Sunteți pe pagina 1din 271

0749_Frame_FM Page iii Wednesday, February 19, 2003 4:53 PM

Finite
Element
Analysis
Thermomechanics of Solids

David W. Nicholson

CRC PR E S S
Boca Raton London New York Washington, D.C.

2003 by CRC CRC Press LLC

0749_Frame_FM Page iv Wednesday, February 19, 2003 4:53 PM

Library of Congress Cataloging-in-Publication Data


Nicholson, D.W.
Finite element analysis : thermomechanics of solids / David W. Nicholson.
p. cm.
Includes bibliographical references and index.
ISBN 0-8493-0749-X (alk. paper)
1. Thermal stressesMathematical models. 2. Finite element method. I. Title.
TA418.58.N53 2003
620.1121dc21

2002041419

This book contains information obtained from authentic and highly regarded sources. Reprinted material
is quoted with permission, and sources are indicated. A wide variety of references are listed. Reasonable
efforts have been made to publish reliable data and information, but the author and the publisher cannot
assume responsibility for the validity of all materials or for the consequences of their use.
Neither this book nor any part may be reproduced or transmitted in any form or by any means, electronic
or mechanical, including photocopying, microfilming, and recording, or by any information storage or
retrieval system, without prior permission in writing from the publisher.
The consent of CRC Press LLC does not extend to copying for general distribution, for promotion, for
creating new works, or for resale. Specific permission must be obtained in writing from CRC Press LLC
for such copying.
Direct all inquiries to CRC Press LLC, 2000 N.W. Corporate Blvd., Boca Raton, Florida 33431.
Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are
used only for identification and explanation, without intent to infringe.

Visit the CRC Press Web site at www.crcpress.com


2003 by CRC CRC Press LLC
No claim to original U.S. Government works
International Standard Book Number 0-8493-0749-X
Library of Congress Card Number 2002041419
Printed in the United States of America 1 2 3 4 5 6 7 8 9 0
Printed on acid-free paper

2003 by CRC CRC Press LLC

0749_Frame_FM Page v Wednesday, February 19, 2003 4:53 PM

Dedication
To Linda and Mike, with profound love

2003 by CRC CRC Press LLC

0749_Frame_FM Page vii Wednesday, February 19, 2003 4:53 PM

Preface
Thousands of engineers use finite-element codes, such as ANSYS, for thermomechanical and nonlinear applications. Most academic departments offering advanced
degrees in mechanical engineering, civil engineering, and aerospace engineering offer
a first-level course in the finite-element method, and by now, almost all undergraduates
of such programs have some exposure to the finite-element method. A number of
departments offer a second-level course. It is hoped that this text will appeal to
instructors of such courses. Of course, it hopefully will also be helpful to engineers
engaged in self-study on nonlinear and thermomechanical finite-element analysis.
The principles of the finite-element method are presented for application to the
mechanical, thermal, and thermomechanical response, both static and dynamic, of
linear and nonlinear solids. It provides an integrated treatment of:
Basic principles, material models, and contact models (for example, linear
elasticity, hyperelasticity, and thermohyperelasticity).
Computational, numerical, and software-design aspects (such as finiteelement data structures).
Modeling principles and strategies (including mesh design).
The text is designed for a second-level course, as a reference work, or for self
study. Familiarity is assumed with the finite-element method at the level of a firstlevel graduate or advanced undergraduate course.
A first-level course in the finite-element method, for which many excellent books
are available, barely succeeds in covering static linear elasticity and linear heat transfer.
There is virtually no exposure to nonlinear methods, which are topics for a secondlevel course. Nor is there much emphasis on coupled thermomechanical problems.
However, many engineers could benefit from a text covering nonlinear problems
and the associated continuum thermomechanics. Such a text could be used in a
formal class or for self-study. Many important applications have significant nonlinearity, making nonlinear finite-element modeling necessary. As a few examples, we
mention polymer processing; metal forming; rubber components, such as tires and
seals; biomechanics; and crashworthiness. Many applications combine thermal and
mechanical response, such as rubber seals in hot engines. Engineers coping with
such applications have access to powerful finite-element codes and computers.
However, they often lack and urgently need an in-depth but compact exposition of
the finite-element method, which provides a foundation for addressing future problems. It is hoped this text also fills this need.
Of necessity, a selection of topics has been made, and topics are given coverage
proportional to the authors sense of their importance to the readers understanding.
Topics have been selected with the intent of giving a unified and complete, but still
compact and tractable, presentation. Several other excellent texts and monographs

2003 by CRC CRC Press LLC

0749_Frame_FM Page viii Wednesday, February 19, 2003 4:53 PM

have appeared over the years, from which the author has benefited. Four texts to
which the author is indebted are:
1. Zienkiewicz, O.C. and Taylor R.L., The Finite Element Method, Vols. 1
and 2, McGraw Hill, London, 1989.
2. Kleiber, M., Incremental Finite Element Modeling in Nonlinear Solid
Mechanics, Chichester, Ellis Horwood, Ltd., 1989.
3. Bonet, J. and Wood, R.D., Nonlinear Continuum Mechanics for Finite
Element Analysis, Cambridge, Cambridge University Press, 1997.
4. Belytschko, T., Lui, W.K, and Moran, B., Nonlinear Finite Elements for
Continua and Structures, New York, John Wiley and Sons, 2000.
This text has the following characteristics:
1. Emphasis on the use of Kronecker Product notation instead of tensor,
tensor-indicial, Voigt, or traditional finite-element, matrix-vector notation.
2. Emphasis on integrated and coupled thermal and mechanical effects.
3. Inclusion of elasticity, hyperelasticity, plasticity, and viscoelasticity with
thermal effects.
4. Inclusion of nonlinear boundary conditions, including contact, in an integrated incremental variational formulation.
Kronecker Product algebra (KPA) has been widely used in control theory for many
years (Graham, 1982). It is highly compact and satisfies simple rules: for example,
the inverse of a Kronecker Product of two nonsingular matrices is the Kronecker
Product of the inverses. Recently, a number of extensions of KPA have been introduced and shown to permit compact expressions for otherwise elaborate quantities
in continuum and computational mechanics. Examples include:
1. Compact expressions for the tangent-modulus tensors in hyperelasticity
(invariant-based and stretch-based; compressible, incompressible, and nearincompressible), thermohyperelasticity, and finite-strain plasticity.
2. A general, compact expression for the tangent stiffness matrix in nonlinear
FEA, including nonlinear boundary conditions, such as contact.
KPA with recent extensions can completely replace other notations in most cases
of interest here. In the authors experience, students experience little difficulty in
gaining a command of it.
The first three chapters concern mathematical foundations, and Kronecker Product
notation for tensors is introduced. The next four chapters cover relevant linear and
nonlinear continuum thermomechanics to enable a unified account of the finite-element
method. Chapters 8 through 15 represent a compact presentation of the finite-element
method in linear elastic, thermal, and thermomechanical media, including solution
methods. The final five chapters address nonlinear problems based on a unified set of
incremental variational principles. Material nonlinearity is treated also, as is geometric
nonlinearity and nonlinearity due to boundary conditions. Several numerical issues in
nonlinear analysis are discussed, such as iterative triangularization of stiffness matrices.

2003 by CRC CRC Press LLC

0749_Frame_FM Page ix Wednesday, February 19, 2003 4:53 PM

Acknowledgment
Finite Element Analysis: Thermomechanics of Solids is the culmination of many
years of teaching and research, made possible through the love and patience of my
wife Linda and son Michael, to whom this is dedicated. To my profound love, I
cheerfully add my immense gratitude. Much of my understanding of the advanced
topics addressed in this book arose through the highly successful research of my
then-doctoral student, Dr. Baojiu Lin. He continued his invaluable support by providing comments and corrections on the manuscript. Also deserving of acknowledgment is Ashok Balasubranamian, whose careful attention to the manuscript has
greatly helped to make it achieve my objectives. Thanks are due to hundreds of
graduate students in my courses in continuum mechanics and finite elements over
the years. I tested and refined the materials through the courses and benefited
immensely from their strenuous efforts to gain command of the course materials.
Finally, I would like to thank CRC Press for taking a chance on me and especially
for Cindy Renee Carelli for patience and many good suggestions.

2003 by CRC CRC Press LLC

0749_Frame_FM Page xi Wednesday, February 19, 2003 4:53 PM

About the Author


David W. Nicholson, Ph.D. is a professor in the Department of Mechanical, Materials, and Aerospace Engineering at the University of Central Florida, Orlando, Florida,
where he has served as the Chair (19901995) and Interim
Chair (20002002). He earned an S.B. degree at MIT in
1966 and a doctorate at Yale in 1971. He has 31 years of
experience since the doctorate, including research in
industry (Goodyear, 7 years) and government (the Naval
Surface Weapons Center, 6 years), and faculty positions
at Stevens Institute of Technology and UCF. He has
authored over 140 articles on the finite-element method,
continuum mechanics, fracture mechanics, and dynamics.
He has served as a technical editor for Applied Mechanics
Review, as an associate editor of Tire Science and Technology, and as a member of
scientific advisory committees for a number of international conferences. He has
been the instructor in over 20 short courses and workshops. Recently, he served as
Executive Chair of the XXIst Southeastern Conference on Theoretical and Applied
Mechanics and as co-editor of Developments in Theoretical and Applied Mechanics,
Volume XXI.

2003 by CRC CRC Press LLC

0749_Frame_FM Page xiii Wednesday, February 19, 2003 4:53 PM

Table of Contents
Chapter 1
1.1

1.2

1.3

1.4

Introduction....................................................................................................1
1.1.1
Range and Summation Convention ................................................1
1.1.2
Substitution Operator ......................................................................1
Vectors............................................................................................................2
1.2.1
Notation ...........................................................................................2
1.2.2
Gradient, Divergence, and Curl ......................................................4
Matrices..........................................................................................................5
1.3.1
Eigenvalues and Eigenvectors.........................................................8
1.3.2
Coordinate Transformations............................................................9
1.3.3
Transformations of Vectors .............................................................9
1.3.4
Orthogonal Curvilinear Coordinates.............................................11
1.3.5
Gradient Operator..........................................................................16
1.3.6
Divergence and Curl of Vectors....................................................17
Appendix I: Divergence and Curl of Vectors in Orthogonal
Curvilinear Coordinates................................................................................18
Derivatives of Base Vectors ..........................................................18
Divergence.....................................................................................19
Curl ................................................................................................20
Exercises ......................................................................................................20

Chapter 2
2.1
2.2

2.3
2.4
2.5
2.6

Mathematical Foundations: Vectors and Matrices...............................1

Mathematical Foundations: Tensors ..................................................25

Tensors .........................................................................................................25
Divergence, Curl, and Laplacian of a Tensor .............................................27
2.2.1
Divergence.....................................................................................27
2.2.2
Curl and Laplacian ........................................................................28
Invariants......................................................................................................29
Positive Definiteness....................................................................................30
Polar Decomposition Theorem....................................................................31
Kronecker Products on Tensors...................................................................32
2.6.1
VEC Operator and the Kronecker Product ...................................32
2.6.2
Fundamental Relations for Kronecker Products...........................33
2.6.3
Eigenstructures of Kronecker Products ........................................35
2.6.4
Kronecker Form of Quadratic Products........................................36
2.6.5
Kronecker Product Operators for Fourth-Order Tensors..............36
2.6.6
Transformation Properties of VEC and TEN22 ............................37
2.6.7
Kronecker Product Functions for Tensor Outer Products ............38

2003 by CRC CRC Press LLC

0749_Frame_FM Page xiv Wednesday, February 19, 2003 4:53 PM

2.6.8

2.7

Kronecker Expressions for Symmetry Classes


in Fourth-Order Tensors ................................................................40
2.6.9
Differentials of Tensor Invariants .................................................41
Exercises ......................................................................................................42

Chapter 3
3.1
3.2

3.3

Introduction to Variational Methods............................................................43


Newton Iteration and Arc-Length Methods ................................................47
3.2.1
Newton Iteration............................................................................47
3.2.2
Critical Points and the Arc-Length Method .................................48
Exercises ......................................................................................................49

Chapter 4
4.1

4.2

4.3
4.4
4.5
4.6
4.7
4.8

5.2
5.3

5.4
5.5
5.6

Kinematics of Deformation ...............................................................51

Kinematics ...................................................................................................51
4.1.1
Displacement .................................................................................51
4.1.2
Displacement Vector......................................................................52
4.1.3
Deformation Gradient Tensor .......................................................52
Strain ............................................................................................................53
4.2.1
F, E, EL and u in Orthogonal Coordinates....................................53
4.2.2
Velocity-Gradient Tensor, Deformation-Rate Tensor,
and Spin Tensor.............................................................................56
Differential Volume Element .......................................................................60
Differential Surface Element .......................................................................61
Rotation Tensor............................................................................................63
Compatibility Conditions For EL and D .....................................................64
Sample Problems .........................................................................................67
Exercises ......................................................................................................69

Chapter 5
5.1

Introduction to Variational and Numerical Methods.........................43

Mechanical Equilibrium and the Principle of Virtual Work .............73

Traction and Stress ......................................................................................73


5.1.1
Cauchy Stress ................................................................................73
5.1.2
1st Piola-Kirchhoff Stress .............................................................75
5.1.3
2nd Piola-Kirchhoff Stress............................................................76
Stress Flux ...................................................................................................77
Balance of Mass, Linear Momentum, and Angular Momentum................79
5.3.1
Balance of Mass ............................................................................79
5.3.2
Rayleigh Transport Theorem ........................................................79
5.3.3
Balance of Linear Momentum ......................................................79
5.3.4
Balance of Angular Momentum....................................................80
Principle of Virtual Work ............................................................................82
Sample Problems .........................................................................................85
Exercises ......................................................................................................89

2003 by CRC CRC Press LLC

0749_Frame_FM Page xv Wednesday, February 19, 2003 4:53 PM

Chapter 6
6.1
6.2

6.3

6.4
6.5

Stress-Strain Behavior: Classical Linear Elasticity ....................................95


Isothermal Tangent-Modulus Tensor...........................................................97
6.2.1
Classical Elasticity ........................................................................97
6.2.2
Compressible Hyperelastic Materials ...........................................97
Incompressible and Near-Incompressible Hyperelastic Materials..............99
6.3.1
Incompressibility ...........................................................................99
6.3.2
Near-Incompressibility ................................................................102
Nonlinear Materials at Large Deformation...............................................103
Exercises ....................................................................................................104

Chapter 7
7.1

7.2
7.3

7.4

9.2

9.3

Introduction to the Finite-Element Method.....................................117

Introduction................................................................................................117
Overview of the Finite-Element Method ..................................................117
Mesh Development ....................................................................................118

Chapter 9
9.1

Thermal and Thermomechanical Response.....................................107

Balance of Energy and Production of Entropy.........................................107


7.1.1
Balance of Energy .......................................................................107
7.1.2
Entropy Production Inequality ....................................................108
7.1.3
Thermodynamic Potentials..........................................................109
Classical Coupled Linear Thermoelasticity ..............................................110
Thermal and Thermomechanical Analogs of the Principle
of Virtual Work ..........................................................................................113
7.3.1
Conductive Heat Transfer ...........................................................113
7.3.2
Coupled Linear Isotropic Thermoelasticity ................................114
Exercises ....................................................................................................116

Chapter 8
8.1
8.2
8.3

Stress-Strain Relation and the Tangent-Modulus Tensor ..................95

Element Fields in Linear Problems .................................................121

Interpolation Models..................................................................................121
9.1.1
One-Dimensional Members ........................................................121
9.1.2
Interpolation Models: Two Dimensions......................................124
9.1.3
Interpolation Models: Three Dimensions ...................................127
Strain-Displacement Relations and Thermal Analogs ..............................128
9.2.1
Strain-Displacement Relations: One Dimension ........................128
9.2.2
Strain-Displacement Relations: Two Dimensions ......................129
9.2.3
Axisymmetric Element on Axis of Revolution ..........................130
9.2.4
Thermal Analog in Two Dimensions..........................................131
9.2.5
Three-Dimensional Elements......................................................131
9.2.6
Thermal Analog in Three Dimensions........................................132
Stress-Strain-Temperature Relations in Linear Thermoelasticity.............132
9.3.1
Overview .....................................................................................132
9.3.2
One-Dimensional Members ........................................................132
9.3.3
Two-Dimensional Elements ........................................................133

2003 by CRC CRC Press LLC

0749_Frame_FM Page xvi Wednesday, February 19, 2003 4:53 PM

9.4

9.3.4
Element for Plate with Membrane and Bending Response .......135
9.3.5
Axisymmetric Element................................................................135
9.3.6
Three-Dimensional Element .......................................................136
9.3.7
Elements for Conductive Heat Transfer .....................................137
Exercises ....................................................................................................137

Chapter 10 Element and Global Stiffness and Mass Matrices ..........................139


10.1
10.2
10.3

10.3

Application of the Principle of Virtual Work............................................139


Thermal Counterpart of the Principle of Virtual Work.............................141
Assemblage and Imposition of Constraints ..............................................142
10.3.1 Rods.............................................................................................142
10.3.2 Beams ..........................................................................................146
10.3.3 Two-Dimensional Elements ........................................................147
Exercises ....................................................................................................149

Chapter 11 Solution Methods for Linear Problems ...........................................153


11.1

11.2
11.3
11.4
11.5

11.6

Numerical Methods in FEA ......................................................................153


11.1.1 Solving the Finite-Element Equations: Static Problems ............153
11.1.2 Matrix Triangularization and Solution of Linear Systems.........154
11.1.3 Triangularization of Asymmetric Matrices.................................155
Time Integration: Stability and Accuracy .................................................156
Newmarks Method ...................................................................................157
Integral Evaluation by Gaussian Quadrature ............................................158
Modal Analysis by FEA ............................................................................159
11.5.1 Modal Decomposition .................................................................159
11.5.2 Computation of Eigenvectors and Eigenvalues ..........................162
Exercises ....................................................................................................164

Chapter 12 Rotating and Unrestrained Elastic Bodies.......................................167


12.1
12.2
12.3

Finite Elements in Rotation.......................................................................167


Finite-Element Analysis for Unconstrained Elastic Bodies......................169
Exercises ....................................................................................................171

Chapter 13 Thermal, Thermoelastic, and Incompressible Media ......................173


13.1

13.2

13.3
13.4
13.5

Transient Conductive-Heat Transfer..........................................................173


13.1.1 Finite-Element Equation .............................................................173
13.1.2 Direct Integration by the Trapezoidal Rule ................................173
13.1.3 Modal Analysis............................................................................174
Coupled Linear Thermoelasticity ..............................................................175
13.2.1 Finite-Element Equation .............................................................175
13.2.2 Thermoelasticity in a Rod...........................................................177
Compressible Elastic Media ......................................................................177
Incompressible Elastic Media ...................................................................178
Exercises ....................................................................................................180

2003 by CRC CRC Press LLC

0749_Frame_FM Page xvii Wednesday, February 19, 2003 4:53 PM

Chapter 14 Torsion and Buckling.......................................................................181


14.1
14.2

14.3

Torsion of Prismatic Bars..........................................................................181


Buckling of Beams and Plates ..................................................................185
14.2.1 Euler Buckling of Beam Columns..............................................185
14.2.2 Euler Buckling of Plates .............................................................190
Exercises ....................................................................................................193

Chapter 15 Introduction to Contact Problems....................................................195


15.1
15.2
15.3
15.4

Introduction: the Gap.................................................................................195


Point-to-Point Contact ...............................................................................197
Point-to-Surface Contact ...........................................................................199
Exercises ....................................................................................................199

Chapter 16 Introduction to Nonlinear FEA........................................................201


16.1
16.2
16.3
16.4

16.5
16.6

Overview....................................................................................................201
Types of Nonlinearity ................................................................................201
Combined Incremental and Iterative Methods: a Simple Example..........202
Finite Stretching of a Rubber Rod under Gravity: a Simple Example ....203
16.4.1 Nonlinear Strain-Displacement Relations...................................203
16.4.2 Stress and Tangent Modulus Relations.......................................204
16.4.3 Incremental Equilibrium Relation...............................................205
16.4.4 Numerical Solution by Newton Iteration....................................208
Illustration of Newton Iteration.................................................................211
16.5.1 Example.......................................................................................212
Exercises ....................................................................................................213

Chapter 17 Incremental Principle of Virtual Work ............................................215


17.1
17.2
17.3
17.4
17.5
17.6
17.7
17.8
17.9
17.10

Incremental Kinematics .............................................................................215


Incremental Stresses ..................................................................................216
Incremental Equilibrium Equation ............................................................217
Incremental Principle of Virtual Work ......................................................218
Incremental Finite-Element Equation .......................................................219
Incremental Contributions from Nonlinear Boundary Conditions ...........220
Effect of Variable Contact .........................................................................221
Interpretation as Newton Iteration.............................................................223
Buckling.....................................................................................................224
Exercises ....................................................................................................226

Chapter 18 Tangent-Modulus Tensors for Thermomechanical


Response of Elastomers ...................................................................227
18.1
18.2
18.3

Introduction................................................................................................227
Compressible Elastomers ..........................................................................227
Incompressible and Near-Incompressible Elastomers ..............................228
18.3.1 Specific Expressions for the Helmholtz Potential ......................230

2003 by CRC CRC Press LLC

0749_Frame_FM Page xviii Wednesday, February 19, 2003 4:53 PM

18.4
18.5
18.6

18.7

18.8

18.9

Stretch Ratio-Based Models: Isothermal Conditions................................231


Extension to Thermohyperelastic Materials..............................................233
Thermomechanics of Damped Elastomers................................................234
18.6.1 Balance of Energy .......................................................................235
18.6.2 Entropy Production Inequality ....................................................235
18.6.3 Dissipation Potential ...................................................................236
18.6.4 Thermal-Field Equation for Damped Elastomers.......................237
Constitutive Model: Potential Functions...................................................238
18.7.1 Helmholtz Free-Energy Density .................................................238
18.7.2 Specific Dissipation Potential .....................................................239
Variational Principles.................................................................................240
18.8.1 Mechanical Equilibrium..............................................................240
18.8.2 Thermal Equilibrium ...................................................................240
Exercises ....................................................................................................241

Chapter 19 Inelastic and Thermoinelastic Materials..........................................243


19.1

19.2

19.3
19.4
19.5
19.6

Plasticity.....................................................................................................243
19.1.1 Kinematics...................................................................................243
19.1.2 Plasticity ......................................................................................243
Thermoplasticity ........................................................................................246
19.2.1 Balance of Energy .......................................................................246
19.2.2 Entropy-Production Inequality ....................................................247
19.2.3 Dissipation Potential ...................................................................248
Thermoinelastic Tangent-Modulus Tensor ................................................249
19.3.1 Example.......................................................................................250
Tangent-Modulus Tensor in Viscoplasticity ..............................................252
Continuum Damage Mechanics ................................................................254
Exercises ....................................................................................................256

Chapter 20 Advanced Numerical Methods ........................................................257


20.1

20.2
20.3

Iterative Triangularization of Perturbed Matrices .....................................257


20.1.1 Introduction .................................................................................257
20.1.2 Notation and Background ...........................................................258
20.1.3 Iteration Scheme..........................................................................259
20.1.4 Heuristic Convergence Argument ...............................................259
20.1.5 Sample Problem ..........................................................................260
Ozawas Method for Incompressible Materials ........................................262
Exercises ....................................................................................................263

Monographs and Texts ........................................................................................265


Articles and Other Sources.................................................................................267

2003 by CRC CRC Press LLC

0749_Frame_C01 Page 1 Wednesday, February 19, 2003 4:55 PM

Mathematical
Foundations: Vectors
and Matrices

1.1 INTRODUCTION
This chapter provides an overview of mathematical relations, which will prove useful
in the subsequent chapters. Chandrashekharaiah and Debnath (1994) provide a more
complete discussion of the concepts introduced here.

1.1.1 RANGE

AND

SUMMATION CONVENTION

Unless otherwise noted, repeated Latin indices imply summation over the range
1 to 3. For example:
3

ai bi =

a b = a b + a b
i i

1 1

2 2

+ a3 b3

(1.1)

i =1

aij bjk = ai1b1k + ai 2 b2 k + ai 3b3k

(1.2)

The repeated index is summed out and, therefore, dummy. The quantity aijbjk in
th
Equation (1.2) has two free indices, i and k (and later will be shown to be the ik
entry of a second-order tensor). Note that Greek indices do not imply summation.
Thus, a b = a1b1 if = 1.

1.1.2 SUBSTITUTION OPERATOR


The quantity, ij, later to be called the Kronecker tensor, has the property that
1
ij =
0

i= j
i j

(1.3)

For example, ijvj = 1 vi, thus illustrating the substitution property.

2003 by CRC CRC Press LLC

0749_Frame_C01 Page 2 Wednesday, February 19, 2003 4:55 PM

Finite Element Analysis: Thermomechanics of Solids

3
v3

v
e3
e2

v2

e1
v1
1
FIGURE 1.1 Rectilinear coordinate system.

1.2 VECTORS
1.2.1 NOTATION
Throughout this and the following chapters, orthogonal coordinate systems will be
used. Figure 1.1 shows such a system, with base vectors e1, e2, and e3. The scalar
product of vector analysis satisfies
e i e j = ij

(1.4)

e k i j and ijk in right-handed order

e i e j = e k i j and ijk not in right-handed order

0 i = j

(1.5)

The vector product satisfies

It is an obvious step to introduce the alternating operator, ijk, also known as the
th
ijk entry of the permutation tensor:

ijk = [e i e j ] e k
1

= 1

2003 by CRC CRC Press LLC

ijk distinct and in right-handed order


ijk distinct but not in right-handed order
ijk not distinct

(1.6)

0749_Frame_C01 Page 3 Wednesday, February 19, 2003 4:55 PM

Mathematical Foundations: Vectors and Matrices

Consider two vectors, v and w. It is convenient to use two different types of


notation. In tensor indicial notation, denoted by (*T), v and w are represented as
v = vi e i

*T)

w = wi e i

(1.7)

Occasionally, base vectors are not displayed, so that v is denoted by vi. By


displaying base vectors, tensor indicial notation is explicit and minimizes confusion
and ambiguity. However, it is also cumbersome.
In this text, the default is matrix-vector (*M) notation, illustrated by
v1

v = v2

v3

*M)

w1

w = w2

w3

(1.8)

It is compact, but also risks confusion by not displaying the underlying base
T
T
vectors. In *M notation, the transposes v and w are also introduced; they are
displayed as row vectors:
*M)

v T = {v1

v2

v3}

w T = {w1

w2

w3}

(1.9)

The scalar product of v and w is written as


v w = (vi e i ) ( w j e j )
= vi w j e i e j
*T)

= vi w j ij
= vi wi

(1.10)

v = vv

(1.11)

The magnitude of v is defined by


*T)

The scalar product of v and w satisfies


*T)

v w = v w cos vw

(1.12)

in which vw is the angle between the vectors v and w. The scalar, or dot, product is
*M)

2003 by CRC CRC Press LLC

v w vTw

(1.13)

0749_Frame_C01 Page 4 Wednesday, February 19, 2003 4:55 PM

Finite Element Analysis: Thermomechanics of Solids

The vector, or cross, product is written as


v w = vi w j e i e j
*T)

= ijk vi w j e k

(1.14)

Additional results on vector notation are presented in the next section, which
introduces matrix notation. Finally, the vector product satisfies
v w = v w sin vw

*T)

(1.15)

and vxw is colinear with n the unit normal vector perpendicular to the plane containing
v and w. The area of the triangle defined by the vectors v and w is given by 12 v w .

1.2.2 GRADIENT, DIVERGENCE,

AND

CURL

The derivative, d/dx, of a scalar with respect to a vector x is defined implicitly by


d =

*M)

d
dx
dx

(1.16)

and it is a row vector whose i entry is d/dxi. In three-dimensional rectangular


coordinates, the gradient and divergence operators are defined by
th

*M)

( )
x
( )
( ) =
y

( )
z

(1.17)

and clearly,
T

*M)

d ( ) = ( )
dx

(1.18)

The gradient of a scalar function satisfies the following integral relation:

dV = n dS

(1.19)

The expression v T will be seen to be a tensor (see Chapter 2). Clearly,


v T = [v1

2003 by CRC CRC Press LLC

v2

v3 ]

(1.20)

0749_Frame_C01 Page 5 Wednesday, February 19, 2003 4:55 PM

Mathematical Foundations: Vectors and Matrices

from which we obtain the integral relation

v dV = nv dS
T

(1.21)

Another important relation is the divergence theorem. Let V denote the volume
of a closed domain, with surface S. Let n denote the exterior surface normal to S,
and let v denote a vector-valued function of x, the position of a given point within
the body. The divergence of v satisfies
*M)

dx dV = n vdS
dv

(1.22)

The curl of vector v, v, is expressed by


( v)i = ijk

v
x j k

(1.23)

which is the conventional cross-product, except that the divergence operator replaces
the first vector. The curl satisfies the curl theorem, analogous to the divergence
theorem (Schey, 1973):

v dV = n v dS

(1.24)

Finally, the reader may verify, with some effort that, for a vector v(X) and a
path X(S) in which S is the length along the path,

v dX(S) = n v dS .

(1.25)

The integral between fixed endpoints is single-valued if it is path-independent,


in which case n v must vanish. However, n is arbitrary since the path is
arbitrary, thus giving the condition for v to have a path-independent integral as
v = 0.

(1.26)

1.3 MATRICES
An n n matrix is simply an array of numbers arranged in rows and columns, also
known as a second-order array. For the matrix A, the entry aij occupies the intersection
th
th
of the i row and the j column. We may also introduce the n 1 first-order array a,
th
T
in which ai denotes the i entry. We likewise refer to the 1 n array, a , as first-order.

2003 by CRC CRC Press LLC

0749_Frame_C01 Page 6 Wednesday, February 19, 2003 4:55 PM

Finite Element Analysis: Thermomechanics of Solids

In the current context, a first-order array is not a vector unless it is associated with
a coordinate system and certain transformation properties, to be introduced shortly.
In the following, all matrices are real unless otherwise noted. Several properties of
first- and second-order arrays are as follows:
The sum of two n n matrices, A and B, is a matrix, C, in which cij = aij + bij.
The product of a matrix, A, and a scalar, q, is a matrix, C, in which cij = qaij.
T
The transpose of a matrix, A, denoted A , is a matrix in which aijT = a ji A is
T
T
called symmetric if A = A , and it is called antisymmetric if A = A .
The product of two matrices, A and B, is the matrix, C, for which
*T)

cij = aik bkj

(1.27)

Consider the following to visualize matrix multiplication. Let the first-order


th
th
array a iT denote the i row of A, while the first-order array bj denotes the j column
of B. Then cij can be written as
*T)

cij = a iT b j

(1.28)

The product of a matrix A and a first-order array c is the first-order array d


th
in which the i entry is di = aijcj.
th
The ij entry of the identity matrix I is ij. Thus, it exhibits ones on the diagonal
positions (i = j) and zeroes off-diagonal (i j). Thus, I is the matrix
counterpart of the substitution operator.
The determinant of A is given by

*T)

det(A) =

1
a a a
6 ijk pqr ip jq kr

(1.29)

Suppose a and b are two non-zero, first-order n 1 arrays. If det(A) = 0, the


matrix A is singular, in which case there is no solution to equations of the form
Aa = b. However, if b = 0, there may be multiple solutions. If det(A) 0, then there
is a unique, nontrivial solution a.
Let A and B be n n nonsingular matrices. The determinant has the following
useful properties:
det(AB) = det(A) det(B)
*M)

det(A T ) = det(A)
det(I) = 1

2003 by CRC CRC Press LLC

(1.30)

0749_Frame_C01 Page 7 Wednesday, February 19, 2003 4:55 PM

Mathematical Foundations: Vectors and Matrices

7
1

If det(A) 0, then A is nonsingular and there exists an inverse matrix, A ,


for which
*M)

AA 1 = A 1A = I

(1.31)

The transpose of a matrix product satisfies


*M)

(AB) T = BT A T

(1.32)

The inverse of a matrix product satisfies


*M)

(AB) 1 = B 1A 1

(1.33)

If c and d are two 3 1 vectors, the vector product c d generates the vector
c d = Cd, in which C is an antisymmetric matrix given by

*M)

C = c3

c2

c3
0
c1

c2

c1

(1.34)

Recalling that c d = ikjckdj, and noting that ikjck denotes the (ij) component
of an antisymmetric tensor, it is immediate that [C]ij = ikjck.
th

If c and d are two vectors, the outer product cd generates the matrix C given by

*M)

c1d1

C = c2 d1

c3 d1

c1d2
c2 d 2
c3 d2

c1d3

c2 d 3

c3 d3

(1.35)

We will see later that C is a second-order tensor if c and d have the transformation
properties of vectors.
An n n matrix A can be decomposed into symmetric and antisymmetric
matrices using
A = As + Aa ,

2003 by CRC CRC Press LLC

As =

1
[A + A T ],
2

Aa =

1
[A A T ]
2

(1.36)

0749_Frame_C01 Page 8 Wednesday, February 19, 2003 4:55 PM

Finite Element Analysis: Thermomechanics of Solids

1.3.1 EIGENVALUES

AND

EIGENVECTORS

In this case, A is again an n n tensor. The eigenvalue equation is


( A j I )x j = 0

(1.37)

The solution for xj is trivial unless A jI is singular, in which event det(A jI) =
0. There are n possible complex roots. If the magnitude of the eigenvectors is set to unity,
they may likewise be determined. As an example, consider
2
A=
1

(1.38)

The equation det(A jI) = 0 is expanded as (2 j) 1, with roots 1,2 = 1, 3, and


2

1
A 1I =
1

1
A 2 I =
1

(1.39)

Note that in each case, the rows are multiples of each other, so that only one row
is independent. We next determine the eigenvectors. It is easily seen that magnitudes
of the eigenvectors are arbitrary. For example, if x1 is an eigenvector, so is 10x1.
T
Accordingly, the magnitudes are arbitrarily set to unity. For x1 = {x11 x12} ,
x11 + x12 = 0

2
2
x11
+ x12
=1

(1.40)

from which we conclude that x1 = {1 1}T / 2 . A parallel argument furnishes


x 2 = {1 1}T / 2 .
If A is symmetric, the eigenvalues and eigenvectors are real and the eigenvectors
are orthogonal to each other: x iT x j = ij . The eigenvalue equations can be stacked
up, as follows.

A[x1 : x 2 :K x n ] = [x1 : x 2 :K x n ]

n1

. (1.41)

With obvious identifications,


AX = X

2003 by CRC CRC Press LLC

(1.42)

0749_Frame_C01 Page 9 Wednesday, February 19, 2003 4:55 PM

Mathematical Foundations: Vectors and Matrices

and X is the modal matrix. Let yij represent the ij entry of Y = X X


th

yij = x iT x j = ij

(1.43)
1

so that Y = I. We can conclude that X is an orthogonal tensor: X = X . Further,


T

X T AX =

A = X XT

(1.44)

and X can be interpreted as representing a rotation from the reference axes to the
principal axes.

1.3.2 COORDINATE TRANSFORMATIONS


Suppose that the vectors v and w are depicted in a second-coordinate system whose
base vectors are denoted by e j . Now, e j can be represented as a linear sum of the
base vectors ei:
e j = q ji e i

*T)

(1.45)

But then e i e j = qij = cos( ij ). It follows that ij = e i e j = (qikek) (qjlel) =


qikqjlkl, so that
qik q jk = qik qkjT
*T)

= ij

In *M) notation, this is written as


QQ T = I

*M)

(1.46)

in which case the matrix Q is called orthogonal. An analogous argument proves that
T
T
T
2
Q Q = I. From Equation (1.30), 1 = det(QQ ) = det(Q)det(Q ) = det (Q). Righthanded rotations satisfy det(Q) = 1, in which case Q is called proper orthogonal.

1.3.3 TRANSFORMATIONS

OF

VECTORS

The vector v is the same as the vector v, except that v is referred to e j , while v is
referred to ei. Now
v = v j e j
*T)

= v j q ji e i
= vi e i

2003 by CRC CRC Press LLC

(1.47)

0749_Frame_C01 Page 10 Wednesday, February 19, 2003 4:55 PM

10

Finite Element Analysis: Thermomechanics of Solids

It follows that vi = v j q ji , and hence


*M)

v = Q T v (a )

th

v = Qv ( b)

(1.48)

in which qij is the ji entry of Q .


We can also state an alternate definition of a vector as a first-order tensor. Let
v be an n 1 array of numbers referring to a coordinate system with base vectors
ei. It is a vector if and only if, upon a rotation of the coordinate system to base
vectors e j , v transforms according to Equation (1.48).
Since ( ddx ) dx is likewise equal to d,

*M)

d = d Q T
dx
dx

(1.49)

for which reason d/dx is called a contravariant vector, while v is properly called a
covariant vector.
Finally, to display the base vectors to which the tensor A is referred (i.e., in
tensor-indicial notation), we introduce the outer product
ei e j

(1.50)

with the matrix-vector counterpart e i e Tj . Now


A = aij e i e j

(1.51)

Note the useful result that


e i e j e k = e i jk
In this notation, given a vector b = bkek,
Ab = aij e i e j bk e k
= aij bk e i e j e k
= aij bk e i jk
= aij b j e i
as expected.

2003 by CRC CRC Press LLC

(1.52)

0749_Frame_C01 Page 11 Wednesday, February 19, 2003 4:55 PM

Mathematical Foundations: Vectors and Matrices

11

1.3.4 ORTHOGONAL CURVILINEAR COORDINATES


The position vector of a point, P, referring to a three-dimensional, rectilinear, coordinate system is expressed in tensor-indicial notation as RP = xiei. The position vector
connecting two sufficiently close points P and Q is given by
R = R P R Q dR x

(1.53)

dR x = dxi e i

(1.54)

dSx = dxi dxi

(1.55)

where

with arc length

Suppose now that the coordinates are transformed to yj coordinates: xi = xi(yj).


The same position vector, now referred to the transformed system, is
3

dR y =

dy g

g = h
dx j dx j

h =

dy dy
dxi
dy

dx j dx j
dy dy

(1.56)

ei

1 dxi
e
h dy i

in which h is called the scale factor. Recall that the use of Greek letters for indices
implies no summation. Clearly, is a unit vector. Conversely, if the transformation
is reversed,
dR y = g i dyi
=

2003 by CRC CRC Press LLC

dyi
g dx
dx j i j

(1.57)

0749_Frame_C01 Page 12 Wednesday, February 19, 2003 4:55 PM

12

Finite Element Analysis: Thermomechanics of Solids

then the consequence is that


ej =

dyi
g =
dx j i

dx

dy

(1.58)

We restrict attention to orthogonal coordinate systems yj, with the property that
T =

(1.59)

The length of the vector dRy is now


dSy = h dyi dyi

(1.60)

Under restriction to orthogonal coordinate systems, the initial base vectors ei can be
expressed in terms of using

e i = Tj e i j
=

1 xi

hi y j j

1 xi x k
e
hi h j y j y j k

(1.61)

and furnishing
1 xi x k
= ik
hi h j y j y j

(1.62)

Also of interest is the volume element; the volume determined by the vector
dRy is given by the vector triple product
dVy = (h1dy1 1 ) [h2 dy2 2 h3 dy3 3 ]
= h1h2 h3 dy1dy2 dy3

(1.63)

and h1h2h3 is known as the Jacobian of the transformation. For cylindrical coordinates
using r, , and z, as shown in Figure 1.2, x1 = rcos , x2 = rsin , and x3 = z. Simple
manipulation furnishes that hr = 1, h = r, hz = 1, and
e r = cos e1 + sin e2

e = sin e1 + cos e2

e z = e3

(1.64)

which, of course, are orthonormal vectors. Also of interest are the relations der =
e d and de = er d.

2003 by CRC CRC Press LLC

0749_Frame_C01 Page 13 Wednesday, February 19, 2003 4:55 PM

Mathematical Foundations: Vectors and Matrices

13

x3

ez
x2
e

er

x1

FIGURE 1.2 Cylindrical coordinate system.

Transformation of the coordinate system from rectilinear to cylindrical coordinates can be viewed as a rotation of the coordinate system through . Thus, if the
vector v is referred to the reference rectilinear system and v is the same vector
referred to a cylindrical coordinate system, then in two dimensions,
cos

Q( ) = sin

v = Q( )v

sin
cos
0

(1.65)

If v is differentiated, for example, with respect to time t, there is a contribution


from the rotation of the coordinate system: for example, if v and are functions of
time t,
d
d
dQ( )
v = Q( ) v +
v
dt
dt
dt
=

dQ( ) T
v +
Q ( )v
t
dt

(1.66)

where the partial derivative implies differentiation with instantaneously held fixed
and
sin
dQ( )
= cos
dt

2003 by CRC CRC Press LLC

cos
sin
0

0
d
0
dt
1

(1.67)

0749_Frame_C01 Page 14 Wednesday, February 19, 2003 4:55 PM

14

Finite Element Analysis: Thermomechanics of Solids

x3
e

er

x2
e

x1

FIGURE 1.3 Spherical coordinate system.

Now
since

dQ ( )
dt

Q T ( ) is an antisymmetric matrix (to be identified later as a tensor)

0=

d
dQ( ) T
dQ( ) T
Q ( ) +
Q ( )
(Q( )Q T ( )) =
dt
dt
dt

(1.68)

In fact,
0

dQ( ) T
Q ( ) = 1
dt

1
0
0

0
d
0
dt
0

(1.69)

It follows that
d

v = v + v
dt
t

(1.70)

in which is the axial vector of .


Referring to Figure 1.3, spherical coordinates r, , and are introduced by the
transformation
x1 = r cos cos

2003 by CRC CRC Press LLC

x 2 = r sin cos

x3 = r sin

(1.71)

0749_Frame_C01 Page 15 Wednesday, February 19, 2003 4:55 PM

Mathematical Foundations: Vectors and Matrices

15

The position vector is given by


r = x1e1 + x 2 e 2 + x3 e 3
= r cos cos e1 + r sin cos e 2 + r sin e 3

(1.72)

Now er has the same direction as the position vector: r = rer. Thus, it follows that
e r = cos cos e1 + sin cos e 2 + sin e 3

(1.73)

Following the general procedure in the preceding paragraphs,


x1
= cos cos
r
x 2
= sin cos
r
x3
= sin
r

x1
= r sin cos

x 2
= r cos cos

x3
=0

x1
= r cos sin

x 2
= r sin sin

x3
= r cos

(1.74)

The differential of the position vector furnishes


dr = dr e r + r cos d e + rd e

(1.75)

e r = cos cos e1 + sin cos e 2 + sin e 3 e1 = cos cos e r sin e sin cos e
e = sin e1 + cos e 2

e 2 = sin cos e r + cos e sin sin e

e = sin [cos e1 + sin e 2 ] + cos e 3

e 3 = sin e r + cos e .

The scale factors are hr = 1, h = rcos , h = r.


Consider a vector v in the rectilinear system, denoted as v when referred to a
spherical coordinate system:
v = v1e1 + v2 e 2 + v3 e 3

v = vr e r + v e + v e .

(1.76)

Eliminating e1, e2, e3 in favor of er, e, e and using *M notation permits writing

v = Q( , )v,

2003 by CRC CRC Press LLC

cos cos

Q( , ) = sin

sin cos

sin cos
cos
sin sin

sin

0 .

cos

(1.77)

0749_Frame_C01 Page 16 Wednesday, February 19, 2003 4:55 PM

16

Finite Element Analysis: Thermomechanics of Solids

Suppose now that v(t), , and are functions of time. As in cylindrical coordinates,
d

v = v + v
dt
t
dQ ( )
dt

where is the axial vector of

Q T ( ). After some manipulation,

sin cos
dQ( )
= cos
dt

sin sin

cos cos

sin

cos sin

cos sin

+
0

cos cos

sin sin
0

d
dt

cos

dt
sin

sin cos

cos

dQ( ) T
Q ( ) = cos
dt

(1.78)

0
d
+ 0
sin
dt
1
0

0
sin

(1.79)

d
dt

1.3.5 GRADIENT OPERATOR


In rectilinear coordinates, let be a scalar-valued function of x: (x), starting with
the chain rule
d =
*T)

dx
xi i

= [ ] dr,

dr = e i dxi

= e i

xi

(1.80)

Clearly, d is a scalar and is unaffected by a coordinate transformation. Suppose


that x = x(y): dr = gi dyi. Observe that
d =
=

dx
xi i

2003 by CRC CRC Press LLC

h1 y

h y

h dy

h dy

(1.81)

0749_Frame_C01 Page 17 Wednesday, February 19, 2003 4:55 PM

Mathematical Foundations: Vectors and Matrices

17

implying the identification


( ) =

(1.82)

For cylindrical coordinates in tensor-indicial notation with er = r, e = , ez = z,

e
+
+ ez
r
z
r

(1.83)

e e z

+
+
r r cos
r

(1.84)

= e r
and in spherical coordinates
= e r

1.3.6 DIVERGENCE

AND

CURL

OF

VECTORS

Under orthogonal transformations, the divergence and curl operators are invariant
and satisfy the divergence and curl theorems, respectively. Unfortunately, the transformation properties of the divergence and curl operators are elaborate. The reader
is referred to texts in continuum mechanics, such as Chung (1988). The development
is given in Appendix I at the end of the chapter. Here, we simply list the results.
Let v be a vector referred to rectilinear coordinates, and let v denote the same vector
referred to orthogonal coordinates. The divergence and curl satisfy
( v ) =

1
h1h2 h3

(h2 h3 v1 ) +
(h3 h1v2 ) +
(h1h2 v3 )

y
y
y
2
3
1

(1.85)

and

( v ) =

1
h1h2 h3

(h3 v3 )
(h2 v2 ) 1
h1
y3
y2

h2
(h3 v3 )
(h1v1 ) 2
y3
y1

+ h3
(h2 v2 )
(h1v1 ) 3
y 2
y1

(1.86)

and in cylindrical coordinates:


( v ) =

2003 by CRC CRC Press LLC

1 (rvr ) 1 v vz
+
+
r r
r
z

(1.87)

0749_Frame_C01 Page 18 Wednesday, February 19, 2003 4:55 PM

18

Finite Element Analysis: Thermomechanics of Solids

and
( v ) =

1 vz (rv )
(rv ) vr
vz vr

re +
e
e r

r
z
z
z
r
r

(1.88)

APPENDIX I: DIVERGENCE AND CURL OF VECTORS


IN ORTHOGONAL CURVILINEAR COORDINATES
DERIVATIVES

OF

BASE VECTORS

In tensor-indicial notation, a vector v can be represented in rectilinear coordinates


as v = vkek. In orthogonal curvilinear coordinates, it is written as v = v =

v gh .

A line segment dr = dxiei transforms to dr = dykgk. Recall that


yl
g =
x k l

ek =

g =

x
e
yk k

y
x k

h = g g

(a.1)

(a.2)

From Equation (a.1),


g
2 x
=
e
y j
yk y j k
=

h ,

2 x y
= h
yk y j x k

The bracketed quantities are known as Cristoffel symbols. From Equations (a.1 and a.2),
dh
dg
=
dy j
dy j

Continuing,

(a.3)


1 g h
=

h y j h y j
y j
=

cj ,

2003 by CRC CRC Press LLC

cj =

1
(1 )
h

(a.4)

0749_Frame_C01 Page 19 Wednesday, February 19, 2003 4:55 PM

Mathematical Foundations: Vectors and Matrices

19

DIVERGENCE
The development that follows is based on the fact that
dv
dv
v = v = tr
= tr
dr
dr

(a.5)

The differential of v is readily seen to be


dv = dv j j + v j d j

(a.6)

First, note that


dv j j =
=

v j
yk

j dyk

j (h dy )

1 v j

v j

h1 y

(h dy )

1 v j

j dr

(a.7)

Similarly,
vjd j = vj
=

v j j

(h dy )

v j j

2003 by CRC CRC Press LLC

v j j
(h dy )
y

dyk

yk

v
j
h

dr
y

c j dr

(a.8)

0749_Frame_C01 Page 20 Wednesday, February 19, 2003 4:55 PM

20

Finite Element Analysis: Thermomechanics of Solids

Consequently,
dv = 1 v j + v c ,
j j
dr h y j

v =

h1 y

+ v j c j

(a.9)

CURL
In rectilinear coordinates, the individual entries of the curl can be expressed as a
th
divergence, as follows. For the i entry,
[ v]i = ijk
=

vk
x j

(i )
w
x j j

w (ji ) = jki vk

= w (i )

(a.10)

Consequently, the curl of v can be written as


w (1)

v = w (2)

w ( 3)

(a.11)

The transformation properties of the curl can be readily induced from Equation (a.9).

1.4 EXERCISES
1. In the tetrahedron shown in Figure 1.4, A1, A2, and A3 denote the areas
of the faces whose normal vectors point in the e1, e2, and e3 directions.
Let A and n denote the area and normal vector of the inclined face,
respectively. Prove that
n=

A
A1
A
e + 2e + 3e
A 1 A 2 A 3

2. Prove that if is a symmetric tensor with entries ij, that

ijk jk = 0,

i = 1, 2, 3.

3. If v and w are n 1 vectors, prove that v w can be written as


v w = Vw

2003 by CRC CRC Press LLC

0749_Frame_C01 Page 21 Wednesday, February 19, 2003 4:55 PM

Mathematical Foundations: Vectors and Matrices

21

A
n
e3
A1

A2
A3

e1

e2

1
FIGURE 1.4 Geometry of a tetrahedron.

in which V is an antisymmetric tensor and v is the axial vector of V.


Derive the expression for V.
4. Find the transposes of the matrices
1
A=
1/ 3

1/ 2

1/ 4

1
B=
1/ 2

1/ 3

1/ 4

(a) Verify that AB BA.


T
T T
(b) Verify that (AB) = B A .
5. Consider a matrix C given by
a
C=
c

b
d

Verify that its inverse is given by


C1 =

1 d

ad bc c

6. For the matrices in Exercise 4, find the inverses and verify that
(AB) 1 = B 1 A 1
7. Consider the matrix
cos
Q=
sin

2003 by CRC CRC Press LLC

sin

cos

0749_Frame_C01 Page 22 Wednesday, February 19, 2003 4:55 PM

22

Finite Element Analysis: Thermomechanics of Solids

Verify that
T
T
(a) QQ = Q Q
T
1
(b) Q = Q
(c) For any 2 1 vector a
Qa = a
[The relation in (c) is general, and Qa represents a rotation of a.]
8. Using the matrix C from Exercise 5, and introducing the vectors (onedimensional arrays)
q
a=
r

s
b=
t

verify that
a T Cb = b T C T a
9. Verify the divergence theorem using the following block, where
x y
v=

x + y
Y

FIGURE 1.5 Test figure for the divergence theorem.

10. For the vector and geometry of Exercise 9, verify that

n vdS = vdV
2003 by CRC CRC Press LLC

0749_Frame_C01 Page 23 Wednesday, February 19, 2003 4:55 PM

Mathematical Foundations: Vectors and Matrices

23

11. Using the geometry of Exercise 9, verify that

n AdS = A dV
T

using
a11 = x + y + x 2 + y 2
a12 = x + y + x 2 y 2
a21 = x + y x 2 y 2
a22 = x y x 2 y 2
12. Obtain the expressions for the gradient, divergence, and curl in spherical
coordinates.

2003 by CRC CRC Press LLC

0749_Frame_C02 Page 25 Wednesday, February 19, 2003 5:00 PM

Mathematical
Foundations: Tensors

2.1 TENSORS
We now consider two n 1 vectors, v and w, and an n n matrix, A, such that v =
Aw. We now make the important assumption that the underlying information in this
relation is preserved under rotation. In particular, simple manipulation furnishes that
v = Qv

)

= QAw
= QAQTQw
= QAQT w .

(2.1)

The square matrix A is now called a second-order tensor if and only if A = QAQ .
Let A and B be second-order n n tensors. The manipulations that follow
1
T
demonstrate that A , (A + B), AB, and A are also tensors.
T

(A T ) = (QAQT )T
T

= QT ATQT

(2.2)

A B = (QAQ T )(QBQ T )
= QA(QQ T )BQ T
= QABQ T

(2.3)

(A + B) = A + B
= QAQ T + QBQ T
= Q(A + B)Q T

(2.4)

A 1 = (QAQT )1
1

= Q T A 1Q 1
= QA 1Q T .

(2.5)

25

2003 by CRC CRC Press LLC

0749_Frame_C02 Page 26 Wednesday, February 19, 2003 5:00 PM

26

Finite Element Analysis: Thermomechanics of Solids

Let x denote an n 1 vector. The outer product, xx , is a second-order tensor since


T

( xx T ) = x x T
= (Qx )(Qx ) T
= Q( xx T )Q T

(2.6)

Next,
d
d
H = .
dx d x
T

d 2 = dx T H dx

(2.7)

However,
dx T H dx = (Q dx ) T H Q dx
= dx T (Q T H Q) dx,

(2.8)

from which we conclude that the Hessian H is a second-order tensor.


Finally, let u be a vector-valued function of x. Then, du = u dx, from which
x

u
du T = dx T
x

(2.9)

and also
T


du T = dx T u T .
x

(2.10)

We conclude that
T

u T
u
= T.
x
x

(2.11)

Furthermore, if du is a vector generated from du by rotation in the opposite


sense from the coordinate axes, then du = Qdu and dx = Qdx. Hence, Q is a tensor.

Also, since du = u dx , it is apparent that


x

u
u T
=Q
Q ,
x
x
from which we conclude that
are tensors.

2003 by CRC CRC Press LLC

u
x

(2.12)

is a tensor. We can similarly show that I and 0

0749_Frame_C02 Page 27 Wednesday, February 19, 2003 5:00 PM

Mathematical Foundations: Tensors

27

2.2 DIVERGENCE, CURL, AND LAPLACIAN


OF A TENSOR
Suppose A is a tensor and b is an arbitrary, spatially constant vector of compatible
dimension. The divergence and curl of a vector have already been defined. For later
purposes, we need to extend the definition of the divergence and the curl to A.

2.2.1 DIVERGENCE
Recall the divergence theorem c Tn dS = Tc dV. Let c = A T b, in which b is an
arbitrary constant vector. Now

b T An dS =
=

(A b)dV
T

A dV b
T

= b T [ T A T ]T dV .

(2.13)

Consequently, we must define the divergence of A such that


*

AndS [ A ] dV = 0.

)

T T

(2.14)

In tensor-indicial notation,

b a n dS b [[ A ] ] dV = 0.
T

i ij j

T T

(2.15)

Application of the divergence theorem to the vector cj = bi aij furnishes


bi
aij [ T A T ]T
x j

] dV = 0.
i

(2.16)

Since b is arbitrary, we conclude that


[ T A T ]i =

aij =
a T.
x j
x j ji

(2.17)

Thus, if we are to write A as a (column) vector, mixing tensor- and matrixvector notation,
A = [ T A T ]T .

2003 by CRC CRC Press LLC

(2.18)

0749_Frame_C02 Page 28 Wednesday, February 19, 2003 5:00 PM

28

Finite Element Analysis: Thermomechanics of Solids

It should be evident that ( ) has different meanings when applied to a tensor


as opposed to a vector.
Suppose A is written in the form
1T
A = T2 ,
T
3

(2.19)

th

in which iT corresponds to the i row of A:[ iT ] j = aij . It is easily seen that


T A T = ( T 1

2.2.2 CURL

AND

T 2

T 3 ).

(2.20)

LAPLACIAN

The curl of vector c satisfies the curl theorem c dV = n c dS. Using tensorindicial notation,

n cdS =

ijk

n j akl bl dS

= ijk n j akl dS bl

c n dSb ,
ij

cij = ijk akl bl .

(2.21)

From the divergence theorem applied to the tensor cij = ijk akl bl,
n AbdS =

( ijk akl bl )dV


j

a
= ijk kl dV bl
x j

= A T dV b ,

if [ A]il = ijk

a .
x j kl

(2.22)

Let lT denote the row vector (array) corresponding to the l row of A: [ lT ]k =


alk . It follows that
th

A = 1

2003 by CRC CRC Press LLC

3 .

(2.23)

0749_Frame_C02 Page 29 Wednesday, February 19, 2003 5:00 PM

Mathematical Foundations: Tensors

29

th
If I is the array for the I column of A, then

A T = 1

3 .

(2.24)

The Laplacian applied to A is defined by


[ 2 A]ij = 2 aij .

(2.25)

It follows, therefore, that


2 A = [ 2 1

2 2

2 3 ].

(2.26)

The vectors i satisfy the Helmholtz decomposition


2 i = ( i ) i .

(2.27)

Observe from the following results that


2 A = ( A T ) [ A T ]T .

(2.28)

An integral theorem for the Laplacian of a tensor is now found as

AdV = (n )AdS n [ A ] dS.


2

T T

(2.29)

2.3 INVARIANTS
Letting A denote a nonsingular, symmetric, 3 3 tensor, the equation det(A l) =
0 can be expanded as

3 I12 + I2 I3 = 0,

(2.30)

in which
I1 = tr(A)

I2 =

1 2
[tr (A) tr(A 2 )]
2

I3 = det(A).

(2.31)

Here, tr (A) = ijaij denotes the trace of A. Equation 2.30 also implies the CayleyHamilton theorem:
A 3 I1A 2 + I2 A I3 I = 0,

2003 by CRC CRC Press LLC

(2.32)

0749_Frame_C02 Page 30 Wednesday, February 19, 2003 5:00 PM

30

Finite Element Analysis: Thermomechanics of Solids

from which
1
I3 = [tr(A 3 ) I1tr(A 2 ) + I2 tr(A)]
3
A

1
3

(2.33)

= I [A I1A + I2 I]
2

The trace of any n n symmetric tensor B is invariant under orthogonal transformations (rotations), such as tr(B) = tr(B), since
a pq pq = q pr qqs ars pq
= ars q pr qqs
= ars rs .
2

(2.34)
2

Likewise, tr(A ) and tr(A ) are invariant since A, A , and A are tensors, thus I1,
I2, and I3 are invariants. Derivatives of invariants are presented in a subsequent section.

2.4 POSITIVE DEFINITENESS


In the finite-element method, an attractive property of some symmetric tensors is
positive definiteness, defined as follows. The symmetric n n tensor A is positivedefinite, written A > 0, if, for all nonvanishing n 1 vectors x, the quadratic product
T
q(A, x) = x Ax > 0. The importance of this property is shown in the following
T
T
example. Let = 12 x Ax x f, in which f is known and A > 0. After some simple
manipulation,
d T d
d 2 = dx T
dx
dx dx
= dx T A dx.

(2.35)

It follows that is a globally convex function that attains a minimum when Ax = f


(d = 0).
The following definition is equivalent to the statement that the symmetric n
n tensor A is positive-definite if and only if its eigenvalues are positive. For the sake
of demonstration,
x T Ax = x T XT x
= y T y,
=

y .
2
i i

2003 by CRC CRC Press LLC

(y = X T x )
(2.36)

0749_Frame_C02 Page 31 Wednesday, February 19, 2003 5:00 PM

Mathematical Foundations: Tensors

31

The last expression can be positive for arbitrary y (arbitrary x) only if i > 0, i =
T
1, 2,, n. The matrix A is semidefinite if x Ax 0, and negative-definite (written
T
T
T
A < 0), if x Ax < 0. If B is a nonsingular tensor, then B B > 0, since q(B B, x) =
T T
T
x B Bx = y y > 0 (in which y = Bx and denotes the quadratic product). If B is

singular, for example if B = yy where y is an n 1 vector, B B is positive-semidefinite


since a nonzero eigenvector x of B can be found for which the quadratic product
T
q(B B, x) vanishes.
Now suppose that B is a nonsingular, antisymmetric tensor. Multiplying through
T
Bxj = j xj with B furnishes
BT Bx j = j BT x j
= j Bx j
= 2j x j .

(2.37)

Since B B is positive-definite, it follows that 2j > 0. Thus, j is imaginary:


2
j = i j using i = 1. Consequently, B2 x j = 2j x j = 2j x j, demonstrating that B
is negative-definite.

2.5 POLAR DECOMPOSITION THEOREM


For an n n matrix B, B B > 0. If the modal matrix of B is denoted by Xb, we can
write
T

BT B = X Tb b X b
1

= X Tb ( b ) 2 YY T ( b ) 2 X b
T

1
1

= X Tb ( b ) 2 Y X Tb ( b ) 2 Y ,

(2.38a)

in which Y is an (unknown) orthogonal tensor. In general, we can write


1

B = Y T ( b ) 2 X b .

(2.38b)

To justify Equation 2.38b, we introduce the square root

1
2
b

BT B = X Tb X b ,

2003 by CRC CRC Press LLC

1
b2 =

BT B using
.

. ,

(2.38c)

0749_Frame_C02 Page 32 Wednesday, February 19, 2003 5:00 PM

32

Finite Element Analysis: Thermomechanics of Solids

in which the positive square roots are used. It is easy to verify that ( BT B ) 2 = B and
that BT B > 0 . Note that

B B T B

1
2

B B T B

1
2

B B

1
2

T
[B B]

B B

= I.

1
2

(2.38d)

Thus, B( BT B ) 1/2 is an orthogonal tensor, called, for example, Z, and hence we


can write
B = Z BT B
1

= ZX Tb b2 X b .

(2.38e)

Finally, noting that (ZX Tb )(ZX Tb ) T = Z(X Tb X b )Z T = ZZ T = I , we make the identification Y T = ZX Tb in Equation 2.38b. Equation 2.38 plays a major role in the
interpretation of strain tensors, a concept that is introduced in subsequent chapters.

2.6 KRONECKER PRODUCTS ON TENSORS


2.6.1 VEC OPERATOR

AND THE

KRONECKER PRODUCT

Let A be an n n (second-order) tensor. Kronecker product notation (Graham, 1981)


reduces A to a first-order n 1 tensor (vector), as follows.
VEC(A) = {a11

a21

a31

an,n1

ann}T .

(2.39)

The inverse VEC operator, IVEC, is introduced by the obvious relation


IVEC(VEC(A)) = A. The Kronecker product of an n m matrix A and an r s
matrix B generates an nr ms matrix, as follows.
a11B

a21B

AB= .

a B
n1

a12 B

a1m B

. .

anm B

(2.40)

If m, n, r, and s are equal to n, and if A and B are tensors, then A B


2
2
transforms as a second-order n n tensor in a sense that is explained subsequently.

2003 by CRC CRC Press LLC

0749_Frame_C02 Page 33 Wednesday, February 19, 2003 5:00 PM

Mathematical Foundations: Tensors

33

Equation 2.40 implies that the n 1 Kronecker product of two n 1 vectors a


and b is written as
2

a1b
a2 b

ab = . .
.

an b

2.6.2 FUNDAMENTAL RELATIONS

FOR

(2.41)

KRONECKER PRODUCTS

Six basic relations are introduced, followed by a number of subsidiary relations.


The proofs of the first five relations are based on Graham (1981).
th
Relation 1: Let A denote an n m real matrix, with entry aij in the i row and
th
j column. Let I = (j 1)n + i and J = (i 1)m + j. Let Unm denote the nm nm
matrix, independent of A, satisfying
1,
u JK =
0,

K=I
KI

1,
u IK =
0,

K=J
KJ

(2.42)

Then,
VEC(A T ) = U nm VEC(A).

(2.43)

Note that uJK = uJI = 1 and uIK = uIJ = 1, with all other entries vanishing. Hence
if m = n, then uJI = uIJ, so that Unm is symmetric if m = n.
Relation 2: If A and B are second-order n n tensors, then
tr(AB) = VEC T (A T )VEC(B).

(2.44)

Relation 3: If In denotes the n n identity matrix, and if B denotes an n n


tensor, then
I n BT = (I n B) T .

(2.45)

Relation 4: Let A, B, C, and D, respectively, denote m n, r s, n p, and s


q matrices. Then,
(A B) (C D) = AC BD.

2003 by CRC CRC Press LLC

(2.46)

0749_Frame_C02 Page 34 Wednesday, February 19, 2003 5:00 PM

34

Finite Element Analysis: Thermomechanics of Solids

Relation 5: If A, B, and C are n m, m r, and r s matrices, then


VEC(ACB) = BT AVEC(C).

(2.47)

Relation 6: If a and b are n 1 vectors, then


a b = VEC([ab T ]T ).

(2.48)

As proof of Relation 6, if I = ( j 1)n + i, then the I entry of VEC(ba ) is


th
T
T T
bi aj. It is also the I entry of a b. Hence, a b = VEC(ba ) = VEC([ab ] ).
T
Symmetry of Unn was established in Relation 1. Note that VEC(A) = Unn VEC(A ) =
2
U nn VEC(A) if A is n n, and hence the matrix Unn satisfies
th

1
.
U nn = U Tnn = U nn

U 2nn = I n2

(2.49)

Unn is hereafter called the permutation tensor for n n matrices. If A is symmetric, then
(U nn I n2 ) VEC(A) = 0. If A is antisymmetric, then (Unn + Inn)VEC(A) = 0.
If A and B are second-order n n tensors, then
tr(AB) = VEC T (B)VEC(A T )
= VEC T (B)U nn VEC(A)
= [U nn VEC(B)]T VEC(A)
= VEC T (BT )VEC(A)
= tr(BA),

(2.50)

thereby recovering a well-known relation.


If In is the n n identity tensor and in = VEC(In), VEC(A) = In Ain since
2
VEC(A) = VEC(AIn). If Inn is the identity tensor in n -dimensional space, then In In =
Inn since VEC(I n ) = VEC(I n I n ) = In InVEC(In). Now, in = Ini, thus In In = I n 2 .
If A, B, and C denote n n tensors, then
VEC(ACBT ) = I n AVEC(CBT )
= (I n A)(B I n )VEC(C)
= B AVEC(C).

(2.51)

However, by a parallel argument,


VEC[(ACBT )T ] = VEC(BCT A T )
= A BVEC(CT )
= A BU n2 VEC(C).

2003 by CRC CRC Press LLC

(2.52)

0749_Frame_C02 Page 35 Wednesday, February 19, 2003 5:00 PM

Mathematical Foundations: Tensors

35

The permutation tensor arises in the relation


VEC[(ACBT )T ] = U n2 VEC(ACBT ).

(2.53)

Consequently, if C is arbitrary,
U n2 B AVEC(C) = A BU n2 VEC(C),

(2.54)

and, upon using the relation U 2 = U 21 , we obtain an important result:


n
n
B A = U n 2 A BU n 2 .

(2.55)

If A and B are nonsingular n n tensors, then


(A B)(A 1 B 1 ) = AA 1 BB 1
= In In
= I n2 .

(2.56)

The Kronecker sum and difference appear frequently (for example, in control
theory) and are defined as follows:
A B = A In + In B

A  B = A I n I n B.

(2.57)

The Kronecker sum and difference of two n n tensors are n n tensors, as


explained in the following section.
2

2.6.3 EIGENSTRUCTURES

OF

KRONECKER PRODUCTS

Let j and k denote the eigenvalues of A and B, and let yj and zk denote the
corresponding eigenvectors. The Kronecker product, sum, and difference have the
following eigenstructures:
expression

jk th eigenvalue

jk th eigenvector

AB

j k

y j zk

AB

j + k

y j zk

j k

yj zk

2003 by CRC CRC Press LLC

(2.58)

0749_Frame_C02 Page 36 Wednesday, February 19, 2003 5:00 PM

36

Finite Element Analysis: Thermomechanics of Solids

As proof,

j y j k zk = j k y j zk
= Ay j Bz k
= (A B)(y j z k ).

(2.59)

Now, the eigenvalues of A In are 1 j, while the eigenvectors are yj wk,


in which wk is an arbitrary unit vector (eigenvector of In). The corresponding
quantities for In B are k 1 and vj zk , in which vj is an arbitrary eigenvector
of In. Upon selecting wk = zk and vj = yj, the Kronecker sum has eigenvalues j +
k and eigenvectors yj zk .

2.6.4 KRONECKER FORM

OF

QUADRATIC PRODUCTS

Let R be a second-order n n tensor. The quadratic product a Rb is easily derived:


if r = VEC(R), then
T

a T Rb = tr[ba T R]
= VEC T ([ba T ]T )VEC(R)
= VEC T (ab T )VEC(R)
= b T a T r.

(2.60)

2.6.5 KRONECKER PRODUCT OPERATORS


FOR FOURTH-ORDER TENSORS
Let A and B be second-order n n tensors, and let C be a fourth-order n n n n
tensor. Suppose that A = CB, which is equivalent to aij = cijklbkl in which the range
of i, j, k, and l is (1, n). The TEN22 operator is introduced implicitly using
VEC(A) = TEN 22(C)VEC(B).

(2.61)

Note that
TEN 22(ACB)VEC(D) = VEC(ACBD)
= I n A VEC(CBD)
= I n A TEN 22(C) VEC(BD)
= I n A TEN 22(C)I n B VEC(D),

2003 by CRC CRC Press LLC

(2.62)

0749_Frame_C02 Page 37 Wednesday, February 19, 2003 5:00 PM

Mathematical Foundations: Tensors

37
1

hence, TEN22(ACB) = In ATEN22(C)In B. Upon writing B = C A, it is obvious


1
that VEC(B) = TEN22(C )VEC(A). However, TEN22(C)VEC(B) = VEC(A), thus
1
1
1
VEC(B) = [TEN22(C)] VEC(A). We conclude that TEN22(C ) = TEN22 (C).
T
T

Furthermore, by writing A = CB , it is also obvious that Un a = TEN22(C)Unb,


) = U TEN22(C) U . The inverse of the TEN22 operator is introduced
thus TEN22(C
n2
n2
using the relation ITEN22(TEN22(C)) = C.

2.6.6 TRANSFORMATION PROPERTIES

OF

VEC

AND

TEN22

Suppose that A and B are true second-order n n tensors and C is a fourth-order


n n n n tensor such that A = CB. All are referred to a coordinate system
denoted as Y. Let the unitary matrix (tensor) Qn represent a rotation that gives rise
to a coordinate system Y. Let A, B, and C denote the counterparts of A, B, and C.
T
Now, since A = Qn AQ n,
VEC(A ) = Q QVEC(A).
1

(2.63)
1

However, note that (Q Q) = Q Q = Q Q = (Q Q) . Hence,


2
Q Q is a unitary matrix (tensor) in an n vector space. However, not all rotations
2
in n dimensional space can be expressed in the form Q Q. It follows that VEC(A)
2
transforms as an n 1 vector under rotations of the form Q Q.
Now write A = C B, from which
T

Q Q VEC(A) = TEN 22(C)Q QVEC(B).

(2.64a)

TEN 22(C) = Q Q TEN 22(C)(Q Q) T ,

(2.64b)

It follows that

thus TEN22(C) transforms a second-order n n tensor under rotations of the form


Q Q.
Finally, letting Ca and Cb denote third-order n n n tensors, respectively,
thereby satisfying relations of the form A = Cab and b = Cb A, it is readily shown
that TEN21(Ca) and TEN12(Cb) satisfy
2

TEN 21(Ca ) = Q QTEN 21(C a )Q T

n2 n

TEN12(Cb ) = QTEN12(C b )Q T Q T

n n2 ,

which we call tensors of order (2,1) and (1,2), respectively.

2003 by CRC CRC Press LLC

(2.65)

0749_Frame_C02 Page 38 Wednesday, February 19, 2003 5:00 PM

38

Finite Element Analysis: Thermomechanics of Solids

2.6.7 KRONECKER PRODUCT FUNCTIONS


FOR TENSOR OUTER PRODUCTS
Tensor outer products are commonly used in continuum mechanics. For example,
Hookes Law in isotropic linear elasticity with coefficients and can be written as
E
Tij = ( ik jl + il jk ) Ekl +
ij kl kl

i, j = 1, 2, 3,

(2.66)

in which Tij and Eij are entries of the (small-deformation) stress and strain tensors
denoted by T and E. Here, ij denotes the substitution (Kronecker) tensor. Equation 2.66
exhibits three tensor outer products of the identity (Kronecker) tensor I: ik jl, il jk,
and ijkl. In general, let A and B be two nonsingular n n second-order tensors
with entries a ij and bij; let a = VEC(A) and b = VEC(B). There are 24 permutations
of the indices ijkl corresponding to outer products of tensors A and B. Recalling the
definitions of the Kronecker product, we introduce three basic Kronecker-product
functions:
C1 (A, B) = ab T

C 2 (A, B) = A B

C 3 (A, B) = A BUn2 .

(2.67)

Twenty-four outer product-Kronecker product pairs are obtained as follows:


TEN 22( aij bkl ) = C1 (A, B)

TEN 22(bij akl ) = C1 (B, A)

TEN 22( a ji bkl ) = C1 (A T , B)

TEN 22(bji akl ) = C1 (BT , A)

TEN 22( aij blk ) = C1 (A, BT )

TEN 22(bij alk ) = C1 (B, A T )

TEN 22( a ji blk ) = C1 (A T , BT )

TEN 22(bji alk ) = C1 (BT , A T )

TEN 22( aik bjl ) = C2 (B, A)

TEN 22(bik a jl ) = C2 (A, B)

TEN 22( aki bjl ) = C2 (B, A T )

TEN 22(bki a jl ) = C2 (A, BT )

TEN 22( aik blj ) = C2 (BT , A)

TEN 22(bik alj ) = C2 (A T , B)

TEN 22( aki blj ) = C2 (BT , A T )

TEN 22(bki alj ) = C2 (A T , BT )

TEN 22( ail bjk ) = C3 (B, A)

TEN 22(bil a jk ) = C3 (A, B)

TEN 22( ali bjk ) = C3 (B, A T )

TEN 22(bli a jk ) = C3 (A, BT )

TEN 22( ail bkj ) = C3 (BT , A)

TEN 22(bil akj ) = C3 (A T , B)

TEN 22( ali bkj ) = C3 (BT , A T )

TEN 22(bli akj ) = C3 (A T , BT )

2003 by CRC CRC Press LLC

(2.68)

0749_Frame_C02 Page 39 Wednesday, February 19, 2003 5:00 PM

Mathematical Foundations: Tensors

39

With t = VEC(T) and e = VEC(E), and noting that U9e = e (since E is symmetric),
we now restate Equation 2.66 as
t = [ [C2 (I, I) + C3 ( I, I)] + C1 (I, I)]e

(2.69)

The proof is presented for several of the relations in Equation 2.68. We introduce
tensors R and S with entries rij and sij. Also, let s = VEC(S) and r = VEC(R).
a.) Suppose that sij = aij bkl rkl. However, aij bkl rkl = aij b lk rkl , in which case b kl
th
T
T
is the kl entry of B . It follows that S = tr(B R)A. Hence,
T

s = VEC T [(BT ) T ]VEC(R)a


= VEC T (B)VEC(R)a
= ab T r
= C1 (A, B)r.

(2.70)

Since s = TEN22(aij bkl)r, it follows that TEN 22( aij bkl ) = C1 (A, B), as
shown in Equation 2.68.
T
T
b.) Suppose that sij = aik bjl rkl. However, aik bjl rkl = aik rkl b lj, thus S = ARB . Now
s = VEC(ARBT )
= I AVEC(RBT )
= I AB Ir
= B Ar
= C 2 (B, A)r,

(2.71)

as shown in Equation 2.68.


T
T T
c.) Suppose sij = ail bjk rkl. However, ail bjk rkl = ail r lk bkj, thus S = AR B . Now
s = VEC(AR T BT )
= I AB IVEC(R T )
= B AU 9 r
= C3 (B, A)r,
as shown in Equation 2.68.

2003 by CRC CRC Press LLC

(2.72)

0749_Frame_C02 Page 40 Wednesday, February 19, 2003 5:00 PM

40

Finite Element Analysis: Thermomechanics of Solids

2.6.8 KRONECKER EXPRESSIONS FOR SYMMETRY CLASSES


IN FOURTH-ORDER TENSORS
Let C denote a fourth-order tensor with entries cijkl. If the entries observe
cijkl = c jikl

( a)

cijkl = cijlk

(b) ,

cijkl = cklij

(c )

(2.73)

then C is said to be totally symmetric. A fourth-order tensor C satisfying Equation


2.73a but not 2.73b or c is called symmetric.
Kronecker-product conditions for symmetry are now stated. The fourth-order
tensor C is totally symmetric if and only if
TEN22(C) = TEN22T (C)

( a)

Un2 TEN22(C) = TEN22(C)

(b) .

TEN22(C)Un2 = TEN22(C)

(c )

(2.74)

Equation 2.74a is equivalent to symmetry with respect to exchange of ij and kl


in C. Total symmetry also implies that, for any second-order n n tensor B, the
corresponding tensor A = CB is symmetric. Thus, if a = VEC(A) and b = VEC(B),
then a = TEN22(C)b. However, U 2 a = TEN 22(C)b . Multiplying through the later
n
expression with Un2 implies Equation 2.74b. For any n n tensor A, the tensor B =
1
1
1
C A is symmetric. It follows that b = TEN22(C )a = TEN22 (C)a, and Un2 b =
1
1
1
1
TEN22 Ca. Thus, TEN22(C ) = Un2 TEN22 (C). Also, TEN22(C) = [ Un2TEN22
1
(C)] = TEN22(C) Un2 . We now draw the immediate conclusion that Un2 TEN22(C)
Un2 = TEN22(C) if C is totally symmetric.
We next prove the following:
C

is totally symmetric if C is totally symmetric.

(2.75)

1
1
Note that TEN22(C) Un2 = TEN22(C) implies that Un2 TEN22(C ) = TEN22(C ),
1
1
while Un2 TEN22(C) = TEN22(C) implies that TEN22(C ) Un2 = TEN22(C ).
Finally, we prove the following: for a nonsingular n n tensor G,
T

GCG is totally symmetric if C is totally symmetric.


T

(2.76)

Equation 2.76 implies that TEN22(GCG ) = I G TEN22(C)I G , so that


T
TEN22(GCG ) is certainly symmetric. Next, consider whether A given by
A = GCG T B

2003 by CRC CRC Press LLC

(2.77)

0749_Frame_C02 Page 41 Wednesday, February 19, 2003 5:00 PM

Mathematical Foundations: Tensors

41

is symmetric, in which case B is a second-order, nonsingular n n tensor. However,


we can write
G 1 A G T = CG 1 B G T .
1

(2.78)

Now G AG is symmetric since C is totally symmetric, and therefore A is


symmetric. Next, consider whether B given by the following is symmetric:
B = G T C 1G 1 A

(2.79)

G T B G = C 1G 1 A G.

(2.80)

However, we can write

Since C is totally symmetric, it follows that G BG is symmetric, and hence


T
B is symmetric. We conclude that GCG is totally symmetric.

2.6.9 DIFFERENTIALS

TENSOR INVARIANTS

OF

Let A be a symmetric 3 3 tensor, with invariants I1(A), I2(A), and I3(A). For a
scalar-valued function f(A),
df (A) =

f
f

da = tr
dA ,
A
aij ij

f
f
.

=
A ij aij

(2.81)

However, with a = VEC(A), we can also write


f T
df (A) = VEC T VEC( dA)
A
=

f
da.
a

(2.82)

Taking this further,


I1
T
(i a ) = i T
=
a a
I2
1 T 2
(i a) a T a
=
a a 2

= I1i T a T

2003 by CRC CRC Press LLC

(2.83)

0749_Frame_C02 Page 42 Wednesday, February 19, 2003 5:00 PM

42

Finite Element Analysis: Thermomechanics of Solids

and
dI3 = tr(A 2 dA) I1tr(A dA) + I2 dA
= tr(A 1dA / I3 )

(2.84)

so that
I3
= VEC(A 1 ) / I3
a

(2.85)

2.7 EXERCISES
1. Given a symmetric n n tensor , prove that
tr( tr( ) In /n) = 0.
2. Prove that if is a symmetric tensor with entries ij,

ijk jk = 0,

i = 1, 2, 3.

3. Verify using 2 2 tensors that


tr(AB) = tr(BA) .
4. Express I3 as a function of I1 and I2.
5. Using 2 2 tensors and 2 1 vectors, verify the six relations given for
Kronecker products.
6. Write out the 9 9 quantity TEN22(C) in Equation 2.66.
7. Using a 2 2 tensor A, write out the differential of ln(A).

2003 by CRC CRC Press LLC

0749_Frmae_C03 Page 43 Wednesday, February 19, 2003 5:01 PM

Introduction to
Variational and
Numerical Methods

3.1 INTRODUCTION TO VARIATIONAL METHODS


Let u(x) be a vector-valued function of position vector x, and consider a vectorvalued function F(u(x), u(x),x), in which u(x) = u/x. Furthermore, let v(x) be a
function such that v(x) = 0 when u(x) = 0 and v(x) = 0 when u(x) = 0, but which
is otherwise arbitrary. The differential d F measures how much F changes if x
changes. The variation F measures how much F changes if u and u change at
fixed x. Following Ewing, we introduce the vector-valued function (e: F) as follows
(Ewing, 1985):
(e:F) = F(u( x ) + ev( x ), u( x ) + ev( x ), x ) F(u( x ), u( x ), x )

(3.1)

The variation F is defined by


d
F = e
,
de e=0

(3.2)

with x fixed. Elementary manipulation demonstrates that

F =
F

F
F

ev + tr
ev ,
u
u

(3.3)

in which u ev = u ev ij . If F = u, then F = u = ev. If F = u , then F = u =


ij
ev. This suggests the form

F=

F
F

u + tr
u .
u

(3.4)

The variational operator exhibits five important properties:


1. (.) commutes with linear differential operators and integrals. For example, if S denotes a prescribed contour of integration:

( )dS = ( )dS.

(3.5)

43

2003 by CRC CRC Press LLC

0749_Frmae_C03 Page 44 Wednesday, February 19, 2003 5:01 PM

44

Finite Element Analysis: Thermomechanics of Solids

2. ( f ) vanishes when its argument f is prescribed.


3. (.) satisfies the same operational rules as d(.). For example, if the scalars
q and r are both subject to variation, then

(qr ) = q (r ) + (q)r.

(3.6)

4. If f is a prescribed function of (scalar) x, and if u (x) is subject to variation,


then

( fu) = fu.

(3.7)

5. Other than for number 2, the variation is arbitrary. For example, for two
T
vectors v and w, v d w = 0 implies that v and w are orthogonal to each
T
other. However, v w implies that v = 0, since only the zero vector can
be orthogonal to an arbitrary vector.
As a simple example, Figure 3.1 depicts a rod of length L, cross-sectional area
A, and elastic modulus E. At x = 0, the rod is built in, while at x = L, the tensile
force P is applied. Inertia is neglected. The governing equations are in terms of
displacement u, stress S, and (linear) strain E:
du
dx

strain-displacement

E=

stress-strain

S = EE

equilibrium

d
=0
dx

(3.8)

Combining the equations furnishes


EA

d 2u
= 0.
dx 2

(3.9)

The following steps serve to derive a variational equation that is equivalent to the
differential equation and endpoint conditions (boundary conditions and constraints).

E,A
P
L

FIGURE 3.1 Rod under uniaxial tension.

2003 by CRC CRC Press LLC

0749_Frmae_C03 Page 45 Wednesday, February 19, 2003 5:01 PM

Introduction to Variational and Numerical Methods

45

Step 1: Multiply by the variation of the variable to be determined (u) and


integrate over the domain.

u EA

d 2u
Adx = 0.
dx 2

(3.10)

Differential equations to be satisfied at every point in the domain are replaced


with an integral equation whose integrand includes an arbitrary function.
Step 2: Integrate by parts, as needed, to render the argument in the domain
integral positive definite.

d
dx

uEA du du EA du dx = 0

dx dx
dx

(3.11)

However, the first term is the integral of a derivative, so that

du
du
du
dx EA dx dx = uEA dx .
0

(3.12)

Step 3: Identify the primary and secondary variables. The primary variable is
present in the endpoint terms (rhs) under the variational symbol, l, and is
u. The conjugate secondary variable is EA du
.
dx
Step 4: Satisfy the constraints and boundary conditions. At x = 0, u is prescribed, thus u = 0. At x = L, the load P = EA du
is prescribed. Also, note
dx
du = ( 1 EA( du )2 ) .
that ( du
)
E
A
2
dx
dx
dx
Step 5: Form the variational equation; the equations and boundary conditions
are consolidated into one integral equation, F = 0, where
F=

1 du
EA
dx Pu( L).
2 dx

(3.13)

th

The j variation of a vector-valued quantity F is defined by


d j
jF = e j j .
de e=0

(3.14)

It follows that u = 0 and u = 0. By restricting F to a scalar-valued function


F and x to reduce to x, we obtain
2

2 F = { u T

u
u T}H
,
u

T
F

u u
H=
T

u u F

and H is known as the Hessian matrix.

2003 by CRC CRC Press LLC


F

u u
,
T

F

u u

(3.15)

0749_Frmae_C03 Page 46 Wednesday, February 19, 2003 5:01 PM

46

Finite Element Analysis: Thermomechanics of Solids

Now consider G given by


G=

F(x, u(x), u(x))dV + h (x)u(x) dS,


T

(3.16)

in which V again denotes the volume of a domain and S denotes its surface area. In
addition, h is a prescribed (known) function on S. G is called a functional since it
generates a number for every function u(x). We first concentrate on a three-dimensional,
rectangular coordinate system and suppose that G = 0, as in the Principle of
Stationary Potential Energy in elasticity. Note that
F

F

u = tr
u +
F u.

u
x u
x u

(3.17)

The first and last terms in Equation 3.17 can be recognized as divergences of
vectors. We now invoke the divergence theorem to obtain
0 = G
F

u u + tr u u dV + h (x)u(x) dS

u x u u dV + n

F
udS + h T ( x ) u( x ) dS.
u

(3.18)

For suitable continuity properties of u, arbitrariness of u implies that G = 0


is equivalent to the following Euler equation, boundary conditions, and constraints
(the latter two are not uniquely determined by the variational principle):
F F

= 0 T.
u x u
u( x ) prescribed

T F
T
T
n u + h1 ( x ) = 0 on S

(3.19)
x on S1
x on S S1

Let D > 0 denote a second-order tensor, and let denote a vector that is a
nonlinear function of a second vector u, which is subject to variation. The function
F = 12 T D satisfies

F = uT

T
D
u T

2 F = uT

T
T

D
u + u T T uD .
T
u
u
u u
(3.20)

Despite the fact that D > 0, in the current nonlinear example, the specific vector

u satisfying F = 0 may correspond to a stationary point rather than a minimum.

2003 by CRC CRC Press LLC

0749_Frmae_C03 Page 47 Wednesday, February 19, 2003 5:01 PM

Introduction to Variational and Numerical Methods

47

3.2 NEWTON ITERATION AND ARC-LENGTH METHODS


3.2.1 NEWTON ITERATION
Letting f and x denote scalars, consider the nonlinear algebraic equation f(x; ) = 0,
in which is a parameter we will call the load intensity. Such equations are often
solved numerically by a two-track process: the load intensity is increased progressively using small increments. At each increment, the unknown x is computed using
th
an iteration procedure. Suppose that at the n increment of , an accurate solution
is achieved as xn. Further suppose for simplicitys sake that xn is close to the actual
st
(0)
solution xn+1 for the (n + 1) increment of . Using x = xn as the starting value,
Newton iteration provides iterates according to the scheme
1

df
x ( j +1) = x ( j )
f ( x ( j ) ).
dx x ( j )

(3.21)

Let n+1,j denote the increment xn( +j +11) xn( +j )1. Then, to first-order in the Taylor series
1

df 1
df
n +1,j n +1,j1 = f(x(j) )
f(x(j1) )
dx x(j1)

dx x(j)
1

df
f(x(j) ) f(x(j1) )
dx x(j)
1

df
df
n +1,j1 + 0 2
(j)
dx
dx
x x(j)
n +1,j1 + 0 2

(3.22)

in which 0 refers to second-order terms in increments. It follows that n+1,j 0 . For


this reason, Newton iteration is said to converge quadratically (presumably to the
correct solution if the initial iterate is sufficiently close). When the iteration scheme
converges to the solution, the load intensity is incremented again. Consider f (x) =
(0)
(0)
(x 1)2. If x = 1/2, the iterates are 1/2, 3/4, 7/8, and 15/16. If x = 2, the iterates
are 3/2, 5/4, 9/8, and 17/16. In both cases, the error is halved in each iteration.
The nonlinear, finite element poses nonlinear, algebraic equations of the form
2

u T [(u) v] = 0,

(3.23)

in which u and are n 1 vectors, v is a constant n 1 unit vector, and represents


st
load intensity. The Newton iteration scheme provides the ( j + 1) iterate for un+1 as



n +1,j =
u u(nj+)1

2003 by CRC CRC Press LLC

[ ( u ) v ] ,
( j)
n +1

j+1

u(jn ++11) u(nj+)1 = n +1,j .

(3.24)

0749_Frmae_C03 Page 48 Wednesday, February 19, 2003 5:01 PM

48

Finite Element Analysis: Thermomechanics of Solids

in which, for example, the initial iterate is un. One can avoid the use of an explicit
matrix inverse by solving the linear system

( j)
u ( j ) n+1, j = u n+1 v j +1
un +1

3.2.2 CRITICAL POINTS

AND THE

u (nj++11) = u (nj+)1 + n+1, j .

(3.25)

ARC-LENGTH METHOD

A point at which the Jacobian matrix J = u is singular is called a critical point,


and corresponds to important phenomena such as buckling. There often is good
reason to attempt to continue calculations through critical points, such as to compute
a postbuckled configuration. Arc-length methods are suitable for doing so. Here, we
present a version with a simple eigenstructure.
Suppose that the change in load intensity is regarded as a variable. Introduce
th
the constraint on the size of the increment for the n load step:

(u, ) = 2 [ n ] + v T (u u n ) 2 = 0,

(3.26)

in which is interpreted as the arc length in n + 1 dimensional space of u and .


Also, > 0. Now,
2

(u) v 0

= .

(3.27)

Newton iteration now is expressed as

( )
(

u ( j ) ( j ) v
u (nj++11) u (nj+)1
n +1
n +1

,
J
=

( j)
( j +1) ( j )

u
,
n+1

n +1
n +1

J
J =
v T

v
.
2

(3.28)

An advantage is gained if J can be made nonsingular even though J is singular.


2
> 0. Then
Suppose that J is symmetric and we can choose such that J + v v /
J admits the triangularization
J + v v T / 2
J =
0T

v / I
T
v /

0
.

(3.29)

2
2
). Ideally, is chosen to maximize
The determinant of J is now det(J + v v /
det (J).

2003 by CRC CRC Press LLC

0749_Frmae_C03 Page 49 Wednesday, February 19, 2003 5:01 PM

Introduction to Variational and Numerical Methods

49

3.3 EXERCISES
1. Directly apply variational calculus to F, given by
F=

1
du
EA dx Pu( L )
dx
2

to verify that F = 0 gives rise to the Euler equation


EA

d 2u
=0
dx 2

What endpoint conditions (not unique) are compatible with F = 0?


2. The governing equation for an Euler-Bernoulli beam in Figure 3.2 is

EI

d 4w
=0
dx 4

in which w is the vertical displacement of the neutral (centroidal) axis.


The shear force V and the bending moment M satisfy
M = EI

d 2w
dx 2

V = EI

d 3w
dx 3

Using integration by parts twice, obtain the function F such that F = 0


is equivalent to the foregoing differential equation together with the
boundary conditions for a cantilevered beam of length L:
w(0) = w (0) = 0
Z

M ( L) = 0, V( L) = V0 .
neutral axis
v0

V
E,I

x
L
FIGURE 3.2 Cantilevered beam.

2003 by CRC CRC Press LLC

0749_Frame_C04 Page 51 Wednesday, February 19, 2003 5:33 PM

Kinematics
of Deformation

The current chapter provides a review of the mathematics for describing deformation
of continua. A more complete account is given, for example, in Chandrasekharaiah
and Debnath (1994).

4.1 KINEMATICS
4.1.1 DISPLACEMENT
In finite-element analysis for finite deformation, it is necessary to carefully distinguish between the current (or deformed) configuration (i.e., at the current time or
load step) and a reference configuration, which is usually considered strain-free.
Here, both configurations are referred to the same orthogonal coordinate system
characterized by the base vectors e1, e2, e3 (see Figure 1.1 in Chapter 1). Consider a
body with volume V and surface S in the current configuration. The particle P
occupies a position represented by the position vector x, and experiences (empirical)
temperature T. In the corresponding undeformed configuration, the position of P is
described by X, and the temperature has the value T0 independent of X. It is now
assumed that x is a function of X and t and that T is also a function of X and t. The
relations are written as x(X, t) and T(X, t), and it is assumed that x and T are
continuously differentiable in X and t through whatever order needed in the subsequent development.

deformed

x
e2
X

undeformed

e1
FIGURE 4.1 Position vectors in deformed and undeformed configurations.

51

2003 by CRC CRC Press LLC

0749_Frame_C04 Page 52 Wednesday, February 19, 2003 5:33 PM

52

Finite Element Analysis: Thermomechanics of Solids

Q'
ds

P'

Q *
P *

dS

FIGURE 4.2 Deformed and undeformed distances between adjacent points.

4.1.2 DISPLACEMENT VECTOR


The vector u(X) represents the displacement from position X to x:
u(X, t ) = x X.

(4.1)

Now consider two close points, P and Q, in the undeformed configuration. The
vector difference XP XQ is represented as a differential d X with squared length
2
T
dS = d X d X. The corresponding quantity in the deformed configuration is dx, with
2
T
dS = dx dx.

4.1.3 DEFORMATION GRADIENT TENSOR


The deformation gradient tensor F is introduced as
dx = Fd X

F=

x
X

(4.2)

F satisfies the polar-decomposition theorem:


F = UV T ,

(4.3)

in which U and V are orthogonal and is a positive definite diagonal tensor whose

2003 by CRC CRC Press LLC

0749_Frame_C04 Page 53 Wednesday, February 19, 2003 5:33 PM

Kinematics of Deformation

53

entries j, the singular values of F, are called the principal stretches.


1

= 0

2
0

(4.4)

Based on Equation 4.3, F can be visualized as representing a rotation, followed


by a stretch, followed by a second rotation.

4.2 STRAIN
The deformation-induced change in squared length is given by
ds2 dS2 = d X T 2 Ed X

1
E = [FT F I],
2

(4.5)

in which E denotes the Lagrangian strain tensor. Also of interest is the Right CauchyT
Green strain C = F F = 2E + I. Note that F = I + u / X. If quadratic terms in
u / X are neglected, the linear-strain tensor EL is recovered as
EL =

T
1 u u
+

2 X X

(4.6)

Upon application of Equation 4.3, E is rewritten as


1
E = V ( 2 I)V T
2

(4.7)

Under pure rotation x = QX, F = Q and E = 2 [Q Q I] = 0. The case of pure


rotation in small strain is considered in a subsequent section.

4.2.1 F, E, EL

AND

IN

ORTHOGONAL COORDINATES

Let Y1, Y2, and Y3 be orthogonal coordinates of a point in an undeformed configuration, with y1, y2, y3 orthogonal coordinates in the deformed configuration. The
2,
3 and 1,2,3.
corresponding orthonormal base vectors are 1,
4.2.1.1 Deformation Gradient and Lagrangian Strain Tensors
Recalling relations introduced in Chapter 1 for orthogonal coordinates, the differential
position vectors are expressed as
dR =

dY H

2003 by CRC CRC Press LLC

dr =

dy h

(4.8)

0749_Frame_C04 Page 54 Wednesday, February 19, 2003 5:33 PM

54

Finite Element Analysis: Thermomechanics of Solids

dr =

dy h

dy h (

dy h q

h dy
H dY
H dY

h dy
( ) H dY ,
H dY

(4.9)

in which denotes the base vector in the curvilinear system used for the undeformed
configuration.
This can be written as
dr = Q T F dR,

T
[Q T ] = q

[F ] =

h y
,
H Y

(4.10)

in which Q is the orthogonal tensor representing transformation from the undeformed


to the deformed coordinate system. It follows that
1
1
E = [ F T F I ] = [ F T F I ],
2
2
from which
[ E]ij =

h 2 y y

H H Y Y
ij .

i j
i
j

(4.11)

Displacement Vector
The position vectors can be written in the form R = Zii, r = z jj. The displacement
vector referred to the undeformed base vectors is
u = [z j q ji Z i ] i ,

q ji = j i .

(4.12)

Cylindrical Coordinates
In cylindrical coordinates,
u = [r cos( ) R] e R + r sin( )e + ( z Z ) e Z .

2003 by CRC CRC Press LLC

(4.13)

0749_Frame_C04 Page 55 Wednesday, February 19, 2003 5:33 PM

Kinematics of Deformation

55
2

We now apply the chain rule to ds in cylindrical coordinates:


ds 2 = dr dr = dr 2 + r 2 d 2 + dz 2
dr
dr d
r d
d
1 dr
= dR +
Rd +
dZ + r
dR +
Rd + r
dZ
R d
dZ dR
R d
dZ
dR
2

dz
1 dz
dz
Rd +
dZ
+ dR +
R d
dZ
dR

Rd

= {dR

dR

dZ} C Rd ,

dZ

(4.14)

in which
dr
d
dz
cRR = + r +
dR dR dR
2

eRR =

1
(c 1)
2 RR

e =

1
(c 1)
2

eZZ =

1
(c 1)
2 ZZ

dr 1 dr d r d dz 1 dz
+ r
+
cR =
dR R d dR R d dR R d

eR =

1
c
2 R

1 dr dr r d d 1 dz dz
+
+
cZ =
r
R d dZ R d dZ R d dZ

eZ =

1
c
2 Z

dr dr
d
d
dz dz
cZR = + r r +
dZ dR dZ dR dZ dR

eZR =

1
c .
2 ZR

1 dr r d 1 dz
+
c =
+
R d R d R d
2

dr
d
dz
cZZ = + r +
dZ dZ dZ
2

(4.15)
4.2.1.2 Linear-Strain Tensor in Cylindrical Coordinates
If quadratic terms in the displacements and their derivatives are neglected, then
uR r R
du
dr
1+ R
dR
dR
1 dr
1 du R

R d R d
dr du R

dZ
dZ

u

R

uZ z Z ,

u
d
d u du
R =
dR
dR R
dR
R
R + uR
u
1 du
1 du
r d
=
1+
1+ R +
R d
R R d
R R d
d du

r
dZ
dZ
r

2003 by CRC CRC Press LLC

(4.16)
dz duZ

dR
dR
1 dz
1 duZ

R d R d
du
dz
1+ Z
dZ
dZ

0749_Frame_C04 Page 56 Wednesday, February 19, 2003 5:33 PM

56

Finite Element Analysis: Thermomechanics of Solids

giving rise to the linear-strain tensor


du R

dR

du
u
1
1 u R

EL = +

2 dR
R
R R

du
1 du Z

+ R

dZ
2 dR

1
2

du

dR
uR

1 u R

R
R R
1 du

R
R d
1 du Z
1 du

+
2 dZ

1 du Z

2 dR
1 du
2

dZ +

du R

dZ

1 du Z

. (4.17)

R d

du Z
dZ

The divergence of u in cylindrical coordinates is given by u = trace ddur =


trace(EL), from which
u =

duR uR 1 du duZ
+
+
+
,
dR R R d dZ

(4.18)

which agrees with the expression given in Schey (1973).

4.2.2 VELOCITY-GRADIENT TENSOR, DEFORMATION-RATE TENSOR,


AND SPIN TENSOR
We now introduce the particle velocity v = x/ t and assume that it is an explicit
function of x(t) and t. The velocity-gradient tensor L is introduced using dv = Ldx,
from which
L=
=

dv
dx
dv dX
dX dx

1 .
= FF

(4.19)

Its symmetric part, called the deformation-rate tensor,


1
D = [L + LT ],
2

(4.20)

can be regarded as a strain rate referred to the current configuration. The corresponding strain rate referred to the undeformed configuration is the Lagrangian strain rate:
1
E = [F T F + F T F]
2
1 1
= F T [FF
+ F T F T ]F
2

= F T DF.

2003 by CRC CRC Press LLC

(4.21)

0749_Frame_C04 Page 57 Wednesday, February 19, 2003 5:33 PM

Kinematics of Deformation

57

The antisymmetric portion of L is called the spin tensor W:


1
W = [L LT ].
2

(4.22)

Suppose the deformation consists only of a time-dependent, rigid-body motion:


x(t ) = Q(t )X + b(t ),

Q T (t )Q(t ) = I.

(4.23)

Clearly, F = Q and E = 0. Furthermore,


T,
L = QQ

(4.24)

T + QQ
T = QQ
T + (QQ
T )T .
which is antisymmetric since 0 = I = [QQ T ] = QQ
T

Hence, D = 0 and W = QQ , thus explaining the name of W.


4.2.2.1 v, L, D, and W in Orthogonal Coordinates
The velocity v(y, t) in orthogonal coordinates is given by
v(y, t ) =

dr
=
dt

v = h

dy
.
dt

(4.25)

Based on what we learned in Chapter 1, with v denoting the velocity vector in


orthogonal coordinates,

dv
1 v j
[L] = =
j + v j c j .

dr h y

c j =

1
(1 )
h

(4.26)

2 x y
= h
yk y j x k

j
1

Of course, [D] = [[L] + [L] ], [W] = [[L] [L] ].


4.2.2.2 Cylindrical Coordinates
The velocity vector in cylindrical coordinates is
v=

dr
d
dz
e +r
e + e
dt r
dt dt z

= vr e r + v e + vz e z .

2003 by CRC CRC Press LLC

(4.27)

0749_Frame_C04 Page 58 Wednesday, February 19, 2003 5:33 PM

58

Finite Element Analysis: Thermomechanics of Solids

Observe that
dv = dvr e r + dv e + dvz e z + vr de r + v de
= [dvr v d ]e r + [dv + vr d ]e + dvz e z .

(4.28)

Converting to matrix-vector notation, we get


v
dv
1 dv r
dvr

dr +
rd + r dz rd
dr

r d
dz
r
dv

dv
dv
v
1
dv = dr + rd + dz + r rd
r d
dz
r
dr

dv z
dv z
1 dv z

dr +
rd +
dz

dr
r d
dz

dr

L rd ,

dz

dvr
dr
dv
L=
dr
dv z

dr

1 dv r
r d
1 dv

v
r
vr

r d
r
1 dv z
r d

dz

dv
.
dz
dv z

dz
dv r

(4.29)

4.2.2.3 Spherical Coordinates


Now
v=

dr
d
d
e r + r cos
e + r
e
dt
dt
dt

= vr e r + v e + v e
e r = cos (cos e1 + sin e 2 ) + sin e 3

(4.30)

e = sin e1 + cos e 2
e = sin (cos e1 + sin e 2 ) + cos e 3 .
Next,

v = Qv,

vr

v = v ,

v

cos cos

Q = sin

sin cos

2003 by CRC CRC Press LLC

v referred to e1, e2 , e3
(4.31)
cos sin
cos
sin sin

sin

cos

0749_Frame_C04 Page 59 Wednesday, February 19, 2003 5:33 PM

Kinematics of Deformation

59

dv = QQT dv * + dQQT v
dvr

dv* = dv referred to e r , e , e

dv

= dv * + dQQT v,

(4.32)

Recall that
0

dQ( ) T
Q ( ) = cos
dt

cos

0
0
d
sin
+ 0
dt
1
0

0
sin

0
0
0

1
d
.
0
dt
0

(4.33)

Thus, it follows that


0
1
cos
dQQ T =
r cos
0

cos

0
0

1
sin r cos d + 0
r

1
0

0
sin

0
0
0

0 rd

0
(4.34)

and
cos v d + v d

dQQ T v = cos vr d + sin v d

sin v d vr d
0
1
0
=
r
0

v
vr + tan v
tan v

v
dr

0 r cos d .

vr r d

(4.35)

dr
d

dv
r cos d

dv rd

(4.36)

Finally,
dvr
dvr dr
dv
dv* = dv =
dr
dv dv

dr

2003 by CRC CRC Press LLC

dv r

1 dv r

r cos d
dv
1

r
1

r cos d
dv
1

r
1

r cos d

0749_Frame_C04 Page 60 Wednesday, February 19, 2003 5:33 PM

60

Finite Element Analysis: Thermomechanics of Solids

and
dvr

dr
dv
L=
dr
dv

dr
dvr
dr

dv
=
dr
dv

dr

dv r

r cos d
dv
1

r
1

r cos d
dv
1

r
1

r cos d

1
1

dv 1
+ 0
d r

dv
0

1 dv r

dv r

r cos d

vr + tan v
tan v

d
r

1 dv
.

r d
dv
vr

d
r

1 dv r

r cos d
r
v r tan v
dv

r cos d
dv
1

r
tan v

vr

(4.37)

The divergence of v is given by trace(L), thus,


v =

dvr
1 dv vr tan v 1 dv
+

+
,
dr r cos d
r
r d

(4.38)

which is again in agreement with Schey (1973).

4.3 DIFFERENTIAL VOLUME ELEMENT


The volume spanned by the differential-position vector dR is given by the vector
triple-product
dV0 = d X1 d X 2 d X 3 = d X1d X 2 d X 3
d X1 = dX1e1

d X 2 = dX2 e 2

The vectors dXi deform into dxj =


verified to be

dx j
dx i

d X 3 = dX3 e 3 .

(4.39)

ej dXi. The deformed volume is now readily

dV = dx1 dx 2 dx 3
1

= JdV0 ,

J = det (F) = det 2 (C),

(4.40)

and J is called the Jacobian. To obtain J for small strain, we invoke invariance and
J = det(C ) to find
1

det 2 (C) = det 2 [I + 2 E]


= (1 + 2 EI )(1 + 2 EII )(1 + 2 EIII ),

2003 by CRC CRC Press LLC

(4.41)

0749_Frame_C04 Page 61 Wednesday, February 19, 2003 5:33 PM

Kinematics of Deformation

61

in which EI, EII, EIII are the eigenvalues of EL, assumed to be much less that unity.
The linear-volume strain follows as
1

evol = det 2 (C) 1


= 1 + 2( EI + EII + EIII ) + quadratic terms 1
tr( EL ),

(4.42)

using the approximation 1 + x 1 + x/2 if x << 1.


The time derivative of J is prominent in incremental formulations in continuum
mechanics. Recalling that J = I3 ,
d
1 dI3
J=
dt
2 J dt
=

J
)
tr(C 1C
2

J
tr(F 1 F T [F T F + F T F])
2

J
tr(F 1 F + F 1 F T F T F)
2

J
[tr(F 1 F ) + tr(F 1 F T F T F)]
2

F 1 F + F T F T
= J tr

= J tr(D).

(4.43)

4.4 DIFFERENTIAL SURFACE ELEMENT


Let dS denote a surface element in the deformed configuration, with exterior unit
normal n, as illustrated in Figure 4.3. The corresponding quantities from the reference configuration are dS0 and n0. A surface element dS obeys the transformation
(Chandrashekharaiah and Debnath, 1994):
n dS = JF T n 0 dS0 ,

(4.44)

from which we conclude that


dS = J n T0 C 1 n 0 dS0

2003 by CRC CRC Press LLC

n=

F Tn0
n T0 C 1 n 0

(4.45)

0749_Frame_C04 Page 62 Wednesday, February 19, 2003 5:33 PM

62

Finite Element Analysis: Thermomechanics of Solids

n0

dx2

dS0

dX2

dS
dX1
dx1
FIGURE 4.3 Surface patches in undeformed and deformed configurations.

During deformation, the surface normal changes direction, a fact which is important, for example, in contact problems. In incremental variational methods, we
consider the differential dn and d(ndS):
dt

d
dJ T
d F T
[n dS] =
F n 0 dS0 + J
n 0 dS0 .
dt
dt
dt

(4.46)

However, recalling Equation 4.43,


dJ
= J tr(D)
dt
dJ T
F n 0 dS0 = tr(D) JF T n 0 dS0
dt
= tr(D)n dS.

(4.47)

Also, since d(F F ) = 0, then


T

dF T
dF T T
F
= F T
dt
dt
= LT F T .

(4.48)

d[ndS]
= [tr(D)I LT ]ndS.
dt

(4.49)

Finally, we have

Next, we find with some effort that


1

dn dF
=
dt
dt

n0
n T0 C 1 n 0

= [(n T Dn)I LT ]n.

2003 by CRC CRC Press LLC

F n0

n T0

dC
dt

T
0

n0

n T0 C 1 n 0 n C n 0
(4.50)

0749_Frame_C04 Page 63 Wednesday, February 19, 2003 5:33 PM

Kinematics of Deformation

63

4.5 ROTATION TENSOR


Elementary manipulation furnishes
du =

du
dX
dX

= [ EL + ]d X,

(4.51)

in which the antisymmetric rotation tensor appears:


1 u u

2 X X

(4.52)

Consider pure rotation with small angle :


cos

x = sin

0
2
1 2

sin
cos
0

0X

0 X,

(4.53)

thus,
2
2

u= 0

2
2

EL = 0

0
0

0 X +

0
0

0 ,

0
0

0X

0
0

0 .

(4.54)

(4.55)

Evidently, the normal strains do not vanish, but are second-order in , while
is first-order in . Under the assumption of small deformation, nonlinear terms are
neglected so that EL is regarded as vanishing.

2003 by CRC CRC Press LLC

0749_Frame_C04 Page 64 Wednesday, February 19, 2003 5:33 PM

64

Finite Element Analysis: Thermomechanics of Solids

L
FIGURE 4.4 Element in undeformed and deformed configurations.

Consider as a second example the following derivation in which the sides of the
unit square rotate toward each other by 2.
The deformation can be described by the relations
2
x = (1 cos ) X + sin Y 1 X + Y
2

(4.56)

2
y = sin X + (1 cos )Y X + 1 Y ,
2

from which we conclude that


2
2

EL =

0 .

(4.57)

Upon neglecting quadratic terms, we conclude that linear-shear strain is a measure of how much the axes rotate toward each other, while the rotation tensor is a
measure of how much the axes rotate in the same sense.

4.6 COMPATIBILITY CONDITIONS FOR EL AND D


The following paragraphs present the compatibility equations for the linear-strain
tensor, allowing EL to be integrated to produce a displacement field u(X), which is
unique to within a rigid-body motion. It should be evident that a parallel argument
produces exactly the same result for the velocity vector v(x) starting from a given
deformation-rate tensor D(x).

2003 by CRC CRC Press LLC

0749_Frame_C04 Page 65 Wednesday, February 19, 2003 5:33 PM

Kinematics of Deformation

65

There are two major paths to solutions of problems in continuum mechanics.


For example,
Assume/approximate u. Apply the strain-displacement relations to estimate E,
apply the stress-strain relations to estimate S, and then apply equilibrium
relations to obtain the field equation. Solve the field equation applying
boundary conditions and constraints.
Assume/approximate S, and then estimate E using the strain-displacement
relations. Apply compatibility relations to obtain the field equation, and
then solve the field equation and boundary conditions and constraints.
To understand the second solution path, we focus on the following compatibility
relation. Suppose the linear strains are known as functions of the position X. Under
what circumstances can the strain-displacement relations be integrated to produce
a displacement field that is unique to within a rigid-body displacement? Recall that
du
dX
dX

du =

= [ EL + ]d X

(4.58)

Evidently, EL is only part of the derivative of u. In 2-D rectilinear coordinates,


we will see that the compatibility equation, guaranteeing unique u to within a rigidbody motion, is given by:
2 E12
2 E22
1 2 E
= 211 +

x1x2 2 x2
x12

(4.59)

The proof of the compatibility equation is as follows. We consider a path in the


physical (X1, X2, X3) space, along which the arc length is denoted by . Accordingly,
the position vector X is a function of . The integral of du, taken from = 0 to * is
u(X( *)) = u(0) +

X ( *)

X ( *)

= u(0) +

du
dX
dX
E(X( *))dX +

X ( *)

(X)dX.

(4.60)

The term u(0) can be interpreted as the rigid-body translation. The first integral
can be evaluated since E(X) and X() are given functions. The second integral can
be rewritten using an elementary transformation as

X ( *)

(X)dX =

2003 by CRC CRC Press LLC

X ( *)

(X * X)d(X * X).

(4.61)

0749_Frame_C04 Page 66 Wednesday, February 19, 2003 5:33 PM

66

Finite Element Analysis: Thermomechanics of Solids

Note the following:


d[(X)(X * X)] = (X)d (X * X) + [(X * X)T d T (X * X)]T
= (X)d (X * X) [(X * X)T d(X * X)]T
T

d(X * X)
d ( X * X ) .
= (X)d (X * X) (X * X)T
d (X * X)

(4.62)
Consequently, integration by parts furnishes

X ( *)

(X)d (X * X) = (X(0))(X( *) X(0))

X ( * ) X ( 0 )

(X * X) T

d(X * X)
d (X * X) . (4.63)
d (X * X)

The term (X(0))(X(*) X(0)) can be identified as a rigid-body rotation since


the rotation tensor (X(0)) is independent of position. We now focus on the second
term. Now that rigid-body translation and rotation have been accommodated, the
integral should give a unique value regardless of the path. Consider two paths, path
1 and path 2, from = 0 to *. The integral over these paths should produce the same
unique values. Therefore, the closed path 0 to * along path 1 followed by * to 0 in
a negative sense along path 2 should vanish. Furthermore, path 1 and path 2 are
arbitrary, except for terminating at the same points. It follows that along any closed path

(X * X)

d(X * X)
d (X * X) = 0 T .
d (X * X)

(4.64)

1(X) 2(X) 3(X)}, in which i denotes


Let us now write (X) as (X) = {
th
the vector corresponding to the i column of . The compatibility condition can
now be written as

(X * X)

d j (X * X)
d (X * X)

d (X * X) = 0,

j = 1, 2, 3.

(4.65)

dX = 0 over
For a suitable differential function (X), the condition for 
an arbitrary contour is that the curl of vanishes, from which we express the
compatibility conditions as
d j (X)

(X * X)T
= 0,
(
X * X)
d

2003 by CRC CRC Press LLC

j = 1, 2, 3.

(4.66)

0749_Frame_C04 Page 67 Wednesday, February 19, 2003 5:33 PM

Kinematics of Deformation

67

This can be expanded at considerable effort to derive 81 compatibility equations,


including Equation 4.59.

4.7 SAMPLE PROBLEMS


1. Referring to Figure 4.5, determine u, F, J, and E as functions of X and Y.
Use H = 1.0, W = 1.0, a = 0.1, b = 0.1, c = 0.3, d = 0.2, e = 0.2, f = 0.1.
Assume a unit thickness in the Z-direction in both the deformed and
undeformed configurations.
Solution: Since straight sides are deformed into straight sides, the deformation
pattern can be assumed in the form
x = X + Y + XY

y = X + Y + XY ,

and it is necessary to determine , , , , , and from the coordinates


of the vertices in the deformed configuration.
At ( X , Y ) = (W , 0) : W + a = W

b = W

e = H

H + f = H

At ( X , Y ) = (0, H ) :

At ( X , Y ) = (W , H ) : W + c = H + H + WH

H + d = W + H + WH

After elementary manipulation,


a
= 1.1
W
f
= 1+
= 1.2
H

= 1+

e
= 0.2
H
W + c (W + a ) e
=
= 0.4
WH

b
= 0.1
W
H + d b (H + f )
=
= 0.
WH

c
e
H

H
f
b
W
undeformed plate

W
a
deformed plate

FIGURE 4.5 Plate elements in undeformed and deformed states.

2003 by CRC CRC Press LLC

0749_Frame_C04 Page 68 Wednesday, February 19, 2003 5:33 PM

68

Finite Element Analysis: Thermomechanics of Solids

The 2 2 deformation gradient tensor F and its determinant are now:

+ X 1.1 + 0.4Y
=
+ X 0.1

+ Y
F=
+ Y

0.2 + 0.4 X

1.2
.

J = 1.3 + 0.48Y 0.04 X


The displacement vector is
x X ( 1) X + Y + XY 0.1X + 0.2Y + 0.4 XY
u=
=
.
=
0.1X + 0.2Y
y Y X + ( 1)Y + XY

The Lagrangian strain E =

E=

1 1.1 + 0.4Y

2 0.2 + 0.4 X

1 [FTF
2

I] is

0.1 1.1 + 0.4Y

1.2 0.1

0.11 + 0.44Y + 0.08Y 2


=
0.17 + 0.22 X + 0.04Y + 0.08 XY

0.2 + 0.4 X 1 1

1.2
2 0

0.17 + 0.22 X + 0.04Y + 0.08 XY


0.24 + 0.08 X + 0.08 X 2

2. Figure 4.6 shows a square element at time t and at t + dt. Estimate L, D,


and W at time t. Use a = 0.1dt, b = 1 + 0.2dt, c = 0.2dt, d = 1 + 0.4dt,
e = 0.05dt, f = 0.1dt, g = 1 0.1dt, h = 1 + 0.5dt. Assume a unit thickness
in the Z-direction in both the deformed and undeformed configurations.
Y

d
c

g
h
1

e
X
1
element at time t

X
b
a
element at time t+dt

FIGURE 4.6 Element experiencing rigid body motion and deformation.

2003 by CRC CRC Press LLC

0749_Frame_C04 Page 69 Wednesday, February 19, 2003 5:33 PM

Kinematics of Deformation

69

Solution: First, represent the deformed position vectors in terms of the undeformed position vectors using eight coefficients determined using the given
geometry. In particular,
x = + X + Y + XY

y = + X + Y + XY .

Following procedures analogous to Problem 1, we find:

= 0.1dt

= 0.05dt

= 1 + 0.2 dt

= 0.1dt

= 0.2 dt

= 1 0.1dt

= 0.5dt

= 0.05dt.

The velocities can be estimated using vx

xX
dt

and vy

yY
,
dt

from which

v x = 0.1 + 0.2 X + 0.2Y 0.5 XY


v y = 0.05 + 0.1X 0.1Y 0.05 XY .
The tensors L, D, and W are now readily found as:
0.2 0.5Y
L=
0.1 0.05Y

0.2 0.5 X

0.1 0.05 X

0.2 0.5Y

D=
0.15 0.25 X 0.025Y
0

W=
0.4 + 0.25 X 0.025Y

0.15 0.25 X 0.025Y

0.1 0.05 X

0.4 0.25 X + 0.025Y


.
0

4.8 EXERCISES
1. Consider a one-dimensional deformation in which x = (1 + )X. What
value of is the linear-strain L error by 5% relative to the Lagrangian
strain E? Use the error measure
error =

E EL
.
E

2. A 1 1 square plate has a constant (2 2) Lagrangian-strain tensor E.


What is the deformed length of the diagonal? What is the volume change?

2003 by CRC CRC Press LLC

0749_Frame_C04 Page 70 Wednesday, February 19, 2003 5:33 PM

70

Finite Element Analysis: Thermomechanics of Solids

If the linear strain EL is now approximated as E, what is the diagonal and


what is the volume change? Take
0.1
E=
0.05

0.05
.
0.1

3. Verify that the expressions in the text for F and E in cylindrical coordinates
are consistent with the equation
F = Q F ,

T
[Q T ] = q

[F ] =

h y
,
H Y

in which Q is the orthogonal tensor representing transformation from the


undeformed to the deformed coordinate system, and also with

[ E]ij =

h2 y y

H H Y Y
ij .

i j i
j

4. Obtain expressions for u, F, E, and EL in spherical coordinates.


5. For cylindrical coordinates, determine the Lagrangian strain in the following two cases:
( a) Pure radial expansion:

r = R, = , z = Z

(b) Torsion:

r = R, = + Z , z = Z

6. In spherical coordinates, determine the Lagrangian strain for pure radial


expansion.
r = R, = , =
7. Obtain v, L, D, and W in spherical coordinates.
8. For cylindrical coordinates, find L, D, and W for the following flows:
(a) pure radial flow:

vr = f (r ),

v = 0,

vz = 0

(b) pipe flow:

vr = 0,

v = 0,

vz = f (r )

(c) cylindrical flow:

vr = 0,

v = f (r ),

vz = 0

(d) torsional flow:

vr = 0,

v = f ( z ),

vz = 0

9. In spherical coordinates, find L, D, and W for pure radial expansion:


vr = f (r ),

2003 by CRC CRC Press LLC

v = 0,

v = 0

0749_Frame_C04 Page 71 Wednesday, February 19, 2003 5:33 PM

Kinematics of Deformation

71

10. For linear strain in rectilinear coordinates and 2-D, the compatibility
relation is
2 E12
2 E22
1 2 E
= 211 +
.
x1x2 2 x2
x12
Find the implications of this relation for a linear-strain field assumed to
be given by
E11 = a1 X 2 + a2 XY + a3Y 2
E22 = b1 X 2 + b2 XY + b3Y 2
E12 = c1 X 2 + c2 XY + c3Y 2 .
11. Consider a square L L plate (undeformed configuration) with linear
strains
E11 = a1 + a2 X + a3Y
E22 = b1 + b2 X + b3Y
E12 = c1 + c2 X + c3Y .
Assuming that the origin does not move, find the deformed position of
(X,Y) = (L,L).

2003 by CRC CRC Press LLC

0749_Frame_C05 Page 73 Wednesday, February 19, 2003 5:03 PM

Mechanical Equilibrium
and the Principle
of Virtual Work

5.1 TRACTION AND STRESS


5.1.1 CAUCHY STRESS
Consider a differential tetrahedron enclosing the point x in the deformed configuration. The area of the inclined face is dS, and dSi is the area of the face whose exterior
normal vector is ei. Simple vector analysis serves to derive that ni = dSi/dS (see
Exercise 1 in Chapter 1). Next, let dP denote the force on a surface element dS, and
(i )
let dP denote the force on area dSi. The traction vector is introduced by = dP/dS.
As the tetrahedron shrinks to a point, the contribution of volume forces, such as
inertia, decays faster than surface forces. Balance of forces on the tetrahedron now
requires that
dPj =

dP .
i
j

(5.1)

The traction vector acting on the inclined face is defined by

dP
,
dS

dPj( i )

dPj( i ) dSi
dSi dS

(5.2)

from which

j =

dS

= Tij ni .

(5.3)

73

2003 by CRC CRC Press LLC

0749_Frame_C05 Page 74 Wednesday, February 19, 2003 5:03 PM

74

Finite Element Analysis: Thermomechanics of Solids

dF
n

dF

(2)

dS

(2)

dS
dS

dF

(1)

dF (1)
dS

(3)

(3)

1
FIGURE 5.1 Equilibrium of a tetrahedron.

with
Tij =

dPj( i )
dSi

(5.4)

It is readily seen that Tij can be interpreted as the intensity of the force acting
in the j direction on the facet pointing in the i direction, and is recognized as the
th
ij entry of the Cauchy stress T. In matrix-vector notation, the stress-traction relation
is written as
= T T n.

(5.5)

The next section will show that T is symmetric by virtue of the balance of angular
T
momentum. Equation 5.5 implies that T is a tensor, thus, it follows that T is a
tensor. To visualize T, consider a differential cube. Positive stresses are shown on
faces pointing in positive directions, as shown in Figure 5.2.
In traditional depictions, the stresses on the back faces are represented by arrows
pointing in negative directions. However, this depiction can be confusingthe
arrows actually represent the directions of the traction components. Consider the
one-dimensional bar in Figure 5.3. The traction vector te1 acts at x = L, while the
traction vector te1 acts at x = 0. At x = L, the corresponding stress is t11 = te1 e1 = t.
At x = 0, the stress is given by (te1) (e1) = t. Clearly, the stress at both ends, and
in fact throughout the bar, is positive (tensile).
We will see later that the stress tensor is symmetric by virtue of the balance of
angular momentum.

2003 by CRC CRC Press LLC

0749_Frame_C05 Page 75 Wednesday, February 19, 2003 5:03 PM

Mechanical Equilibrium and the Principle of Virtual Work

75

T33
T32

T31

T23

T13
T11

T21

T22

T12

1
FIGURE 5.2 Illustration of the stress tensor.

Z
FIGURE 5.3 Tractions on a bar experiencing uniaxial tension.

5.1.2 1ST PIOLA-KIRCHHOFF STRESS


The transformation to undeformed coordinates is now considered. In Chapter 3, we
saw that
dS = T ndS
= T JF T n 0 dS0
= S T n 0 dS0 ,

S = JF 1T.

(5.6)

S is known as the 1st Piola-Kirchhoff stress tensor, and it is not symmetric.

2003 by CRC CRC Press LLC

0749_Frame_C05 Page 76 Wednesday, February 19, 2003 5:03 PM

76

Finite Element Analysis: Thermomechanics of Solids

5.1.3 2ND PIOLA-KIRCHHOFF STRESS


We next derive the stress tensor that is conjugate to the Lagrangian-strain rate, i.e.,
corresponds to the correct amount of work per unit undeformed volume. At a segment
dS at x on the deformed boundary, assuming static conditions, the rate of work dW
of the tractions is
dW = dP T u
= T u dS.

(5.7)

Over the surface S, shifting to tensor-indicial notation and invoking the divergence
theorem, we find
W =
=
=

u dS
T

T n u dS
ij i

u j

Tij dV + u j

Tij
xi

We will shortly see that static equilibrium implies that


us to conclude that

W =

u j

Tij
x i

Tij dV .

= 0, which enables

(5.9)

Returning to matrix-vector notation, since

=
W

(5.8)

dV .

u j
x i

= [L]ji, then

tr(TL )dV
T

tr[T(D W)]dV ,

tr[TD]dV

(5.10)

since the trace vanishes for the product of a symmetric and antisymmetric tensor.

2003 by CRC CRC Press LLC

0749_Frame_C05 Page 77 Wednesday, February 19, 2003 5:03 PM

Mechanical Equilibrium and the Principle of Virtual Work

77

To convert to undeformed coordinates,


=
W

tr(TD)dV

tr( JTF

tr(SE )dV,

E F T )dV0

(5.11)
1

tr( JF TF E )dV0
S = JF T TF 1 .

The tensor S is called the 2nd Piola-Kirchhoff stress tensor. It is symmetric if


T is symmetric.
1 ) = J tr(F 1TF ) = tr( S F ), so that the 1st PiolaNote that J tr(TL) = J tr(TFF
Kirchhoff stress is said to be conjugate to the deformation-gradient tensor F.

5.2 STRESS FLUX


Consider two deformations, x1 and x2, differing only by a rigid body motion:
x 2 = V(t )x 1 + b(t ) ,

(5.12)

in which V(t) is orthogonal. A tensor A(x) is objective (see Eringen) if


A( x 2 ) = VA( x1 )V T .

(5.13)

It turns out that the matrix of time derivatives of the Cauchy stress, T , is not
objective, while the deformation rate tensor, D, is. Thus, if is a fourth-order tensor,
= D would be senseless. Instead, the time
a constitutive equation of the form T
derivative is replaced with an objective stress flux, as explained in the following.
First, note that
F2 = VF1

(5.14)

L 2 = F 2 F21
T VF + VF ]F 1 V T
= [VV
1
1
1
= + VL1 V T ,

T.
= VV

T
The tensor is antisymmetric since dl/dt = 0 = + .

2003 by CRC CRC Press LLC

0749_Frame_C05 Page 78 Wednesday, February 19, 2003 5:03 PM

78

Finite Element Analysis: Thermomechanics of Solids

We seek a stress flux affording the simplest conversion from deformed to undeformed coordinates. Note that
dS d
= [JF T TF 1 ]
dt dt
= Jtr(D)F T TF 1 + JF T TF 1 + JF T TF 1 + JF T TF 1 .

(5.15)

1 and F T = F T F T F T . Now,
However, (F 1 F). = 0 so that F 1 = F 1 FF
dS

= JF T T F 1 ,
dt

(5.16a)

o
T = T + tr( D)T LT TLT ,

(5.16b)

in which

is known as the Truesdell stress flux. Under pure rotation, F = Q and S = JQ T Q T .


o
To prove the objectivity of T , note that
o

T 2 = T2 + T2 tr(D 2 ) L 2 T2 T2 LT2

= [VT1 V T ] + VT V T tr(D1 ) [VL1 V T + ]VT1V T VT1T V T VLT1 V T


= VT1V T + VT1V T VT1V T + VT1V T tr(D1 )
VL1T1 V T VT1LT1 V T VT1 V T + VT1 V T
o

= V T 1 VT .

(5.16c)

The choice of stress flux is not unique. For example, the stress flux given by


T = T Ttr(D)
= T LT T LT

(5.17)

is also objective, as is the widely used Jaumann stress flux:

T = T + WT T W,

2003 by CRC CRC Press LLC

W=

1
(L LT ).
2

(5.18)

0749_Frame_C05 Page 79 Wednesday, February 19, 2003 5:03 PM

Mechanical Equilibrium and the Principle of Virtual Work

79

5.3 BALANCE OF MASS, LINEAR MOMENTUM,


AND ANGULAR MOMENTUM
5.3.1 BALANCE

OF

MASS

Balance of mass requires that the total mass of an isolated body not change. For
example,
d
dt

dV = 0,

(5.19)

in which ( x, t ) is the mass density. Since dV = JdV0 , it follows that J = 0 . In


addition,
0=
=

d
( J )
dt

( J ) + ( J )v( x, t ),
t
x

v ( x, t ) =

x
.
t

(5.20)

5.3.2 RAYLEIGH TRANSPORT THEOREM


Let w( x, t ) denote a vector-valued function. Conversion of the volume integral from
deformed to undeformed coordinates is achieved as

w(x, t)dV =
V

V0

0 w( x, t )dV0 .

(5.21)

The Rayleigh Transport Theorem is


d
dt

5.3.3 BALANCE

OF

w(x, t)dV = dt w(x, t)dV = 0 .

(5.22)

LINEAR MOMENTUM

In a fixed-coordinate system, balance of linear momentum requires that the total


force on a body with volume V and surface S be equal to the rate of change of linear
momentum:
P=

dS = dt dt dV .
d

du

(5.23)

Invoking Equation 5.22 yields

2003 by CRC CRC Press LLC

dS =

d 2u
dV .
dt 2

(5.24)

0749_Frame_C05 Page 80 Wednesday, February 19, 2003 5:03 PM

80

Finite Element Analysis: Thermomechanics of Solids

In current coordinates, application of the divergence theorem furnishes the equilibrium equation

dS = T n dS
T

= [ T T T ]T dV .

(5.25)

Equation 5.24 now becomes

T T
d 2uT
T
dV = 0 T .

dt 2

(5.26)

Since this equation applies not only to the whole body, but also to arbitrary
subdomains of the body, the argument of the integral in Equation 5.26 must vanish
pointwise:
TT T =

d 2 uT
.
dt 2

(5.27)

To convert to undeformed coordinates, the 1st Piola-Kirchhoff stress is invoked


to furnish

S T n 0 dS0 =

d 2uT
dV0 ,
dt 2

(5.28)

and the divergence theorem furnishes


T0 S T = 0

d 2uT
,
dt 2

(5.29)

in which denotes the divergence operator referred to undeformed coordinates.


This equation will later be the starting point in the formulation of incremental
variational principles.

5.3.4 BALANCE

OF

ANGULAR MOMENTUM

Assuming that only surface forces are present, relative to the origin, the total moment
of the traction is equal to the rate of change of angular momentum:

x dS = dt x x dV.
d

2003 by CRC CRC Press LLC

(5.30)

0749_Frame_C05 Page 81 Wednesday, February 19, 2003 5:03 PM

Mechanical Equilibrium and the Principle of Virtual Work

81

To examine this principle further, it is convenient to use tensor-indicial notation.


First, note using the divergence theorem that

x dS =

ijk

x j k dS

ijk

x j Tlk nl dS

x (

ijk

x j Tlk )dV

ijk

jl Tlk dV + ijk x j

ijk

Tjk dV + ijk x j

T dV
xl lk

T dV .
xl lk

(5.31)

Continuing,
d
dt

x x dV = x x dV + x x dV
=

x x dV

ijk

x j
x k dV .

(5.32)

Balance of angular momentum can thus be restated as

0=

ijk

Tjk dV + ijk x j
Tlk
x k dV .
xl

(5.33)

The second term vanishes by virtue of balance of linear momentum (see Equation
5.27), leaving

ijk Tjk = 0 ,

(5.34)

which implies that T is symmetric: T = T (see exercises in Chapter 1). It follows


from Equation 5.6 and Equation 5.11 that S is also symmetric, but that S is not
symmetric.
T

2003 by CRC CRC Press LLC

0749_Frame_C05 Page 82 Wednesday, February 19, 2003 5:03 PM

82

Finite Element Analysis: Thermomechanics of Solids

5.4 PRINCIPLE OF VIRTUAL WORK


The balance of linear and angular momentum leads to auxiliary variational principles
that are fundamental to the finite-element method. Variational methods were introduced in Chapter 3. We recall the balance of linear momentum in rectilinear coordinates as

T uk = 0 ,
xl kl

(5.35)

and Tlk = Tkl by virtue of the balance of angular momentum. We have tacitly assumed
that x = uk , which is to say that deformed positions are referred to a coordinate
system that does not translate or rotate. A variational principle is sought from

u x T
k

kl

uk dV = 0 ,

(5.36)

in which u k is an admissible variation of uk . Consider the spatial dependence of


uk to be subject to variation, but not the temporal dependence. For example, if uk
can be represented, at least locally, as
u = N T ( x) ( t ) ,

(5.37a)

u k = [N T ( x )]kl l ( t ).

(5.37b)

then

The second term in the variational equation remains as uk uk dV . The first


term is integrated by parts only once for reasons that will be identified shortly: it
becomes [u T ]dV u k T dV . From the divergence theorem,
x l

x l

k lk

lk

x [u T ]dV = n [u T ]dS
k lk

k lk

= uk k dS,

(5.38)

which can be interpreted as the virtual work of the tractions on the exterior boundary.
Next, since T is symmetric,

uk
T dV = 3kl Tlk dV
xl lk

u
1 u
3kl = k + l ,
2 xl x k

2003 by CRC CRC Press LLC

(5.39)

0749_Frame_C05 Page 83 Wednesday, February 19, 2003 5:03 PM

Mechanical Equilibrium and the Principle of Virtual Work

83

and we call 3kl the Eulerian strain. The term 3kl Tlk dV can be called the virtual
work of the stresses. Next, to evaluate uk k dS, we suppose that the exterior
boundary consists of three zones: S = S1 + S2 + S3 . On S1, the displacement uk is
prescribed, causing the integral over S1 to vanish. On S2, suppose that the traction is
prescribed as k (s). The contribution is S uk k (s)dS . Finally, on S3, suppose that tk =
2
k (s) [A(s)]kl ul (s), thus furnishing S3 uk k (s)dS S3 uk [A(s)]kl ul (s)dS. The various contributions are consolidated into the Principle of Virtual Work (Zienkiewicz
and Taylor, 1989) as

3 T dV + u u dV = u (s)dS + u (s)dS u [ A(s)] u (s)dS


kl lk

k k

S2

s3

k k

s3

kl

(5.40)
Now consider the case in which, as in classical elasticity,
Tlk = d lkmn 3mn

T = D3,

(5.41)

in which dlkmn are the entries of a fourth-order, constant, positive-definite tensor D.


The first term in Equation 5.40 now becomes 1 3kl d lkmn 3mn dV , with a positive2
definite integrand. Achieving this outcome is the motivation behind integrating by
parts just once. In the language of Chapter 3, uk is the primary variable and k is
the secondary variable: on any boundary point, either uk or k should be specified,
or, more generally, k should be specified as a function of uk. Finally, application of
T
the interpolation model (see Equation 5.37) and cancellation of furnishes the
ordinary differential equation
M
+ [K + H] = f (t ) ,

(5.42)

in which

M=

N(x)N (x)dV

H=

N(x)AN (x)dV

S3

K=
f (t ) =

BDB dV

S2 + S3

(5.43)

N(s) k (s)dS

and in which VEC(T ) = VEC(3) , = TEN22(D). In addition, VEC(3) = BT ( x ) .


B is derived from the strain-displacement relations, and M is the positive-definite
mass matrix. K is the positive-definite matrix representing the domain contribution
to the stiffness matrix. H is the boundary contribution to the stiffness matrix, and f
is a consistent force vector. These notions will be addressed in greater detail in
subsequent chapters.

2003 by CRC CRC Press LLC

0749_Frame_C05 Page 84 Wednesday, February 19, 2003 5:03 PM

84

Finite Element Analysis: Thermomechanics of Solids

To convert the Principle of Virtual Work to undeformed coordinates,

u u dV u u dV
k

S2

uk k (s)dS +

S3

0 k

uk k (s)dS

S20

uk k0 (s0 )dS0 +

S30

(5.44)

uk k0 (s0 )dS0 ,

using, from Chapter 3,

= J n T0 C 1 n 0 .

k0 = k

(5.45)

Next,

u [A(s)] u (s)dS u [A(s)] u (s )dS .


S3

kl

S3

kl l

(5.46)

Some manipulation is required to convert the virtual work of the stresses.


Observe that
3 =

1 u u
+

2 x x

T
1 u X u X
+

2 X x X x

1
[FF 1 + F T F T ]
2

1
= F T [F T F + F T F] F 1
2

= F T EF 1 .

(5.47)

Now

T dV =

kl lk

tr( 3T)dV

tr(F

tr( EJ F

tr( ES)dV

E F 1 T) JdV0
1

T F T )dV0

= E ji Sij dV0 .

2003 by CRC CRC Press LLC

(5.48)

0749_Frame_C05 Page 85 Wednesday, February 19, 2003 5:03 PM

Mechanical Equilibrium and the Principle of Virtual Work

85

Consolidating the terms, the Principle of Virtual Work in undeformed coordinates


is

E S dV + u u dV =
ji ij

0 k

S20

uk k0 (s0 )dS0 +

S30

S30

uk k0 (s0 )dS0

uk [ A(s 0 )]kl ul (s0 )dS0 .

(5.49)

5.5 SAMPLE PROBLEMS


1. The following plate is subject to the tractions shown. At B, there is a
frictionless roller support. Find the reactions at A and B. Assume that the
plate has thickness W.
y

= 0[ ex + ey ]

x
L

H
x
A

B
L

FIGURE 5.4 Plate under traction.

Solution: The reactions are denoted as RAx , RAy , RBx , and RBy and are all
assumed to be positive. Clearly, RBx = 0. Force equilibrium requires that
RAx tW
RAy + RBy tW

x
tW
dx = 0 RAx =
2L
L
x
tW
.
dx = 0 RAy + RBy =
2L
L

Balance of moments about A requires that


LRBy + tW
Accordingly, RAy =

2003 by CRC CRC Press LLC

5 tW
6 L

x2
tW
.
dx = 0 RBy =
3L
L

0749_Frame_C05 Page 86 Wednesday, February 19, 2003 5:03 PM

86

Finite Element Analysis: Thermomechanics of Solids

2. The figure shows a square plate of uniform thickness W. It experiences


Cauchy stresses given by
T11

T = T21

T12
0
0

0 .

On diagonal AB, find the total transverse force and moment about A.
y
B

x
A

FIGURE 5.5 Plate element under stress.

Solution: The axes are rotated through 45 degrees so that the x-axis coincides
with the diagonal (x-axis). Now,
Tx x =

Txx
+ T xy
2

Tyy =

Txx
T xy
2

The transverse force is


T

Fy = xx T xy 2 WL.
2

The moment about A is


M=

2003 by CRC CRC Press LLC

2
LFy .
2

Tx y =

Txx
.
2

0749_Frame_C05 Page 87 Wednesday, February 19, 2003 5:03 PM

Mechanical Equilibrium and the Principle of Virtual Work

87

3. The plate illustrated has thickness W, width dx, and height dy. The stresses
Txz, Tyz, and Tzz vanish. Find the equations resulting from balancing forces
and moments. In particular, prove that
Txy = Tyx .
Tyy + dTyy

Tyx + dTyx

Txy + dTxy
Txx + dTxx

dy

Txx
Txy

dx
Tyx

Tyy
FIGURE 5.6 Moment balance on a plate element under nonuniform stress.

Solution: For the x-direction forces,


dTyx
dTyx
dT
dT

W Txx + xx dx Txx dy + W Tyx +


dy Tyx dx = 0 xx +
=0.
dx
dy
dx
dy

For y-direction forces,


dTxy
dTyy
dTxy dTyy

W Txy +
dx Txy dy + W Tyy +
dy Tyy dx = 0
+
= 0.
dx
dy
dx
dy

For moments about A,


dTyx
dT

dy
W Txx + xx dx Txx dy
W Tyx +
dy dxdy
dx
dy

dTxy
dTyy

dx
dx dydx + W Tyy +
dy Tyy dx
= 0.
+ W Txy +
dx
dy
2

2003 by CRC CRC Press LLC

0749_Frame_C05 Page 88 Wednesday, February 19, 2003 5:03 PM

88

Finite Element Analysis: Thermomechanics of Solids

Consequently, neglecting terms of third order in differentials,


WTyx dxdy + WTxy dydx = 0 Txy = Tyx .
4. At point (0,0,0), the tractions 1, 2, 3 act on planes with normal vectors
n1, n2, and n3. Find the Cauchy stress tensor T given
n1 =

1
[e + e 2 + e 3 ]
3 1

1 =

1
[6e 1 + 9e 2 + 15e 3 ]
3

n2 =

1
[e + e 2 e 3 ]
3 1

2 =

1
[0e 1 + 1e 2 + 3e 3 ]
3

n3 =

1
[e e 2 e 3 ]
3 1

3 =

1
[4e 1 + 5e 2 + 7e 3 ].
3

Solution: This problem requires application of the stress-traction relation.


Now
6 T11
1
9 = T
3 21
12 T31
T11 + T12 + T13 = 6

T12
T22
T32

T13
1
1
T23
1
3
T33
1

(1) T21 + T22 + T23 = 9

(2) T31 + T32 + T33 = 1 (3) .

Similarly,

1
2
T11 + T12 = 3

3 T11

5 = T21

7 T31

( 4)

(7)

2003 by CRC CRC Press LLC

T22
T32

T21 + T22 = 5

4 T11
1
5 = T
3 21
6 T31
T11 T12 T13 = 4

T12

T12
T22
T32

T13
1
1
T23
1
2
T33
0
T31 + T32 = 7

(5)

(6)

T13
1
1
T23
1
3
T33
1

T21 T22 T23 = 5

(8)

T31 T32 T33 = 6

(9)

0749_Frame_C05 Page 89 Wednesday, February 19, 2003 5:03 PM

Mechanical Equilibrium and the Principle of Virtual Work

89

The solution is straightforward. For example, adding (1) and (7) furnishes T11 =
1, from which (4) implies T12 = 2.
Solving the equations furnishes us with
1

T = 2

4 .

2
3
4

5. Given the Cauchy stress T, find S and S if x(t) = Q(t)X, Q(t)Q(t) = I.


T

Solution: Clearly F = Q, thus J = 1. Also, F


T
Q(t)T and S = Q(t)TQ(t) .

= Q. It is immediate that S =

5.6 EXERCISES
1. In classical linear elasticity, introduce the isotropic stress and isotropic
linear strain as
s = tr( S)

e = tr( EL ),

and introduce the deviatoric stress and strain using


1
sd = s si
3

1
ed = e esi.
3

Verify that, if s = ze + tr( )i


i Tsd = 0

iTed = 0

sd = 2 ed
s = (2 + 3 )e .
2. Find the reaction forces in the following problem. There is a frictionless
roller support at B.

2003 by CRC CRC Press LLC

0749_Frame_C05 Page 90 Wednesday, February 19, 2003 5:03 PM

90

Finite Element Analysis: Thermomechanics of Solids

= 0[ex + ey ][1 y/H]

H
x
A

B
L

FIGURE 5.7 Plate element under traction.

3. Find the net force and moment about A (thickness = W) on the diagonal
in the following plate, given the Cauchy stresses
x
T = y
0

y
xy
0

0
0 .
0

y
B

A
L
FIGURE 5.8 Plate element experiencing stress.

4. At point (0,0,0), the tractions 1, 2, 3 act on planes with normal vectors


n1, n2, and n3. Find the Cauchy stress tensor T, given,
n1 =

1
[e + e 2 + e 3 ]
3 1

1 =

n2 =

1
[e + e 2 ]
2 1

2 =

n3 =

1
[e e 2 e 3 ]
3 1

3 =

2003 by CRC CRC Press LLC

1
[2e 1 5e 2 + 6e 3 ]
3
1
[e e 2 + e 3 ]
2 1
1
[6e 1 1e 2 + 2e 3 ].
3

0749_Frame_C05 Page 91 Wednesday, February 19, 2003 5:03 PM

Mechanical Equilibrium and the Principle of Virtual Work


st

nd

5. Given the Cauchy stress tensor T, find the 1 and 2 Piola-Kirchhoff stress
(t)X, in which
tensors if x(t) = Q(t)
1 + at

(t ) = 0

0
1 + bt
0

0 .

1 + ct

6. The Cauchy stress tensor is given by


T11

T = T12

T12
T22
0

0 .

Now suppose the 1-2 axes are rotated through + degrees around the
z-axis, to furnish 1 2 axes.
Prove that
T1 1 =

T11 + T22 T11 T22


+
cos 2 + T12 sin 2
2
2

T22 =

T11 + T22 T11 T22

cos 2 T12 sin 2


2
2

T1 2 =

T11 T22
sin 2 + T12 cos 2
2

7. Verify that
Eij( L ) =

Tij
Tkk ij

2
2 + 3

is the inverse of Tij = 2 Eij( L ) + Ekk( L ) ij .


8. In undeformed coordinates and Kronecker Product (VEC) notation, the
2nd Piola-Kirchhoff stress for incompressible hyperelastic materials can
be written as
s=

w
= 21n1 + 2 2 n 2 pI3 n 3 .
e

Find the corresponding expression in deformed coordinates. Derive


1 and 2 , in which direct transformation furnishes
t = 2 1m1 + 2 2 m 2 pi .

2003 by CRC CRC Press LLC

91

0749_Frame_C05 Page 92 Wednesday, February 19, 2003 5:03 PM

92

Finite Element Analysis: Thermomechanics of Solids

9. Prove that, under uniaxial tension in an isotropic, linearly elastic material,


E22 = E33.
10. Obtain in terms of E and . (See Problem 7.)
( L)
11. The bulk modulus is defined by tkk = 3ekk
. Obtain as a function of
E and . (See Problem 7.)
12. The 2" 2" 2" cube shown in Figure 5.9 is confined on its sides facing
the x faces by rigid, frictionless walls. The sides facing the z faces are
free. The top and bottom faces are subjected to a compressive force of
7
100 lbf. Take E = 10 psi and = 1/3. Find all nonzero stresses and strains.
What is the volume change? What are the principal stresses and strains?
What is the maximum shear stress?

x
z

FIGURE 5.9 Strain in a constrained cube.

13. Write out the proof of the Principle of Virtual Work in deformed coordinates.
14. Write out the conversion of the Principle of Virtual Work to undeformed
coordinates.
15. Write out the steps whereby the Principle of Virtual Work leads to
M
+ [K + H] = f (t ) .
16. Suppose that u = N (x) (t) and assume two dimensions. In 2-D, find B
T
in terms of the derivatives of N (x), in which
T

2003 by CRC CRC Press LLC

0749_Frame_C05 Page 93 Wednesday, February 19, 2003 5:03 PM

Mechanical Equilibrium and the Principle of Virtual Work

311

321
T
= B ( x ) (t ),
312

322
1
N T (x) =
0

u1
T
= N ( x ) (t )
u2

y 0

is a 1 n constant matrix (row vector); is n 1.

2003 by CRC CRC Press LLC

93

0
.

0749_Frame_C06 Page 95 Wednesday, February 19, 2003 5:06 PM

Stress-Strain Relation and


the Tangent-Modulus
Tensor

6.1 STRESS-STRAIN BEHAVIOR: CLASSICAL


LINEAR ELASTICITY
Under the assumption of linear strain, the distinction between the Cauchy and PiolaKirchhoff stresses vanishes. The stress is assumed to be given as a linear function
of linear strain by the relation
Tij = cijkl Ekl( L ) ,

(6.1)

in which cijkl are constants and are the entries of a 3 3 3 3 fourth-order tensor,
C. If T and EL were not symmetrical, C might have as many as 81 distinct entries.
However, due to the symmetry of T and EL there are no more than 36 distinct entries.
Thermodynamic arguments in subsequent sections will provide a rationale for the
Maxwell relations:
Tij
Ekl

Tkl
.
Eij

(6.2)

It follows that cijkl = cklij, which implies that there are, at most, 21 distinct coefficients. There are no further arguments from general principles for fewer coefficients.
Instead, the number of distinct coefficients is specific to a material, and reflects the
degree of symmetry in the material. The smallest number of distinct coefficients is
achieved in the case of isotropy, which can be explained physically as follows.
Suppose a thin plate of elastic material is tested such that thin strips are removed
at several angles and then subjected to uniaxial tension. If the measured stress-strain
curves are the same and independent of the orientation at which they are cut, the
material is isotropic. Otherwise, it exhibits anisotropy, but may still exhibit limited
types of symmetry, such as transverse isotropy or orthotropy. The notion of isotropy
is illustrated in Figure 6.1.
In isotropic, linear-elastic materials (which implies linear strain), the number of
distinct coefficients can be reduced to two, m and l, as illustrated by Lames equation,
Tij = 2 Eij( L ) + Ekk( L ) ij .

(6.3)
95

2003 by CRC CRC Press LLC

0749_Frame_C06 Page 96 Wednesday, February 19, 2003 5:06 PM

96

Finite Element Analysis: Thermomechanics of Solids

T
34
1

2
3
4

E
FIGURE 6.1 Illustration of isotropy.

This can be inverted to furnish


Eij( L ) =

T
T .
2 ij 2 + 3 kk ij

(6.4)

The classical elastic modulus E0 and Poissons ratio n represent response under
uniaxial tension only, provided that T11 = T, Tij = 0. Otherwise,
E=

T11
E11( L )

( L)
E( L)
E22
= 33
.
( L)
E11
E11( L )

(6.5)

=
,
E 2 2 + 3

(6.6)

It is readily verified that


1
1

=
1

E 2 2 + 3
from which it is immediate that
1 1+
=
,
2
E

(6.7)

Leaving the case of uniaxial tension for the normal (diagonal) stresses and
strains, we can write
E11( L ) =
=

1
1

1
T
(T + T )
2 2 + 3 11 2 2 + 3 22 33
1
[T (T22 + T33 )],
E 11
( L)
E22
=

2003 by CRC CRC Press LLC

1
[T (T33 + T11 )],
E 22

(6.8)
(6.9)

0749_Frame_C06 Page 97 Wednesday, February 19, 2003 5:06 PM

Stress-Strain Relation and the Tangent-Modulus Tensor

97

and
1
[T (T11 + T22 )],
E 33

( L)
E33
=

(6.10)

and the off-diagonal terms satisfy


E12( L ) =

1+
T
E 12

( L)
E23
=

1+
T
E 23

( L)
E31
=

1+
T .
E 31

(6.11)

6.2 ISOTHERMAL TANGENT-MODULUS TENSOR


6.2.1 CLASSICAL ELASTICITY
Under small deformation, the fourth-order tangent-modulus tensor D in linear elasticity is defined by
dT = DdEL .

(6.12)

In linear isotropic elasticity, the stress-strain relations are written in the Lames
form as
T = 2 EL + tr( EL )I.

(6.13)

Using Kronecker Product notation from Chapter 2, this can be rewritten as


VEC(T ) = [2 I I + iiT ]VEC( EL ),

(6.14)

from which we conclude that


D = ITEN22(2 I I + iiT ).

(6.15)

6.2.2 COMPRESSIBLE HYPERELASTIC MATERIALS


In isotropic hyperelasticity, which is descriptive of compressible rubber elasticity,
nd
the 2 Piola-Kirchhoff stress is taken to be derivable from a strain-energy function
that depends on the principal invariants I1, I2, I3 of the Right Cauchy-Green strain
tensor:
S=

dw
dw
=2
dE
dC

s = VEC( S)

2003 by CRC CRC Press LLC

sT =

dw
dw
=2
de
dc

e = VEC( E)

(a)
(6.16)
c = VEC( C ).

(b)

0749_Frame_C06 Page 98 Wednesday, February 19, 2003 5:06 PM

98

Finite Element Analysis: Thermomechanics of Solids

Now,
s = 2i ni ,

i =

w
,
I i

niT =

I i
.
c

(6.17)

From Chapter 2,
n1 = i

n2 = I1i c

n3 = VEC(C1 ) I3 .

(6.18)

The tangent-modulus tensor Do, referred to the undeformed configuration, is


given by
dS = DodE

ds = TEN 22( Do )dE,

(6.19)

and now
TEN 22( Do ) = 4ij ni nTj + 4i A i ,

Ai =

dni
.
dc

(6.20)

Finally,
A1 =

di
= 0T
dc

A2 =

d
I
dc 2

d
[ I i c]
dc 1

= iiT I 9
A3 =
=

(a)

(b)

(6.21)

d
[VEC(C 1 ) I3 ]
dc
d
[ I i I1c + VEC(C 2 )]
dc 2

= inT2 ciT + C C
= I1[iiT I 9 ] [ic T + ciT ] + C C

(c)

In deriving A3 we have taken advantage of the Cayley-Hamilton theorem (see


Chapter 2).

2003 by CRC CRC Press LLC

0749_Frame_C06 Page 99 Wednesday, February 19, 2003 5:06 PM

Stress-Strain Relation and the Tangent-Modulus Tensor

99

6.3 INCOMPRESSIBLE AND NEAR-INCOMPRESSIBLE


HYPERELASTIC MATERIALS
Polymeric materials, such as natural rubber, are often nearly incompressible. For
some applications, they can be idealized as incompressible. However, for applications involving confinement, such as in the corners of seal wells, it may be necessary
to accommodate the small degree of incompressibility to achieve high accuracy.
Incompressibility and near-incompressibility represent internal constraints. The
principal (e.g., Lagrangian) strains are not independent, and the stresses are not
determined completely by the strains. Instead, differences in the principal stresses
are determined by differences in principal strains (Oden, 1972). An additional field
must be introduced to enforce the internal constraint, and we will see that this internal
field can be taken as the hydrostatic pressure (referred to the current configuration).

6.3.1 INCOMPRESSIBILITY
The constraint of incompressibility is expressed by the relation J = 1. Now,
J = det F
= det 2 (F)
= det(F) det(F T )
= det 2 (F T F)
= I3 ,

(6.22)

and consequently,
I3 = 1.

(6.23)

The constraint of incompressibility can be enforced using a Lagrange multiplier


(see Oden, 1972), denoted here as p. The multiplier depends on X and is, in fact,
the additional field just mentioned. Oden (1972) proposed introducing an augmented
strain-energy function, w, similar to
w = w( I1, I2 )

1
p( I3 1),
2

I1 =

I1
,
I31/ 3

I2 =

I2
,
I32 / 3

(6.24)

in which w is interpreted as the conventional strain-energy function, but with dependence on I3 (=1) removed. I1 and I2 are called the deviatoric invariants. For reasons
to be explained in a later chapter presenting variational principles, this form serves

2003 by CRC CRC Press LLC

0749_Frame_C06 Page 100 Wednesday, February 19, 2003 5:06 PM

100

Finite Element Analysis: Thermomechanics of Solids

to enforce incompressibility, with S now given by


s=

w
e

= 21n1 + 2 2 n 2 pI 3 n 3

1 =

w
I1

2 =

w
I2

(6.25)
I
n1 = 1
c

I
n 2 = 2 .
c
1

To convert to deformed coordinates, recall that S = JF T F . It is left to the


reader in an exercise to derive 1 and 2 in which
t = VEC(T ) = 2 1m1 + 2 2 m 2 pi

(6.26)

in which iTm1 = 0 and iTm2 = 0. It follows that p = tr(T)/3 since


tr(T ) = i T t
= 2 1i T m 1 + 2 2 i T m 2 pi T i
= i T ip
= 3 p.

(6.27)

Evidently, the Lagrange multiplier enforcing incompressibility is the true


hydrostatic pressure.
Finally, the tangent-modulus tensor is somewhat more complicated because dS
depends on dE and dp. We will see in subsequent chapters that it should be defined

as D using
ds
de
T
TEN 22( D ) =
ds
dp

ds
dp

ds
= 4[1A1 + 2 A 2 + 11
n1 (n1 ) T + 12
n1 (n 2 ) T
de

(6.28)

+ 21
n 2 (n1 ) T + 22
n 2 (n 2 ) T ]
ds
= I3 n3 .
dp
Example: Uniaxial Tension
Consider the Neo-Hookean elastomer satisfying
w = [I1 3]

2003 by CRC CRC Press LLC

and subject to

I3 = 1.

(6.29)

0749_Frame_C06 Page 101 Wednesday, February 19, 2003 5:06 PM

Stress-Strain Relation and the Tangent-Modulus Tensor

101

We seek the relation between s1 and e1, which will be obtained twice: once by
enforcing the incompressibility constraint a priori, and the second by enforcing the
constraint a posteriori.
a priori: Assume for the sake of brevity that e2 = e3. I3 = 1 implies that
c2 = 1/ c1 . The strain-energy function now is w = [c1 + 2c 3]. The stress, s1, is
1
now found as

s1 = 2

dw
1
= 2 1 3/2 .
dc1
c1

(6.30)

a posteriori: Use the augmented function


w = [ I1 3]

p
[ I 1].
2 3

(6.31)

Now

s1 = 2

dw
= 2 p/c1
dc1

0 = s2 = 2

dw
= 2 p/c2
dc 2
(6.32)

dw
= 2 p/c3
0 = s3 = 2
dc3
0=

dw
= 0 I 3 = 1.
dp

Thus, it follows that c2 = c3 = 1/ c1 and p/c2 = 2 . We conclude that


s1 = 2 p / c1
= 2

c2
p / c2
c1

c
= 2 1 2
c1

1
= 2 1 3/ 2 .
c1

2003 by CRC CRC Press LLC

(6.33)

0749_Frame_C06 Page 102 Wednesday, February 19, 2003 5:06 PM

102

Finite Element Analysis: Thermomechanics of Solids

a posteriori with deviatoric invariants: Consider the augmented function with


deviatoric invariants:

w = I1 I31/ 3 3
s1 = 2

p
[ I 1]
2 3

(a)

1
I3
dw
1 I1 I3
= 2 1 3
p
43
dc1
I
3
I
c
c
3
1
1
3

(b)
(6.34)

1
I3
dw
1 I1 I3
0 = s2 = 2
= 2 1 3
p
43
dc2
I
3
I
c
c
3
2
2
3

(c)

1
I3
dw
1 I1 I3
= 2 1 3
p
43
dc3
3 I3 c3
c3
I3

(d)

0 = s3 = 2

Hence, c2 = c3. Furthermore,


dw
= 0 I3 = 1.
dp

(6.34e)

Equation 6.34 now implies that


2 c2 c1 p

= ,
3 c2 c2

(6.35)

and substitution into Equation 6.34b furnishes

1
s1 = 2 1 3/ 2 ,
c1

(6.36)

as in the a priori case and in the first a posteriori case. Now, the Lagrange multiplier
p can be interpreted as the pressure referred to current coordinates.

6.3.2 NEAR-INCOMPRESSIBILITY
As will be seen in Chapter 18, the augmented strain-energy function
w = w( I1, I2 )

1
p2
p[ I3 1]
2
2

(6.37)

(6.38)

serves to enforce the constraint

p = I3 1 .

2003 by CRC CRC Press LLC

0749_Frame_C06 Page 103 Wednesday, February 19, 2003 5:06 PM

Stress-Strain Relation and the Tangent-Modulus Tensor

103

Here, k is the bulk modulus, and it is assumed to be quite large compared to,
for example, the small strain-shear modulus. The tangent-modulus tensor is now
ds
de
T
=
ds
dp

( )

TEN 22 D *

ds
dp
.
1

(6.39)

6.4 NONLINEAR MATERIALS AT LARGE


DEFORMATION
Suppose that the constitutive relations are measured at a constant temperature in the
current configuration as
o

T = DD,

(6.40)

in which the fourth-order tangent-modulus tensor D can, in general, be a function


of stress, strain, temperature, and internal-state variables (discussed in subsequent
o
chapters). This form is attractive since T and D are both objective. Conversion to
undeformed coordinates is realized by
S = JF 1 DDF T
1

(6.41)

= JF DF E F 1F T .
If s = VEC(S) and e = VEC(E), then
s = JI F 1VEC( DF T E F 1F T )
= JI F 1F 1 ITEN 22( D)VEC(F T E F 1 )
= JF 1 F 1TEN 22( D)(F T I)(I F T )VEC( E )
= JF 1 F 1TEN 22( D)F T F T e
= TEN 22( Do )e .

(6.42)

Recalling Chapter 2, it follows that


S = Do E ,

(6.43)

Do = ITEN 22( JF 1 F 1TEN 22( D)F T F T )

(6.44)

in which

is the tangent-modulus tensor Do referred to undeformed coordinates.

2003 by CRC CRC Press LLC

0749_Frame_C06 Page 104 Wednesday, February 19, 2003 5:06 PM

104

Finite Element Analysis: Thermomechanics of Solids

Suppose instead that the Jaumann stress flux is used and that

T = DD.

(6.45)

Now

S = JF T T + Ttr( D) (L + W)T T (LT W) F 1

= JF T DF T E F 1 + Ttr(F T E F 1 ) (L + W)T T (LT W) F 1.

(6.46)

For this flux, there does not appear to be any way that S can be written in the
form of Equation 6.42, i.e., is determined by E . To see this, consider
VEC( S ) = TEN 22( D )VEC( E ) TEN 22( D )VEC(L + W)

(a)

TEN 22( D ) = [I (F T ) 2 TEN 22( JF T DF T ) + VEC(T )VEC T (FF T )]

(b)

TEN 22( D ) = J[I F 1 ]2 [T I I TU 9 ]

(c)
(6.47)

In the second term in Equation 6.47a, VEC(L + W) cannot be eliminated in favor


of VEC(D), and hence in favor of VEC( E ) . Note that
VEC(D) =

1
1
VEC(L + LT ) = (I + U 9 )VEC(L).
2
2

(6.48)

I + U9 is singular, as seen in the following argument. Recall that U9 is symmetric


and thus has real eigenvalues. However, U 29 = I and thus the eigenvalues of U9 are
either 1 or 1. Some of the eigenvalues must be 1. Otherwise, U9 would be the
identity matrix, in which case it would not, in general, have the permutation property
identified in Chapter 2. Thus, some of the eigenvalues of I + U9 vanish. Instead, we
write
) D VEC(D) D VEC( D + 2 W),
VEC( S ) = D VEC(E

(6.49)

and we see that if the Jaumann stress flux is used, S is determined by E and the
spin W. Recall that the spin does not vanish under rigid-body rotation.

6.5 EXERCISES
1. In classical linear elasticity, introduce the isotropic stress and isotropic
linear strain as
s = tr( S)

2003 by CRC CRC Press LLC

e = tr( EL ),

0749_Frame_C06 Page 105 Wednesday, February 19, 2003 5:06 PM

Stress-Strain Relation and the Tangent-Modulus Tensor

105

and introduce the deviatoric stress and strain using


1
sd = s si
3

1
ed = e ei.
3

Verify that
iTsd = 0

iTed = 0

sd = 2 ed
s = (2 + 3 )e.
2. Verify that Equation 6.3 can be inverted to furnish
e=

t
(i T t )i,

2 (2 + 3 )

t = VEC(T )

3. In classical linear elasticity, under uniaxial stress, the elastic modulus E


and Poissons ratio are defined by
E=

11
11

22
= 33 ,
11
11

11 0

ij = 0

ij 11 .

Prove that

1
1

=
1

E 2 (2 + 3 )

2( + )

E
.
2(1 + )

4. Obtain v, L, D, and W in spherical coordinates.


5. For cylindrical coordinates, find L, D, and W for the following flows:
(a) pure radial flow:

vr = f (r ),

v = 0,

vz = 0

(b) pipe flow:

vr = 0,

v = 0,

vz = f (r )

(c) cylindrical flow:

vr = 0,

v = f (r ),

vz = 0

(d) torsional flow:

vr = 0,

v = f ( z ),

vz = 0

nd

6. In undeformed coordinates, the 2 Piola-Kirchhoff stress for an incompressible, hyperelastic material is given by
s=

2003 by CRC CRC Press LLC

w
= 21n1 + 2 2 n2 pI3n3 .
e

0749_Frame_C06 Page 106 Wednesday, February 19, 2003 5:06 PM

106

Finite Element Analysis: Thermomechanics of Solids

Find the corresponding expression in deformed coordinates: derive


1 and 2 , in which direct transformation furnishes
t = 2 1m1 + 2 2 m2 pi.
7. Prove that under uniaxial tension in an isotropic linearly elastic material
e22 = e33. (Problem 9, Chapter 5.)
8. Obtain in terms of E and . (Problem 10, Chapter 5.)
( L)
9. The bulk modulus K is defined by tkk = 3ekk
. Obtain K as a function
of E and . (Problem 11, Chapter 5.)
10. The 2" 2" 2" shown in Figure 6.2 is confined on its sides facing the
x faces by rigid, frictionless walls. The sides facing the z faces are
free. The top and bottom faces are subjected to a compressive force of
100 lbf. Take E = 10 ^ 7 psi and = 1/3. Find all nonzero stresses and
strains. What is the volume change? What are the principal stresses and
strains? What is the maximum shear stress? (Problem 12, Chapter 5.)
v
E

x
z

FIGURE 6.2 Strain in a constrained plate.

2003 by CRC CRC Press LLC

0749_Frame_C07 Page 107 Wednesday, February 19, 2003 5:07 PM

Thermal and
Thermomechanical
Response

7.1 BALANCE OF ENERGY AND PRODUCTION


OF ENTROPY
7.1.1 BALANCE

OF

ENERGY

The total energy increase in a body, including internal energy and kinetic energy, is
equal to the corresponding work done on the body and the heat added to the body.
In rate form,
K + = W + Q ,

(7.1)

in which:
is the internal energy with density
=

dV;

(7.2a)

W is the rate of mechanical work, satisfying


W =

u dS,
T

(7.2b)

Q is the rate of heat input, with heat production h and heat flux q, satisfying
Q =

hdV n q dS, and


T

(7.2c)

K is the rate of increase in the kinetic energy,


K =

u T

du
dV .
dt

(7.2d)

It has been tacitly assumed that all work is done on S, and that body forces do
no work.
107

2003 by CRC CRC Press LLC

0749_Frame_C07 Page 108 Wednesday, February 19, 2003 5:07 PM

108

Finite Element Analysis: Thermomechanics of Solids

Invoking the divergence theorem and balance of linear momentum furnishes

du

+ u dt T tr(T D) h + qdV = 0.
T

(7.3)

The inner bracketed term inside the integrand vanishes by virtue of the balance
of linear momentum. The relation holds for arbitrary volumes, from which the local
form of balance of energy, referred to undeformed coordinates, is obtained as

= tr(TD) T q + h.

(7.4)

To convert to undeformed coordinates, note that

n qdS = q J F
T

n 0 dS0

q n dS
T
0

q 0 = J F 1q.

(7.5)

In undeformed coordinates, Equation 7.3 is rewritten as

[ tr(SE) h + q ] dV
0

T
0

= 0,

(7.6a)

furnishing the local form

0 tr( SE ) 0 h + T0 q 0 = 0.

(7.6b)

7.1.2 ENTROPY PRODUCTION INEQUALITY


Following the thermodynamics of ideal and non-ideal gases, the entropy production
inequality is introduced as follows (see Callen, 1985):
H =

dV

h
dV
T

nTq
dS,
T

(7.7a)

in which H is the total entropy, is the specific entropy per unit mass, and T is the
absolute temperature. This relation provides a framework for describing the irreversible nature of dissipative processes, such as heat flow and plastic deformation.
We apply the divergence theorem to the surface integral and obtain the local form
of the entropy production inequality:

T h T q + q T T/T.

2003 by CRC CRC Press LLC

(7.7b)

0749_Frame_C07 Page 109 Wednesday, February 19, 2003 5:07 PM

Thermal and Thermomechanical Response

109

The corresponding relation in undeformed coordinates is

0 T h T0 q 0 + q T0 0 T/ T.

(7.7c)

7.1.3 THERMODYNAMIC POTENTIALS


The Balance of Energy introduces the internal energy , which is an extensive
variableits value accumulates over the domain. The mass and volume averages
of extensive variables are also referred to as extensive variables. This contrasts with
intensive, or pointwise, variables, such as the stresses and the temperature. Another
extensive variable is the entropy H. In reversible elastic systems, the heat flux is
completely converted into entropy according to
Q = TH .

(7.8)

(We shall consider several irreversible effects, such as plasticity, viscosity, and heat
conduction.) In undeformed coordinates, the balance of energy for reversible processes can be written as
+ T .
0 = tr( SE)
0

(7.9)

We call this equation the thermal equilibrium equation. It is assumed to be


integrable, so that the internal energy is dependent on the current state represented
by the current values of the state variables E and . For the sake of understanding,
we can think of T as a thermal stress and as a thermal strain. Clearly, = 0 if
there is no heat input across the surface or generated in the volume. Consequently,
the entropy is a convenient state variable for representing adiabatic processes.
In Callen (1985), a development is given for the stability of thermodynamic
equilibrium, according to which, under suitable conditions, the strain and the entropy
density assume values that maximize the internal energy. Other thermodynamic
potentials, depending on alternate state variables, can be introduced by a Lorentz
transformation, as illustrated in the following equation. Doing so is attractive if the
new state variables are accessible to measurement. For example, the Gibbs Free
Energy (density) is a function of the intensive variables S and T:

0 g = 0 tr( SE) 0 T,

(7.10a)

0 g = tr( ES ) 0T .

(7.10b)

from which

Stability of thermodynamic equilibrium requires that S and T assume values that


minimize g under suitable conditions. This potential is of interest in fluids experiencing adiabatic conditions since the pressure (stress) is accessible to measurement
using, for example, pitot tubes.

2003 by CRC CRC Press LLC

0749_Frame_C07 Page 110 Wednesday, February 19, 2003 5:07 PM

110

Finite Element Analysis: Thermomechanics of Solids

In solid continua, the stress is often more difficult to measure than the strain.
Accordingly, for solids, the Helmholtz Free Energy (density) f is introduced using

0 f = 0 0 T,

(7.11a)

0 f = tr( SE ) 0T .

(7.11b)

furnishing

It is evident that f is a function of both an intensive and an extensive variable.


At thermodynamic equilibrium, it exhibits a (stationary) saddle point rather than a
maximum or a minimum. Finally, for the sake of completeness, we mention a fourth
potential, known as the enthalpy 0h = 0x tr(SE), and

0 h = tr( ES ) + 0 T .

(7.12)

The enthalpy also is a function of an extensive variable and an intensive variable


and exhibits a saddle point at equilibrium. It is attractive in fluids under adiabatic
conditions.

7.2 CLASSICAL COUPLED LINEAR THERMOELASTICITY


The classical theory of coupled thermoelasticity in isotropic media corresponds to
the restriction to the linear-strain tensor, E E L and to the stress-strain temperature
relation
T = 2 E + [tr( E) (T T0 )]I.

(7.13)

Here, is the volumetric coefficient of thermal expansion, typically a small


number in metals. If the temperature increases without stress being applied, the
strain increases according to evol = tr(E) = (T T0). Thermoelastic processes are
assumed to be reversible, in which case, q + h = T . It is also assumed that
the specific heat at constant strain, ce, given by
ce = T

,
T

(7.14)

is constant. The balance of energy is restated as

0 = tr(TE ) T

(7.15)

Recalling that is a function of the extensive variables E and , to convert to


E and T as state variables, which are accessible to measurement, we recall the

2003 by CRC CRC Press LLC

0749_Frame_C07 Page 111 Wednesday, February 19, 2003 5:07 PM

Thermal and Thermomechanical Response

111

Helmholtz Free Energy f = e + T. Since f is an exact differential, to ensure path


independence, we infer the Maxwell relation:

T
.
=
E T T E

(7.16)

Returning to the energy-balance equation,



E +
= tr

E
E


= tr
+
E +
T.
E E T
E T E
E

(7.17)

Also, note that

T=

T=

,
E

(7.18a)

thus


= tr
+
E +
T
E E T
E T E
E

T

= tr T T
T.
E + T
T E
T E

(7.18b)

We previously identified the coefficient of specific heat, assumed constant, as ce =


T T , so that
E

= tr T T
E + ce T.

(7.19)

From Equation 7.13, upon approximating T as T0, T0 TT = T0 I. From Fouriers Law, q = kT. Thus, the thermal-field equation now can be written as
) + c T .
k2 T = T0tr(E
e

2003 by CRC CRC Press LLC

(7.20)

0749_Frame_C07 Page 112 Wednesday, February 19, 2003 5:07 PM

112

Finite Element Analysis: Thermomechanics of Solids

The balance of linear momentum, together with the stress-strain and straindisplacement relations of linear isotropic thermoelasticity, imply that

x j

1 u u
u

u
2 i + j + k (T T0 ) ij = 2i ,

x
x

t
2

j
i
k

(7.21)

from which we obtain the mechanical-field equation (Naviers Equation for


Thermoelasticity):

2 u + ( + )tr( E) T =

u
.
t 2

(7.22)

The thermal-field equation depends on the mechanical field through the term
T0 tr( E ). Consequently, if E is static, there is no coupling. Similarly, the mechanical field depends on the thermal field through T, which often is quite small in,
for example, metals, if the assumption of reversibility is a reasonable approximation.
We next derive the entropy. Since ce = T is constant, we conclude that it
E
has the form

= 0 + c e ln(T/T0 ) + * ( E),

(7.23)

where *(E) remains to be determined. We take 0 to vanish. However, E = E* =


T
TT E = I, implying that

= 0 + ce ln(T/T0 ) + tr( E).

(7.24)

Now consider f, for which the fundamental relation in Equation 7.11b implies
f
=
T E

f
= T.
E T

(7.25)

Integrating the entropy,

f = f0 ce [T ln(T/T0 ) 1] tr(E)T + f * ( E),

(7.26)

in which f (E), remains to be determined. Integrating the stress,

f = tr( E 2 ) +

2003 by CRC CRC Press LLC

2
tr (E) tr( E)(T T0 ) + f ** (T),
2

(7.27)

0749_Frame_C07 Page 113 Wednesday, February 19, 2003 5:07 PM

Thermal and Thermomechanical Response

113

**

in which f (T) also remains to be determined. However, reconciling the two forms
furnishes (taking f0 = 0)

f = tr(E2 ) +

2
tr (E) tr(E)T ce [T ln(T/T0 ) 1].
2

(7.28)

7.3 THERMAL AND THERMOMECHANICAL ANALOGS


OF THE PRINCIPLE OF VIRTUAL WORK
7.3.1 CONDUCTIVE HEAT TRANSFER
For a linear, isotropic, thermoelastic medium, the Fourier Heat Conduction Law
becomes q = kTT, in which kT is the thermal conductance, assumed positive.
Neglecting coupling to the mechanical field, the thermal-field equation in an isotropic
medium experiencing small deformation can be written as
T kT T + ce T = 0.

(7.29)

We now construct a thermal counterpart of the principle of virtual work. Multiplying by T and using integration by parts, we obtain

Tk TdV + Tc T dV = Tn qdS
T

(7.30)

Clearly, T is regarded as the primary variable, and the associated secondary


variable is q. Suppose that the boundary is decomposed into three segments: S = SI +
SII + SIII. On SI, the temperature T is prescribed as, for example, T0. It follows that
T = 0 on SI. On SII, the heat flux q is prescribed as q0. Consequently, TnTq
TnTq0. On SIII, the heat flux is dependent on the surface temperature through a
heat-transfer vector: h: q = q0 h(T T0). The right side of Equation 7.30 now becomes

Tn qdS =
T

SII

Tn T q 0 dS +

SIII

Tn T q 0 dS

SIII

Tn T h(T T0 )dS.

(7.31)

We now suppose that T is approximated using an interpolation function of the


form
T T0 ~ N TT ( x ) (t )

T ~ N TT ( x ) (t ),

(7.32)

T ~ BTT ( x ) (t ),

(7.33)

from which we obtain


T ~ BTT ( x )(t )

2003 by CRC CRC Press LLC

0749_Frame_C07 Page 114 Wednesday, February 19, 2003 5:07 PM

114

Finite Element Analysis: Thermomechanics of Solids

in which BT is the thermal analog of the strain-displacement matrix. Upon substitution


of the interpolation models, the thermal-field equation now reduces to the system
of ordinary differential equations:
K T (t ) + M T (t ) = fT ,

(7.34)

in which
K T = K T1 + K T 2

Thermal Stiffness Matrix

K T1 =

B (x)k B (x)dV

Conductance Matrix

K T2 =

Surface Conductance Matrix

MT =
fT =

SIII

T
T

N T ( x )n T hN TT ( x )dS

N (x)c N (x)dV
T

SII

Thermal Mass Matrix; Capacitance


Matrix

T
T

N T ( x )n T q 0 dS +

SIII

N T ( x )n T q 0 dS

Consistent Thermal Force;


Consistent Heat Flux

7.3.2 COUPLED LINEAR ISOTROPIC THERMOELASTICITY


The thermal-field equation is repeated as
kT 2 T = T0 tr( E ) + c e T .

(7.35)

Following the same steps used for conductive heat transfer furnishes the variational principle

Tk TdV + Tc T dV + TT tr( E )dV = Tn qdS.


T

(7.36)

The principle of virtual work for the mechanical field is recalled as

tr(ET )dV + u udV = u (s)dS.


T

(7.37)

Also, recall that T = 2E + [tr(E) (T T0)]I. Consequently,

tr( E[2E + tr(E)I])dV tr( E (T T )I)dV + u udV = u (s)dS.


T

(7.38)

2003 by CRC CRC Press LLC

0749_Frame_C07 Page 115 Wednesday, February 19, 2003 5:07 PM

Thermal and Thermomechanical Response

115

Then introduce the interpolation models,


u( x ) = N T ( x ) (t ),

(7.39)

from which we can derive B(x) and b(x), thus satisfying


VEC( E) = BT ( x ) (t )

tr( E) = b T ( x ) (t ).

(7.40)

It follows that
K (t ) + M
(t ) T (t ) = f .

(7.41)

We assume that the traction (S) is specified everywhere as 0(S) on S. Here,


K=

B (x)DB(x)dV

Stiffness Matrix

M=

N(x)N (x)dV

Mass Matrix

(7.42)

= b ( x )N T ( x )dV Thermoelastic Matrix


f=

N(x)dS

Consistent Force Vector .

For the thermal field, assuming that the heat flux q is specified as q0 on the
surface, variational methods, together with the interpolation models, furnish the
equation
1
1
1
K T (t ) +
M T (t ) + (t ) =
f ,
T0
T0
T0 T

fT =

N (x)n q dS.
T

(7.43)

The combined equations for a thermoelastic medium are now written in state
(first-order) form as
(t )
(t ) f (t )


d
W1 (t ) + W2 (t ) = 0
dt


(t )
(t ) T10 fT (t )

W1 = 0

0
K
0

2003 by CRC CRC Press LLC

0 ,

1
MT
T0

W2 = K

K
0
0

0 .

1
MT
T0

(7.44)

0749_Frame_C07 Page 116 Wednesday, February 19, 2003 5:07 PM

116

Finite Element Analysis: Thermomechanics of Solids

Note that W1 is positive-definite and symmetric, while 12 (W2 + W2T ) is positivesemidefinite, implying that coupled linear thermoelasticity is at least marginally
stable, whereas a strictly elastic system is strictly marginally stable. Thus, thermal
conduction has a stabilizing effect, which can be shown to be analogous to viscous
dissipation.

7.4 EXERCISES
1. Express the thermal equilibrium equation in:
(a) cylindrical coordinates
(b) spherical coordinates
2. Derive the specific heat at constant stress, rather than at constant strain.
3. Write down the coupled thermal and elastic equations for a one-dimensional
member.

2003 by CRC CRC Press LLC

0749_Frame_C08 Page 117 Wednesday, February 19, 2003 5:08 PM

Introduction to the FiniteElement Method

8.1 INTRODUCTION
In thermomechanical members and structures, finite-element analysis (FEA) is typically invoked to compute displacement and temperature fields from known applied
loads and heat fluxes. FEA has emerged in recent years as an essential resource for
mechanical and structural designers. Its use is often mandated by standards such as
the ASME Pressure Vessel Code, by insurance requirements, and even by law. Its
acceptance has benefited from rapid progress in related computer hardware and software, especially computer-aided design (CAD) systems. Today, a number of highly
developed, user-friendly finite-element codes are available commercially. The purpose
of this chapter is to introduce finite-element theory and practice. The next three chapters
focus on linear elasticity and thermal response, both static and dynamic, of basic
structural members. After that, nonlinear thermomechanical response is considered.
In FEA practice, a design file developed using CAD is often imported into finiteelement codes, from which point little or no additional effort is required to develop
the finite-element model and perform sophisticated thermomechanical analysis and
simulation. CAD integrated with an analysis tool, such as FEA, is an example of
computer-aided engineering (CAE). CAE is a powerful resource with the potential of
identifying design problems much more efficiently and rapidly than by trial and error.
A major FEM application is the determination of stresses and temperatures in
a component or member in locations where failure is thought most likely. If the
stresses or temperatures exceed allowable or safe values, the product can be redesigned and then reanalyzed. Analysis can be diagnostic, supporting interpretation of
product-failure data. Analysis also can be used to assess performance, for example,
by determining whether the design-stiffness coefficient for a rubber spring is attained.

8.2 OVERVIEW OF THE FINITE-ELEMENT METHOD


Consider a thermoelastic body with force and heat applied to its exterior boundary.
The finite-element method serves to determine the displacement vector u(X, t) and
the temperature T(X, t) as functions of the undeformed position X and time t. The
process of creating a finite-element model to support the design of a mechanical
system can be viewed as having (at least) eight steps:
1. The body is first discretized, i.e., it is modeled as a mesh of finite elements
connected at nodes.
2. Within each element, interpolation models are introduced to provide
approximate expressions for the unknowns, typically u(X, t) and T(X, t),
117

2003 by CRC CRC Press LLC

0749_Frame_C08 Page 118 Wednesday, February 19, 2003 5:08 PM

118

Finite Element Analysis: Thermomechanics of Solids

3.

4.

5.
6.
7.

8.

in terms of their nodal values, which now become the unknowns in the
finite-element model.
The strain-displacement relation and its thermal analog are applied to the
approximations for u and T to furnish approximations for the (Lagrangian)
strain and the thermal gradient.
The stress-strain relation and its thermal analog (Fouriers Law) are
applied to obtain approximations to stress S and heat flux q in terms of
the nodal values of u and T.
Equilibrium principles in variational form are applied using the various approximations within each element, leading to element equilibrium equations.
The element equilibrium equations are assembled to provide a global
equilibrium equation.
Prescribed kinematic and temperature conditions on the boundaries (constraints) are applied to the global equilibrium equations, thereby reducing
the number of degrees of freedom and eliminating rigid-body modes.
The resulting global equilibrium equations are then solved using computer
algorithms.

The output is postprocessed. Initially, the output should be compared to data


or benchmarks, or otherwise validated, to establish that the model correctly represents the underlying mechanical system. If not satisfied, the analyst can revise the
finite-element model and repeat the computations. When the model is validated,
postprocessing, with heavy reliance on graphics, then serves to interpret the results,
for example, determining whether the underlying design is satisfactory. If problems
with the design are identified, the analyst can then choose to revise the design. The
revised design is modeled, and the process of validation and interpretation is repeated.

8.3 MESH DEVELOPMENT


Finite-element simulation has classically been viewed as having three stages: preprocessing, analysis, and postprocessing. The input file developed at the preprocessing stage consists of several elements:
1.
2.
3.
4.
5.
6.

control information (type of analysis, etc.)


material properties (e.g., elastic modulus)
mesh (element types, nodal coordinates, connectivities)
applied force and heat flux data
supports and constraints (e.g., prescribed displacements)
initial conditions (dynamic problems)

In problems without severe stress concentrations, much of the mesh data can be
developed conveniently using automatic-mesh generation. With the input file developed, the analysis processor is activated and raw output files are generated. The
postprocessor module typically contains (interfaces to) graphical utilities, thus facilitating display of output in the form chosen by the analyst, for example, contours of
the Von Mises stress. Two problems arise at this stage: validation and interpretation.

2003 by CRC CRC Press LLC

0749_Frame_C08 Page 119 Wednesday, February 19, 2003 5:08 PM

Introduction to the Finite-Element Method

119

The analyst can use benchmark solutions, special cases, or experimental data to
validate the analysis. With validation, the analyst gains confidence in, for example,
the mesh. He or she still may face problems of interpretation, particularly if the
output is voluminous. Fortunately, current graphical-display systems make interpretation easier and more reliable, such as by displaying high stress regions in vivid
colors. Postprocessors often allow the analyst to zoom in on regions of high
interest, for example, where rubber is highly confined. More recent methods based
on virtual-reality technology enable the analyst to fly through and otherwise become
immersed in the model.
The goal of mesh design is to select the number and location of finite-element
nodes and element types so that the associated analyses are sufficiently accurate.
Several methods include automatic-mesh generation with adaptive capabilities,
which serve to produce and iteratively refine the mesh based on a user-selected error
tolerance. Even so, satisfactory meshes are not necessarily obtained, so that model
editing by the analyst may be necessary. Several practical rules are as follows:
1. Nodes should be located where concentrated loads and heat fluxes are
applied.
2. Nodes should be located where displacements and temperatures are constrained or prescribed in a concentrated manner, for example, where pins
prevent movement.
3. Nodes should be located where concentrated springs and masses and their
thermal analogs are present.
4. Nodes should be located along lines and surface patches, over which
pressures, shear stresses, compliant foundations, distributed heat fluxes,
and surface convection are applied.
5. Nodes should be located at boundary points where the applied tractions
and heat fluxes experience discontinuities.
6. Nodes should be located along lines of symmetry.
7. Nodes should be located along interfaces between different materials or
components.
8. Element-aspect ratios (ratio of largest to smallest element dimensions)
should be no greater than, for example, five.
9. Symmetric configurations should have symmetric meshes.
10. The density of elements should be greater in domains with higher gradients.
11. Interior angles in elements should not be excessively acute or obtuse, for
example, less than 45 or greater than 135.
12. Element-density variations should be gradual rather than abrupt.
13. Meshes should be uniform in subdomains with low gradients.
14. Element orientations should be staggered to prevent bias.
In modeling a configuration, a good practice is initially to develop the mesh locally
in domains expected to have high gradients, and thereafter to develop the mesh
in the intervening low-gradient domains, thereby reconciling the high-gradient
domains.

2003 by CRC CRC Press LLC

0749_Frame_C08 Page 120 Wednesday, February 19, 2003 5:08 PM

120

Finite Element Analysis: Thermomechanics of Solids

There are two classes of errors in finite-element analysis:


Modeling error ensues from inaccuracies in such input data as the material
properties, boundary conditions, and initial values. In addition, there often are
compromises in the mesh, for example, modeling sharp corners as rounded.
Numerical error is primarily due to truncation and round-off. As a practical
matter, error in a finite-element simulation is often assessed by comparing
solutions from two meshes, the second of which is a refinement of the first.
The sensitivity of finite-element computations to error is to some extent controllable. If the condition number of the stiffness matrix (the ratio of the maximum
to the minimum eigenvalue) is modest, sensitivity is reduced. Typically, the condition
number increases rapidly as the number of nodes in a system grows. In addition,
highly irregular meshes tend to produce high-condition numbers. Models mixing
soft components, for example, rubber, with stiff components, such as steel plates,
are also likely to have high-condition numbers. Where possible, the model should
be designed to reduce the condition number.

2003 by CRC CRC Press LLC

0749_Frame_C09 Page 121 Wednesday, February 19, 2003 5:09 PM

Element Fields in Linear


Problems

This chapter presents interpolation models in physical coordinates for the most part,
for the sake of simplicity and brevity. However, in finite-element codes, the physical
coordinates are replaced by natural coordinates using relations similar to interpolation models. Natural coordinates allow use of Gaussian quadrature for integration
and, to some extent, reduce the sensitivity of the elements to geometric details in
the physical mesh. Several examples of the use of natural coordinates are given.

9.1 INTERPOLATION MODELS


9.1.1 ONE-DIMENSIONAL MEMBERS
9.1.1.1 Rods
The governing equation for the displacements in rods (also bars, tendons, and
shafts) is

EA

2u
2u
=

A
,
x 2
t 2

(9.1)

in which u(x, t) denotes the radial displacement, E, and are constants, x is the
spatial coordinate, and t denotes time. Since the displacement is governed by a
second-order differential equation, in the spatial domain, it requires two (timedependent) constants of integration. Applied to an element, the two constants can
be supplied implicitly using two nodal displacements as functions of time. We now
approximate u(x, t) using its values at xe and xe+1, as shown in Figure 9.1.
The lowest-order interpolation model consistent with two integration constants
is linear, in the form

u( x, t ) = Tm1 ( x )m1 m1 (t ),

ue (t )
m1 (t ) =
,
ue+1 (t )

Tm1 ( x ) = (1

x ).

(9.2)

We seek to identify m1 in terms of the nodal values of u. Letting ue = u(xe) and


ue+1 = u(xe+1), furnishes
ue (t ) = (1

xe )m1 m1 (t ),

ue+1 (t ) = (1

xe+1 )m1 m1 (t ).

(9.3)

121

2003 by CRC CRC Press LLC

0749_Frame_C09 Page 122 Wednesday, February 19, 2003 5:09 PM

122

Finite Element Analysis: Thermomechanics of Solids

ue

ue+1
x

xe

xe+1

FIGURE 9.1 Rod element.

wet

wet +1

we

we+1
x

xe

xe+1

FIGURE 9.2 Beam element.

However, from the meaning of m1(t), we conclude that


1
m1 =
1

xe
,
1

xe
1 xe+1
=
le 1
xe+1

le = xe+1 xe .

(9.4)

9.1.1.2 Beams
The equation for a beam, following Euler-Bernoulli theory, is:

EI

4w
2u
+ A 2 = 0,
4
x
t

(9.5)

in which w(x, t) denotes the transverse displacement of the beams neutral axis, and
I is a constant. In the spatial domain, there are four constants of integration. In an
element, they can be supplied implicitly by the values of w and w = w/x at each
of the two element nodes. Referring to Figure 9.2, we introduce the interpolation
model for w(x, t):

w( x, t ) = b1 ( x )b1 b1 (t ),

2003 by CRC CRC Press LLC

b1 ( x ) = 1

x2

x3 ,

we

we
m1 (t ) =
. (9.6)
we+1

we+1

0749_Frame_C09 Page 123 Wednesday, February 19, 2003 5:09 PM

Element Fields in Linear Problems

123

Enforcing this model at xe and at xe+1 furnishes


1

0
b1 =
1

xe

x e2

2 x e

x e+1

x e2+1

2 x e+1

x e3

3 x e2
.
x e3+1

2
3 x e+1

(9.7)

9.1.1.3 Beam Columns


Beam columns are of interest, among other reasons, in predicting buckling according
to the Euler criterion. The zdisplacement w of the neutral axis is assumed to depend
only on x and the xdisplacement. Also, u is modeled as
u( x, z ) = u0 ( x ) z

w( x )
,
x

(9.8)

in which u0(x) represents the stretching of the neutral axis. It is necessary to know
u0(x), w(x) and w(xx ) at xe and xe+1. The interpolation model is now
u( x, z, t ) = (1

x )m1 m1 z(1

ue (t )
ml =
,
ue+1 (t )

x 3 )b1 b1,

x2

(9.9)

we

we
bl =
,
we+1

we+1

and

m1 =

1
le

x e+1

0
b1=
1

xe
,
1

xe

x e2

2 x e

x e+1

x e2+1

2 x e+1

x e3

3 x e2
.
x e3+1

3 x e2+1

9.1.1.4 Temperature Model: One Dimension


The temperature variable to be determined is T T0, in which T0 is a reference
temperature assumed to be independent of x. The governing equation for a onedimensional conductor is
kA

2003 by CRC CRC Press LLC

2 T
T
= Ace
.
2
x
t

(9.10)

0749_Frame_C09 Page 124 Wednesday, February 19, 2003 5:09 PM

124

Finite Element Analysis: Thermomechanics of Solids

z
Xe+2,Ye+2
Y
Xe+1,Ye+1

Xe,Ye

middle surface
FIGURE 9.3 Triangular plate element.

This equation is formally the same as for a rod equation (see Equation 9.1),
furnishing the interpolation model for the element as
T( x, t ) T0 = Tm1 ( x )m11 (t ),

Te (t ) T0
1 (t ) =
.
Te+1 (t ) T0

(9.11)

9.1.2 INTERPOLATION MODELS: TWO DIMENSIONS


9.1.2.1 Membrane Plate
Now suppose that the displacements u(x, y, t) and v(x, y, t) are modeled on the
triangular plate element in Figure 9.3, using the values ue(t), ve(t), ue+1(t), ve+1(t), ue+2(t),
and ve+2(t). This element arises in plane stress and plane strain, and is called a
membrane plate element. A linear model suffices for each quantity because it provides
three coefficients to match three nodal values. The interpolation model now is

u( x, y, t ) m 2

v( x, y, t ) 0 T

ue (t )

u (t ) = ue+1 (t ) ,

ue+2 (t )

0 T m2

m 2 0 T

ve (t )

v (t ) = ve+1 (t ) ,

ve+2 (t )

m2

1

= x ,

y

0 T u (t )

m2 v (t )

m 2

= 1

(9.12)

xe
xe+1
xe + 2

ye

ye+1 .

ye+2

9.1.2.2 Plate with Bending Stresses


In a plate element experiencing bending only, the in-plane displacements, u and v,
are expressed by

u( x, y, z, t ) = z

2003 by CRC CRC Press LLC

w
w
, v( x, y, z, t ) = z
,
x
y

(9.13)

0749_Frame_C09 Page 125 Wednesday, February 19, 2003 5:09 PM

Element Fields in Linear Problems

125

in which z = 0 at the middle plane. The out-of-plane displacement, w, is assumed


to be a function of x and y only.
An example of an interpolation model is introduced as follows to express w(x, y)
throughout the element in terms of the nodal values of w, wx and wy :
w( x, y, t ) = Tb 2 ( x, y)b 2 b 2 (t )
Tb 2 ( x, y) = 1

Tb 2 (t ) = we

b21 = 0

x2

( )

w
x

w
y

e +1

y2

xy

w
y
w e+2

(9.14)
1 2
( x y + y2 x)
2

x3

we +1

( )

w
x

e+2

w
x

( )

e +1

w
y

e+2

xe ye ( xe + ye )

xe

ye

xe2

xe ye

ye2

xe3

2 xe

ye

3 xe2

( xe ye + 12 ye2 )

xe

2 xe

( 12 xe2 + xe ye )

xe +1

ye +1

xe2+1

xe +1 ye +1

ye2+1

xe3+1

2 xe +1

ye +1

xe +1

2 ye +1

xe + 2

ye + 2

xe2+ 2

xe + 2 ye + 2

ye2+ 2

xe3+ 2

2 xe + 2

ye + 2

xe + 2

2 ye + 2

3 x

3 x

1
2

xe +1 ye +1 ( xe +1 + ye +1 )
( xe +1 ye +1 + 12 ye2+1 )

2
e +1

2
e+2

1
2

( 12 xe2+1 + xe +1 ye +1 )
1
2

y3

xe + 2 ye + 2 ( xe + 2 + ye + 2 )
( xe + 2 ye + 2 + 12 ye2+ 2 )
( 12 xe2+ 2 + xe + 2 ye + 2 )

ye3

3ye2

ye3+1

0 .

3ye2+1

ye3+ 2

3ye2+ 2

It follows that
z b 2
u(x, y, z, t)
x



v(x, y, z, t) = z b 2 ( t ).
b2 b2
y


w(x, y, z, t) b 2

(9.15)

9.1.2.3 Plate with Stretching and Bending


Finally, for a plate experiencing both stretching and bending, the displacements are
assumed to satisfy
u( x, y, z, t ) = u0 ( x, y, z, t ) z

2003 by CRC CRC Press LLC

w
,
x

v( x, y, z, t ) = v0 ( x, y, z, t ) z

w
y

(9.16)

0749_Frame_C09 Page 126 Wednesday, February 19, 2003 5:09 PM

126

Finite Element Analysis: Thermomechanics of Solids

and w is a function only of x, y, and t(not z). Here, z = 0 at the middle surface, while
u0 and v0 represent the in-plane displacements. Using the nodal values of u0, v0, and
w0, a combined interpolation model is obtained as
z b 2
u( x, y, z, t ) u0 ( x, y, t )
x



v( x, y, z, t ) = v ( x, y, t ) + z b 2 (t )
0
b2 b2


w( x, y, z, t ) w0 ( x, y, t ) b 2


m2

= 0T

T
0

z xb 2 m 2

b 2
z y 0

b 2 0

0T
m 2
0T

0
m 2
0

0 u (t )

0 v (t ) .

b 2 w (t )

(9.17)

9.1.2.4 Temperature Field in Two Dimensions


In the two-dimensional, triangular element illustrated in Figure 9.3, the linear interpolation model for the temperature is
T T0 = m 2 m 2 2 ,

2 = (Te T0

Te+1 T0

Te+2 T0 ).

(9.18)

9.1.2.5 Axisymmetric Elements


An axisymmetric element is displayed in Figure 9.4. It is applicable to bodies that
are axisymmetric and are submitted to axisymmetric loads, such as all-around
pressure. The radial displacement is denoted by u, and the axial displacement is
denoted by w. The tangential displacement v vanishes, while radial and axial displacements are independent of . Now u and w depend on r, z, and t.
There are two distinct situations that require distinct interpolation models. In
the first case, none of the nodes are on the axis of revolution (r = 0), while in the
second case, one or two nodes are, in fact, on the axis. In the first case, the linear
e+2
e+1
r

FIGURE 9.4 Axisymmetric element.

2003 by CRC CRC Press LLC

0749_Frame_C09 Page 127 Wednesday, February 19, 2003 5:09 PM

Element Fields in Linear Problems

127

interpolation model is given by


0 T aT1

Ta1 0 T

T
u(r, z, t ) a1

=
w(r, z, t ) 0 T

Ta1 = (1

z ),

a1 = 1

0 T ua1 (t )

.
aT1 wa1 (t )

ze

ze+1 ,

ze+2

re
re+1
re+2

ua1

ue

= ue+1 ,

ue+2

(9.19)

wa1

we

= we+1

we + 2

Now suppose that there are nodes on the axis, and note that the radial displacements are constrained to vanish on the axis. For reasons shown later, it is necessary
to enforce the symmetry constraints a priori in the formulation of the displacement
model. In particular, suppose that node e is on the axis, with nodes e + 1 and e + 2
defined counterclockwise at the other vertices. The linear interpolation model is now
0 T aT2

Ta 2 0 T

T
u(r, z, t ) a 2

=
w(r, z, t ) 0 T

T
a2

= (r

0 T ua 2 (t )

aT2 wa 2 (t )

re+1
a 2 =
re+2

z ze ),

(9.20)

ze+1 ze
.
ze+2 ze

A similar formulation can be used if two nodes are on the axis of symmetry so
that the u displacement in the element is modeled using only one nodal displacement,
with a coefficient vanishing at each of the nodes on the axis of revolution.

9.1.3 INTERPOLATION MODELS: THREE DIMENSIONS


We next consider the tetrahedron illustrated in Figure 9.5.
A linear interpolation model for the temperature can be expressed as
T T0 = 3TT 3T 3
3TT = (1
1

1
3T =
1

z)

xe

ye

xe+1

ye+1

xe + 2

ye+2

xe + 3

ye+3

3TT = {Te To

2003 by CRC CRC Press LLC

(9.21)

Te+1 To

ze

ze+1

ze+2

ze+3

Te+2 To

Te+3 To}

0749_Frame_C09 Page 128 Wednesday, February 19, 2003 5:09 PM

128

Finite Element Analysis: Thermomechanics of Solids

z
e+3

e+2

e+1

FIGURE 9.5 Tetrahedral element.

For elasticity with displacements u, v, and w, the corresponding interpolation


model is
T
u( x, y, z, t ) 3

T
v( x, y, z, t ) = 0


w( x, y, z, t ) 0T

ue (t )

ue+1 (t )
u (t ) =
,
ue+2 (t )

ue+3 (t )

0T
3T
0T

0T 3

0T 0

3T 0

ve ( t )

ve+1 (t )
v (t ) =
,
ve+2 (t )

ve+3 (t )

0
3
0

0 u (t )

0 v (t )

3 w (t )

(9.22)

we (t )

we+1 (t )
w (t ) =

we+2 (t )

we+3 (t )

9.2 STRAIN-DISPLACEMENT RELATIONS


AND THERMAL ANALOGS
9.2.1 STRAIN-DISPLACEMENT RELATIONS: ONE DIMENSION
For the rod, the strain is given by = E11 = ux . An estimate for implied by the
interpolation model in Equation 9.3 has the form

( x, t ) = Tm1 ( x )m1 m1 ( t ),

2003 by CRC CRC Press LLC

(9.23)

0749_Frame_C09 Page 129 Wednesday, February 19, 2003 5:09 PM

Element Fields in Linear Problems

129

Tm1 =

d Tm1
= (0
dx

1)

For the beam, the corresponding relation is

( x , z, t ) = z

w
,
x

(9.24)

from which the consistent approximation is obtained:

( x, z, t ) = z Tb1 ( x )b1 b1 ( t ),
Tb1 =

d Tb1
= (0
dx

(9.25)

3x 2 )

2x

For the beam column, the strain is given by

( x, z, t ) = z Tmb1 ( x )mb1 mb1 ( t ),

Tmb1 = Tm1

m1
z Tb1
0

(9.26)

0 m1 (t )

mb1 b1 (t )

9.2.2 STRAIN-DISPLACEMENT RELATIONS: TWO DIMENSIONS


In two dimensions, the (linear) strain tensor is given by
Exx
E( x, y) =
Exy

Exy
=
Eyy 12

u
x
u + v
y x

1 u + v
2 y x
v
y

).

(9.27)

We will see later the two important cases of plane stress and plane strain. In the
latter case, Ezz vanishes and szz is not needed to achieve a solution. In the former
case, szz vanishes and Ezz is not needed for solution.
In traditional finite-element notation, we obtain [cf. (Zienkiewicz and Taylor, 1989)]
Exx

u2

= Eyy = Tm 2

m2

v2
E
xy

2003 by CRC CRC Press LLC

(9.28)

0749_Frame_C09 Page 130 Wednesday, February 19, 2003 5:09 PM

130

Finite Element Analysis: Thermomechanics of Solids

m 2

m 2
,
y
1 m 2

2 x

m 2
x

= 0
1
ym 2
2

= m2

m2
0

m 2

The prime in e is introduced temporarily to call attention to the fact that it does
not equal VEC(EL). Hereafter, the prime will not be displayed.
For a plate with bending stresses only,
2w
x 2
2
e ( x, y, z, t ) = z w2 ,
y

2w
xy

(9.29)

from which

e ( x, y, z, t ) = z Tb 2 ( x )b 2 b 2 (t ),

b2

2Tb 2
x 2
2T
b2
=
.
2
y
2Tb 2
xy

(9.30)

For a plate experiencing both membrane and bending stresses, the relations can
be combined to furnish
e ( x, y, z, t ) = Tmb 2 ( x, y, z )mb 2 mb 2 (t )

Tmb 2 = Tm 2

m 2
zTb 2 , mb 2 =
0

9.2.3 AXISYMMETRIC ELEMENT

ON

AXIS

OF

(9.31)

0
m 2 ( t )
, mb 2 ( t ) =

b 2
b2 (t )

REVOLUTION

For the toroidal element with a triangular cross section, it is necessary to consider
two cases. If there are no nodes on the axis of revolution, then

e(r, z, t ) =

1
2

2003 by CRC CRC Press LLC

u
r
u
r
w
z

a1
= Ta1

( uz + wr )

0 ua1 (t )

a1 wa1 (t )

(9.32)

0749_Frame_C09 Page 131 Wednesday, February 19, 2003 5:09 PM

Element Fields in Linear Problems

12
Ta1 =
0

131

z
r

1
2

1
2

in which the prime is no longer displayed. If element e is now located on the axis
of revolution, we obtain
a 2
e(r, z, t ) = baT2
0

Ta 2

9.2.4 THERMAL ANALOG

1
=
0

IN

0 ua 2 (t )

a1 wa1 (t )

z ze
r

1
2

1
2

(9.33)

TWO DIMENSIONS

The thermal analog of the strain is the temperature gradient satisfying


0
TT2 =
0

T = TT 2 T 2 2 ,

1
0

0
.
1

(9.34)

9.2.5 THREE-DIMENSIONAL ELEMENTS


Recalling the tetrahedral element in the previous section, the strain relation can be
written as
Exx

Eyy

Ezz
e= =
Exy 12

Eyz 12

E 1
zx 2

2003 by CRC CRC Press LLC

u
x
v
y

(
(

w
z
u + v
y x
v + w
z
y

= 3T 0

)
)

( wx + uz )

0
3
0

0 u 3

0 v 3

3 w 3

(9.35)

0749_Frame_C09 Page 132 Wednesday, February 19, 2003 5:09 PM

132

Finite Element Analysis: Thermomechanics of Solids

9.2.6 THERMAL ANALOG

IN

THREE DIMENSIONS

Again referring to the tetrahedral element, the relation for the temperature gradient is

T = 3TT 3T 3 ,

3TT

= 0

0 0 .

0 1
(9.36)
0

9.3 STRESS-STRAIN-TEMPERATURE RELATIONS


IN LINEAR THERMOELASTICITY
9.3.1 OVERVIEW
If S is the stress tensor under small deformation, the stress-strain relation for a
linearly elastic, isotropic solid under small strain is given in Lames form by
S = 2 EL + tr( EL )I,

(9.37)

in which I is the identity tensor. The Lames coefficients are denoted by and ,
and are given in terms of the familiar elastic modulus E and Poisson ratio as

E
,
2(1 + )

E
.
(1 2 )(1 + )

(9.38)

Letting s = VEC(S) and e = VEC(EL), the stress-strain relations are written using
Kronecker product operators as
s = De,

D = 2 I 9 + ii T ,

(9.39)

and D is the tangent-modulus tensor introduced in the previous chapters.

9.3.2 ONE-DIMENSIONAL MEMBERS


For a beam column, recalling the strain-displacement model,
S11 = ( x, z, t ) = E = z E Tmb1 ( x )mb1 mb1 (t ).

(9.40)

The cases of a rod and a beam are recovered by setting m1 or b1 equal to zero
vectors, respectively.

2003 by CRC CRC Press LLC

0749_Frame_C09 Page 133 Wednesday, February 19, 2003 5:09 PM

Element Fields in Linear Problems

133

9.3.3 TWO-DIMENSIONAL ELEMENTS


9.3.3.1 Membrane Response
In two-dimensional elements, several cases can be distinguished. We first consider
elements in plane stress. It is convenient to use Hookes Law in the form
Exx =

1
[ S ( Syy + Szz )]
E xx

Eyy =

1
[ S ( Syy + Szz )]
E yy

Ezz =

1
[ S ( Sxx + Syy )]
E zz

(9.41)

1+
Exy =
S
E xy
Eyz =

1+
S
E yz

Ezx =

1+
S
E zx

Under plane stress, Sxz = Syz = Szz = 0. Now, Ezz 0, but we will see that it is
not of present interest since it does not influence the solution process. Later on, for
reasons such as tolerances, we may wish to calculate it, but this is a postprocessing
issue. Consequently, in plane stress, the stress-strain relations reduce to
Exx =

1
( S Syy )
E xx

Eyy =

1
( S Sxx )
E yy

Exy =

1+
S
E xy

(9.42)

In traditional finite-element notation, this can be written as


Sxx
Exx


Syy = Dm 21 Eyy .


S
E
xy
xy

(9.43)

in which

Dm21

= E

2003 by CRC CRC Press LLC

1
0

1 +

1
E

=
12
0

1
0

1 +

(9.44)

0749_Frame_C09 Page 134 Wednesday, February 19, 2003 5:09 PM

134

Finite Element Analysis: Thermomechanics of Solids

can be called the tangent-modulus matrix under plane stress (note that, unlike the
previous definition, it is not based on the VEC operator and is not a tensor).
Under plane strain, Exz = Eyz = Ezz = 0. Now, Szz 0. However, this stress is not
of great interest since its value is not needed for attaining a solution. The quantity
may be of interest later, e.g., to check whether yield conditions in plasticity are met,
but it can be obtained at that point by a postprocessing procedure. The equations of
plane strain are
Sxx
E xx


Syy = Dm22 E yy ,


S
E
xy
xy

Dm22 =

E
1 *2

*
1
0

(9.45)

0 , E * =
, * =
2
1

1 + *

and Dm22 is the tangent-modulus tensor in plane strain.


However, in the finite-element method, we will see that we are interested in a
slightly different quantity from Dm21 or Dm22 , as follows. The Principle of Virtual
Work uses the strain-energy density given by 12 Sij Eij. However, elementary manipulation serves to prove that

1
1
Sij Eij = ( Sxx
2
2

Syy

Szz )0

0
1
0

0 Exx

0 Eyy .

2 Exy

(9.46)

In the Principle of Virtual Work, the tangent-modulus tensor in plane stress is


replaced by

Dm21

1
E

=
12
0

1
0

2(1 + )
0

(9.47)

and similarly for plane strain. (This peculiarity is an artifact of traditional finiteelement notation and does not appear if VEC notation is used.) The stresses are now
given in terms of nodal displacements by
Sxx (x, y, t)
Exx


u2

Syy (x, y, t) = Dm 2 i Eyy = Dm 2 i Tm 2


,
m2


2
S (x, y, t)
E
xy

xy

2003 by CRC CRC Press LLC

1
i=
2

plane stress
(9.48)
plane strain

0749_Frame_C09 Page 135 Wednesday, February 19, 2003 5:09 PM

Element Fields in Linear Problems

135

9.3.3.2 Two-Dimensional Members: Bending Response


Thin plates experiencing bending only are assumed to be in a state of plane stress.
The tangent-modulus matrix is given in Equation 9.47, and an approximation for
the strain is obtained as

Sxx
Exx


(t ).
Syy = Dm 21 Eyy z Dm 21 Tb 2
b2 b2


S
E
xy
xy

(9.49)

9.3.4 ELEMENT FOR PLATE WITH MEMBRANE


AND BENDING RESPONSE
Plane stress is also applicable, consequently:

Sxx
Exx


(t ).
Syy = Dm 21 Eyy = Dm 21Tmb 2 ( x, y, z )
mb 2 mb 2


S
E
xy
xy

(9.50)

9.3.5 AXISYMMETRIC ELEMENT


It is sufficient to consider the case in which none of the nodes of the element are
located on the axis of revolution.

Srr
Err

S
E
a1
T
= Da = Da a1
0
Szz
Ezz



S
E
rz
rz
1

Da = E

2003 by CRC CRC Press LLC

0 ua1 (t )

,
wa1 (t )

a1

1 +

(9.51)

0749_Frame_C09 Page 136 Wednesday, February 19, 2003 5:09 PM

136

Finite Element Analysis: Thermomechanics of Solids

For use in the Principle of Virtual Work, Da is modified to furnish Da , given by


1

0
Da = E
0

0 1

2 0

1 +

0
. (9.52)
0

9.3.6 THREE-DIMENSIONAL ELEMENT


All six stresses and strains are now present. Using traditional notation, we write
Sxx
Exx


Syy
Eyy
3

Szz
Ezz
T
= D3 = D33 0

Sxy
Exy
0

Syz
Eyz


S
E
zx
zx
1

D3 = E
0

0 u 3

0 v 3

3 w 3

3
0

1+

1+

1+

1+

1 +

(9.53)

and for the Principle of Virtual Work,


1

D3 = EJ 6
0

2003 by CRC CRC Press LLC

0
J6
0

1 +

(9.54)

0749_Frame_C09 Page 137 Wednesday, February 19, 2003 5:09 PM

Element Fields in Linear Problems

0
J6 =
0

9.3.7 ELEMENTS

FOR

137

CONDUCTIVE HEAT TRANSFER

Assuming the isotropic version of Fouriers Law, the heat-flux vector, which can be
considered the thermal analog of the stress, satisfies
T 1T 11

q = k T 2 T 2 2

T 3 T 3 3

1 D
2D

(9.55)

3 D.

9.4 EXERCISES
1. Modify the rod element to replace the physical coordinate x with the
natural coordinate in which

= ax + b,

2.

3.
4.
5.
6.

( x e ) = 1,

( x e+1 ) = +1.

Rewrite the interpolation model using natural coordinates, and perform


the inverse to obtain m1.
Rewrite the Euler-Bernoulli equation for the beam using the previous
transformation. Rewrite the interpolation model using natural coordinates,
and perform the inverse to obtain b1.
Verify that the inverse given in Equation 9.44 is correct.
Invert Equation 9.39 to express E as a function of s. Find the correct
1
expression for D .
1
Write out the 9 9 tensors D and D that correspond to Lames equation.
For the axisymmetric triangular element, find the strain-displacement
T
matrix in the cases in which two of the nodes are located on the axis
of revolution. Note that for axisymmetric problems, ur = 0 at r = 0.

2003 by CRC CRC Press LLC

0749_Frame_C09 Page 138 Wednesday, February 19, 2003 5:09 PM

138

Finite Element Analysis: Thermomechanics of Solids

7. The stress-strain relations in two-dimensional plane stress can be written


as
Exx
1
1
Eyy =
E
E
0
xy

1
0

0 Sxx

0 Syy

1 + Sxy

Prove that in plane strain


Exx
1
1
Eyy = * *
E
E
0
xy

*
1
0

0 Sxx

0 Syy ,

1 + * Sxy

where
E* = E /(1 2 ), * = /(1 ) .
8. If
Exx 1 1 2
=
Eyy E (1 + )

(1 + ) Sxx
,
1 2 Syy

find the matrix D such that


Sxx
Exx
E
.
D
=
Syy (1 + )(1 2 ) Eyy
What happens if 1/2?

2003 by CRC CRC Press LLC

0749_Frame_C10 Page 139 Wednesday, February 19, 2003 6:04 PM

10

Element and Global


Stiffness and Mass
Matrices

10.1 APPLICATION OF THE PRINCIPLE


OF VIRTUAL WORK
Elements of variational calculus were discussed in Chapter 3, and the Principle of
Virtual Work was introduced in Chapter 5. Under static conditions, the principle is
repeated here as

E S dV + u u dV = u dS.
ij ij

(10.1)

i i

As before, represents the variational operator. We assume for our purposes


that the displacement, the strain, and the stress satisfy representations of the form
u( x, t ) = T ( x ) ( t ),

E = T ( x ) ( t ),

S = DE,

(10.2)

in which E and S are written as one-dimensional arrays in accordance with traditional


finite-element notation. For use in the Principle of Virtual Work, we need D, which
introduces the factor 2 into the entries corresponding to shear. We suppose that the
boundary is decomposed into four segments: S = SI + SII + SIII + SIV . On SI, u is
prescribed, in which event u vanishes. On SII , the traction is prescribed as 0.
On SIII, there is an elastic foundation described by = 0 A(x)u, in which A(x) is
a known matrix function of x. On SIV , there are inertial boundary conditions, by
virtue of which = 0 B. The term on the right now becomes

u dS =
T

i i

SII +SIII +SIV

T ( x ) 0 dS

T ( x )A T ( x )dS (t )

T ( x )B T ( x )dS
(t ).

SIII

(10.3)

SIV

139

2003 by CRC CRC Press LLC

0749_Frame_C10 Page 140 Wednesday, February 19, 2003 6:04 PM

140

Finite Element Analysis: Thermomechanics of Solids

The term on the left in Equation 10.1 becomes

E S dV = K (t),

K=

M=

ij ij

ui u i dV = M ( t ),
T

(x)D (x) dV
T

(10.4)

(x) (x) dV,


T

in which K is called the stiffness matrix and M is called the mass matrix. Canceling
the arbitrary variation and bringing terms with unknowns to the left side furnishes
the equation as follows:
(K + K S ) (t ) + (M + M S )
(t ) = f
f=

KS =

( x )A ( x )dS

MS =

T ( x )B T ( x )dS.

SII +SIII

T ( x ) 0 dS

(10.5)
T

SIII

SIV

Clearly, elastic supports on SIII furnish a boundary contribution to the stiffness


matrix, while mass on the boundary segment SIV furnishes a contribution to the mass
matrix.
Sample Problem 1: One element rod
Consider a rod with modulus E, mass density , area A, and length L. It is built in
at x = 0. At x = L, there is a concentrated mass m to which is attached a spring of
stiffness k, as illustrated in Figure 10.1. The stiffness and mass matrices, from the
domain, reduce to the scalar values K EA/L, M AL/3, MS m, KS k.
AL
+ k ) + ( 3 + m) = f .
The governing equation is ( EA
L
Sample Problem 2: Beam element
Consider a one-element model of a cantilevered beam to which a solid disk is welded
at x = L. Attached at L is a linear spring and a torsional spring, the latter having
the property that the moment developed is proportional to the slope of the beam.
P

E,A,L,

FIGURE 10.1 Rod with inertial and compliant boundary conditions.

2003 by CRC CRC Press LLC

0749_Frame_C10 Page 141 Wednesday, February 19, 2003 6:04 PM

Element and Global Stiffness and Mass Matrices

141

z
y

V0
M0

m
x

E,I,A,L,

kT
k

FIGURE 10.2 Beam with translational and rotational inertial and compliant boundary conditions.

The shear force V0 and the moment M0 act at L. The interpolation model, incorporating the constraints w(0, t) = w (0, t) = 0 a priori, is
L2
w( x , t ) = ( x x )
2 L
2 3

L3 w( L, t )

.
3 L2 w ( L, t )

(10.6)

The stiffness and mass matrices, due to the domain, are readily shown to be
K=

EI 12

L3 6 L

6L
,
4 L2

13
35
M = AL
11
210
L

L
.
2
1
105 L

11
210

(10.7)

The stiffness and mass contributions from the boundary conditions are
k
KS =
0

0
,
kT

m
MS =
0

0
.
mr 2
2

(10.8)

The governing equation is now


( L, t )
w
w( L, t ) V0
(M + M S )
+ (K + K S )
= .
( L, t )
w ( L, t ) M0
w

(10.9)

10.2 THERMAL COUNTERPART OF THE PRINCIPLE


OF VIRTUAL WORK
For our purposes, we focus on the equation of conductive heat transfer as
k 2 T = ce

2003 by CRC CRC Press LLC

T
.
t

(10.10)

0749_Frame_C10 Page 142 Wednesday, February 19, 2003 6:04 PM

142

Finite Element Analysis: Thermomechanics of Solids

Multiplying by the variation of T T0, integrating by parts, and applying the


divergence theorem furnishes

TkTdV + Tc
T

T
dV = Tn T qdS.
t

(10.11)

Now suppose that the interpolation models for temperature in the current element
furnish a relation of the form
T T0 = TT ( x )T (t ),

T = TT ( x )T (t ),

q = kTT ( x )T (t ) .

(10.12)

The terms on the left in Equation 10.11 can now be written as

TkTdV (t)K (t),


T

T
Tce
dV T (t )MT (t ),
t

K T = kTTT TTT dV
(10.13)

c dV.

MT =

T
T

T
T

KT and MT can be called the thermal stiffness (or conductance) matrix and thermal
mass (or capacitance) matrix, respectively.
Suppose that the boundary S has four zones: S = SI + SII + SIII + SIV . On SI, the
temperature is prescribed as T1, from which we conclude that T = 0. On SII, the
T
T
T
heat flux is prescribed as n q1. On SIII, the heat flux satisfies n q = n q1 h1 (T
T
T
T0), while on SIV , n q = n q1 h2 d T/dt. The governing finite-element equation is
now
[M T + M TS ] (t ) + [K T + K TS ]( t) = fT (t )
M TS =

fT (t ) =

n q dS,

SIV

TT T h2 TT T dS,
T
T

K TS =

SIII

(10.14)

TT T h1 TT T dS

= SII + SIII + SIV

10.3 ASSEMBLAGE AND IMPOSITION


OF CONSTRAINTS
10.3.1 RODS
Consider the assemblage consisting of two rod elements, denoted as e and e + 1
[see Figure 10.3(a)]. There are three nodes, numbered n, n + 1, and n + 2. We first
consider assemblage of the stiffness matrices, based on two principles: (a) the forces
at the nodes are in equilibrium, and (b) the displacements at the nodes are continuous.
Principle (a) implies that, in the absence of forces applied externally to the node, at
node n + 1, the force of element e + 1 on element e is equal to and opposite the force
of element e on element e + 1. It is helpful to carefully define global (assemblage

2003 by CRC CRC Press LLC

0749_Frame_C10 Page 143 Wednesday, February 19, 2003 6:04 PM

Element and Global Stiffness and Mass Matrices

143

Pn+2

Pn

e+1

n+1
a. forces in global system

P1(e)

P1(e+1)

e+1
P2 (e+1)

P2 (e)

b. forces in local system

FIGURE 10.3 Assembly of rod elements.

level) and local (element level) systems of notation. The global system of forces is
shown in (a), while the local system is shown in (b). At the center node,
P1( e ) P2( e+1) = 0
(e)
2

since no external load is applied. Clearly, P


The elements satisfy
1
k (e)
1
1
k ( e+1)
1
and in this case, k = k
equations:
(e)

(e+1)

(10.15)
( e +1)
1

= Pn and P

1 un P2( e )

=
1 un+1 P1( e )
1 un+1 P2( e+1)

=
1 un+2 P1( e+1)

(10.16)

= EA/L. These relations can be written as four separate


k ( e ) un k ( e )un+1 = P2( e )

(i)

k ( e ) un + k ( e )un+1 = P1( e )

(ii)

k ( e+1) un+1 k ( e+1)un+2 = P2( e+1)

(iii)

k ( e+1) un+1 + k ( e+1)un+2 = P1( e+1)

2003 by CRC CRC Press LLC

= Pn+2.

(10.17)

(iv)

0749_Frame_C10 Page 144 Wednesday, February 19, 2003 6:04 PM

144

Finite Element Analysis: Thermomechanics of Solids

Add (ii) and (iii) and apply Equation 10.15 to obtain


k ( e ) un k ( e )un+1 = P2( e )

(i)

k ( e ) un + [k ( e ) + k ( e+1) ]un+1 k ( e+1)un+2 = 0

(ii + iii)

k ( e+1) un+1 k ( e+1)un+2 = P1( e+1)

(10.18)

(iv)

and in matrix form


k (e)

k ( e )

k (e)
k ( e ) + k ( e+1)
k ( e+1)

un Pn

k ( e+1) un+1 = 0

k ( e+1) un+2 Pn+1


0

(10.19)

The assembled stiffness matrix shown in Equation 10.19 can be visualized as


an overlay of two element stiffness matrices, referred to global indices, in which
there is an intersection of the overlay. The intersection contains the sum of the lowest
entry on the right side of the upper matrix and the highest entry on the left side of
the lowest matrix. The overlay structure is depicted in Figure 10.4.
K(1)
K(2)
K(3)
K(4)
K(5)
(4)

(5)

k22 +k11

K(N2)
K(N1)
K(N)

FIGURE 10.4 Assembled beam stiffness matrix.

2003 by CRC CRC Press LLC

0749_Frame_C10 Page 145 Wednesday, February 19, 2003 6:04 PM

Element and Global Stiffness and Mass Matrices

145

Now, the equations of the individual elements are written in the global system as
( e ) , K ( e+1) K
( e+1)
K (e) K

(e)
K

= k ( e ) 1

0 ,

1
0

( e+1)
K

= k ( e+1) 0

0
1
1

1 .

(10.20)

(g)

The global stiffness matrix (the assembled stiffness matrix K of the two-element
(e) + K
( e+1) .
member) is the direct sum of the element stiffness matrices: K ( g ) = K
(g)
(
e
)
. In this notation, the strain energy in the two elements
Generally, K = e K
can be written in the form
V( e ) =
V( e+1) =

1 ( e ) ( e ) ( e )
K
2
1 ( e ) ( e+1) ( e )
K
2

( e ) = (un

un+1

(10.21)

un+2 ).

The total strain energy of the two elements is 12 T K ( g ) .


(g)
Finally, notice that K is singular: the sum of the rows is the zero vector, as is
the sum of the columns. In this form, an attempt to solve the system will give rise
to rigid-body motion. To illustrate this reasoning, suppose, for simplicitys sake,
(e)
(e+1)
that k = k , in which case equilibrium requires that Pn = Pn+2. If computations
were performed with perfect accuracy, the equation would pose no difficulty. However, in performing computations, errors arise. For example, Pn is computed as
Pn = Pn + n and Pn+1 = Pn+1 + n+1 , Pn+1 = Pn = P . Computationally, there is now an
unbalanced force, n+1 n. In the absence of mass, this, in principle, implies infinite
accelerations. In the finite-element method, the problem of rigid-body motion can
be detected if the output exhibits large deformation.
The problem is easily suppressed using constraints. In particular, symmetry
implies that un+1 = 0. Recalling Equation 10.19, we now have
1
EA
1
L
0

1
2
1

un Pn

0 = R ,

un+2 Pn+1

(10.22)

in which R is a reaction force that arises to enforce physical symmetry in the presence
of numerically generated asymmetry. The equation corresponding to the second
equation is useless in predicting the unknowns un and un+2 since it introduces the

2003 by CRC CRC Press LLC

0749_Frame_C10 Page 146 Wednesday, February 19, 2003 6:04 PM

146

Finite Element Analysis: Thermomechanics of Solids

new unknown R. The first and third equations are now rewritten as
0 un P + n

,
=
1 un+1 P + n+1

EA 1

L 0

(10.23)

with the solution


EA
,
L

un = [ P + n ]

un+2 = [ P + n+1 ]

EA
.
L

(10.24)

To preserve symmetry, it is necessary for un + un+1 = 0. However, the sum is


computed as
un + un+1 =

n + n+1

( ELA )

(10.25)

The reaction force is given by R = [n + n+1], and, in this case, it can be considered
as a measure of computational error.
Note that Equation 10.23 can be obtained from Equation 10.22 by simply striking
out the second row of the matrices and vectors and the second column of the matrix.
The same assembly arguments apply to the inertial forces as to the elastic forces.
Omitting the details, the kinetic energies T of the two elements are
T=

1 ( e )T ( e ) ( e+1) ( e )
[M + M ]
2

1
( e ) = m( e )
M
1/ 2
1
m( e ) = Al
3

( e ) = {ue

(10.26)

1/ 2

(e)

ue+1

ue+2}

10.3.2 BEAMS
A similar argument applies for beams. The potential energy and stiffness matrix of
th
the e element can be written as
V (e) =

(e)

1 T (e)
K b
2 b

(e)
K11

=
( e )T
K 21

(e)

K12

(e)
K 22

Tb = {we we we+1 we+1}

2003 by CRC CRC Press LLC

(10.27)

0749_Frame_C10 Page 147 Wednesday, February 19, 2003 6:04 PM

Element and Global Stiffness and Mass Matrices

147

In a two-element beam model analogous to the previous rod model, V


(e+1)
V , implying that

(g)

(e)
K 11
( e )T
= K 21
0

K
.

(e)
K 12
( e +1)
K + K 11
K (21e+1)T

( e +1)
12
( e +1)
22

(e)
22

(g)

Generally, in a global coordinate system, K

(g)

=V

(e)

(10.28)

(e)
= e K .

10.3.3 TWO-DIMENSIONAL ELEMENTS


We next consider assembly in 2-D. Consider the model depicted in Figure 10.5
consisting of four rectangular elements, denoted as element e, e + 1, e + 2, and e + 3.
The nodes are also numbered in the global system. Locally, the nodes in an element
are numbered in a counterclockwise fashion. Suppose there is one degree of freedom
per node (e.g., x-displacement) and one corresponding force.
e+3,4

e+3,3

e+2,4

e+3

e+3,1

e+2

e+3,2

e+2,2

e+2

5
e

e+2,1

e+3

e,4

e+2,3

e+1

e,3

e+1,3
e+1,4

e,1

e+1

e,2

FIGURE 10.5 2-D assembly process.

2003 by CRC CRC Press LLC

e+1,1

e+1,2

0749_Frame_C10 Page 148 Wednesday, February 19, 2003 6:04 PM

148

Finite Element Analysis: Thermomechanics of Solids

In the local systems, the force on the center node induces displacements according to
fe,3 = k3(,e1)ue,1 + k3(,e2)ue,2 + k3(,e3)ue,3 + k3(,e4)ue,4
fe+1,4 = k4( e,1+1)ue+1,1 + k4( e,2+1)ue+1,2 + k4( e,3+1)ue+1,3 + k4( e,4+1)ue+1,4
fe+2,1 = k1(,e1+2 )ue+2,1 + k1(,e2+2 )ue+2,2 + k1(,e3+2 )ue+2,3 + k1(,e4+2 )ue+2,4

(10.29)

fe+3,2 = k2( e,1+3)ue+3,1 + k2( e,2+3)ue+3,2 + k2( e,3+3)ue+3,3 + k2( e,4+3)ue+3,4


Globally,
ue,1 u1

ue,2 u2

ue,3 u5

ue,4 u6

ue+1,1 u2

ue+1,2 u3

ue+1,3 u4

ue+1,4 u5

ue+2,1 u5

ue+2,2 u4

ue+2,3 u9

ue+2,4 u8

ue+3,1 u6

ue+3,2 u5

ue+3,3 u8

ue+3,4 u7

(10.30)

Adding the forces of the elements on the center node gives

[
]u + [k

f5 = k3(,e1)u1 + k3(,e2) + k4( e,1+1) u2 + k4( e,2+1)u3 + k4( e,3+1) + k1(,e2+2 ) u4

+ k3(,e3) + k4( e,4+1) + k1(,e1+2 ) + k2( e,2+3)

(e)
3,4

+ k2( e,1+3) u6

+ k2( e,4+3)u7 + k1(,e4+2 ) + k2( e,3+3) u8 + k1(,e3+2 )u9

(10.31)

Taking advantage of the symmetry of the stiffness matrix, this implies that the
fifth row of the stiffness matrix is

5T = k3(,e1)

[k

(e)
3,3

[k

(e)
3,2

+ k4( e,1+1)

k4( e,2+1)

[k

( e +1)
4 ,3

+ k1(,e2+2 )

+ k4( e,4+1) + k1(,e1+2 ) + k2( e,2+3) K symmetry K

]
(10.32)

Finally, for later use, we consider damping, which generates a stress proportional
to the strain rate. In linear problems, it leads to a vector-matrix equation of the form
M
+ D + K = f (t ).

(10.33)

At the element level, the counterpart of the kinetic energy and the strain energy
(e)
is the Rayleigh Damping Function, D , given by D ( e ) = 12 ( e ) T D ( e ) ( e ) , and the conth
sistent damping force on the e element is
fd( e ) (t ) = D( e ) = D( e ) ( e ) .

2003 by CRC CRC Press LLC

(10.34)

0749_Frame_C10 Page 149 Wednesday, February 19, 2003 6:04 PM

Element and Global Stiffness and Mass Matrices

149

The Rayleigh Damping Function is additive over the elements. Accordingly, if


( e ) is the damping matrix of the eth element referred to the global system, the
D
assembled damping matrix is given by
D( g ) =

(e)

(10.35)

It should be evident that the global stiffness, mass, and damping matrices have
the same bandwidth; the force on one given node depends on the displacements
(velocities and accelerations) of the nodes of the elements connected at the given
node, thus determining the bandwidth.

10.3 EXERCISES
1. The equation of static equilibrium in the presence of body forces, such
as gravity, is expressed by
Sij
X j

= bi .

Without the body forces, the dynamic Principle of Virtual Work is derived as

Eij Sij dV + ui

2ui
dV = ui i dS.
t 2

How should the second equation be modified to include body forces?


2. Consider a one-dimensional system described by a sixth-order differential
equation:

d 6q
= 0,
dx 6

Q a constant.

Consider an element from xe to xe+1. Using the natural coordinate = 1


when x = xe = +1 when x = xe+1, for an interpolation model with the
minimum order that is meaningful, obtain expressions for (), , and ,
serving to express q as
q( ) = T ( ) .
3. Write down the assembled mass and stiffness matrices of the following
three-element configuration (using rod elements). The elastic modulus is
E, the mass density is , and the cross-sectional area is A.

2003 by CRC CRC Press LLC

0749_Frame_C10 Page 150 Wednesday, February 19, 2003 6:04 PM

150

Finite Element Analysis: Thermomechanics of Solids

L/2

4. Assemble the stiffness coefficients associated with node n, shown in the


following figure, assuming plane-stress elements. The modulus is E, and
(1)
(2)
(3)
the Poissons ratio is . K , K , and K denote the stiffness matrices
of the elements.

K (3)
K (1)

K (2)

5. Suppose that a rod satisfies = 0, in which is given by


=

1 du
Pu( L)
.
dx
E
2 dx

Use the interpolation model


u( x ) = {1

u( x e , t )
x}
.
u( xe+1, t )

For an element xe < x < xe+1, find the matrix Ke such that
u( xe , t ) fe
Ke
= .
u( xe+1, t ) fe+1
6. Redo this derivation for Ke in the previous exercise, using the natural
coordinate , whereby = ax + b, in which a and b are such that 1 =
axe + b, +1 = axe+1 + b.
7. Next, regard the nodal-displacement vector as a function of t. Find the
matrix Me such that

u( x e , t )
u( xe , t ) fe
+
Me
K

= ,
e
u( xe+1, t )
u( xe+1, t ) fe+1

2003 by CRC CRC Press LLC

0749_Frame_C10 Page 151 Wednesday, February 19, 2003 6:04 PM

Element and Global Stiffness and Mass Matrices

151

in which is the mass density. Derive Me using both physical and natural
coordinates.
8. Show that, for the rod under gravity, a two-element model gives the exact
answer at x = 1, as well as a much better approximation to the exact displacement distribution.

E
A

g
L

9. Apply the method of the previous exercise to consider a stepped rod, as


shown in the figure, with each segment modeled as one element. Is the
displacement at x = 2L still exact?

E1

A1

L1

1
g

E2

A2
2

2003 by CRC CRC Press LLC

L2

0749_Frame_C11 Page 153 Wednesday, February 19, 2003 5:13 PM

11

Solution Methods
for Linear Problems

11.1 NUMERICAL METHODS IN FEA


11.1.1 SOLVING THE FINITE-ELEMENT EQUATIONS: STATIC PROBLEMS
Consider the numerical solution of the linear system K = f, in which K is the
positive-definite and symmetric stiffness matrix. In many problems, it has a large
T
dimension, but is also banded. The matrix can be triangularized: K = LL , in
which L is a lower triangular, nonsingular matrix (zeroes in all entries above the
T
diagonal). We can introduce z = L and obtain z by solving Lz = f. Next, can be
T
computed by solving L = z. Now Lz = f can be conveniently solved by forward
substitution. In particular, Lz = f can be expanded as
l11

l21

l31

l
n1

l22

l32

l33

ln 2

0 z1 f1

. z 2 f2

. z3 f3
= .
. . .

0 . .

lnn zn fn

(11.1)

Assuming that the diagonal entries are not too small, this equation can be solved,
starting from the upper-left entry, using simple arithmetic: z1 = f1/l11, z2 = [ f2 l21z1]/
l22, z3 = [ f3 l31z1 l32z2]/l33, .
T
Next, the equation L = z can be solved using backward substitution. The
equation is expanded as
l11

l12

l22

ln2,n2

ln2,n1

ln1,n1

l1n 1 f1

. 2 .

. 3 .
=
.
ln2,n . fn2

ln1,n . fn1

lnn n fn

(11.2)

153

2003 by CRC CRC Press LLC

0749_Frame_C11 Page 154 Wednesday, February 19, 2003 5:13 PM

154

Finite Element Analysis: Thermomechanics of Solids

Starting from the lower-right entry, solution can be achieved using simple arithmetic as n = fn/lnn,

n1 = [ fn1 ln1,1 n ]/ln1,n1 , n2 = [ fn2 ln2,n n ln2,n1 n1 ]/ln2,n2 , K .


In both procedures, only one unknown is encountered in each step (row).

11.1.2 MATRIX TRIANGULARIZATION


OF LINEAR SYSTEMS

SOLUTION

AND

We next consider how to triangularize K. Suppose that the upper-left ( j 1) ( j


1) block Kj1 has been triangularized: K j 1 = L j 1LTj 1 . In determining whether the
j j block Kj can be triangularized, we consider
K j 1
Kj = T
k j

k j L j 1
=
k jj Tj

0 LTj 1

l jj 0 T

j
,
l jj

(11.3)

in which kj is a ( j 1) 1 array of the first j 1 entries of the j column of Kj.


Simple manipulation suffices to furnish kj and ljj.
th

k j = L j 1 j
(11.4)

l jj = k jj Tj j

Note that j can be conveniently computed using forward substitution. Also,


note that ljj = k jj kTj K j 11k j . The fact that Kj > 0 implies that ljj is real. Obviously,
st
the triangularization process proceeds to the ( j + 1) block and on to the complete
stiffness matrix.
As an example, consider
1

A 3 = 12

13

1
2
1
3
1
4

1
4 .

1
5

1
3

(11.5)

Clearly, L1 = LT1 1. For the second block,


1

2003 by CRC CRC Press LLC

0 1

l22 0

2 1
=
l22 12

,
1
3

1
2

(11.6)

0749_Frame_C11 Page 155 Wednesday, February 19, 2003 5:13 PM

Solution Methods for Linear Problems

155

from which 2 = 1/2 and l22 = 1/3 (1/2) 2 = 1/ 12 . Now


1
L2 =
1/2

0
.
1/ 12

(11.7)

We now proceed to the full matrix:


1

12

13

1
2
1
3
1
4

T
1
4 = L3L3

1
5

1
3

= 12

l31

0 1

0 0

l33 0

0
1/ 12
l32

= 12

l31

1
2

1
2

1/ 12
0

l31

l32

l33

l31 / 2 + l32 / 12

2
2
2
l31
+ l32
+ l33

l31

1
3

l31 / 2 + l32 / 12

2
We conclude that l31 = 1/3, l32 = 1/ 12 , l33
= 1/5 1/9 1/12 =

11.1.3 TRIANGULARIZATION

(11.8)

1
180

ASYMMETRIC MATRICES

OF

Asymmetric stiffness matrices arise in a number of finite-element problems, including problems with unsteady rotation and thermomechanical coupling. If the matrix
is still nonsingular, it can be decomposed into the product of a lower-triangular and
an upper-triangular matrix:
K = LU.

(11.9)

th

Now, the j block of the stiffness matrix admits the decomposition


K j 1
Kj = T
k2 j

k1 j L j 1
=
k jj Tj

L j 1U j 1
= T
j U j 1

0 U j 1

l jj 0 T

Tj u j + u jj l jj

uj

u jj

L j 1u j

(11.10)

in which it is assumed that the ( j 1) block has been decomposed in the previous
step. Now, uj is obtained by forward substitution using Lj1uj = k1j, and j can be
obtained by forward substitution using U Tj 1 j = k 2 j . Finally, u jj l jj = k jj Tj u j , for
th

2003 by CRC CRC Press LLC

0749_Frame_C11 Page 156 Wednesday, February 19, 2003 5:13 PM

156

Finite Element Analysis: Thermomechanics of Solids

which purpose ujj can be arbitrarily set to unity. An equation of the form Kx = f can
now be solved by forward substitution applied to Lz = f, followed by backward
substitution applied to Ux = z.

11.2 TIME INTEGRATION: STABILITY AND ACCURACY


Much insight can be gained from considering the model equation:
dy
= y,
dt

(11.11)

in which is complex. If Re() > 0, for the initial value y(0) = y0, y(t) = y0 exp(t),
then clearly y(t) 0. The system is called asymptotically stable in this case.
We now consider whether numerical-integration schemes to integrate Equation
11.11 have stability properties corresponding to asymptotic stability. For this purpose, we apply the trapezoidal rule, the properties of which will be discussed in a
subsequent section. Consider time steps of duration h, and suppose that the solution
th
has been calculated through the n time step, and we seek to compute the solution at
st
the (n + 1) time step. The trapezoidal rule is given by
dy yn+1 yn

,
dt
h

yn+1 =

1 h/2
y
1 + h/2 n

[ y + yn ].
2 n+1

(11.12)

Consequently,

1 h/2
=
y0
1 + h/2

(11.13)

Clearly, yn+1 0 if 11 + hh//22 < 1, and yn+1 if 11 + hh//22 > 1, in which || implies
the magnitude. If the first inequality is satisfied, the numerical method is called A-stable
(see Dahlquist and Bjork, 1974). We next write = r + ii, and now A-stability
requires that
ih

ih

1 r h
2

1 + r h
2

< 1.

(11.14)

A-stability implies that r > 0, which is precisely the condition for asymptotic
stability.
Consider the matrix-vector system arising in the finite-element method:
M
+ D + K = 0,

2003 by CRC CRC Press LLC

(0) = 0 ,

(0) = 0 ,

(11.15)

0749_Frame_C11 Page 157 Wednesday, February 19, 2003 5:13 PM

Solution Methods for Linear Problems

157

in which M, D, and K are positive-definite. Elementary manipulation serves to derive


that
d
dt

1 T M + 1 T K = T D < 0.
2

(11.16)

It follows that 0 and 0. We conclude that the system is asymptotically


stable.
Introducing the vector p = , the n-dimensional, second-order system is written
in state form as the (2n)-dimensional, first-order system of ordinary differential
equations:
M

0 p
D
+
I
I

K p f
= .
0 0

(11.17)

We next apply the trapezoidal rule to the system:


M

0 h (p n+1 p n ) D

I 1 ( n+1 n ) I
h

1
1

K 2 (p n+1 + p n )

= 2 (fn+1 + fn ) .

0 1 ( n+1 + n )
0

(11.18)

From the equation in the lower row, pn+1 = 2h [n+1 n] pn. Eliminating pn+1
in the upper row furnishes a formula underlying the classical Newmark method:

h
h2
K D = M + D +
K
2
4

K D n+1 = rn+1 ,

rn+1

(11.19)

h
h2
h
h
h2
(f + f )
= M + D
K yn + M + D p n +
2
4
2 2
4 n+1 n

and KD can be called the dynamic stiffness matrix. Equation 11.19 can be solved
by triangularization of KD, followed by forward and backward substitution.

11.3 NEWMARKS METHOD


To fix the important notions, consider the model equation
dy
= f ( y).
dx

(11.20)

Suppose this equation is modeled as

yn+1 + yn + h[ fn+1 + fn ] = 0.

2003 by CRC CRC Press LLC

(11.21)

0749_Frame_C11 Page 158 Wednesday, February 19, 2003 5:13 PM

158

Finite Element Analysis: Thermomechanics of Solids

We now use the Taylor series to express yn+1 and fn+1 in terms of yn and fn. Noting
that yn = fn and yn = fn, we obtain
0 = [ yn + yn h + ynh 2 /2] + yn + h [ yn + ynh] + hyn .

(11.22)

For exact agreement through h , the coefficients must satisfy

+ =0

+ + = 0

/2 + = 0.

(11.23)

We also introduce the convenient normalization + = 1. Simple manipulation


serves to derive that = 1, = 1, = 1/2, = 1/2, thus furnishing
yn+1 yn 1
= [ f ( yn+1 ) + f ( yn )],
2
h

(11.24)

which can be recognized as the trapezoidal rule.


The trapezoidal rule is unique and optimal in having the following three characteristics:
It is a one-step method using only the values at the beginning of the current
time step.
2
It is second-order-accurate; it agrees exactly with the Taylor series through h .
Applied to dy/dt + y = 0, with initial condition y(0) = y0, it is A-stable
whenever a system described by the equation is asymptotically stable.

11.4 INTEGRAL EVALUATION BY


GAUSSIAN QUADRATURE
There are many integrations in the finite-element method, the accuracy and efficiency
of which is critical. Fortunately, a method that is optimal in an important sense,
called Gaussian quadrature, has long been known. It is based on converting physical
coordinates to natural coordinates. Consider ba f ( x )dx. Let = b 1 a [2x (a + b)].
Clearly, maps the interval [a, b] into the interval [1,1]. The integral now becomes
1 1
Now consider the power series
b a 1 f ( )d.
f ( ) = 0 + 1 + 2 2 + 3 3 + 4 4 + 5 5 + L,

(11.25)

from which

f ( )d = 2 1 + 0 +

2
2
+ 0 + 5 + 0 + L.
3 3
5

(11.26)

The advantages illustrated for integration on a symmetric interval demonstrate that,


st
with n function evaluations, an integral can be evaluated exactly through (2n 1) order.

2003 by CRC CRC Press LLC

0749_Frame_C11 Page 159 Wednesday, February 19, 2003 5:13 PM

Solution Methods for Linear Problems

159

Consider the first 2n 1 terms in a power-series representation for a function:


g( ) = 1 + 2 + L + 2 n 2 n1 .

(11.27)

Assume that n integration (Gauss) points i and n weights are used as follows:

g( )d =

g( )w = w + w + L + w
i

i =1

i =1

2 n 1
i i

2n

i =1

(11.28)

i =1

Comparison with Equation 11.26 implies that

wi = 1,

i =1

i =1

wi = 0,

wii2 = 2/3, K,

i =1

wii2 n2 =

i =1

2
,
2n 1

2 n 1
i i

= 0.

i =1

(11.29)
It is necessary to solve for n integration points, i, and n weights, wi. These are
2ni
universal quantities. To integrate a given function, g(), exactly through , it is
necessary to perform n function evaluations, namely to compute g(i).
As an example, we seek two Gauss points and two weights. For n = 2,
w1 + w2 = 2

(i),

w11 + w22 = 0

(ii)

2
3

(iii),

w113 + w223 = 0

(iv)

w112 + w222 =

(11.30)

From (ii) and (iv), w11[12 22 ] = 0, leading to 2 = 1. From (i) and (iii), it
follows that 2 = 1 = 1/ 3. The normalization w1 = 1 implies that w2 = 1.

11.5 MODAL ANALYSIS BY FEA


11.5.1 MODAL DECOMPOSITION
In the absence of damping, the finite-element equation for a linear mechanical
system, which is unforced but has nonzero initial values, is described by
M
+ K = 0,

(0) = 0 ,

(0) = 0 .

(11.31)

Assume a solution of the form = exp(t ), which furnishes upon substitution


[K + 2 M] = 0.

(11.32)

The j eigenvalue, j, is obtained by solving det(K + 2j M) = 0, and a corresponding eigenvector vector, j, can also be computed (see Sample Problem 2).
th

2003 by CRC CRC Press LLC

0749_Frame_C11 Page 160 Wednesday, February 19, 2003 5:13 PM

160

Finite Element Analysis: Thermomechanics of Solids

For the sake of generality, suppose that j and j are complex. Let Hj denote the
complex conjugate (Hermitian) transpose of j. Now, 2j satisfies

2j =

Hj K j
Hj M j

Since M and K are real and positive-definite, it follows that j is pure imaginary:
j = ij. Without loss of generality, we can take j to be real and orthonormal.
Sample Problem 1
As an example, consider
k12
.
k22

(11.33)

(2 )2 + [k11 + k22 ]2 + [k11k22 + k122 ] = 0,

(11.34)

1
M=
0

k11
K=
k12

Now det[K + I] = 0 reduces to


2

with the roots

[
1
= [[k
2

2+, =

1
[k11 + k22 ] [k11 + k22 ]2 4[k11k22 + k122 ]
2
11

+ k22 ] [k11 k22 ]2 4 k122

]
(11.35)

so that both 2+ and 2 are negative (since k11 and k22 are positive).
th
th
We now consider eigenvectors. The eigenvalue equations for the i and j
eigenvectors are written as
[K + 2j M]g j = 0

[K + k2 M]g k = 0

(11.36)

It is easily seen that the eigenvectors have arbitrary magnitudes, and for convenience, we assume that they have unit magnitude Tj j = 1. Simple manipulation
furnishes that
Tk K j Tj K k [ 2j Tk M j k2 Tj M k ] = 0.

2003 by CRC CRC Press LLC

(11.37)

0749_Frame_C11 Page 161 Wednesday, February 19, 2003 5:13 PM

Solution Methods for Linear Problems

161

Symmetry of K and M implies that


Tk K j Tj K k = 0,

[ 2j Tk M j k2 Tj M k ] = ( 2j k2 ) Tk M j = 0.
(11.38)

Assuming for convenience that the eigenvalues are all distinct, it follows that
Tj M k = 0,

Tj K k = 0,

j k.

(11.39)

The eigenvectors are thus said to be orthogonal with respect to M and K. The
th
th
quantities j = Tj M j and j = Tj K j are called the ( j ) modal mass and ( j )
modal stiffness.
Sample Problem 2
Consider
0 1

1 2

k 2

m 1

1 1 0
= .
1 2 0

(11.40)

2
2
2
2
Let = j / 0 , 0 = k/m. For the determinant to vanish, 1 2 = 1/ 2 .
Using 1 2 = 1/ 2 , the first eigenvector satisfies

(1)
1 1 0

=
,
(1)
1/ 2 2 0

[ ] + [ ]
(1) 2
1

(1) 2
2

= 1,

(11.41)

implying that 1(1) = 1/ 3 , 2(1) = 2 / 3 . The corresponding procedures for the second eigenvalue furnish that 1( 2 ) = 1/ 3 , 2( 2 ) = 2 / 3 . It is readily verified that
T

1 = 4/3,

1 =

( 2 ) M (1) = (1) M ( 2 ) = 0,
( 2 ) K (1) = (1) K ( 2 ) = 0,

2 = 4/3,

4
[1 1/ 2 ],
3

2 =

4
[1 + 1/ 2 ]
3

(11.42)

The modal matrix X is now defined as


X = [ 1

2003 by CRC CRC Press LLC

n ].

(11.43)

0749_Frame_C11 Page 162 Wednesday, February 19, 2003 5:13 PM

162

Finite Element Analysis: Thermomechanics of Solids


th

Since the jk entries of X MX and X KX are Tj M k and Tj K k , respectively,


it follows that
1

X T MX = .

X T KX = .

. . (11.44)

The modal matrix is said to be orthogonal with respect to M and K, but it is


1
T
not purely orthogonal since X X .
The governing equation is now rewritten as
+ X T KX = g,
X T MX

= X 1 ,

g = XTf,

(11.45)

implying the uncoupled modes

j j + j j = g j (t ).

(11.46)

Suppose that gj(t) = gj0 sin ( t). Neglecting transients, the steady-state solution
th
for the j mode is

j =

gj0

j 2 j

sin( t ).

(11.47)

It is evident that if 2 ~ 2j = j / j (resonance), the response amplitude for the


j mode is much greater than for the other modes, so that the structural motion under
this excitation frequency illustrates the mode. For this reason, the modes can easily
be animated.
th

11.5.2 COMPUTATION

OF

EIGENVECTORS

AND

EIGENVALUES

Consider K j = 2j M j , with Tj j = 1. Many methods have been proposed to


compute the eigenvalues and eigenvectors of a large system. Here, we describe a
method that is easy to visualize, which we call the hypercircle method. The vectors
Kj and Mj must be parallel to each other. Furthermore, the vectors K j / Tj K 2 j
and M j / Tj M2 j must terminate at the same point in a hypersphere in nth
dimensional space. Suppose that (j ) is the iterate and that the two vectors

2003 by CRC CRC Press LLC

0749_Frame_C11 Page 163 Wednesday, February 19, 2003 5:13 PM

Solution Methods for Linear Problems

163

do not coincide in direction. Another iterate can be attained by an interval-halving


method:

M (j +1) =

M ( )
K ( )
1
j
j
+
T
2
(

)
2 ( ) T M 2 ( )
K ( )
j
j
j
j

(j +1) = (j +1)

T
(j +1) M2 (j +1) .

(11.48)
Alternatively, note that
C( j ) =

Tj K

M j

Tj K 2 j

Tj M 2 j

(11.49)

is the cosine of the angle between two unit vectors, and as such it assumes the
maximum value of unity when the vectors coincide. A search can be executed on
the hypersphere in the vicinity of the current point, seeking the path along which C(j)
increases until the maximum is attained.
Once the eigenvector j is found, the corresponding eigenvalue is found from
2
j = ( Tj K j )/( Tj M j ). Now, an efficient scheme is needed to deflate the system
so that 12 ,and 1 no longer are part of the eigenstructure in order to ensure that the
solution scheme does not converge to values that have already been calculated.
Given 1 and 12 , we can construct a vector p2 that is M-orthogonal to 1 by
using an intermediate vector p 2 : p2 = p 2 1T Mp 2 1. Clearly, 1T Mp2 = 1T Mp 2
1T Mp 2 1T M 1 = 0 assuming 1 = 1. However, p2 is also clearly orthogonal to K1
since it is collinear with M1. A similar procedure leads to vectors pj, which are orthogonal to each other and to M1 and K1. For example, with p2 set to unit magnitude,
p 3 = p 3 p T2 p 3p 2 1T Mp 3 1 . Now, p T2 p 3 = 0 and 1T Mp 3 = 0.
th
Introduce the matrix X1 as follows: X1 = [1 p2 p3 pn]. For the k eigenvalue, we can write

[X KX
T
1

2k X1TMX1 X11 k = 0,

(11.50)

which decomposes to
12

0T

0
1
k2

0 T
K
n 1

0 0 0
.
=

0
M
n 1
n 1

(11.51)

This implies the deflated eigenvalue problem


M
] = 0.
[K
n 1
n 1
n 1
k

(11.52)

The eigenvalues of the deflated system are also eigenvalues of the original system.
The eigenvector n1 can be used to compute the eigenvectors of the original system

2003 by CRC CRC Press LLC

0749_Frame_C11 Page 164 Wednesday, February 19, 2003 5:13 PM

164

Finite Element Analysis: Thermomechanics of Solids

using transformations involving the matrix X1. The eigenvalues and eigenvectors of
the deflated system can be computed by, for example, the hypercircle method
described previously.

11.6 EXERCISES
1. Verify that the triangular factors L3 and LT3 for A3 in Equation 11.5 are correct.
2. Using A3 in Equation 11.5, use forward substitution followed by backward
substitution to solve
1

A 3 = 1 .

1
3. Triangularize the matrix
36

K = 30

18

30
41
23

18

23.

14

4. For the model equation dy/dx = f(y), develop a two-step numerical-integration


model:
(yn+1 + yn + yn1 ) + h[ f ( yn+1 ) + f ( yn ) + f ( yn1 )] = 0.
What is the order of the integration method (highest power in h with exact
agreement with the Taylor series)?
5. Find the integration (Gauss) points and weights for n = 3.
6. In the damped linear-mechanical system
M
+ D + K = f (t ),
th
suppose that (t) = n at the n time step. Derive KD and rn+1 such that
st
at the (n + 1) time step satisfies

K D n+1 = rn+1 .
7. For the linear system
36

30

24

30
41
32

24 1 1

32 2 = 2

27 3 3

triangularize the matrix and solve for 1, 2, 3.

2003 by CRC CRC Press LLC

0749_Frame_C11 Page 165 Wednesday, February 19, 2003 5:13 PM

Solution Methods for Linear Problems

165

8. (a) Find the modal masses 1, 2 and the modal stiffnesses 1 and 2 of
the system
1
3
0

0 1
1
+ 27
2 2
1

1 1 10
= sin(10t ).
2 2 20

(b) Determine the steady-state response of the system (i.e., particular


solution to the equation).
9. Triangularize
2 A/ L

0
4 AL/Y 2
2 A/ L

2 A/ L .

10. Put the following equations in state form, apply the trapezoidal rule, and
triangularize the ensuing dynamic stiffness matrix:
M
+ K = f ,

2003 by CRC CRC Press LLC

T = 0.

0749_Frame_C12 Page 167 Wednesday, February 19, 2003 6:08 PM

12

Rotating and Unrestrained


Elastic Bodies

12.1 FINITE ELEMENTS IN ROTATION


We first consider rotation about a fixed axis. The coordinate system is embedded in
the fixed point and rotates. The undeformed position vector, X in the rotated system
is related to its counterpart, X, in the unrotated system by X = Q(t )X. The counterpart
for the deformed position is x = Q(t )x. The displacement also satisfies u = Q(t )u.
(t )Q T (t )(*).
The rotation is represented by the axial vector, , satisfying (*) = Q
T
(t )Q (t ) is antisymmetric). The time derivatives in rotating coordinates
(Recall that Q
satisfy
du u
=
+ u
dt
t
d 2 u 2 u
= 2 + 2 u + u + u
dt 2
t

(12.1)

(.)
where = t and t imply differentiation with the coordinate system instantaneously fixed. The rightmost four terms in (12.1) are called the translational, Coriolis,
centrifugal, and angular accelerations, respectively.
2
Applied to the Principle of Virtual Work, the inertial term becomes u T d 2
dt
[u + X]dV . Assuming that u = T (X) (t ),

u T

d 2
d
d 2 u
T
dV
=

+ G1
+ (G 2 + A )
2

2
dt
dt
dt

G1 = T T d V,

= T T d V.

M = T T d V,
G2 = T 2 T d V,

(12.2)

The matrix M is the conventional positive-definite and symmetric mass matrix; the
Coriolis matrix G1 is antisymmetric; the centrifugal matrix G2 is negative-definite; and
= QQ
T, A =
.
the angular acceleration matrix A is antisymmetric. Also,
167

2003 by CRC CRC Press LLC

0749_Frame_C12 Page 168 Wednesday, February 19, 2003 6:08 PM

168

Finite Element of Analysis: Thermomechanics of Solids

r
E,A,L,

f0 sin t

FIGURE 12.1 Elastic rod on rotating rigid shaft.

There is also a rigid-body force term:

u T

d2
Xd V = T frot ,
dt 2

frot = T [2 + A]Xd V.

(12.3)

The governing equation is now


M

d2
d
+ G1
+ [K + G 2 + A] = f frot .
dt 2
dt

(12.4)

Consider a rod attached to a thin shaft rotating steadily at angular velocity


(see Figure 12.1), with f0 = 0.
If r is the undeformed position along the shaft, the governing equation is
EA

d 2u
= 2 [r + u].
dr 2

(12.5)

Assuming a one-element model with u(r, t ) = ru( L, t ) L , we obtain


EA 2 AL
2
L
u( L, t ) = A
3

=
=

2 AL2
3

r
rdr
L
(12.6)

Clearly, u(L, t) becomes unbounded if becomes equal to the natural frequency


AL , in which case, is called a critical speed.
EA
3
3

2003 by CRC CRC Press LLC

0749_Frame_C12 Page 169 Wednesday, February 19, 2003 6:08 PM

Rotating and Unrestrained Elastic Bodies

12.2

169

FINITE-ELEMENT ANALYSIS FOR


UNCONSTRAINED ELASTIC BODIES

Consider the response of an elastic body that has no fixed points, as in spacecraft.
It is assumed that the tractions are prescribed on the undeformed surface. The
response is referred to the body axes corresponding to axes embedded in the
corresponding rigid body, for example, the principal axes of the moment of inertia
tensor. The position vector r of a point in the body can be decomposed as follows:
r = rc + + u ,

(12.7)

in which rc is the position vector to the center of mass; is the relative positionvector from the center of the mass to the undeformed position of the current point,
referred to rotating axes; and u is the displacement from the undeformed to the
deformed position in the rotating system (see Figure 12.2).
The balance of linear momentum becomes
Sij
j

= [
rci + i + ui ].

(12.8)

Recall that = , and the corresponding variational relation is = u ,


= . It follows that r = rc + ( + u) + u, in which u is the variation of
u with the axes instantaneously held fixed. The quantities rc, , and u can be varied
independently since there is no constraint relating them. For rc, we have
Sij

r
ci

rci + i + ui ] dV = 0.
[

(12.9)

z
u(,t)
(t)

rc(t)

y
x
FIGURE 12.2 3-D unconstrained 3-D element.

2003 by CRC CRC Press LLC

0749_Frame_C12 Page 170 Wednesday, February 19, 2003 6:08 PM

170

Finite Element of Analysis: Thermomechanics of Solids

Now rci i dV = rci i dV = 0 by virtue of the definition of the center of mass.


Next rci ui dV = 0 on the assumption, for which a strong argument can be made,
that deformation does not affect the position of the center of mass. Then,
rci
rci dV = rci
rci dV = rci mrci . Finally,
Sij

ci

dV = rci

Sij

dV

= rci Sij n j dS

= rci i dS
= rci Pi

(12.10)

Consequently P = mrc, as expected.

Consider the variation with respect to : d = d

i ij [
rci + i + ui ] dV = 0.
j

(12.11)

First,

r dV = r dV,
i

ci

ci

(12.12)

and [ i dV ] = 0, Recall that = and the corresponding variational relation


is = , = . Now,

u d V = [ ] u dV = udV.
T

(12.13)

It is common to assume that udV = 0, which also implies that ui dV = 0.


This assumption implies that the body axes in the deformed configuration are not
affected by deformation and are obtained by rotating the undeformed axes according
to rigid-body relations.
Next, consider i:

dV = [ + ]dV
T

= T [ + ]dV

= T [ J + J ]

2003 by CRC CRC Press LLC

(12.14)

0749_Frame_C12 Page 171 Wednesday, February 19, 2003 6:08 PM

Rotating and Unrestrained Elastic Bodies

171

in which J = 2 dV, and (.) = (.). Clearly, as the matrix corresponding to a


2
vector product, is antisymmetric. We determined in an earlier discussion that is
negative-definite, thus, the moment-of-inertia tensor J is positive-definite. Finally,
Sij

dV =

( S )dV S dV
i ij

ij

= (i Sij ni )dS

= iti dS
= T m,

m = t dS

(12.15)

The vector-valued quantity m is recognized as the moment of the tractions


referred to the center of mass. We thus obtain the Euler equation of a rigid body as
m = J + J .

(12.16)

Equations 12.10 and 12.16 can be solved separately from the finite-element
equation for the displacements, to be presented next, to determine the origin and
orientation of the body axes.
Now consider the variation u:
ij

u
i

[
rci + i + ui ] dV = 0.

(12.17)

First, note that uidVrci = 0 since uidV = 0. The remaining terms are exactly
the same as for a body with a fixed point, and consequently it reduces to
M

d2
d
+ G1
+ [K + G 2 + A] = f frot
dt 2
dt

G2 = T 2 T dV,

frot = T [2 + A]XdV

G1 = T T dV,
A = T A T dV,

(12.18)

12.3 EXERCISES
1. Consider a one-element model for a steadily rotating rod (see Figure 12.1)
to which is applied an oscillatory force, as shown. Find the steady-state
solution.
2. Find the exact solution for u(r) in Exercise 1. Does it agree with the oneelement solution? Try two elements. Is the solution better? By how much?

2003 by CRC CRC Press LLC

0749_Frame_C12 Page 172 Wednesday, February 19, 2003 6:08 PM

172

Finite Element of Analysis: Thermomechanics of Solids

3. Prove that the inertia tensor J is positive-definite.


Find J for the brick element with density .

b
a

4. Write down the proof that for an unrestrained elastic body,


if

udV = 0 and udV = 0,

then
F = mr0

M = J + J.

5. Consider a thin beam column that is rotating unsteadily around a shaft.


Its thin (local z) direction points in the direction of the motion, giving
rise to Coriolis effects in bending. Derive the ensuing one-element model.
Note that this situation couples extension and bending.

2003 by CRC CRC Press LLC

0749_Frame_C13 Page 173 Wednesday, February 19, 2003 5:19 PM

13

Thermal, Thermoelastic,
and Incompressible Media

13.1 TRANSIENT CONDUCTIVE-HEAT TRANSFER


13.1.1 FINITE-ELEMENT EQUATION
The governing equation for conductive-heat transfer without heat sources, assuming
an isotropic medium, is
k 2 T = ce T .

(13.1)

With the interpolation model T(t) T0 = TT (x)T( t ) and T = TT T (t ), the


finite-element equation assumes the form
K T + MT = q(t )

K T = TT T k TT T dV ,

MT = TT T ceTT TT dV

(13.2)

This equation is parabolic (first-order in the time rates), and implies that the
temperature changes occur immediately at all points in the domain, but at smaller
initial rates away from where the heat is added. This contrasts with the hyperbolic
(second-order time rates) solid-mechanics equations, in which information propagates into the medium as finite velocity waves, and in which oscillatory response
occurs in response to a perturbation.

13.1.2 DIRECT INTEGRATION

BY THE

TRAPEZOIDAL RULE

Equation 13.1 is already in state form since it is first-order, and the trapezoidal rule
can be applied directly:
MT

q + qn
n+1 n
+ n
+ K T n+1
= n+1
,
h
2
2

(13.3)

from which
K DT n+1 = rn+1
K DT = MT +

(13.4)

h
h
h
K r = MT n K T n (q n+1 + q n )
2 T n+1
2
2
173

2003 by CRC CRC Press LLC

0749_Frame_C13 Page 174 Wednesday, February 19, 2003 5:19 PM

174

Finite Element Analysis: Thermomechanics of Solids

For the assumed conditions, the dynamic thermal stiffness matrix is positivedefinite, and for the current time step, the equation can be solved in the same manner
as in the static counterpart, namely forward substitution followed by backward
substitution.

13.1.3 MODAL ANALYSIS


Modes are not of much interest in thermal problems since the modes are not
oscillatory or useful to visualize. However, the equation can still be decomposed
into independent single degree of freedom systems. First, we note that the thermal
system is asymptotically stable. In particular, suppose the inhomogeneous term
T
vanishes and that at t = 0 does not vanish. Multiplying the equation by and
elementary manipulation furnishes that
d T MT
T

= K T < 0.
dt 2

(13.5)

T
Clearly, the product MT decreases continuously. However, it only vanishes if
vanishes.
To examine the modes, assume a solution of the form (t) = 0j exp(jt). The
eigenvectors 0j satisfy

j =k
,
T0 j MT 0 k = Tj
jk
0

j = k
T0 j K T 0 k = Tj
jk
0

(13.6)

and we call Tj and Tj the j modal thermal mass and j modal thermal stiffness,
01 0n ], and again
respectively. We can also form the modal matrix = [
th

Tj

TMT = .

th

Tj

. ,

Tj

TK T = .

Tj

. .

.
(13.7)

1
T
T
Let = and g(t) = q(t). Pre- and postmultiplying Equation 13.2 with
and , respectively, furnishes the decoupled equation

Tj j + Tj j = g j .

2003 by CRC CRC Press LLC

(13.8)

0749_Frame_C13 Page 175 Wednesday, February 19, 2003 5:19 PM

Thermal, Thermoelastic, and Incompressible Media

175

Suppose, for convenience, that gj is a constant. Then, the general solution is of


the form

j = j 0 exp Tj
Tj

t +

exp Tj (t T ) gj d ,
0
Tj

(13.9)

illustrating the monotonically decreasing nature of the response. Now there are n
uncoupled single degrees of freedom.

13.2 COUPLED LINEAR THERMOELASTICITY


13.2.1 FINITE-ELEMENT EQUATION
The classical theory of coupled thermoelasticity accommodates the fact that the
thermal and mechanical fields interact. For isotropic materials, assuming that temperature only affects the volume of an element, the stress-strain relation is
Sij = 2 Eij + ( Ekk (T T0 )) ij ,

(13.10)

in which denotes the volumetric thermal-expansion coefficient. The equilibrium


S
equation is repeated as xij = ui . The Principle of Virtual Work implies that
j

E [2E + E ] dV + u dV E (T T )
ij

ij

kk ij

ij

ij

dVo = u j t j dSo .
(13.11)

Now consider the interpolation models


u = N T ( x ) (t ),

E ij E = BT ( x ) (t ),

T T0 = v T ( x ) (t ), T = BTT ( x ) (t ),
(13.12)

in which E is the strain written as a column vector in conventional finite-element


notation. The usual procedures furnish the finite-element equation
M
(t ) + K (t ) (t ) = f (t ),

= B T dVo .

(13.13)

The quantity is the thermomechanical stiffness matrix. If there are nm displacement degrees of freedom and nt thermal degrees of freedom, the quantities appearing
in the equation are
M, K: nm nm ,

2003 by CRC CRC Press LLC

(t ), f (t ): nm 1,

: nm nt ,

(t ): nt 1.

0749_Frame_C13 Page 176 Wednesday, February 19, 2003 5:19 PM

176

Finite Element Analysis: Thermomechanics of Solids

We next address the thermal field. The energy-balance equation (from Equation
7.35), including mechanical effects, is given by
k 2 T = ce T + T0 tr( E ).

(13.14)

Application of the usual variational methods imply that

K T (t ) + MT (t ) + T0 T (t ) = q,

q=

n qdS.

(13.15)

Case 1:
Suppose that T is constant. At the global level, (t ) = T0 K T1 T (t ) + T0 K T1 q. Thus,
the thermal field is eliminated at the global level, giving the new governing equation
as
(t ) + T0 K T1 T (t ) + K (t ) = f (t ).
M

(13.16)

Conductive-heat transfer is analogous to damping. The mechanical system is


now asymptotically stable rather than asymptotically marginally stable.
We next put the global equations in state form:
Q1z + Q 2 z = f
M

Q1 = 0

Q 2 = K

0
K
0
K
0
0

M T / T0
0

K T / T0

(13.17)


z =


f

f = 0

q / T0

Clearly, Equation 13.17 can be integrated numerically using the trapezoidal rule:
Q + h Q z = Q h Q z + h [f + f ].
1 2 2 n+1 1 2 2 n 2 n+1 n

2003 by CRC CRC Press LLC

(13.18)

0749_Frame_C13 Page 177 Wednesday, February 19, 2003 5:19 PM

Thermal, Thermoelastic, and Incompressible Media

177

Now, consider asymptotic stability, for which purpose it is sufficient to take f = 0,


T
z(0) = z0. Upon premultiplying Equation 13.17 by z , we obtain
d 1 T
z Q1z = z T Q 2 z

dt 2
= z T

1
Q + Q 2T z
2 2

= T K T

(13.19)

and z must be real. Assuming that 0, it follows that z 0, and hence the system
is asymptotically stable.

13.2.2 THERMOELASTICITY

IN A

ROD

Consider a rod that is built into a large, rigid, nonconducting temperature reservoir
at x = 0. The force, f0, and heat flux, q0, are prescribed at x = L. A single element
models the rod. Now,
u( x, t ) = x (t )/ L,

E( x, t ) = (t )/ L,

T T0 = x (t )/ L,

dT
= (t )/ L.
dx

(13.20)

The thermoelastic stiffness matrix becomes = B dV = A/2. The


governing equations are now
T

1
AL EA
+
A = f0
2
3
L
1
1 ce AL 1 kA
+
+ A = q 0
2
T0 3
T0 L

(13.21)

13.3 COMPRESSIBLE ELASTIC MEDIA


For a compressible elastic material, the isotropic stress Skk and the dilatational strain
Ekk are related by Skk = 3Ekk, in which the bulk modulus satisfies = E/[3(1 2)].
Clearly, as 1/2, the pressure, p = Skk /3, needed to attain a finite compressive
volume strain (Ekk < 0) becomes infinite. At the limit = 1/2, the material is said
to satisfy the internal constraint of incompressibility.
Consider the case of plane strain, in which Ezz = 0. The tangent modulus matrix
D is readily found from
12
Sxx


E
(1 + )
Syy =
(1 + )(1 2 )

0
Szz

2003 by CRC CRC Press LLC

(1 + )
12
0

0 Exx

0 Eyy .

1 2 Ezz

(13.22)

0749_Frame_C13 Page 178 Wednesday, February 19, 2003 5:19 PM

178

Finite Element Analysis: Thermomechanics of Solids

Clearly, D becomes unbounded as 1/2. Furthermore, suppose that for a


material to be nearly incompressible, is estimated as .495, while the correct value
is .49. It might be supposed that the estimated value is a good approximation for
1
the correct value. However, for the correct value, (1 2) = 50. For the estimated
1
value, (1 2) = 100, implying 100 percent error!

13.4 INCOMPRESSIBLE ELASTIC MEDIA


In an incompressible material, a pressure field arises that serves to enforce the
constraint. Since the trace of the strains vanishes everywhere, the strains are not
sufficient to determine the stresses. However, the strains together with the pressure
are sufficient. In FEA, a general interpolation model is used at the outset for the
displacement field. The Principle of Virtual Work is now expressed in terms of
the displacements and pressure, and an adjoining equation is introduced to enforce
the constraint a posteriori. The pressure can be shown to serve as a Lagrange
multiplier, and the displacement vector and the pressure are varied independently.
In incompressible materials, to preserve finite stresses, we suppose that the
second Lame coefficient satisfies as tr(E) 0 in such a way that the product
is an indeterminate quantity denoted by p:

tr( E) p.

(13.23)

The Lame form of the constitutive relations becomes


Sij = 2 Eij p ij ,

(13.24)

together with the incompressibility constraint Eijij = 0. There now are two independent principal strains and the pressure with which to determine the three principal
stresses.
w
In a compressible elastic material, the strain-energy function w satisfies Sij = Eij ,
and the domain term in the Principle of Virtual Work can be rewritten as EijSijdV =
wdV. The elastic-strain energy is given by w = Ei j Ei j + 2 Ekk2. For reasons
explained shortly, we introduce the augmented strain-energy function
w = Eij Eij pEkk

(13.25)

and assume the variational principle

wdV + u u dV = u dS .
T

2003 by CRC CRC Press LLC

(13.26)

0749_Frame_C13 Page 179 Wednesday, February 19, 2003 5:19 PM

Thermal, Thermoelastic, and Incompressible Media

179

Now, considering u and p to vary independently, the integrand of the first term
becomes w = Eij[2Eij pij ] pEkk , furnishing two variational relations:

E S dV + u u dV = u dS
T

ij ij

(a)

(13.27)

pE dV = 0
kk

(b)

The first relation is recognized as the Principle of Virtual Work, and the second
equation serves to enforce the internal constraint of incompressibility.
We now introduce the interpolation models:
u = N T ( x ) g (t )
Ekk = bT ( x ) g(t )

e = BT ( x ) g ( t )
p( x ) = x T ( x ) p(t )

(13.28)

Substitution serves to derive that


M(t ) + K (t ) (t ) = f

= b T dVo ,

(13.29)

T (t ) = 0

Assuming that these equations apply at the global level, use of state form
furnishes
M

0 T

0
K
0T

0 (t ) 0


d

0 (t ) + K

dt


0 (t ) T

K
0
0T

(t ) f (t )

0 (t ) = 0 .


0 (t ) 0

(13.30)

The second matrix is antisymmetric. Furthermore, the system exhibits marginal


asymptotic stability; namely, if f(t) = 0 while (0), (0), and (0) do not all vanish,
then

d 1 T
( (t )
dt 2

2003 by CRC CRC Press LLC

T (t )

T (t )) 0

0 T

0
K
0T

0 (t )

0 (t ) = 0

0 (t )

(13.31)

0749_Frame_C13 Page 180 Wednesday, February 19, 2003 5:19 PM

180

Finite Element Analysis: Thermomechanics of Solids

13.5 EXERCISES
1. Find the exact solution for a circular rod of length L, radius r, mass density
, specific heat ce , conductivity k, and cross-sectional area A = r2. The
initial temperature is T0, and the rod is built into a large wall at fixed
temperature T0 (see figure below). However, at time t = 0, the temperature
T1 is imposed at x = L. Compare the exact solution to the one- and twoelement solutions. Note that for a one-element model,

c AL
kA
( L, t ) + e ( L, t ) = q( L).
3
L
L
r

T0

T1,t>0

2. State the equations of a thermoelastic rod, and put the equations for the
thermoelastic behavior of a rod in state form.
3. Put the following equations in state form, apply the trapezoidal rule, and
triangularize the ensuing dynamic stiffness matrix, assuming that the triangular factors of M and K are known.
+ K = f,
M

T = 0.

4. In an element of an incompressible square rod of cross-sectional area A,


it is necessary to consider the displacements v and w. Suppose the length
is L, the lateral dimension is Y, and the interpolation models are linear for
the displacements (u linear in x, with v,w linear in y) and constant for the
pressure. Show that the finite-element equation assumes the form
2 A / L

0
4 AL / Y 2
2 AL / Y

u( L) f


2 AL / Y v(Y ) = 0


0 p 0

and that this implies that 3 u( L) = f (which can also be shown by an a


L
priori argument).

2003 by CRC CRC Press LLC

0749_Frame_C14 Page 181 Wednesday, February 19, 2003 5:41 PM

14

Torsion and Buckling

14.1 TORSION OF PRISMATIC BARS


Figure 14.1 illustrates a member experiencing torsion. The member in this case is
cylindrical with length L and radius r0. The base is fixed, and a torque is applied at
the top surface, which causes the member to twist. The twist at height z is (z), and
at height L, it is 0.
Ordinarily, in the finite-element problems so far considered, the displacement is
the basic unknown. It is approximated by an interpolation model, from which an
approximation for the strain tensor is obtained. Then, an approximation for the stress
tensor is obtained using the stress-strain relations. The nodal displacements are
solved by an equilibrium principle, in the form of the Principle of Virtual Work. In
the current problem, an alternative path is followed in which stresses or, more
precisely, a stress potential, is the unknown. The strains are determined from the
stresses. However, for arbitrary stresses satisfying equilibrium, the strain field may
not be compatible. The compatibility condition (see Chapter 4) is enforced, furnishing

z
T
0
r0
section after twist

section before twist


L

x
z
y

x
FIGURE 14.1 Twist of a prismatic rod.
181

2003 by CRC CRC Press LLC

0749_Frame_C14 Page 182 Wednesday, February 19, 2003 5:41 PM

182

Finite Element Analysis: Thermomechanics of Solids

a partial differential equation known as the Poisson Equation. A variational argument


is applied to furnish a finite-element expression for the torsional constant of the
section.
For the member before twist, consider points X and Y at angle and at radial
position r. Clearly, X = r cos and Y = r sin . Twist induces a rotation through angle
(z), but it does not affect the radial position. Now, x = r cos( + ), y = r sin( + ).
Use of double-angle formulae furnishes the displacements, and restriction to small
angles furnishes, to first order,
u = Y ,

v = X .

(14.1)

It is also assumed that torsion does not increase the length of the member, which
is attained by requiring that axial displacement w only depends on X and Y. The
quantity w(X, Y) is called the warping function.
It is readily verified that all strains vanish except Exz and Eyz, for which
Exz =

1 w

y ,

2 x
z

1 w

+ x .

2 y
z

Eyz =

(14.2)

Equilibrium requires that


Sxz Syz
+
= 0.
x
y

(14.3)

The equilibrium relation can be identically satisfied by a potential function y


for which
Sxz =

,
y

Syz =

.
x

(14.4)

We must satisfy the compatibility condition to ensure that the strain field arises
from a displacement field that is unique to within a rigid-body translation and
rotation. (Compatibility is automatically satisfied if the displacements are considered
the unknowns and are approximated by a continuous interpolation model. Here, the
stresses are the unknowns.) From the stress-strain relation,
Exz =

1
1
Sxz =
,
2
2 y

Eyz =

1
1
Syz =
.
2
2 x

Compatibility (integrability) now requires that

2 w
xy

2 w ,
yx

furnishing

1 1 d 1 1 d
+ y
+

x
= 0,
y 2 y 2 dz x 2 x 2 dz

2003 by CRC CRC Press LLC

(14.5)

(14.6)

0749_Frame_C14 Page 183 Wednesday, February 19, 2003 5:41 PM

Torsion and Buckling

183

n
ds

dy

d z

n x = cos = dy/ds
n y = sin = dx/ds

dx

FIGURE 14.2 Illustration of geometric relation.

which in turn furnishes Poissons Equation for the potential function y :


2 2
d
+ 2 = 2
.
x 2
y
dz

(14.7)

For boundary conditions, assume that the lateral boundaries of the member are
unloaded. The stress-traction relation already implies that x = 0 and y = 0 on the
lateral boundary S. For traction z to vanish, we require that

z = n x Sxz + ny Syz = 0 on S.

(14.8)

Upon examining Figure 14.2, it can be seen that nx = dy/ds and ny = dx/ds, in
which s is the arc length along the boundary at z. Consequently,

z =

dy
dx
S
S
ds xz ds yz

dy dx
+
ds y ds x

d
ds

(14.9)

Now, on S, dds = 0, and therefore is a constant, which can, in general, be taken


as zero.
We next consider the total torque on the member. Figure 14.3 depicts the cross
section at z. The torque on the element at x and y is given by
dT = xSyz dxdy ySxz dxdy
d
d
= x
y
dxdy
dx
dy

2003 by CRC CRC Press LLC

(14.10)

0749_Frame_C14 Page 184 Wednesday, February 19, 2003 5:41 PM

184

Finite Element Analysis: Thermomechanics of Solids

dy
dx

sxz

syz

y
x

x
FIGURE 14.3 Evaluation of twisting moment.

Integration furnishes
d
d
T = x
dxdy
+y
dy
dx

d ( x ) d ( y )
dx dy
=
+
dx + dy dxdy
dx
dy

x
= dxdy + 2 dxdy
y

(14.11)

Application of the divergence theorem to the first term leads to [xnx + yny]ds,
which vanishes since y vanishes on S. Finally,

T = 2 dxdy.

(14.12)

We apply variational methods to the Poisson Equation, considering the stresspotential function y to be the unknown. Now,

[ + 2 ]dxdy = 0.
2003 by CRC CRC Press LLC

(14.13)

0749_Frame_C14 Page 185 Wednesday, February 19, 2003 5:41 PM

Torsion and Buckling

185

Integration by parts, use of the divergence theorem, and imposition of the


constraint y = 0 on S furnishes

( ) dxdy = 2 dxdy.

(14.14)
th

The integrals are evaluated over a set of small elements. In the e element,
approximate as TT ( x, y) e e , in which T is a vector with dimension (number of
rows) equal to the number of nodal values of . The gradient has a corresponding
interpolation model = TT ( x, y) e e , in which T is a matrix. The finite-element
counterpart of the Poisson Equation at the element level is written as
K (Te ) e = 2 fT( e )

K (Te ) = eT T TT dxdy e

(14.15)

fT( e ) = eT T ( x, y)dxdy
and the stiffness matrix should be nonsingular, since the constraint y = 0 on S has
1
already been used. It follows that, globally, g = 2 K (Tg ) fT( g ) . The torque satisfies

T = 2 dxdy
= 2 Tg fT( g )
1

= 4 fT( g ) K(Tg ) fT( g )


T

(14.16)

In the theory of torsion, it is common to introduce the torsional constant J, for


T
1
which T = 2J . It follows that J = 2 fT( g ) K(Tg ) fT( g ) .

14.2 BUCKLING OF BEAMS AND PLATES


14.2.1 EULER BUCKLING

OF

BEAM COLUMNS

14.2.1.1 Static Buckling


Under in-plane compressive loads, the resistance of a thin member (beam or plate)
can be reduced progressively, culminating in buckling. There are two equilibrium
states that the member potentially can sustain: compression only, or compression
with bending. The member will snap to the second state if it involves less potential
energy than the first state. The notions explaining buckling are addressed in detail in
subsequent chapters. For now, we will focus on beams and plates, using classical
equations in which, by retaining lowest-order corrections for geometric nonlinearity,
in-plane compressive forces appear.

2003 by CRC CRC Press LLC

0749_Frame_C14 Page 186 Wednesday, February 19, 2003 5:41 PM

186

Finite Element Analysis: Thermomechanics of Solids

Q0

E,I,A,L,

M0

FIGURE 14.4 Euler buckling of a beam column.

For the beam shown in Figure 14.4, the classical Euler buckling equation is
= 0,
EIw iv + Pw + Aw

(14.17)

and P is the axial compressive force. The interpolation model for w(x) is recalled
T
. Following the usual variational procedures (integration by parts)
as w(x) = (x)
furnishes

wAw dx M ,
T

w[EIw

iv

M = T ( x )A T ( x )dV

(14.18)

+ Pw ]dx = w EIw dx w Pw dx
[( w )( Pw EIw )]0L [( w )( EIw )]0L

At x = 0, both w and w vanish, while the shear force V and the bending
moment M are identified as V = EIw and M = EIw. The effective shear force
Q is defined as Q = Pw EIw.
For the specific case illustrated in Figure 14.3, for a one-element model, we can
use the interpolation formula

w( x ) = ( x

L2
x )
2 L
3

L3
(t ),
3 L2

w( L )
(t ) =
.
w ( L)

(14.19)

The mass matrix is shown, after some algebra, to be

M = ALK 0 ,

2003 by CRC CRC Press LLC

13
35
K0 =
11
210 L

L
.
1 L2
105
11
210

(14.20)

0749_Frame_C14 Page 187 Wednesday, February 19, 2003 5:41 PM

Torsion and Buckling

187

Similarly,

w Pw dx = T

wEIwdx =

P
K ,
L 1

65
K1 =
1
10
L

EI
K ,
L3 2

12
K2 =
6 L

2 2
15 L
1
10

(14.21)

6L

4 L2

The governing equation is written in finite-element form as


EI K P K + ALK
0 = f,
L3 2 L 1

Q0
f = .
M0

(14.22)

In a static problem,
= 0, the solution has the form
2

PL
cof K2
K1
EI

f,
=
2

PL
det K2
K1
EI

in which cof denotes the cofactor, and for values of


2
det(K2 PL
EI K1 ) = 0.

(14.23)

PL2
EI

which render

14.2.1.2 Dynamic Buckling


In a dynamic problem, it may be of interest to determine the effect of P on the
resonance frequency. Suppose that f(t) = f0 exp(it), in which f0 is a known vector.
The displacement function satisfies (t) = 0 exp(it), in which the amplitude vector
0 satisfies
EI K P K 2 ALK = f .
0 0
0
L3 2 L 1

(14.24)

Resonance occurs at a frequency 0, for which


EI
P
det 3 K 2 K1 02 ALK 0 = 0.
L
L

(14.25)

1
Clearly, 02 is an eigenvalue of the matrix AL
K 01 / 2 [ EL3I K 2 PL K1 ]K 01 / 2 . The
2
resonance frequency 0 is reduced by the presence of P and vanishes precisely at
the critical value of P.

2003 by CRC CRC Press LLC

0749_Frame_C14 Page 188 Wednesday, February 19, 2003 5:41 PM

188

Finite Element Analysis: Thermomechanics of Solids

V0

P
M0
L

FIGURE 14.5 Buckling of a clamped-clamped beam.

14.2.1.3 Sample Problem: Interpretation of Buckling Modes


Consider static buckling of a clamped-clamped beam, as shown in Figure 14.5.
This configuration can be replaced with two beams of length L, for which the
right beam experiences shear force V1 and bending moment M1, while the left beam
experiences shear force V0 V1 and bending moment M0 M1. The beam on the
right is governed by

65
K1 =
1
10
L

P
w Pw dx =
K ,
L 1
T

wEIwdx =

EI
K ,
L3 2

12
K2 =
6 L

2 L2
15
1
10

(14.26)

6L

4 L2

The governing equation is written in finite-element form as


12
EI
3
L 6 L

6 L P 65

1 L
4 L2 L 10

V1
f = ,
M1

L
= f
2 2
15 L
1
10

(14.27)

w( L )
=

w ( L)

Consider the symmetric case in which M0 = 0, with the implication that w(L) = 0.
The equation reduces to
12 EI 6 P w( L) = V ,
1
L3 5 L

(14.28)

from which we obtain the critical buckling load given by P1 = 10EI/L .


2

2003 by CRC CRC Press LLC

0749_Frame_C14 Page 189 Wednesday, February 19, 2003 5:41 PM

Torsion and Buckling

189

Next, consider the antisymmetric case in which V0 = 0 and w(L) = 0. The counterpart of Equation 14.28 is now
4 EI 2 PL ( w ( L)) = M ,
1
L 15

(14.29)

and P2 = 30EI/L .
If neither the constraint of symmetry nor axisymmetry is applicable, there are two
critical buckling loads, to be obtained in Exercise 2 as 27.8 EI2 and 8.9 EI2 . These values
L
L
are close enough to the symmetric and antisymmetric cases to suggest an interpretation
of the two buckling loads as corresponding to the two pure buckling modes.
Compare the obtained values with the exact solution, assuming static conditions. Consider the symmetric case. Let w(x) = wc(x) + wp(x), in which wc(x) is
the characteristic solution and wp(x) is the particular solution reflecting the perturbation. From the Euler buckling equation demonstrated in Equation 14.17, wc(x)
has a general solution of the form wc(x) = + x + cos x + sin x, in which
= PL2 / EI . Now, w = w = 0 at x = 0, w(L) = 0, and EIw(L) = V1, expressed
as the conditions
2

1 + 0 + 1 + 0 = w p (0)
0 + 1 + 0 + = w p (0)
(14.30)

0 + L sin L + cosL = w p ( L)
0 + 0 + 3 sin L 3 cosL = EIw
p ( L ) + V1
or otherwise stated
w p (0)

w p (0)

Bz =
,
w p ( L)

EIw ( L) + V

p
1

0
B=
0

sin L

3 sin L

,
cosL

3 cosL
0




z=



(14.31)

For the solution to blow up, it is necessary for the matrix B to be singular,
which it is if the corresponding homogeneous problem has a solution. Accordingly,
we seek conditions under which there exists a nonvanishing vector z, for which Bz = 0.
Direct elimination of and furnishes = and = . The remaining
coefficients must satisfy
sin L

sin L

2003 by CRC CRC Press LLC

cosL L 0
= .
cosL 0

(14.32)

0749_Frame_C14 Page 190 Wednesday, February 19, 2003 5:41 PM

190

Finite Element Analysis: Thermomechanics of Solids

A nonvanishing solution is possible only if the determinant vanishes, which


reduces to sin L = 0. This equation has many solutions for kL, including kL = 0.
2
2
2
The lowest nontrivial solution is kL = p, from which Pcrit = EI/L = 9.87 EI/L .
2
Clearly, the symmetric solution in the previous two-element model (Pcrit = 10 EI/L )
gives an accurate result.
For the antisymmetric case, the corresponding result is that tan L = L. The
lowest meaningful root of this equation is kL = 4.49 (see Brush and Almroth, 1975),
2
giving Pcrit = 20.19 EI/L . Clearly, the axisymmetric part of the two-element model is
not as accurate, unlike the symmetric part. This issue is addressed further in the
subsequent exercises.
Up to this point, it has been implicitly assumed that the beam column is initially
perfectly straight. This assumption can lead to overestimates of the critical buckling
load. Consider a known initial distribution w0(x). The governing equation is
d2
d2
d2
EI 2 ( w w0 ) + P 2 ( w w0 ) = 0,
2
dx
dx
dx

(14.33)

d2
d2
d2
d2
d2
d2
w .
I
w
P
w
I
w
P
E
+
=
E
+
0
dx 2
dx 2 0
dx 2
dx 2
dx 2
dx 2

(14.34)

or equivalently,

The crookedness is modeled as a perturbation. Similarly, if the cross-sectional


properties of the beam column exhibit a small amount of variation, for example,
EI(x) = EI0[1 + sin( x/L)], the imperfection can also be modeled as a perturbation.

14.2.2 EULER BUCKLING

OF

PLATES

The governing equation for a plate element subject to in-plane loads is


Eh 2
2 w
2 w
2 w
4 w + Px 2 + Py 2 + Pxy
=0
2
12(1 )
x
y
xy

(14.35)

(see Wang 1953), in which the loads are illustrated as shown in Figure 14.6. The usual
z

y
Pxy
Pyx

Px
FIGURE 14.6 Plate element with in-plane compressive loads.

2003 by CRC CRC Press LLC

Py

0749_Frame_C14 Page 191 Wednesday, February 19, 2003 5:41 PM

Torsion and Buckling

191

variational methods furnish, with some effort,

w w dA = w(n ) w dS w (nP )w dS + tr( WW)dA,


4

(14.36)
in which W = w (a matrix!). In addition,
T

2 w
2 w
2 w
w Px 2 + Py 2 + Pxy
dA = dw(nx
y
xy
x

ny )p dS

{ wx

wx
wy}P dA,
wy

[
[

P w + 1 P
x x
2 xy
p=
1 P w + P
y
2 xy x

w
y
w
y

] ,
]

Px
P=
12 Pxy

1
2

(14.37)

Pxy

Py

For simplicitys sake, assume that w( x, y) = Tb 2 b 2 b 2 from which we can obtain


the form
wx
w = = 1Tb 2 b 2 b 2 ,
wy

VEC(W) = T2 b 2 b 2 b 2 .

(14.38)

We also assume that the secondary variables (n ) w, (n )w, and also


[ P w + 1 P w ]
x x
2 xy y

are prescribed on S.
w

w
1
[ Pxy
+
P
]
y y
2
x
These conditions serve to obtain
2

[K b 21 K b 22 ] b 2 = f
K b 21 =

(14.39)

Eh 2
T T dAb 2
12(1 2 ) b 2 2 b 2 2 b 2

K b 22 = bT2 1b 2 P1Tb 2 dVb 2


and f reflects the quantities prescribed on S.
As illustrated in Figure 14.7, we now consider a three-dimensional loading space
in which Px, Py , and Pxy correspond to the axes, and seek to determine a surface in
the space of critical values at which buckling occurs. In this space, a straight line

2003 by CRC CRC Press LLC

0749_Frame_C14 Page 192 Wednesday, February 19, 2003 5:41 PM

192

Finite Element Analysis: Thermomechanics of Solids

Pxy

Py

Px
FIGURE 14.7 Loading space for plate buckling.

emanating from the origin represents a proportional loading path. Let the load
intensity, , denote the distance to a given point on this line. By analogy with
spherical coordinates, there exist two angles, and , such that
Px = cos cos ,

Py = sin cos ,

Pxy = sin

(14.40)

Now,
( , )
K b 22 = K
b 22

( , ) = T P ( , )T dV
K
b 22
b2
b2
1b 2
1b 2
cos cos
P ( , ) =
sin

(14.41)

sin

sin cos

For each pair (q, f), buckling occurs at a critical load intensity, crit(, ),
satisfying
] = 0.
det[K b 21 crit ( , )K
b 22

(14.42)

A surface of critical load intensities, crit(, ), can be drawn in the loading space
shown in Figure 14.7 by evaluating crit(, ) over all values of (q, f) and discarding
values that are negative.

2003 by CRC CRC Press LLC

0749_Frame_C14 Page 193 Wednesday, February 19, 2003 5:41 PM

Torsion and Buckling

193

14.3 EXERCISES
1. Consider the triangular member shown to be modeled as one finite element. Assume that

( x, y) = (1

y)1

x1
x2
x3

y1

y2

y3

1

2 .

3

Find KT , f T , and the torsional constant T.


y

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

Triangular shaft cross section.

2. Find the torsional constant for a unit square cross section using two
triangular elements.
3. Derive the matrices K0, K1, and K2 in Equations 14.20 and 14.21.
4. Compute the two critical values in Equation 14.23.
5. Use the four-element model shown in Figure 14.5, and determine how much
improvement, if any, occurs in the symmetric and antisymmetric cases.
6. Consider a two-element model and a four-element model of the simplesimple case shown in the following figure. Compare Pcrit in the symmetric
and antisymmetric cases with exact values.
V0

P
M0
L

2003 by CRC CRC Press LLC

0749_Frame_C14 Page 194 Wednesday, February 19, 2003 5:41 PM

194

Finite Element Analysis: Thermomechanics of Solids

7. Consider a cantilevered beam with a compressive load P, as shown in the


following figure. The equation is

EI

d 4w
d 2w
+ P 2 = 0.
4
dx
dx

The primary variables at x = L are w(L) and w(L), and

w EIw dx = T

EI 12

L3 6 L

6L
,
4 L2

w Pw dx = T

P 12

L 6 L

6L
.
4 L2

Find the critical buckling load(s).


V0
M0

2003 by CRC CRC Press LLC

0749_Frame_C15 Page 195 Wednesday, February 19, 2003 5:20 PM

15

Introduction to Contact
Problems

15.1 INTRODUCTION: THE GAP


In many practical problems, the information required to develop a finite-element
model, for example, the geometry of a member and the properties of its constituent
materials, can be determined with little uncertainty or ambiguity. However, often
the loads experienced by the member are not so clear. This is especially true if loads
are transmitted to the member along an interface with a second member. This class
of problems is called contact problems, and they are arguably the most common
boundary conditions encountered in practical problems. The finite-element community has devoted, and continues to devote, a great deal of effort to this complex
problem, leading to gap and interface elements for contact. Here, we introduce gap
elements.
First, consider the three-spring configuration in Figure 15.1. All springs are of
stiffness k. Springs A and C extend from the top plate, called the contactor, to the
bottom plate, called the target. The bottom of spring B is initially remote from the
target by a gap g. The exact stiffness of this configuration is

<g

2 k
kc =
3k

(15.1)

g.

From the viewpoint of the finite-element method, Figure 15.1 poses the following
difficulty. If a node is set at the lowest point on spring B and at the point directly
P

contactor

A
k

k
g

target

FIGURE 15.1 Simple contact problem.

195

2003 by CRC CRC Press LLC

0749_Frame_C15 Page 196 Wednesday, February 19, 2003 5:20 PM

196

Finite Element Analysis: Thermomechanics of Solids

contactor

A
k

k
kg

target

FIGURE 15.2 Spring representing contact element.

below it on the target, these nodes are not initially connected, but are later connected
in the physical problem. Furthermore, it is necessary to satisfy the nonpenetration
constraint whereby the middle spring does not move through the target. If the nodes
are considered unconnected in the finite-element model, there is nothing to enforce
the nonpenetration constraint. If, however, the nodes are considered connected, the
stiffness is artificially high.
This difficulty is overcome in an approximate sense by a bilinear contact element.
In particular, we introduce a new spring, kg, as shown in Figure 15.2.
The stiffness of the middle spring (B in series with the contact spring) is now
denoted as km, and
1

1 1
km = + .
k kg

(15.2)

It is desirable for the middle spring to be soft when the gap is open (g > ) and
to be stiff when the gap is closed (g ):
k /100 g >
kg =
100 k g .

(15.3)

Elementary algebra serves to demonstrate that


2 k + 0.01k
kc
2 k + 0.99k

g>
g .

(15.4)

Consequently, the model with the contact is too stiff by 0.5% when the gap
is open, and too soft by 0.33% when the gap is closed (contact). One conclusion
that can be drawn from this example is that the stiffness of the gap element should
be related to the stiffnesses of the contactor and target in the vicinity of the contact
point.

2003 by CRC CRC Press LLC

0749_Frame_C15 Page 197 Wednesday, February 19, 2003 5:20 PM

Introduction to Contact Problems

197

15.2 POINT-TO-POINT CONTACT


Generally, it is not known what points will come into contact, and there is no
guarantee that target nodes will come into contact with foundation nodes. The gap
elements can be used to account for the unknown contact area, as follows. Figure 15.3
shows a contactor and a target, on which are indicated candidate contact areas, dSc
and dSt, containing nodes c1, c2,..,cn, t1, t2,..,tn. The candidate contact areas
must contain all points for which there is a possibility of establishing contact.
th
The gap (i.e., the distance in the undeformed configuration) from the i node of
th
the contactor to the j node of the target is denoted by gij. (For the purpose of this
th
discussion, the gap is constant, i.e., not updated.) In point-to-point contact, the i
node on the contactor is connected to each node of the target by a spring with a
bilinear stiffness. (Clearly, this element may miss the edge of the contact zone when
it does not occur at a node.) It follows that each node of the target is connected by
a spring to each of the nodes on the contactor. The angle between the spring and
the normal at the contactor node is ij, while the angle between the spring and the
th
normal to the target is ji. Under load, the i contactor node experiences displacement
th
th
uij in the direction of the j target node, and the j target node experiences displaceth
th
ment uji. For example, the spring connecting the i contactor node with the j target
node has stiffness kij, given by
kijlower
kij =
kijupper

ij < gij

(15.5)

ij gij ,

in which ij = uij + uji is the relative displacement. The force in the spring connecting
th
th
the i contactor node and j target node is fij = kij(gij)ij. The total normal force
th
experienced by the i contactor node is fi = j fij cos(ij).

candidate contactor contact surface


dSc
c1

c2

c3
31
k(g31)

t1

t2
dSt

candidate target contact surface


FIGURE 15.3 Point-to-point contact.

2003 by CRC CRC Press LLC

t3

0749_Frame_C15 Page 198 Wednesday, February 19, 2003 5:20 PM

198

Finite Element Analysis: Thermomechanics of Solids

As an example of how the spring stiffness might depend upon the gap, consider
the function
kij ( gij ij )

= k0 + (1 2 ) 2 tan 1 ( gij ij ) 2 ( gij ij ) ,

(15.6)

where , , and are positive parameters selected as follows. When gij ij > 0,
kij attains the lower-shelf value, k0, and we assume that << 1. If gij ij < 0, kij
approaches the upper-shelf value, k0 (1 ). We choose to be a small value to
attain a narrow transition range from the lower- to the upper-shelf values. In the
range 0 < gij ij < , there is a rapid but continuous transition from the lower-shelf
(soft) value to the upper-shelf (stiff) value. If we now choose such that = 1,
kij becomes k0 /2, when the gap closes (gij = ij). The spring characteristic is illustrated
in Figure 15.4.
The total normal force on a contactor node is the sum of the individual contactelement forces, namely
Nc

ftj =

k (g
ij

ij

ij ) ij cos( ij ).

(15.7)

Clearly, significant forces are exerted only by the contact elements that are
closed.

k ij
k 0

k0

k 0 /2

k 0

FIGURE 15.4 Illustration of a gap-stiffness function.

2003 by CRC CRC Press LLC

ij gij

0749_Frame_C15 Page 199 Wednesday, February 19, 2003 5:20 PM

Introduction to Contact Problems

199

candidate target contact surface


dSt
t1

t2

t3

element connecting node


t3 with nodes c1 and c2

c2
dSc

c1

c3

candidate contactor contact surface


FIGURE 15.5 Element for point-to-surface contact.

15.3 POINT-TO-SURFACE CONTACT


We now briefly consider point-to-surface contact, illustrated in Figure 15.5 using a
triangular element. Here, target node t3 is connected via a triangular element to
contactor nodes c1 and c2. The stiffness matrix of the element is written as k([g1 1],
, in which g1 1 is the gap between nodes t1 and c1, and K
is the
[g2 2]) K
geometric part of the stiffness matrix of a triangular elastic element. The stiffness
matrix of the element can be made a function of both gaps. Total force normal to
the target node is the sum of the forces exerted by the contact elements on the
candidate contactor nodes.
In some finite-element codes, the foregoing scheme is used to approximate the
tangential force in the case of friction. Namely, an elastic-friction force is assumed
in which the tangential tractions are assumed proportional to the normal traction
through a friction coefficient. This model does not appear to consider sliding and
can be considered a bonded contact. Advanced models address sliding contact and
incorporate friction laws not based on the Coulomb model.

15.4 EXERCISES
1. Consider a finite-element model for a set of springs, illustrated in the
following figure. A load moves the plate on the left toward the fixed plate
on the right.
What is the load-deflection curve of the configuration?
For a finite-element model, an additional bilinear spring is supplied, as
shown. What is the load-deflection curve of the finite-element model?
Identify a kg value for which the load-deflection behavior of the finiteelement model is close to the actual configuration.
Why is the new spring needed in the finite-element model?

2003 by CRC CRC Press LLC

0749_Frame_C15 Page 200 Wednesday, February 19, 2003 5:20 PM

200

Finite Element Analysis: Thermomechanics of Solids

k
k

k
k
H
L

2. Suppose a contact element is added in the previous problem, in which the


stiffness (spring rate) satisfies
k ( gij ij )

= k0 + (1 2 ) 2 tan 1 ( gij ij )2 ( gij ij ) .


2

Suppose = 1, kL = k/100, and ku = 100k. Compute the stiffness k for


the configuration as a function of the deflection .

2003 by CRC CRC Press LLC

0749_Frame_C16 Page 201 Wednesday, February 19, 2003 5:23 PM

16

Introduction to
Nonlinear FEA

16.1 OVERVIEW
The previous section addressed finite-element methods for linear problems. Applications that linear methods serve to analyze include structures under mild loads,
disks and rotors spinning at modest angular velocities, and heated plates. However,
a large number of problems are nonlinear. Plasticity is a nonlinear materials theory
suited for metals in metal forming, vehicle crash, and ballistics applications. In
problems with high levels of heat input, mechanical properties, such as the elastic
modulus, and thermal properties, such as the coefficient of specific heat, can be
strongly temperature-dependent. Rubber seals and gaskets commonly experience
strains exceeding 50%. Soft biological tissues typically are modeled as rubberlike.
Many problems involve variable contact, for example, meshing gear teeth. Heat
conducted across electrical contacts can be strongly dependent on normal pressures.
Fortunately, much of the linear finite-element method can be adapted to nonlinear
problems, as explained in this chapter. The next chapter focuses on isothermal
problems. The extension to thermomechanical problems will be presented in a
subsequent chapter.

16.2 TYPES OF NONLINEARITY


There are three major types of nonlinearity in thermomechanical boundary-value
problems: material nonlinearity, geometric nonlinearity, and boundary-condition
nonlinearity. Nonlinearities can also be present if the formulation is referred to
deformed coordinates, possibly introducing stress fluxes and converted coordinates.
Material nonlinearity can occur through nonlinear dependence of the stress on
the strain or temperature, including temperature dependence of the tangent modulus
tensor. Metals undergoing plastic flow exhibit nonlinear material behavior.
Geometric nonlinearity occurs because of large deformation, especially in problems referred to undeformed coordinates. Rubber components typically exhibit large
deformation and require nonlinear kinematic descriptions. In this situation, a strain
measure needs to be chosen, as does the stress conjugate to it. Boundary-condition
nonlinearity occurs because of nonlinear supports on the boundary and contact. An
example of a nonlinear support is a rubber pad placed under a machine to absorb
vibrations.
For finite-element methods for nonlinear problems, the loads or time steps are
applied in increments. Then, incremental variational principles, together with interpolation models, furnish algebraic (static) or ordinary differential equations (dynamic)
201

2003 by CRC CRC Press LLC

0749_Frame_C16 Page 202 Wednesday, February 19, 2003 5:23 PM

202

Finite Element Analysis: Thermomechanics of Solids

in terms of vector-valued incremental displacements or incremental temperatures.


For mechanical systems, a typical equation is
K( ) + ( )
= f,

(16.1)

in which is the incremental displacement vector, f is the incremental force


vector, K() is the tangent stiffness matrix, and M() is the tangent mass matrix. It
will be seen that this type of equation is a realization of the optimal Newton iteration
method for nonlinear equations.

16.3 COMBINED INCREMENTAL AND ITERATIVE


METHODS: A SIMPLE EXAMPLE
Consider a one-dimensional rod of nonlinear material under small deformation in which
the elastic modulus is a function of strain: E = E0(1 + ), = E11. Under static loading,
the equilibrium equation is
AE 0

1 + = P .

L
L

(16.2)

Suppose the load is applied in increments, jP, and that the load at the n load
th
step is Pn. Suppose further that the solution n is known at the n load. We now
consider the steps necessary to determine the solution n+1 for Pn+1. We introduce the
th
increment n = g n+1 n. Subtracting Equation 16.2 at the n step from the equation
st
th
at the (n + 1) step, the incremental equilibrium equation for the n increment is now
th

AE 0
+ n
1 + 2 n+1
= nP .
n
L
L

(16.3)

This equation is quadratic in the increments. In fact, geometric nonlinearity


generally leads to a quadratic function of increments. The error of neglecting the
nonlinearity may be small if the load increment is sufficiently small. However, we
will retain the nonlinear term for the sake of illustrating the use of iterative procedures.
In particular, the previous equation can be written with obvious identifications in
the form
g( x ) = x + x 2 = 0.

(16.4)

Newton iteration (discussed in Chapter 3) furnishes the optimal iteration scheme


th
for the iterate:
[x + x ]
,
+ 2 x
2

x +1 = x

2003 by CRC CRC Press LLC

x 0 = / .

(16.5)

0749_Frame_C16 Page 203 Wednesday, February 19, 2003 5:23 PM

Introduction to Nonlinear FEA

203

SAMPLE PROBLEM 1
The efficiency of this scheme is addressed as follows. Consider = 1, = 12, and
= 1. The correct solution is x = 0.0828. Starting with the initial value x = 0, the
1
first two iterates are, approximately, 121 = 0.833, and 121 (1 144
) = 0.0828 . This illustrates the rapid quadratic convergence of Newton iteration.

16.4 FINITE STRETCHING OF A RUBBER ROD UNDER


GRAVITY: A SIMPLE EXAMPLE
Figure 16.1 shows a rubber rod under gravity. It is assumed to attain finite strain.
The rod is also assumed to experience uniaxial tension. The figure refers to the
undeformed configuration, with the element occupying the interval (Xe, Xe+1). Its
length is le, its cross-sectional area is Ae, and its mass density is . It is composed
of rubber and is stretched axially by the loads Pe and Pe+1. Prior to stretching, a given
material particle is located at X. After deformation, it is located at x(X), and the
displacement u(X ) is given by u(X ) = x(X ) X.

16.4.1 NONLINEAR STRAIN-DISPLACEMENT RELATIONS


The element is assumed short enough that a satisfactory approximation for the
displacement u(X ) is provided by the linear interpolation model
u( X ) = ue + ( X Xe )(ue+1 ue ) le
= NTae ,

NT =

1
{x x
le e

x x e},

(16.6)

in which a eT = {ue
ue+1}. Now, u(Xe) = ue and u(Xe+1) = ue+1 are viewed as the
unknowns to be determined using the finite-element method. The Lagrangian strain

Pe
xe,ue
g

xe+1,ue+1
Pe+1

FIGURE 16.1 Rubber rod under load.

2003 by CRC CRC Press LLC

0749_Frame_C16 Page 204 Wednesday, February 19, 2003 5:23 PM

204

Finite Element Analysis: Thermomechanics of Solids

E = Exx is given by
E=

u 1 u
+
x 2 x

ue+1 ue 1 ue+1 ue
+
2 le
le

1
{1
le

1} a e +

1
1 T 1
ae 2
2 le 1

1
ae .
1

(16.7)

Note that
dE = BT da e ,
BL =

1 1

le 1

B = B L + B NL a e
B NL =

1
le2

(16.8)

1
.
1

The vector BL and the matrix BNL are the strain-displacement matrices.
The following sections illustrate two different approaches to formulating a finiteelement model for the rubber rod. One is based on satisfying the incompressibility
constraint a priori. The second is based on satisfying the constraint a posteriori,
which is typical of finite-element code practice. In addition, we also include a slight
variant involving a near-incompressibility constraint in an a posteriori manner.

16.4.2 STRESS

AND

TANGENT MODULUS RELATIONS

The Neohookean strain-energy density function, w, accommodating incompressibility, is expressed in terms of the eigenvalues c1, c2, and c3 of C = 2E + I, as follows:
w=

D
(c + c + c3 3),
2 1 2

subject to c1c2 c3 1 = 0 ,

(16.9)

in which D is the (small-strain) elastic modulus. We can show, with some effort,
that under uniaxial tension, c2 = c3.
We first enforce the incompressibility constraint a priori by using the substitution
c2 = c3 =

1
.
c1

(16.10)

After elementary manipulations,


w=

2003 by CRC CRC Press LLC

D
[2 E + 1 + 2 /(2 E + 1) 1/ 2 3] .
2

(16.11)

0749_Frame_C16 Page 205 Wednesday, February 19, 2003 5:23 PM

Introduction to Nonlinear FEA

205

nd

The (2 Piola-Kirchhoff) stress, S, defined in Chapter 5, is now obtained as


S=
=

w
E
1
D[1 (2 E + 1)3/ 2 ].
3

(16.12)

The tangent modulus, D, is also required:


S
E

DT =

= D(2 E + 1)5/ 2 .

(16.13)

If the strain E is small compared to unity, DT reduces to D.


We next satisfy the incompressibility constraint a posteriori. An augmented
strain-energy function w* is introduced by
w* =

D
p
[c + c + c3 3] (c1c2 c3 1) ,
2 1 2
2

(16.14)

in which the Lagrange multiplier, p, is the (true) hydrostatic pressure. The augmented
energy is stationary with respect to p as well as c1, c2, and c3, from which it follows
that c1c2c3 1 = 0 (incompressibility).
The stresses satisfy
S1 =

w*
w* D p c1c2c3
=2
=
2 2 c1
E1
c1

S2 =

w*
w* D p c1c2c3
=2
=
=0
2 2 c2
E2
c2

S3 =

w *
w * D p c1c2c3
=2
=
= 0.
2 2 c3
E3
c3

(16.15)

It is readily shown that c2 = c3. Enforcement of the stationary condition for


p (c1c2c3 1 = 0), with S2 = 0, now furnishes that p = D / c1 . It follows that
S1 = D[1 c13/ 2 ]/ 2 , in agreement with Equation 16.12.

16.4.3 INCREMENTAL EQUILIBRIUM RELATION


The Principle of Virtual Work states the condition for static equilibrium as

Xe +1

Xe

BSe AdX P e g

2003 by CRC CRC Press LLC

Xe +1

Xe

N T AdX a e = 0,

Pee
Pe =
,
Pe
e+1

(16.16)

0749_Frame_C16 Page 206 Wednesday, February 19, 2003 5:23 PM

206

Finite Element Analysis: Thermomechanics of Solids

in which the third term represents the weight of the element, while Pe represents the
forces from the adjacent elements. In an incremental formulation, we can replace
the loads and displacements by their differential forms. In particular,

d Pe =

Xe +1

BdSAe dX +

Xe

Xe +1

Xe

d BSAe dX

= K e da e
= {K1e + K 2 e + K 3e + K 4 e} da e

(16.17)

in which
K1e = Ae

Xe +1

Xe

Xe

B L DT BTL dX
1

DT Ae
le

K 2 e = Ae

Xe +1

DT B L a eTB NL + B NL a e BTL dX
1

DT Ae [ue+1 ue ] 1
2

le
le
1

K 3e = Ae

(16.18a)

Xe +1

Xe

B NL a ea eTBTNL dX
1

2
DT Ae ue+1 ue 1

le le 1

K 4 e = Ae

(16.18b)

Xe +1

Xe

(16.18c)

B L SdX

SAe 1

le 1

(16.18d)

and
DT =

2003 by CRC CRC Press LLC

1
le

Xe +1

Xe

DT dX ,

S=

1
le

Xe +1

Xe

SdX.

(16.19)

0749_Frame_C16 Page 207 Wednesday, February 19, 2003 5:23 PM

Introduction to Nonlinear FEA

207

Combining Equations 16.18 and 16.19 produces the simple relation


1
Ke =
1

1
,
1

DT Ae
le

ue+1 ue ue+1 ue SAe


1
2
+
+

l
+ l .
le


e
e

(16.20)

Suppose (ue+1 ue)/le is small compared to unity, with the consequence that E
is also small compared to unity. It follows in this case that DT D, S S. As a
result, Ke reduces to the stiffness matrix for a rod element of linearly elastic material
experiencing small strain:
Ke =

EAe
le

1
.
1

(16.21)

Several special cases illuminate additional aspects of finite-element modeling.


Sample Problem 1: Single Element Built in at One End
Figure 16.2 depicts a single-element model of a rod that is built in at X = 0: Xe =
X0 = 0. At the opposite end, Xe+1 = X1 = 1. The rod is submitted to the load P. The
displacement at X = 0 is subject to the constraint u(0) = u0 = 0, so that Equation
16.20 becomes
1

1 0 dPr
=
,
1 du1 dP

(16.22)

in which dPr is an incremental reaction force that is not known a priori (of course,
from equilibrium, dPr = dP). For the only unknown reaction, this last equation
condenses to
kdu1 = dP .

x1

A,

u1

FIGURE 16.2 Rubber rod element under load: built in at top.

2003 by CRC CRC Press LLC

(16.23)

0749_Frame_C16 Page 208 Wednesday, February 19, 2003 5:23 PM

208

Finite Element Analysis: Thermomechanics of Solids

For later use, the current shape function degenerates to the expression N N =
X/X1.
The (Lagrangian) strain is given by
2

E(u1 ) =

u1 1 u1
+
,
X1 2 X1

(16.24)

and the strain-displacement matrices reduce to


B B=

u
1
+ 12 .
X1 X1

(16.25)

nd

The (2 Piola-Kirchhoff) stress S is now a function of u1:


3

2
2

1
1 u
u

S(u1 ) = D1 2 1 + 1 1 .
3 X1 2 X1

16.4.4 NUMERICAL SOLUTION

BY

(16.26)

NEWTON ITERATION

Equation 16.22 can be solved by Newton iteration, sometimes called load balancing, the reason for which is explained shortly. We introduce the residual function
(u1) = 0 as follows. The Principle of Virtual Work implies that if gravity is considered in Figure 16.2,

(u1 ) =

X1

BSA dX P g

X1

N T A dX a e

2
2
1 u1
1 u1 1 u1
X
= + 2 AX1 D1 2 + 1 P gA 1 u1
2
le le
3 X1 2 X1

= 0.

(16.27)
th

when u1 is the solution. Consider an iteration process in which the j iterate u1j has
been determined. Newton iteration determines the next iterate, u1j +1, using

(u1j ) j u1 = (u1j )
u1j +1 = u1j + j u1.

(16.28)

Convergence of this scheme to the correct value u1 is usually rapid, provided


that the initial iterate u10 is sufficiently close to u1. A satisfactory initial iterate is

2003 by CRC CRC Press LLC

0749_Frame_C16 Page 209 Wednesday, February 19, 2003 5:23 PM

Introduction to Nonlinear FEA

209

often obtained using an incremental procedure. Suppose that the solution has been
obtained at previous load steps. A starting iterate for the current load step is obtained
by extrapolating from the previous solutions. As an example, suppose the solution
is known at the load P = kP, in which P is a load increment. The load is now
incremented to produce P = (k + 1)P, and the solution u1,k+1 at this load is determined
using Newton iteration. It is frequently satisfactory to start the iteration using
u10,k +1 = u1. This numerical procedure is illustrated more extensively at the end of
this chapter.
Sample Problem 2: Assembled Stiffness Matrix
for a Two-Element Model
Assemblage procedures are used to combine the element-equilibrium relations to
obtain the global equilibrium relation for an assemblage of elements. For elements
e and e + 1, the element-equilibrium relations are expanded as
k11e due + k12e due+1 = dPee

(i)

e
e
k21
due + k22
due+1 = dPee+1

(ii)
(16.29)

k11e+1due+1 + k12e+1due+2 = dPee++11

(iii)

e +1
e +1
k21
due+1 + k22
due+2 = dPee++21

(iv).

The superscript indicates the element index, and the subscript indicates the node
index. If no external force is applied at xe+1, the interelement force balance is
expressed as
dPee+1 + dPee++11 = 0 .

(16.30)

Adding Equation 16.29(ii) and Equation 16.29(iii) furnishes

e
e
k21
due + k22
+ k11e+1 due+1 + k12e+1due+2 = 0 .

(16.31)

Equations 16.29, 16.30, and 16.31 are expressed in matrix form as


k11e

e
k21

k12e
e
k22
+ k11e+1
e +1
k21

0 due dPee


k12e+1 due+1 = 0 .

e+1
e +1
k22
due+2 dPe+2

(16.32)

Equation 16.32 illustrates that the (incremental) global stiffness matrix is formed
by overlaying Ke and Ke+1, with the entries added in the intersection.

2003 by CRC CRC Press LLC

0749_Frame_C16 Page 210 Wednesday, February 19, 2003 5:23 PM

210

Finite Element Analysis: Thermomechanics of Solids

Pee

xe,ue

P,Ae

e
Pe+1
e+1
Pe+1

xe+1,ue+1

P,Ae+1

xe+2,ue+2

e+1
e+2

x
FIGURE 16.3 Two rubber rods under load.

Sample Problem 3: Two Identical Elements under Gravity under


Equal End-Loads
If Ke = Ke+1 and dPee = dPee++11 = dP , overlaying the element matrices leads to the
global (two-element) relation
1

1
2
1

0 du1 dP

1 du2 = 0 .

1 du3 dP

(16.33)

Note that Equation 16.33 has no solution since the global stiffness matrix has
no inverse: the second row is the negative of the sum of the first and last rows. This
suggests that, due to numerical errors in the load increments, the condition for static
equilibrium is not satisfied numerically, and therefore that the body is predicted to
accelerate indefinitely (undergo rigid-body motion). However, we also know that the
configuration under incremental loads is symmetric, implying a constraint du2 = 0.
This constraint permits condensation, that is, reducing Equation 16.33 to a system
with two unknowns by eliminating rows and columns associated with the middle
incremental displacement:
1

1 du1 dP
=
.
1 du3 dP

(16.34)

The condensed matrix is now proportional to the identity matrix, and the system
has a solution. In general, stiffness matrices can easily be singular or nearly singular

2003 by CRC CRC Press LLC

0749_Frame_C16 Page 211 Wednesday, February 19, 2003 5:23 PM

Introduction to Nonlinear FEA

211

(with a large condition number) unless constraints are used to suppress rigid-body
modes.

16.5 ILLUSTRATION OF NEWTON ITERATION


Unfortunately, since elastomers or metals experiencing plasticity are typically quite
compliant, it is common to encounter convergence problems, for which there are
four major approaches:
1.
2.
3.
4.

Increasing stiffness, such as by introducing additional constraints if available


Reducing load-step sizes and reforming the stiffness matrix after each iteration
Switching to displacement control rather than load control
Using an arc-length method (described in Chapter 3)

For an equation of the form

( x ) = 0,

(16.35)

Newton iteration seeks a solution through an iterative process given by


1

d ( x j )
x j +1 = x j
(x j ) .
dx

(16.36)

Clearly, using two sequential iterates,


1

d ( x j )
d ( x j 1 )
x j +1 x j = x j x j 1
(xj )

( x j 1 )
dx
dx
1
d ( x *) d 2 ( x *)

(
*)
= x j x j 1
+
x

x
j

dx 2

dx

( x *) +

d ( x *) d 2 ( x *)

d ( x *)
( x j 1 x*)
( x j x*)
+
dx 2
dx
dx

( x *) +

d ( x *)

( x j 1 x*) + L
dx

d ( x *) d ( x *)
( x x j 1 ) + O([ x j x j1 ]2 )
= 1
dx j
dx

for some x* between x j and x j1.

2003 by CRC CRC Press LLC

(16.37)

0749_Frame_C16 Page 212 Wednesday, February 19, 2003 5:23 PM

212

Finite Element Analysis: Thermomechanics of Solids

encounter difficulties with convergence


Y
1

X
FIGURE 16.4 Illustration of the inverse tangent function.

16.5.1 EXAMPLE
The performance of Newton iteration procedure is illustrated using a simple example
showing convergence problems. Consider

( x) =

2
tan 1 ( x ) y = 0.

(16.38)

The goal is to find the solution of x as y is incremented in the range (0,1). Clearly,
x approaches infinity as y approaches unity, so that the goal is to generate the x(y)
relationship accurately as close as possible to y = 1. The curve appears as shown in
Figure 16.4. As y is incremented by small amounts just below unity, the differences
in x due to the increment are large, so that the solution value is far from the initial
iterate.
th
Suppose that y is incremented such that the n value of y is
yn = ny.

(16.39)

To obtain the solution of the (n + 1) step, the Newton iteration procedure


st
th
generates the ( + 1) iterate from the iterate, as follows:
st

x n++11

x n+1

( )

d x n+1

x n+1 ,
dx

( )

(16.40)

in which x n+1
is the iterate for the solution xn+1.
0
Of course, a starting iterate, x n+1
, is needed. An attractive candidate is xn+1.
However, this may not be good enough when convergence difficulties appear.
Newton iteration was implemented for this example in a simple problem using
increments of 0.001. A stopping criterion of 10 iterations was used. Convergence
was found to fail near y = 0.999.
th

2003 by CRC CRC Press LLC

0749_Frame_C16 Page 213 Wednesday, February 19, 2003 5:23 PM

Introduction to Nonlinear FEA

213

16.6 EXERCISES
1. Doing the algebra, make judicious choices of starting iterates, and develop
the first few iterates for the roots of the equations
x 2 + 3x + 2 = 0
x2 1 = 0
( x 1)2 = 0
2. Write a simple program implementing Newton iteration, and apply it to
the examples in Exercise 1.
3. Using Example 5.1, write a simple program using Newton iteration to
compute X(Y ) using increments y = 0.001. Set the stopping criterion as
the first of: (a) 10 iterates or (b) an absolute difference of less than 0.001
in the successive values of x. Note the value of x at which convergence
fails, and compare the values of x with accurate values available from the
formulae and tables in Abramawitz and Stegun (1997).

2003 by CRC CRC Press LLC

0749_Frame_C17 Page 215 Wednesday, February 19, 2003 5:24 PM

17

Incremental Principle
of Virtual Work

17.1 INCREMENTAL KINEMATICS


Recall that the displacement vector u(X) is assumed to admit a satisfactory approxT
(t). Also recall that the
imation at the element level in the form u(X) = (X)
deformation-gradient tensor is given by F = xx . Suppose that the body under study
is subjected to a load vector, P, which is applied incrementally via load increments,
th
jP. The load at the n load step is denoted as Pn. The solution, Pn, is known, and
the solution of the increments of the displacements is sought. Let nu = un+1 un,
T
n. By suitably arranging the derivatives of nu with respect
so that nu = (X)
to X, a matrix, M(X), can easily be determined for which VEC(nF) = M(X)n.
T
We next consider the Lagrangian strain, E(X) = 12 (F F I). Using Kronecker
Product algebra from Chapter 2, we readily find that, to first order in increments,
n e = VEC( n E)
1
= VEC [FT n F + n FT F]
2

1
= [I FT + FT IU]VEC( n F)
2
= G n g,

1
GT = [I FT + FT IU]M (X) n g.
2

(17.1)

This form shows the advantages of Kronecker Product notation. Namely, it


enables moving the incremental displacement vector to the end of the expression
outside of domain integrals, which we will encounter subsequently.
Alternatively, for the current configuration, a suitable strain measure is the
T 1
Eulerian strain, = 12 (I F F ), which refers to deformed coordinates. Note that
1
1
1
1
T
T
T T
since n(FF ) = 0, nF = F nFF . Similarly, nF = F nF F . Simple
manipulation furnishes that
1
VEC( n ) = [F T F 1 F T U + F 1 F T F 1 ]M n g.
2

(17.2)

There also are geometric changes for which an incremental representation is


1
useful. For example, since the Jacobian J = det(F) satisfies dJ = Jtr(F dF), we obtain

215

2003 by CRC CRC Press LLC

0749_Frame_C17 Page 216 Wednesday, February 19, 2003 5:24 PM

216

Finite Element Analysis: Thermomechanics of Solids

the approximate formula


J = Jtr(F 1 n F)
= JVEC T (F 1 )VEC( n F)
T

= VEC T (F T ) JM n .

(17.3)

Also of interest are


d[n dS]
= [tr(D)I LT ]n dS
dt
dn
= [(nT Dn)I LT ]n
dt

(17.4)

d
dS = [tr(D) nT Dn]dS
dt
Using Equation 17.4, we obtain the incremental forms
n [ndS] = dS[nVEC T (F T ) nT F T U]M n
n n = [n(nT F T ) nT nT F T U]M n

(17.5)

n dS = dS[nVEC T (F T ) n(nT F T ) nT ]M n .

17.2 INCREMENTAL STRESSES


For the purposes of deriving an incremental variational principle, we shall see that
st
the incremental 1 Piola-Kirchhoff stress, n S , is the starting point. However, to
o
formulate mechanical properties, the objective increment of the Cauchy stress, n T,
is the starting point. Furthermore, in the resulting variational statement, which we
called the Incremental Principle of Virtual Work, we find that the quantity that
nd
appears is the increment of the 2 Piola-Kirchhoff stress, nS.
From Chapter 5, we learned that S = SF T , from which, to first order,
n S = n SF T + S n F T .

(17.6)

For the Cauchy stress, the increment must take into account the rotation of the
underlying coordinate system and thereby be objective. We recall the objective
o
Truesdell stress flux, T / t, introduced in Chapter 5:
o

T/t = T/t + Ttr(D) LT TLT .

2003 by CRC CRC Press LLC

(17.7)

0749_Frame_C17 Page 217 Wednesday, February 19, 2003 5:24 PM

Incremental Principle of Virtual Work

217

Among the possible stress fluxes, it is unique in that it is proportional to the


nd
rate of the 2 Piola-Kirchhoff stress, namely
o

S/t = J F 1 T/tF T .

(17.8)

An objective Truesdell stress increment is readily obtained as


o
VEC( n T ) = 1 F T FVEC( n S).

(17.9)

Furthermore, once VEC( n T ) has been determined, the (nonobjective) increment of the Cauchy stress can be computed using
o

n T = n T + Ttr( n FF 1 ) n FF 1T T F T n F T ,

(17.10)

from which
o

VEC( nT ) = VEC( n T ) + [TVEC T (F T ) (T F T ) I I (T F T )]M n . (17.11)

17.3 INCREMENTAL EQUILIBRIUM EQUATION


We now express the incremental equation of nonlinear solid mechanics (assuming
that there is no net rigid-body motion). In the deformed (Eulerian) configuration,
equilibrium at tn requires

T n dS = u dV.
T

(17.12)

Referred to the undeformed (Lagrangian) configuration, this equation becomes

S n dS = u dV ,
T

(17.13)

st

in which, as indicated before, S is the 1 Piola-Kirchhoff stress, S denotes the surface


(boundary) in the deformed configuration, and n0 is the surface normal vector in the
undeformed configuration. Suppose the solution for S is known as Sn at time tn and
is sought at tn+1. We introduce the increment n S to denote Sn+1 Sn . A similar
definition is introduced for the increment of the displacements. Now, equilibrium
applied to Sn+1 and Sn implies

S n
T

2003 by CRC CRC Press LLC

dS0 =

u dV .
0

(17.14)

0749_Frame_C17 Page 218 Wednesday, February 19, 2003 5:24 PM

218

Finite Element Analysis: Thermomechanics of Solids

Application of the divergence theorem furnishes the differential equation


T .
T n S T = 0 n u

(17.15)

17.4 INCREMENTAL PRINCIPLE OF VIRTUAL WORK


To derive a variational principle for the current formulation, the quantity to be varied is
the incremental displacement vector since it is now the unknown. Following Chapter 5,
Equation 17.15 is multiplied by (nu) .
Integration is performed over the domain.
The Gauss divergence theorem is invoked once.
Terms appearing on the boundary are identified as primary and secondary
variables.
Boundary conditions and constraints are applied.
T

The reasoning process is similar to that in the derivation of the Principle of


Virtual Work in finite deformation in which u is the unknown, and furnishes

tr( E S)dV + FS F dV + u u dV = u dS ,
T

n 0

(17.16)

in which 0 is the traction experienced by dS0. The fourth term describes the virtual
external work of the incremental tractions. The first term describes the virtual internal
work of the incremental stresses. The third term describes the virtual internal work
of the incremental inertial forces. The second term has no counterpart in the previously formulated Principle of Virtual Work in Chapter 5, and arises because of
geometric nonlinearity. We simply call it the geometric stiffness integral. Due to the
importance of this relation, Equation 17.16 is derived in detail in the equations that
follow. It is convenient to perform the derivation using tensor-indicial notation:

u X ( S )dV = X [ u ( S )]dV X [ u ] S dV
n i

n ij

n i

n ij

n i

n ij

(17.17)

= nui 0 nui dV0 .

The first term on the right is converted using the divergence theorem to

X [ u ( S )]dV = u (n S )dS
n i

n ij

n i

n ij

= n ui 0 j dS0
which is recognized as the fourth term in Equation 17.16.

2003 by CRC CRC Press LLC

(17.18)

0749_Frame_C17 Page 219 Wednesday, February 19, 2003 5:24 PM

Incremental Principle of Virtual Work

219

To first-order in increments, the second term on the right is written, using tensor
notation, as

X [ u ] S dV = tr( F S )dV
n i

n ij

= tr( n F[ n SFT + S n FT ])dV0

= tr(FT n F n S)dV0 + tr( n FS n FT )dV0

(17.19)

The second term is recognized as the second term in Equation 17.16.


The first term now becomes
1

tr(F F S)dV = tr 2 [F F + F F] S dV
T

= tr( n T nS)dV0

(17.20)

which is recognized as the first term in Equation 17.16.

17.5 INCREMENTAL FINITE-ELEMENT EQUATION


For present purposes, let us suppose constitutive relations in the form
n S = D( X, n ) n E,

(17.21)

in which D(X, ) is the fourth-order tangent modulus tensor. It is rewritten as


n s = ( X , n ) n e
s = VEC( S),

e = VEC( E),

(17.22)

= TEN22( D).

Also for present purposes, we assume that 0 is prescribed on the boundary S0, a
common but frequently unrealistic assumption that is addressed in a subsequent section.
In VEC notation, and using the interpolation models, Equation 17.16 becomes

n T [(K T + K G ) n + M n n f ] = 0

(17.23)

KT =

M GG M dV

KG =

M S IM dV

M=

dV

n f =

dS

2003 by CRC CRC Press LLC

0749_Frame_C17 Page 220 Wednesday, February 19, 2003 5:24 PM

220

Finite Element Analysis: Thermomechanics of Solids

KT is now called the tangent modulus matrix, KG is the geometric stiffness matrix,
M is the (incremental) mass matrix, and nf is the incremental force vector.

17.6 INCREMENTAL CONTRIBUTIONS


FROM NONLINEAR BOUNDARY CONDITIONS
Again, let Ii denote the principal invariants of C, and let i = VEC(I), c2 = VEC(C ),
n iT = Ii /c, and Ai = ni /c. Recall from Chapter 2 that
2

n1 = i

n2 = I1i c

n3 = I2 i I1c + c2

I9 = I I

A1 = 0

A2 = ii T I9

A3 = I C + C I (ic T + ciT ) + I1 (iiT I 9 ).

(17.24)

Equation 17.23 is complete if increments of tractions are prescribed on the


undeformed surface S0. We now consider the more complex situation in which is
referred to the deformed surface S, on which they are prescribed functions of u.
From Chandrasekharaiah and Debnath (1994), conversion is obtained using
dS = 0 dS0

n T q dS = n T0 q 0 dS0

dS = dS0

= J n C n0 =
T
0

(17.25)
n n n3
T
0

T
0

and from Nicholson and Lin (1997b)


d = m T dc m T c,

m T = n T0 n T0 A 3 / 2 .

(17.26)

Suppose that is expressed on S as follows:


= A TM u.

(17.27)

Here, d is prescribed, while AM is a known function of u. Also, S0 is the


undeformed counterpart of S. These relations are capable of modeling boundary
conditions, such as support by a nonlinear elastic foundation.
From the fact that dS = 0 dS0 = 0 dS, we conclude that = 0. It follows that

0 =
=

2003 by CRC CRC Press LLC

( ATM u) 2 mT c

(17.28)

0749_Frame_C17 Page 221 Wednesday, February 19, 2003 5:24 PM

Incremental Principle of Virtual Work

221

From the Incremental Principle of Virtual Work, the rhs term is written as
1

u dS = u ( A u)
T

T
M

mT c dS0 .

(17.29)

Recalling the interpolation models for the increments, we obtain an incremental


force vector plus two boundary contribution to the stiffness terms. In particular,

u dS = f [ K
T

f =

dS ,
1

K BF = NATM N T dS0 ,

BF

+ KBN ]

K BN =

(17.30)

mT G dS0

The first boundary contribution is from the nonlinear elastic foundation coupling
the traction and displacement increments on the boundary. The second arises from
geometric nonlinearity when the traction increment is prescribed on the current
configuration.

17.7 EFFECT OF VARIABLE CONTACT


In many, if not most, real-world problems, loads are transmitted to the member of
interest via contact with other members, for example, gear teeth. The extent of the
contact zone is an unknown to be determined as part of the solution process. Solution
of contact problems, introduced in Chapter 15, is a difficult problem that has absorbed
the attention of many investigators. Some algorithms are suited primarily for linear
kinematics. Here, a development is given for one particular formulation, which is
mostly of interest for explicitly addressing the effect of large deformation.
Figure 17.1 shows a contactor moving into contact with a foundation that is
assumed to be rigid. We seek to follow the development of the contact area and the
tractions arising throughout it. From Chapter 15, we recall that corresponding to a
point x on the contactor surface there is a target point y(x) on the foundation to
which the normal n(x) at x points. As the contactor starts to deform, n(x) rotates
and points toward a new value, y(x). As the point x approaches contact, the point
y(x) approaches the foundation point, which comes into contact with the contactor
point at x.
We define a gap function, g, using y(x) = x + gn. Let m be the surface normalvector to the target at y(x). Let Sc be the candidate contact surface on the contactor,
whose undeformed counterpart is S0c. There also is a candidate contact surface Sf
on the foundation.
We limit our attention to bonded contact, in which particles coming into contact
with each other remain in contact. Algorithms for sliding contact with and without
friction are available. For simplicitys sake, we also assume that shear tractions, in

2003 by CRC CRC Press LLC

0749_Frame_C17 Page 222 Wednesday, February 19, 2003 5:24 PM

222

Finite Element Analysis: Thermomechanics of Solids

contactor

n
g
m
y(x)

foundation

FIGURE 17.1 Contact.

the osculating plane of point of interest, are negligible. Suppose that the interface
can be represented by an elastic foundation satisfying the incremental relation
n = k ( g)un .

(17.31)

Here, n = n and un = n u are the normal components of the traction and


displacement vectors. Since the only traction to consider is the normal traction (to
the contactor surface), the transverse components of u are not needed (do not result
from work). Also, k(g) is a nonlinear stiffness function given in terms of the gap by,
for example,
T

k ( g) =

kH

arctan( g ) + k ,
r
L
k
2

k H /k L >> 1.

(17.32)

As in Chapter 15, when g is positive, the gap is open and k approaches kL, which
should be chosen as a small number, theoretically zero. When g becomes negative,
the gap is closed and k approaches kH, which should be chosen as a large number,
theoretically infinity to prevent penetration of the rigid body).
Under the assumption that only the normal traction on the contactor surface is
important, it follows that = tnn, from which
= n n + n n.

2003 by CRC CRC Press LLC

(17.33)

0749_Frame_C17 Page 223 Wednesday, February 19, 2003 5:24 PM

Incremental Principle of Virtual Work

223

The contact model contributes the matrix Kc to the stiffness matrix as follows
(see Nicholson and Lin, 1997b):

u dS = u
T

n dS0

= T K c

(17.34)

K c = 2 Nn n mT T dSc 0 + kc ( g)NnnT N T dSc 0 + N n hT dSc 0 .


To update the gap, use the following relations proved in Nicholson and Lin (1997-b).
T
The differential vector, dy, is tangent to the foundation surface, hence, m dy = 0. It
follows that
0 = mT du + gmT d n + m Tn dg
g =

(17.35)

mT u + gmT n
.
mTn

Using Equation 17.5, we may derive, with some effort, that


g = T ,

T =

mT N + ghT
,
mTn

hT = [n[(nT F T ) nT ] nT F T ]MT .
(17.36)

17.8 INTERPRETATION AS NEWTON ITERATION


The (nonincremental) Principle of Virtual Work can be restated in the undeformed
configuration as

tr( ES)dV + u u dV = u dS .
T

(17.37)

We assume for convenience that is prescribed on So. The interpolation model


satisfies the form

e = BTL + BTNL ( )
BTL =
BTNL =
Fu =

2003 by CRC CRC Press LLC

1
(I I + I IU)M(X)
2

1
I FuT + Fu IU M(X)
2
u
.
X

(17.38)

0749_Frame_C17 Page 224 Wednesday, February 19, 2003 5:24 PM

224

Finite Element Analysis: Thermomechanics of Solids

Clearly, Fu and BNL are linear in .

Upon cancellation of the variation d , an algebraic equation is obtained as

dVo ,
( , f ) = [(B L + B NL ( )]s dVo + N T u

f = N T o dSo .

(17.39)

At the load step, Newton iteration is expressed as

()

J=
n+1, fn+1

(nv++11) = (n+)1 J 1 (n+)1,fn+1 ,

(17.40)

or as a linear system

J (nv++11) (nv+)1 = (nv+)1, fn+1 ,

(17.41)

(nv++11) = (nv+)1 + (vn++11) (nv+)1 .


If the load increments are small enough, the starting iterate can be estimated as the
th
solution from the n load step. Also, a stopping (convergence) criterion is needed
to determine when the effort to generate additional iterates is not rewarded by
increased accuracy.
Careful examination of the relations from this and the incremental formulations
uncovers that
J = KT + KG ,

(17.42)

so that the incremental stiffness matrix is the same as the Jacobian matrix in Newton
iteration. This, of course, is a satisfying result. The Jacobian matrix can be calculated
by conventional finite-element procedures at the element level followed by conventional assembly procedures. If the incremental equation is only solved once at each
load increment, the solution can be viewed as the first iterate in a Newton iteration
scheme. The one-time incremental solution can potentially be improved by additional
iterations, as shown in Equation 17.41, but at the cost of computing the residual
at each load step.

17.9 BUCKLING
Finite-element equations based on classical buckling equations for beams and plates
were addressed in Chapter 14. In the classical equations, geometrically nonlinear
terms appear through a linear correction term, thereby furnishing linear equations.
Here, in the absence of inertia and nonlinearity in the boundary conditions, we briefly
present a general viewpoint based on the incremental equilibrium equation
(K T + K G ) n+1 = fn+1 .

2003 by CRC CRC Press LLC

(17.43)

0749_Frame_C17 Page 225 Wednesday, February 19, 2003 5:24 PM

Incremental Principle of Virtual Work

225

This solution predicts a large incremental displacement if the stiffness matrix


KT + KG is ill-conditioned or outright singular. Of course, in elastic media, KT is
positive-definite. However, in the presence of in-plane compression, KG may have
a negative eigenvalue whose magnitude is comparable to the smallest positive eigenvalue of KT . To see this recall that
KT =

M GG M dV
T

KG =

M S IM dV .
T

(17.44)

We suppose that the element in question is thin in a local z (out-of-plane


direction). This suggests the assumption of plane stress. Now, in plate-and-shell
theory, it is necessary to add a transverse shear stress on the element boundaries to
allow the element to support transverse loads. We assume that the transverse shear
stresses only appear in the incremental force term and the tangent stiffness term,
and that the geometric stiffness term strictly satisfies the plane-stress assumption. It
follows that if the z-direction is out of the plane, in the geometric stiffness term,
S11I

S I S12 I

0 I

S12 I
S22 I
0I

0I

0I .

0 I

(17.45)

In classical buckling, it is assumed that loads applied proportionately induce


proportionate in-plane stresses. Thus, for a given load path, only one parameter,
the length of the straight line the stress point traverses in the space of in-plane
stresses, arises in the eigenvalue problem for the critical buckling load. In nonlinear
problems, there is no assurance that the stress point follows a straight line. Instead,
if l denotes the distance along the line followed by the load point in proportional
loading, the stresses become numerical functions of l.
As a simple alternative to the general case, we consider buckling of a single
element and suppose that the stresses appearing in Equation 17.46 are applied in a
compressive sense along the faces of the element in a proportional manner whereby
( S11 )I

S I ( S12 )I

0I

( S12 )I
( S22 )I
0I

0I

0 I ,

0 I

(17.46)

in which the circumflex implies a reference value along the stress path at which l = 1,
and the negative signs on the stresses are present since buckling is associated with
compressive stresses. At the element level, the equation now becomes
) = f .
(K T K
G
n +1
n +1

2003 by CRC CRC Press LLC

(17.47)

0749_Frame_C17 Page 226 Wednesday, February 19, 2003 5:24 PM

226

Finite Element Analysis: Thermomechanics of Solids

At a given load increment, the critical buckling load for the current path, as a
function of two angles determining the path in the stress space illustrated in
) singular.
Chapter 14, is obtained by computing the l value rendering (K T K
G

17.10 EXERCISES
1. Assuming linear interpolation models for u,v in a plane triangular
membrane element with vertices (0,0),(1,0),(0,1), obtain the matrix M,
G, BL, and BNL.
2. Repeat Exercise 1 with linear interpolation models for u, v, and w in a
tetrahedral element with vertices (0,0,0),(1,0,0),(0,1,0),(0,0,1).

2003 by CRC CRC Press LLC

0749_Frame_C18 Page 227 Wednesday, February 19, 2003 5:25 PM

18

Tangent-Modulus Tensors
for Thermomechanical
Response of Elastomers

18.1 INTRODUCTION
Within an element, the finite-element method makes use of interpolation models for
the displacement vector u(X, t) and temperature T(X, t) (and pressure p = trace()/3
in incompressible or near-incompressible materials):
u( X , t ) = NT ( X ) (t ),

T( X , t ) T0 = T ( X ) (t ),

p = T ( X ) (t ),

(18.1)

in which T0 is the temperature in the reference configuration, assumed constant.


Here, N, , and are shape functions and , , and are vectors of nodal values.
Application of the strain-displacement relations and their thermal analogs furnishes
f1 = VEC(F I) = M1 g = U T M2 g, f2 = VEC(F T I) = M2 g, e = b T g,
b = M2G, GT =

(18.2)

1 T
(F I + I F T U), T = bTT q
2

in which U is a 9 9 universal permutation tensor such that VEC(A ) = UVEC(A),


and e = VEC(E) is the Lagrangian strain vector. Also, is the gradient operator
referred to the deformed configuration. The matrix and the vector T are typically
expressed in terms of isoparametric coordinates.
T

18.2 COMPRESSIBLE ELASTOMERS


The Helmholtz potential was introduced in Chapter 7 and shown to underlie the
relations of classical coupled thermoelasticity. The thermohyperelastic properties of
compressible elastomers are also derived from the Helmholtz free-energy density
(per unit mass), which is a function of T and E. Under isothermal conditions it is
conventional to introduce the strain energy density w(E) = 0 (T, E) (T constant),
in which 0 is the density in the undeformed configuration. Typically, the elastomer
is assumed to be isotropic, in which case can be expressed as a function of T, I1,
I2, and I3. Alternatively, it may be expressed as a function of T and the stretch ratios
1, 2, and 3.
227

2003 by CRC CRC Press LLC

0749_Frame_C18 Page 228 Wednesday, February 19, 2003 5:25 PM

228

Finite Element Analysis: Thermomechanics of Solids

With known as a function of T, I1, I2, and I3, the entropy density per unit
mass and the specific heat ce at constant strain are obtained as
ce = T

T E

.
T

(18.3)

The 2 Piola-Kirchhoff stress, s = VEC(S), is obtained from


nd

s T = 0

=2
T

n ,
0 i

i =

.
Ii

(18.4)

Also of importance is the (isothermal) tangent-modulus matrix


DT =

s
=4
e T

n n + 4 A ,
0 ij

T
j

0 i

ij =

2
.
Ii I j

(18.5)

An expression for DT has been derived by Nicholson and Lin (1997c) for
compressible, incompressible, and near-incompressible elastomers described by
strain-energy functions (Helmholtz free-energy functions) and based on the use of
stretch ratios (singular values of F) rather than invariants.

18.3 INCOMPRESSIBLE AND NEAR-INCOMPRESSIBLE


ELASTOMERS
When the temperature T is held constant, elastomers often satisfy the constraint of
incompressibility or near-incompressibility. The constraint is accommodated by
augmenting with terms involving a new parameter similar to a Lagrange multiplier.
Typically, this new parameter is related to the pressure p. The thermohyperelastic
properties of incompressible and near-incompressible elastomers can be derived
from the augmented Helmholtz free energy, which is a function of E, T, and p. The
constraint introduces additional terms into the governing finite-element equations
and requires an interpolation model for p.
If the elastomer is incompressible at a constant temperature, the augmented
Helmholtz function, , can be written as

= d ( J1, J2 , T) ( J , T)/0 ,

J1 = I1/ I31/ 3 ,

J2 = I2 / I32 / 3 ,

(18.6)

where is a material function satisfying the constraint (J, T ) = 0 and J = I31 2 = det(F).
It is easily shown that d depends on the deviatoric Lagrangian strain Ed, due to the
introduction of the deviatoric invariants J2 and J3. The Lagrange multiplier is, in
fact, the (true) pressure p:
p = trace(T )/3 =

2003 by CRC CRC Press LLC

.
J T

(18.7)

0749_Frame_C18 Page 229 Wednesday, February 19, 2003 5:25 PM

Tangent-Modulus Tensors for Thermomechanical Response of Elastomers

229

For an elastomer that is near-incompressible at a constant temperature, can be


written as

0 = 0d ( J1, J2 , T) p ( J , T) p2 /2 ,

(18.8)

in which 0 is a constant. The near-incompressibility constraint is expressed by


/p = 0, which implies
p = ( J , T).

(18.9)

The bulk modulus is given by

.
=0
J T
J T

(18.10)

Chen et al. (1997) presented sufficient conditions under which near-incompressible models reduce to the incompressible case as . Nicholson and Lin (1996)
formulated the relations

( J , T) = f 3 (T) J 1, d = 1( J1 , J 2 ) + 2 (T), 2 (T) = ce T(1 ln(T/T0 )), (18.11)


with the consequence that
p = 0 ( f 3 (T) J 1),

= f 3 (T) 0 .

(18.12)

Equation 18.12 provides a linear pressure-volume relation in which thermomechanical effects are confined to thermal expansion expressed using a constant-volume
coefficient . If the constraint is assumed to be satisfied a priori, the Helmholtz free
energy is recovered as

( I1, I2 , I3 ) = d ( J1, J2 , T) + 0 ( f 3 (T) 1)2 /2 0 .

(18.13)

Alternatively, the latter term results from retaining the lowest nonvanishing term
3
in a Taylor-series representation of about f (T)J 1.
Given Equation 18.13, the entropy now includes a term involving p:

d
+ f 4 (T)/0 ,
T

= p/ f 3 (T).

(18.14)

The stress and the tangent-modulus matrices are correspondingly modified:


s T = 0

d
e

f 3 (T)n3T / J
T,

(18.15)
T

s
d
DTP =
= 0
f 3 (T) 2A3 / J n3n3T / J 3
e e
e T,

2003 by CRC CRC Press LLC

0749_Frame_C18 Page 230 Wednesday, February 19, 2003 5:25 PM

230

Finite Element Analysis: Thermomechanics of Solids

18.3.1 SPECIFIC EXPRESSIONS

FOR THE

HELMHOLTZ POTENTIAL

There are two broad approaches to the formulation of Helmholtz potential:


To express as a function of I1, I2, and I3, and T (and p)
To express as a function of the principal stretches 1, 2, and 3, and T (and p).
The latter approach is thought to possess the convenient feature of allowing direct
use of test data, for example, from uniaxial tension. We will now examine several
cases.
18.3.1.1 Invariant-Based Incompressible Models:
Isothermal Problems
The strain-energy function depends only on I1, I2, and incompressibility is expressed
by the constraint I3 = 1, assumed to be satisfied a priori. In this category, the most
widely used models include the Neo-Hookean material:

= C1 ( I1 3),

I3 = 1

(18.16)

and the (two-term) Mooney-Rivlin material:

= C1 ( I1 3) + C2 ( I2 3),

I3 = 1,

(18.17)

in which C1 and C2 are material constants. Most finite-element codes with hyperelastic elements support the Mooney-Rivlin model. In principle, Mooney-Rivlin
coefficients C1 and C2 can be determined independently by fitting suitable loaddeflection curves, for example, uniaxial tension. Values for several different rubber
compounds are listed in Nicholson and Nelson (1990).
18.3.1.2 Invariant-Based Models for Compressible Elastomers
under Isothermal Conditions
Two widely studied strain-energy functions are due to Blatz and Ko (1962). Let G0
be the shear modulus and v0 the Poissons ratio, referred to the undeformed configuration. The two models are:
0
1 2 0
1 + 0
1
1 2 0
01 = G0 I1 +
I3 I3

0
0
2

(18.18)

1 I
0 2 = G0 2 + 2 I3 5
2 I3

Let w denote the Helmholtz free energy evaluated at a constant temperature, in which
case it is the strain energy. We note a general expression for w which is implemented

2003 by CRC CRC Press LLC

0749_Frame_C18 Page 231 Wednesday, February 19, 2003 5:25 PM

Tangent-Modulus Tensors for Thermomechanical Response of Elastomers

231

in several commercial finite-element codes (e.g., ANSYS, 2000):


w( J1, J2 , J ) =

C ( J 3) ( J 3) + ( J 1) /D ,
i

ij

Jr = J /(1 + Eth ),

(18.19)
in which Eth is called the thermal expansion strain, while Cij and Dk are material
constants. Several codes also provide software for estimating the model coefficients
from user-supplied data.
Several authors have attempted to uncouple the response into isochoric (incompressible) and volumetric parts even in the compressible range, giving rise to functions of the form = 1(J1, J2) + 2(J). A number of proposed forms for 2 are
discussed in Holzappel (1996).
18.3.1.3 Thermomechanical Behavior under Nonisothermal
Conditions
Now we come to the accommodation of coupled thermomechanical effects. Simple
extensions of, for example, the Mooney-Rivlin material have been proposed by
Dillon (1962), Nicholson and Nelson (1990), and Nicholson (1995) for compressible
elastomers, and in Nicholson and Lin (1996) for incompressible and near-incompressible elastomers. From the latter,

0 = C1 ( J1 2) + C2 ( J2 3) + 0ce T(1 ln(T/To ) ( f 3 (T) J 1) 2 /2 ,


(18.20)
in which = p/f (T ). As previously mentioned, a model similar to Nicholson and
Lin (1996) has been proposed by Holzappel and Simo (1996) for compressible
elastomers described using stretch ratios.
3

18.4 STRETCH RATIO-BASED MODELS:


ISOTHERMAL CONDITIONS
For compressible elastomers, Valanis and Landel (1967) proposed a strain-energy
function based on the decomposition

(1 , 2 , 3 , T) = (1 , T) + ( 2 , T) + (3 , T),

T fixed.

(18.21)

Ogden (1986) has proposed the form


N

0 ( , T) =

2003 by CRC CRC Press LLC

1),

T fixed.

(18.22)

0749_Frame_C18 Page 232 Wednesday, February 19, 2003 5:25 PM

232

Finite Element Analysis: Thermomechanics of Solids

In principle, in incompressible isotropic elastomers, stretch ratio-based models


have the advantage of permitting direct use of archival data from single-stress tests,
for example, uniaxial tension.
We now illustrate the application of Kronecker Product algebra to thermohyperelastic materials under isothermal conditions and accommodate thermal effects.
From Nicholson and Lin (1997c), we invoke the expression for the differential of a
tensor-valued isotropic function of a tensor. Namely, let A denote a nonsingular
n n tensor with distinct eigenvalues, and let F(A) be a tensor-valued isotropic
function of A, admitting representation as a convergent polynomial:

F(A) =

A .
j

(18.23)

Here, j are constants. A compact expression for the differential dF(A) is presented using Kronecker Product notation.
The reader is referred to Nicholson and Lin (1997c) for the derivation of the
following expression. With f = VEC(F) and a = VEC(A),
df ( a) =

1 T
F Fda + Wd
2

F ( A ) =

j A
j

j 1

df
dF
= ITEN 22
dA
da

(18.24)

1
W = ( F AF/2)T  ( F AF/2) + ( AT F F T A)
2
= VEC(d
), in which d
is an antisymmetric tensor representing the
Also, d
rate of rotation of the principal directions. The critical step is to determine a matrix J
= Jda. It is shown in Dahlquist and Bjork that J = [AT  A]1W, in
such that Wd
T
I
which [A  A] is the Morse-Penrose inverse [(Dahlquist and Bjork(1974)]. Thus,
df /da = F T F/2 [AT  A] W.

(18.25)

We now apply the tensor derivative to elastomers modeled using stretch ratios,
especially in the model presented by Ogden (1986). In particular, a strain-energy
function, w, was proposed, which, for compressible elastomers and isothermal
response, is equivalent to the form

w = tr

[C

I ] ,

(18.26)

in which ci are the eigenvalues of C, and i, i are material properties. The tangentmodulus tensor appearing in Chapter 17 for the incremental form of the Principle

2003 by CRC CRC Press LLC

0749_Frame_C18 Page 233 Wednesday, February 19, 2003 5:25 PM

Tangent-Modulus Tensors for Thermomechanical Response of Elastomers

233

of Virtual Work is obtained as

=4

( 1)C

i 2

i i

C i 2 /2 + 4

Wi =

3 i
2

[A
i i

 A] Wi

(18.27)

C i C i +

i 1 i 2
2
[C C C C i ].
2

(18.28)

18.5 EXTENSION TO THERMOHYPERELASTIC


MATERIALS
Equations 18.27 and 18.28 can be extended to thermohyperelastic behavior as follows,
based on Nicholson and Lin (1996). The body initially experiences temperature T0
uniformly. It is assumed that temperature effects occur primarily as thermal expansion, that volume changes are small, and that volume changes depend linearly on
temperature. Thus, materials of present interest can be described as mechanically
nonlinear but thermally linear.
Due to the role of thermal expansion, it is desirable to uncouple dilatational and
deviatoric effects as much as possible. To this end, we introduce the deviatoric Cauchy = C / I 1/ 3 in which I3 is the third principal invariant of C. Now, modGreen strain C
3
ifying w and expanding it in J 1, ( J = I31/ 2 ) and retaining lowest-order terms gives

w = tr

[C

1
I] + ( J 1)2 ,
2

(18.29)

in which is the bulk modulus. The expression for in Equation 18.27 is affected
by these modifications.
To accommodate thermal effects, it is necessary to recognize that w is simply
the Helmholtz free-energy density o under isothermal conditions, in which o is
the mass density in the undeformed configuration. It is assumed that = 0 in the
undeformed configuration. As for invariant-based models, we can obtain a function
with three terms: a purely mechanical term M, a purely thermal term T , and a
mixed term TM. Now, with entropy, , satisfies the relations
s T = o

e T

.
T e

(18.30)

Following conventional practice, the specific heat at constant strain, ce = T /Te,


is assumed to be constant, from which we obtain

T = ce T [ln(T/T0 ) + 1].

(18.31)

On the assumption that thermal effects in shear (i.e., deviatoric effects) can be

2003 by CRC CRC Press LLC

0749_Frame_C18 Page 234 Wednesday, February 19, 2003 5:25 PM

234

Finite Element Analysis: Thermomechanics of Solids

neglected relative to thermal effects in dilatation, the purely mechanical effect is


equated with the deviatoric term in Equation 18.29:

M = tr

[C

] .

(18.32)

Of greatest interest is TM. The development of Nicholson and Lin (1996) furnishes

TM =

[ 3 (T) J 1]2
2

(T) = (1 + (T/T0 )/3) 1 .

(18.33)

The tangent-modulus tensor = s/ e now has two parts: XM + XTM, in which


XM is recognized as , derived in Equation 18.29. Without providing the details,
Kronecker Product algebra furnishes the following:

TM =

n nT
3 3
2 n3n3T + ( 3 J 1)A3 3 2 3 .
J J
J

(18.34)

The foregoing discussion of stretch-based thermohyperelastic models has been


limited to compressible elastomers. However, many elastomers used in applications,
such as seals, are incompressible or near-incompressible. For such applications, as
we have seen, an additional field variable is introduced, namely, the hydrostatic
pressure (referred to deformed coordinates). It serves as a Lagrange multiplier
enforcing the incompressibility and near-incompressibility constraints. Following
the approach for invariant-based models, Equations 18.33 and 18.34 can be extended
to incorporate the constraints of incompressibility and near-incompressibility.
The tangent-modulus tensor presented here only addresses the differential of
stress with respect to strain. However, if coupled heat transfer (conduction and
radiation) is considered, a general expression for the tangent-modulus tensor is
required, expressing increments of stress and entropy in terms of increments of strain
and temperature. A development accommodating heat transfer for invariant-based
elastomers is given in Nicholson and Lin (1997a).

18.6 THERMOMECHANICS OF DAMPED ELASTOMERS


Thermoviscohyperelasticity is a topic central to important applications, such as
rubber mounts in hot engines. The current section introduces a thermoviscohyperelastic constitutive model thought to be suitable for near-incompressible elastomers exhibiting modest levels of viscous damping following a Voigt model. Two
potential functions are used to provide a systematic treatment of reversible and
irreversible effects. One is the familiar Helmholtz free energy in terms of the strain
and the temperature; it describes reversible, thermohyperelastic effects. The second
potential function, based on the model of Ziegler and Wehrli (1987), models viscous

2003 by CRC CRC Press LLC

0749_Frame_C18 Page 235 Wednesday, February 19, 2003 5:25 PM

Tangent-Modulus Tensors for Thermomechanical Response of Elastomers

235

dissipation and arises directly from the entropy-production inequality. It provides a


consistent thermodynamic framework for describing damping in terms of a viscosity
tensor that depends on strain and temperature.
The formulation leads to a simple energy-balance equation, which is used to
derive a rate-variational principle. Together with the Principle of Virtual Work,
variational equations governing coupled thermal and mechanical effects are presented. Finite-element equations are derived from the thermal-equilibrium equation
and from the Principle of Virtual Work. Several quantities, such as internal energy
density, , have reversible and irreversible portions, indicated by the subscripts r
and i: = r + i. The thermodynamic formulation in the succeeding paragraphs is
referred to undeformed coordinates.
There are several types of viscoelastic behaviors in elastomers, especially if they
contain fillers such as carbon black. For example, under load, elastomers experience
stress softening and compression set, which are long-term viscoelastic phenomena.
Of interest here is the type of damping that is usually assumed in vibration isolation
in which the stresses have an elastic and a viscous portion reminiscent of the classical
Voigt model, and the viscous portion is proportional to strain rates. The time constants are small. This type of damping is viewed as arising in small motions superimposed on the large strains, which already reflect long-term viscoelastic effects.

18.6.1 BALANCE

OF

ENERGY

The conventional equation for the balance of energy is expressed as

0 = s T e 0T q 0 + 0 h
= srT e + siT e 0T q 0 + 0 h

(18.35)

where s = VEC(S) and e = VEC(E). Here, is the internal energy per unit mass, q0
is the heat-flux vector, 0 is the divergence operator referred to undeformed coordinates, and h is the heat input per unit mass, for simplicitys sake, assumed independent of temperature. The state variables are thus e and T. The Helmholtz free
energy, r per unit mass, and the entropy, per unit mass, are introduced using

r = T.

(18.36)

T0 q 0 0 h = sTr e + sTi e 0 T 0T 0 r .

(18.37)

Now,

18.6.2 ENTROPY PRODUCTION INEQUALITY


The entropy-production inequality is stated as

0 T T0 q 0 + 0 h + q0T T/T
0 r sTr e siT e + 0 T + 0T + q T0 T/T

2003 by CRC CRC Press LLC

(18.38)

0749_Frame_C18 Page 236 Wednesday, February 19, 2003 5:25 PM

236

Finite Element Analysis: Thermomechanics of Solids

The Helmholtz potential is assumed to represent reversible thermohyperelastic


effects. We decompose into reversible and irreversible portions: = r + i. Now,
r , r , and i are assumed to be differentiable functions of E and T. Furthermore,
we suppose that i = i1 + i2 and

0 T i 2 = T0 q0 + 0 h i .

(18.39)

This allows us to say that the viscous dissipation is absorbed as heat. We also
suppose that reversible effects are absorbed as a portion of the heat input as follows:

0 T r = T0 q0 + 0 h r .

(18.40)

In addition, from conventional arguments,

r /e = sTr

r /T = r ,

(18.41)

it follows that
sTi e q T0 0 T/T 0i1T .

(18.42)

Inequality as shown in Equation 18.42 can be satisfied if 0i T 0 and


sTi e 0

(a)

q T0 T/T 0

(18.43)

(b).

Inequality as shown in Equation 18.43b is conventionally assumed to express


the fact that heat flows irreversibly from cold to hot zones. Inequality as shown in
Equation 18.43a requires that viscous effects be dissipative.

18.6.3 DISSIPATION POTENTIAL


Following Ziegler and Wehrli (1987), the specific dissipation potential (q0 , e , e, T)
= 0i T is introduced, for which
siT = 0 i /e

(a).

T0 T/T = t 0 /q 0 .

(b).

(18.44)

The function is selected such that i and t are positive scalars, in which case
the inequalities in Equations 18.44a and 18.44b require that
(/e ) e 0

(/q 0 )q 0 0.

(18.45)

This can be interpreted as indicating the convexity of a dissipation surface in (e , q0 )


space. Clearly, to state the constitutive relations, it is sufficient to specify r and .

2003 by CRC CRC Press LLC

0749_Frame_C18 Page 237 Wednesday, February 19, 2003 5:25 PM

Tangent-Modulus Tensors for Thermomechanical Response of Elastomers

237

A simple illustration is now provided showing how the dissipation potential in


Equation 18.45 provides a framework for describing dissipative effects. On the
expectation that properties governing heat transfer are not affected by strain, we
introduce the decomposition
= i + t

0 t = [q0T q0 ]2 [ t /2].

(18.46)

Now, t represents thermal effects, and we assume for simplicitys sake that t is
a material constant. Inequality in Equation 18.46 implies that
0 T/T = q 0 / t .

(18.47)

This is essentially the conventional Fourier law of heat conduction, with t


recognized as the thermal conductivity. As an elementary example of viscous dissipation, suppose that
i = (T, J1, J2 )e Te / 2

i = 1,

(18.48)

in which (T, J1, J2) is the viscosity. Hence,


si = (T, J1, J2 )e ,

(18.49)

and Equation 18.44a requires that be positive.

18.6.4 THERMAL-FIELD EQUATION

FOR

DAMPED ELASTOMERS

The energy-balance equations of thermohyperelasticity (i.e., the reversible response)


are now reappearing in terms of a balance law among reversible portions of the
stress, entropy, and internal energy. Equation 18.40 is repeated as

0 r = sTr e 0r T.

(18.50)

The ensuing Maxwell relation is


srT /T = 0 r /e.

(18.51)

Conventional operations furnish the reversible part of the equation of thermal


equilibrium (balance of energy):
[T0 q0 + 0 h]r = T(sTr /T)e + 0ce T ,

ce = Tr /T.

(18.52)

For the irreversible part, we recall the relations

T0 q 0 0 h i = sTi e + 0ci T

2003 by CRC CRC Press LLC

(18.53)

0749_Frame_C18 Page 238 Wednesday, February 19, 2003 5:25 PM

238

Finite Element Analysis: Thermomechanics of Solids

and

0 Ti 2 = 0 T i 2 e + 0 T i 2 T
e
T

(18.54)

and
siT = 0 T

i 2
,
e

ci = T

i 2
.
T

(18.55)

Upon adding the relations, we obtain the thermal-field equation


T0 q 0 + 0 h = T sTr /Te siTe + 0 (ce + ci )T .

(18.56)

It is easily seen that Equation 18.55 directly reduces to a well-known expression


in classical linear thermoelasticity. In addition, under adiabatic conditions in which
T0 q 0 + 0 h = 0, most of the viscous work, s iT e , is absorbed as a temperature
increase controlled by 0(ce + ci), while a smaller portion is absorbed into the
elastic strain-energy field.

18.7 CONSTITUTIVE MODEL: POTENTIAL


FUNCTIONS
18.7.1 HELMHOLTZ FREE-ENERGY DENSITY
In the moderately damped, thermohyperelastic material, the elastic (reversible) stress
is assumed to satisfy a thermohyperelastic constitutive relation suitable for nearincompressible elastomers. In particular,

r = rm ( I1 , I2 , I3 ) + rt (T) + rtm (T, I3 ) + ro .

(18.57)

Here, rm represents the purely mechanical response and can be identified as the
conventional, isothermal, strain-energy density function associated, for example,
with the Mooney-Rivlin model. Again, I1, I2, I3 are the principal invariants of the
(right) Cauchy-Green strain tensor. The formulation can easily be adapted to stretch
ratio-based models, such as the Ogden (1986) model. The function rt(T) represents
the purely thermal portion of the Helmholtz free-energy density. Finally, rtm(T, I3)
represents thermomechanical effects, again based on the assumption that the primary
coupling is through volumetric expansion. The quantity ro represents the Helmholtz
free energy in the reference state, and, for simplicitys sake, is assumed to vanish.
The forms of rt and rtm are introduced in the current presentation:

rt (T) = ce T[1 ln(T/T0 )]


rtm (T, I3 ) =

[ f 3 (T) J 1]2
2 0

J = I31/ 2 = det( F) f (T) = 1 + ( T T0 )


3

2003 by CRC CRC Press LLC

(18.58)
(18.59)
1

0749_Frame_C18 Page 239 Wednesday, February 19, 2003 5:25 PM

Tangent-Modulus Tensors for Thermomechanical Response of Elastomers

239

and is the volumetric coefficient of thermal expansion. For the sake of illustration,
for rm(I1, I2, I3), we display the classical two-term Mooney-Rivlin model:

rm ( I1, I2 , I3 ) = C1 ( J1 1) + C2 ( J2 1),

I3 = 1,

(18.60)

in which J1 = I1/ I31/ 3 and J2 = I2 / I32 / 3 are the deviatoric invariants of C. The reversible stress is obtained as
sr = 2 j n j
n1 = i,

i = VEC(I)

j = r /I j

n1 = I1i c,

c = VEC(C)

n3 = I3VEC(C1 )

(18.61)

18.7.2 SPECIFIC DISSIPATION POTENTIAL


Fouriers law of conduction is obtained from:

t = [q0T q0 ]1/ 2 [kt /2].

(18.62)

The viscous stress si depends on the shear part of the strain rate as well as the
temperature. However, since the elastomers of interest are nearly incompressible, to
good approximation si can be taken as a function of the (total) Lagrangian strain rate.
The current framework admits several possible expressions for i, of which an
example was already given in Section 6.3. Here, taking a more general viewpoint,
we seek expressions of the form i = 12 e T Dv (e, T)e , in which Dv is called the
viscosity tensor; it is symmetric and positive-definite. (Of course, the correct expression is determined by experiments.) The simplest example was furnished in Section
18.6.3. As a second example, to ensure isotropy, suppose that i is a function of J1
and J2 : i ( J1 , J2 ), and note that
I
J1 = m1T e, m1T = 2 iT 1 I1 n3T I31 3
3 3

I
J2 = mT2 e, mT2 = 2 nT2 23 I2 n3T I32 3 .

(18.63)

For an expression reminiscent of the two-term Mooney-Rivlin strain-energy function, let us consider the specific form

i = 1 (T) C1v J12 /2 + C2 v J2 2 ,

(18.64)

in which C1v and C2v are positive material coefficients. We obtain the viscosity tensor,

Dv = (T) C1v m1m1T + C2 v m2 mT2 .

(18.65)

Unfortunately, this tensor is only positive-semidefinite. As a second example,


suppose that the dissipation potential is expressed in terms of the deformation rate

2003 by CRC CRC Press LLC

0749_Frame_C18 Page 240 Wednesday, February 19, 2003 5:25 PM

240

Finite Element Analysis: Thermomechanics of Solids

tensor D, in particular, i = (T)tr(D )/2, which has the advantage in that the
deformation rate tensor D is in the observed (current) configuration. With d =
VEC(D),
2

d = F T F T e

Dv = (T)(2 + I )1 (2 + I )1,

(18.66)

which is positive-definite.

18.8 VARIATIONAL PRINCIPLES


18.8.1 MECHANICAL EQUILIBRIUM
In this section, we present one of several possible formulations for the finite-element
equations of interest, neglecting inertia. Application of variational methods to the
mechanical field furnishes the Principle of Virtual Work in the form

tr(ES )dV = u dS tr(ES )dV ,


T

(18.67)

in which 0 denotes the traction vector on the undeformed surface S0. As illustrated
in the examples in the previous section, we expect that the dissipation potential has
the form i = 12 e T Dv(e, T) e , from which
si = Dv e ,

(18.68)

in which Dv is again the viscosity tensor, and it will be taken as symmetric and
positive-definite. (It is positive-definite since siTe 0 for all e .) Equation 18.67 is
thus rewritten as

e D edV = u t dS e s dV .
T

(18.69)

18.8.2 THERMAL EQUILIBRIUM


The equation for thermal equilibrium is rewritten as

0 ( r + i 2 ) = T0 q0 + 0 h / T,

(18.70)

and note that it is in rate form, in contrast to the equation of mechanical equilibrium
(see Equation 18.69). For the sake of a unified rate formulation, we first introduce
the integrated form of this relation. The current value of r + i2, assuming that the
initial values of the entropies vanish, is now given by

0 (r + i 2 ) =

2003 by CRC CRC Press LLC

[[ q + h]/ T]dt.
T
0 0

(18.71)

0749_Frame_C18 Page 241 Wednesday, February 19, 2003 5:25 PM

Tangent-Modulus Tensors for Thermomechanical Response of Elastomers

241

The corresponding variational principle is stated as

T ( +
0

i2

[[ q + h]/ T]dt dV .

)dV0 = T

T
0 0

(18.72)

With some effort, a rate (incremental) variational principle can be obtained in


the form

T[Ts /Te s e + (c + c )T ]/ TdV = T[[ q


T
r

T
i

T
0

] ]

+ 0 h / T dV0 .

(18.73)

(The motivation behind writing Equations 18.71 and 18.72 is simply to establish
how a rate principle is obtained in the thermal field as the counterpart of the rate
(incremental) principle for the mechanical field.) Upon approximating T in the
denominator by T0 and letting k denote the thermal conductivity, we can obtain the
thermal-equilibrium equation in the form

T[s /Te s e (c + c )T ]/ T dV
T
r

T
i

+ [k (0 T)T 0 T/T0 ]dV0 = T0 h/T0 dV0 nT ( Tq/T0 )dS0

(18.74)

Using interpolation models for displacement and temperature, Equations 18.69


and 18.74 reduce directly into finite-element equations for the mechanical and
thermal fields.

18.9 EXERCISES
1. Derive explicit forms of the stress and tangent-modulus tensors using the
Helmoltz potential in Equation 18.20.
2. Derive the quantities m1 and m2 in Equation 18.63.
3. Verify Equation 18.58 by recovering ce upon differentiating twice with
respect to T.
4. Substitute the required interpolation models in Equations 18.69 and 18.74
to obtain an element-level finite-element equation for the mechanical and
thermal fields.

2003 by CRC CRC Press LLC

0749_Frame_C19 Page 243 Wednesday, February 19, 2003 5:34 PM

19

Inelastic and
Thermoinelastic Materials

19.1 PLASTICITY
Plasticity and thermoplasticity are topics central to the analysis of important applications, such as metal forming, ballistics, and welding. The main goal of this section
is to present a model of plasticity and thermoplasticity, along with variational and
finite-element statements, accommodating the challenging problems of finite strain
and kinematic hardening.

19.1.1 KINEMATICS
Elastic and plastic deformation satisfies the additive decomposition
D = D r + Di ,

(19.1)

from which we can formally introduce strains:


=

Ddt

r =

D dt
r

i =

D dt.
i

(19.2)

The Lagrangian strain E satisfies the decomposition


E = F T DF

Er = F T D r Fdt

Ei = F T D i Fdt.

(19.3)

Typically, plastic strain is viewed as permanent strain. As illustrated in Figure 19.1,


in a uniaxial tensile specimen, the stress, S11, can be increased to the point A, and
then unloaded along the path AB. The slope of the unloading portion is E, the same
as that of the initial elastic portion. When the stress becomes equal to zero, there
still is a residual strain, Ei, which is identified as the inelastic strain. However, if
instead the stress was increased to point C, it would encounter reversed loading at
point D, which reflects the fact that the elastic region need not include the zerostress value.

19.1.2 PLASTICITY
We will present a constitutive equation for plasticity to illustrate how the tangent
modulus is stated. The ideas leading to the equation will be presented subsequently
in the section on thermoplasticity. With e = ITEN22(De) and De denoting the
243

2003 by CRC CRC Press LLC

0749_Frame_C19 Page 244 Wednesday, February 19, 2003 5:34 PM

244

Finite Element Analysis: Thermomechanics of Solids

C
S11

A
E
E
D

E
B
Ei

E11

FIGURE 19.1 Illustration of inelastic strain.

tangent-modulus tensor relating the elastic-strain rate to the stress rate (assuming
a linear relation), the constitutive equation of interest is
e i = Ci s,

e e = Ces,
T

i
Ci = i
H,
s s

Ce = e ,

k = Ke i

T
H = i + i K i .
k s
ei

(19.4)

In Equation 19.4, i is the yield function. i = 0 determines a closed convex


surface in stress space called the yield surface. (We will see later that i also serves
as a [complementary] dissipation potential.) The stress point remains on the yield
surface during plastic flow, and is moving toward its exterior. The plastic strain rate,
expressed as a vector, is typically assumed to be normal to the yield surface at the
stress point. If the stress point is interior to, or moving tangentially on, the yield
surface, only elastic deformation occurs. On all interior paths, for example, due to
unloading, the response is only elastic. Plastic deformation induces hardening,
corresponding to a nonvanishing value of Ci. Finally, k is a history-parameter vector,
introduced to represent dependence on the history of plastic strain, for example,
through the amount of plastic work.
The yield surface is distorted and moved by plastic strain. In Figure 19.2(a), the
conventional model of isotropic hardening is illustrated in which the yield surface
expands as a result of plastic deformation. This model is unrealistic in predicting a
growing elastic region. Reversed plastic loading is encountered at much higher
stresses than isotropic hardening predicts. An alternative is kinematic hardening (see
Figure 19.2[b]), in which the yield surface moves with the stress point. Within a few
percentage points of plastic strain, the yield surface may cease to encircle the origin.

0749_Frame_C19 Page 245 Wednesday, February 19, 2003 5:34 PM

Inelastic and Thermoinelastic Materials

245

Sll

path of stress point

Sl

Slll

principal stresses Sl Sll Slll

FIGURE 19.2(a) Illustration of yield-surface expansion under isotropic hardening.

Sll

path of stress point

Sl

Slll
FIGURE 19.2(b) Illustration of yield-surface motion under kinematic hardening.

A reference point interior to the yield surface, sometimes called the back stress,
must be identified to serve as the point at which the elastic strain vanishes. Combined isotropic and kinematic hardening are shown in Figure 19.2(c). However, the
yield surface contracts, which is closer to actual observations (e.g., Ellyin [1997]).

0749_Frame_C19 Page 246 Wednesday, February 19, 2003 5:34 PM

246

Finite Element Analysis: Thermomechanics of Solids

Sll

path of stress point

Sl

Slll
FIGURE 19.2(c) Illustration of combined kinematic and isotropic hardening.

The rate of movement must exceed the rate of contraction for the material to remain
stable with a positive tangent modulus.
Combining the elastic and inelastic portions furnishes the tangent-modulus tensor:

= e1 + Ci

= [I + eCi ]1 e .

(19.5)

Suppose that in uniaxial tension, the elastic modulus is Ee and the inelastic
modulus-relating stress and inelastic strain increments are Ei, and Ei << Ee. The total
Ei
uniaxial modulus is then
.
(1 + E i / E e )

19.2 THERMOPLASTICITY
As in Chapter 18, two potential functions are introduced to provide a systematic
way to describe irreversible and dissipative effects. The first is interpreted as the
Helmholtz free-energy density, and the second is for dissipative effects. To accommodate kinematic hardening, we also assume an extension of the Green and Naghdi
(G-N) (1965) formulation, in which the Helmholtz free energy decomposes into
reversible and irreversible parts, with the irreversible part depending on the plastic
strain. Here, it also depends on the temperature and a workless internal state variable.

19.2.1 BALANCE

OF

ENERGY

The conventional equation for energy balance is augmented using a vector-valued, workless internal variable, 0, regarded as representing microstructural rearrangements:

o 0 = sT e r + sT e i T0 q0 + o h + T0 0 ,

(19.6)

0749_Frame_C19 Page 247 Wednesday, February 19, 2003 5:34 PM

Inelastic and Thermoinelastic Materials

247

where o is the internal energy per unit mass in the undeformed configuration and
s = VEC(S), e = VEC(E), and 0 is the flux per unit mass associated with 0. However,
r. Also, c is the internal energy per unit mass, q0 is the
note that = 0, thus i =
heat-flux vector referred to undeformed coordinates, and h is the heat input per unit
mass, for simplicitys sake, assumed independent of temperature. The state variables
are Er , Ei , T, and 0.
The next few paragraphs will go over some of the same ground as for damped
elastomers in Chapter 18, except for two major points. In that chapter, the stress
was assumed to decompose into reversible and irreversible portions in the spirit of
elementary Voigt models. In the current context, the strain shows the decomposition
in the spirit of the classical Maxwell model. In addition, as seen in the following,
it proves beneficial to introduce a workless internal variable to give the model the
flexibility to accommodate phenomena such as kinematic hardening.
The Helmholtz free energy, , per unit mass and the entropy, , per unit mass
are introduced using

= T.

(19.7)

T0 q0 0 h = sT e r + sT e i 0 T 0T 0 + T0 0 .

(19.8)

Now,

19.2.2 ENTROPY-PRODUCTION INEQUALITY


The entropy now satisfies

0 T T0 q0 + 0 h + q0T T/ T
0 sT e r sT e i + 0 T + 0T + q0TT/ T T0 0 .

(19.9)

Viewing r as a differentiable function of er , T, and 0, we conclude that


r / e r = sT

r / T = 0r

r / 0 = T0 r .

(19.10)

Extending the G-N formulation, let s* = 0i /ei and assume that i = i /T


0 = T0i . Now,
and 0i /
T

s*T = 0 i / ei

i = i / T

0 i / 0 = T0 i .

(19.11)

The entropy-production inequality (see Equation 19.9) is now restated as


(sT s*T )e i q0T 0 T/ T 0.

(19.12)

0749_Frame_C19 Page 248 Wednesday, February 19, 2003 5:34 PM

248

Finite Element Analysis: Thermomechanics of Solids

The inequality shown in Equation 19.12 can be satisfied if


(sT s*T )e i 0

(a)

q0T 0 T/ T 0

(b).

(19.13)

The first inequality involves a quantity, s* = VEC(S*), with dimensions of stress.


In the subsequent sections, s* will be viewed as a reference stress, often called the
back stress, which is interior to a yield surface and can be used to characterize the
motion of the yield surface in stress space. In classical kinematic hardening in which
the hyperspherical yield surface does not change size or shape but just moves, the
reference stress is simply the geometric center. If kinematic hardening occurs, as
stated before, the yield surface need not include the origin even with small amounts
of plastic deformation. Thus, there is no reason to regard e r as vanishing at the
origin. Instead, e r = 0 is now associated with a moving-reference stress interior to
the yield surface, identified here as the back stress s*.

19.2.3 DISSIPATION POTENTIAL


As in Chapter 18, we introduce a specific dissipation potential, i, for which
e iT = 0 i i /

(i)

T0 T / T = t 0 i / q 0

(ii)

= s s* (iii),
(19.14a)

from which, with i > 0 and t > 0,

0 i ( / ) + 0 t ( / q 0 )q 0 > 0.

(19.14b)

On the expectation that properties governing heat transfer are not affected by
strain, we introduce the decomposition into inelastic and thermal portions:

= i + t

0 t =

t
q T0 q 0 ,
2

(19.15a)

where i represents mechanical effects and is identified in the subsequent sections.


The thermal constitutive relation derived from the dissipation potential implies
Fouriers law:
0 T / T = q 0 / t .
The inelastic portion is discussed in the following section.

(19.15b)

0749_Frame_C19 Page 249 Wednesday, February 19, 2003 5:34 PM

Inelastic and Thermoinelastic Materials

249

19.3 THERMOINELASTIC TANGENT-MODULUS


TENSOR
The elastic strain rate satisfies a thermohypoelastic constitutive relation:
e r = Cr (s s*) + ar T .

(19.16)

Cr is a 9 9 second-order, elastic compliance tensor, and ar is the 9 1


thermoelastic expansion vector, with both presumed to be known from measurements. Analogously, for rate-independent thermoplasticity, we seek tensors Ci and
ai , depending on , ei , and T such that
e i = Ci (s s*) + ai T

(19.17a)

e = [Cr + Ci ](s s*) + ( ar + ai )T.

(19.17b)

During thermoplastic deformation, the stress and temperature satisfy a thermoplastic yield condition of the form
i (, ei , k, T, i 2 ) = 0,

(19.18)

and i is called the yield function. Here, the vector k is introduced to represent the
effect of the history of inelastic strain, e i , such as work hardening. It is assumed to
be given by a relation of the form
k = K (ei , k, T)e i .

(19.19)

= 0 during thermoplastic flow, from


The consistency condition requires that
i
which
d i
d i
d i d i d i
+
e +
k+
T+
= 0.
d
dei i dk
dT
di i

(19.20)

We introduce a thermoplastic extension of the conventional associated flow rule,


whereby the inelastic strain-rate vector is normal to the yield surface at the current
stress point,
d i
e i = i

(a)

d i
i = i
di

(b).

(19.21)

0749_Frame_C19 Page 250 Wednesday, February 19, 2003 5:34 PM

250

Finite Element Analysis: Thermomechanics of Solids

Equation 19.14a suggests that the yield function may be identified as the dissipation potential: i = 0i. Standard manipulation furnishes
e i = Ci + ai T
T

i
Ci = i
H


ci = i
T

i 2 = biT + ci T
T

i
ai = bi = i
H
T

(19.22)

T
i
i
H = i + i K i +

k
i 2 T
ei

and H must be positive for i to be positive. Note that, in the current formulation,
the dependence of the yield function on temperature accounts for ci. The thermodynamic inequality shown in Equation 19.13a is now satisfied if H > 0.
T
Next, note that s* depends on ei , T, and 0 since s* = 0i /ei. For simplicitys
sake, we neglect dependence on 0 and assume that a relation of the following form
can be measured for s*:
T

s* = e i + T

ei ei I

T = 2 i / ei T.

(19.23)

From Equations 19.16 and 19.17, the thermoinelastic tangent-modulus tensor


and thermal thermomechanical vector are obtained as
e = Cs + aT
C = (C r + Ci )[I (I + Ci ) 1 Ci ]
a = [ ar + ai + (C r + Ci )ai (C r + Ci )(I + Ci ) 1 ]

(19.24a)

(19.24b)

If appropriate, the foregoing formulation can be augmented to accommodate


plastic incompressibility.

19.3.1 EXAMPLE
We now provide a simple example using the Helmholtz free-energy density function
and the dissipation-potential function to derive constitutive relations. The following
expression involves a Von Mises yield function, linear kinematic hardening, linear
work hardening, and linear thermal softening.

0749_Frame_C19 Page 251 Wednesday, February 19, 2003 5:34 PM

Inelastic and Thermoinelastic Materials

251

i. Helmholtz free-energy density:

= r + i

0i = k3eiT ei

0r = eTr Cr1e r / 2 a T Cr1 (T T0 )e r + 0 cr T(1 ln(T / T0 ))

(19.25)

in which cr is a known constant. From Equation 19.28,


= 0 (r / e r )T = Cr1[e r ar ( T T0 )].

(19.26)

Finally,
cr = T

2 r
= c
T 2

(19.27)

ii. Dissipation potential:


= i + t

0 t =

t
q T0 q 0
2

i = T [k0 + k1k k2 (T T0 )] = 0

(19.28)

k =T e i

(19.29)

Straightforward manipulations serve to derive


H = k1 T = k1[k0 + k1k k2 (T T0 )]
Ci =

T
H
T

ci = k22 H

ai = bi =

k2
T

(19.30)

(19.31)

Consider a two-stage thermomechanical loading, as illustrated schematically in


nd
Figure 19.3. Let SI , SII , SIII denote the principal values of the 2 Piola-Kirchhoff
stress, and suppose that SIII = 0. In the first stage, with the temperature held fixed
at T0, the stresses are applied proportionally well into the plastic range. The center
of the yield surface moves along a line in the (S1, S2) plane, and the yield surface
expands as it moves. In the second stage, suppose that the stresses S1 and S2 are
fixed, but that the temperature increases to T1 and then to T2 and T3. The plastic
strain must increase, thus, the center of the yield surface moves. In addition, strain
hardening tends to cause the yield surface to expand, while the increased temperature
tends to make it contract. However, in this case, thermal softening must dominate
strain hardening, and contraction must occur since the center of the yield surface
must move further along the path shown even as the yield surface continues to kiss
the fixed stresses SI and SII.

0749_Frame_C19 Page 252 Wednesday, February 19, 2003 5:34 PM

252

Finite Element Analysis: Thermomechanics of Solids

Sll
T3
T1

T2

T0

T4

T0
T0
T0

T0

Sl
T0

Slll
FIGURE 19.3 Effect of load and temperature on yield surface.

Unfortunately, accurate finite-element computations in plasticity and thermoplasticity often require close attention to the location of the front of the yielded
zone. This front will occur within elements, essentially reducing the continuity order
of the fields (discontinuity in strain gradients). Special procedures have been developed in some codes to address this difficulty.
The shrinkage of the yield surface with temperature provides an element of
the explanation of the phenomenon of adiabatic shear banding, which is commonly
encountered in some materials during impact or metal forming. In rapid processes,
plastic work is mostly converted into heat and on into high temperatures. There
is not enough time for the heat to flow away from the spot experiencing high
deformation. However, the process is unstable while the stress level is maintained.
Namely, as the material gets hotter, the rate of plastic work accelerates, thanks to
the softening evident in Figure 19.3. The instability is manifested in small, periodically spaced bands, in the center of which the material is melted and resolidified,
usually in a much more brittle form than before. These bands can nucleate brittle
failure.

19.4 TANGENT-MODULUS TENSOR


IN VISCOPLASTICITY
The thermodynamic discussion in the previous section applies to thermoinelastic
deformation, for which the first example given concerned quasi-static plasticity and
thermoplasticity. However, it is equally applicable when rate sensitivity is present, in
which case viscoplasticity and thermoviscoplasticity are attractive models. An example

0749_Frame_C19 Page 253 Wednesday, February 19, 2003 5:34 PM

Inelastic and Thermoinelastic Materials

253

of a constitutive model, for example, following Perzyna (1971), is given in undeformed


coordinates as

ei =< 1

k
>
i

i


i

<1

1 k ,
k

>= i
i
0,

k
0
i
otherwise,
if

(19.32)
and i(, ei, k, T, i) is a loading surface function. The elastic response is still
considered linear in the form

e r = r1s + r T.

(19.33)

Recall from thermoplasticity that


T

ei ei i

s* = e i + T

T = 2 i /ei T.

(19.34)

Corresponding to , there is a reference stress s and a corresponding vector =


s s* such that i( , ei, k, T, i) = k(ei, k, T, i) determines a quasi-static, referenceyield surface. The vectors and have the same origin and direction, but the latter
terminates at the reference surface, while the latter terminates outside the reference
surface if inelastic flow is occurring. Interior to the surface, no inelastic flow occurs.
If exterior to the surface, inelastic flow occurs at a rate dependent on the distance to
the exterior of the reference surface. This situation is illustrated in Figure 19.4.
Sll

stress point
Sl

Slll
FIGURE 19.4 Illustration of reference surface in viscoplasticity.

0749_Frame_C19 Page 254 Wednesday, February 19, 2003 5:34 PM

254

Finite Element Analysis: Thermomechanics of Solids

It should be evident that viscoplasticity and thermoviscoplasticity can be formulated to accommodate phenomena such as kinematic hardening and thermal
shrinkage of the reference-yield surface.
The tangent-modulus matrix now reduces to elastic relations, and viscoplastic
effects can be treated as an initial force (after canceling the variation) since


k
s = r e + r r T r < 1
>
i
v

i


i

(19.35)

In particular, the Incremental Principle of Virtual Work is now stated, to first order, as

e edV + e TdV + u udV


T


k
= uT dSo + eT r < 1
>
v
i

i


i

dVo

(19.36)

19.5 CONTINUUM DAMAGE MECHANICS


Ductile fracture occurs by processes associated with the notion of damage. An
internal-damage variable is introduced that accumulates with plastic deformation.
It also manifests itself in reductions in properties, such as the experimental values
of the elastic modulus and yield stress. When the damage level in a given element
reaches a known or assumed critical value, the element is considered to have failed.
It is then removed from the mesh (considered to be no longer supporting the load).
The displacement and temperature fields are recalculated to reflect the element
deletion.
There are two different schools of thought on the suitable notion of a damage
parameter. One, associated with Gurson (1977), Tvergaard (1981), and Thomasson
(1990), considers damage to occur by a specific mechanism occurring in a threestage process: nucleation of voids, their subsequent growth, and their coalescence
to form a macroscopic defect. The coalescence event is used as a criterion for element
failure. The parameter used to measure damage is the void-volume fraction f. Models
and criteria for the three processes have been formulated. For both nucleation and
growth, evolution of f is governed by a constitutive equation of the form
f = ( f , ei , T ),

(19.37)

0749_Frame_C19 Page 255 Wednesday, February 19, 2003 5:34 PM

Inelastic and Thermoinelastic Materials

255

S11

Ep1

Sy1

Ep3

Ep2

Sy3

Sy2
E2

E3

E1

E11
FIGURE 19.5 Illustration of effect of damage on elastic-plastic properties.

for which several specific forms have been proposed. To this point, a nominal stress
is used in the sense that the reduced ability of material to support stress is not
accommodated.
The second school of thought is more empirical in nature and is not dependent
on a specific mechanism. It uses the parameter D, which is interpreted as the fraction
of damaged area Ad to total area Ao that the stress (traction) acts on. Consider a uniaxial
tensile specimen with damage, but experiencing elastic behavior. Suppose that damaged area Ad can no longer support a load (is damaged). For a given load P, the true
stress at a point in the undamaged zone is S = A P A = 1 1 D AP = 1 1 D S . Here, S
o
D
o
is a nominal stress, but is also the measured stress. If E is the elastic modulus
measured in an undamaged specimen, the modulus measured in the current specimen
will be E = E(1 D), demonstrating that damage is manifest in small changes in
properties.
As an illustration of damage, suppose specimens are loaded into the plastic
range, unloaded, and then loaded again. Without damage, the stress-strain curve
should return to its original path. However, due to the damage, there are slight
changes in the elastic slope, in the yield stress, and in the slope after yield (exaggerated in Figure 19.5).
From the standpoint of thermodynamics, damage is a dissipative internal variable.
In reality, the amount of mechanical or thermal energy absorbed by damage is probably
small, so that its role in the energy-balance equation can be neglected. At the risk of
being slightly conservative, in dynamic (adiabatic) problems, the plastic work can be
assumed to be completely converted into heat. Even so, for the sake of a consistent
framework for treating dissipation, a dissipation potential, d, can be introduced for
damage, as has been done, for example, by Bonora (1997). The contribution to the
irreversible entropy production can be introduced in the form D D 0, in which D
is the force associated with flux D . Positive dissipation is assured if we assume

d
,
D

d =

1
(e , T, k ) D 2 ,
2 d i

d (ei , T, k ) > 0.

(19.38)

0749_Frame_C19 Page 256 Wednesday, February 19, 2003 5:34 PM

256

Finite Element Analysis: Thermomechanics of Solids

An example of a satisfactory function is d (ei , T, k) = do (s s*) e i dt, do a


positive constant, showing damage to depend on plastic work.
Specific examples of constitutive relations for damage are given, for example,
in Bonora (1997).
At the current values of the damage parameter, the finite-element equations are
solved for the nodal displacements, from which the inelastic strains can be computed.
This information can then be used to update the damage-parameter values. Upon
doing so, the damage-parameter values are compared to critical values. As stated
previously, if the critical value is obtained, the element is deleted. The path of deleted
elements can be viewed as a crack.
The code LS-DYNA Ver. 9.5, (2000) incorporates a material model that includes
viscoplasticity and damage mechanics. It can easily be upgraded to include thermal
effects, assuming that all viscoplastic work is turned into heat. Such a model has
been shown to reproduce the location and path of a crack in a dynamically loaded
welded structure (see Moraes [2002]).

19.6 EXERCISES
1. In isothermal plasticity, assuming the following yield function, find the
uniaxial stress-strain curve:

i = (s k1ei )T (s k1ei ) ko + k2

e i T e i dt .

Assume small strain and that the plastic strain is incompressible: tr(Ei ) = 0.
2. Regard the expression in Exercise 1 as defining the reference-yield surface
in viscoplasticity, with viscosity hv. Find the stress-strain curve under
uniaxial tension if a constant strain rate is imposed.

2003 by CRC CRC Press LLC

0749_Frame_C20 Page 257 Wednesday, February 19, 2003 5:35 PM

20

Advanced Numerical
Methods

In nonlinear finite-element analysis, solutions are typically sought using Newton


iteration, either in classical form or augmented as an arc-length method to bypass
critical points in the load-deflection behavior. Here, two additional topics of interest
are briefly presented.

20.1 ITERATIVE TRIANGULARIZATION


OF PERTURBED MATRICES
20.1.1 INTRODUCTION
In solving large linear systems, it is often attractive to use Cholesky triangularization
followed by forward and backward substitutions. In computational problems, such as
in the nonlinear finite-element method, solutions are attained incrementally, with the
stiffness matrix slightly modified whenever it is updated. The goal here is to introduce
and demonstrate an iterative method of determining the changes in the triangular
factors ensuing from modifying the stiffness matrix. A heuristic convergence argument is given, as well as a simple example indicating rapid convergence. Apparently, no
efficient iterative method for matrix triangularization has previously been established.
The finite-element method often is applied to problems requiring solution of
large linear systems of the form K00 = f0, in which the stiffness matrix K0 is positivedefinite, symmetric, and may be banded. As discussed in a previous chapter, an
attractive method of solution is based on Cholesky decomposition (triangularization),
T
in which K0 = L0 L 0 and L0 is lower-triangular, and it is also banded if K0 is banded.
The decomposition enables an efficient solution process consisting of forward substitution followed by backward substitution. Often, however, the stiffness matrix is
updated during the solution process, leading to a slightly different (perturbed) matrix,
K = K0 + K, in which K is small when compared to K0. For example, this situation
may occur in modeling nonlinear problems using an updated Lagrangian scheme
and load incrementation. Given the fact that triangular factors are available for K0,
it would appear to be attractive to use an iteration scheme for the perturbed matrix
K, in which the initial iterate is L0. The iteration scheme should not involve solving
intermediate linear systems except by using current triangular factors. A scheme is
introduced in the following section and produces, in a simple example, good estimates
within a few iterations.
The solution of perturbed linear systems has been the subject of many investigations. Schemes based on explicit matrix inversion include the Sherman-MorrisonWoodbury formulae (see Golub and Van Loan [1996]). An alternate method is to
carry bothersome terms to the right side and iterate. For example, the perturbed linear
257

2003 by CRC CRC Press LLC

0749_Frame_C20 Page 258 Wednesday, February 19, 2003 5:35 PM

258

Finite Element Analysis: Thermomechanics of Solids

system can be written as


K 0 = f K 0 K ,

(20.1)

and an iterative-solution procedure, assuming convergence, can be employed as


K 0 ( j +1) = f K 0 K ( j ) .

(20.2)

Unfortunately, in a typical nonlinear problem involving incremental loading, especially in systems with decreasing stiffness, it will eventually be necessary to update
the triangular factors frequently.

20.1.2 NOTATION

AND

BACKGROUND

A square matrix is said to be lower-triangular if all super-diagonal entries vanish.


Similarly, a square matrix is said to be upper-triangular if all subdiagonal entries
vanish. Consider a nonsingular real matrix A. It can be decomposed as
A = A l + diag(A) + A u,

(20.3)

in which diag(A) consists of the diagonal entries of A, with zeroes elsewhere; Al


coincides with A below the diagonal with all other entries set to zero; and Au coincides
with A above the diagonal, with all other entries set to zero. For later use, we
introduce the matrix functions:
lower(A) = Al +

1
diag(A),
2

upper(A) = Au +

1
diag(A).
2

(20.4)

Note that: (a) the product of two lower-triangular matrices is also lower-triangular,
and (b) the inverse of a nonsingular, lower-triangular matrix is also lower-triangular.
Likewise, the product of two upper-triangular matrices is upper-triangular, and the
inverse of a nonsingular, upper-triangular matrix is upper-triangular. In proof of (a),
(1)
(2)
th
let L and L be two n n lower-triangular matrices. The ij entry of the product
(1)
(1) ( 2 )
matrix is given by k =1,n lik lkj . Since L is lower-triangular, lik(1) vanishes unless
k i. Similarly, lkj( 2 ) vanishes unless k j. Clearly, all entries of k =1,n lik(1)lkj( 2 ) vanish
(1) (2)
unless i j, which is to say that L L is lower-triangular.
In proof of (b), let A denote the inverse of a lower-triangular matrix L. We multiply
th
T
the j column of A by L and set it equal to the vector ej (ej = {0.. 1..0} with
th
unity in the j position): now,
l11a1 j

=0

l21a1 j + l22 a2 j

=0

l31a1 j + l32 a2 j + l33a3 j

=0

M
l j1a1 j + l j 2 a2 j + l j 3a3 j + L + l jj a jj = 1

2003 by CRC CRC Press LLC

(20.5)

0749_Frame_C20 Page 259 Wednesday, February 19, 2003 5:35 PM

Advanced Numerical Methods

259
1

Forward substitution establishes that akj = 0, if k < j, and ajj = l jj , thus, A = L is


lower-triangular.

20.1.3 ITERATION SCHEME


Let K0 denote a symmetric, positive-definite matrix, for which the unique triangular
T
factors are L0 and L 0. If K0 is banded, the maximum width of its rows (the bandwidth)
equals 2b 1, in which b is the bandwidth of L0. The factors of the perturbed matrix
K can be written as
[K 0 + K] = [L 0 + L][LT0 + LT ].

(20.6)

We can rewrite Equation 20.6 as


[I + L01 L][I + LT L0T ] = L01 [K 0 + K]L0T ,

(20.7)

L01 L + LT L0T = L01 KL0T L01 LLT L0T .

(20.8)

from which

Note that L0 L is lower-triangular. It follows that

L = L 0 lower L01 KL0T L01 LLT L0T .

(20.9)

The factor of 1/2 in the definition of the lower and upper matrix functions is
motivated by the fact that the diagonal entries of L01 L and LT L0T are the same.
Furthermore, for banded matrices, if L and L0 have the same semibandwidth,
b, it follows that, for the correct value of L, L01 KL0T L01 LLT L0T is also
banded, with a bandwidth no greater than b. Unfortunately, it is not yet clear how
to take advantage of this behavior.
An iteration scheme based on Equation 20.9 is introduced as

L( j +1) = L 0 lower L01 KL0T L01 L( j ) L( j ) L0T

L(1) = L 0 lower L01 KL0T

(20.10)

Explicit formation of the fully populated inverses L01 and L0T can be avoided
by using forward and backward substitution. In particular, L01 K = [L01 k1
L01 k 2 K L01 k n ], where k1 is the first column of K. We can now solve for
b j = L01k j by solving the system L0 bj = kj.

20.1.4 HEURISTIC CONVERGENCE ARGUMENT


For an approximate convergence argument, we use the similar relation
A = K 2A 1 ( A) ( A),

2003 by CRC CRC Press LLC

(20.11)

0749_Frame_C20 Page 260 Wednesday, February 19, 2003 5:35 PM

260

Finite Element Analysis: Thermomechanics of Solids

in which (A) is the solution (converged iterate) for A. Consider the iteration
scheme
A ( j +1) = K 2A 1 ( A) ( A)( j ) .

(20.12)

Subtraction of two successive iterates and application of matrix-norm inequalities furnish


A ( j +2 ) A ( j +1) = 2A 1 A [ A ( j +1) A ( j ) ].

(20.13)

Convergence is assured in this example if (A 1 A ) < 12 , in which s denotes the


spectral radius (see Dahlquist and Bjork [1974]). An approximate convergence
criterion is obtained as
max j ( A ) <
j

1
min (A),
2 k k

(20.14)

in which j(A) denotes the j eigenvalue of the n n matrix A. Clearly, convergence


is expected if the perturbation matrix has a sufficiently small norm. Applied to the current
problem, we also expect convergence will occur if max j | j ( L ) |< 12 mink | k (L) |.
th

20.1.5 SAMPLE PROBLEM


Let L0 and K0 be given by
a
L0 =
b

0
,
c

ab
.
b 2 + c 2

a2
K0 =
ab

(20.15)

Now suppose that the matrices are perturbed according to


a
L=
b

0
,
c + d

a2
K0 =
ab

b 2 + (c + d )2
ab

(20.16)

so that
0
K =
0

2003 by CRC CRC Press LLC

.
d (2c + d )
0

(20.17)

0749_Frame_C20 Page 261 Wednesday, February 19, 2003 5:35 PM

Advanced Numerical Methods

261

We are interested in the case in which d/c << 1, for example, d/c = 0.1, ensuring
that the perturbation is small. We also use the fact that
L01 =

1 c

ac b

0
.
a

(20.18)

The correct answer, which should emerge from the iteration scheme, is
0
L =
0

0
.
d

(20.19)

The initial iterate is found from straightforward manipulation as

1 d .

d 1+

2 c

0
L =
0

(1)

(20.20)

The ratio of the norms of the error is


error =

norm( L(1) ) norm( L )


norm( L )

1 d
d 1 +
d
2 c
=
d

= 5%

(20.21)

Letting = d (1 + 12 dc ), the second iterate is found, after straightforward manipulation, as


0
L( 2 ) =
0

0
1

0
=
0

2 c

0
d 1

d
2
c

1 d3
8 c
3

(20.22)

The relative error is now


error =

d2
c2

1 + 1
8

d
c

1
= 0.011 +
80
Clearly, this is a significant improvement over the initial iterate.

2003 by CRC CRC Press LLC

(20.23)

0749_Frame_C20 Page 262 Wednesday, February 19, 2003 5:35 PM

262

Finite Element Analysis: Thermomechanics of Solids

20.2 OZAWAS METHOD FOR INCOMPRESSIBLE


MATERIALS
In this section, thermal and inertial effects are neglected and the traction is assumed
to be prescribed on the undeformed exterior boundary. Using a two-field formulation
for an incompressible elastomer leads to an incremental relation, in global form, as
follows (see Nicholson, 1995):
K MM
T
K MP

K MP d dfM
= =
.
0 d 0

(20.24)

If KMM is singular, it can be replaced with KMM = KMM + KMP K MP , and can be
chosen to render KMM positive-definite (see Zienkiewicz, 1989).
The presence of zeroes on the diagonal poses computational difficulties, which
have received considerable attention. Here, we discuss a modification of the Ozawa
method discussed by Zienkiewicz and Taylor (1989). In particular, Equation 20.24
is replaced with the iteration scheme
T

K MM

K TMP

K MP d

I / a d

j +1

dfM
= j
,
d / a

(20.25)

th

in which the superscript j denotes the j iterate and ra is an acceleration parameter.


This scheme converges rapidly for suitable choices of ra.
If the assumed pressure fields are discontinuous at the element boundaries, this
method can be used at the element level to eliminate pressure variables (see Hughes,
1987). In this event, the global equilibrium equation only involves displacement
degrees-of-freedom.
For each iteration, it is necessary to solve a linear system. Computation can be
expedited using a convenient version of the LU decomposition. Let L1 and L2 denote
lower-triangular matrices arising in the following Cholesky decompositions:
KMM = L1L1

I/ra + K MPL1L1 KMP = L2L2.


T

(20.26)

Then, a triangularization is attained as


K MM

K TMP

K MP L1
=
I / a K TMP L1T

0 LT1

L2 0T

L11K MP
.
LT2

(20.27)

Forward and backward substitution can now be exploited to solve the linear
system arising in the incremental finite-element method.

2003 by CRC CRC Press LLC

0749_Frame_C20 Page 263 Wednesday, February 19, 2003 5:35 PM

Advanced Numerical Methods

263

20.3 EXERCISES
1. Examine the first two iterates for the matrices
a

K = b

d
a

K + K = b

d
a2

= ab

ad

0 a

0 0

f 0

0
c
e
0
c
e+g

d a2

e = ab

f ad

b
c
0

0 a

0 0

f 0
ab

ab
b2 + c2
bd + ce

bd + ce

d 2 + e 2 + f 2
ad

e + g

b
c
0

bd + c(e + g) .

d 2 + (e + g) 2 + f 2
ad

b2 + c2
bd + c(e + g)

2. Verify that the product and inverse of lower-triangular matrices are lowertriangular using
a
L1 =
b

0
,
c

d
L2 =
e

0
.
f

3. Verify the triangularization scheme in the matrix


2

A = 1

1
2
1

1.

Use the triangular factors to solve the equation


2

Ab = 1.

1

2003 by CRC CRC Press LLC

0749_Frame_Ref Page 265 Wednesday, February 19, 2003 5:38 PM

Monographs and Texts


Abramawitz, M. and Stegun, I., Eds, Handbook of Mathematical Functions, with Formulas,
Graphs and Mathematical Tables, Dover Publications, Inc., New York, 1997.
Belytschko, T., Liu, W.K., and Moran, B., Nonlinear Finite Elements for Continua and
Structures, John Wiley and Sons, New York, 2000.
Bonet, J. and Wood, R., Nonlinear Continuum Mechanics for Finite Element Analysis, Cambridge University Press, Cambridge, 1997.
Brush, D. and Almroth, B., Buckling of Bars, Plates and Shells, McGraw-Hill Book Company,
New York, 1975.
Callen, H.B., Thermodynamics and an Introduction to Thermostatistics, 2nd ed., John Wiley
and Sons, New York, 1985.
Chandrasekharaiah, D. and Debnath, L., Continuum Mechanics, Academic Press, San Diego,
1994.
Chung, T.J., Continuum Mechanics, Prentice-Halls, Englewood Cliffs, NJ, 1988.
Dahlquist, G. and Bjork, A., Numerical Methods, Prentice-Hall, Englewood Cliffs, NJ, 1974.
Ellyin, F., Fatigue Damage, Crack Growth and Life Prediction, Chapman & Hall, FL, 1997.
Eringen, A.C., Nonlinear Theory of Continuous Media, McGraw-Hill, New York, 1962.
Ewing, G.M., Calculus of Variations with Applications, Dover Publications, Mineola, N.Y.,
1985.
Gent, A.N., Ed, Engineering with Rubber: How to Design Rubber Components, Hanser, New
York, 1992.
Golub, G. and Van Loan, C., Matrix Computations, Johns Hopkins University Press, Baltimore,
MD, 1996.
Graham, A., Kronecker Products and Matrix Calculus with Applications, Ellis Horwood, Ltd.,
Chichester, 1981.
Hughes, T., The Finite Element Method: Linear Static and Dynamic Analysis, Dover Publishers Inc., New York, 2000.
Kleiber, M., Incremental Finite Element Modeling in Non-Linear Solid Mechanics, Ellis
Horwood, Ltd., Chichester, 1989.
Oden, J.T., Finite Elements of Nonlinear Continua, McGraw-Hill Book Company, New York,
1972.
Schey, H.M., DIV, GRAD, CURL and All That, Norton, New York, 1973.
Thomason, P.F., Ductile Fracture of Metals, Pergamon Press, Oxford, 1990.
Wang, C-T., Applied Elasticity, The Maple Press Company, York, PA, 1953.
Zienkiewicz, O.C. and Taylor, R.L., The Finite Element Method. Volume 2, 4th ed, McGrawHill Book Company, London, 1989.

265

2003 by CRC CRC Press LLC

0749_Frame_Ref Page 267 Wednesday, February 19, 2003 5:38 PM

Articles and Other Sources


ANSYS User Manual, ver 6.0, Swanson Analysis Systems, 2000.
Blatz, P.J. and Ko, W.L., Application of finite elastic theory to the deformation of rubbery
materials, Trans. Soc. Rheol., 6, 223, 1962.
Bonora, N., A nonlinear CDM model for ductile failure, Engineering Fracture Mechanics,
58, 1/2, 11, 1997.
Chen, J., Wan, W., Wu, C.T., and Duan, W., On the perturbed Lagrangian formulation for
nearly incompressible and incompressible hyperelasticity, Comp. Methods Appl. Mech.
Eng., 142, 335, 1997.
Dillon, O.W., A nonlinear thermoelasticity theory, J. Mech. Phys. Solids, 10, 123, 1962.
Green, A. and Naghdi, P., A general theory of an elastic-plastic continuum, Arch. Rat. Mech.
Anal., 18/19, 1965.
Gurson, A.L., Continuum theory of ductile rupture by void nucleation and growth, part 1: yield
criteria and flow rules for porous ductile media, J. Eng. Mats Technol., 99, 2, 1977.
Holzappel, G., On large strain viscoelasticity: continuum applications and finite element
applications to elastomeric structures, Int. J. Numer. Meth. Eng., 39, 3903, 1996.
Holzappel, G. and Simo, J., Entropy elasticity of isotropic rubber-like solids at finite strain,
Comp. Methods Appl. Mech. Eng., 132, 17, 1996.
LS-DYNA, ver. 95, Livermore Software Technology Corporation, Livermore, CA, 2000.
Moraes, R. and Nicholson D.W., Local damage criterion for ductile fracture with application to welds under dynamic loads, in Advances in Fracture and Damage Mechanics, Guagliano, M. and Aliabadi, M.H., Eds., 2nd ed. Hoggar Press, Geneva, 277,
2002.
Nicholson, D.W. and Nelson, N., Finite element analysis in design with rubber (Rubber
Reviews), Rubber Chem. Tech., 63, 638, 1990.
Nicholson, D.W., Tangent modulus matrix for finite element analysis of hyperelastic materials,
Acta Mech., 112, 187, 1995.
Nicholson, D.W. and Lin, B., Theory of thermohyperelasticity for near-incompressible elastomers, Acta Mech., 116, 15, 1996.
Nicholson, D.W. and Lin, B., Finite element method for thermomechanical response of nearincompressible elastomers, Acta Mech., 124, 181, 1997a.
Nicholson, D.W. and Lin, B., Incremental finite element equations for thermomechanical of
elastomers: effect of boundary conditions including contact, Acta Mech., 128, 12,
81, 1997b.
Nicholson, D.W. and Lin, B., On the tangent modulus tensor in hyperelasticty, Acta Mech.,
131, 1997c.
Nicholson, D.W., Nelson, N., Lin, B., and Farinella, A., Finite element analysis of hyperelastic
components, Appl. Mech. Rev., 51, 5, 303, 1999.
Ogden, R.W., Recent advances in the phenomenological theory of rubber elasticity, Rubber
Chem. Tech., 59, 361, 1986.
Perzyna, P., Thermodynamic theory of viscoplasticity, Adv. Appl. Mech., 11, 1971.
Valanis, K. and Landel, R.F., The strain energy function of a hyperelastic material in terms
of the extension ratios, J. Appl. Physics, 38, 2997, 1967.

267

2003 by CRC CRC Press LLC

0749_Frame_Ref Page 268 Wednesday, February 19, 2003 5:38 PM

268

Finite Element Analysis: Thermomechanics of Solids

Tvergaard, V., Influence of voids on shear band instabilities under plane strain conditions,
Int. J. Fracture, 17, 389407, 1981.
Ziegler, H. and Wehrli, C., The derivation of constitutive relations from the free energy and
the dissipation function, Adv. Appl. Mech., 25, 187, 1987.

2003 by CRC CRC Press LLC

S-ar putea să vă placă și