ξjδnp Γ`

mn vp +∂ξi

∂η`

∂2η`

∂ξj∂ξk

∂ξk

∂ηpvp

=∂ξi

∂η`

∂ηm

∂ξjΓ`

mn vn +∂ξi

∂η`

∂2η`

∂ξj∂ξk

∂ξk

∂ηpvp (A.10)

Hence, we have

∂vi

∂ξj+ Γi

jkvk =

∂ηm

∂ξj

∂vn

∂ηm

∂ξi

∂ηn+

∂ηm

∂ξjvn ∂2ξi

∂ηm∂ηn

+∂ξi

∂η`

∂ηm

∂ξjΓ`

mn vn +∂ξi

∂η`

∂2η`

∂ξj∂ξk

∂ξk

∂ηpvp

=∂ηm

∂ξj

∂vn

∂ηm

∂ξi

∂ηn+

∂ξi

∂η`

∂ηm

∂ξjΓ`

mn vn

+∂ηm

∂ξjvn ∂2ξi

∂ηm∂ηn+

∂ξi

∂η`

∂2η`

∂ξj∂ξk

∂ξk

∂ηpvp

=∂ξi

∂ηn

∂ηm

∂ξj

(∂vn

∂ηm+ Γn

mpvp

)

+

(∂ηm

∂ξj

∂2ξi

∂ηm∂ηn+

∂ξi

∂η`

∂2η`

∂ξj∂ξk

∂ξk

∂ηn

)vn. (A.11)

195

We will show that the last line vanishes. Note that

∂ηm

∂ξj

∂2ξi

∂ηm∂ηn=

∂ξj

(∂ξi

∂ηn

)(A.12)

∂ξi

∂η`

∂2η`

∂ξj∂ξk

∂ξk

∂ηn=

∂ξi

∂η`

∂ξk

∂ηn

∂ξj

(∂η`

∂ξk

). (A.13)

We will view the partial derivatives as functions of (ξ1, ξ2, ξ3). Let

Aij =

∂ξj

∂ηj,

(A−1

)i

j=

∂ηi

∂ξj, (A.14)

as these collections of partial derivatives do form inverse matrices. Then the coeffi-

cient of vn is∂Ai

n

∂ξj+ Ai

`Akn

∂ (A−1)`k

∂ξj. (A.15)

Since (A−1)`p Ap

q = δ`q,

∂ξj

(A−1

)`

pAp

q = 0

Apq

∂ξj

(A−1

)`

p= − (

A−1)`

p

∂ξjAp

q

(A−1

)q

kAp

q

∂ξj

(A−1

)`

p= − (

A−1)q

k

(A−1

)`

p

∂ξjAp

q

δpk

∂ξj

(A−1

)`

p= − (

A−1)q

k

(A−1

)`

p

∂ξjAp

q

∂ξj

(A−1

)`

k= − (

A−1)q

k

(A−1

)`

p

∂ξjAp

q (A.16)

The coefficient of vn can be written as

∂Ain

∂ξj− Ai

`Akn

(A−1

)q

k

(A−1

)`

p

∂ξjAp

q

=∂Ai

n

∂ξj− δi

p δqn

∂ξjAp

q

=∂Ai

n

∂ξj− ∂Ai

n

∂ξj

= 0. (A.17)

The result is that

∂vi

∂ξj+ Γi

jkvk =

∂ξi

∂ηn

∂ηm

∂ξj

(∂vn

∂ηm+ Γn

mpvp

). (A.18)

196

APPENDIX B

SYMMETRY OF P · F T

In many introductions, such as Gurtin (1981) and Ogden (1984), this is derived

in terms of the Cauchy stress. We follow the derivation in Section X11.7 of Antman

(1995).

With the first Piola-Kirchhoff stress known, Eq. 5.10 can be written as

∂(XΨ)(B)

χ× (P · ν) dS +

(XΨ)(B)

(χ× f − ρ0χ× χtt) dV = 0, (B.1)

where ν is the outward-pointing unit normal to the surface. This holds for all

well-behaved subsets Ω ⊂ Ψ(B):

∂X(Ω)

χ× (P · ν) dS +

X(Ω)

(χ× f − ρ0χ× χtt) dV = 0. (B.2)

To localize the balance of angular momentum, we must convert the surface in-

tegral to a volume integral. However, χ × (P · ν) is not the flux of a vector field

through the surface, so the standard Divergence Theorem for vector fields is not

applicable. For a fixed vector b, though, b · P is a vector field, and b · P · ν is the

flux of that vector field. We can apply the Divergence Theorem to such vector fields.

We take the dot-product of this equation with a×b, where a and b are constant

vectors.

(a× b) · (χ× (P · ν)) = (a · χ) (b · P · ν)− (a · P · ν) (b · χ) (B.3)

= b · ((P · ν)⊗ χ) · a− a · ((P · ν)⊗ χ) · b= ((P · ν)⊗ χ) : (a⊗ b− b⊗ a) (B.4)

(a× b) · (χ× f) = (f ⊗ χ) : (a⊗ b− b⊗ a) (B.5)

(a× b) · (χ× χtt) = (χtt ⊗ χ) : (a⊗ b− b⊗ a) (B.6)

197

DIV ((a · χ) (b · P )) = (b · P ) ·GRAD (a · χ) + (a · χ) DIV (b · P )

= (b · P ) · (a · F ) + (a · χ) (b ·DIVP )

= b · (P · F T + (DIVP )⊗ χ) · a

=(P · F T + (DIVP )⊗ χ

): a⊗ b (B.7)

DIV ((b · χ) (a · P )) =(P · F T + (DIVP )⊗ χ

): b⊗ a (B.8)

∂X(Ω)

(a× b) · (χ× (P · ν)) dS

=

∂X(Ω)

((a · χ) (b · P · ν)− (a · P · ν) (b · χ)) dS

=

X(Ω)

(DIV ((a · χ) (b · P ))−DIV ((b · χ) (a · P ))) dV

=

X(Ω)

(P · F T + (DIVP )⊗ χ

): (a⊗ b− b⊗ a) dV (B.9)

The cross-product of Eq. (B.1) with a× b can now be written as∫

X(Ω)

(P · F T + (DIVP + f − ρ0χtt)⊗ χ

)dV : (a⊗ b− b⊗ a) = 0 (B.10)

for all fixed vectors a and b. By the balance of linear momentum as given in

Eq. (5.18),

DIVP + f − ρ0χtt ≡ 0, (B.11)

so we have ∫

X(Ω)

P · F T dV : (a⊗ b− b⊗ a) = 0 (B.12)

Since a and b are arbitrary,∫

X(Ω)

P · F T dV : B = 0 (B.13)

for each skew-symmetric Lagrangian tensor B. By the near-arbitrariness of the set

Ω, we conclude that almost every pointwise value of P ·F T is symmetric: P ·F T =

F · P T .

198

APPENDIX C

POLAR DECOMPOSITION OF A TWO-POINT TENSOR

The existence of a polar decomposition of an invertible linear transformation is

frequently used in finite elasticity.

Theorem C.0.4 If F is an arbitrary linear transformation on a finite-dimensional

inner-product space, then there is a uniquely determined positive semi-definite linear

transformation U and an isometry R such that F = R ·U. If F is invertible, then

R is also unique.

This is quoted, with a slight change of notation, from Halmos (1958). Given this

result, elementary linear algebra can be used to show that if the linear transformation

F is invertible, then U is positive-definite. It can also be shown that if F is a linear

transformation on a real inner-product space with positive determinant, then R is

an orthogonal matrix, i.e. RT ·R = I, the identity.

The proof involves the properties of FT ·F. In the case of a real linear transfor-

mation F with positive determinant, then FT ·F is symmetric and positive-definite,

and U is the unique positive-definite symmetric square root of FT · F.

It should be noted that this theorem holds for a linear transformation from an

inner-produce space into the same inner-product space, yet it is the deformation

gradient, a two-point tensor, that is most frequently subjected to the polar decom-

position in finite elasticity.

Each deformation gradient is a map from the tangent space at X ∈ (X Ψ)(B)

to the tangent space at the deformed point x = χ(X), and it is most often expressed

initially with respect to the dual basis Ej at X and the coordinate basis ei at x:

F =∂ζ i

∂ξjei ⊗Ej, ei =

∂x

∂ζ i, Ej dual to Ej =

∂X

∂ξj.

199

If we follow the proof of the existence of the polar decomposition, we consider

F T · F :

F T · F =

(∂ζk

∂ξ`E` ⊗ ek

)·(

∂ζ i

∂ξjei ⊗Ej

)

= eki∂ζk

∂ξ`

∂ζ i

∂ξjE` ⊗Ej.

F T · F is symmetric and positive-definite and maps from the tangent space at

X into that same tangent space, i.e. it is Lagrangian. However, F T · F actually

maps tangent vectors Ej (contravariant) to cotangent vectors E` (covariant). At a

point in E3, cotangent vectors and tangent vectors are of the same kind, but they

have opposite transformation properties and correspond to very different objects in

Ψ(B), the space of coordinates. F T · F can be expressed as a map from tangent

vectors to tangent vectors as follows:

F T · F = eki∂ζk

∂ξ`

∂ζ i

∂ξjE` ⊗Ej = eki

∂ζk

∂ξ`

∂ζ i

∂ξjE`m Em ⊗Ej.

Since F T · F is a symmetric, positive-definite, Lagrangian tensor, it has a sym-

metric, positive-definite Lagrangian square root. It would seem natural to let U be

this Lagrangian square root. However, Theorem 6.4.1 states that a response func-

tion P (the function that assigns values of the first Piola-Kirchhoff stress tensor)

must satisfy

P (F ) = R · P (U ),

where F = R · U is the polar decomposition. If F is a two-point tensor and U is

Lagrangian, they cannot both be used as arguments of P . It would seem that U

must be a two-point square root. But it is not clear in what sense a two-point tensor

can be symmetric. In addition, the commonly-used notation U 2 is meaningless if

U is a two-point tensor:

U ·U =(U i

j ei ⊗EJ) · (Uk

` ek ⊗E`)

is undefined, as Ej and ek belong to different tangent spaces, so the dot-product

Ej · ek is undefined.

200

APPENDIX D

MATHEMATICA CODE FOR THE

RIEMANN-CHRISTOFFEL CURVATURE TENSOR

The following Mathematica commands compute the Riemann-Christoffel curvature

tensor for metric tensor presented in Eq. (4.21). In our experience, Mathematica

simplifies expressions more rapidly when computing the coefficients

Rqijk =∂Γikq

∂ξj− ∂Γijq

∂ξk+ Γp

ijΓkqp − ΓpikΓjqp,

where Γijq are the Christoffel symbols of the first kind:

Γijq =1

2

(∂Eiq

∂ξj+

∂Ejq

∂ξi− ∂Eij

∂ξq

).

and Γkij are the Christoffel symbols of the second kind:

Γkij = EkqΓijq.

We use the symbol g for the metric tensor, as E is protected for other uses in

Mathematica.

(* Define cylindrical coordinates. *)

coords = R,Θ,Z(* Define the metric tensor with respect to basis dual to coordinate basis. *)

g[R , Θ , Z ] :=

(r′[R])2 + r[R]2θ(1,0)[R, Θ]2, r[R]2θ(1,0)[R, Θ]θ(0,1)[R, Θ], 0

,r[R]2θ(1,0)[R, Θ]θ(0,1)[R, Θ], r[R]2θ(0,1)[R, Θ]2, 0

,

0, 0, 1

201

(* Define inverse of metric tensor. *)

ginv[R , Θ , Z ] := Simplify[Inverse[g[R, Θ, Z]]]

(* Define Christoffel symbols of the first kind. *)

Γfirst[i , j , q , R , Θ , Z ] =1

2(D[g[R, Θ, Z][[i]][[q]], coords[[j]]]

+ D[ g[R, Θ, Z][[j]][[q]], coords[[i]] ]

− D[ g[R, Θ, Z][[i]][[j]], coords[[q]] ])

(* Define Christoffel symbols of the second kind. The last integer index is the

raised index. *)

Γsecond[i , j , q , R , Θ , Z ]

:= Sum[ ginv[R, Θ, Z][[k]][[q]] ∗ Γfirst[i, j, q, R, Θ, Z], q, 1, 3 ]

(* Define the Riemann-Christoffel curvature tensor. *)

Riem[q , i , j , R , Θ , Z ]

= D[ Γfirst[i, k, q, R, Θ, Z], coords[[j]] ]

− D[ Γfirst[i, j, q, R, Θ, Z], coords[[k]] ]

+ Sum[ Γsecond[i, j, p, R, Θ, Z] ∗ Γfirst[k, q, p, R, Θ, Z], p, 1, 3 ]

− Sum[ Γsecond[i, k, p, R, Θ, Z] ∗ Γfirst[j, q, p, R, Θ, Z], p, 1, 3 ]

(* Print out the components of the Riemann-Christoffel curvature tensor. *)

For[q = 1, m < 4, m++,

For[i = 1, i < 4, i++,

For[j = 1, j < 4, j++,

For[k = 1, k < 4, k++,

Print[”Riem[”, q, ”,”, i, ”,”, j, ”,”, k, ”,R,Θ,Z] = ”,Riem[q,i,j,k,R,Θ,Z]]

]

]

]

]

202

REFERENCES

Antman, S. S. (1995). Nonlinear Problems of Elasticity. Springer, Berlin.

Antman, S. S. and P. V. Negron-Marrero (1987). The remarkable nature of radiallysymmetric equilibrium states of nonlinearly elastic bodies. Journal of Elasticity,18, pp. 131–164.

Antman, S. S. and J. E. Osborn (1979). The principle of virtual work and integrallaws of motion. Archive for Rational Mechanics and Analysis, 69, pp. 321–262.

Ball, J. M. (1982). Discontinuous equilibrium solutions and cavitation in nonlinearelasticity. Philosophical Transactions of the Royal Society of London. Series A,Mathematical and Physical Sciences, 306, pp. 557–611.

Ben Amar, M. and A. Goriely (2005). Growth and instabilities in elastic tissues.Journal of the Mechanics and Physics of Solids, 53, pp. 2284–2319.

Bernoulli, J. (1789). Essai theorique sur les vibrations de plaques elastiques rectan-gulaires et libres. Nova Acta Academiae Scientiarum Imperialis Petropolitanae,5, pp. 197–219.

Bloom, F. (1979). Modern Differential Geometric Techniques in the Theory of Con-tinuous Distributions of Dislocations. Springer-Verlag, Berlin.

Brauer, F. and J. A. Nohel (1989). The Qualitative Theory of Ordinary DifferentialEquations. Dover Publications, New York.

Chen, Y.-C. and A. Hoger (2000). Constitutive functions of elastic materials in finitegrowth and deformation. Journal of Elasticity, 59, pp. 173–193.

Ciarlet, P. G. (2005). An Introduction to Differential Geometry with Applicationsto Elasticity. Springer Netherlands.

DiCarlo, A., P. Podio-Guiduli, and W. O. Williams (2001). Shells with thicknessdistension. International Journal of Solids and Structures, 38, pp. 1201–1225.

DiCarlo, A. and S. Quiligotti (2002). Growth and balance. Mechanics ResearchCommunications, 29, pp. 449–456.

Dollhofer, J., A. Chiche, V. Muralidharan, C. Creton, and C. Y. Hui (2004). Surfaceenergy effects for cavity growth and nucleation in an incompressible neo-Hookeanmaterial — modeling and experiment. International Journal of Solids and Struc-tures, 41, pp. 6111–6127.

203

Eckart, C. (1948). The Thermodynamics of Irreversible Processes. IV. The Theoryof Elasticity and Anelasticity. Physical Review, 73(4), pp. 373–382.

Efrati, E., Y. Klein, H. Aharoni, and E. Sharon (2007a). Shaping of elastic sheetsby prescription of non-Euclidean metrics. Science, 315(5816), pp. 1116–1120.

Efrati, E., Y. Klein, H. Aharoni, and E. Sharon (2007b). Shaping of elastic sheets byprescription of non-Euclidean metrics: supplementary materials. Science website.http://www.sciencemag.org/content/vol315/issue5815/images/data/

1116/DC1/1135994S1.mpg.

Efrati, E., Y. Klein, H. Aharoni, and E. Sharon (2007c). Spontaneous buckling ofelastic sheets with a prescribed non-Euclidean metric. Physica D, 235, pp. 29–32.

Efrati, E. and R. Kupferman (2009). Buckling transition and boundary layer innon-Euclidean plates. Physical Review E, 80, pp. 016602–1–016602–18.

Efrati, E., E. Sharon, and R. Kupferman (2009). Elastic theory of unconstrainednon-Euclidean plates. Journal of the Mechanics and Physics of Solids, 57, pp.762–775.

Epstein, M. and G. A. Maugin (2000). Thermomechanics of volumetric growth inuniform bodies. International Journal of Plasticity, 16, pp. 951–978.

Euler, L. P. (1766). De motu vibratorio tympanorum. Novi Commentarii AcademiaePetropolitanae, 10, pp. 243–260.

Fond, C. (2001). Cavitation criterion for rubber materials: a review of void-growthmodels. Journal of Polymer Science Part B: Polymer Physics, 39, pp. 2081–2096.

Foppl, A. (1907). Vorlesungen uber technischke Mechanik. B. G. Teubner, Leipzig.

Frankel, T. (1997). The Geometry of Physics: An Introduction. Cambridge Univer-sity Press, Cambridge.

Friesecke, G., R. D. James, and S. Muller (2006). A hierarchy of plate modelsderived from nonlinear elasticity by Gamma-convergence. Archive for RationalMechanics and Analysis, 180(2), pp. 183–236.

Gent, A. N. and P. B. Lindley (1959). Internal rupture of bonded rubber cylindersin tension. Proceedings of the Royal Society of London. Series A: Mathematicaland Physical Sciences, 249, pp. 195–205.

Goriely, A. and M. Ben Amar (2005). Differential growth and instabilities in elasticshells. Physical Review Letters, 94(19), p. 198103.

204

Goriely, A., M. Destrade, and M. Ben Amar (2006). Instabilities in elastomers andsoft tissues. Quarterly Journal of Mechanics and Applied Mathematics, 59, pp.615–630.

Goriely, A., R. Vandiver, and M. Destrade (2008). Nonlinear Euler Buckling. Pro-ceedings of the Royal Society of London Series A, 464, pp. 3003–3019.

Gurtin, M. (1981). An Introduction to Continuum Mechanics. Academic Press, NewYork.

Halmos, P. R. (1958). Finite-Dimensional Vector Spaces, 2nd edition. D. VanNostrand Company, Princeton, NJ.

Han, Q. (2006). Isometric Embedding of Riemannian Manifolds in Euclidean Spaces.American Mathematical Society, Providence, RI.

Haughton, D. M. (1986). On non-existence of cavitation in incompressible elasticmembranes. Quarterly Journal of Mechanics and Applied Mathematics, 39, pp.289–296.

Haughton, D. M. (2001). Elastic membranes. In Fu, Y. B. and R. W. Ogden(eds.) Nonlinear Elasticity: Theory and Applications, pp. 233–267. CambridgeUniversity Press, Cambridge.

Haughton, D. M. and R. W. Ogden (1978). On the incremental equations in non-linear elasticity — I. Membrane theory. Journal of the Mechanics and Physics ofSolids, 26, pp. 93–110.

Hoger, A. (1997). Virtual configurations and constitutive equations for residuallystressed bodies with material symmetry. Journal of Elasticity, 48, pp. 125–144.

Hoger, A. (1999). An incremental mechanical theory of growth for soft biological tis-sues. Unpublished notes. See also A. Hoger, “An incremental theory of volumetricgrowth for soft biological tissues,” 1999 ASME Mechanics and Materials Confer-ence, June 27–30, 1999, Virginia Polytechnic and State University, Blacksburg,Virginia.

Horgan, C. O. (2001). Equilibrium Solutions for Compressible Nonlinearly ElasticMaterials. In Fu, Y. B. and R. W. Ogden (eds.) Nonlinear Elasticity: Theory andApplications, pp. 135–159. Cambridge University Press, Cambridge.

Horgan, C. O. and D. A. Polignone (1993). Cavitation for incompressible anisotropicnonlinearly elastic spheres. Journal of Elasticity, 33, pp. 27–65.

Horgan, C. O. and D. A. Polignone (1995). Cavitation in nonlinearly elastic solids:A review. Applied Mechanics Reviews, 48, pp. 471–485.

205

Humphrey, J. D. (2002). Cardiovascular Solid Mechanics: Cells, Tissues, and Or-gans. Springer-Verlag, New York.

Humphrey, J. D. (2006). Towards a Theory of Vascular Growth and Remodeling. InHolzepfel, G. and R. W. Ogden (eds.) Mechanics of Biological Tissue, pp. 3–16.

Kirchhoff, G. R. (1876). Vorlesungen uber mathematische Physik. B. G. Teubner,Leipzig.

Klarbring, A., T. Olsson, and J. Stalhand (2007). Theory of residual stresses withapplication to an arterial geometry. Archives of Mechanics, 59, pp. 341–364.

Klisch, S. M., S. S. Chen, R. S. Sah, and A. Hoger (2003). A growth mixture theoryfor cartilage with application to growth-related experiments on cartilage explants.Journal of Biomechanical Engineering, 125(2), pp. 169–179.

Kroner, E. (1959). Allgemeine Kontinuumstheorie der Versetzungen und Eigenspan-nungen (General Continuum Theory of Dislocations and Residual Stresses).Archive for Rational Mechanics and Analysis, 4(1), pp. 273–334.

Lee, E. H. and D. T. Liu (1967). Finite-Strain Elastic-Plastic Theory with Appli-cation to Plane-Wave Analysis. Journal of Applied Physics, 38(1), pp. 19–27.

Lee, J. M. (1997). Riemannian Manifolds: An Introduction to Curvature. Springer-Verlag, Dordrecht.

Lin, I.-E. and L. A. Taber (1995). A model for stress-induced growth in the devel-oping heart. Journal of Biomechanical Engineering, 117(3), pp. 343–349.

Liu, I.-S. (2002). Continuum Mechanics. Springer, Berlin.

Lubarda, V. A. and A. Hoger (2002). On the mechanics of solids with a growingmass. International Journal of Solids and Structures, 39.

Marion, J. B. and S. T. Thornton (1988). Classical Dynamics of Particles andSystems, 3rd edition. Harcourt Brace Jovanovich, Forth Worth, TX.

Marsden, J. E. and T. J. R. Hughes (1983). Mathematical Foundations of Elasticity.Prentice-Hall.

Noll, W., R. A. Toupin, and C.-C. Wang (1968). Continuum Theory of Inhomo-geneity in Simple Bodies: A Reprint of Six Memoirs. Springer-Verlag, New York.

Ogden, R. W. (1984). Non-Linear Elastic Deformations. Dover Publications, Mi-neola, NY.

206

Pericak-Spector, K. A., J. Sivaloganathan, and S. J. Spector (2002). An explicitradial cavitation solution in nonlinear elasticity. Mathematics and Mechanics ofSolids, 7, pp. 87–93.

Perko, L. (1996). Differential Equations and Dynamical Systems, 2nd ed. Springer,New York.

Rachev, A., N. Stergiopulos, and J.-J. Meister (1998). A model for geometric andmechanical adaptation of arteries to sustained hypertension. Journal of Biome-chanical Engineering, 120(1), pp. 9–17.

Rodriguez, E. K., A. Hoger, and A. D. McCulloch (1994). Stress-dependent finitegrowth law in soft elastic tissue. Journal of Biomechanics, 27, pp. 455–467.

Scott, N. H. (2007). The incremental bulk modulus, Young’s modulus and Poisson’sratio in nonlinear isotropic elasticity: physically reasonable response. Mathematicsand Mechanics of Solids, 12, pp. 526–524.

Simonite, T. (2009). Chemical “caterpillar” points to electronics-free robots. NewScientist website. http://www.newscientist.com/article/dn16910-chemical-caterpillar-points-to-electronicsfree-robots.html.

Sivaloganathan, J. (1999). On cavitation and degenerate cavitation under internalhydrostatic pressure. Proceedings of the Royal Society A: Mathematical, Physicaland Engineering Sciences, 455, pp. 3645–3664.

Sivaloganathan, J. and S. J. Spector (2002). On cavitation, configurational forcesand implications for fracture in a nonlinearly elastic material. Journal of Elastic-ity, 67, pp. 25–49.

Skalak, R., S. Zargaryan, R. J. Jain, P. A. Netti, and A. Hoger (1996). Compatibilityand the genesis of residual stress by volumetric growth. Journal of MathematicalBiology, 34, pp. 889–914.

Steigmann, D. J. (1992). Cavitation in elastic membranes—an example. Journal ofElasticity, 28, pp. 277–287.

Strang, G. (1980). Linear Algebra and Its Applications. Academic Press.

Strasburger, E. (1895). The Development of Botany in Germany During the Nine-teenth Century. Botanical Gazette, 20(5), pp. 193–204.

Szilard, R. (2004). Theories and Applications of Plate Analysis: Classical, Numer-ical, and Engineering Methods. John Wiley & Sons, Hoboken, NJ.

207

Taber, L. A. (1998a). Biomechanical growth laws for muscle tissue. Journal ofTheoretical Biology, 193(2), pp. 201–213.

Taber, L. A. (1998b). A model of aortic growth based on fluid shear and fiberstresses. Journal of Biomechanical Engineering, 120(3), pp. 348–354.

Taber, L. A. and D. W. Eggers (1996). Theoretical study of stress-modulated growthin the aorta. Journal of Theoretical Biology, 180.

Taber, L. A. and J. D. Humphrey (2001). Stress-modulated growth, residual stress,and vascular homogeneity. Transactions of the American Society of MechanicalEngineers, 123, pp. 528–535.

Taber, L. A. and R. Perucchio (2000). Modeling Heart Development. Journal ofElasticity, 61, pp. 165–197.

Todhunter, I. and K. Pearson (1960). A History of the Theory of Elasticity, volumesI and II. Dover Publications, New York.

Truesdell, C. (1978). Some challenges offered to analysis by rational thermome-chanics. In de la Penha, G. M. and L. A. J. Medeiros (eds.) ContemporrayDevelopments in Continuum Mechanics and Partial Differential Equations, pp.495–603.

Tvergaard, V. (1996). Effect of void size difference on growth and cavitation insta-bilities. Journal of the Mechanics and Physics of Solids, 44, pp. 1237–1253.

Vandiver, R. (2009). Morphoelasticity: The Mechanics and Mathematics of ElasticGrowth. Ph.D. thesis, The University of Arizona, Tucson, Arizona.

Vandiver, R. and A. Goriely (2008). Tissue tension and axial growth of cylindricalstructures in plants and elastic tissues. Europhysics Letters, 84, pp. 58004–p1–58004–p6.

Ventsel, E. and T. Krauthammer (2001). Thin Plates and Shells: Theory, Analysis,and Applications. Marcel Dekker, New York.

von Karman, T. (1910). Festigkeitsprobleme im Maschinenbau. In Encyklopadie derMathematischen Wissenschaften, volume 4, pp. 348–352.