Você está na página 1de 458

FOUNDATIONS OF SOLID MECHANICS

SOLID MECHANICS AND ITS APPLICATIONS


Volume 3
Series Editor:

G.M.L. GLADWELL

Solid Mechanics Division, Faculty 0/ Engineering


University o/Waterloo
Waterloo, Ontario. Canada N2L 3Gl

Aims and Scope of the Series


The fundamental questions arising in mechanics are: Why?, How?, and How much?
The aim of this series is to provide lucid accounts written by authoritative researchers giving vision and insight in answering these questions on the subject of
mechanics as it relates to solids.
The scope of the series covers the entire spectrum of solid mechanics. Thus it
includes the foundation of mechanics; variational formulations; computational
mechanics; statics, kinematics and dynamics of rigid and elastic bodies; vibrations
of solids and structures; dynamical systems and chaos; the theories of elasticity,
plasticity and viscoelasticity; composite materials; rods, beams, shells and
membranes; structural control and stability; soils, rocks and geomechanics;
fracture; tribology; experimental mechanics; biomechanics and machine design.
The median level of presentation is the first year graduate student. Some texts are
monographs defining the current state of the field; others are accessible to final
year undergraduates; but essentially the emphasis is on readability and clarity.

For a list o/related mechanics titles, see final pages.

Foundations of
Solid Mechanics

by
P. KARASUDHI
Asian Institute o/Technology, Bangkok, Thailand

SPRINGER SCIENCE+BUSINESS MEDIA, B.V.

Library of Congress Cataloging-in-Publication Data


Ptsidhi Karasudhi.
Foundations of solId mechanics / by Plsidhi Karasudhi.
p.
cm. -- (Sol id mechanics and Its applications
v. 3)
Includes bibliographical references and indexes.
ISBN 978-94-010-5695-3
ISBN 978-94-011-3814-7 (eBook)
DOl 10.1007/978-94-011-3814-7

1. Strength of materials.
Series.
TA405.P54 1990
620. 1 '05--dc20

n.

2. Mechanics. Applied.

I.

Title.
90-48510

ISBN 978-94-010-5695-3

Printed on acid-free paper

All Rights Reserved


1991 Springer Science+Business Media Dordrecht
Originally published by Kluwer Academic Publishers in 1991
Softcover reprint of the hardcover 1st edition 1991
No part of the material protected by this copyright notice may be reproduced or
utilized in any form or by any means, electronic or mechanical,
including photocopying, recording or by any information storage and
retrieval system, without written permission from the copyright owner.

..... in memory of, and dedicated to

my father.

TABLE OF CONTENTS
PREFACE
LIST OF SYMBOLS
I

Tensors and continuum mechanics


Scalars and vectors
Indicia! notation
Algebra of Cartesian tensors
Matrices and determinants
Linear equations and Eigenvalue problem
Theorems on tensor fields
Differential geometry
Dirac-delta and Heaviside step functions
Bessel functions
Laplace transforms
Inverse Laplace transforms
One-to-one mappings
Curvilinear coordinates
Derivatives with respect to curvilinear coordinates
Exercise problems

1
1
5
7
9
12
17
19
23
25
28
32
34
36
41
44

STRESS AND STRAIN TENSORS

2.1
2.2
2.3
2.4
2.5
2.6
2.7
2.8
2.9
2.10
2.11

2.12

xiii

MATHEMATICAL FOUNDATIONS

1.1
1.2
1.3
1.4
1.5
1.6
1.7
1.8
1.9
1.10
1.11
1.12
1.13
1.14
1.15
1.16

xi

Introduction
Force distribution and stresses
Stress vector and equations of mation
Euler's laws of motion
Stress tensor
Stationary shear stresses
Octahedral shear stress and stress deviator
Strain tensor
Compatibility conditions
Cylindrical and spherical coordinates
Problems
2.11.1 Stress or strain computation from three different
normal components
2.11.2 Cylindrical and spherical rotation components
2.11.3 Rigid-body rotation and translation components
Exercise problems

47
47
49
52
54
57
59
61
64

67
73
73

74
74
76

LINEAR ELASTICITY

3.1
3.2
3.3
3.4

Strain energy function


Orthotropic and isotropic elastic solids
Young's moduli and Poisson's ratios for orthotropic elastic solids
Solution schemes

86
91
95
98

Foundations of Solid Mechanics

viii

3.5
3.6

IV

4.3
4.4
4.5

4.6
4.7

102
102
103

104
106
107
107
108

Plane problems of orthotropic elastic materials


Airy function for isotropic plane problems
Isotropic elastic plane problems in cylindrical coordinates
Displacement for a given bihannonic function
Examples of infinite plane problems
Particular solutions for concentrated forces
Exercise problems
Table 4.1
Complementary and particular solutions for
elastostatics of isotropic planes

111
113
114
116
121
129
132

147

BENDING OF ELASTIC TIHN PLATES


5.1
5.2
5.3
5.4
5.5
5.6
5.7
5.8
5.9
5.10

VI

100

ELASTOSTATIC PLANE PROBLEMS


4.1
4.2

Field equations in tenns of displacements


Problems
3.6.1
Beltrami-Michell compatibility conditions
3.6.2
Conservative forces and potentials
3.6.3
Elastostatic displacement potentials
3.6.4
Elastodynamic displacement potentials
3.6.5
Positive definiteness of the strain energy function
3.6.6
Stress and strain computation from measured results
3.6.7
Saturated porous elastic media

Basic assumptions
Equilibrium, boundary conditions and stress resultants
Physical meaning of stress resultants
Governing conditions for isotropic plates
Solutions for rectangular plates
Closed fonn solutions for circular plates
Series solutions for circular plates
Polygonal plates supported at comers
Plates on elastic foundation
Exercise problems
Complementary and particular solutions for
Table 5.1
elastostatic bending of thin isotropic plates

154
156
159

162
164
168
175

178
180
183
190

ELASTOSTATICS WITH DISPLACEMENTS AS UNKNOWNS


6.1
6.2
6.3
6.4
6.5
6.6
6.7

6.8
6.9

Field equations for plane problems


Solution scheme for large planes
Solution scheme for large spaces
Homogeneous half planes and half spaces
Concentrated force inside a half space
Load transfer problems
Infinite elements for multilayered half spaces
Saturated large spaces
Exercise problems

196
197
203
207

211
217
225
229
231

Tables of Contents

vn

LINEAR VISCOELASTICITY
7.1
7.2
7.3
7.4
7.5
7.6
7.7
7.8
7.9
7.10
7.11
7.12

VIII

Linear elasticity and Newtonian viscosity


Creep and relaxation
Compliance and modulus of mechanical models
Differential equations for stress-strain relationship
Steady state harmonic oscillation
Tbermorheologically simple solids
Three-dimensional theory
Quasi-static solution by separation of variables
Steady state harmonic solution scheme
Integral ttansform methods and their limitations
Three-dimensional thermoviscoelasticity
Problems
7.12.1 Reciprocal theorem for harmonic oscillation
7.12.2 Vibration of a bar with a viscoelastic support
7.12.3 Indentation on a viscoelastic half space
7.12.4 Torsional oscillation of a hollow cylinder
7.12.5 Quasi-static torsional oscillation of a hollow cylinder
7.12.6 Dynamic response of an incompressible cylinder
7.12.7 Isothermal harmonic vibration
7.12.8 Isothermal effects on stretched string
7.12.9 Varying temperature effects on stretched string
7.12.10 Heating of an infinite slab

235
237
240
245
253
260
265
268
269
270
273
274
274
275
277
277
278
280
282
283
284
286

WAVE PROPAGATION
8.1
8.2
8.3
8.4
8.5
8.6
8.7
8.8
8.9
8.10
8.11
8.12
8.13
8.14

IX

ix

Terminology in wave propagation


Wavefront and jumps
Velocity jumps in isotropic elastic domains
Reflection and tansmission at interfaces and boundaries
Waves in isotropic viscoelastic media
In-plane harmonic surface waves
Antiplane harmonic surface waves
Vibration of multilayered elastic half spaces
Asymmetric vibration of a homogeneous half space
Axisymmetric torsion of a layered half space
Total solution to vibration of half planes
Vibration of viscoelastic half spaces
Infinite elements for a homogeneous half space
Exercise problems

288
296
305
309
318
320
326
329
335
343
346
347
348
352

PLASTICITY
9.1
9.2
9.3
9.4
9.5
9.6

Facts from simple tests


Basic assumptions and common characteristics of various theories
Various yield functions
Hardening and flow rules
Incremental formulation for isotropic hardening
Viscoplasticity

355
358
361
364
369
374

Foundations of Solid Mechanics

FINITE DEFORMATION

10.1
10.2
10.3
10.4
10.5
10.6
10.7
10.8
10.9
10.10
10.11
10.12

Different descriptions of changing configuration


Material derivative and conservation of mass
Stress tensors in different descriptions
Equations of motion in different descriptions
Finite strain tensors
Reformed Lagrangian description
Strain tensors in curvilinear coordinates
Equilibrium equations and stress tensors in curvilinear coordinates
Physical components of vectors and tensors
Boundary conditions and constitutive relationship in curvilinear coordinates
Compatibility conditions
Problems
10.12.1 Constitutive law in Eulerian description
10.12.2 Maxwell-Betti reciprocal theorem for fmite deformation

378
380
386
391

394
396

401
402
406
410
412
416
416
416

REFERENCES

419

AUTHOR INDEX

429

SUBJECT INDEX

433

PREFACE
This book has been written with two purposes, as a textbook for engineering
courses and as a reference book for engineers and scientists. The book is an outcome
of several lecture courses. These include lectures given to graduate students at the
Asian Institute of Technology for several years, a course on elasticity for University of
Tokyo graduate students in the spring of 1979, and courses on elasticity, viscoelasticity
and ftnite deformation at the National University of Singapore from May to November
1985.
In preparing this book, I kept three objectives in mind: ftrst, to provide sound
fundamental knowledge of solid mechanics in the simplest language possible; second,
to introduce effective analytical and numerical solution methods; and third, to impress
on readers that the subject is beautiful, and is accessible to those with only a standard
mathematical background.
In order to meet those objectives, the ftrst chapter of the book is a review of
mathematical foundations intended for anyone whose background is an elementary
knowledge of differential calculus, scalars and vectors, and Newton's laws of motion.
Cartesian tensors are introduced carefully. From then on, only Cartesian tensors in the
indicial notation, with subscript as indices, are used to derive and represent all theories.
Any combination of indicial and Gibbs notations is avoided except in the sections on
curvilinear coordinates in the :ftnite deformation chapter. Conditions under small
deformation for cylindrical and spherical coordinates are put in explicit symbols. Most
of the pertinent theorems and formulas are compiled, proved and/or verifted. The only
theorems and formulas which are quoted without proofs are those which can be seen
in standard mathematical books. Whenever possible, presentations are made by
induction processes, i.e. emerging from the simplest special cases to the most general.
New analytical tools and methods such as the Dirac-delta distribution, integral transforms
and integral equations are introduced along with their limitations. It is emphasized that
an effective solution must have a rational basis. Where deemed appropriate, tables are
provided to save mundane though straightforward operations. Approximate formulas
and proven numerical algorithms are brought to the attention of the reader. All exercise
problems are accompanied by hints and/or answers.
Readers will learn that the major conditions governing the mechanics of a solid
domain are the equilibrium equations, strain-displacement relationships, constitutive
relationships, and boundary conditions. Chapters II to VIII of this book are concerned
with linear solid continua, Chapters IX and X with nonlinear. Both static and dynamic
linear problems, and static and quasi-static nonlinear problems are treated.
I have used the book as a textbook in three graduate courses:
1) Introduction to Solid Mechanics - mostly concerned with linear elastostatics.
This uses Sections 1.1 to 1.10 and 1.16 and Chapters II to VI.
2) Advanced Solid Mechanics - time dependent constitutive relationships,
xi

xii

Foundations of Solid Mechanics

material nonlinearity, and geometrical nonlinearity. This uses Sections 1.11


to 1.15 and Chapters VIT, IX and X.
3) Wave Propagation in Elastic and Viscoelastic Media - special emphasis on
half spaces. This uses Chapter VIll.
A lecturer using this book should fmd it possible to present it to students
equation-by-equation, page-by-page, and section-by-section. The book is also supposed
to contain all the essentials of solid mechanics normally expected from a good reference
book on the subject. In addition, the arrangement of the contents in this book, together
with the Author and Subject Indices and the List of Symbols, should make it simple to
use.
My appreciation of solid mechanics was enhanced through many years of association with Professors J.D. Achenbach, J. Dundurs, L.M. Keer, and S.L. Lee of
Northwestern University. I would like to acknowledge my indebtedness to the lectures
and publications of those professors. I am grateful for the warm friendship, constructive
criticism, and excellent hospitality provided by Professor F. Nishino of the University
of Tokyo during my sabbatical stay at the University of Tokyo in 1979. I cannot fmd
suitable words to record my heartfelt gratitude to Professor S.L. Lee. He was my
professor at Northwestern, my colleague at the Asian Institute of Technology, my host
during the sabbatical leave spent at the National University of Singapore in 1985, my
close friend for more than two decades, and he has given me many valuable suggestions
and constant encouragement.
I am indebted to Professor G.M.L. Gladwell of the University of Waterloo for
reviewing the book and suggesting several changes in it.
This book has been written in memory of the person who was the ftrst one who
taught me reading, writing and arithmetic, and who was always ready to give me love
and care. That was my own late father, to whom this piece of work is humbly dedicated.

PK.

LIST OF SYMBOLS

Common symbols are defined when they first appear, and when used at other
places stand for the following (unless specified differently):

C.P.V.

bulk creep compliance;


Laplace transform of bulk creep compliance;
Cauchy's'principal value;

cp

pressure wave speed;

Cs

shear wave speed;

Young's modulus, Eqs. 3.36 to 3.38;

Laplace transform of Young's modulus;

eij

strain deviator, Eq. 2.47;

H(

Heaviside step function, Eqs. 1.104 and 1.106;

H2>(

a Hankel function or third kind Bessel function of ( );

H:;>( )

another Hankel function or third kind Bessel function of ( );

$(

imaginary part of ( );

imaginary number, i2 = -1;


J

J
J m(

shear creep compliance;


Laplace transform of shear creep compliance;
first kind Bessel function of ( );

Mr

bulk modulus;
Laplace transform of bulk modulus;
a plate bending moment in cylindrical coordinates;

Mre

plate twisting moment in cylindrical coordinates;

M;x,My

plate bending moments in Cartesian coordinates, Eqs. 5.lOa and b;

M",

plate twisting moment in Cartesian coordinates, Eq. 5.1Oc;

Me

another plate bending moment in cylindrical coordinates;

nj

unit normal vector to a surface;

Laplace transform parameter;

Qr

a transverse shear force in plate bending in cylindrical coordinates;

Q;x,Qy

transverse shear forces in plate bending in Cartesian coordinates,

Eqs.5.20;
xiii

Foundations of Solid Mechanics

xiv

Qa

another transverse shear force in plate bending in cylindrical

coordinates;
radial spherical coordinate, Fig. 2.12;
real part of ( );

~(

radial cylindrical coordinate, Fig. 2.12;

surface;
stress deviator, Eq. 2.34;

Sij

temperature; or
transpose of a matrix when appears as a superscript;
time;
displacement component in X -direction;

t
U

displacement vector;
displacement vector;

Ui
V

volume;

Vi

velocity jump, Eq. 8.72;

V,

a supplemented shear force in plate bending in cylindrical

VX,V)/

coordinates, Eq. 5AOa;


supplemented shear forces in plate bending in Cartesian coordinates;

Va

Eqs.5.26;
another supplemented shear force in plate bending in cylindrical

coordinates, Eq. 5AOb;


displacement component in y-direction;

Vi

Cartesian components of velocity;

displacement component in z -direction, plate deflection;

body force in x-direction;

surface force in x -direction;

body force vector;

Xi

body force vector;

surface force vector;

Xi

surface force vector;

Xl;

Xi

original position vector, Lagrangian coordinates;

Ym(
y

second kind Bessel function of (

);

List of Symbols

Yi

position vector after defonnation, Eulerian coordinates;

<X/11)

a radical quantity related to pressure wave, Eq. 8.167a;

a (11)

i<Xp (11), Eq. 8.173;

<X.(11)

a radical quantity related to shear wave, Eq. 8.167b;

<Xs(rt)

i<X.(11), Eq. 8.171;

virtual quantity, variation symbol;


Dirac-delta function, Eqs. 1.102, 1.103 and 1.106;
Kronecker delta, Kronecker symbol, Eq. 1.23;

B(

Green's strain tensor, Eq. 10.98;


Almansi's strain tensor, Eqs. 2.44 and 10.99;
pennutation tensor, Eq. 1.33;
a complex variable, 11 + is;
transfonn parameter, 9t(~);
dimensionless surface Love wavenumber;
dimensionless pressure wavenumber, Eq. 8.164b;
dimensionless surface Rayleigh wavenumber;
11.

dimensionless shear wavenumber, Eq. 8.164c;


angular cylindrical coordinate = an angular spherical coordinate,
Fig. 2.12;
curvilinear coordinates, Fig. 1.14;
an elastic constant in plane problems, Eqs. 4.8;
a Lame's constant;
Laplace transfonn of A;

Jl
V

another Lame's constant, shear modulus;


Laplace transfonn of shear modulus;
Poisson's ratio;

Laplace transfonn of Poisson's ratio;

Z(~);

IT

product series, Eq. 7.29g;

mass density;

xv

Foundations of Solid Mechanics

xvi
~ij

Kirchhoff stress tensor, Eq. 10.58;

(Jij

Eulerian stress tensor, Figs. 2.2 and 10.6;

<I>

'I'

another angular spherical coordinate, Fig. 2.12; or


stress function, Chapter IV;
wavefront function, Eqs. 8.49;
angular frequency;

ro
VZ

Laplace operator, Eq. 1.85d; and

[]

jump = discontinuity at wavefront, Eq. 8.39b.

CHAPTER I
MATHEMATICAL FOUNDATIONS

1.1 TENSORS AND CONTINUUM MECHANICS

Continuum mechanics is that part of physical science dealing with the defonnation
and motion of continuous material media under the influence of external agencies. The
foundations of the theory involve laws of motion and constitutive laws that are conveniently stated in tenns of tensors. As a mathematical entity, a tensor has an existence
independent of any coordinate system, yet it may be specified in a particular coordinate
system by a certain set of quantities known as its components. Components of a tensor
in different coordinate systems are related by the law of transformations, to be elaborated
later.
When attention is restricted to transfonnations from one rectangular coordinate
system to another, the tensors involved are referred to as Cartesian tensors. Since much
of the theory of continuum mechanics may be developed in tenns of Cartesian tensors,
the word 'tensor' in this text means 'Cartesian tensor' unless specifically stated otherwise. Physical laws of continuum mechanics are expressed by tensor equations. If
such an equation is valid in one (Cartesian) coordinate system, it is valid in any other
(Cartesian) coordinate system.
Tensors are classified by rank or order according to the number of components
they possess. In three-dimensional Euclidean space such as the ordinary physical space,
a tensor of order N has 3N components. Accordingly, a tensor of order zero is
specified in any coordinate system by one component. Such quantities are also known
as scalars. A tensor of order one has three coordinate components in three dimensional
Euclidean space and known as a vector. A second order tensor is called a dyadic. Third
order tensors, triadics, and fourth order tensors, tetradics, also appear in the mathematics
of continuum mechanics.
1.2 SCALARS AND VECTORS
Certain physical quantities, such as mass, length, time, temperature, energy, etc.
possess magnitude only and are scalars, tensors of order zero. Quantities such as force,

Foundations of Solid Mechanics

displacement, velocity, acceleration, etc. which possess both magnitude and direction
are vectors, ftrst order tensors. In three-dimensional space, a vector may be represented
as an arrow pointing in the appropriate direction and having a length proportional to
the magnitude of the vector. Two vectors, regardless of their position, are equal if they
are in the same direction and have the same magnitude. The negative of a vector is
that vector having the same magnitude but opposite direction. A unit vector is a vector
of unit magnitude. In the symbolic or Gibbs notation, scalars are denoted by italic
letters such as a, b, c etc., vectors by bold-faced letters such as a, b, etc. Unit
vectors are often distinguished by a caret placed over the bold-faced letter, e.g. at. In
Fig. 1.1, arbitrary vectors a and b are shown along with a unit vector
and a pair
of equal vectors c and d . The magnitude of any vector a is written as a.

Fig. 1.1 Pictorial views of vectors in three-dimensional space.

Vector addition obeys the parallelogram law, which deftnes a vector sum of two
vectors as the diagonal of a parallelogram having the summed vectors as adjacent sides.
This law is the same as the triangle rule which defmes the sum of two vectors as the
vector extending from the tail of the ftrst to the head of the second when the summed
vectors are adjoined head to tail. The graphical construction for the addition of a and
b by the parallelogram law is shown in Fig. 1.2(a). Algebraically, the addition process
is expressed by the vector equation
(1.1)
a+b=b+a=c
-b

a+b+q=h

(a)

(b)

Fig. 1.2 Vector addition and subtraction.

(c)

Mathematical Foundations

Vector subtraction is accomplished by addition of the negative vector as shown, for


example, in Fig. 1.2(b) where the triangular rule is used. Thus
a-b=-b+a=d
(1.2)
Thus Eqs. 1.1 and 1.2 state that the operations of vector addition and subtraction are
commutative. Extensions of such operations to more than two tensors are immediate,
since such operations are also associative, as illustrated by the case in Fig. 1.2(c) for
which the appropriate equations are
(a+b)+q =a+(b+q) =h
(1.3)
Multiplying a vector by the reciprocal of its magnitude results in a unit vector in
the direction of the original vector. This relationship is expressed by the equation
b=blb
(1.4)
The dot or scalar product of two vectors a and b is the scalar
(1.5)
A=a b=b a=ab cosO
in which 0 is the angle between the two vectors. Evidently [Fig. 1.3(a)], a b may be
regarded as the quantity obtained by multiplying the magnitude a of a by the projection
b cos 0 of b upon a, or as the quantity obtained by multiplying b by the projection
of a upon b.

v=axb

a
(b) Cross product.

(a) Dot product.


Fig. 1.3 Products of vectors.

The cross or vector product of a into b is the vector v given by


v=a xb=-bxa= (ab sinO)v
(1.6)
in which 0 is the angle less than 1t between the vectors a and b, and v is perpendicular to the plane of a and b so sensed that a, b and v form a right-handed
system, i.e. a right threaded screw rotating through the angle 0 from a to b will
advance in the direction of v. Note that the magnitude of ax b is equal to the area
of the parallelogram whose adjacent sides are a and b shown shaded in Fig. 1.3(b).

Foundations of Solid Mechtmics

The cross product is not commutative. In general, three vectors a, b and c which have
coincident initial points are said to fann a right-handed system or dextral system if a
right threaded screw rotated through the angle less than 1t from a to b will advance
in the direction c as shown in Fig. 1.4.

a
Fig. 1.4 A right-handed or dextral system fanned by a, b and c.
The pre/e"ed coordinate systems for a Euclidean space of any dimension (to be
described more in Section 1.13) are'rectangular Cartesian coordinate systems, or simply
rectangular coordinate systems for brevity. For a three dimensional Euclidean space,
the well-known rectangular Cartesian coordinate system is often represented by mutually
perpendicular axes x, y, z shown in Fig. 1.5. It should be noted that, in this text, only
right-handed coordinate systems are used unless stated otherwise. Any vector v in
such a system may be expressed as a linear combination of three arbitrary, nonzero,
noncoplanar vectors of the system, which are called base vectors. The most frequent
choice of base vectors for the rectangular Cartesian system is the set of unit vectors f,
], k along the coordinate axes as shown in Fig. 1.5. These base vectors constitute a
right-handed unit vector triad, for which
f f=].] =k k= 1,
f.]=].k=k.f=o

(1.7)

and
A=-":

kXI=J,
co

Ixi=jxj=kxk=O
(1.8)
In terms of unit triads f, j, k, the vector v shown in Fig. 1.5 may be expressed
by

v=v%l+v,.i+v,k

(1.9)

in which the Cartesian components


v% =vf=vcosa

(1.10a)

Mathematical Foundations

v)/ =v.j=vcosP

(1.10b)

v. = V k = V cosy

(1.1Oc)

are the projections of v on to the coordinate axes. The unit vector in the direction of
v is given according to Eq. 1.4 by
(1.11)
V =v/v

=(cosa.)f+ (cosP)j + (cosy)k

(1.12)

It follows that the unit vector corresponding to a given vector has the direction cosines
of that vector as its Cartesian components.

z
v
~_""...L..-

__.. y

x
Fig. 1.S Cartesian unit base vectors and direction cosines of a vector v.
1.3 INDICIAL NOTATION
The components of a tensor of any order, and indeed the tensor itself, can be
represented clearly and concisely by the use of indicial notation. In this notation, letter
indices (either subscripts or superscripts) are appended to the generic or kernel letter
representing the tensor quantity of interest. Typical examples illustrating the use of
indices are the tensor symbols

ai' b i , Tij' Uk' RPf.


When an index occurs unrepeated in a term, that index is known as a free index
and can take on one of the values 1,2, ... , N, where N is a specified integer that
determines the range of the index. The tensor rank of a given term is equal to the
number of free indices appearing in that term. A correctly written tensor equation must
have the same letters as free indices in every term.
The repetition of an index in a term denotes a contraction, i.e. summation with
respect to that index over its range. In such so-called Einstein summation convention,

Foundations of Solid Mechanics

repeated indices are often referred to as dummy or umbral indices, since their
replacement by any other letter not appearing as a free index does not change the
meaning of the term in which they occur. No index occurs more than twice in a properly
written term. First order tensors (Le. vectors) are denoted by kernel letters bearing one
free index. Thus a vector a is represented by a symbol having a single subscript or
superscript, Le. in one or the other of the two forms,
j

aj' a.

The following terms, having only one free index, are also recognized as first order
tensors,

aijbj, Fikk' eijkUjvk


Second order tensors are denoted by symbols having two free indices. Thus a dyadic
D will appear in one of the two possible forms,
Dij' Dij.

By a logical continuation of the above scheme, third order tensors are expressed by
symbols with three free indices, while symbols such as A and ea which have no free
indices represent scalars or zero order tensors.
When only rectangular Cartesian coordinates are considered, the distinction
between covariance and contravariance (to be discussed later in Section 1.14) is
immaterial, and we can choose to write all indices of a tensor as subscripts. In curvilinear coordinate systems, however, such distinction must be recognized by using
subscripts for covariance and superscripts for contravariance.
In ordinary physical space, a basis is composed of three non-coplanar vectors, so
any vector in this space is completely specified by its three components in the directions
of those base vectors, and the range of an index is 3. Accordingly, the symbol aj
(where i = 1,2,3) is understood to represent a vector a completely, and also the i th
component of the vector, Le. the component of a in the direction of the i th base vector.
For such a range of three on both indices, the symbol Aij represents the second order
tensor (dyadic) A, which has nine components. The tensor Aij is often presented
explicitly by the nine components in a square array, i.e.
All A12
[Aiji = [A2l An
A31

A32

Al~

(1.13)

A23
A3

In the same way, the component of a first order tensor (vector) in three-dimensional
spaces may be displayed explicitly by a row or column arrangement of the form

(<>;) =(a,

a, aJ

or

{<>;}

{j

In general, an n th order tensor with a range N has Nil components.

(1.14)

Mathematical Foundations

The usefulness of the indicial notation in presenting systems of equations in


compact fann is illustrated by the following two examples. For a range of three on
both i and j the indicial equation
(1.15)
represents the three equations
Xl

=O'Unl + 0'1211z + 0'13~

=0'21 n l + O'zzllz + 0'23~


X 3 =0'31 n l + 0'3211z + 0'33~
X2

(1.16)

For a range of two on all indices, the indicial equation


Au

=BipCi/Jpq

(1.17)

represents the four equations


Au =BuCuDu +BUC1P12 + B12CuD2l +B12C1PZZ'
A12

=BUC21Du + BUCzP12 +B12C21D2l + B12CzPZZ'

A2l =B21CuD U+B21C1P12+BzzCuD2l +BZZC1Pzz,

A zz =B21C2PU + B21CzP12 + BZZC21D2l + BzzCzPzz

(1.18)

1.4 ALGEBRA OF CARTESIAN TENSORS

x;

Let Xl X2 ~ and ~ x~ represent two right-handed rectangular Cartesian coordinate systems with a common origin at 0 as shown in Fig. 1.6. If the symbol au
denotes the cosine of the angle between i th unprimed and j th primed coordinate axes,
i.e. aij =COS(Xi, x;) , the relative orientation of individual axes of each system with respect
to the other is conveniently given by the table inset into Fig. 1.6. An arbitrary vector
v in the same figure has vector components Vi; i =1,2,3 in Xl' X2 and ~ directions
respectively. Alternatively, the same vector can be considered as composed of vector
components v;; i =1,2,3 in x;, ~ and x~ directions respectively. From the definition of au and in view of Eqs. 1.10, each component in the primed system can be
obtained from those in the unprimed system as

v2=
V3=a

VI =aUv l +tlzlV2+~lV3'
a 12v l +tlzzV2+~2V3'
13v l

+ tZz3v2+ ~3V3

(1.19)

or, in indicial notation, as


(1.20)
By interchanging the roles of the primed and unprimed vectors in the above develop-

Foundations of Solid Mechanics

x'
X'
I

x'
2

x'
3

xI

a"

0 12

0 13

x2

0 21

0 22

0 23

x3

0 31

0 32

0 33

x2
I

-I

cos

~=--+--~X2

x'I
Fig. 1.6 Transfonnation for ftrst order tensor.
ment, the inverse of Eq. 1.20 is found to be
(1.21)
It is important to note that in Eq. 1.20 the free index on aij appears as the second
index; in Eq. 1.21 the free index appears as the ftrst.
Substitution of Eq. 1.20 into Eq. 1.21 with an appropriate choice of dummy indices
yields

=aijal;jvk
1.22 must give the identity Vj =Vj, the coefficient

(1.22)

Vj

Since Eq.
the Kronecker delta

Bib

aijal;j must be identical to

defined by
I

Bit =Bkj ={0,'


i.e.
aijal;j

fori =k
fori .k

(1.23)

=Bit

(1.24)

In expanded form, Eq. 1.24 consists of nine equations which are known as the ortho
gonality or orthonormality conditions on the direction cosines aij.
On the other hand, if Eq. 1.21 is substituted into Eq. 1.20 to produce =afiajkV;,
then orthogonality conditions appear in an alternative fonn afiajk =Bjk , thus
aijal;j = afiajk = Bit
(1.25)

v;

Equations 1.20 and 1.21 are called the transformation laws for first order tensors.
According to Eq. 1.20, a dyad UjVj has components in the primed coordinate
system given by
u;v; = (apju p) (aqjvq) = apjaqjupvq

(1.26)

In an obvious generalization of Eq. 1.26, any second order Cartesian tensor Tij obeys
the transformation law

Mathematical Foundations

T~ =apjaqjT pq

(1.27)

With the help of the orthogonality conditions (Eq. 1.25) it is a simple calculation to
invert Eq. 1.27, thereby obtaining the transfonnation rule from primed components to
unprimed components as
(1.28)
Tij =aipajqT~
The transfonnation can be generalized for Nth order tensor as

T~" ... =apiaqja", ... T pqr...

(1.29)

Cartesian tensors of the same order may be added (or subtracted) component by
component in accordance with the rule
Aij"... Bij"... =Tij"".
(1.30)
The sum is a tensor of the same order as those added.
Multiplication of every component of a tensor by a given scalar produces a new
tensor of the same order. For the scalar multiplier A., typical examples written in both
indicial and symbolic notations are
b=Aa;
or

B = A.A.
The outer product of two tensors of arbitrary order is the tensor fonned by simply
setting down the factor tensors in juxtaposition, for examples:
(a) ajb j =Tij'
(c) DijTbn =(f)ijkm'
Bij = }.Aij'

or

(b) vjFjk flu".


(d) eq"VIII aijbn'
This process produces a tensor having an order which is the sum of the orders of the
factor tensors.
The inner product of two tensors of first order is the result of a contraction,
involving one index from each tensor, perfonned on the outer product of the two tensors.
Such a scalar product written in both indicia! and symbolic notations is
(1.31)
ajbj =8' b
The cross product 8 x b is expressed in indicial notation by

Eu"ajb" =8 X b

where

(1.32)

eq" is the permutation tensor, i.e.

I; if i,j,k permute like, 1,2,3, etc.


ifi,j,kpermutelike,3,2,I,etc.
0; if two or more of the indices, i, j, k have the same value.

Eu,,= { -1;

(1.33)

1.5 MATRICES AND DETERMINANTS


A rectangular array of elements, enclosed by square brackets and subject to certain

Foundations 0/ Solid Mechanics

10

laws of combination is called a matrix. An M xN matrix is one having M (horizontal)


rows and N (vertical) columns of elements. In the symbol A jj, used to represent the
typical element of a matrix, the flfStsubscript denotes the row, the second subscript the
column occupied by the element. The matrix itself is designated by enclosing the
elements in square brackets [ ]. For example, a 3 x 3 matrix [Au"! is the array given
by
Au AI2 AI~
[AiiJ =[ A21 Azz ~
A31 A32 A3
A matrix for which M N is called a square matrix. A 1 x N matrix is called
a row matrix and is written by enclosing the elements in parentheses ( ) as follows
(Vi) (VI v2 ... vN)
(1.34)

while a column matrix is an M x 1 matrix and is written by enclosing the elements in


braces { }, i.e.

(1.35)

{V;l=

Thus every tensor of order two or less can be represented by a matrix; a tensor of
second order by a square matrix, a vector by a row or a column matrix, and a scalar
by a single term. A matrix having only zeros as elements is called the zero matrix. A
square matrix with zeros everywhere except on the main diagonal (Au, An> A 33 , .... , AMM )
is called a diagonal matrix. If the nonzero elements of a diagonal matrix are all unity,
the matrix is called the unity or identity matrix. The N xM matrix [AiiJT, formed by
interchanging rows and columns of the M xN matrix [Au"!, is called the transpose
matrix of [AiiJ. A square matrix [AiiJ is called symmetric if [Au"!T =[Au"!.
Matrices having the same number of rows and columns may be added (or subtracted) element by element. Multiplication of the matrix [Avl by a scalar A. results in
the matrix [AAvl. The product of two matrices is defmed only if the matrices are
conformable, i.e. if the pre/actor matrix has the same number of columns as the post/actor matrix has rows. The product of an M x P matrix multiplied into a P x N
matrix is an M xN matrix.
Matrix multiplication is usually denoted by simply setting down the matrix symbols
in juxtaposition as follows:
(1.36)
[AiiJ [B.;J =[C;J
in which the element Clk =AvBik and the summation convention applies from 1 to P

Mathematical Foundations

11

for the repeating index j. It follows from the deflnition of matrix multiplication, but
is left to the reader to prove, that the transpose of a product of matrices equals the
product of the transposed matrices taken in the reverse order, e.g.
[[AuJ [BjJ [CtJ]T = [CtJT[BjJT[AuJT

(1.37)

Matrix multiplication is not commutative in general. In three-dimensional Euclidean


space, the transformation of Cartesian vectors and second order tensors in Eqs. 1.20,
1.21, 1.27 and 1.28 can be rewritten respectively in matrix form as follows;
{Vi}

=[ajJ

{V)

(1.38)

{vJ

= [auJ {v;J

(1.39)

= [apJ T [Tpq] [aJ


= [ai;! [T;J [aJ T

(1.40)

,T

[T~
[TuJ

(1.41)

The determinant of a square matrix [AuJ is normally denoted by the symbol IAijl
or det[AijI , and can be determined by a standard method, which involves the concept
of cofactors and minors. The cofactor of an element Aij of a square matrix [AijI,
denoted by
is deflned by
A: = (-li+jMij
(1.42)

A;,

in which Mij is the minor of Aij, i.e. the determinant of the square array remaining
after the row and column of Aij are deleted. The value of a determinant is then defmed
as sum of the products of the elements in any row (or column) by their corresponding
cofactors, i.e. Laplace expansion of the form
N

det[AuJ

=j~1 AvA: '

(i is optional and not summed)

(1.43a)

=i=l
LAvA:, Uis optional and not summed)

(1.43b)

where N is the number of rows (and columns as well) of the matrix [AuJ.
Here are some important theorems on determinants.
Theorem 1.1. The value of a determinant remains the same if rows and columns
are interchanged. In symbols,
det [AuJ

=det[AuJ T

(1.44)

Theorem 1.2. An interchange of any two rows (or columns) changes the sign of
the determinant.
Theorem 1.3. If any two rows (or columns) are the same or proportional, the
determinant is zero.
Theorem 1.4. If [AijI and [BijI are square matrices of the same order, then
det [[AuJ [BjJ] = det [AuJ det [BuJ

(1.45)

Foundations of Solid Mechanics

12

A square matrix [Ag] whose determinant is zero is called a singular matrix.


The determinant of a 3 x 3 matrix [Tg] can be put in indicial notation in various
ways, as follows
(1.46)
(1.47)
(1.48)

(1.49)
(1.50)

=(,Ejjl,E/mIITi/TjmTkn

(1.51)

where Vk is the permutation tensor in three dimensional space as defined in Eq. 1.33.
The adjoint matrix of [Ag] is obtained by replacing each element by its cofactor
and then interchanging rows and columns. If a square matrix [Ag] is non-singular, it
possesses a unique inverse matrix [Ag]-l which is defined as the adjoint matrix of [Ag]
divided by the determinant of [Ag]. Thus
[A)-l = [A;J
I Alii
From the inverse matrix definition (Eq. 1.52), it may be shown that
[Ag]-l [Ag] = [A~ [A~-l =!

(1.52)

(1.53)

where ! is the identity matrix.


1.6 LINEAR EQUATIONS AND EIGENVALUE PROBLEMS
Let [crg] be an n x n matrix, and {xa and {b;} column matrices each with n
elements. A set of equations having the form
[cr~ {Xj} ={b;}
(1.54a)
or, in indicial notation,
(1.54b)
is called a system of n linear equations in the n unknowns Xl' Xl' ... , X". If b l , bl , ... , b"
are all zero the system is called homogeneous. If they are not all zero it is called
nonhomogeneous. Any set of XbXl, ... ,x", which satisfies Eqs. 1.54 is called a solution
of the system and may be put in the form
{x;} = [cr~-l {bj}

(1.55)

Mathematical Foundations

13

More explicitly, such a solution can be put according to the Cramer's rule as follows
{}

Xi

{L\}

(1.56a)

L\

(1.56b)

det[O'iJ

or, in indicial notation,


x=-I
det[O'iJ

where L\, i =1,2 ...... ,n is the determinant obtained from [O'iJ by removing the i th
column and replacing it by the column matrix {b;}. The following four cases can
arise.
Case 1, det [O'iJ 0, {b;} {O}. In this case there will be a unique solution where
not all Xi will be zero.
Case 2, det[O'iJ O, {b;} ={O}. In this case the only solution is the trivial
solution, i.e. {x;} ={OJ .
Case 3, det [O'iJ =0, {bi} ={O}. In this case there will be infmitely many solution
other than the trivial solution. This means that at least one of the equations can be
obtained from the others, i.e. the equations are linearly dependent. To be discussed
more later is this case which is normally called an eigenvalue problem.
Case 4, det[O'iJ =0, {b;} {OJ. In this case infmitely many solutions will exists
if and only if all L\ in Eqs. 1.56 vanish. Otherwise there will be no solution.
Next, consider a homogeneous system of n equations of the form
[O'vl {x) -A.{x;} =0
(1.57a)
or, in indicial notation,
(1.57b)
where A. is a number, and Bii denotes the Kronecker symbol as defmed in Eq. 1.23,
or in matrix form [BiJ is an identity matrix, i.e.
100
0
010
o
(1.58)
[Bvl = .

As discussed previously, the system of linear equations, Eq. 1.57a or b will have
non-trivial solutions if and only if
det[O'ii - Mvl =0
(1.59)
which is a polynomial equation of degree n in A.. The problem characterized by Eqs.
1.57 is normally called an eigenvalue problem. The polynomial equation in A., Eq.
1.59, is the characteristic equation, and its roots are eigenvalues or characteristic values
of the matrix [O'iJ . Corresponding to each eigenvalue there will be a solution
{Xi} {O}, i.e. a non-trivial solution, which is called an eigenvector or characteristic

Foundations of SoUd Mechanics

14

vector. Writing Eq. 1.57a for the case where the eigenvalue is equal to

~,

we have

[avl {xjl>.l =A.I:{xr>.l ,

(k not summed)

(1.60a)

aifXjl)=A.~r),

(k not summed)

(1.60b)

or, in indicial notation,


Note that the equations in Eq. 1.60a or b are linearly dependent, thus x~I:),x~I:), .. ,x!l:)
cannot be obtained explicitly but rather as known multiples of one of them, say x,(I:)
which must be non-zero.
It is proved in standard texts of mathematics (e.g. Spiegel, 1971) that a symmetric
real matrix has real eigenvalues and orthogonal eigenvectors. The latter property for
a symmetric real [avl can be put as

xjl>X}')=0,

(k

I)

(1.61)

xJ")

into a unit eigen-

(k not summed)

(1.62a)

(k not summed)

(1.62b)

It will be found useful to transform each of these eigenvectors


vector
by the following normalization

{w

or, in matrix form,

Since

Xa:Xa:=I,

(k not summed)

(1.63)

we fmd, on combining the last equation with Eq. 1.61, that we can write the ortho-

gonality condition for the unit eigenvectors in the form


XuXa: = 81:1

(1.64a)

or, in matrix form,


(k =/)

=0,
(k l)
In terms of these unit eigenvectors, Eq. 1.60b can be put in the form

ai/Xjl = A.I:Xa:,
(k not summed)
Multiplying the last equation by 'Xu, and using Eq. 1.64a, we fmd
aijXu'ljl = A.I:81A: ,

(k not summed )

(1.64b)
(1.64c)
(1.65)
(1.66)

or, in matrix form,

=0,

(k=/)

(1.67a)

(k /)

(1.67b)

15

Mathematical Foundations

Equation 1.66 can be rewritten in matrix fonn as


~ 0 0

o Az

o
o
(1.68)

Here [X;J is an n x n matrix with unit eigenvectors assembled as columns in the


matrix, i.e. M denotes the value of Xi for the eigenvalue At. Equation 1.68 shows
how the symmetric real matrix [oJ may be transformed into the diagonal matrix constructed from the eigenvalues of [oJ. Thus introducing the linear transformation
{xi}

=[xiJ {YJ

or

(1.69a)

='ljkYt

(1.69b)
into Eq. 1.54a, premultiplying the result by [Lf, and using Eq. 1.66 or 1.68, we fmd
Xi

y,=

t(u}T {bJ

A,

(I not summed)

(1.70a)

(I not summed)

(1.70b)

or, in indicial form,


Y,

b{Xu

=-A, .

This enables us to say that the system of linear equations of Eqs. 1.54 may be uncoupled
by the transformation defmed in Eq. 1.69. For this reason, {yJ are frequently referred
to as the normal coordinates, to distinguish them from the generalized coordinates Xi'
Next, consider the quadratic form of generalized coordinates
V

={x;lT [oJ {xi}

(1.71 a)
(1.71b)

The transformation defined in Eq. 1.69 will readily change this quadratic form into a
new form with no cross product tenns, i.e.

=A1Y: + Azy: + ... + AllY:

(1.72)

which is called the canonical form. A symmetric real matrix [oJ and its quadratic
form as defined in Eqs. 1.71 are said to be positive definite if V> 0 for all real
{xJ {OJ. If [oJ is positive defmite, then the tenns on the main diagonal of [oJ
must all be positive; for if one, say 22 were negative or zero, then V would be
negative or zero when Xz is the only non-zero coordinate. This condition is insufficient

Foundations of Solid Mechanics

16

to ensure the positive defmiteness. A necessary and sufficient condition is that (Frazer,
Duncan and Collar, 1938)
(1.73)
Alternatively, the canonical fonn defmed in Eq. 1.72 gives another necessary and
sufficient condition for this purpose, i.e. the positive defmiteness is ensured provided
that all eigenvalues are positive.
Before the advent of digital computers, the solution to the eigenvalue problem for
a square matrix with a large dimension (n) had to be done manually using an appropriate iterative scheme. The use of digital computers and appropriate softwares for this
purpose has greatly simplified the task.
For n 2, the characteristic equation, Eq. 1.59, becomes a quadratic equation of
the fonn

(1.74)
the roots of which are
A,1'

When n

~ =~{crll + cr22 [(crll-cr22i+4~ll2}

(1.75a,b)

=3, the characteristic equation is

A,3 - I1A,2 + 12A, - 13 =0

(1.76)

where
(1.77a)
(1.77b)

13 =det[crijl

(1.77c)

New symbols may be introduced as follows

1
3

a =-(/1 -3/J

(1.78a)

1
3
b = 27 (211 - 9112 + 27/3 )

(1.78b)

and we may show that 27b2-4a3~0. If 27b 2-4a 3=0, Eq. 1.76 will have three real
roots, of which two at least are equal, i.e.
I
A,1 = (4b )1/3 +..!
(1.79a)

Ib)l/3

~=~=1.2

+iI

(1.79b,c)

Mathematical Foundations

17

On the other hand, if 27b z - 4a 3 < 0, there will be three real and unequal roots, i.e.

+i

I
A.I = 2( 3a )112cos(cp)
3

(1.80a)

(a J'2cosex+cp)
II
Az=23
- 3 - +3

(1.80b)

(a JI2 cos(4X+CP)
II
- 3 - +3

(1.80c)

~=2 3

where
3-{3b
coscp=-20312

(1.81)

1.7 THEOREMS ON TENSOR FIELDS


A tensor field assigns a tensor T(x,t) to every pair (x,t) where the position
vector x( == xJ varies over a particular region of space and t varies over a particular
interval of time. The tensor field is said to be a continuous (or differentiable) function
of Xi and t if the components of the field are continuous (or differentiable) functions
of Xi and t. If the components are functions of Xi only, the tensor field is said to be

steady.
With respect to a rectangular Cartesian coordinate system, for which the position
Xi' tensor fields of various orders are represented in
indicial and symbolic notations as follows:
(a) scalar field; cP = CP(Xi' t) or cP = cp(x, t).
(b) vector field; Vi = Vi(Xj, t) or v = v(x, t).
(c) second order tensor field; Tij = Tij(xk' t) or T = T(x, t).
Note the following identity

vector of an arbitrary point is

a( ) a( )ax;
alXi alXj' alXi '

--=----

x;

where
is the position vector with reference to another rectangular coordinate system
(Fig. 1.6). Applying the transformation law, Eq. 1.20, to
we find that the equation
above becomes

x;,

(1.82)
Since XI' X2 and X3 are independent variables, a derivative of one of them with respect
to another must take the form
(1.83)

Foundations of Solid Mechanics

18

where Bjk is the Kronecker symbol defined by Eq. 1.23. Thus Eq. 1.82 becomes
a() a()
-a-=aij-a'
!Xj

(1.84)

!Xj

Comparing this equation with Eq. 1.21, which is the transformation law for a first-order
tensor, we see that the operator a/aXj is transformed in the same way a first-order
tensor. For this reason we represent such differentiation by a comma-subscript as
illustrated by the following examples:

=<I>.j.

(a)

:.

(b)

-a
=Vj j'
!Xj

(c)

ax. =v

(d)

(Wj

(Wj

j j

We note that the operator a/aXj produces a tensor of one order higher if i remains a
free index, and a tensor of one order lower if i becomes a dummy index in the
derivative.
Differential operators which appear often in continuum mechanics are given here
for reference.
gradej>= Vej>
or q,'i'
divv=V v

or

curl v = Vxv

or EijkV".j'

Laplacianej> = V2ej> = V . Vej>

Viii'

or ej>,"

(1.85)

Gauss's divergence theorem relates a volume integral to a surface integral. In its


traditional form the theorem says that for the vector field v = v(x) ,

Iv divvdV = Is n . vdS

(1.86)

where n is the outward unit normal to the bounding surface S , of the volume V in
which the vector field is defined. In the indicial notation, Eq. 1.86 is written as

Iv vj,jdV =Is vjnjdS

(1.87)

For an arbitrary tensor field T jjk the theorem is

rT.." ...n dS
JvrT..",,.PdV =Js"
p

(1.88)

Green's theorem in the plane. If R is a closed region of the X 1X2 plane bounded
by a simple closed curve (a curve which does not intersect itself anywhere) C and if

Mathematical Foundations

19

M and N are continuous functions of


then

Xl

and

X2

having continuous derivative in R ,


(1.89)

This theorem also holds for a multiply-connected region, i.e. region bounded by a fInite
number of simple closed curves which do not intersect.
1.8 DIFFERENTIAL GEOMETRY

Differential geometry is the study of space curves and surfaces. If in ordinary


physical space a position vector Xi is a function of a single scalar variable u, i.e.
~=~~t

~=~~t

~=~~t

the terminal point of Xi describes a space curve as u changes. As shown in Fig. 1.7,
if

x.= x. (u+au)-x.{u)
II

Fig. 1.7 Space curve.

Ax dx

Lim-'=-'
Au --+ 0Llu
du
exists, the limit will be a vector in the direction of the tangent to the space curve at
point (XI,X2,X3) and will have Cartesian components (dxl/du, dx:!du, dxidu). If u
is the time t, dx/dt represents the velocity Vi with which the terminal of vector Xi
describes the curve. Similarly, dv/dt =d2x/dt 2 represents its acceleration along the
curve.
If the scalar u is taken as the arc length s measured from some fIxed point on
the space curve, then dx/ds is a unit tangent vector to the curve and denoted by Ti
(see Fig. 1.8). If Ni and Bi are the unit principal normal and binormal vectors to the

Foundations of Solid Mechanics

20

/ - - - - - - - - - - - X2

Fig. 1.8 Tangent and nonnal unit vectors to a space curve.


curve respectively, the three unit vectors T j , N j and B j are mutually orthogonal such
that
(1.90)
B j =EijkTjNk
Other relations among them are called Frenet-Serret formUlas, i.e.
dTj
dB j
(1.91)
cis =KNj , di=-rNj
and
dNj
-=tB-KT.
cis
I
I

(1.92)

where K and 't denote the curvature and torsion respectively, and the reciprocal of
these, p =K- I and 0" ='t-I are called the radius of curvature and radius of torsion
respectively.
The parametric equations which ensure that point (XI> x z, X3) is on the straight line
joining points (Yl> Yz, Y3) and (ZI> Zz, Z3) are
XI - YI XZ- Yz X3 - Y3
(1.93)
--=--=-See the illustration in Fig. 1.9 (a) in which numbers 1, 2, 3 denote the coordinate
axes. In Fig. 1.9 (b), another position vector Xj which is equal to the vector running
from (YI' Yz, Y3) to (ZI> zz, Z3) is illustrated, thus
(1.94)
Xj =Zj-yj
The equation of the plane passing through three points (ai, Uz, ~),
("(I' "(Z, "(3), which are not on the same straight line, has the fonn

(13z, 13z, ~) and

21

Mathematical Foundations

(a) End of

Xi

on the straight line joining

end of Yi with end of

(b) Position vector

Xi

=Zi -

Yi .

Zi'

Fig. 1.9 Equations of a straight line.

-F---~.

Fig. 1.10 A plane passing through tenninal points of <X;,


ax 1 + bX 2 + cx] + 1

=0

Pi

and "(;.
(1.95)

in which a, b and c are constants to be detennined from the condition that all three
prescribed points be on the plane. This condition gives the equation of the plane in
the fonn

Foundations of Solid Mechanics

22

Fig. 1.11 Plane perpendicular to

Yi

and through tenninal point of

xl-a l

~-~

X3-~

(31-~

~-~

(33-~ =0

Zi'

(1.96)

'Yl-al 'Y2-~ 'Y3-~


See the illustration in Fig. 1.10. On the other hand, an equation for the plane perpendicular to a vector Yi and passing through the tenninal point of another position vector
Zi (see Fig. 1.11) is
or
(1.97)
or
XIYI

+ X2Y2 + X3Y3 =YIZI + Y2Z2 + Y3 Z3'

Another way of writing the equation of a plane is

xini=d

(1.98)

where ni is the unit vector nonnal to the plane and d is the shortest distance from
the origin 0 to the plane.
On a surface ~(Xi) constant, its total differential at a position defmed by the
vector Xi is

d4l

~ =-dx
OXi

This shows that the vector

~.i

'

=~,"dx =O.

is perpendicular to the vector dxi and therefore to the

Mathematical Foundations

23

surface. In other words, the gradient of a surface is a normal vector to the surface.
Moreover, a unit vector nonnal to the surface <I>(xJ = constant can be obtained by the
usual nonnalization

+ <I>.i
ni=- -V<l>A.j

(1.99)

The sign in Eq. 1.99 is often chosen so that the vector ni is directed 'outward' from
the region on one particular side of the surface <I>(Xi) = constant. Often we can express
the position vector of a point on a surface in the fonn

Xi =Xi(a,~)
(1.100)
are two independent parameters of the sUrface. Noting that ax/aa

where a and ~
and ax/a~ are vectors lying in the tangent plane to the surface, we see that the direction
of the nonnal to the surface is the same as the following non-zero vector,
aXjaXk
Eijk

aa a~

':.

(1.101)

1.9 DIRAC-DELTA AND HEAVISIDE STEP FUNCTIONS


It has been found advantageous in some problems, such as those involving concentrated forces and impulsive forces, to introduce the Dirac-delta /unction, B(t - a),
which is singular but integrable as follows
B(t-a)=oo, (t=a)
(1.102a)

=0,

Ie B(t-a)dt= 1,

a)

(1.102b)

(b <a <c)

(1.102c)

(t

':.

--

8U-o)

I
/1-- 2

Fig. 1.12 Graphical representation of Dirac-delta function.

Foundations of Solid Mechanics

24

If I>(t) is a continuous function pf t in the neighborhood of t =a, then

Ie

I>(t)S(t -a)dt

=I>(a),

(b < a < c)

(1.103)

The Dirac-delta function is not an ordinary function, but rather a limit of a special
sequence of ordinary functions (Stakgold, 1967). In Fig. 1.12 we show the Dirac-delta
function as the limit of a pulse. In characterizing a concentrated force by l)(t - a), t is
a spatial variable. On other occasions, when l)(t - a) characterizes an impulsive force,
t is the time variable.
Since all time functions involved are normally in the state of complete quietude
before the time t is reckoned as zero, it is appropriate to study the Heaviside step
function which has the properties
(1.104a)
(t >0)
H(t) 1,

(1.104b)
=0,
(t <0)
Graphically, we may put H(t-a) as shown in Fig. 1.13.
To study the derivatives of H(t) and l)(t), let us consider the following integral

Ie I>(t)dH~-a)

where b < a < c and I>(t) is continuous at t


fmd

Ie I>(t)dH~-a)

dt

=a.

dt,
Performing integration by parts, we

=I>(c)(I)-I>(b)(O)=I>(c)-

Ie

Ie H(t-a)~t)

dt

d!>(t)

=I>(c)-I>(c) + I>(a)
=I>(a)

(1.105)

Comparing Eq. 1.105 with Eq. 1.103 leads to the 'in effect' identity as follows
dH(t-a) S(t-a)
dt
Another 'in effect' identity induced from Eqs. 1.102 and 1.103 is

(1.106)

I>(t)S(t - a) I>(a )l)(t - a)


For a function I>(t) which is differentiable, we can show that

(1.107)

Ie

I>(t)S(l)(t -a)dt

=[-t!l(l)(t)],=a'

(b<a<c)

(1.108)

in which a superscript in parentheses denotes a derivative with respect to t. As a more


general identity, we can show by mathematical induction for a function I>(t) which is
differentiable n times that

25

Mathematical Foundations

H (t-o)

Fig. 1.13 Graphical representation of Heaviside step function.

La c!>(t)8(")(t -a)dt

=[(-I)"c!>(")(t)],=a'

(b<a<c)

(1.109)

More 'in effect' identities should be noted here, i.e.


t"8(")(t) = (-I)"n(n -1) (n - 2) ... (1)8(t)

(m >n)

d
dt [c!>(t)H(t)] =c!>(0)8(t) + c!>(I)(t)H(t)

~ [c!>(t)H(t)] ="fc!>(,,-I-i)(0)8(i)(t) + c!>(")(t)H(t)


dt"

i=O

(1.11Oa)
(1.11Ob)
(1.11Oc)
(1.11Od)

1.10 BESSEL FUNCfIONS


We present a brief review of Bessel functions. Further details may be found in
the books by Abramowitz and Stegun (1964), and Watson (1966). The Bessel equation
is the following
(1.111)

In general, z is a complex variable, and m is a complex constant. Solutions to the


equation above are the Bessel functions of the first kind Jm(z), of the second kind (also
called Weber's function) Y",(z), and of the third kind (also called Hankel functions)
H2)(z), H~)(z). Each is a regular (analytic or holomorphic) function of z throughout
the z-plane cut along the negative real axis, and for fIxed z (: 0) each is an entire

Foundations of Solid Mechanics

26

(integral) function of m. When m is a positive or negative integer, Jm(z) has no


branch point at (z = 0) and is an entire function of z. See Churchill (1960) for the
defmitions of the teons analytic, entire, branch point.
Listed below are recurrence formulas,
2m
z

Cm_1(Z)+Cm+1(z) =-Cm(z)
dCm(z)

Cm_l(Z)-Cm+l(Z)=2~

dCm(z)

m
=Cm_1(z)--;C m(z)

dCm(z)

~ =-Cm+1(z)+-;Cm(z)

(1.112a)
(1.112b)
(1.112c)
(1.112d)

in which C denotes J, Y, H(l>, H\}.> or any linear combination of these functions,


the coefficients in which are independent of Z and m. These functions of the same
order m are related as the following

+ iYm(z)

(1.113a)

H!:>(z) =Jm(z)-iYm(z)

(1.113b)

H2>(z) = Jm(z)

(i2 =-1).

where i is the imaginary number


To follow are analytic continuation foonulas,
J m(zelllfC) = elmllfCJm(z)
Ym(ze lllfC) = e-lmIIfCym(z)

+ 2isin(mmt)cot(m1t)Jm(z)

sin(m1t)H2>(ze lllfC) = -sin[(n -1)m1t]H2>(z) - e-lmfCsin(mn1t)H!:>(z)


sin(m1t)H!:>(ze lll") = sin[(n

+ 1)m1t]H!:>(z) + elmfCsin(mn1t)H2>(z)

H2>(ze~

=-e -imtcH!:>(z)

H!:>(ze~ =-elmfCH2>(z)

(1.114a)
(1.114b)
(1.114c)
(1.114d)
(1.114e)
(1.1140

In Eqs. 1.114a to d, m is an integer.


When m is fixed and I z I~ 0,
(z/2)m
Jm(z) == r(m + 1)'

(m

negative integer)

Yo(z) == -iHJ1>(Z) == iHf>(z) == ~ logz


1t

(1.115a)
(1.115b)

Mathematical Foundations

27

[9t(m) > 0]

(1.115c)

In these equations, the symbol == means 'is approximately equal to'. On the other
hand, if m is fIxed and I Z I~ 00,
Jm(z)

=(:Z JTco{z - ~1t -~ )+eIS(Z~ 0(1 Z 1-1)].

(I argz 1< 1t)

(1.116a)

(~ JTsin(z - ~1t -~ )+eIS(Z~ 0(1 z rl)].

(I argz 1< 1t)

(1. 116b)

Ym(z) =

H~)(z) == [2/(1tz)] 1I2e(Z-T-~) ,

(-1t < argz < 21t)

(1.116c)

(-21t < argz < 1t)

(1.116d)

In Eqs. 1.115 and 1.116, f'(z) is a gamma function (discussed in the next section for
real and non-negative z), 9t(z) and 5(z) are real and imaginary parts of z respectively, argz =tan- 1[5(z)/9t(z)] , and O( ) denotes the order of the truncation error.
Closely related to Bessel functions are modified Bessel functions, Kelvinfunctions,
and spherical Bessel functions. For the modifIed Bessel functions, the pertinent relationship is
(-1t < argz

=e3il1rl2J

Km(z)

i;

(ze- 3itr12 ),

~ 1tI2)

(1tI2 < argz

(1.117a)

~ 1t)

(1.117b)

= eimtrl2H~)(zei1rl2),

(-1t < argz

~ 1tI2)

(1.118a)

__ i1t -imtrl2H (2)( -itrI2)


- 2e
m ze
,

(-1tI2 < argz < 1t)

(1. 118b)

The Kelvin functions, of a real and non-negative argument z and a real order m,
can be expressed in terms of Bessel functions as
berm(z) + ibeim(z) =J m(ze 3i1r14 )
(1.119a)

=~ H~)(ze3i1r14)
Spherical Bessel functions are defIned for m =0, 1, 2, ...
jm(z) =[1tI(2Z)]1/2J J(z)
m+i
ker m(z) + ikeim(z)

(1.119b)
by the equations
(1.120a)
(1.120b)
(1.120c,d)

Foundations of Solid Mechanics

28

(1. 120e,t)
Here j"., y". and h". are spherical Bessel functions of the first kind, second kind, and
third kind, respectively. In addition, h2) and h~) are also known as spherical Hankel

functions.
With the help of the recurrence fannulas (Eqs. 1.112) and the asymptotic fann
when m is fixed and Iz I~ 00 (Eqs. 1.116), we can derive the following
C".(z )lzl: == 0

(1.121 a)
(1.121b)
(1.121c)

where k ~ I, and C denotes J, Y, H(l) or H(J.), provided the value of argz falls
within the respective ranges stipulated in Eqs. 1.116. Due to Eqs. 1.120, the same
conditions also hold for j, y, h(l) or h(J.).
Computation of these special functions has been simplified since the establishment
of their polynomial approximations. Such approximations are presented in the handbook
by Abramowitz and Stegun (1964).
1.11 LAPLACE TRANSFORMS
If the function v(t) is defmed for the time interval 0 < t < 00, then its Laplace
transform L{v(t)} or v(P) is defined as

L{v(t)} =v(P)=

L-

v (t)e-P'dt

(1.122)

Note that the lower limit in the integral above is not a plane 0 but rather 0-, which
is a quantity infmitesimally less than O. This is used to accommodate the state of
complete quietude of v(t) before the time t is reckoned as zero, i.e.
v (0-) =v(I)(Ol =v(J.)(Ol =.... =0
(1.123)
in which a superscript n in parentheses denotes an n th derivative with respect to t.
Laplace transforms of some elementary functions are listed below:

1
L{H(t)} =-

(1.124a)

L{B(t)} = 1

(1. 124b)

L{B(II)(t)}

=p"

L{e"'} = (p -art,

(1.124c)

(p-a>O)

(1.124d)

29

Mathematical Foundations

L{sinat} =al(pz+a z)

(1.124e)

=pl(pz +a z)

(1.1241)

L{cosat}

L{sinhat} =al(pz_a z),

=pl(pz_a z),
L{t}=r(a + l)lp a+l,

L{coshat}

(p>(lal)

(1.124g)

(p>(lal)

(1.124h)

(a> 0)

(1.124i)

In the last equation above r(y) denotes a gamma junction, i.e.


r(y) = i~ xY-1e-Xdx,

(y > 0)

(1.125)

Accordingly, it can be shown that


(1.126)
r(y + 1) = yr(y), y 0,-1,-2,-3, ....
If Y is equal to a positive integer n,
r(n) = (n -I)!
(1.127a)
where n! denotes the factorial n, i.e.
n! =n(n -l)(n -2) ... (1)
(1.127b)
Note that
11 =O! = 1
(1.128)
Here are some important theorems on Laplace transforms:
Theorem 1.5. Linearity property. If C1 and Cz are any constants then

L{C1V1(t)+CZvz(t)} =C1V1(P)+CZvz(P)

(1.129)

Theorem 1.6. First translation or shifting property. If L{ v(t)} = v(P) then


L{ealv(t)} =v(p -a)

(1.130)

Theorem 1.7. Second translation or shifting property. If L{v(t)} =v(P) and


f(t) =v (t - a)H (t - a), where H (t) denotes the Heaviside step function, then
L{f(t)} =e-pav(p)

(1.131)

Theorem 1.8. Laplace transform of derivatives. If L{v(t)} =v(P), then

L{-dllV}
=pv
dt"

11-

(1.132)

in which the state of complete quietude (Eqs. 1.123) has been incorporated.
Theorem 1.9. Laplace transform of integrals. If L{v(t)} = v(P), then

L{L v(S)ds} = V~)

(1.133)

Theorem 1.10. Laplace transform of convolution integrals. If L{v(t)} =v(P)


and a convolution integral is symbolized by

30

Foundations of Solid Mechanics

then
(1.135)
Since one often wants to detennine the derivative of a convolution integral, it is
appropriate to record here that
d
Ov2(t-s)
(1.136)
dtJO_ VI(S)V2(t-S)ds= JO_VI(S)
at
ds+VI(t)viO~

r t rt

where 0+ denotes a quantity infInitesimally greater than zero. A more general fonnula
known as Leibnitz's rule for differentiating an integral is the following
d
dt

ib(t) f(t,s)ds =ib(t):\af(t,s)ds + f(t, b) dbdt - f(t, a) dadt


a(t)

a(t)

ot

(1.137)

Note that such convolution integrals obey the commutative law, i.e.
VI *v2

=v2*v1

(1.138a)

the associative law, i.e.


(1.138b)
and the distributive law, i.e.
V1*(V2 +v3 )

=VI*V2 +VI*V3

(1. 138c)

Theorem 1.11. Lerch's uniqueness theorem. If v(t) is sectionally continuous in


every fInite interval 0 < t < N and of exponential order for t > N, then the inverse
Laplace transform of v(t), i.e. L-1{ v(P)} = v(t) is unique.
Theorem 1.12. Initial-value theorem. If the indicated limits exist, then

Limv(t) = Limpv(P)
t~O+

p-+-

(1.139)

Theorem 1.13. Final-value theorem. If the indicated limits exist, then

Lim v(t) = Lim pv(P)


1-+00

p-+o+

(1.140)

Theorem 1.14. Fourier transform. For a function v(t) defined for the interval
0< t < 00, its Fourier transfonn F{v(t)} or v.(ro) is defined as
F{v(t)} =v.(ro)= i~v(t)e-ioltdt

(1.141)

A sufficient condition for the existence of v.(ro) is that the function v(t) must be
absolutely integrable, i.e.

i~' v(t)1 dt <

00

(1.142)

31

Mathematical Foundations

Comparing Eq. 1.141 with Eq. 1.122, we get


v.(eo) = v(ieo)

(1.143)

It should be noted that Eq. 1.142 is a sufficient but not necessary condition for
the existence of v.(eo). Functions which do not satisfy Eq. 1.142 may have Fourier
transforms. Here are examples of such functions (Hsu, 1970);
~
1
F{H(t)} = l H(t)e-'OlIdt =:-+1tO(eo)
(1.144a)

0-

leo

(1.144b)
Note that the second term on the right-hand side of each of Eqs. 1.144 is an addition
to that readily given by Eq. 1.143. In fact, Eq. 1.144b can be obtained by differentiating
Eq. 1.144a with respect to eo.
Equating the real parts in Eq. 1.144a gives
o(eo) = -11~ coseotdt
1t 0
while doing the same for the imaginary parts gives

(1.145)

1 l~ sineotdt
-=

eo

(1.146)

Taking the inverse cosine transform of Eq. 1.145 leads to


1 = 21~ o(eo)coseotdeo,

(t > 0),

or
r~
1
)0 O(eo)coseotdeo=Z

(1.147)

On the other hand, the inverse sine transform of Eq. 1.146 gives
1 = ~ r~ sineot deo,
1t)0

(t > 0),

eo

or
1 ll~sineotd
H()
t =-+- - eo
2

1t 0

eo

(1.148)

or
1t
--deo=-sgnt
o eo
2
where sgn t (read as signum t) is defined as
sgn t = 1, (t > 0)

(1.150a)

< 0)

(1.150b)

~sineot

=-1,

(t

(1.149)

Foundations of Solid Mechanics

32

1.12 INVERSE LAPLACE TRANSFORMS


For a given Laplace transform v(P), its inverse symbolized by

L-1{v(P)} =v(t)

(1.151)

may be found from standard mathematical textbooks, e.g. by Spiegel (1965). Otherwise,
one may use Bromwich's integral formula, i.e.

1.
v(t)=-2

Y+ 1-

1tJ y-Ioo

(1.152)

ePtv(p)dp

The integration in the equation above is to be performed in the complex plane where
p = q + ir along a line q = 'Y. The real number 'Y is chosen so that q 'Y lies to the
right of all singularities (poles, branch points, or essential singularities) but is otherwise
arbitrary. The formula provides direct means for calculating v(t) for a given v(P).
When the methods mentioned above are not possible or practical, approximate
methods may be adopted, but their accuracy and efficiency must be checked or tested
case by case. Cost (1964) started his investigation on some approximate methods with
the n th derivative with respect to p of the Laplace transform, Eq. 1.122, i.e.

d"v(P)
- = (-1)"
dp"

or
(-1)

"p,,+ld"v(p)
---=
nl dp"

1-

v (t)t"e-ptdt,

0-

10-

v(t)

[p"+lt"e-pt J
dt

nl

(1.153)

Considering the term in the brackets on the right-hand side of the equation above, i.e.
p,,+lt "e-pt

nl
we note that it has a single peak at t =nip, and due to the definitions of the gamma
functions and factorials, Eqs. 1.125 and 1.127, we may write
P "+l1t"e-ptdt
nl

0-

=1

(1.154)

Moreover, the Stirling's formula, (Spiegel, 1971) which says that


n 1== V21te-"nlJ+ll2

(1.155)

implies the following


(1.156a)

0 , (t ~ nip)
(1.156b)
Comparing these results with the properties of the Dirac-delta function given in Eqs.
1.102, we can write

33

Mathematical Foundations
. pn+1 tne-PI

Lun

n!

n~~

n)

=8 t--

(1.157)

Substituting Eq. 1.157 into Eq. 1.153, in view of Eq. 1.103, leads to
. {(_1ypIl+1dny(p)]

v(t)=Lun

n.
dp" p="/I
which is known as the Widder's general inversion formula (Widder, 1946).
Alfrey's approximation (Alfrey, 1944). Putting n = 1 in Eq. 1.158 yields
n~

V(t)==[_p2d~(P)J

P P=II
Haar's approximation (ter Haar, 1951). Rewrite Eq. 1.122 in the fonn
r~V(t)[pm+1tme-PI]dt
Jo- t m
ml

pm+1v(p)

m!

where m is a positive integer, and put n =m in Eq. 1.157 to get


pm+1 t me-PI {

m)

t--

ml
p
Substituting the equation above into Eq. 1.160 yields
pm+ly(p)
,

m.

(1.158)

(1.159)

(1.160)

(1.161)

[V(t)]
t m I=mlp '

or
(1.162)
For m = 1 , we get
(1.163)

Schapery's approximation. In 1962, R.A. Schapery proposed the following


approximate fonnula
v(t) == py(P)1 P=O.SiI

(1.164)

Collocation method. Suppose the given Laplace transfonn has an approximate


inverse of the fonn
~I

v(t)==v (t)= L gje


j=l

(1.165)

where N is a positive integer and aj(i = 1 to N) are positive constants, gj are unknown
constants. The values of N and aj are specified basing on the nature of the problem
involved, while gj are to be detennined by the techniques of the minimum square error,
i.e. from the condition

Foundations of Solid Mechanics

34

a Loo[vet) - v*(t)] dt =0,

:\
ugj

or

or, due to Eq. 1.165,

0-

L0-oo[vet) -

L
OO

0-

av*(t)

v*(t)]-:l-dt =0,
ugj

[v(t)-v*(t)]e~ dt =0,

or
(1.166)
Substituting Eq. 1.165 again into the right-hand side of the equation above, while noting
its left-hand side is V(<lj) , we can write
N
g.
v(<lj) :E _J_,
(i 1,2, ... , N)
(1.167)

j=l~+aj

Thus we have N equations in the fann as above to determine gj(i 1 to N).


The approximation methods presented above have been found quite successful for
quasi-static problems. Beside the collocation method, all of them automatically give
exact initial and final values, since Eqs. 1.139 and 1.140 are satisfied exactly. The same
favorable characteristics can be achieved in the collocation method also by a proper
selection of v*(t) in Eq. 1.165 and assigning appropriate constraint conditions among
gj before carrying out the procedure of the minimum square error. For problems of
wave propagation in elastic solids, Swanson (1980) proposed to use the numerical
inversion techniques developed by Bellman, Kalaba and Lockett (1966).
1.13 ONE-TO-ONE MAPPINGS
A set of three independent variables Xl' X2' X3 may be thought of as specifying
the coordinates of a point in a three dimensional space. A transfannation from Xl' X2 ,
X3 to a set of new variables Yl' Y2' Y3 through the equations
Yj yj(Xl'Xz,~) , (i 1,2,3)
(1.168a)

specifies a transformation of coordinates, and may be thought as specifying the coordinates of a point in another three dimensional space. The points specified by Yj and
Xj may not belong to the same space, and if they do belong to the same space they
may not define the same point. The inverse transfannation
Xj

=Xj(Yl'Y2'Y3)

(1.168b)

proceeds in the reverse direction. By the standard chain rule of differentiation, we can

35

Mathematical Foundations

derive the following identities


dA(Yi) dA dYk
--=--,
dXj
dYk dXj

dB (Xi) dB dXk
--=-dYj
dXk dYj

(1.l68c,d)

dYi dXk = 5.. = dXi dYk


(1.l68e,f)
dXk dYj " dYk dXj
When an infmitesimal quantity dxi in the domain of the variables Xi is uniquely
mapped into dYi in the domain of the variables Yi we can write
dYi
dy.=-dx.
(1.l69a)
, dXj J
or, reversely

dXi
dx.=-dy.
, dYj J

(1.l69b)

Each of the last two equations can be rewritten as a system of three linear equations
(Sections 1.6) as the following

Yi
dX']
= [' {dy}
{dyJ = [ d
dXj ] {dx), {dx}
,
dYj
J

(1.1 69c,d)

where

dYl
dXl
dYz
dXl
dY3
dXl
dXl

dYl
dXz
dYz
dXz
dY3
dXz
dXl

dYl
dXz
dYl
dX3

dYz dY3
dXz dXz
dYz dY3
dX3 dX3
dYz dY

dYl

dYl
dX3
dYz
dX3
dY3
dX

(1.l70a)

dXl
(1.170b)

(1.l70c)
Thus in order to insure that the transformation is reversible and in one-to-one
correspondence in a certain domain V of the variables Xi' i.e. in order that each Xi
defines a unique Yi for Xi in V and vice versa, it is sufficient that :
(a) The functions Yi are single-valued, continuous, and possess continuous first

Foundations of Solid Mechanics

36

partial derivatives in V, and


(b) The Jacobian determinant J =loy/ox) does not vanish at any point in V.
Naturally, since the equations of the inverse transformation, Eq. 1.168b, are to be
uniquely solvable for Xi' the Jacobian determinant of the inverse transformation, i.e.
Iox/oYjl , must also be different from zero throughout V.
Coordinate transformation with the properties (a) and (b) named above are called
admissible transformations. If the Jacobian is positive everywhere, then a right-hand set
of coordinates is transformed into another right-hand set, and the transformation is said
to be proper. If the Jacobian is negative everywhere, a right-hand set of coordinates is
transformed into a left-hand one, and the transformation is said to be improper. In this
book, we shall tacitly assume admissible and proper transformations. Note that the
transformation principle described above can be readily generalized for a set of any
number of independent variables.
1.14 CURVll..INEAR COORDINATES
Suppose that any point pO in a three dimensional space (see Fig. 1.14) can be
located by two different sets of coordinates, the usual rectangular Cartesian coordinates
(xt, X2, X3) and the curvilinear coordinates (at, a2, e3), so that

Xl =xl(et, e2, e\

e e3),

x2=x2(e l , 2,
X3

=x 3(et, e2, e\

or
(1.171)
Curve

8 2 Curve

Fig. 1.14 Coordinate surfaces and coordinate curves.

Mathematical Foundations

37

Fig. 1.15 Vector dx with components in directions of three coordinate curves.


and, in reverse,

Oi =Oi(X 1,X2,X3)

(1.172)

The sets of equations above (Eqs. 1.171 and 1.172) defme a transformation of coordinates, when the functions involved make the transfoonation admissible and proper.
In practice, this assumption may not apply at certain points and special consideration
is required.
In Fig. 1.14, the surfaces 01 = Cl> ~ = C2 and 03 = C3 where Cl' C2 and C3 are
constants, are called coordinate surfaces. The intersections of these surfaces are known
as coordinate curves or lines. Ifa neighboring point QO is located by the position vector
x + dx , the vector dx drawn from po to QO must have a mathematical foon as follows

ax lax 2 ax 3

dx=oOl dO +o02 dO +o03 dO ,


or
dx=g~dO~

where

(1.173)

is known as the covariant base vector, i.e.

ax

g~=~

(1.174)

Thus, as described in Section 1.8, ~ is a vector tangent to the O~ curve at point pO.
Accordingly, an infmitesimal parallelepiped with pOQo as a diagonal is constructed and

Foundations of Solid Mechanics

38

shown in Fig. 1.15.


At this stage, we may notice that some notation rules are different from those
already used in rectangular Cartesian coordinate systems. We now have to show the
difference between covariance and contravariance by using subscripts for covariant
tensors and superscript for contravariant tensors. Equating and summing two covariant
indices or two contravariant indices do not yield a tensor of lower order and not a proper
contraction. An equation does not have a correct tensor form if having a free index as
a subscript on one side but as a superscript on the other side. However, the Einstein
summation convention is still in use, while vectors and tensors are written in symbolic
(Gibbs) notation together with indexed scripts. An example of this point is Eq. 1.173
where the notation is essentially symbolic but with the added feature of the summation
convention. In such a 'combination' style of notations, which seems to be the most
efficient way when dealing with curvilinear coordinates, tensor character is not given
by the free index rule prevailing in a true indicial notation.
The position vector x can be put explicitly in terms of three Cartesian components
as follows

x=xlf+X2j+X3k

(1.175)

where f, j and k are unit base vectors of axes Xl, x and x respectively, and
have the product properties given in Eqs. 1.7 and 1.8. Accordingly, Eq. 1.174 becomes
2

dX I " dX2~ dX 3
gA = d9~ 1+ d9~J + a9~ k

3,

(1.176)

In general, gl' g2 and g3 are not mutually orthogonal nor unit vectors. When they are
mutually orthogonal, they form an orthogonal curvilinear coordinate system. The length
ds of the infmitesimalline element pOQo in Fig. 1.15 is given by

(ds)2 =dx. dx =dxidxi

or, due to Eq. 1.173,


(dsl=gAjl.d9~d91l

(1.177)

where gAjl. is the covariant metric tensor and defined as

gAjl. =g~. gil

(1.178)

The covariant metric tensor is a symmetric tensor, i.e.


gAjl.=g~

(1.179)

Substituting Eq. 1.176 into Eq. 1.178 yields

dXidX i

gAjl. = d9~a91l

(1.180)

or, in matrix form,


(1.181)
where 'Xij denotes the ith Cartesian components of gj (given in Eq. 1.176), i.e.

39

Mathematical Foundations

e1

Curve

Fig. 1.16 Infinitesimal tetrahedron fonned by three line elements


along coordinate curves.
(1.182)
Moreover, we can show that

g" . (g" x gj1) = 0,

( A not summed)

{i = gl . (g2 x g3) = g2 . (g3 x gl) = g3 . (gl X gz)

(1.183a)
(1.183b)
(1.183c)

=det[X~

and

g = det[g~ = det[XkiX,J

(1.183d)

In a general N dimensional space, a curvilinear coordinate system involved must


have N coordinates (a1, a2, ... , eN) , and an infmitesimallineelement with a length ds
is given by the metric form

(dsi=gAj1da"daj1,

(A,Il=1,2, ... ,N)

(1.184)

If there exists a transfonnation into a rectangular coordinate system (xi,x 2, .... x N ) such
that the metric fonn becomes

(ds)2=dx i dx i (i = 1.2 ..... N)


(1.185)
the space is called an N dimensional Euclidean space, otherwise the space is Rie-

mannian.
In three dimensional Euclidean space like the ordinary physical space. the volume

40

Foundations of Solid Mechanics

of the infInitesimal parallelepiped in Fig. 1.15 is given by


dVo =
de 1dezde3

Vi

(1.186)
where g is as defIned in Eqs. 1.183b to 1.183d. Further, passing a plane through corners
RO, SO and TO of the same parallelepiped, we obtain an inflnitesimal tetrahedron as
shown in Fig. 1.16, and the following relation can be noted
2n:,dS~ = (gzdez- glde1) x (g3de3 - glde1)
(1.187)

where dS~ is the area of the triangle ROSoro and n:, is the unit vector nonnal to dS~.
Perfonning a dot product of Eq. 1.187 with gl' in view of Eqs. 1.183a and b, yields
(1.188a)
2dS~gl . n:, =
dezd93

"i

and, in a similar manner,

2dS~gz . n:, = "ide3de 1

(1.188b)

2dS~g3 . n:, =

(1.188c)

"ide1dez

At this stage, it is appropriate to introduce two very important contravariant tensors.


Firstly, the contravariant metric tensor g'-!L may be introduced as follows
g'-!L

=g).p.

(1.189)

where g~ is, as defmed in Eq. 1.42, the cofactor of an element g'-!L of the square
matrix [g~. Thus it can be shown that
g'-!L=g"'(1.190a)
g'AlCg1q1 = a~

(1.190b)

where a~ is the Kronecker symbol. Secondly, the contravariant base vector


be introduced as follows

may

(1.191)

t=g'-!Lg"

There are interesting identities, namely


g'-!L=ig"

(1. 192a)

g'A =g'-!Lg"

(1.192b)

g,,' i=g'A g" =a~

(1. 192c)

is orthogonal to
The last equation means physically that a contravariant base vector
the coordinate surface eA = constant. Using this property for A. = 1 , we can obtain the
area of the side pOsoro of the infmitesimal tetrahedron in Fig. 1.16 as

dS~ =~glgl' (gzx g3)dezde3,

41

Mathematical Foundations

where gl is the magnitude of gl. Due to Eqs. 1.183b and 1.192a, the equation above
can be written in the form

dS~ = !-Vggll de2de3

2
Likewise, the area of any side, where e)" = constant, of the same tetrahedron is

dS;=~-VggUdeJ.ldex.

(A not summed, A:;f: Jl:;f: lC)

(1.193)

1.15 DERIVATIVES WITH RESPECf TO CURVILINEAR COORDINATES


Henceforth, curvilinear coordinates will be indexed by Greek scripts and a partial
differentiation with respect to e)" will follow the comma-subscript convention. The
derivative of the covariant base vector gJ.l with respect to e v is again a vector, and can
be put in the form
(1.194)
where

{~}

is the Christoffel symbol of the second kind relating the component of gJ.l,v

in the direction of g),..

Recalling the definition of gJ.l in Eq. 1.174, we can write


gJ.l,v = 1v,J.I
(1.195)

thus
(1.196)
In addition, the Christoffel symbol can be expressed in terms of metric tensors as

{JlvA} ='21g ),.1<(g"",v+gKY,J.I-gJ.lV,1<)

(1.197)

and the derivative of the contravariant base vector gJ.l with respect to e v can be obtained
as
(1.198)
A vector u can be resolved into three vectors in the directions of g),. as followsy
),.

u = v g),.

(1.199)

where v),. is called the contravariant component of the vector u. Differentiating u


with respect to e v and using Eq. 1.194 lead to
),.

u, v = v;v g),.

(1.200)

where v;~ is the covariant derivative of the contravariant component v),. and given by

Foundations of Solid Mechanics

42
V,A = VA

.v

.v

+{pV

(1.201)

I.}VP

On the other hand, the same vector u can be resolved into three vectors in the directions
of gA as follows
U=V~A

where

VA

is the covariant component of

(1.202)

and related to

VA

by the relation

=g"-fLv lL
1.202 with respect to e yields

(1.203)

VA

Differentiating Eq.

(1.204)

u,v = V).;vgA

where

v).;v

is the covariant derivative of the covariant component

v).;v=VA.,V-{~Jvp

VA

and given by
(1.205)

For a general tensor T~::,~;, its covariant derivative can be put as

T~"'Ar =T~"'Ar +
IL\.. ,IL,;V

IL\ .. ,IL,.V

T~"'~-\P~+\"'Ar {Ai}
pV

i=l IL\........................ IL,

-i= T~ ......................
j

Ar{
1 IL\.. ,ILj-\PILj+\.. ,IL,

p}

Ilj V

(1.206)

Two applications of the last equation are illustrated in Eqs. 1.201 and 1.205. Another
is that the covariant derivatives of the covariant and contravariant metric tensors vanish,
i.e.,

gAjL;v =0,

gt: =0

Moreover, the covariant derivative of a product of two tensors


formula,

(1.207)

S;:::T;":: follows a simple


(1.208)

The covariant derivative of Eq. 1.203, due to Eqs. 1.207 and 1.208 takes the form
v).;v

=g"-fLv;~

(1.209)

For an orthogonal curvilinear coordinate system, gA and gA are in the same


direction, and the following conditions hold:
g"-fL;;;; g"-fL = 0,

(I. Il)

(1.21Oa)
(1.21Ob)
(1.21Oc)

Mathematical Foundations

43

(1.21Od)

(1.210e)

{:J

=0,

(1.210)

In Eqs. 1.210, a repetition of an index script does not imply a summation.


A rectangular Cartesian coordinate system may be viewed as a special case of an
orthogonal curvilinear system. In this special case, " and gA become an identical unit
vector, i.e.

(1.211a)
(1.211b)
Consequently, the contravariant and covariant metric tensors become identical and equal
to Kronecker symbol, g defmed in Eqs. 1.183b to d becomes unity, and the Christoffel
symbol of the second kind given in Eq. 1.197 vanishes, i.e.
g=g=~

=1

(1.212a)
(1.212b)
(1.212c)

Moreover, the contravariant and covariant components of a vector such as in Eqs. 1.199
and 1.202 become identical, and a covariant partial differentiation such as in Eqs. 1.201,
1.205 and 1.206 is reduced into a usual partial differentiation, i.e.
A

=VA

(1.213a)

AA
V;v= v,v

(1.213b)

v~v = v).,v

(1.213c)

r'>-t A,

1LJIl,:v

A,
=r'>-t
1l1Il"v

(1 213d)

Strictly speaking, equations such as Eqs. 1.211 a, 1.212a and 1.213a are inadmissible,
since the free index scripts are not of the same type.
A special type of curvilinear coordinate systems is a system of skew coordinates,
in which every coordinate (SA) has a unit of length and has a ftxed orientation with
respect to the global (Xi) coordinate system. Hence, in such a system: a covariant base
vector gIL is independent of any skew coordinate SA; a Christoffel symbol of the

Foundations of Solid Mechanics

44

second kind due to Eq. 1.194 vanishes as put in Eq. 1.212c; a covariant partial dif
ferentiation as in Eqs. 1.201, 1.205 and 1.206 is reduced into a usual partial differentiation as in Eqs. 1.213; and we can put each and every Cartesian coordinate as a linear
function of the skew coordinates as follows
Xi

=XijEY + ci

(1.214)

ci

where
is a constant Cartesian component of the position vector of the origin of the
skew coordinate system, and 'Xij is the cosine of the angle between axes Xi and EY
Thus
XqXij = 1,
(j not summed)
(1.215)
In fact, 'Xij is as defined in Eq. 1.182 for any curvilinear coordinate systems.
1.16 EXERCISE PROBLEMS
1.16.1

Prove the equivalence between Gibbs and indicial notations as in Eqs. 1.31,
1.32 and 1.85.

1.16.2

Prove the identities that involve the permutation tensor and Kronecker delta
as
ijk

=~)j2~k3 + ~jl~k2~i3 + ~klBi2Bj3

-BjlBi2Bk3 - BklBj2Bi3 - BilBk2Bj3

~jki!m =Bjl~km - Bjm~k1


ijkk1j

(1.216a)
(1.216b)
(1.216c)

= 2Bi!

and Eqs. 1.46 to 1.51.


1.16.3

(a) If [CJ = [A.J [B~, where the ranges of indices i, j and k are 2, 3
and 4, respectively; find each and every element Cij in terms of elements
Ai,t and Big. (b) Write the set of simultaneous equations

3a1 - 2az+ 2a3 = 10,

4a1 +az+~ =3.

in matrix form, and in indicial notation. Find the determinant and the
inverse of the square matrix involved, and solve the equations for a 1 , az
and ~.
Answer: Determinant 35 ,/a1 2, az -3, ~ -1 .

45

Mathematical Foundations

1.16.4

Find eigenvalues and eigenvectors of the following matrix

[a~ =[~~ ~

Rewrite the generalized coordinates in tenns of the nonnal coordinates.


Show that [aijJ is positive deftnite by Eq. 1.72 and by Eq. 1.73.
Answer.

=16, ~,=4, ~ ~ r!1 =: rr\

~O
1.16.5

where

(a) Use Gauss's divergence theorem to $how that

Is xjnjdS =V8ij.

(b) Verify Green's theorem in the plane for


[(X1X2+X;)dx1+xJdxzl,

where C is the closed curve of the region bounded by


as shown in Fig. 1.17.

Xl

=Xz

and

X2

=xl

Fig. 1.17.
1.16.6

Find a unit nonnal and an equation for the tangent plane to the surface
2x1x:-3x1Xz-4xl=7 at the point 0,-1,2).
Answer: A unit nonnal (nl'nz,~) =(7,-3,8)/~122, and an equation for
the tangent plane is
7(x1 -l)-3(Xz+ 1)+8(x3 -2)=O.

1.16.7

Find an equation for the plane perpendicular to the position vector


(Y1> Y2' Y3) =(2,3,6) and passing through the tenninal point of the position
vector (Z1> Z2' Z3) = (1,5,3) . Find also the distance from the origin to the
plane.

Foundations 0/ Solid Mechanics

46

Answer: An equation for the plane is 2xl + 3.xz

+ &3 =35, and the plane

is 5 from the origin.


1.16.8

By means of Gauss's divergence theorem show that

Is nx(axx)dS =2aV,

where V is the volume enclosed by the surface S having the outward unit
normal n. The position vector to any point in V is x, and a is an
arbitrary constant vector.
Hint: Write the expression in indicial notation.
1.16.9

Find the tangent and normal vectors, and verify the Frenet-Serret formulas
of the following space curves: (a) a plane circle defined by Xl acose,
X2 =a sine, X] =b where a and b are constants while e varies. (b) a
cylindrical helix defmed by Xl acose, .xz a sine, X3 be.

CHAPTER II
STRESS AND STRAIN TENSORS

2.1 INTRODUCTION
The present book is restricted to small deformation of solid continua until Chapter
X, where the fmite defonnation theory will start. Indicial notation rules, using subscripts
to denote Cartesian components of tensors, will be followed until Section 10.6. Unless
specified otherwise, subscripts 1, 2 and 3 denote Cartesian components in Xl' ~ and
x3-directions respectively. Many textbooks denote Xl> X2' X3 by x, y, z respectively.
Figure 2.1 shows the equivalence between the notations.

Fig. 2.1 Various notations of Cartesian coordinate axes.


2.2 FORCE DISTRIBUTION AND STRESSES
One of the main concerns in the study of continuous media is how forces are
transmitted through the media. These forces can be classified into two classes, i.e. body
forces and sUrface tractions. A body force acts directly on the distribution of the matter
in the domain of interest, and thus has the dimensions of force per unit volume of
material. (In finite deformation theory, Chapter X, we will specify a body-force dis-

Foundations oj Solid Mechanics

48

o~

__ .....

Fig. 2.2 Cartesian components of stress tensor.


tribution per unit mass, since mass is the most basic characteristic of the material, and
must be conserved.)
For a continuum there may be some actual boundary that encloses the domain of
interest, or we may specify a domain of interest being enclosed by an artificial boundary.
There can be physical forces acting on an actual boundary, while forces on an artificial
boundary can be generated by the material immediately outside the domain of interest.
In either case, there will be a surface force distribution or a surface traction which has
the dimensions of force per unit area.
Consider an infmitesimal rectangular parallelepiped with its edges parallel to Xl'
X2 , X3 axes, as in Fig. 2.2. The surface traction on each rectangular boundary plane can
be decomposed into three Cartesian components. Each of these components can be
represented by the symbol (Jij (i,j = 1,2,3), where the first subscript denotes the
direction of the normal to the plane, while the second subscript denotes the direction
of the force. The stress component (Jij has the dimensions of force per unit area. As
an example, (J12 is the X2 -component of the stress tensor on a plane normal to the
Xl-direction. We will call the stress components with repeated numeral subscripts
normal stresses, since the force corresponding to each of these stresses is normal to the
surface upon which the force acts; and we call the remaining stress components shear

stresses.
A positive normal stress is directed outward from the material domain of interest,
i.e. a tensile stress. A shear stress is positive, if it is directed in a positive coordinate
direction when its associated normal stress (on the same rectangular boundary plane) is
also positive in another positive coordinate direction. A better understanding may be
obtained from Fig. 2.3, which is an illustration of the convention for a two-dimensional
case. Moreover it should be a good exercise for a reader to complete Fig. 2.2 with

49

Stress and Strain Tensors

positive stress components on the remaining three faces of the infmitesimal parallelepiped.
We may display the stress components in the matrix

[O"ijI=[::: :: ::1=[:: :: ~1
0"31

0"32

0"3J

O"u

O"zy

(2.1)

O"J

-Fig. 2.3 Positive stresses for two-dimensional case.


where the terms in the main diagonal are the normal stresses, while the off-diagonal
terms are the shear stresses.
2.3 STRESS VECTOR AND EQUATIONS OF MOTION
First consider the two-dimensional situation with Xl and X z as coordinates of the
frame of reference, as in Fig. 2.4. The plane AB is infinitesimally close to the point
P and is parallel to the x3-axis. Let nl and nz denote the direction cosines of the unit
vector nj normal to the plane AB, and let Xl and Xz be Cartesian components of the
surface traction across the plane AB. If 0"31 and 0"32 vanish, the equilibrium of forces
in the xl-direction on the element PAB requires
Xl(AB)-O"udxz-O"21dxl =0,
or
or
(2.2)

Foundations of Solid Mechanics

50

Fig. 2.4 Stress vector for a two-dimensional case.

Fig. 2.5 Two-dimensional neighborhood of point P in equilibrium.


Similarly. equilibrium in the xz-direction gives

Xz =a1Zn1+ azznz

(2.3)

Consequently. the surface traction Xi. being a vector but having the same dimensions
as a stress component ali' is called a stress vector.
Each stress component can be a function of position so that the same stress

Stress and Strain Tensors

51

components on any two opposing boundary plane need not be equal. Next, consider the
equilibrium of forces acting on an infmitesimally small neighborhood of the point
P(xI'X~ subjected to a body force Xi and surface tractions CJji as shown in Fig. 2.5.
The equilibrium of forces in Xl and x2 -directions yields, respectively, the following
aCJn

aCJ21

aXI

aX2

aCJI2

aCJzz

aXI

aXl

-+-+XI =0
-+-+X2 =0

(2.4a)
(2.4b)

On the other hand, the equilibrium of moments yields the stress symmetry, i.e.
CJ12

=CJlI

(2.5)

After understanding the two-dimensional case, we can easily obtain the corresponding three dimensional conditions at any point P located by (XI' Xl' x3). For the
stress vector, the illustration is depicted in Fig. 2.6 and the formulas can be shown to
be
XI

=CJnnl + CJ21nz+ CJ31 n 3,

Xl = CJllnl + CJzznz + CJ32~'


X3

=CJI3n l + CJ23nz + CJ33~'

or, in indicial notation,

Xi =CJjinj

(2.6)

where ni is the unit vector normal to the plane where Xi acts, thus

nini == n;+ni+n;= 1

(2.7)

Moreover, ni must be 'outward', i.e. pointing away from the material domain of interest.
For static equilibrium, the equation is
CJft,}
.. . +X., =0
(2.8)
and the symmetry of stresses is
~=~
~~
The last equation implies that the matrix [CJijI shown in Eq. 2.1 is a symmetric matrix,
i.e. [CJijI = [CJijlT, and that there are six (not nine) distinct stress components. For
dynamic equilibrium, Newton's second law of motion (1 Newton, 1687) gives, in place
of Eq. 2.8, the following equation

CJji,i +Xi =pa i


(2.10)
where Xi is the body force, Ui the displacement vector (see Section 2.8), p the mass
density (mass per unit volume), and a super dot (') a derivative with respect to the
time t. Equation 2.10 can be rewritten in the form of Eq. 2.8 if the reversed mass
acceleration density -pa j , the so-called D' Alembert' s force (J. Ie R. D' Alembert, 1743),
is added to the actual body force.

52

Foundations of Solid Mechanics

}--,
2

Fig 2.6 Stress vector for a three-dimensional case.


2.4 EULER'S LAWS OF MOTION
The laws that state that the time rate of the linear momentum is equal to the
applied force, and that the time rate of the moment of momentum is equal to the applied
torque, are known as Euler's first and second laws of motion, respectively (L. Euler,
1752). In fact, Euler's ftrst law of motion is identical to Newton's second law. For
an infmitesimally small domain of volume dV which is completely enclosed by a
surface dS, the particular law can be expressed as
d
XidV +XidS =dt (pvidV)
(2.11)
where

Vi

is the velocity, i.e. clx/dt. For small deformation, it can be shown that
aUi

at

v=I

(2.12)

and that the changes of p and dV with time are negligible. Thus Eq. 2.11 becomes
XidV +XidS

= puidV,

or, if considering Xi to include both actual and D' Alembert's forces,


XidV +XidS

=0

Substituting the condition on the surface S, Eq. 2.6, into Eq. 2.13 yields

(2.13)

53

Stress and Strain Tensors

XidV + ajinjdS =O.


Applying Gauss's divergence theorem, Eq. 1.88, to the second term in the equation
above, we get
XidV + ajiidV =0,
which holds for any dV at a particular point P(XttXz,x3). Thus the common factor dV
can be removed from the equation above, and the latter becomes
aJI.l
.. .+X=O
(2.14)
I

which is the equation of motion (Eq. 2.8) derived before. If the equation of motion is
used in Eq. 2.13, we shall obtain
--<JjiJdV +XidS =0,
or, due to Gauss's divergence theorem,
--<JjinjdS +XidS =0,

or

Xi =ajinj'
which is the formula of the stress vector (Eq. 2.6). Hence, we can see that for any
domain V enclosed completely by a surface S, the equation of motion for any point in
V can be derived from the condition of surface traction on S. and in the reverse the
condition of surface traction at any point on S can be derived from the equation of
motion in V.
Euler's second law of motion may be expressed as
-d
evtXj(XkdV +XkdS) =dt (evtXjPvkdV)

=evtXjpukdV

+ pevkvjvkdV.

However, the second term on the right-hand side of the equation above must vanish,
since evkVjVk denotes a cross product of two identical vectors. Thus, we have
evtXj(XkdV +XkdS)

=0

(2.15)

in which Xi includes both actual and D' Alembert's forces. In fact, Eq. 2.15 can be
obtained directly from Eq. 2.13 by taking the moment of the latter. Proceeding in the
same way as Eq. 2.14 has been obtained, one should arrive at the following
t;;k(XPIk),1 + evtX}Xk

=0,

or
t;;tXj(alk,l + Xk) + t;;tXj,lalk

=O.

The first term vanishes due to the equation of motion, Xj,l in the second term is Sjl, so
that

Foundations of Solid Mechanics

54

This gives for i =1,


for i =2,
and for i =3,

=(J21'

(J12

Thus the symmetry of stresses, Eq. 2.9, is proved once more.


2.5 STRESS TENSOR
At this stage, it is appropriate to demonstrate that
start with the stress vector formula

(J1i

is, in fact, a tensor. We

(Jjinj =Xj

The right-hand side is a vector, so that, upon applying the transformation law, Eq. 1.21,
we can rewrite the equation above as

=au){~,

(Jjinj

where a prime (') denotes a Cartesian component in the rectangular coordinate system
(x;, x;, x~). Applying the stress vector formula to
in the last equation yields

X;

(Jjinj

=ajJ:(Ju.nl'

Applying the transformation law (Eq. 1.20) for the vector


,

(Jjinj

But

nj

=ajJ:ajl(Ju.nj'

is an arbitrary unit vector, so that


(Jji

n; changes the equation into

=aj1ajJ:(Ju.

(2. 16a)

The inverse to Eq. 2.16a can be obtained by the same procedure, i.e.
,
~=~~~

~1~)

Thus we have shown that (J1i is a tensor of second order, since it obeys the transformation law for such a tensor, Eqs. 1.27 and 1.28. Note that the transformed stress is
symmetric as well, i.e. <iii =c.ip.
Letting i =j in Eq. 2.16a or 2.16b and using the orthogonality conditions, Eq.
1.25, we fmd
,

(Jii

(JkA;'

which shows that the scalar quantity (Jii does not change with the reference coordinate
system. Any term which has such property is normally referred to as an invariant. The
following are also stress invariants;
(J1i(J ji.

(J1i(Jjk(Jkj.

(J1i(Jjk(J/d(J/i,

etc.

55

Stress and Strain Tensors

The three quantities, 11' 12, 13 given in Eqs. 1.17, and known as the first, second and
third stress invariants, are combinations of these invariants; in fact

(2. 17a)
(2. 17b)

(2. 17c)

}-,
2

Fig. 2.7 A principal plane and corresponding principal stress.


Among all the planes passing through a point P, there can be one that has no
shear traction. Such a plane is called a principal plane, the normal traction on the plane
is known as a principal stress, and the direction of the outward normal vector is called
a principal direction or principal axis. In Fig. 2.7, ABC is a principal plane with nj as
the unit outward nonnal vector, and cr is the value (scalar) of the principal stress.
Following the stress vector formula, Eq. 2.6, we can write
crnj

=crjjnj'

or
(crjj - crajj)nj

=O.

Since the matrix [crvl is symmetric, the analysis of Section 1.6 shows that there are
three real values of cr, i.e. three principal stresses, denoted by cr1' cr2, cr3, for which this
equation has a non-trivial solution for nj. These values, the eigenvalues, are the roots
of the equation

Foundations of Solid Mechanics

56

det[crij - crBijI
or

=0,

cr -/ d'+/ cr-I = 0,
1

as given in Eq. 1.76. Stress invariants given by Eqs. 1.77 or Eqs. 2.17 can be expressed
in tenns of principal stresses as
(2.18a)
(2.18b)

13 =cr1cr2cr3

(2.18c)

If the eigenvalues are distinct, each eigenvalue defines a unique eigenvector nj'
and the three eigenvectors are mutually orthogonal. In other words, when all three
principal stresses are distinct, there is a unique triad of mutually orthogonal principal
planes, and a corresponding unique triad of mutually orthogonal principal axes.
On the other hand, if two principal stresses are equal, only the principal axis in
the direction of the distinct principal stress is unique. Using principal axes as reference
Cartesian coordinate axes we have

(2.19)
or, in matrix fonn,

0 OJ

cr1
[crijl =[ 0 cr2 0,
o 0 cr

where we suppose cr3=cr2 cr1. In the primed coordinate system, the stress components
are given by Eq. 2.16b, i.e.

cr~ =alia1pl + (OziOzj + ~i~)cr2


For

Xl

and x~ being identical, and

axes, the transfonnadon tensor

x;

and

x;

(2.20)

being any pair of mutually orthogonal

~; = [~ ~ ~1.
o

~2

aJ

Then, in view of the orthogonal property of aij' Eq. 1.25, we can show that Eq. 2.20
gives

so that the primed axes are also principal.

57

Stress and Strain Tensors

In the hydrostatic case, when all three principal stresses are equal, then, referred
to principal axes
(2.21)

or, in matrix form,

raJ

=[~ ~ ~.

For any other set of axes, Eq. 2.16b gives

cr~ = crBij.
Thus any three mutually orthogonal axes form a set of principal axes.
2.6 STATIONARY SHEAR STRESSES
Using principal axes as the reference Cartesian coordinate axes, we can write the
stress tensor crij in the form
crij =cr1BilBj1 + cr2Bj2Bj2 + cr3Bj3Bj3
(2.22)
or, in matrix form,

[crijl

cr1

=[ 0

0 OJ

cr2 0
0 cr

(2.23)

The stress vector on a plane with the unit normal vector nj, as shown in Fig. 2.8, is
given by Eq. 2.6, and its components normal to and tangential to the plane are
respectively
(2.24)

Sj=Xj-Nj

(2.25)

Substituting Eqs. 2.6 and 2.24 into Eq. 2.25, and using Eq. 2.22, we fmd

Sj =cr1Biln1+ cr2Bj2~ + cr3Bj3n3

- (cr1n12+ cr2~2+ cr3~2)nj


The square of the magnitude of the shear traction is

(2.26)

SjSj =(~n~ + ~n; + ~n;)


2
2
22
- (cr1n1 + cr2~ + cr3~)
The components of the unit vector nj satisfy

n;= 1-(n~+n;)

(2.27)
(2.28)

Foundations of Solid Mechanics

S8

)-,
2

Fig. 2.8 Principal axes as reference coordinate axes for plane


ABC with nonnal vector nj.

n:

Eliminating from Eq. 2.27, taking derivatives of the result SjSj with respect to nt and
liz, then equating these derivatives to zero; we obtain
[ (O't - 0'3)nt
(O't - 0'3)nz

(0'2- 0'3)nt]
(0'2 - 0'3)nz

In;} =.!{nt(O'll - 0'33)1


n;

(2.29)

2 nz(0'22 - 0'33)J

The direction nj are specified by nt and liz, which are governed by the equation above.
We seek nj for which SjSj is a maximum or minimum, or more generally, stationary.
One obvious solution to Eq. 2.29 is that nt = liz = 0, which is the case of a principal
stress 0'3' Another possible solution is that nt = 0, liz = 1I-../2 and ~ 1/-.J2. for which
Eq. 2.27 gives
(2.30)
and Eq. 2.24 gives
(2.31)
Investigating for all possible solutions, we obtain, as compiled in Table 2.1, the three
cases of principal stresses, and three cases of stationary shear stresses. Note that a
stationary shear stress occurs on a plane bisecting the angle between two principal
directions and has a value equal to half the difference between those two principal
stresses. In fact, each stationary shear stress may be obtained individually from a

59

Stress and Strain Tensors

Table 2.1

n1

1I...[2

1I...[2

rIz

1I...[2

1I...[2

n3

1I...[2

1I...[2

Stationary shear stresses


Shear
Stress

0'2- 0'3
-2

<J3- <J1
-2

<J1-<J2
-2

<J3

--

0'2+ 0'3
2

-3 2-1

<J +<J

--

Principal stresses
Normal
Stress

0'1

<J2

0'1 + 0'2
2

two-dimensional problem with two principal axes forming a plane of reference. As an


example, consider the fifth column in Table 2.1, and take axes l' and 2' as shown in
Fig. 2.9 and the 3'-axis along the I-axis. The equilibrium of forces acting on the element
PAB in l' and 2'-directions give, respectively, the normal stress 0'~1 and the stationary
shear stress 0'~2 the listed values in the fifth column, i.e. 0'~1 =(0'2 + 0'3)/2 and
~2 =(0'3 - <Jz}/2.
3

Fig. 2.9 A stationary shear stress treated as a two-dimensional problem.


2.7 OCfAHEDRAL SHEAR STRESS AND STRESS DEVIATOR
There can be eight planes making an equal angle with each principal direction as

60

Foundations of Solid Mechanics

shown in Fig. 2.10. Each such plane is called an octahedral plane, and has direction
cosines, with respective to the principal axes, given by
2
2
2 1
n1 = nz = ~ =(2.32)
3
Equation 2.27 shows that the magnitude of the shear stress on each octahedral plane,
called the octahedral shear stress 'to, is given by

9io = (0'1 -

O'i + (0'2 -

0'3i + (0'3 -

O'l

We shall now show that ~ is a stress invariant.


We introduce the stress deviator
1

sij =O'ij -'3O'kkOij

(2.33)

(2.34)

This equation, which may be rewritten as


1

O'ij =sij +'3O'kkOij

(2.35)

shows the stress tensor O'ij decomposed into a deviatoric stress tensor sij and a
hydrostatic or spherical stress tensor O'kkB/3. We note that O'kk is identical to the ftrst
stress invariant Ii' Equation 2.34 may be rewritten in matrix form

O'kk

0'-22

(2.36)

Note that sij is a symmetric second order tensor i.e.

sij =sji
(2.37)
Now the first, second and third deviatoric stress invariants are, respectively, deftned as
the following

(2.38)
(2.39)
(2,40)

While J 1 is identically zero, J 2 and J 3 can be expressed in terms of other stress


invariants as
(2,41 a)

Stress and Strain Tensors

61

Fig. 2.10 Octahedral planes.

(2.41b)
Thus we obtain the relationship between the octahedral shear stress and the second
deviatoric stress invariant in the fonn
(2.42)
In 1913, Richard von Mises proposed that yielding of some materials occurs at a constant
value of 1 l , then Nadai (1915) apparently because of Eq. 2.42 interpreted that the
yielding occurs at a constant value of the octahedral shear stress.
2.8 STRAIN TENSOR
As usual, let us begin with a two-dimensional case with Xl and Xl as reference
axes, and consider the situation that a point P located by (Xl> xz) at a reference time
moves to a point P' located by (Xl +U I , xl+uz) at any time t. In other words UI and
Uz are displacements of the point P in Xl and xl-directions respectively. Similarly,
other points Q and R infInitesimally close to P move to points Q' and R'
respectively as shown in Fig. 2.11.
From the figure, we may notice 'small defonnation' as follows; the change of the
line element PQ per unit original length is approximately aUl/aXI' the change of the

62

Foundations of Solid Mechanics

o---------;~ U I

+ ax
aUI

dX 2

Fig. 2.11 Displacement and defonnation in a two-dimensional case.


line element PR per unit original length is approximately dujd:XZ and the change of
the angle QPR is approximately dUl!dX2 + dujdXl Denote these changes by new
symbols as follows
(2.43a)
(2.43b)

(2.43c)

In their common names, Ell is the normal strain in Xl -direction, ~ the normal strain
in X2 -direction, and E12 the shear strain in X1X2 -plane. Note, however, the shear strain
defmed above is only half of the whole angle change. The latter (whole angle change)
is nonnally called an engineering shear strain.
For three-dimensional cases, strains can be put in terms of displacement components Ui as follows
1
F. .. =-(u. . +U .)
(2.44)
II 2 J,}
},I

and engineering shear strains

63

Stress and Strain Tensors

=U ,2 + Uz,1
123 =Uz,3 + ~,2
131 =~,1 + U1,3
112

Thus, we have symmetry in strains, i.e.

(2.4Sa)

(2.4Sb)
(2.4Sc)

eu =eji

(2.46a)

and
(2.46b)
which means that there are six (not nine) distinct strain components. It was shown in
Section 1.7 that differentiation of a tensor with respect to Xi yields tensor of one order
higher if the index i remains a free index. The quantity 9;, consisting of such a
derivative of the displacement vector, is therefore a tensor of second order. So all ideas
developed for CJij' e.g. stress invariants, principal stresses, etc., have counterparts for
eu ' i.e. strain invariants, principal strains, etc.
In particular, the fIrst strain invariant elk is called the dilatation since it represents
the proportional volume change of an infinitesimal element, and the strain deviator is
defmed by
(2.47)
The tensor defmed by

1
" 2

Cll =-(u..

',I

(2.48)

-u .. )
I,'

is the rotation tensor, since each component of it represents an average angle of rotation
about a coordinate axis. In Fig. 2.11, the line PQ rotates through an angle OUz/OX1
about the X3 -axis to become P'Q' and the line PR rotates through an angle -OuiChz
about the same axis to become P'R'. Hence the average rotation of the material body
framed by lines PQ and PR about the X3 -axis is (OUz/OX1 - ou/ihz)/2 =-IDt2' The
rotation tensor is antisymmetric, i.e.
Cllij = -C1)ji
(2.49)
In a matrix form,

[Cllul

has zero elements on its main diagonal, since


Cllll = ~ = 0>:33 = 0

(2.50)

Thus Cllij has only three distinct components corresponding to the angles of rotation
about Xl' Xz and X3 axes. The rotation vector Clli having these angles as its Cartesian
components can be written as
1
(2.S1a)
Clli =ZBu.tCllq
1

=-P".I;(u.t

4'"

.-u'.I;)

,I

J.

(2.S1b)

64

Foundations of Solid Mechanics

or, more explicitly,


001=

1
0>:32 ="2 (U:l,2 - Uz,3)
1

ffiz = 0013 ="2 (U I,3 - U:l,I)


1

0>:3 = ffizl ="2(Uz,1 - UI,Z>

(2.52a)
(2.52b)
(2.52c)

When there is displacement, but there is zero strain and constant rotation, the
material body is displaced as a rigid body. The Cartesian components of the rigid-body
displacement vector will be given by
(2.53)
Uj =Cj + ijJ1r k
where B j and Cj are arbitrary constants. Thus
so that

2.9 COMPATIBILITY CONDmONS


If we take the strain components as continuous functions of coordinates of material
points, the deformation will be single-valued and continuous only if the following
conditions are satisfied
ev,kI

+ EkI,ij -

Eik,jl - Ejl,ik

=0

(2.54)

which are known as compatibility conditions. The proof that these conditions are necessary can be made simply by differentiating Eq. 2.44 which is the definition of the
strain, but the proof of their sufficiency is more involved, as will be shown next,
following Cesaro (1906) and Sokolnikoff (1956).
Let pO(xf,x~,x~) be a point at which displacements Uj and ev are known. Then
the displacement of any other point pa in the domain can be obtained by a line integral
along a continuous curve C from pO to pa as follows

ut=u o+ J.pO

p.

du j

(2.55)

65

Stress and Strain Tensors

The derivative of the rotation tensor


CJ.) .. I;

v.

CJ.)ij'

defined in Eq. 2.48, with respect to

XI;

is

=-21 (u . .. - uh it)
I~

1
2

1
2

=-(u. ",-u ... )--(u ...,-u ... )

I....

~.v

hOI

"'"

=u:J - Ejl;.j

(2.56)
In the last integral in Eq. 2.55, we can replace dx j by d(xj - Xj4) , then integrate by
parts, to obtain

or, due to Eq. 2.56,

(2.57)

where
(2.58)
For singlevaluedness and continuity of Uj, the integral in Eq. 2.57 must be independent
of the integrating path C. This is possible only if Uitdx" is an exact differential, say
Uitdx,,=dVj

or, by expansion,
Then, since

Xl'

Uildxl + Uj'}.dx'}. + Uj3dx 3=Vj.ldx l + Vj.'}.~ + Vj.3~'

Xz and

X3

are independent variables, the last equation implies that


Uil = Vj l

or, in a compact form,


(2.59)

Uij=VjJ

Differentiating the equation above with respect to


Uij.1; =VjJI;

XI;

yields
(2.60)

Foundations of Solid Mechanics

66

but Vi,jk in the last equation is identical to Vi,li' hence we have


Uij,,,

=Uik,j

(2.61)

Substituting Eq. 2.58 into Eq. 2.61 yields


Eu.,1 - Biik,j - jk)

+ (xj" - x) (ik,jl -

jk,iI)

or
(X~-X,)(,<
'1-'kil-iI
,<+,/,<)=0.
J
J
"'./
J
.1"
J .'"

Since Xj" - Xj is in general not zero, its factor in the equation above must vanish leading
to the sufficiency (and necessity as well) of the compatibility conditions as defined in
Eq.2.54.
Actually, only six of the total eighty-one compatibility conditions are independent,
while the rest are identities or repetitions due to the symmetry of the strain tensor. These
six distinct conditions can be easily identified by contracting Eq. 2.54 with respect to
two indices, say k and 1, upon which we get more compact form of compatibility
conditions as

eu,a + a,ij - ik,jk - jk,ik =0,

which is symmetric with respect to the remaining two free indices i and j , Moreover,
writing the equation above explicitly as a set of six independent equations, we can say
that the six independent equations can be obtained from the equation above even with
k not summed. In conclusion, the most compact form for the compatibility conditions
is
ij,a

+ a,ij -

ik,jk - jk,ik

=0,

(k

i,

j, k not summed)

A more formal derivation of Eq. 2.62 from Eq. 2.54 can start by defming
absolute value of the left-hand side of Eq. 2.54, i.e.
RijkJ

The tensor

RijkJ

=Iij,kJ + kJ,ij -

RijkJ

Rijld

as the

ik,jl - jl,ikl.

has following symmetry properties


RijkJ =RkJij =RjikJ =Rijlk

Moreover,

(2.62)

=R ikjl =RIj/ci'

vanishes if more than two of its subscripts are equal, i.e.


RllWR2223'

== O.

Hence Eq. 2.54 is equivalent to the following distinct equations,

=0,
R 2311 =0,
RU22

=0,
R3122 =0,
R2233

=0,
R l233 =O.
R3311

which can be put collectively as Eq. 2.62. The tensor RijkJ is the
covariant curvature tensor to be described in Section 10.11.

Riemann-Christoffel

67

Stress and Strain Tensors

2.10 CYLINDRICAL AND SPHERICAL COORDINATES


The most frequently used curvilinear coordinate systems are the cylindrical
coordinate (r, e, z ) and the spherical coordinate (R, cI>, e) systems as shown in Fig. 2.12.
Both of them are orthogonal curvilinear coordinate systems. Although the governing
equations for these systems may be found using the general tensor analysis for orthogonal curvilinear systems, it is convenient to derive them in explicit (not indicial)
notation.

Fig. 2.12 Cartesian (x, y, z ), cylindrical (r, e, z) and spherical (R, cI>, e )
coordinates.
Cartesian and cylindrical coordinate systems are related by

x = r cos e,
which may be reverted to give

y = r sin e, z = z

r = +(x 2+ y2)112, e = tan-1l , z = z


x

(2.63a,b,c)
(2.64a,b,c)

The positive sign in front of the square root in Eq. 2.64a denotes that r always assumes
a non-negative value. The derivatives of cylindrical coordinates with respect to Cartesian
coordinates are

ar x
ax r

-=-=cose,

ae

y
r2

sine

ar =l= sine
ay r
x cose
ay r2 r
ae

-=--=---, - = - = - -

ax

(2.65a,b)
(2.65c,d)

Any derivative in Cartesian equations can be put in terms of cylindrical coordinates as


a
a ar a ae
a sin e a
- = - - + - - =cose---ax ar ax ae ax
ar r ae

(2.66a)

68

Foundations of Solid Mechanics

(2.66b)
Since r, 0 and z fonn a right-hand orthogonal system, the rules for transfonnation among
Cartesian tensor components and cylindrical tensor components described in Section 1.4
hold. As an example, the relationship among displacement components with reference
to these two different coordinate systems can be written as
{u;} =[aijl {u)

(2.67a)

or
(2.67b)
where

{xJ

{uJ

=t}. {xJ =t}

(2.68a,b)

={:}. ={j

(2.68c,d)

{uj

COSo

[aijl

= s~O
[

-sinO o~
cosO 0

(2.68e)

Figure 2.13 shows the decomposition of a vector u into alternative orthogonal components u, v and u" Us, thus reconflnning Eqs. 2.67. For tensors of second order, e.g.
the strain tensor, we have
y

~_---.J'---

______

~_

Fig. 2.13 Orthogonal components of a vector u in xy or rO plane.

69

Stress and Strain Tensors

(2.69)
where

[E~

E"

=[

(2.70)
Symm.

Substituting Eq. 2.44, in view of Eqs. 2.66 to 2.68 and Eqs. 2.70, into Eq. 2.69 yields
the strain-displacement relationship in cylindrical coordinates as

dU r
1 (dUe)
dU,
E"=a;:' fge=;:-l"r+d9 ' E.. =azEre

Ea. = .& =
.

![!
(du
1+ dUe]
2rd9
dr
!(!rd9dU, + dUe)
dZ

=Eer =

r _

Us

=e,. =!(dU
+ dUO)
2 dZ dr
r

(2.71a,b,c)
(2.71d)
(2.71e)
(2.711)

The dilatation, which is a strain invariant, can be written

1[

d
dUe] dU,
Ek.t=E"+fge+e..=;:- dr(rur )+ d9 +az-

(2.72)

Substituting the Cartesian rigid-body displacement components, Eq. 2.53, into Eq. 2.67b
leads to the cylindrical rigid-body displacement components
(2.73a)
Ur =C1 cos9+C2 sin9+ z(-B1 sin 9 +B2cos9)
Us=-C 1 sin9+C2cos9+B3r - z(B l cos 9 +B2 sin 9)

(2.73b)

u, =C3 +r(-B2cos9+B 1 sin 9)

(2.73c)

By similar procedures, the Cartesian equation of motion, Eq. 2.14, can be transformed to obtain those in r, 9 and z directions, respectively as
(2.74a)
(2.74b)
(2.74c)
in which Xr

X& and X, are cylindrical components of the body force, thus related to

Foundations of Solid Mechanics

70

the Cartesian components by


(2.75)
where

Xe x.l
Figure 2.14 depicts positive cylindrical components of the stress tensor.
Cylindrical and spherical coordinate systems are related by
{X;} =(Xr

(2.76)

r=Rsincl>, 9=9, z=Rcoscl>

(2.77a,b,c)

In reverse,
(2.78a)

cl>=tan-l~, 9=9
z
Note that R and cl> are non-negative, and the maximum value of cl> is

(2.78b,c)
1t.

. . . . . . - - - - - - - - - - - - ........ x

Fig. 2.14 Positive cylindrical components of stress tensor.


Spherical and cylindrical components of displacements and body forces can be related
by
(2.79a)
UR = ur sincl> + u.coscl>
(2.79b)

71

Stress and Strain Tensors

(2.79c)
(2.80a)
X, =X,cose!> -X. sine!>

(2.80b)

Xe=Xe

(2.8Oc)

Figure 2.15 shows the decomposition of a vector u into its cylindrical and spherical
components, thus reconftrming Eqs. 2.79 and 2.80. The strain-displacement relationship
can be written
(2.81a,b)
(2.81c)
(2.81d)

(a

1 Ue
au,)
e+& = ea. = 2R de!> - uecote!> + cosec e!> ae

(2.81e)
(2.81f)

The dilatation is
ekk = eRR + e.. + Eoo

1 a

=R2aR (R

cosece!>[ a
.
aUe]
uR)+-R- ae!> (u,sme!+ ae

o
R

o - direction

Fig. 2.15 Orthogonal components of a vector u in rz or Re!> plane.

(2.82)

Foundations of Solid Mechanics

72

The spherical rigid-body displacement components are

=sinell(C1 cos e+ C"sin e) + C3 cos ell


U. =cosell(C1 cos 9 + C" sin 9) - C3 sinell- B1R sin e+ B,.R cos e
Us =-C sin e+ C" cos e- R cos eIl(B1 cos e+ B" sin e) + B/l. sin ell

UR

(2. 83a)
(2.83b)
(2.83c)

where Bi and Ci are given in Eq. 2.53. The equations of equilibrium in R, ell and
e directions are, respectively

aaRR

1 OO'Re 100'R. 20'RR-O'ee-0',,-O'R.Cotell


oR + R sin ell Ta+li 0cIl +
R
+XR =0

Fig. 2.16 Positive spherical components of stress tensor.

(2.84a)

Stress and Strain Tensors

73

OO'R.

00'&+

100'.. 30'R++(0'.. -0'ee)cotcll

OO'Re

oO'ee

100'+& 30'Re+ 20'9+cotcll

~+R~cIl08+R~+

oR +Rsincll 08 +R ~ +

+~=O

+Xe=O

(2.84b)
(2.84c)

Figure 2.16 depicts positive spherical components of the stress tensor.


2.11 PROBLEMS

2.11.1 Stress or Strain Computation from Three Different In-Plane Normal Components
When the axes Xl and X2 are rotated about the X3 axis through an angle 8 to
become axes x~ and x~ as shown in Fig 2.17, the 'in-plane' strain components of the
primed coordinate system can be shown to be in terms of those of the nonprimed one
as
x'2

x'3

Fig. 2.17
28 12 + sm
. 28 f22
11 = cos28 Et1 + sm

(2.85a)
(2.85b)

.1
12 = '2 sin 28(f22 - 11) + cos 2812

(2.85c)

while the 'out-of-plane' components are


~3=~3

(2.85d)

~ = cos 8~ - sin 8~1

(2.85e)

e;1 = sin 8~ + cos 8~1

(2.850

Foundations of Solid Mechanics

74

The equations above hold also for stress components, i.e. if E is replaced by 0' in
those equations. In practice, it is not easy to measure a shear component directly. Instead,
normal strain components are measured in a plane in three different directions. Then
Eq. 2.85a is applied for each measured component to establish a set of three linear
equations for determining Eu, ~ and E12 , which are in-plane strain components
aligned with designated coordinate axes Xl and x2
2.11.2 Cylindrical and Spherical Rotation Components

Being mutually orthogonal components of a vector, cylindrical rotation components


can be obtained from Cartesian ones by means of the usual transformation formula for
vectors, i.e.
{roC} = [aijf {roJ

(2.86)

where [aiJ is the transformation matrix given in Eq. 2.68e, and


{roC} = (ro,

roe

ro.)T,

{roJ = (ro.,

ro)'

rool

(2.87a,b)

e.g. roe is the rotation angle about the e -direction. Substituting Eqs. 2.52, 2.66 and
2.68e into Eq. 2.86 yields
(2.88a)
(2.88b)
(2.88c)
Spherical rotation components can be derived from the cylindrical components in
a similar way, thus
ro =
1l

1 [a(Uesin<l
2R sin <I>
a<

[aU

au,]
ae

a(R

1
Ue)]

R
ro'=2Rsin<l> ae-sm<l>~
roe = _1
2R

[a(R u,) _au


aR

R]

a<

(2.89a)
(2. 89b)
(2.89c)

2.11.3 Rigid-Body Rotation and Translation Components

When a continuum rotates as a rigid body about the X -axis, following conditions
arise

Stress and Strain Tensors

75

=-:\=-:}",
aw av Eyz=O
(2.90a,b)
"uy
uz
in which a super-asterisk (.) denotes a rigid body component. For rigid-body rotation
about the y-axis,
CO

au aw
co=-=--,
Y
az
ax

(2.91a,b)

Ezx=O

and about the z-axis,

av au
co =-a =-:)' Exy =0
(2.91c,d)

!X
uy
In cylindrical coordinates; for rigid-body rotation about the r-direction,

au.

co, = rae =

auo
--az-'

Eo. = 0

(2.92a,b)

about the e-direction,


(2.92c,d)
and about the z-axis,

aUe

Ue

au,

co = - = - - ar r rae'

(2.92e,f)

E,o=O

In spherical coordinates; for rigid-body rotation about the R -direction,

COR

au.)

aUe
1(
= Rae!> = R sine!> lCOS e!>uo- ae '

E4>9 = 0

(2.93a,b)

EaR =0

(2.93c,d)

(2.93e,f)
These formulas of rigid-body rotation components in terms of displacement components
are useful in the derivation of governing equations of bending of plate or shell structures
which have small thickness in one direction. In such problems, it is very common to
assume that a normal in the same direction as the small thickness remains normal after
deformation. This is equivalent to the assumption that the thin cross section rotates as
a rigid body. We adopt this assumption in the thin plate problems in Chapter V.
We showed in Section 2.8 that if the body is displaced as a rigid body, according
to Eq. 2.53, then the Cartesian rigid-body rotation components are
(2.94a,b,c)

Foundations 0/ Solid Mechanics

76
If it is merely translated (Le. if Bj

ponents can be put as

=0), then its Cartesian rigid-body translation comv = Cz , w' = C3

u' = C1 ,

(2.95a,b,c)

The corresponding equations in cylindrical coordinates are


u; =C1 cos9+Czsin9

(2.96a)

U; =-C 1sin 9 + Czcos9

(2.96b)
(2.96c)
(2.97a)

00; =-B

(2.97b)

sin9+Bzcos9

(2.97c)
and in spherical coordinates

u; =sincl>(C1 cos9+Czsin9)+C coscl>

(2.98a)

u; =cos cI>(C1cos 9 + Czsin 9) - C sin cI>

(2.98b)

u; =-C 1sin9+Czcos9

(2.98c)

00; =sin cI>(B1cos 9 + B z sin 9) + B3 cos cI>


00; =cos cI>(B1cos 9 + Bz sin 9) - B3 sin cI>
00; =-Bl sin 9 + Bzcos 9

(2.99a)
(2.99b)
(2.99c)

2.12 EXERCISE PROBLEMS


With permission of the publisher, McGraw-Hill Book Company, Problems 2.12.1
to 2.12.27 were taken from the book by Mase (1970) and Problems 2.12.28 to 2.12.30
from the book by Wang (1953).

2.12.1

The state of stress at a point is given by the stress tensor


[crvl

=[a:

a;

~~l,

bcr ccr crJ


where a, b, c are constants and cr is some stress value. Determine the
constants a, b, c so that the stress vector vanishes on the plane which
makes equal angles with the reference Cartesian axes.
Answer: a = b = c = -1/2 .

Stress and Strain Tensors

2.12.2

77

yn

The stress tensor at a point P is given by the array


[aJ

=[ ~5

Determine the stress vector on the plane passing through P and parallel to
the plane ABC shown in Fig 2.18.
Answer: <Xl,XZ,X 3) =(-9,5, 10)17.

Fig. 2.18
2.12.3

The state of stress throughout a continuum is given with respect to the


Cartesian axes by the array
3X1XZ

[crijl = 5x;
[

5x; 0 ]
0 2xz ,
2x3 0

Determine the stress vector acting at the point P(2, 1, f3) of the plane that
is tangent to the cylindrical surface x;+x;=4 at P (Fig. 2.19).

Fig. 2.19

Foundations 0/ Solid Mechanics

78

2.12.4

For the distribution of the state of stress given in Problem 2.12.3, what form
must the body force components have if the equilibrium equations are to
be satisfied everywhere.
Answer: Xl =-13x2' X2 =-2, X3 =0 .

2.12.5

The state of stress at a point is given with respect to the Cartesian axes by
the array
[aJ

[2 -2 OJ

O.

-2...J2
o 0

--./2.

Determine the stress tensor cij for the rotated axes related to the unprimed
axes by the transformation tensor

raJ

=[11~

11...J2
112
11...J2 -1/2

Answer:

-11~
112.
-112

2J

-1

1-...J2
-1

2.12.6

+-./2.

The stress tensor at a point P is given as

1
0
1 2
Determine the principal stress values and
Answer:
[aJ

2.12.7

=[ 31

IJ
2.
Q
the principal stress directions.

(n~l), nJI),njl)

=(0, 1,-I)/...Ji,

(n~2),nJ2),nj2)

=(l,-I,-I)/...j3,

(n~3),nJ3),nj3)

=(-2,-1,-1)/-.J6.

The stress tensor at a point is given by

79

Stress and Strain Tensors

[a~=[~ ~~

-H

Detennine the principal stresses and the maximum shear stress at the point.
Answer:

Fig. 2.20
2.12.8

The state of stress throughout a body is given by the stress tensor

[crvl

=[c:o

C;3
-CXl

-~Xl]'
0

where C is an arbitrary constant. (a) Show that the equilibrium equations


are satisfied if body forces are zero. (b) At the point P(4, -4, 7) calculate
the stress vector on the plane
and on the sphere

(c) Detennine the principal stresses, and maximum shear stresses at P.


Answer: (b) For the plane, (Xl>XZ'X 3) =(14,18, -8)C/3; for the sphere,
(-28,0, 16)C/9. (c) (crl>crz,cr3)=(1,0,-1)V65C, cr.=V65C.
2.12.9

At point P stress tensor is

[crvl =[

~4

-7

;1
0

~7].
35

Detennine the stress vector on the plane at P parallel to planes (a) BGE ,
(b) BGFC of the small parallelepiped shown in Fig. 2.21.

Foundations of Solid Mechanics

80

6"
B
~-----'=-----::>IC

I
A~-+I---"""""D
4"

I
I

G /~~-------~7,,=------~X2
F

Fig. 2.21
Answer: (a) (X 1
2.12.10

,.xZ,X3) = (11, 12,9), (b)

(21,14,21)1'''5.

Detennine the normal and shear stress components on the plane BGFC of
~blem 2.12.9.
Answer: ON 63/5, Os 7~15 .

2.12.11

The principal stresses at point P are 0 1 = 12, Oz = 3, 0 3= -6 . Detennine


the stress vector and its nonnal component on the octahedral plane at P.
Answer: ~1,.xZ,.x3) =(12, 3,-6)/V3; ON =3.

2.12.12

Detennine the principal stress values for

(.)[a~ =[! ~ ~

and

~)[a~ =[! ~ ~

and show that both have the same principal directions.


Answer: (a) 0 1 =2, Oz=03=-1, (b) 0 1 =4, OZ=03=1.
2.12.13

Show that the normal component of the stress vector on the octahedral plane
is equal to one third the frrst invariant of the stress tensor.

2.12.14

The stress tensor at a point is given as

[aJ=~

+n

with Ozz unspecified. Detennine Ozz so that the stress vector on some plane
at the point will be zero. Give the unit normal for this traction-free plane.
Answer: Ozz = 1; (n 1,"z,fl]) = (1,-2, 1)1~.

81

Stress and Strain Tensors

2.12.15

In a continuum, the stress field is given by the tensor

X~X2

(l-x;)x1

[OJ= [ (l-x;)x1

(xi-3x:J/3

0]
O.

2x32

Determine (a) the body force distribution if the equilibrium equations are
to be satisfied throughout the field, (b) the principal stress values at the
point P (a,O,2-[ci), (c) the maximum shear stress at P.
Answer: (a) X3=-4X3' (b) a,-a,8a, (c) 4.5a.
2.12.16

For a displacement field given by


(Ul'~'~) =Xl - X2)2, (X2 + x3i, -X IX2).
Determine at point P(O,2,-1) , (a) the linear strain tensor, and (b) the
change in length per unit length in the direction
(nl'~'~)

Answer:
(a)

=(8,-1,4)/9.

[ 20 2 -2]1,
-2

(b) -2/27.

2.12.17

Use the results of Problem 2.12.16 to compute the change in the angle
between two unit line elements (nb~'~) = (8,-1,4)/9 and (mb"'-2, fn])
= (4,4,-7)/9 at the point P.
Answer: 106/27 radians.

2.12.18

A linear (small strain) deformation is specified by


u l =4x l -x 2+3x3, ~ =xl +7X2' ~ =-3x l + 4x 2+ 4x 3
Determine the principal strains.
Answer: 8, 4, 3.

2.12.19

A 45 strain-rosette measures longitudinal strain along the axes shown in


Fig. 2.22. At a point P ,
x'I

X2

45
45

Fig 2.22

XI

82

Foundations of Solid Mechanics

tl = 5 X 10-4, e'l1 = 4 X 10-4, ~ = 7 X 10-4 in/in.


Detennine the shear strain el2 at the point.
Answer: t2=-2x 10-4.

2.12.20

The state of strain throughout a continuum is specified by

x: x;
[
[Evl = x; X~
XI~

X3

XIX3]
x~.
XI

Are the compatibility equations for strain satisfied?


Answer: Yes.

2.12.21

For a homogeneous defonnation the small strain tensor is given by

0.01

- 0.005
0]
0.02
0.01.
o
0.01
-0.03
What is the change in the 90 angle ADC depicted by the small tetrahedron
OABC in Fig. 2.23 if OA = OB = OC , and D is the midpoint of AB ?

[Evl = [ -0.005

Answer:

-o.OllV3.

Fig 2.23

XI

2.12.22

At a point the strain tensor is given by

[e,]

=[!J
-1

Show that it is in principal form as

[~=[~

-1
4
0
0
4
0

-1]o .
4

Stress and Strain Tensors

83

Calculate the strain invariants for each of these tensors and show their
equivalence.
Answer: 11 = 13, 12 =54, 13 =72.
2.12.23

Show that the displacement field


u1=Ax1+3x2, ~=3X1-Bx2' ~ =5,
gives a state of plane strain (E:!3 =E:!1 =E:!2 =0) and determine the relationship
between A and B for which the deformation is isochoric (constant volume
deformation).
Answer: A =B.

2.12.24

A so-called delta-rosette for measuring longitudinal surface strains has the


shape of the equilateral triangle /1 and records normal strains ell> e~l' e;l in
the directions shown in Fig. 2.24. If Et1 =a, e~l =b. e;l =c. determine e12
and ~ at the point.
Answer: e12 = (b - c )/-{'3 and ~2 = (-a + 2b + 2c )/3.

Fig. 2.24
2.12.25

The strain tensor at a point is given by

1
[
[evl = - 3

-3
1

...j2

-...j2

Determine: (a) the normal strain in the direction of a unit vector nj. of which
(nlo ~, ~) =(1, -1, ...j2)/2; (b) the shear strain between nj and another unit
vector mj, of which (m1'~'~) =(-1,1, ...j2)/2; (c) determine the
principal-axes form of E;j and note that nj and mj are principal directions;
(b) calculate the three strain invariants.
Answer: (a) 6, (b) 0, (c) (elo~' E:!) =(6, 2, -2). (d) (/1,12,13) =(6, -4, -24).
2.12.26

A displacement field is given by


U1 =3x 1xi, ~ =2x~1'

~ =xI - X1X2

Foundations of Solid Mechanics

84

Detennine the strain tensor


ditions are satisfied.
Answer:

Ev and check whether the compatibility con-

3X 1X2+ X3

x/2

- X12]

x1/2.
~

Of course, the compatibility conditions are assured because the strains arise
from a displacement field which is given.
2.12.27

For a delta-strain-rosette the nonnal strains are found to be those shown in


Fig. 2.25. Detennine 12 and ~ at the location.
Answer: ~ 1 X 10-4, Et2 -0.2885 X 10-4.

2.12.28

Show that
XX k(X2 + y2), e" ky2, "(19 2kxy, E..
="(u 0,
where k is a small constant, is a possible state of strain; while
XX kz(x2+y2), e,,=k y 2z, "(19=2kxyz, e..="(.,.="(u=O,
is not a possible one.

=, (.,.

2.12.29

An elastic solid is heated nonunifonnly to a temperature distribution


T(x,y,z) , where T(x,y,z) is a function of x,y,z. If each element in the
body has unrestrained thennal expansion, the strain components will be

e.:., = e" = n = aT,


"(19 = "(.,. = "(U = 0,
where a is the coefficient of thennal expansion and is a constant. Prove
that this can occur only when T is a linear function of x,y,z.
2.12.30

Detennine the relations among the constants A o , AI' B o , B 1 , Co, C1 ,


C2 , so that the following is a possible system of strains:

e.:., =Ao+Al(x2+ y2) + (x 4 + /),

Stress and Strain Tensors

85

e" =Bo + B 1(X 2 + l) + (X 4 + y\


"f", =Co + C1XY(X 2 + y2 + CJ,

e.. ="fyo ="fa =O.

Answer: A1 +B l -2C2 =O, C1 =4.

CHAPTER III
LINEAR ELASTICITY

3.1 STRAIN ENERGY FUNCTION


So far we have discussed some general concepts of mechanics of solids under
small defonnation. A solid is classified as elastic, if it possesses a homogeneous
stress-free natural state, and if in an appropriately defined finite neighborhood of this
state there exists a one-to-one correspondence between the stress tensor C1 jj and the
strain tensor 9;. This definition is based on Cauchy's approach (A. L. Cauchy, 1827).
It is understood here that the natural state is a state of thermodynamic equilibrium in
which all components of stress and strain are zero throughout the body. In other words,
the definition incorporates the idea that the body returns to the natural state when all
loads are removed.
Consider the case where the body force Xi acts on an infinitesimal body of volume
dV at point P , and the body is deformed in such a way that point P moves through
the distances dui . Then the increment of work done by the body force Xi is
dWb =XiduidV
(3.1)
At the same time, the work done by the surface force Xi on the infmitesimal surface
dS (enclosing dV) is
or, due to Eq. 2.6,

dW. =C1jjnidujdS

By Gauss's divergence theorem, we can write


dW, =(C1 jjdu j )}V,
or, after perfonning the indicated differentiation,
dW, =(C1 jjidu j + C1 jjdu j i )dV

(3.2)

If the process is adiabatic, i.e. without loss or gain of heat, the whole work done will
be stored up inside the material body as potential energy, or strain energy as it is called
in the mechanics of defonnable bodies. Denoting such energy per unit volume of the
material by U, we have

87

Linear Elasticity

dUdV=dWb+dW.

(3.3)

or, due to Eqs. 3.1 and 3.2,


dUdV =[(X.J + cr!I,I.)du.J + cr!I..duJ,iJJdV,
or, due to the equilibrium (Eq. 2.14),
dU =crijduj,i

(3.4)

It should be noted that Eq. 3.3 is valid for conservation of energy in dynamics if the
body force Xi in Eq. 3.1 is taken as including the D' Alembert's force -PUi' On the
other hand, if Xi is taken strictly as the actual body force only, Eq. 3.3 should be
written
(3.5)
dU dV + dTdV = dWb + dW.
where T is the kinetic energy per unit volume of the material, i.e.
I

.2

(3.6)

T='2 PUi
Thus

=PduiUi
Substituting the equation above together with Eqs. 3.1 and 3.2 into Eq. 3.5, in view of
Eq. 2.10, yields the same result as before, i.e. Eq. 3.4. Rewrite the latter as

dU

='2 (crijduj,i + crjidui),

or, due to the symmetry of crij'


(3.7)
If the stress components are unique homogeneous functions of the strain components,
there exists the strain energy function U as a scalar homogeneous function of the strain
components, i.e.
(3.8)
Accordingly, its total differential is

au

dU=-dE ..
af;j

!I

(3.9)

In Eqs. 3.7 and 3.9, each of the nine components of the strain tensor are considered as
an independent variable in the function U. Comparing these two equations, we have

au
aev

crij =

(3.10)

The existence of a strain energy function such that the stress components are unique

Foundations of Solid Mechanics

88

homogeneous functions of the strain components is a basic property of elastic solids.


For one-dimensional cases, the stress-strain relationship for elastic solids are shown in
Fig. 3.1, where the loading and unloading paths coincide as implied by the theory.

~Oading

Loading

o
(a) Linear.

(b) Nonlinear.

Fig. 3.1 One-dimensional stress-strain relationship for elastic solids.


When Hooke's law (R. Hooke, 1678) is obeyed the material is classified as linear
elastic. This law states: Stresses are linear and homogeneous functions of strains, i.e.
(Jij

=Cij/d/d

(3.11)

We shall now show that the symmetry of the stress and strain tensors implies that there
are only 21 independent components among the 81 Cij/d'
Cij/d/d

au

=-a
'
U

the derivative of which with respect to e"". is


Cij/d8bt8",

(fu

=-,
aeuae",..

or
(fu

Cij"",

Interchanging

=aeuae",..

(3.12)

with m and j with n in the equation above gives

C .. = (fu
IIUIIf

ae",..aeu

(3.13)

Linear Elasticity

89

which shows that CijlrJ =C lrJij Inasmuch as the components (1ij are symmetric, an
interchange of the idices i and j in Eq. 3.11 does not alter the formula, thus CijlrJ =CjilrJ
and consequently,
(3.14)
CijlrJ =ClrJij =CjilrJ =Cij/k
Substituting Eq. 3.11 into Eq. 3.7 yields
dU =CijlrJlrJdeij

or, due to Eq. 3.14,

="2 CijlrJd(ijIrJ).
Integrating the equation above from the natural state, where all stresses and strains are
identically zero, leads to

="2(1ijij

(3.15b)

To verify Eq. 3.10, let us take the derivative of Eq. 3.15a with respect to e".", i.e.

au

ae"."

a;j)

( a1rJ
ae"." v + ae"." 1rJ

="2 CijlrJ
1

= "2 CijlrJ(0km0",ij + imjnlrJ)

or, due to Eq. 3.14,

=(1 .... ,

according to Eq. 3.11.

90

Foundations of Solid Mechanics

On the other hand, we can take the advantage of the symmetry of the stress and
strain tensors and consider each of these tensors to be tensors of the fIrst order but with
a range 6 in the form of matrices as follows

crll
crzz
cr33
,
{cr.} =
cr12
cr23
cr3

ell
En
{e.} =

~3

(3. 16a,b)

"{12
"{23

"{3

By expansion, we can show that Eq. 3.7 can be rewritten in the form
dU = cri~' (i summed from 1 to 6)

(3.17)

Note that the engineering (not the tensorial) shear strain components must be used.
Assuming the existence of the strain energy function U as a function of the six
independent strain components as listed in Eq. 3.16b, we can write
(3.18a)
U =U(eJ, (i = 1 to 6)
and
dU=dU de.

dj

Comparing Eq. 3.18b with Eq. 3.17 leads to


dU
cri =dq

(3.18b)

(3.19)

which is essentially identical to Eq. 3.10 obtained earlier.


For linear elastic solids, the stress-strain relationship, Eq. 3.11, can be put in the
form

cri =Cijej

(3.20)

where Cij is the stiffness tensor. Substituting Eq. 3.20 into Eq. 3.19 leads to
dU

Cijej = dj'

the derivative of which with respect to e", is

azu

CijBjm=-d
d '
je",
or

C.

un

= azu

djde",

(3.21)

Interchanging i and m in the equation above, we can show the symmetry of Cij' i.e.
Cij =Cft. Thus Hooke's law for the most general kind of anisotropy has the matrix

91

Linear Elasticity

fonn

Cu

au

a22
a33
a12
a23
a3

C13 C 14
C23 C24
C 33
C34
C44

=
Symm.

CIS

C 16

Czs

C26 En

C 3S

C 36

C 4S

C46

C ss

C S6

"{23

"{3

~3

(3.22)

"{12

The matrix [Cvl, being symmetric, has 21 independent elastic constants. Since the strain
energy function is
(3.23a)

(3.23b)
and is positive for any non-trivial set of strain components, the matrix [Cvl must be
positive definite as described in Section 1.6 and elaborated in Section 3.6.5 for various
types of linear elastic solids.
3.2 ORTHOTROPIC AND ISOTROPIC ELASTIC SOLIDS
Almost all of the materials used for engineering purposes exhibit certain degrees
of elastic symmetry, and require less than 21 independent elastic constants. First let
the 1,2, 3 reference frame in Fig. 3.2 be such that the 1- 2 plane is a plane of symmetry.
Let another frame l', i, 3' be generated by a 1800 rotation about the 3-axis. The
transfonnation tensor aij linking the two systems is

[a~ =[~1 ~1 ~
Transfonning stresses according to Eq. 1.27 yields
,

,
a 33

a23 =-Q'23

(3.24)

=a 33,

Similarly, the transfonnation of strains gives

,
"{12

En =En,
,

~3=~3'

="{12
(3.25)

92

Foundations of Solid Mechanics

/ ________

I
I

I
I
I
I

I ('

Fig. 3.2 Transfonnation for orthotropic symmetry with respect to plane 1- 2.


Substituting Eqs. 3.24 and 3.25 into Eq. 3.22 leads to
,

0"11

Cll

C13 C14
CZ3 CZ4
C33 C34
C44

0"22
0"33
O"lZ

Symm.

0"Z3
,

-CIS
-Czs
-C3S
-C4S
Css

0"3

-C16 11
,
-Cu ~
,
-C36 ~3
,

-C46 11Z
CS6
1Z3
,
C66
13

(3.26)

Since the stress and strain relationships of 1 2 3 and l' 2' 3' systems must be
identical, the stiffness matrix [Cy] in Eq. 3.26 must be equal to the one in Eq. 3.22.
This implies that the negative elements Cij in Eq. 3.26 must be zero, i.e.
Cll

[Cy]

CIZ
C22

=
Symm.

C13 C14
CZ3 CZ4
C33 C34
C44

0
0
0
0
0
0
0
0
Css CS6
C

(3.27)

Note that there remain 13 essential elastic constants for the case of a single plane of
symmetry.
An additional plane of symmetry, say the plane 2-3, will further reduce the
stiffness matrix to

Linear Elasticity

93

Cn
[CJ

C 12

C 13

Cl l

C 23

0
0
0

C 33

0
0
0
0

C44
Symm.

C ss

0
0
0
0
0

(3.28)

C
If the plane 3 -1 is also the plane of symmetry, no further reduction in the number
of elastic constants can be made. Such an elastic solid, which has three mutually
orthogonal planes of symmetry, is said to be orthotropic; and it has nine elastic constants.
Wood and fiber reinforced plastics are examples of orthotropic elastic solids.
Materials, which behave mechanically in the same manner for all directions are
called isotropic materials. An isotropic elastic solid has more elastic symmetry than an
orthotropic one, and therefore has fewer elastic constants. We must have, for an isotropic elastic solid,
Cn = Cl l= C33 ,
C 12 =Cn

=C

31 ,

and
~=~=~

~~

Equations 3.29 show that there are three essential elastic constants, but they are not
independent, as we now show. Rotate the coordinates 1, 2, 3 about the 3 -axis through
an angle 45, to give 1', i, 3' as shown in Fig. 3.3. The transformation tensor is

11-5 -11-5 01
[aJ = 11-: 11: ~J
[

,,

I'

//

Fig. 3.3 A transformation for isotropic solid.

94

Foundations of Solid Mechanics

Repeating the analysis already used for orthotropy yields the relation
2C 44 C u - C 1Z

(3.30)

Equations 3.29 and 3.30 show that the stiffness matrix for isotropic elastic solids has
the form
A.
A.
0 0 0
A.+21l
0 0 0
A.
A.+21l
A.+21l 0 0 0
(3.31)
(CijI
Il 0 0
Symm.
Il 0

where only two constants A. and Il, commonly known as Lame's constants (0. Lame,
1852), remain. Since Il relates a shear stress with a shear strain, it is called the shear
modulus, and it is sometimes replaced by the symbol G. In indicial notation, Hooke's
law (Eq. 3.11) for an isotropic solid can be written as
(3.32)
C1ij = AEUBij + 2j.lij
In fact, Eq. 3.11 can be reduced to Eq. 3.32 upon setting
CijiJ =ABijBiJ + Il(BjkBj / + Bi/Bjk )

(3.33)

With the introduction of the stress deviator sij and the strain deviator eij as in Eqs.
2.34 and 2.47, Hooke's law for isotropic elastic solids separates into two parts, governing
shear and bulk behavior, i.e.
(3.34)
C1 u

=3Ke u

(3.35)

where K, known as the bulk modulus, is equal to A. + 2Jl13 .


For the state of uniaxial stress in the Xl -direction (only the stress C1u is non-zero),
Young's modulus E (T. Young, 1807) and Poisson's ratio v (S. D. Poisson, 1829)
can be introduced through the relationship
(3.36a)
and
(3.36b)
In terms of E and v the constitutive law for isotropic elastic solids becomes

C1ij = I! v (eij + 1 2v Bijeu)

(3.37)

Table 3.1 shows a number of useful relations among the isotropic constants. In the state
of a uniaxial stress where the non-zero stress and strains are as defined in Eqs. 3.36,
real observations show that the signs of C1u and eu are the same but opposite to that
of Zz and 8:J3' Hence we may state that

E>O,

v>O.

Linear Elasticity

95

Table 3.1
A,Jl

E,v

Jl,K

Jl,v

Ev
(1 + v)(l- 2v)

3K -2Jl
3

2Jlv
1-2v

Jl

Jl

2(1 +v)

Jl

Jl

2Jl(1 +v)
3(1-2v)

K
E

2
A+-Jl
3
Jl(3A + 2Jl)
A+Jl
A

3(1-2v)
E

2Jl(1 +v)

Another special state of stress, known as hydrostatic pressure, is defmed by CJij =-POij'
By employing Eq. 3.35 we fmd p =-K kk , where K =E/[3(1- 2v)] . Since a hydrostatic
pressure p should give a negative dilatation, we should have K > 0, i.e.
1-2v>0.
Hence the ranges of values of the engineering elastic constants E and v are
E > 0,
0 ~v ~ 112
(3.38)
When v = 112, the material is called incompressible since the dilatation remains zero
irrespective of the stress components.
3.3

YOUNG'S MODULI AND POISSON'S RATIOS FOR ORTHOTROPIC


ELASTIC SOLIDS
The inverse of Eq. 3.22, for [CJ of orthotropy as defmed in Eq. 3.28, is
""

S11

CJ""

e",

S21 S22 S23 0


S31 S32 S33 0
0
0
0 S44
0
0
0
0
0
0
0
0

CJ)'Y

CJ

CJxy

Sss
0

CJyz

E..
1xy
1yz
'Y,

S12 S13

(3.39)

where
(3.40a)

Sij=Sji
[S~

= [C~-I;

i,j=1,2,3

(3.40b)

Foundations 0/ Solid Mechanics

96

and
(3.4Oc)

An elastic solid in which


0'.. = 0',. =

O'u = 0

(3.41)

is said to be in a state of plane stress. Its essential stress-strain relations become

.a} -_[S11
S21
{e"
'Y",

S12
S22
0

~ {O'''''}
O'yy

0
S

(3.42)

0'",

Taking the inverse of Eq. 3.42 yields,


(3.43a)
(3.43b)
where
(3.44a)
and
Eij =Eji; i,j

=1,2

(3.44b)

Suppose it is found in an experiment that 0'..,. is the only non-zero stress. Equation
3.43a gives
(3.45a)
(3.45b)
or
(3.400)

(3.46b)
From this experiment we may determine tile Young's modulus E" and the Poisson's
ratio v", by using the equations
0'..,.

E=

(3.47a)

v =_ e"
'"
e..,.

(3.47b)

" e..,.

97

Linear Elasticity

Substituting Eq. 3.47a into Eq. 3.46a yields


(3.48a)
Equation 3.4Sb gives

E12
Vxy =E22

(3.48b)

A repetition of this operation with 0' being the only non-zero stress leads to

E~2
E., =E22 -Ell
-

(3.49a)
(3.49b)

It should be noted that

E" Vxy
E., =Vyx

(3.50)

In addition, it is customary to name C44 a shear modulus Gxy, and the 4 distinct elastic
constants for the case of plane stress of orthotropic elastic materials (see Eqs. 3.43) are

E"- E ll = I-vxyv.,,,

(3.51a)
(3.S1b)

E.,
E22 =----'-l-vxyvyx

(3.Slc)

C44 =Gxy

(3.5ld)

Moreover, Sij in Eq. 3.42 can be obtained as


I
S11=-

E"

(3.S2a)

(3.52b)
(3.S2c)
(3.52d)

Foundations of Solid Mechanics

98

Finally, we can write the remaining Sij 's in Eq. 3.39 as


1
E.

(3.52e)

S33=-

v,.

v~

Ey

E.

S23=S32=--=--

(3.520

1
Sss=G,.

(3.52g)
(3.52h)

S66

=-G1

(3.52i)

In total, there are 9 elastic constants for an orthotropic elastic material.


The constitutive law for an orthotropic elastic material can be specialized for an
isotropic elastic material by setting
(3.53a)
Ex=Ey=E.=E
Vxy = v,. =vu=vjiX =v~ =v;u =V

(3.53b)

and
E

Gxy = G,. = G"" = Il= 2(1 +v)

(3.54)

3.4 SOLUTION SCHEMES


It was stated in Section 3.1 that if the process in an elastic system is adiabatic,
the work done by the forces must be equal to the strain energy which is stored up in
the system. This fact can be expressed in another form known as the principle of virtual
work, i.e.
where WE is external virtual work, i.e.
WE

=Iv XjBujdV + Is XjOujdS

(3.56)

and W/ is the internal virtual work or virtual strain energy (Eq. 3.7), i.e.
W/

O'ijOe;jdV

(3.57)

In the equations above, V denotes the material domain, S the surface completely
enclosing V, and B the variational symbol signifying a virtual quantity (Spiegel, 1971).

Linear Elasticity

99

Due to Eq. 2.44, Eq. 3.57 can be written as


WI

=Iv O'ij~(ui)dV,

or, upon applying the divergence theorem (Eq. 1.88),


WI

=Is <1jinj~uidS -

Iv

O'jiJ~uidV

(3.58)

Substituting Eqs. 3.56 and 3.58 into Eq. 3.55 yields

Iv(O'jiJ+Xi)~UidV + Is(Xi-O'jinj)~UidS =0,


which can be satisfied for arbitrary

~Ui

only if

O'jiJ+Xi=O,

(3.59)

in V

and either
(3.6Oa)

~Ui=O, onS

or
(3.6Ob)
Note that Eq. 3.59 is the equilibrium equation, Eq. 2.8, and Eq. 3.6Ob is the stress vector
formula, Eq. 2.6. Equation 3.59 is afield equation. On the other hand, Eqs. 3.60 specify
boundary conditions in a proper way, i.e. at a certain point on the boundary Seither
ui

=ui

or <1jinj =Xi

(3.61)

in which a bar (-) denotes a prescribed quantity. These well-posed boundary conditions
state that, at a certain point on the material boundary, either 1 or X1 , either Uz or
X2 , and either U3 or X3 must be prescribed. IT, as an example, the boundary conditions
are the prescription of U1' Uz and X2 , the solution to the problem may not be possible
since the boundary conditions are not well posed. In other words, boundary conditions
are improper and the solution may not exist if a displacement and a traction in the same
direction are prescribed at a certain point on the boundary of the material domain.
The principle of virtual work and its consequences derived so far hold for nonlinear
as well as linear elastic materials. Each boundary value problem is governed by fifteen
field equations to be solved for flfteen unknown variables. The fifteen field equations
are three equations of motion (Eq. 3.59), six equations for the strain-displacement
relationship (Eq. 2.44), and six equations for the constitutive relationship (e.g. Eq. 3.11
for linear elastic solids). On the other hand, the fifteen unknown variables are three
displacements, six stresses and six strains. The boundary conditions (Eqs. 3.61) are
needed for determining uniquely the constants arising from integrating the equations of
motion and the strain-displacement relations, which are differential equations.
For linear elastic solids, solution methods may be classified into two main schemes,
one in which the basic unknowns are displacements, one in which they are stresses or
strains. In the first scheme, the fifteen field equations can be reduced, by simple

100

Foundations oj Solid Mechanics

substitutions, to three field equations for the three displacements. For an isotropic elastic
solid the equations are
lUI. u,+ (A. + II)Eu, .+X =0.
~tM
~
M.I
I
as described in the next section. In the second scheme, since strains (or stresses) are
chosen to be the basic unknowns, the compatibility conditions (Eq. 2.62) must be
incorporated in the approach to guarantee the uniqueness and continuity of the solution.
In general, the first scheme is more straight-forward and should be more efficient than
the second scheme, since it involves fewer unknowns. The second scheme is effective
for elastostatics of isotropic plane problems, to be discussed in Chapter IV. In Chapter
V a 'mixed' scheme developed for elastic bending of thin plates is presented. In Chapter
VI and vm the first scheme is described for elastostatics and elastodynamics of isotropic
solids.
Sometimes it is convenient to replace one boundary value problem by using the
Maxwell-Betti reciprocal theorem (1. C. Maxwell, 1870 and E. Betti, 1872) which states
that:
'In a linear elastic solid, the work done by a set of forces acting through the
corresponding displacements produced lJy a second set of forces is equal to the work
done lJy the second set of forces acting through the corresponding displacements produced lJy the first set of forces:
Denoting the frrst and second sets by superscripts 1 and 2 respectively, we can
state the theorem as

Jv( X~)u~)dV + Js(r')u~)dS =Jv( X~)u~)dV + Js(rz)u~')dS


II

II

II

(3.62)

II

The proof of this theorem starts with rewriting the left-hand side of the equation above
by using the procedure for obtaining Eq. 3.7. The result is

id:)~)dV,
or, due to Hooke's law (Eq. 3.11),

Jv(CijldEId(I)Gi(2bV
u
Similarly, the right-hand side of Eq. 3.62 can be rewritten as

Jv(CijldEId(2)Gi(I)u..V
The last two expressions are identical since
is proved.

Cijld

CIdij

(Eq. 3.14), and thus the theorem

3.5 FIELD EQUATIONS IN TERMS OF DISPLACEMENTS


In this section, displacement field equations will be presented for the Cartesian,

101

Linear Elasticity

cylindrical and spherical coordinates. The presentation will be limited to linear isotropic
elastic solids, of which the constitutive laws are: for the Cartesian system (Eq. 3.32),

crij =Aekkoij + 2JlEij

(3.63)

for the cylindrical system,

cr"
cree
cr..
crr9
cr9z

A+2f..l

A
A+2f..l

A
A
A+2f..l

Symm.

cr"

00
0
0
00
00
0
2f..l 0 0
2f..l 0
2

Err

fee
E..
Ere

(3.64)

E"

and for the spherical system,

crRR
cr..
cree
crR,

A+2f..l

A
A+2f..l

A
A
A+2f..l

Symm.

cr~

cr8//

00
0
00
0
000
2f..l 0 0
2f..l 0
2

ERR
E..

fee
ER ,

(3.65)

E~

Substituting Eq. 3.63 into the equations of motion (Eq. 2.14), and using the straindisplacement relationship (Eq. 2.44), we obtain the Cartesian displacement field equations known as Navier equations (C. L. M. H. Navier, 1823 and 1827), i.e.

f..lUi,kk + (A + f..l)Ekk,i + Xi =0

(3.66)

in which it has been assumed that the material is homogeneous, i.e. A and f..l are
constants. The cylindrical Navier equations can be obtained, by using Eqs. 2.71, 2.74
and 3.64, as
(3.67a)

(3.67b)
(3.67c)
The spherical Navier equations can be obtained, by using Eqs. 2.81, 2.84 and 3.65, as

2UR

2 dU, 2u,cotc!>

(A+f..l)dR(Ekk)+~VUR-R2-R2d<!>

R2

2 dUe)
R 2 sinc!>da +XR=O,

Foundations of Solid Mechanics

102

In these equations,

VZ

is an invariant operator as dermed in Eq. 1.85d, i.e.


(j

VZ=OXkdxk

(3.69a)

(3.69b)

(3.69c)
and eu is the dilatation defined in Eq. 2.72 for cylindrical coordinates and in Eq. 2.82
for spherical coordinates.
3.6 PROBLEMS

3.6.1 Beltrami-Michell Compatibility Conditions


The stress-strain relationship for an isotropic elastic solid, Eq. 3.37, has the inverse
Eeq =(1 +v)O'ii -vSiiO'u
(3.70)
Thus, for a homogeneous material,

Eeq,u =(1 +v)O'ii,U -vSliO'U,mm

(3.71a)

Eeu,li =(1- 2v)O'u,1i

(3.71b)

Eeu"ii =(l + v)O'i.I:,ii - VO'U,Ii

(3.72a)

Eeii,i.I: =(l +V)O'ii,i.t -VO'U,ii

(3.72b)

A spatial derivative of the equation of equilibrium, Eq. 2.8, can be put as


O'i.t,ik

=-Xi,i

(3.73)

Due to Eqs. 3.72 and 3.73, we can write

E (eu,Jk + eik,i.t) =-(1 + v) (XiJ + Xi,i) - 2vO'U,1i

(3.74)

Substituting Eqs. 3.71 and 3.74 into the compatibility conditions, Eq. 2.62, with k
summed, we obtain
(1 +v)av."'"
...... +0'Mi,V
...... -vSII..O'.Mi.1f'IItI
... + (1 +V)fV
.. +X},I.) 0
(3.75)
\LI..',J

103

Linear Elasticity

Summing the equation above for i =j , we find


2
l+v
V (j/dr; =--I-Xk k

-v '

(3.76)

Substituting Eq. 3.76 into Eq. 3.75 yields the Beltrami-Michell Compatibility conditions
(E. Beltrami, 1892 and J. H. Michell, 19(0), i.e.
1
v !:
V2(j.. +--(j/dr;
.. +--u"Yu+x.
.+X =0
(3.77)
" 1 + v ," 1 - vir, ,
',J
J,'
If the body force Xi is zero or constant, Eqs. 3.76 and 3.77 show that (j/dr; is
harmonic, in the sense that it satisfies Laplace (harmonic) equation,

V2(j/dr; =0

(3.78)

1
V2(j .. +-(j/dr;
.. =0
" 1 +v ,'J

(3.79)

and Eq. 3.77 becomes

Applying the Laplace operator V2 on the equation above, and using Eq. 3.78, we obtain
the biharmonic equation

V'(jii =0

(3.80)

where V4 is the biharmonic operator, i.e.


(3.81)
V'=V2V2
Equation 3.80 is a set of six distinct uncoupled partial differential equations. It can be
shown that any biharmonic function c;I>, i.e. a solution of
V'c;I> =0
(3.82)
can be written
(3.83)

where C is an arbitrary constant tensor, 'P a harmonic scalar field, and 'Pi a harmonic
vector field, i.e.
(3.84a,b)
It is interesting to note that any function X, defined by

X=xix/P,ii

(3.85)

is harmonic if 'P is harmonic.


3.6.2 Conservative Forces and Potentials
If (and only if) the body force field is conservative, then it has a potential V such

that

X.=-V.

"

So that a simple check for Xi to be conservative is the equation

(3.86)

104

Foundations of Solid Mechanics

x.1,/.=X
'I,'

(3.87)

and the compatibility conditions (Eq. 3.77) become

VZO' .. +-1-O'LL .. -~a VZV-2V .. =0


'I
1 +v ....'1 1-v 'I
.'1

(3.88)

The static equilibrium equation. Eq. 2.8. is satisfied exactly if the stress tensor is
expressed in terms of a stress potential cf> as the following
0''I..

=
a. (V2cf> + V) 'I

ell.'1..

(3.89)

Note if Xi vanishes, we can set V to zero without loss of generality. Substituting Eq.
3.89 into Eq. 3.88, we fmd

a. (V'cf> + 1-2v
VZV)+ 1-V(VZell+ 1-2v V) =0
1-v
1+v
1-v
..
'I

Summing the equation above with i

(3.90)

'1

=j

leads to

V'eIl+ 1-2v VZV =0


1-v

(3.91)

Incorporating Eq. 3.91 in Eq. 3.90 yields


1-2v
VZcf>.'1.. =---V
..
1-v''I

(3.92)

Thus for an elastostatic, isotropic, homogeneous, conservative system, the solution


via stresses leads to the search for a single scalar function cf> linked to the stresses by
Eq. 3.89 and satisfying Eq. 3.92. The solution to Eq. 3.92 can be taken as a sum of a
particular solution and a complementary solution. While such a particular solution
satisfies the governing equation exactly without involving any arbitrary constants, any
complementary solution satisfies the homogeneous governing equation, i.e.

VZell.ii 0

(3.93)

The solution to which can be shown to be

(3.94)
ell ' + Cii"xiXjX" + CqX;Xj + C;XiX"x" + Cx"x"
where ' is any harmonic function, and C 's are constant tensors. Note that polynomials of Xi of lower degrees than two are disregarded since they give trivial stress
components. The same solution scheme is described in more detail in Chapter IV,
particularly for two-dimensional cases.

3.6.3 Elastostatic Displacement Potentials


We shall now describe solutions which start from the displacements rather than
the stresses. If the displacement components are given by
(3.95)
u =eIl.
then the dilatation .. becomes

.'

Linear Elasticity

105

(3.96)
and the static equilibrium equations for homogeneous isotropic solids, Eq. 3.66, become
(A.+2J.1)(~(I),i+Xi =0

(3.97)
where Xi is the actual body force vector. The function (I) is called Lame's strain
potential.
Similarly, if

(3.98)
the dilatation becomes
;;

=Vk'l'k,1i
1
='2(Vk'l'k,1i +jik'l'kJi)
1
=-(Eh'l'k
.. )
2 v ,!I.. -I',,'I'k
.."
,I'

(3.99)

=0
and Eq. 3.66 is specialized into
J.1vk("k) J + Xi =0

(3.100)

Lastly, if we set (Galerkin, 1930)


Ui =

A.+2J.1 2
A. + J.1 V Xi - Xk,ki

(3.101)

we can show that the dilatation and the equilibrium equations become, respectively,
kk

=" J.1 (~Xk).


1I.+J.1
,~

(3.lO2)

and

A.+2J.1r
- 'V.+X=O
{ A.+J.1
~
1

(3.lO3)

The vector Xi is called the Galerkin's vector.


The functions (I), 'l'i and Xi as introduced above are known as displacement
potentials. Ifthe actual body force Xi is conservative (Eq. 3.86) and each displacement
component is put as a linear combination of Eqs. 3.95, 3.98 and 3.lO1, i.e.
A.+2J.1~

u=<I>
.+1'-v~
..'1'~,J.+--v
. ..,
1
,I
A. + J.1 'V.-'V
~
IIJ<,..

(3.104)

the equilibrium equations, Eq. 3.66, can be satisfied exactly, provided that <I> satisfies
the equation

(3.105)

106

Foundations oj Solid Mechanics

that 'P j is a hannonic function, i.e.


(3.106)

"j=O
and that Xi is a bihannonic function, i.e.

(3.107)

V'X;=O

Thus Navier equations can be uncoupled by means of the potentials as defmed by Eqs.
3.104 to 3.107.
In Section 6.3, elastostatic displacement potentials appropriate for a homogeneous
isotropic domain infmitely extended radially (0::;; r < 00) are described. Other elastostatic displacement potentials such as those proposed by Papkovich (1932) and Neuber
(1934) appear in several books on solid mechanics, e.g. Sokolnikoff (1956).

3.6.4 Elastotiynamic Displacement Potentials


The introduction of the potential CI> as in Eq. 3.95 yields the dilatation as in Eq.
3.96, and the equations of motion (Eq. 3.66) for the case of zero actual body force as

(A.+2J.L)(VZCl,j=p(~~1j

(3.108)

Similarly, the potential 'P j introduced as in Eq. 3.98 gives the dilatation as in Eq.
3.99, and the equations of motion as

J.Iv"(V~")J=pEv{Z"l

(3.109)

Incorporating the results above, we can state that the displacement vector

uj = eIl,j + Ev,,'P",i

(3.110)

satisfies the equations of motion exactly, provided that ell and 'P j are solutions to the

wave equations
(3.111)
and
2..-.2>T.

c. V T

OZ'P

= at2

(3.112)

Here
(3.113a)
is the pressure wave speed, and
(3. 113b)
the shear wave speed. They are discussed further in Chapter VIll. So we can say that

107

Linear Elasticity

the equations of motion of a homogeneous isotropic elastic domain under zero actual
body force can be uncoupled into two wave equations by means of the potentials as
defmed by Eqs. 3.110 to 3.112.
In Section 8.8, we describe elastodynamic displacement potentials appropriate for
a homogeneous isotropic domain infmitely extended radially (0 :S r < 00 ). More general
forms of cylindrical and spherical displacement components in terms of appropriate
potentials have been listed by Achenbach (1973).

3.6.5 Positive Definiteness of the Strain Energy Function


Since the strain energy function must be positive for any non-trivial set of the
strain components, the stiffness tensor Cij in Eq. 3.22 must satisfy the conditions
stipulated by Eq. 1.73, i.e.
C ll C12 C13
Cll C121
Cll > 0, 'c
C > 0, C21 C22 Cn > 0, ... ,
21
22

C31 C32 C33

C13 C14 CIS C16


Cn

C24

C25

C33

C34

C3S C36
>0
C4S C46
Css CS6

C44
Symm.

Cu
(3.114)

C66
For orthotropic solids (Eq. 3.28), these conditions can be specialized into
Cll C 12

C 13

C21 C22 Cn > 0


C31 C32 C33

Cll > 0,
C44 > 0,

Css > 0,

C66 > 0

(3.115a,b,c)

(3. 115d,e,f)

Simplest among all, are the conditions for isotropic solids (Eq. 3.31), i.e.
1..1.>0, 3A.+21..1.>0
Equation 3.116b states that Poisson's ratio v is less than 112.

(3. 116a,b)

3.6.6 Stress and Strain Computation from Measured Results


At a point 0 on a surface enclosing an isotropic elastic domain with Poisson's
ratio v =0.3 , the axes x and y are drawn tangent to the surface and the axis z normal
outwardly to the domain. A surface traction of a magnitude 10 is applied at point 0

Foundations of Solid Mechanics

108

- - -..
_ x

Fig. 3.4
in the direction of a vector N j with (N 1,N2 ,N3 ) =(1,-1, 1). In addition, three strain
gages placed at point 0 give normal strains as shown in Fig. 3.4. Determine the stresses
and strains at point O.
Answer:
7.530 -0.444 5.774 ]
-5.774,
5.991
[cry] =[
Symm.
5.774

E[ey]

=[

-0.577
4
Symm.

7.506
-7.506.
1.718

3.6.7 Saturated Porous Elastic Media


Biot (1941) proposed the following theory for quasi-statics of porous isotropic
elastic solids completely saturated with fluids. In the absence of body forces, the
equilibrium conditions can be written

crIJ,J.+oIJ,}
. p/.=O
(3.117)
where crij is the effective stress tensor, and PI the excess pore pressure (positive if in

tension). The stress-strain relationship is in the usual form

crij =2J..lEij + oijAea

(3.118)

where A. and f..l are Lame's constants of the bulk material. The strains and the dilatation

109

Linear Elasticity

also take the common fonns, i.e.


P. ..

"

kk

1
2

=-(u. . +u .),
I,'

'oJ

=Ut,t.

where Ui is a displacement component of the solid matrix. The fluid flow is governed
by Darcy's law in the fonn

Ow

at -

-'-kpJ,i
where k is the coefficient of permeability, and
Wi =f(Ui - ui )

(3.119)
(3.120)

in which Ui is a displacement component of the fluid, and f the porosity, i.e. the
volume of the voids per unit total volume. The condition of perfect saturation with an
incompressible fluid implies that
(3.121)
kk=-Wi,i
So that Darcy's law becomes
(3.122)
Substituting Eq. 3.118 into Eq. 3.117 yields the Navier equations for the saturated
porous solid as

J.1~Ui + (1..+ J.1)kk,i + hi = 0

(3.123)

while Darcy's law (Eq. 3.122), due to Eq. 3.121, becomes

cV2
it

=akk

at

(3.124)

where c is the coefficient of consolidation, i.e.

c = (A.+2J.1)k

(3.125)

Note that the Navier equations for solids (Eq. 3.66) and for saturated solids (Eq. 3.123)
become identical if the body force Xi in the fonner is replaced by the excess pore
pressure gradient PI,i' Darcy's law can be put in another fonn after substituting Eq.
3.121 into Eq. 3.124 and then using Eq. 3.119, i.e.
(3.126)
The well-posed boundary conditions for the present case can be obtained by means
of a variational principle. They state that, at a certain point on the boundary, either
u i = 'iii or nj(CJij + SijPf) =)(

(3. 127a)

and, either
(3. 127b)

110

Foundations of Solid Mechanics

Here the bar ( - ) denotes a prescribed quantity, and Q the prescribed rate of the
outflow.
It can be seen that there are four basic unknown functions U 1 , Uz, ~ and PI'
and four governing field equations Eqs. 3.123 and 3.124. The problem becomes identical
to that of pure elastic solids if PI tends to zero, which happens if the system is subjected
to constant forces for a sufficiently large time. In Section 6.8, we discuss problems of
saturated porous elastic domains infinitely extended radially (0 S; r < 00 ).

CHAPTER IV
ELASTOSTATIC PLANE PROBLEMS

4.1 PLANE PROBLEMS OF ORTIIOTROPIC ELASTIC MATERIALS

For orthotropic elastic case of plane stress in the xy-plane, the three stresses on
the plane nonnal to the z-axis vanish as in Eq. 3.41. The remaining stresses are related
to strains as in Eqs. 3.42, with the coefficients Sij derived in Eqs. 3.52a to d. Altogether
there are eight field equations: two equilibrium equations, three stress-strain relations,
and three strain-displacement relations. The equilibrium equations are Eqs. 2.4 rewritten
in explicit symbols as
(4.1a)
(4.1b)
where X and Y are the body forces in the x and y directions, respectively. The
stress-strain relations are Eqs. 3.42. The strain-displacement relations are Eqs. 2.43,
i.e.
(4.2a)
(4.2b)

(4.2c)
Equally, there are eight unknowns: 3 stresses (a"", a)l)l' a..,), 3 strains (.a, e"" 1..,), and
2 displacements (u, v). In addition, since we will encounter some constants of integration in solving the differential equations, Eqs. 4.1 and 4.2, we will have to prescribe
some boundary conditions.
The boundary conditions, Eqs. 3.61, may be written
(4.3a)
X a""n" + a..,n, or u u

Y= a..,n" + a)l)ln,

or

v= v

(4.3b)

Foundations of Solid Mechanics

112

where a bar C) denotes a prescribed quantity on the boundary, and n" and n, are
directions cosines of the outward drawn normal to the boundary, thus

n;+n;= 1

(4.3c)

Rigid-body displacements (Eq. 2.53 for 3-dimensional cases) for plane problems are
u =Ay +B
(4.3d)

v =-Ax+C

(4.3e)

where A, B and C are arbitrary constants.


A material body is said to be in state of plane strain in the xy-plane if the displacement w vanishes and u and v are independent of z. It is found that all the
governing conditions derived already for plane stress problems hold, with the exception
that the stress-strain relations for plane strain should be derived from the general
relationship, Eqs. 3.39 with
(4.4)
=y,. =y", = 0

e..

This leads to the following relationship among the pertinent stresses and strains.

(4.5)

S
where Sij's are given in Eqs. 3.52. It should be noted that 0'0 in a plane strain problem
does not vanish.
For an isotropic elastic material, the stress-strain relations for both plane stress
and plane strain problems can be written in a form known as the unified Hooke's law
for isotropic plane problems, i.e.

~=2~ O';a-T(O';a+O''')J

(4.6a)

1 [3-K
1
e,,=2~ O'''-T(O';a+O''')J

(4.6b)

1
"(, =-0'

(4.6c)

1 [3-K

or, in reverse,

'"

'"

0';a=~1
[(K+ 1)~+(3-K)e,,1
K-

(4.7a)

0'" = ~I [(K+ 1)e" + (3 - K)e,J

(4.7b)

K-

(4.7c)

Elastostatic Plane Problems

113

where

K={(3 -V)/(l +V),

for plane stress


(4.8a,b)
3-4v,
for plane strain
Thus a solution to a plane strain problem with a Poisson's ratio v is equivalent to that
of a plane stress with a Poisson's ratio =v/(1-v).
4.2 AIRY FUNCTION FOR ISOTROPIC PLANE PROBLEMS
This is an approach using stresses as basic unknowns. Equilibrium is governed by
Eqs. 4.1, stress-strain relationship by Eqs. 4.6 or 4.7, strain-displacement relationship
by Eqs. 4.2, and boundary conditions are in the form of Eqs. 4.3. Among the 6
compatibility conditions stated in Eq. 2.62, we have to consider only the condition
iff;" cfe" ifyxy
-+---=0
(4.9)
dy2 dX2 dXdY
Other 5 compatibility conditions are satisfied identically in plane strain problems. For
plane stress problems, the other 5 compatibility conditions will be satisfied if

ife.z ife.. ife..

-=-=-=0

dX2

dy2

dXdy

(4.10)

or
(4.11)
where A, B and C are constants of integration. Such a solution in Eq. 4.11 is rather
an exception than a general solution. Thus the solution to a plane stress problem satisfying Eq. 4.9 is not exact, but should be a good approximation to a thin plate problem.
Substituting Eqs. 4.6 into Eq. 4.9 yields
(1 +K)V2(0' +0' )-2 d20'xy _ ifO'Ja _ if0'" =0
(4.12)
4
Ja"
dXdY dX2 dy2
For the present two-dimensional case, the Laplace operator (harmonic operator) V2 is

V2=~+~
dX2 dy2
Determining ifO'xy/dXdY from Eqs. 4.1, we get
ifO'xy d20'Ja if0'" dX dY
-2 dXdY = dX2 + dy2 + dX + dy

(4.13)

(4.14)

Substituting Eq. 4.14 into Eq. 4.12 leads to the following compatibility condition for
the normal stresses
2

4 (dX dY)
dX +dY

V(O'Ja+0'")=-1+K

If the body force is conservative, then it has a potential V, and

(4.15)

Foundations of Solid Mechanics

114

X=_dV
dX

(4. 16a)

y=_dV
dy

(4. 16b)

In fact, the condition for a conservative body force is

dX dY
dy =dX

(4.17)

Equation 4.15 becomes


(4.18)
In 1862, G.B. Airy proposed that the stresses be expressed in terms of another potential
cI>(x,y) by

(4.19a)
(4.19b)
(4. 19c)
These satisfy the equilibrium equations, Eqs. 4.1, automatically, and change the compatibility condition, Eq. 4.18, into
Vcl>=_2 K - 1 V 2V

K+l
where V4 is the plane biharmonic operator and defined by

(4.20)

(4.21)
V4 =V2V2
We note that cI> in terms of polynomials of x and y of lower degrees than two can
be neglected, since it gives zero stress components. If both body force components
vanish, then V can be set to zero without of loss of generality. Lastly, the tbreedimensional form of Eq. 4.20 is Eq. 3.91.
4.3

ISOTROPIC ELASTIC PLANE PROBLEMS IN CYLINDRICAL


COORDINA1ES

Various conditions in cylindrical coordinates are derived in Sections 2.10 and 3.5.
For plane problems in the re-plane, equations of equilibrium in the r and e directions
can be found from Eqs. 2.74a and 2.74b, respectively, as
dCJrr 1 dCJrS CJrr - CJee
-+--+
+X =0
(4.22a)

dr

r de

115

Elastostatic Plane Problems

(4.22b)
The strain-displacement relations are Eqs. 2.71, while the unified Hooke's law is the
same as in Eqs. 4.6 or 4.7 with r and e replacing x and y, respectively, i.e.

1[

(4.23a)

(4.23b)

3-lC
e"=2J.L 0"--4-(0,,+000)J

[3-K

00= 2J.L 000 - -4-(0" +OOO>J

(4.23c)
(4.24a)
0 00 = ~1 [(lC+ 1)00 + (3 -lC)err]
lC-

(4.24b)

0,0 =2J.Le,0

(4.24c)

The compatibility condition in terms of normal stresses is

4 (ax,
axo)
v2(0,,+000)=--- - +x,- +1l+lC ar r rae
In polar (plane cylindrical) coordinates, the Laplace operator
2

(4.25)

VZ is

la(a)
ta
r ar + r2 ae2
2

=-;: ar

(4.26)

Equation 4.25 is equivalent to Eq. 4.15 in Cartesian coordinates. For a conservative


body force, X, and Xo can be expressed in terms of a potential V as

av
x=-, ar

~n~

lav
Xo=--;:ae

(4.27b)

The condition to check whether the body force is conservative is

ax, a(rXo)
ae=a;:-

(4.28)

Substituting Eqs. 4.27 into Eq. 4.25 leads to


2

4
lC+

V (0"+000)=-1 V V

(4.29)

Foundations 0/ Solid Mechanics

116

The basic unknown stresses can be expressed in the following fonn


10cl> 1 OZ4>

a =--+--+V
2
2

" rar

ao

(4.30a)
(4.30b)

_i.(!
ar

=
Ocl
(4.3Oc)
re
rao
which satisfy the equilibrium equations, Eqs. 4.22, exactly and change the compatibility
condition, Eq. 4.29, into
K-1
(4.31)
~4>=-2-~V
K+1
Note that Eqs. 4.29 and 4.31 are identical to Eqs. 4.18 and 4.20, respectively.
Rigid-body displacements for plane problems in cylindrical coordinates can be
written
(4.32a)
Ur =CI cosO + C2 sin 0
a

(4.32b)
4.4 DISPLACEMENT FOR A GIVEN BIHARMONIC FUNCTION
In the absence of body forces, the Airy /unction 4> is governed by the plane

biharmonic equation
(4.33)

~4>=0
We seek a solution in the fonn of a Fourier series

4>=

.. =0

R.. cosmO+

.. =1

R:sinmO

(4.34)

in which R.. and R: are functions of the radial coordinate r only. Substituting Eq.
4.34 into Eq. 4.33, we obtain, for each m, an ordinary differential equation
where

~".R.. =O

v:. =V!V!, and

2
~.. =!~(r~)rdr dr
r2

The general solution of R.. is, for m

=0 ,

Ro =Ao + Bor 2 + Co log r + Dor 2 log r

(4.35a)

for m=1,
(4.35b)

117

Elastostatic Plane Problems

and for m>1,


(4.35c)
where the constants Am, Bm, Cm, and D". are arbitrary. Similar expressions can be
written for R~. However, we should not have the impression that only the functions in
Eq. 4.34 are plane biharmonic junctions. There are several other functions which are
also biharmonic, since they happen to satisfy Eq. 4.33.
If cjl is biharmonic, and
VZcjl=P(x,y)
(4.36)
then
(4.37)
which means that P is a plane harmonic junction. Consider an analytic junction
(Churchill, 1960) of the fonn
(4.38)
f(z) =P(x ,y) + iQ(x,y)
Z
where Q is another harmonic function, i is the imaginary number (i =-1) , and z
in this particular section is the complex variable, i.e.
(4.39)
z=x+iy
or, in polar coordinates,

z =rei8

(4.40)

We may introduce another analytic function 'I'(z) , the derivative of which is


: ' =f~)

(4.41)

Let 'I' be composed of the real and imaginary parts as follows


'1'= p(x,y)+iq(x,y)
(4.42)
Thus p and q are also harmonic functions, and d'l'/dz can be put in the fonn
d'l' =ap +iaq
dz ax ax

=aq _i ap
ay

ay

(4.43)

Note that we have used the Cauchy-Riemann conditions


ap aq
ap
aq
-=-,
-=-(4.44a,b)
ax ay ay
ax
in obtaining Eq. 4.43. The Cauchy-Riemann conditions in polar coordinates are
ap aq
aq
ap
=rae'
=- rae
(4.45a,b)

ar

ar

These functions p and q are also called conjugate harmonic junctions. Substituting
Eqs. 4.38 and 4.43 into Eq. 4.41 leads to
ap aq P
(4.46a)
ax =ay ="4

Foundations of Solid Mechanics

118

(4.46b)
The stress components for the case of
from Eqs. 4.19 with V =0, i.e.

<\>

governed by Eq. 4.33 can be obtained


(4.47a)

(4.47b)

('J

zy

= -if<\>
--

(4.47c)

axay

Substituting strain-displacement relationship and Eqs. 4.47 into the stress-strain relationship (Eqs. 4.6), in view of Eqs. 4.36 and 4.46a, leads to

au

if<\>

ap

21lax =-ax 2 +(1+lC)ax

av =_if<\>+ (1 + lC) aq
ay
ay2
ay
Jau av) if<\>

21l

(4.48a)
(4.48b)

(4.48c)
~ay +ax =-axay
Integrating Eq. 4.48a with respect to x, and Eq. 4.48b with respect to y leads to,
respectively
2JlU
21lv

=-:

+(1 +lC)p + gl(Y)

(4.49a)

=-: + (1 + lC)q + g2(X)

(4.49b)

where gl(Y) and g2(X) result from the integration. Upon substituting the preceding
equations into Eq. 4.48c, we find

dg 1

dg 2

dy

dx

-=--=A

(4.50)

where A is an arbitrary constant. Integrating the above equation leads to


gl(y)=Ay+B

(4.51a)

g2(x)=-Ax+C

(4.51b)

where B and C are also arbitrary constants. Recalling Eqs. 4.3d and e, we notice that
Eqs. 4.51 are rigid-body displacement components. Ignoring such components for the
time being, Eqs. 4.49 become

Elastostatic Plane Problems

119

2J.I.U

=- ~ + (1 + x:)p

(4.52a)

2j.1v

=- ~~ + (1 + x:)q

(4.52b)

With the relationship among cylindrical and Cartesian coordinates as listed in Eqs. 2.65
to 2.68, we may write displacement components in the cylindrical coordinate system as
2J.I.U,

=- ~ + (1 + x:) (p cos e + q sin e)

2J.I.Ua=- r~ +(1 +x:)(-p sine+q cos e)


Let a real function Pl be denoted by
Pl =<I> - ~[z'l'(z)]
where ~ denotes the real part and

z the conjugate of z, i.e.


z=x-~

Applying

(4.53a)
(4.53b)
(4.54)
~5~

VZ on Eq. 4.54, in view of Eqs. 4.36, 4.42 and 455, yields


V 2Pl =p -[XV2p +yV2q

+2(~ +~~].

or, due to Eqs. 4.46,

V 2Pl =-xV2p _yV2q.


Moreover, the fact that VZp = V 2q = 0 reduces the last equation into

V 2Pl =0

(4.56)

Thus Pl as introduced is a harmonic function.


Then Eq. 4.54 can be rewritten in the fonn
<I> =~[z'l'(z)] + Pl

= ~[(z + z)'I'(z)] - ~[z'l'(z)] + Pl'


or, due to Eqs. 4.39, 4.42 and 455,
(457)
where P2=-~[(Z'l'(z)] +Pl thus is a harmonic function. Alternatively, we can write
Eq. 4.54 in the fonn
<I> =~[(z - z )'I'(z)] + ~[z'l'(z)] + Pl

=2yq+P2

(4.58)

where P2 =~[(z'l'(z)] + Pl thus is also a harmonic function. Equation 458 is another


general fonn of biharmonic functions. On the other hand, if we rewrite Eq. 4.54 as
follows

120

Foundations of Solid Mechanics

we would arrive at
(4.59)
where
P3

=9t[",(z)lz]
=9t[(PX +qy)+i(-py +qx)]/r2

=9t[(P cos9+q sin9)+i(-p sin9+q cos9)]/r


= 9t(P3 + iq3)'

where
P3

=(P cos9+q

q3

=(-p sin 9 + q cos 9)1r

sin9)1r

(4.60a)
(4.60b)

or,
P =Pyc-q3Y

(4.61a)

q =P3y+q,x

(4.61b)

Since P3 and q3 are real and imaginary parts, respectively, of the same analytic function
[",(z)lz] , they are conjugate harmonic functions and related by the Cauchy-Riemann
conditions. Incorporating Eqs. 4.61 and 4.60 into Eqs. 4.52 and 4.53 respectively, we
have
2JlU
2J.1v

=-:

+(1 + lC) (p,x -q3Y)

(4.62a)

=-~ +(1 + lC) (P3Y +q,x)

(4.62b)

2JlUr=-~ +(1+lC)rP3

(4.63a)

2J.1Ue =- r~ + (1 + lC)rq3

(4.63b)

Thus when a harmonic function takes the form of Eq. 4.59, q3 can be detennined from
P3 by simple integration of the Cauchy-Riemann conditions (Eqs. 4.44 or 4.45), then
the displacement components can be simply obtained by Eqs. 4.62 or 4.63. For the case
of biharmonic functions in the form of Eq. 4.57, q should be detennined by the
Cauchy-Riemann conditions, and the displacement components by Eqs. 4.52 or 4.53.
The case of biharmonic functions in the form of Eq. 4.58 can be treated in the same
way as the case of Eq. 4.57 with the exception that P must be detennined from the
known function q using the Cauchy-Riemann conditions. Arbitrary constants arising
from the integration of the Cauchy-Riemann conditions correspond to rigid-body

Elastostatic Plane Problems

121

displacement components, Eqs. 4.3d and e for Cartesian components, and Eqs. 4.32 for
cylindrical components.
Table 4.1 at the end of the present chapter, contains some common harmonic and
biharmonic functions and their corresponding stresses and displacements (but not
rigid-body components). When these functions are combined properly, we have new
harmonic functions. Here are some harmonic functions which can be used in some half
plane problems.
r(1ogr sin6+6cos6) =y logr +x6
(4.64a)

r(1ogrcos6-6sin6) =x logr -y6

(4.64b)

=
rZ(logr cos 26 - 6 sin 26) =(x z - yZ)logr - 2xy6

rZ(1ogr sin 26 + 6 cos 26) 2xy logr + (x z-l)6

(4.64c)

(4.64d)

One may notice that each term in the equation above is a product of two harmonic
functions, hence it may be found expedient to note that a Laplacian of a product of
harmonic functions il(x , y) and Iz(x, y) can be put as

VZu n = J oil olz + oil olz]


V]J2I
4. ox ox oy oy

(4.65)

The total solution to Eq. 4.20 or 4.31 can be a linear combination of complementary solutions and a particular solution. A complementary solution is a biharmonic
function; while a particular solution is one that satisfies the governing equations exactly
without involving any arbitrary constants. In addition to complementary solutions, Table
4.1 also contains particular solutions and their corresponding stresses and displacements
for various common potentials V. The arbitrary constants, some being attached to
complementary solutions while the rest associated with rigid-body displacement components, are to be determined from the condition that the total solution must satisfy all
pertinent boundary conditions.
4.5 EXAMPLES OF INFINITE PLANE PROBLEMS
The first problem to be considered is an infinite plane under uniform tension in
the x -direction as shown in Fig. 4.1. Since there is no body force (X Y 0), we can
set

= =

V=O.
The stress field everywhere is
<1""

So that the Airy stress function

~cI>

=<10 ,

(1)/)/

=0 ,

<1..,

=O.

cI> of Eq. 4.19 satisfies


~cI>

~cI>

oyZ =<10 , oxz =0 , oxoy =0,

Foundations of Solid Mechanics

122

L...-_ _ _ _ _

+-_---O_ X

Fig. 4.1 Infinite plane under unifonn tension.


the solution to which can be shown to be

y2

~=O'0"2+CIX +C2Y +C3,

where C1 to C3 are constants of integration, and can be set to zero since they yield
zero stresses. Thus
(4.66a)
In polar coordinates,

~= ~0(r2-r2coS26)

(4.66b)

and the stresses (see Table 4.1) are


(4.67a)
0'0
0'r9 =--sin28
2

(4.67b)

0'0
0'99 ="2 (1

(4.67c)

- cos 26)

The second example is an infmite plane with a circular hole and under unifonn
tension far away from the hole as shown in Fig. 4.2. The stresses at the far field
(r --+ 00) are identical to those in the previous example. Thus, the complementary
solution may be written

Elastostatic Plane Problems

123

Fig 4.2 Infinite plane with a hole under unifonn tension at infinity.
where cj)1 is the function cj) of the previous example, i.e.
cj)1 = ~o (r2 - r2 cos 20),
and cj)2 should be such that it gives zero stresses as r tends to infinity, and leads to
the satisfaction of the stress free boundary conditions at r =a, i.e.
(4.68a)
O'"(a,O) =0
O',e(a,O)

=0

(4.68b)

Considering all these requirements and after consulting Table 4.1, we fmd
cj)=

~O~2_r2cos20+Aa210gr+Ba4CO;220 +ca 2cos20)

(4.69)

where A, B and C are constants to be determined by substituting Eq. 4.69 into Eqs.
4.68, i.e.
2+2cos20+A -6B cos20-4C cos 20 =0,
-2 sin 20 + 0 - 6B sin 20 - 2C sin 20 =O.
Since the two equations hold for every value of 0, they imply the following three
independent conditions
2+A =0,
2-6B-4C=0,
-2-6B -2C =0.

124

Foundations of Solid Mechanics

Finally, we fmd

c1>= a40[ r2-r 2cos28-2a 210gr- ar2 cos 28 + 2a 2 cOS28J

(4.70)

(1 =(1[I_ a2 +(1_4a2+3a4)cos2el
"2
r2
r2
r4

(4.71 a)

(1

r9

a2 +3 a4 )Sin28
=(1(_1_2
2
r2
r4

(1ee =

~o[ 1 + ;: - (I + 3 ;:)cOS28]

The hoop stress, which is (1ee at r =a , is


(1ee(a,8) =(10(1- 2cos28)

(4.71 b)
(4.71 c)

(4.72)

Its largest tensile and compressive values are shown in Fig. 4.2, i.e.

(1ee(a,O) = (1ee(a, n) =-(10


(1ee(a,1tI2) =(1ee(a, -1tI2) =3(10
Thus the stress concentration factor is 3.

(4.73a,b)
(4.73c,d)

Such a factor would be increased to 4 if a uniform compression in the y-direction


is added to the same plate as shown in Fig. 4.3.

0"0

Fig. 4.3

Infinite plane with a hole, under tension in one direction


and compression in another orthogonal direction.

Elastostatic Plane Problems

125

~--~-----+--~x

Domain I

Fig. 4.4 Infinite plane subjected to unifonn body force.


The last example to be considered in this section is the problem of an infinite
plane subjected to a unifonn body force in a circular region as shown in Fig. 4.4. Let
10 be the unifonn intensity of the body force in the x-direction in domain 2, where
o S; r < a. Domain 1 is the region r > a. Thus for domain 2 we have
X(2) =10,
y(2) =O.
Note that a superscript in the parentheses is used to denote a domain number. A simple
check by Eq. 4.17 shows that the body force is conservative, and Eqs. 4.16 give
ay(2) ay(2)

ax =-10, ay

=0

(4.74a,b)

A general solution to Eqs. 4.74 is y(2) = -foX + C, where C is an arbitrary constant


which can be set to zero without loss of generality, i.e.
y(2)=-!oX =-lorcos9
(4.75)
According to Eqs. 4.30 or Table 4.1, the stresses due to y(2) alone are
(4.76a)
(4.76b)
Gee =

-lor cos 9

The unifled Hooke's law gives the strains


au,
lC-l
aT = -10 4J.L r cos 9

(4.76c)

(4.77a)

126

Foundations of Solid Mechanics

laUe U,
K-l
--+-=-,h--rcOSa
raa r
4f.1
Integrating Eqs. 4.77 yields u, and Ue as

2f.1U, =-4(K-l)for2cosa

(4.77c)

(4.78a)

2J.lUe=-.!.(K-l)for 2 sina
(4.78b)
4
which are the same as listed in Table 4.1. While V(I) vanishes, V(2) does not but
VZV(2) = 0; hence the governing equation for both domains 1 and 2 is V41> = O. The
conditions at the interface (r = a) between these two domains are
d~)(a, a) = d;:(a, a)

(4.79a)

d.~(a,a)=d.~(a,a)

(4.79b)

u~I)(a,a)=u;Z)(a,a)

(4.79c)

u~I)(a,a)=~)(a,a)

(4.79d)

The pattern of stress and displacement fields in domain 2 due to fo is given in Eqs.
4.76 and Eqs. 4.78. In order to satisfy Eqs. 4.79 we need to investigate the following
biharmonic functions listed in Table 4.1
r3 cos a, ra sin a, r log r cos a, cos air.
Since the stresses must be bounded as r ---+ 00 in domain 1, and as r ---+ 0 in domain
2, we take
1>(1)= foa \2Arasina +Br logr cos a + ca 2Co;a)

(4.80a)

I>\l)=fJJr3cosa

(4.80b)

where A , B, C and D are constants to be determined from the boundary conditions


Eqs. 4.79. From the table, we can put the stresses and displacements for the domain 1
as

d,? = foa 2(2A cos air +B cos air -

2C a 2cos 81r 3 )

(4.81a)

d.~ = foa 2(B sin 81r - 2C a 2sin 81r 3 )

(4.81b)

d~=foa2(B cos81r +2Ca 2cos8Ir 3 )

(4.81c)

Elastostatic Plane Problems

2J.1U?)

127

=foa 2{[A(K-1) + B(K+ 1)] 0 sin 0/2


+[A(K+1)+B(K-1)] logrcosO/2-(A +B)cosO/2

(4.81d)

2~1) =foa 2{[A(K-1) + B (K+ 1)] cos 0/2

-[A(K+ 1)+B(K-1)] logr sinO/2-(A +B)sinO/2

+C a 2sin 0/r2 - C~I)sin O}

(4.81e)

in which C~I) corresponds to the appropriate rigid-body displacement. In domain 2, the


Airy function is as given in Eq. 4.80b, for which the stresses and displacements can be
obtained from Table 4.1. These quantities together with those due to the applied body
force 10 in Eqs. 4.76 and 4.78 are listed below.
d~) =10(20 -1)r cos

(4.82a)

d~=2foDr sin a

(4. 82b)

d~=fo(6D -1)rcosa

(4.82c)

2J.1U~2) =fo{[D (K- 2) -~ (K-1)}2 cos a+ a 2C?) cos a}

(4.82d)

2J.1U~)=fo{[D(K+2)-~(K-1)}2sina-a2C~2)sina }

(4.82e)

Before applying the boundary conditions, a closer look at Eqs. 4.81d and e reveals
multivalued displacement terms, 0 sin a and acos a, which are not admissible in the
present problem. In order to delete such terms, we should set their coefficients to zero,
i.e.
A(K-1)+B(K+ 1) =0
(4.83)
Substituting the general expressions of stresses and displacements into the boundary
conditions Eqs. 4.79, and using Eq. 4.83, we find
2A+B-2C=2O-1
B -2C =20

(4. 84a)
(4.84b)

1 1 1
"2 [A(K+ 1)+B(K-1)] loga -"2 (A +B)+C + (C~I)_C~2~ =D(K-2)-"4(K-1) (4.84c)
1 1 1
-"2
[A(K+ 1)+B(K-1)] loga -"2 (A +B)+C _(C~I)_C~2~ =D(K+2)-"4(K-1)

(4.84d)

Foundations of Solid Mechanics

128

Equations 4.83 and 4.84 fonn a set of 5 linear simultaneous equations for detennining
5 constants, i.e. A, B, e, D and ef 1) - ef2). The solution is
1
lC-1
A =-2"'
B =2(lC+ 1)
(4.85a,b)

e=

1
4(lC+ 1) ,

D=

lC
4(lC+ 1)

(4.85c,d)
(4.85e)

The tenn ef 1) - ef) signifies the indefinite displacement. The displacement has a
'weak' singularity as logr as r tends to infmity. Such peculiar situation of the
indefiniteness and unboundedness of the displacement occurs only in statics of infmite
planes.
The solution for a concentrated force P applied to an infmite plane, as shown in
Fig. 4.5, can be obtained by taking the limit a --+ 0, and assuming that
Lim(1ta 2fo) =P.
a ..... 0

Equation 4.80a gives the limiting potential

=21t(:+ 1) [-(lC+ l)re sin e + (lC-1)r logr cos e]

<l>

(4.86)

the stresses and displacements

cr
"

(lC+ 3)P cos e


21t(lC+ 1) r

(4. 87a)

(lC-1)P sine
cr - --'---"'--r9 - 21t(lC+ 1) r

(4.87b)

(lC-1)P cose
cr - --'---"'--ee-21t(lC+1) r

(4.87c)

y
}

p-~
o
Fig. 4.5 Force at a point inside infmite plane.

129

Elastostatic Plane Problems

2J.1Ur = 21t(1C+ 1) (-21Clogr + 1)cos9+ C1 cos9

(4. 87d)

2J.1Ue = 21t(:+ 1) (21Clogr + 1) sin 9 - C1 sin 9

(4.87e)

Observe that the stress decays as r-1 when r increases and is singular at the origin.
The displacement has a weak singularity of logr both at the origin and at infinity.
4.6 PARTICULAR SOLUTIONS FOR CONCENTRA1ED FORCES
The function <I> given by Eq. 4.86 can be taken as a particular solution to any
problem of a concentrated force inside a plane, whether finite or infinite. The first term
gives the appropriate singularity at the point where the force is concentrated, while the
second term suppresses the multivaluedness of the displacement components due to the
first term.
In a problem where the concentrated force is applied at a point on a smooth
boundary of the material domain as shown in Fig. 4.6, a particular solution can assume
the form
(4.88)
<I>=Ar9sin9
The constant A must be determined from the conditions that the force resultant in the
x -direction is P and that in the y -direction is zero, i.e.
Limr (It-a X (r,9)d9+P =0

(4.89a)

Limr (lt Y (r,9)d9=0

(4. 89b)

r-+O

r-+O

J-<l

J-<l

Here X(r,9) and Y(r,9) are components of the stress vector on the boundary where
r is a constant, i.e.

p=-----'----+---+--x

,"

""

Fig. 4.6 Force at a point on smooth boundary.

Foundations of Solid Mechanics

130

p----,~-,.--

Fig. 4.7 Force at a cornered boundary.


X(r,e) = CJ",,(r, e) cos e + CJ",(r, e) sine

(4.90a)

Y(r,e) = CJ11 (r, e) sine + CJ",(r, e) cos e

(4.90b)

In writing Eqs. 4.89, we have incorporated the fact that ell as given by Eq. 4.88 yields
CJre = CJee = 0 everywhere. The Cartesian stress components according to Table 4.1 are

(4.91a,b)
CJ 11

= 2A cose(1-cosZe)r-1

(4.91c)

Substituting Eqs. 4.91 into Eq. 4.89a yields A = -P lx, while Eq. 4.89b is satisfied
automatically. Thus a particular solution is
eIl=_P re sin e

(4.92)

For a concentrated force acting at a cornered boundary, as shown in Fig 4.7, a


particular solution can be found to be

ell =

pr:

l-cos2 -2

pz{[Sin2(P-a)+sin2a-2P.1 sine
+ [cos2(p-a)-cos2a] cose}

(4.93)

11Us satisfies the conditions near the point of application of the force, i.e.
Limrill-uX(r,e)de+p =0
r-+O

(4.94a)

131

Elastostatic Plane Problems

p----"'C~---f-<-

(a)

(b)

Fig. 4.8 Concentrated force P in special alignments with the cornered


boundary.
Lim r l

~-a-

Y(r,e)de =0

-a

r-+O

(4.94b)

If ~ = 1t, the boundary becomes smooth, the problem is as shown in Fig. 4.6, and
Eq. 4.93 is specialized into Eq. 4.92.
If the force P has special alignments as shown in Figs. 4.8a and b, particular
solutions are

'" _ 2P(~ - sin ~)re sin e


'I' 1 _ cos 2~ _ 2~2 ' for Fig. 4.8(a)

(4.95)

= 2P(~+ sin ~)recose, for Fig.4.8(b)

(4.96)

c!>

l-cos2~-2~2

Putting ~ = 1t in Eqs. 4.95, we can get a particular solution for the case of a
concentrated force normal to a smooth boundary, as shown in Fig. 4.9, namely
c!>

=-~resine
1t

(4.97)

On the other hand, Eq. 4.96 can be specialized to give a particular solution for
the case of a shear concentrated force on a smooth boundary, as shown in Fig. 4.10,
namely
P
(4.98)
c!> =-recose
1t

Foundations of Solid Mechanics

132

P---r-~--1-_

Fig. 4.9 Concentrated force nonna! to smooth boundary.

Fig. 4.10 Shear concentrated force on smooth boundary.


4.7

EXERCISE PROBLEMS

With permission o/the respective publishers, Problems 4.7.5,4.7.10 to 4.7.14 were taken
from the book by Boresi and Lynn (1974), and Problems 4.7.18, 4.7.19, and 4.7.21 to
4.7.28 from the book by Timoshenko and Goodier (1970).
4.7.1

Show that the governing equation of an axisymmetric problem without any


body forces has the fonn

!~{
r~[!.!(r ~]} =O.
rdr dr rdr dr
Solve the problem of a thick-walled cylinder whose outer surface is fixed while
the inner surface is subjected to a unifonn fluid pressure.

133

Elastostatic Plane Problems

Hint: Integrating successively yields


~ =Ar210gr +Br2+C logr +D,
A =0, since r 210gr gIves a multivalued Ug, and D =0 since it contributes

zero stresses. The remaining constants B and C are to be determined from


boundary conditions.
4.7.2

A composite thick cylinder is made of two different isotropic elastic materials.


The radii of the hole, the interface and the outer circumference are a, b and
c respectively. Solve the problem when the pressure inside the cylinder is p.

4.7.3

A steel disk of radius b is welded to a rigid concentric circular core of radius


a. A magnetic pole is placed at the center of the core. The magnetic force per
unit volume of steel is equal to k times the reciprocal of the square of the
distance from the pole. Is the magnetic force conservative? If so, fmd the
solution.
Hint:
k
k
V=-
Xr =-r2'

.=

r'

2(lC-l)kr
2
(lC+ 1) +Br +Clogr.

Show that due to the first term of above and V,


k(lC-l)
ur
f.L{lC+ 1) .

4.7.4

Solve the plane stress problem of a ring under a uniform shearing stress applied
at the outer surface of the ring while its inner surface is fixed. Hint: The Airy
stress function. is a function of 9 only.

4.7.5

For a ring loaded as shown in Fig. 4.11, the Airy stress function has the form

=(Ar2+Br4+ ~+D )cos 29 +Fr2+H logr.


Determine the constants A, B, C, D, F and H.
4.7.6

An infinite plate whose middle plane is xy has a circular rigid core of radius
a. If the plate is subjected to a uniform tension 0'0 at x =CO, what are the
slresses and displacements? Hint:

.= :0~2_r2cos29+Aa210gr+Ba4CO;;9

+ca 2cos29}

Set 3 equations for determining 3 unknowns A, B and C.

134

Foundations of Solid Mechanics

Fig. 4.11
4.7.7

Solve the previous problem, if the core is not rigid, but made of an isotropic
material with different properties from the plate. Hint:" For domain 1, a < r < 00,

,.) =~r -r'cos28 +Aa'logr +Ba'oo;;e + Ca'cos 28}


For domain 2, 0 _ r < a,
cjl(2)=

~O(DrZ+FrZcos2a+H ::cos2aJ.

Obtain 6 equations for the 6 unknowns A, B, C, D, F and H.


4.7.8

A heavy circular isotropic elastic disk of mass density p, radius b and thickness
a. The core supports the disk
in such a way that the disk is in the xy-plane and the gravitational force is in
the x-direCtion. Find stresses and displacements. Ifthe support becomes a point,
what is the solution? Hint:
v =-for cos a,
,
.
zcosa
r3
)
cjl=/oa lArasma+BrlogrCOsa+Ca -r-+ D azcosa .
h; has a rigid concentric circular core of radius

Obtain 5 equations for the 5 unknowns A, B, C, D and CI , where CI is a


rigid-body coefficient in Eqs. 4.32.
4.7.9

A homogeneous isotropic elastic plane consists of two regions, as shown in

135

Elastostatic Plane Problems

Fig. 4.12. The inner region is subjected to a body force in the x-direction equal
to fo per unit volume. The edge of the plane is fixed. Find the stresses and
displacements. Hint: For domain I, a < r < b,

()

2(

2COSe

r3

cl>1 =foa Aresme+Brlogrcose+Ca -r-+ D a2COSeJ

For domain 2, 0 S; r < a,

v =-for cos e,
=foF r3 cos e.

cI>(2)

Obtain 7 equations for the 7 unknowns A, B, C, D, F, C~l) and C~2).

'f---f-----+---x

y
Fig 4.12
4.7.10

Pure bending of curved bars. The Airy stress function


cI> =A logr +Br210gr +Cr2+D

may be used to study the problem of pure bending of curved bars (Fig. 4.13).
The constants A, B and C are determined from the boundary conditions

=(J"(b, e) =0,
b (Jea<ir =0,
b (JooTdr =-M,
(J"(a, e)

(Jre =0, on all boundaries.


The constants are given by

4M

=-Na b
22

log

a'b

Foundations of Solid Mechanics

136

Fig. 4.13

2M 2 2
B=-N(b -a),
M

C =N [b 2 -a 2 +2(b 2 10gb -a 2 10ga)],


N

4.7.11

=(b 2 _a 2)2 -4a2b110g~

Circular cantilever beam. The Airy stress function


cjl(r, 0) =f(r) sinO

may be used to study the problem of the circular cantilever beam subject to
end shear (Fig. 4.14), with

f(r)=Ar 3 +B +Cr+Drlogr.
r

The stress components are


(J

"

=(2Ar- 2B + D)SinO

r3

'

137

Elastostatic Plane Problems

Fig. 4.14

Fig 4.15

With boundary conditions,

a" = are = 0 at r = a, b
and the end shear condition,

we get

It may be shown that the Airy stress function


<1> = !(r)cose
yields a solution to the circular cantilever beam subjected to end tension T and
end moment M (Fig. 4.15). By appropriate superposition we may obtain
solutions of the problems illustrated in Figs. 4.16 and 4.17.
4.7.12

A circular cantilever beam is loaded in pure bending (Fig. 4.18). Determine


the displacement of the end (point A). For r = (a + b )/2, e = 0, let the radial
and tangential displacement components vanish,

aU

ur=ua=O and ae =0.

Foundations of Solid Mechanics

138

Fig. 4.16

Fig 4.17

Fig. 4.18
4.7.13

A thick rectangular plate is rolled into a cylindrical shape (Fig. 4.19). Residual
stresses due to the rolling process are removed by annealing. After annealing,
the end planes are a small angle ex. apart. The end planes are then brought
together by applying a moment M to each plane, and the faces are welded
together. Then uniform internal pressure Pi and external pressure Po are applied
to the lateral surface of the cylinder. Derive expressions for radial and tangential
stress components (J" and (Jae. Hint: Before applying pressure, show that the

Elastostatic Plane Problems

139

Fig 4.19

Fig 4.20
solution is the same as Problem 4.7.10, with the moment
Jl.cxN
21t(b 2 - a 2)(1C + 1)
where N is as given in Problem 4.7.10
M=

4.7.14

For a circular ring with a radial slit, B, as shown in Fig. 4.20, the function
r log r sin e is one of the possible candidates for the Airy stress function <1>.
Find the stresses and displacements. Hint: Try
. e +Bre cos
<I>=Arlogrsm

sine D r 3 . e
e+ c a 2 --+-2
sm .

Foundations of Solid Mechanics

140

Show that
A

=2J..l0l[7t(1C+ 1)],

=0,

C =Ab zl[2(a 2+ b 2)] , D =-Aa 2/[2(a 2+b 2)].


Find Us, the shear force P, the nonnal force T and the moment M, at each side
of the slit. Note that the two rigid-body displacement constants can be
detennined by setting Us at the slit to zero at two points only, not at every
point.
4.7.15

Nonnal point load on edge of half plane. The problem of the point load P on
the half plane boundary (Fig. 4.21), with zero stresses as r ~ co, may be
analyzed by means of the stress function
cI>(r,

4.7.16

a) = -~7t ra sin a.

Shear point load on edge of the half plane (Fig. 4.22). Show that
cI>=Q recose
7t

is the solution.

y-~~=----"..--...,

x
Fig. 4.21
4.7.17

Fig 4.22

Verify Hertz's solution


cI>

=i(-r1e 1 sine1 + r2e2 sin e2 + ;;).

where the notation is shown 10 Fig. 4.23. Find its

O'yy

along the x-axis. Answer:

O'yy(x,O) = P 1(7ta), constant tension.

4.7.18

A circular isotropic elastic plate is subjected to two equal but opposite forces
as shown in Fig. 4.24. The solution can be written in the fonn
cI>=-i[r1e1sine1-rze2Siriez+A

~].

141

Elastostatic Plane Problems

r-------2 a -+-------!.:::------'<;--t-- y

L_ _ _ _ ~

Fig. 4.23

""----+---y

x
Fig. 4.24
Detennine the constant A from the boundary conditions at r =a. Find also the
hoop stress cree(a, e). Answer:
P sin2a
A =_ cos a
cree(a,
e)
=
(.
e . )
2 '
1ta

sm +sma

142

4.7.19

Foundations of Solid Mechanics

Verify that the Airy stress function

~=-~{a
r sina-.!.(I-V)rlogrcosa-.!.rasina+~logr _ d (3-V).!.cosa}
1t
4
2
4
32
r
2

is the solution of the problem of a force P acting in a hole in an infInite plate


as shown in Fig. 4.25. Furthermore, show that the circumferential stress around
the hole is
P
1td [2+ (3 -v) cos 8]
except at the point where the force is applied, and that the function ~ corresponds to single-valued displacements. Hint: The question whether ~~ = 0
should be answered only for the term air sin a, which is equal to alr, sin a"
Thus ~(alrl sin a,) = 0, where V4 is a biharmonic operator in the rial-system.
Note that only 0"1'1 is the non-zero stress due to this stress function term. In
order to settle the question of single-valuedness of the displacements, we need
to examine the terms
91r, sin 91 -~(I-V)r logrcos9 -4r9Sin9,
which give possible multivalued displacement components proportional to

11

U,(9,9 1) }
{Sin9} [
{ ----- = 4(lC-l)91 - - - + -~(I-V)(lC+ 1)-~(lC-l)J
Ue(9, 91)
cos 9

sinal
---.
cos 9

Noting that v = (3 -lC)/(lC+ 1) for the present plane stress problem, we can
show that

Fig. 4.25

Elastostatic Plane Problems

143

a1) =u,(a + 2mt, a1 + 2nx),


Ue(a, a1) =Ue(a + 2nx, a1 + 2nx),
u,(a,

where n is an integer. Thus there is no multivalued displacement.


4.7.20

The Airy stress function

~ = ~[rlal cosa +~~: :; rlogr sina- ~ racosa+Cad +Dd2r-1sina]


1

is the solution to the problem of a force acting tangentially in a hole of an


infinite plate as shown in Fig. 4.26. Find the constants A, B, C and D from
appropriate conditions. Furthennore, fmd the circumferential stress around the
hole. Hint: For singlevalued displacements,
Answer: A =B = 1, C = 114, D

2-A -B =0.

=K/[8(K+ 1)].

Fig. 4.26

PIIIIIII~

; "8=.rc tan !
Fig. 4.27

4.7.21

Show that the stress function

144

Foundations of Solid Mechanics

Y ]

P[

cjl=-- (x 2 +y 2)arctan--xy

2n
x
is the solution to the problem of a half plane (-00 < X < 00, ~ y < 00) subjected
to a unifonn pressure p on the surface y =0, as shown in Fig. 4.27. The load
extends indefinitely to the left.
4.7.22

Show that the function


cjl =![.!.y210g(X2 + y2) +xy arctanI-l1
n 2
x

is bihannonic, and is the solution to the problem of a half plane subjected to


a unifonn shear load on the surface y =0, as shown in Fig. 4.28. The load
extends indefinitely to the left. Hint: Note the function defmed by Eq. 4.64a
is hannonic.

Fig. 4.28
4.7.23

By superposition, using the results of Problem 4.7.21, obtain stress and displacement components for pressure p on a segment -a < x < a on the surface
of the half plane. Note the indefiniteness of the displacement, and the 'weak'
singularity of the displacement as r tends to infmity. Answer:
p
~2
(J =
1t1
n [(x _a)2+ y~ [(x +a)2+ y~

4.7.24

If the pressure loading of Problem 4.7.23 is replaced by a shear load s, fmd


stress and displacement components.

x
y
Fig. 4.29

145

Elastostatic Plane Problems

4.7.25

Show that the function

ell =L[(.!x 3+xy2Iarctanl+'!y310g(x2 + y2) _.!X2y]


2xa3
)
x3
3
is biharmonic and is the solution to the problem of a half plane subjected to a
linearly increasing pressure p which extends indefmitely to the left, as shown
in Fig. 4.29. Hint: Note the functions defined by Eqs. 4.64a and c are harmonic.
4.7.26

Show that if the pressure loading of Problem 4.7.25 is replaced by a shear load

s, the appropriate stress function is


'" =_S_[Xy210g(X2+y2) + (x 2y -l)arctanl_3Xy21.

't' 2xa
.
x
J
Hint: Note that the functions defmed by Eqs. 4.64b and d are harmonic.

4.7.27

Show how the distribution of load indicated in Fig. 4.30 may be obtained by
superposition of loadings of the type shown in Fig. 4.29.

p
I-a-----J

/1

I~

I-a+-b-/-a-l

I-a+a-l

Fig. 4.30
4.7.28

Show that the solution to the half plane subjected to the parabolic loading as
shown in Fig. 4.31 is given by the stress function

{xy3

2 2(

y2 ]

y2J}

2
2 1
p --log-r; [a-+-(x
2
1 ( 1--+x2
-+y) 1 -x- + - a+-axp+-ay
2
2
2
1t
3a
r?
42
6a 2a
32
3a 2 a 2
for pressure, and

Fig. 4.31

Foundations of Solid Mechanics

146

+~ayp+~(x2-3y2-3a2)a+
4xy2}
2

!{L(3a2_3x2+y2)lOgr;
1t 6a 2
rt 3
for shear, where

a= 91 -92 = arctan

2ay
2

X +y -a

2'

3a

3a

Elastostatic Plane Problems

147

Table 4.1 Complementary and Particular Solutions for Elastostatics of


Isotropic Planes.
Complementary solutions, V4<1> =O.
I>

A..

TO

0 00

2JtuT

reos6

r sin 6

(lC-1)r

r2
logr

lIr2

-1/r2

-lIr

1/r2

-l/r

r210gr

210gr + 1

210gr+3

(lC-l)r 10gr-r

(lC+ 1)r6

rZO

26

-1

26

(lC-1)r6

-(lC+ l)r 10gr

cos61r2

sin61r2

cos61r

-2 sin 61r 3

2eos61r

-2sin61r 3

2eos6lr]

2sin61r 3

sin61r 2

-eos6lr 2

-2 cos 26

2 sin 26

2cos26

-2reos26

2r sin 26

-2 sin 26

-2 cos 26

2 sin 26

-2rsin26

-2rcos26

--<icos 261r 4

--<i sin 261 r 4

6 cos 261r 4

2eos261r

2 sin 261r 3

--<isin261r 4

6 cos 261r 4

6 sin 261r 4

2 sin 261r 3

-2 cos 261r 3

cos 26

-4 cos 261r 2

-2 sin 261r 2

(lC+ 1) cos 261r

-(lC-1)sin26Ir

sin 26

-4 sin 261r 2

2 cos 261r 2

(lC+ 1) sin 261r

(lC-1)eos26Ir

r 4 eos26

6r 2sin 26

12r 2eos26

-(3 -lC)r 3 cos 26

(3 + lC)r 3 sin 26

r 4 sin 26

--<ir2cos26

12r 2sin26

-(3 -lC)r3 sin 26

-(3 + lC)r 3 cos 26

sin61r
r 2eos26
r 2sin26

cos261r 2
sin 261r 2

-2cos61r

Hannonie functions, V21> = 0 .

cosn9
r"

Hannonic functions, ~ =O.

r"
r"

n(n -1)cosn9

(n + 2)(n -1)sinn9

r,,-2

sinn9

n(n -1)sinn9

(n +2)(n -1)cosn9
r"

r"+2

+ l)n cos n9

(n + l)n sinn9
r"+2
(n

(n + l)n sinn9
r"+2

r"

r"+2

(n+l)ncosn9

r,,-2

cosn9

r"

sinn9
-

--

-(n + l)(n - 2)r" sinn 9

rll+2sinn9
-(n + 1)nr"cosn9

(n + 1)nr"sinn9

r"sinn9

-(n + l)(n -2)r"cosn9

n(n -1)r,,-2sinn9

rll+2cosn9

-n(n -1)r,,-2cosn9

(JrO

-n(n -1)r,,-2cosn 8

(J"

-n(n -1)r,,-2 sinn9

r"cosn9

II>

Complementary solutions, vt~ =0 .

r"+2

+ l)n sinn9

r"

(n-l)(n-2)sinn9

r"

(n -l)(n -2)cosn9

(n

(n + l)n cosn9
r"+2

(n +2)(n + 1)r"sinn9

(n +2)(n + 1)r"cosn9

n(n -1)r"-2sinn8

n(n -1)r,,-2cosn9

(Joo

+ 1 +lC)rIl+ 1sinn8

ncosn9

r,,-l

(n -1-lC)cosn9

(n -I +lC)sinn9

r,,-l

r,,-l

(n -1-lC)sinn9

-~

(n-l+lC)cosn9
r"-l

nsinn9

r"+l

----;:;+1

nsinn9

r"+l

-(n + 1 + lC)r,,+1 cosn9

(n

-nr,,-lcosn9

nr,,-lsinn9

2JUlo

ncosn9

-(n + l-lC)r,,+1sinn9

-(n + l-lC)rIl+1cosn8

-nr,,-lsinn8

-nr,,-lcos n9

2J1U,.

i:

~
~

i~.

00

2rsinO

-2rcosO
0

2r cos 0

2rsinO

2cosO/r

cosO/r

-2sinO/r

sinO/r

r3cosO

r3 sinO

rOsinO

rlogrcosO

rOcosO

rlogr sinO

-cosO/r

sinO/r

cr,e

cr"

c!>

Complementary solutions, V4c!> = 0 .

sinO/r

cosO/r

6rsinO

6rcosO

cree

-sinO-(K+ l)OcosO]

~[(K-l)IOgr sinO

+sinO+(K-l)OcosO]

~[-(K+ l)logr sinO

-COSO+(K+ 1)0 sinO]

1
"2[(K-l)logrcosO

(K-l)logr sin 0

+cosO+ (K+ 1)0 sinO]

1
2[(K-l)logrcosO

-cosO-(K-l)OsinO]

1
"2[ -(K+ l)logrcosO

-sinO+ (K+ l)OcosO]

~[-

-sinO+(K-l)OcosO]

~[-(K+ l)logrsinO

1
"2 [(K+ l)logrcosO
-cosO+ (K-l)OsinO]

-(K+2)r 2 cosO

(K+ 2)r2 sin 0

21lUe

(K-2)r 2 sinO

(K-2)r 2 cosO

2j.1U,

is"

1';"
'"tI

s...~

--

r4

r2

r4

---

r6

r4

- -

r6

r4

--+-

r6

2y _

8x 2y

r2

x 2x
-4

6x 8x 3

r6

r4

y 2x 2y
--+-

2y _ 8x 2y

8x 3

r4

6x
-4

Hannonic functions, V2cp = 0

sinO
-

cosO
-

r2

y 2x 2y
-+-

2x 3

x
r2

r4

r4

2x 2y

r2

r4

2x 2x 3

--+-

r4

2xy

r2

2x 3

r4

_ 2x 2y

r2

r4

1 2x 2
--+-

rlogrsinO

r4

1 2x 2
--+-

2xy

r4

crxY

crzx

--+-

logr cosO

rOsinO

rOcosO

logr

<1>

Complementary solutions, V4<1> = 0 .

r4

r6

r4

r4

r6

6x

2y 8x 2y
--+-

8x 3

--+-

r2

--------

r4

r4

---

2xy
-

r2

1 2x
--+-

2x2

--r2
r4

r4

2xy

1
x2
-(K-1)logr +2
r2

1
xy
--(K+1)0-2
r2

L_ 2x 2y

r2

1
xy
-(K+1)0-2
r2

1
xy
-(K-1)0-2
r2

1
x2
--(K+ l)logr-2
r2

1
x2
-(K-1)logr-2
r2

1
x2
-(K+ 1)Iogr-2
r2

1
xy
-(K-1)0+2
r2

Y
r2

x
r2

x
r2

21lv

r2

2JlU

3x 2x
-4
3

2x 2x
- r-4
r2

r2

r4

2y 2x 2y
--+-

1 2x
--r2
r4

r4

2xy
-

crw

~.

~.

lC-1
lC-1
-T(logr sinO-OcosO) T(logr cos 0 + 0 sin 0)

logr

1
-(lC-1)(logr -l)r
2
1
-(lC-1)rO
2

[1- 2(lC-1)(1-10g r)}_2


(lC+ 1)

2(lC-1)logr -1
r
(lC+ 1)

are

aee

2JlUr

2JlUe

~.

lC-1
lC-1
-2- (logr cos 0 + 0 sin 0) T(logr sinO-OcosO)

logr

a"

r-1 cos 0

r-1 cos 0

r-1 sinO

r-1 sinO

2(lC-1)logr }-2
[1
(lC+ 1)

lC+1
2(lC-1) <I>

(logr)2
2

r-1 sinO

logr
r-1 cosO

X
4x 3
(3+lC)--r2 r 4

Y 4x 2y
-(lC+1)-+r2
r4

2Jlv

r-2

Particular solutions, V4<1>=-2:::~V.

Y 4x 2y
(lC-1)-+r2
r4

12xy 16x3 y
--+-r4
r6

2 16x 2 16x4
---+r2 r 4
r6

4xy _16x 3y
r4
r6

sin 20

X
4x 3
-(3-lC)-+r2 r 4

4 20x 2 16x4
---+r2 r 4
r6

8xy _16x 3y
r4
r6

12x2 16x
r
r6

-4 - -

cos 20

2JlU

a yy

a.a

<I>

a..,

Complementary solutions, V4<1> = 0 .

VI

--

S
to
Ei

J"

(lC-l)[(3 + 2logr)logr -1] cos 29


4(lC+ 1)
(lC-l) [2(2Iogr + 1)logr -1]r sin 29
8(lC+ 1)
(lC-l)(l- 21ogr)r1ogr cos 29
4(lC+ 1)

(lC-l)[l- (3 + 21ogr)logr] sin29


4(lC+ 1)
(lC-l)[2(21ogr + 1)logr -1]rcos29
8(lC+ 1)
(lC-l) (1- 2logr)r logr sin 29
4(lC+ 1)

2(lC-l)r"+1
(lC+ l)(n +2)

2j.1U,

2J.U1e

+ IOgr} sin29

{ (lC-l)
8(lC+ 1) [2 log r(2 log r +5) + 1]

+ IOgr} sin29

crre

+IOgr} cos29

{ (lC-l)
8(lC+ 1) [2 logr(2logr +5)+ 1]

+IOgr} cos29

[1 2(lC-l)(n + 1)}"
(lC+ l)(n + 2)

{ (lC-l)
8(lC+ 1) [1 + 2(2logr - 3)logr]

(21ogr -1) 21
. 29
16
r ogrsm

logrsin29

cree

{i~~li) [1 + 2(21ogr - 3)logr]

(21ogr -1) 2
r logrcos29
16

logrcos29

[1

2(lC-l)
(lC+l)(n+2) r

(n+2i

r"+2

r"
(n :;>!:Oor -2)

crrr

lC+l
2(lC-l) c!>

Particular solutions, V4c!>=-2K-ll~V.


K+

t.

~.

(lC-I)r.. -1logr
lC+l

(lC-l)r.. +1
2(m+ 1)

2J.1Ue
(sinm9, -cosm9)

2(lC-I)m[4+(m2-4)logr]r
(lC+ l)(m2_4)2

-2(m 2 -4)logr]

(lC+ I)[(n +2Y-m1

2(lC-I)mr~+l

2(lC-I)(n +2)r~+1
(lC+ 1)[(n+2)2-m1

+2(ml-4)logr] +logr

+m(m -1)10gr+2m]}

2(lC-I)
r[-m 2-4
(lC+ l)(m2_4)2

{1- 2(lC-l)(n + 2)(n + 1) }r~


(lC+ I)[(n +2)2-m1

2(lC-l)
[-4+ 3m 2
(lC+ l)(m2_4)2

r m- 2{ 1- lC-l [-I
m(lC+ 1)

(lC-l)(1 +m logr)rm - 1
m(lC+ I)

2(lC-l)m(n + l)r~)
(lC+ 1)[(n +2Y-m1

2(lC-l)m[m2+(m 2-4)logr]
(lC+ l)(m2_4)2

2(lC-l)(n+2-m2)} ~
(lC+l)[(n+2r-m1 r

(l-lC)[1 +(m -1)logr]r m- 2


lC+l

{I

(n+2'f-m2

r~+2

(n ::m,n ::m-2,m ::O)

r~

-(ml-4)(ml-2)logr] +logr

2(lC-l)
[4-3m 2
(lC+ l)(m2_4)2

[4+ (m2-4)logr]r2
(m2_4)2

(m::O or 2)

logr

-m(m-l)10gr]}

r"-2{1+ (l-lC) [l
m(lC+ 1)

(lC-I)r.. +1
2(m+ 1)

rm

r"

r"logr
2m

(m::O or 1)

(m::O or -1)

r m- 2

rm

2J.1Ur
(cosm9, sinm9)

(cosm9, sinm9)

Oea

ore
(sinm9, -cosm9)

(cosm9, sinm9)

0,.

-{lC+ 1)c1>
2(lC-l)(cosm9,sinm9)

V
(cosm9,sinm9)

Particular solutions, ~cI> =-2 1<+


1<-11 VZv .

t,,;I

Ul

1=;'

CHAPTER V
BENDING OF ELASTIC THIN PLATES

5.1 BASIC ASSUMPTIONS


The preceding chapters have been concerned, in large part, with so-called theory
of elasticity solution. It was assumed that there were three displacements u, v, W, and
each was, in general, a function of x, y, z. The displacements and corresponding strains
and stresses had to satisfy the stress-strain relations and the equilibrium equations which,
when put all together, gave Navier equations of motion.
This chapter, on the other hand, describes an approximate engineering theory. For
a thin plate in bending, shown in Fig. 5.1, which has a z-dimension (thickness) much
smaller than its other dimensions; we assume that a normal to the undeformed middle
plane of the plate remains normal after deformation (Fig. 5.2), and the deflection W
(displacement in the z-direction) is constant across the plate thickness. Thus
u(x,y,z)

=uO(x,y)-z aw~:,y)
(0,0 )

(5.la)

Fig. 5.1 Rectangular plate with span lengths equal to a and b in x


and y directions respectively.

155

Bending of Elastic Thin Plates

(5.1b)
Here the origin of the Cartesian system is located on the plate middle plane, and the
superscript zero stands for the quantities there. Using Eqs. 5.1, we can show that shear
distortions 1:u and 1,. are neglected.
The only non-zero strains are

E:

rfw

=~-z

dX2

o
rfw
e" =e;, - Z dy2
AP

(5.2)

rfw

1", =r", - 2z dXdY


Note that fu also vanishes and all compatibility conditions are satisfied exactly.
Moreover, it is normal to further assume that the middle plane remains unstrained due
to bending, i.e. e!: = ~ = = O. This leads to

E:

=-z~

e" =-zKy,

(5.3)

1", =-2z1\.y
where

1C

stands for plate curvatures, i.e.


(5.4a)

(5.4b)

Fig. 5.2 Original and displaced configurations of an xz-section together


with a normal (AB ) to middle plane of the plate.

Foundations of Solid Mechanics

156

(5.4c)
5.2 EQUILmRIUM, BOUNDARY CONDmONS, AND STRESS
RESULTANTS
The equilibrium and well-posed boundary conditions of a rectangular plate as
shown in Fig. 5.1 can be derived from the following virtual work principle,
~=~

~~

where Wi and W. are the internal and external virtual works, respectively, and can be
expressed by
(5.6)

[iGi"/2(Oyy& +oyzSu
o

-lI/2

+o,.8w)dzdx]7=b
, .. 0

(5.7)

where 8 is the variational symbol signifying a virtual quantity, q (x, y) is the applied
transverse load, a bar C) denotes a boundary traction but with a stress sign convention,
and f/(x )];:~ == f(aJ - f(al). Performing integration with respect to z in Eqs. 5.6 and
5.7 yields

(5.8)

LqawdA - fIMz{: )+Mxy{~ )-Q z8W]dY

+[M7{~ )+Mxy{: )-Q,8W]dx

(5.9)

where +c denotes the integral along the simple closed curve C enclosing the plate
middle plane, and stress resultants are introduced for the first time, i.e.

157

Bending of Elastic Thin Plates

(5.lOa,b,c)
(5.lOd,e)

(5.11)

and

dMzaw
ffM",]
dx dx + dM",aw
dx dy =[ddX (dMz
dX w)+dyd(dM",
dX ] [ffMz
dX2 W+ dxdy
W

(5.12)

In view of Eq. 5.11 and the Green's theorem (Eq. 1.89), we have

i (MzB~ +M",BK",)dA = - [M",{:}u -Mz{: fY]


-i[d;z{:)+ d:", {:]dA

(5.13)

Similarly, we can get

i (M,B~ + M",BK",)dA = [M",{~ fY -M,{~ }u]


- i[d;,{~)+ d~"'{~;]dA

(5.14)

Due to Eq. 5.12, applying the Green's theorem one more time, we have

i[AdX--dMz{aw)
dX +dM",{aw]
dX- -dy

dA = -

f(dM",
dMdXz )
-Bwdx --Bwdy
cdX
(5.15)

Similarly, we can get

i[Ady--dM,{aw)
dy +dM",{aw]
dy- -dX

dA =

f(dM",
dM,
-Bwdy --Bwdx
CdY
dy )
(5.16)

Foundations of Solid Mechanics

158

In addition, due to integration by parts, we can show that

M..,s(~}x =- a:.., 5wdx

+M..,(a,0)5w(a,0)-M..,(0,0)5w(O,O)

+M..,(O, b)5w(O, b) -M..,(a,b)5w(a,b)

(5.17)

Similarly,

M..,s(~}y =- a:..,

5wdy +M..,(a,b)5w(a,b)-M..,(a,O)5w(a,O)

+M..,(0,0)5w(0,0)-M..,(O,b)5w(0,b)

(5.18)

Due to Eqs. 5.13 to 5.18, Eq. 5.8 becomes

W. = _ (ifMit + 2 ifM.., + ifM")5WdA

JA ax 2 axay ay2

-Mlts(~ + M.,s(~ }x
}y

( a:..,
Q" +

)5wdY _

( a:.., )5Wdx
Q., +

-2M..,(a,b)5w(a, b) + 2M..,(O,b)5w(0, b)
(5.19)

+2M..,(a,O)5w(a,O)-2M..,(O,O)5w(O,O)
where Q" and Q., are defined as

aM" aM..,
ax + ay
aM., aM..,
Q.,= ay + ax

(5.20a)

Q,,=

(5.20b)

It will be shown later that Q" and Q., are shear stress resultants in the same manner
as Eqs. 5.1Od and e. In view of Eqs. 5.17 and 5.18, Eq. 5.9 becomes

w. =

if

q5wdA -

M"s(~}y + M.,s(~}x + (Q" + a:..,

}WdY

r(aM ~~.. - Jcl


Q .,+ ax'" ywdx-2M..,(a,b)5w(a,b)+2M..,(O,b)5w(o,b)
+ 2M..,(a, O)5w(a, 0) - 2M..,(0, 0)5w(0, 0)
Substituting Eqs. 5.19 and 5.21 into Eq. 5.5 leads to

(5.21)

159

Bending of Elastic Thin Plates

OZM"
OZMxy OZM,
--+2--+--+q =0;
dX 2
dXdY dy2

dw

M"=M,,

or -=
dX

M,=M,

or -=
dy

dw

inA

(5.22)

constant; on boundaries x = 0 and a

(5.23a)

constant; on boundaries y = 0 and b

(5.23b)

V"=V,,

or w = constant;

on boundaries x = 0 and a

(5.24a)

V,=V,

or w = constant;

on boundaries y = 0 and b

(5.24b)

R =R or w =constant; at four corners


(5.25)
Equation 5.22 is the equilibrium equation, and Eqs. 5.23 to 5.25 are well-posed boundary
conditions. In these equations,
dMxy
(5.26a)
V,,=Q,,+ay

dM,,,
V,=Q,+ dX

(5.26b)

R =-Mxy-M,,,=-2Mxy; forcorners(O,O)and(a,b)

(5.27a)

=Mxy +M,,, =2Mxy;

for corners (a,O) and (O,b)

(5.27b)

It will be shown later that M" and M, are bending moments, Mxy twisting moment,

V" and V, Kirchhoff or supplemented shear forces (G. Kirchhoff, 1850), and R a
corner force.
5.3 PHYSICAL MEANING OF STRESS RESULTANTS
According to Eqs. 5.10, we can illustrate the stress resultants involved on a small
element of the plate middle plane as in Fig. 5.3. For the equilibrium of forces in
z -direction, it can be shown that

dQ" dQ,
(5.28)
-+-+q=O
dX
dy
while the equilibrium of moments about x and y axes gives Eqs. 5.20b and 5.20a,
respectively; which upon substituting into Eq. 5.28 give the same equilibrium equation
as Eq. 5.22. For the supplemented shear forces on plate boundaries, consider the
boundary x = a, as an example. The shear force Q" and the twisting moment Mxy are,
according to Eqs. 5.10, as shown on the Fig. 5.4(a), while Fig. 5.4(b) depicts an
equivalent system, i.e. a twisting moment is taken as a couple of equal and opposite

Foundations of Solid Mechanics

160

dx

/
-oQ

-----;;;,.------- /----'--~----=--Qx + _x dx
-ax

.,--- M +'aMXd
-- x
X

1}x

-aMxy
Mxy + a;-dx

Fig. 5.3 Small plate element subjected to stress resultants.


Point (a,y)

(Mxy -

Point (a,y)

-aMxy dy
'Oy 2" )dy

Mxy -

'aMxy dy
'ay""2

Qx dy
( Mxy +

(a)

Plate edge subjected to transverse shear force Qx and twisting


moment M",.

'aMxy dy

---ay 2"

(b) Equivalent couple of shear forces


for M",.

Fig. 5.4 Supplemented or Kirchhoff shear force on a plate edge perpendicular to x -axis.
forces. Thus the equivalent force at point (a, y ) per unit length of the plate edge is

161

Bending of Elastic Thin Plates

Mxy

(a) Twisting moments near a comer


(a,b).

(b) Equivalent couple of shear forces


for each twisting moment near
the comer.

Fig. 5.5 Comer force at comer (a,b).


which is defmed as V.. in Eq. 5.26a. For the corner force, consider the comer (a,b)
as an example. The total twisting moments Mxydy and M)!Xdx act on edges near the
comer as shown in Fig. 5.5(a), while Fig. 5.5(b) depicts an equivalent system, i.e. a
twisting moment is taken as a couple of equal and opposite forces. The free force at
the comer is thus
which is defined as R in Eq. 5.27a.
For a non-right angle ('# 1CI2) comer as shown in Fig. 5.6, the comer force can
be shown to be
R

=-M..", + M.."y"

(5.29)

where x',y',z and x",y",z belong to right-hand systems with x' and x" normal to
two different edges forming the comer.

Fig. 5.6 Formulation of comer force at non-right angle.

Foundations of Solid Mechanics

162

S.4 GOVERNING CONDmONS FOR ISOTROPIC PLATES


For isotropic elastic plane stress problems, the stress-strain relations, Eqs. 4.7, are

<1""=--2 (E",,+VEYJI )

(I-v)

<1YJ1

=(1- v2) (Ey, + VE",)


E

(S.30b)
(S.30c)

"l
"'-2(1+v) ..,

<1 -

(S.30a)

Substituting these equations into Eqs. S.lOa to c leads to

M" = - D (K.,. + vlS,)

(S.31a)

My = -D(lS, +vK.,.)

(S.31b)

M.., = My" =-D(l-v)K..,

(S.31c)

where D is the flexural rigidity in bending, i.e.

Eh 3
D =----,,12(1-v2)
and

le'S

(S.32)

are curvatures as defmed in Eqs. S.4. Thus

= _D(ifw +vifw)

= _D(ifw
+vifw)
dy2
dX2

"

dX2

dy2

(S.33)

ifw

M.., =Mp =-D(1-V)dXdY


For a plate with uniform properties; the equilibrium (Eq. S.22), shear stress
resultants (Eqs. S.20 and S.26) and the comer force (Eq. S.27), can be put in terms of
w as

V'w=!L
D

d u2
=
-D-(v w)
"
dX

(S.34)
(S.3Sa)

(S.3Sb)

163

Bending of Elastic Thin Plates

Or

'00
+ __
r

dr

Fig. S.7 Stress resultants in polar coordinates acting on small plate element.

a [azw
azw]
"= - Dax- -+(2-v)ax 2
ay2

(S.36a)

_D~[azw + (2_V)azw]
2

(S.36b)

v
v

=
y

ay ay2

ax

azw

(S.37)

R = 2D(I-V)aXay

In term of polar coordinates, the equilibrium equation is Eq. S.34, while the stress
resultants acting on the plate element as shown in Fig. S.7 are
Mr=

-D[(1-V)~;Z +VV2W]
(S.38)

a (ldw)
;:-ae

Mr9 =Mer =-D(1-V)ar

Qr = -D-(Vw)
ar

I
rae

Q = -D--(Vw)
9

(S.39a)
(S.39b)

164

Foundations of Solid Mechanics

and the supplemented or Kirchhoff shear forces are


loMr9
Vr=Qr+;:-ae

(5.40a)
(5.40b)

In these equations the Laplace operator ~ and the biharmonic operator V4 are given,
respectively, by Eqs. 4.26 and 4.21.
5.5 SOLUTIONS FOR RECTANGULAR PLATES
Consider an infmitely long plate of width a, i.e. occupying -00 < Y < 00, 0 S; x
with a load q which depends only on x. The field equation, Eq. 5.34, is

d4w q(x)
-=D
dx 4
The solution of

particular solution

(5.41)

can be written as the sum of a complementary solution


i.e.

We

We

and a

W P'

(5.42)

W=We+Wp

Thus

S; a,

must satisfy the homogeneous differential equation

dw
_
_e=O
dx 4
4

(5.43)

the solution of which is


(5.44)
a

--"/

Fig. 5.8 An infmitely long plate.

165

Bending of Elastic Thin Plates

in which C1 to C4 are arbitrary constants to be detennined by means of four boundary


conditions along the plate edges, two at x =0 and two at x =a. The particular solution
is any solution of the original field equation, i.e.
d4wp
dx4

q(x)

=D

(S.4S)

The present problem is commonly known as cylindrical bending. As an example,


consider the case where q(x) =qo, one edge of the plate is built-in, while the other is
free. The particular solution wp can be chosen as
q~4

wP=24D
and the pertinent boundary conditions are
w(O)=O, : ; 1,.=0 =0,
M,,(a) =0, V,,(a) =0.

These give the four equations

qo

qo

2C3 +6aC4 =-2D a , C4 =-6D a.


Next, consider the problem of a rectangular plate supported at three corners and
loaded by a force P at the remaining corner, as shown in Fig. S.9. The field equation
is
V'w=O

(S.46)

and boundary conditions are


M,,(O,y)=O,

M,,(a,y) =0

(S.47a,b)

V,,(O,y) =0,

V,,(a,y) =0

(S.47c,d)

M,(x, 0) =0, M,(x,b)=O

(S.47e,f)

V,(x,O)=O,

(S.47g,h)

V,(x,b)=O

w(O, 0) = w(a,O) =w(O,b) =0

(S.47iJ,k)

-2M",(a,b)=P

(S.48)

w=Axy

(S.49)

The solution is found to be


This satifies Eqs. S.46 and S.47.

The arbitrary constant A can be detennined from

166

Foundations of Solid Mechanics

Fig. 5.9 Anticlastic bending.


Eq. 5.48, to give
Pxy
2D(1-v)

w=-~-

(5.50)

Note that all stress resultants vanish everywhere and Mzy is constant; the problem has
been named pure torsion bending or anticlastic bending.
Consider the problem of a rectangular plate simply supported along all four edges,
as shown in Fig. 5.10. The field equation is

~w =q(x,y)
D

(5.51)

and boundary conditions are


w(O,y)=O,

w(a,y)=O

Fig. 5.10 Rectangular plate simply supported along all edges.

(5.52a,b)

Bending of Elastic Thin Plates

167

Mz(O, y)

=0,

Mz(a, y)

=0

(5.52c,d)

w(x,O)=O,

w(x,b)=O

(5.52e,t)

My(x,O)=O,

My(x,b)=O

(5.52g,h)

In order to satisfy Eqs. 5.52 we use the series solution


~
~
. m1tX . nxy
w(x,y)= "'~lll~l w"",Slna-SlnT

(5.53)

The loading function can be expressed as a similar double Fourier series as


~
~
m1tX
nxy
q(x,y)= "'~1II~1 q"",sina-sinT

(5.54)

where
(5.55)
Substituting Eqs. 5.53 and 5.54 into Eq. 5.51 yields, for each pair of harmonics m and

n,
W

q"",

=------"'" Dx 4(m 21a 2+ n 21b 2)2

(5.56)

and the solution is complete. The solution in the form of Eq. 5.53 is commonly known
as Na:vier solution (C. L. M. H. Navier, 1823 and 1827). It is noteworthy that comer
forces exist in this case.
The technique of Fourier series can be applied to any rectangular plate simply
supported on two opposite edges, as shown in Fig. 5.11. In order to satisfy the simply

r,--------~L-~-x

f-a

./

y
Fig. 5.11 Rectangular plate simply supported along two opposite edges.

Foundations of Solid Mechanics

168

supported conditions along the edges where x = 0 and x = a we take the single series
solution
m1tX
(5.57)
W(x,y)= L wm(y)sin-m=l

The loading function can be expanded in a similar Fourier series


q(x,y)=

2i

where

qm(Y)=-

m1tX
L- qm(y)sin-

m=l

(5.58)

m1tX
q(x,y)sin-dx

(5.59)

Substituting Eqs. 5.57 and 5.58 into the field equation, Eq. 5.51, leads to, for each
hannonic m, an ordinary differential equation of the fonn
d4wm
m2~d2wm m 41t4
qm
- - 2 -2- + - w
=(5.60)
4
dy4
a dy2
a
m D
the solution of which can be taken as a sum of complementary and particular solutions
Wm(y) =Wme(y) + wmp(y)

(5.61)

where Wme satisfying Eq. 5.60 with qm =0 takes the fonn


m~
m~.
m~
. m~
m~
m~
wme(y)=Amcosh--+Bm--smh--+Cmsinh--+Dm--cosh-- (5.62a)

or
W (y)=A e-n"cy'a+B m1ty e-n"Cyla+C emfCy1a+D m1tYemfCyla
me
m
m a
m
m a

(5.62b)

Here Am to Dm are arbitrary constants to be detennined by means of four boundary


conditions along the remaining two edges where y =b and y =-c. Equation 5.62a is
convenient in the considemtion of the symmetry with respect to the x-axis, while Eq.
5.62b is more appropriate in the discussion of the boundedness of the solution when
the y-dimension of the plate extends to infmity. The solution fonn in the present case
is known as Levy-Nadai solution (M. Levy, 1899 and A. Nadai, 1925).
Solutions for several useful boundary value problems concerning rectangular plates
are compiled by Timoshenko and Woinowsky-Krieger (1959).
5.6 CLOSED FORM SOLUTIONS FOR CIRCULAR PLATES
Consider axisymmetrical bending of a circular plate subjected to a unifonn load
of an intensity qo in the domain 0 ~ r < b < a and simply supported along its edge
where r =a. We divide the plate into two domains by a circle r =b, then write the
field equations

169

Bending oj Elastic Thin Plates

(5.63a)
and

VW z =0

(b < r < a)

(5.63b)

Here subscripts 1 and 2 are introduced to denote domain 1 where 0 < r < b, and domain
2 where b < r < a, respectively. The biharmonic operator V4 for the present case which
is axisymmetrical assumes the simple form

V4=~! {r ![~!~! )J}

(5.64)

For Eq. 5.63a, the solution is


(5.65)
where Al to Dl are arbitrary constants to be determined from boundary conditions. To
be bounded everywhere inside the domain where it belongs, Eq. 5.65 must not contain
the terms log r and rZlog r since they are singular and/or give singular stress resultants
at r =0. Thus C1 =D 1 =0 and

qo

w1 =64D(r +Al+Blr),

(O~r<b)

(5.66)

At this stage, we can say that a particular solution for any case that has a uniform load
intensity qo is
qor 4
wp= 64D

(5.67)

For domain 2 which is governed by Eq. 5.63b, there is only the complementary solution,
i.e.
Wz

qo
z
z
=64D
(Az+Bzr + C zlog r +D2r logr),

(b < r < a)

(5.68)

The remaining task is to determine the constants At> B1, A2, B2, C2 and D2 from the
boundary conditions at r =a and the continuity conditions at the interface r =b between
the two domains. These conditions are
W

dW11
II r=b -w
21 r=b' -dr

r=b

dW21
- dr

M rl Ir=b -M
r2Ir=b'
w2 1r=a

=0,

(5.69a,b)
r=b

(5.69c,d)
Mr21 r=a

=0

(5.6ge,t)

Note that Vr and Qr are identical in axisymmetrical cases. Table 5.1, at the end of
the present chapter, shows that Eqs. 5.69 may be written

Foundations of Solid Mechanics

170

b 4+Al +bZBI =Az+bzBz+logbCz+b2IogbDz

(5.70a)

4b 3+2bB I =2bBz+b-ICZ+b(1 +2Iogb)Dz

(5.70b)

-4(3+v)b z -2(1 +V)B1 = -2(1 +v)Bz+(1-v)b-zCz


- [2(1 +v)logb +3+v]Dz

(5.7Oc)

-32b = -4b-IDz

(5.7Od)

Az+azBz+logaCz+a2IogaDz=0

(5.70e)

-2(1 +v)Bz+(1-v)a-zCz - [2(1 +v)loga +3+v]Dz =0

(5.70f)

Equations 5.70 fann a set of six linear simultaneous equations for derennining the six
constants Al to Dz.
Now consider the axisymmetrical bending of a plate subjected to a concentrated
force P at the origin of the reference coordinates, as shown in Fig. 5.12. The pertinent
field equation is

Vw =0,

(r >0).

The general solution is the complementary solution, i.e.


1

w =D (AI +Azr +A3 log r +A4r logr).


Table 5.1 contains slopes and stress resultants for variaus common deflection functions.
From that table, we note that the term logr is inadmissible since it is singular at r =0,
thus the constant A3 must be set to zero. The term rZlogr, though it gives singular
moments and a singular shear stress resultant, should be retained since it is appropriate
for a concentrated load. Consider a small neighborhood of the point 0, as shown in Fig.
5.13. The equilibrium of the neighborhood demands

y
Fig. 5.12 Plate subjected to a concentrated force.

171

Bending of Elastic Thin Plates

j
y
Fig. 5.13 Equilibrium near application point of the concentrated load.
P

+ Lim
-+0

2",

(r V r )

_
r-

de =0

(5.71)

With the help of Table 5.1, we can obtain

-~) =0,

P + Al (0) + A2(0) + 21tA4{


or

A4= 81t'

We can also verify that any other admissible displacement functions (axisymmetric as
well as non-axisymmetric) contribute nothing to the second term in Eq. 5.71. Thus
without loss of generality, we can conclude that the term
P
81tD r210gr

(5.72)

is a particular solution for axisymmetric as well as non-axisymmetric problems relating


to a plate subjected to a concentrated force P at the origin.
We may use the particular solution Eq. 5.72 in Michell's problem, the problem of
a circular plate clamped along its edge and subjected to a concentrated load P as shown
in Fig. 5.14. Michell (1902) used the method of inversion to derive the closed form
solution
(5.73)
The first term is a particular solution Eq. 5.72, and the remaining terms are admissible
complementary solution. Even r;10gr2 is indeed admissible inside the plate, since r2 > 0
there. There remains the task of checking the boundary conditions
wi r=a =0, (awlar)1 r=a =0
(5.74a,b)

Foundations of Solid Mechanics

172

Fig. 5.14 Michell's solution.


We start with the observation that Eq. 5.73 expresses w as a function of'l and '2' i.e.
W('I,':z}. Accordingly, its derivatives are
aw awa'i awa'2
a, = a'i ar + ar2 ar

(5.75a)

aw
P (
rl )
a'i = 41tD rllog ~'2

(5.75b)

aw = ~(~2'i-';)
ar2 81tD
r2

(S.75c)

,;=,2+ ~2a2-2~rcos9

(5.76a)

2 2 a2 a
'2=r +~2-2~,cos9

(5.76b)

We note that

, sin 9 =rl sin 91 = r2sin 92


'1 cos 91 =

rcos9- ~

(5.76c,d)
(5.76e)

a
r2cos92=rcos9-~

(5.761)

a
rlcos91 -r2cos92=~-~

(5.76g)

173

Bending of Elastic Thin Plates

and in particular for r = a,

r1 = pr2

(5.77)

Equations 5.75 and 5.77 show that w in Eq. 5.73 satisfies Eqs. 5.74
Next, consider the problem of a clamped plate subjected to a concentrated moment
Ml as shown in Fig. 5.15(a). We may consider the moment Ml to be produced by two
equal and opposite forces at two neighboring points as shown in Fig. 5.15(b), where
Lim(Pe) =Ml
(5.78)
-+0

The solution for the force P acting at the point 01 is given by Eq. 5.73, i.e.
w+(r,O; pa)= 8:V wO(r1,r2,p)

(5.79)

where
(5.80)

Clamped edge

of-- f3a

~I~------~~----~--x
I

(a) Original.

(b) Equivalent.

Fig. 5.15 Clamped circular plate subjected to a concentrated moment MI'


Differentiating Eqs. 5.76a and b, and using Eqs. 5.76e and f, we find

ar

ap

(5.81a)

=-acosO I

(5.81b)
The solution for the other force -P at the point very close to 0 1 in Fig. 5.15(b) is
w-(r,O;

pa +e) =-w+(r,O; pa)-

a;+ ~

(5.82)

Adding the last equation to Eq. 5.79, and using Eq. 5.78, we fmd the required solution

Foundations of Solid Mechanics

174

for the applied moment Ml to be

Ml Owo(rl, r2' P)
w = 81tDa
ap
Ml (Owoari Owoar2 Owo)
arl ap + ar2 ap +arr

= - 81tDa

Because of Eqs. 5.80 and 5.81, Eq. 5.83 is

Ml .[
j rl )
2
w =81tD 2rl lo,"Pr2 cos 9 1 + (rl -

9 1]

2 (cos
P2r2)
p2r2 + Pa
2

(5.83)

(5.84a)

or, due to the identity,

rcos9 r2
cos 92 1
---+--=-----+---,
p2r2 Pa
f32rf Parf
derived from Eqs. 5.76b and 5.76f,

Ml [
j rl )
2
w =81tD 2rl I0l(\pr2 cos 9 1 + (rl -

2 (r2
P2r2)
Parf -

r cos 9]
P2rf

(5.84b)

Being the only singular term in the solution above, the term

Ml

(5.85)

41tD rllogrl cos 9 1

is a particular solution for any plate subjected to a concentrated moment Ml about the
y-axis and at the origin 0 1,
For a circular plate axisymmetrically supported along its edge and subjected to a
concentrated force P at its center, the solution takes the form
P
w(r) 81tD (r 2 10gr +Al +A2r 2)
(5.86)

in which axisymmetry and boundedness of the solution at r =0 have been taken into
account. The arbitrary constants Al and A2 can be determined from the boundary conditions Eqs. 5.74 for clamped plates, or
(S.87a)
(5.87b)

MI
=0
r r=a

for simply supported plates. Use Table 5.1.


When the same plate is subjected, at its center, to a concentrated moment Ml about
the y-axis, its solution takes the form

Ml
3
w(r,9) = 41tD (rlogr+Alr+A2r )cos9

(5.88)

Here we have used Table 5.1 to select the complementary solution and have accounted
for the symmetry and the boundedness of the solution at r O. The boundary conditions,

Bending of Elastic Thin Plates

175

Eqs. 5.74 or 5.87, detennine the arbitrary constants A1 and A 2


The eventual results are the following: For a circular plate simply supported along
its edge and subjected, at its center, to a concentrated force P,
W

-~[

21

- 81tD r og~+

(3+v)(a 2-r2)]

2(1 +v)

(5.89)

while that subjected to a concentrated moment Ml about the y-axis,

[1

(1

r
a+-'-----''--'----'-:,--+v)(a -r )rcosa]
=Ml
- - r og-cos
(5.90)
41tD
a
2(3+v)a 2
For a circular plate clamped along its edge and subjected, at its center, to a concentrated
force P,
W

(5.91)
while that subjected to a concentrated moment M1 about the y-axis,
W

M1 [ r
(a 2 _r2)
r log~cos + 2a 2 r cos

=41tD

aJ1

(5.92)

Equations 5.91 and 5.92 can be obtained by specializing Eqs. 5.73 and 5.84, respectively,
and using
p~o, r 1 ~r, a1 ~a,
92~1t,

r2~00,

pr2~a.

Note that the limiting case of Eq. 5.84a involves an indetenninate quantity (00-00), but
Eq. 5.84b is well behaved.
If the concentrated force P and concentrated moment M1 are applied at an eccentric
point and the plate edge is not clamped, closed fonn solutions (like Eqs. 5.73 and 5.84)
seem not possible. For such cases we use the particular solution given by Eq. 5.72 or
5.85 with a complementary solution in Fourier series fonn, as described in the following
section.
5.7 SERIES SOLUTION FOR CIRCULAR PLATES
The deflection and load functions are expanded in Fourier cosine series, i.e.
w(r,9) =
q(r,9)=

where

L
L

m=O

m=O

wm(r)cosma

(5.93a)

qm(r)cosm9

(5.93b)

Ii"

qo(r) =-

1t 0

q(r,a)d9

(5.94a)

Foundations of Solid Mechanics

176

2LI: q(r, 0) cos mOdO,

q",(r) =-

1t

(5.94b)

(m >0)

Substituting Eqs. 5.93 and 5.94 into the field equation, Eq. 5.34, yields
q",(r)
v:.w",(r)=v

(5.95)

where
2
v2'" =!~(r~)m
rdr dr
r2

(5.96)

The total solution to Eq. 5.95 can be taken as


(5.97)

w",=wp",+w.",

where wpm is a particular solution satisfying Eq. 5.95, and W.'" is the complementary
solution satisfying

v:.w.",(r) =0
For each m,

(5.98)

W.'" can be obtained as


w.o(r) =Ao +Bor 2+ Cologr +Dor 210gr

(5.99a)

w. 1(r) =A1r +B1r 3 +C1r-1 +Dlr logr

(5.99b)

wem (r)=AIIIr"'+B r ...... +Cmr"'+2+D1r ......


+2 (m > I)
ft'

(5.99c)

1ft

where the constants A"" B"" C", and D", are to be detennined so that the complete
solution satisfies the boundary conditions. We may obtain a similar sine series solution.
Table 5.1 (at the end of the present chapter) contains Chv/'iJr and the stress resultants
for all these complementary and particular solutions.
As an example, consider an annular plate with two concentric boundary circles of
radii a and b (a > b). The load on the plate is uniform with an intensity qo. The plate
is supported by k equally spaced columns along its outer edge. Each column is 2aa
wide, where 2<X is the angle subtended by each column at the center of the boundary
circles. It is assumed that the column reaction is only in the form of a force uniformly
distributed over each column width. Choose the reference coordinate system to have its
origin at the center of the boundary circles and the x-axis passing through the center
of a column. The total load on the plate is 1t(a 2 - b 2)qo, thus the reaction intensity on
each column is 1t(a 2 -b 2)qol(2kaa). This reaction can be written as a periodic but
discontinuous function of 0 as follows.

1(0)=

1t(a 2 -b 2)qo

=0,

2kaa

'

(21tj

21tj.

T-<x<O<T+<X'} =0, 1,2, .. ,k

(elsewhere)

(5.IOOa)
(5.IOOb)

The period is equal to 21t1k. The function 1(0) is expanded in a Fourier cosine series

177

Bending of Elastic Thin Plates

with such a period, namely

ao
/(9) ="2 + "~l an cos nk9,
~

where

an =-2ki"'k /(9) cos nk9d9,


1t

or, due to Eqs. 5.100,

Clank

Replacing nk by m in the series above, we get


(5.101)
where
(5.102a)

ao =(a 2 - b 2)qr/a

(5.102b)

The solution can then be written in a similar series as

w(r, 9) =

~~ ~4 + m=o.flk.... wcmcosm9)

(5.103)

in which the particular solution given in Eq. 5.67 has been adopted and wcm(r) is as
defmed by Eqs. 5.99. Boundary conditions for the present case are
Vr (a,9) = -/(9), M r (a,9) = 0
(5.104a,b)
Vr (b,9) =0, M r (b,9) =0

(5.104c,d)

Substituting Eqs. 5.101 and 5.103 into Eqs. 5.104, using Eqs. 5.38 to 5.40 (or Table
5.1) and 5.102, and equating the coefficients for each m in the resulting series, we arrive
at the simultaneous equations; for m =0,
-32a - 4a-1D o =-32(a 2 - b 2)/a

(5.105a)

-4(3 +v)a 2 -2(1 +v)Bo+ (1-v)a- 2Co- [2(1 +v)loga +3 +v]Do =0

(5.105b)

-32b -4b- 1D o=0

-4(3 +v)b 2 -2(1 +v)Bo+ (l-v)b-2CO- [2(1 +v)logb + 3 +v]Do =0

(5.105c)
(5.105d)

Foundations of Solid Mechanics

178

and for m

~2,

(l-v)m 2(m -1)a"'-3A", - (l-v)m 2(m

+ l)a-m- 3B", + (m + l)m[(I-v)m

-4.1a",-IC",+(1-m)m[4+(I-v)m]a-m- ID", =-64(a 2_b 2)sinma


ama
-(I-v)m(m -1)a"'-2A",-(I-v)m(m

+ l)a-m-2B", -em + 1)[4v

+(I-v)(m + 2)]a"'C", + (m -1)[(I-v)(2-m)+4v]a-mD", =0


(l-v)m2(m -1)b"'-3A", - (l-v)m 2(m

(S.106b)

+ l)b-m-3B", + (m + l)m[(I-v)m

-4.1b"'-IC", +(1-m)m[4+(I-v)m]b-m- ID", =0


-(I-v)m(m -1)b"'-2A", -(I-v)m(m

(S.I06a)

(S.106c)

+ l)b-m-2B",-(m + 1)[4v

+(I-v)(m +2)]b"'C",+(m -1)[(I-v)(2-m)+4v]b-mD", =0

(S.106d)

Noting that Eqs. S.1OSa and c are identical, and that Ao does not appear in any conditions above, we need to impose another condition to determine this constant, which
corresponds to a rigid-body deflection of the plate. We choose the condition of zero
deflection at the center of a column, e.g. w(a, 0) 0 which leads to

a4+Ao+Boa2+Cologa +Doa 210 ga

I.

III=k. 2k.

...

(A a'" + B a-m
IN

1ft

+ C a"'+2 +D a-m+2) =0
'"

(S.107)

Then Do can be determined from Eq. S.1OSa or c, Bo and Co from Eqs. S.10Sb and d,
A", to D", (m ~ 2) from Eqs. S.106a to d, and Ao from Eq. S.107. If there is a single
supporting column, the constants for m 1 vanish, i.e.
Al =Bl =C l =Dl =0.

but this case is not valid since the structure is unstable.


Several problems of an annular plate supported by two concentric rings of equally
spaced columns were analyzed by Agabein, Lee and Dundurs (1967). Later, problems
of annular plates supported by columns on the outer boundary and ftxed along the inner
boundary were solved by Kosherick, Dundurs and Lee (1968).

S.8 POLYGONAL PLATES SUPPORTED AT CORNERS


This section deals with problems of polygonal slabs or plates supported directly
on columns at the corners. Such problems are frequently encountered in practice. The
solutions to be presented are approximate, but in closed forms.
As an example, consider the paper by Lee and Ballesteros (1960), which is a study
on a rectangular plate subjected to a uniform load of an intensity qo and supported at

179

Bending of Elastic Thin Plates

its four corners, as shown in Fig. 5.16. For an approximate solution, incorporating the
relevant symmetry, we take a polynomial
w(x,y)=C1 +C~2+C3y2+C~4+C~2y2+C6Y4

Substituting Eq. 5.108 into the field equation, Eq. 5.34, for q(x,Y)

(5.108)

=qo,

qo
24C4+ 8Cs + 24C6 =D

yields
(5.109)

a 'j- a-j

+
b

Fig. 5.16 Rectangular plate supported at corners.


The boundary condition that there be no shear stress resultants along the free edges is
expressed by
(5. 110a,b)
Vz(a,y) =0, Vy(x,b) =0
Substituting Eq. 5.108 into Eqs. 5.110, and using Eqs. 5.36, we find
6C4 + (2-v)Cs =0

(5.111a)

(2-v)Cs + 6C6 = 0

(5.111b)

Solving Eqs. 5.109 and 5.111 for C4 , Cs and C6 and substituting results in Eq. 5.108
yields
2

W(X,y)=Cl+C~ +C3y

qo

+ 48(I-v)D [(2-v)(x

2 2-!

+y )-6x Y J

(5.112)

The twisting moments Mzy and Myz given by Eq. 5.33c are

qoXY

Mzy=Myz=T

(5.113)

and the force at each plate corner given by Eqs. 5.27 is


(5.114)
R =-qab
Note that, irrespective of C 1, C2 and C3, Eq. 5.112 satisfies the field equation, the
condition of zero shear stress resultants at the free edges, and the condition of corner

Foundations oj Solid Mechanics

180

forces. The remaining unknowns Ch C z and C3 can be determined from the condition
of zero deflection at a comer, i.e.
w(a,b)=O
(5.115)
and two conditions stating that the mean moments on the edges are zero,

f'
f'

Mx(a,y)dy =0

(5. 116a)

My(x,b)dx =0

(5. 116b)

Equations 5.116 are approximations to the exact boundary conditions at the free edges,
i.e. Mia,y)=O and My(x,b) =0. Substitute Eq. 5.112 into Eqs. 5.115 and 5.116, and
use Eqs. 5.33a and 5.33b. We find

ZbZ
Z b Z) -2(7v-l)
C1 = qoa Z [ (lO+v-vz) ( az+-Z
48(I-v)D
ba

(5. 117a)

a]

(5. 117b)

bJ

(5. 117c)

b
b
Z)-b
Cz = qoa Z [ (1 +5v)--(6+v-v
24(I-v)D
a

a
Z
C3 = qoab Z [ (1+5v)-b-(6+v-v)24(I-v)D
a

Numerical results obtained from by this approximate solution agree closely with those
obtained numerically by Nadai (1922) and Marcus (1932), and by experiments on
prestressed concrete slabs by Scordelis, Pister and Lin (1956).
A similar approach was adopted by Vijakkhana, Karasudhi and Lee (1973) for
equilateral triangular plates subjected to uniform loads and supported at the plate comers.
Suriyamongkol (1985) studied the bending of regular polygonal plates of general
form, with 3, 4, 5, ... sides, subjected to uniformly distributed loads and central concentrated loads, and having simply supported edges, or free edges with comer supports.
The solution is a sum of a particular solution and a complementary solution. The
particular solution is given by Eq. 5.67 for a uniform load and by Eq. 5.72 for a
concentrated force. The complementary solution is taken as a series. The Gram-Schmidt
orthogonalization process (Wylie, 1975) is applied to solve the series equations. Exact
solutions are obtained for simply supported edges. Approximate moment conditions
(Eqs. 5.116) have to be adopted for free edges. For practical purposes, the series may
be truncated at the third term. Results agree closely with others found by different
reliable methods.
5.9 PLATES ON ELASTIC FOUNDATION
When a transversely loaded plate rests on a yielding foundation, the reaction of

181

Bending of Elastic Thin Plates

the foundation opposes the plate deflection. If the foundation is elastic, the resistance
is proportional to the deflection, i.e. 'YW, where y is the modulus of the foundation and
expressed in force per unit of plate area per unit of deflection. The only change brought
about by the presence of the elastic foundation is that the equilibrium equation for forces
in the z-direction, Eq. 5.28, becomes

aQ% aQ,

ax + ay

-'YW+q=O

(5.118)

Consequently, the field equation in terms of the deflection for a unifonn isotropic plate,
Eq. 5.34, becomes
(5.119)
where
~

=+(ylD )1/4

(5.120)

For cylindrical bending of very long rectangular plates, Eq. 5.41 becomes

d4w +J: 4w =q(x)


dx 4 ~
D

(5.121)

and the complementary solution, in place of Eq. 5.44, is


(5.122)
The plate deflection function for anticlastic or pure torsion bending (Fig. 5.9) is
given by Eq. 5.50. Substituting that equation into Eq. 5.119 yields
q(x,y)

yPxy

=2D(l-v)

(5.123)

Thus the anticlastic or pure torsion bending for plates on elastic foundation is possible
if the distributed load is related to the applied corner force by the equation above.
When a rectangular plate is simply supported along all four edges as in Fig. 5.10,
the Navier solution scheme given by Eqs. 5.52 to 5.55 may be applied. Substituting
these equations into Eqs. 5.119 yields
q"".
=-------------"". DK'(m2/a 2+n2/b 2)2+y

(5.124)

When a rectangular plate is simply supported along two opposite edges as in Fig.
5.11, the Levy-Nadai solution scheme as given by Eqs. 5.57 to 5.59 is also applicable.
Substituting those equations into Eq. 5.119 yields, in place of Eq. 5.60, the following
ordinary differential equation
(5.125)

Foundations of Solid Mechanics

182

The complementary solution for this equation is


w",(Y) = A",cosh /3",y cos a",Y + B", sinh /3mY sina",y
(5. 126a)
or
W"'(Y)= (A",cosa",y

+Bmsinamy)e~"'Y

+(C'" cos a",y + Dm sin amy)l"'Y

(5. 126b)

where

2/3! = [(m1tla)4 + ~~ 112 + (m1tla)2

(5. 127a)

2a! = [(m1tlat + ~~ 112 - (m1tla)2

(5.127b)

In Eqs. 5.126, Am to Dm are arbitrary constants to be determined by means of four


boundary conditions along the remaining two edges where y = b and y = -c. Equation
5.126a is convenient in the consideration of the symmetry with respect to the x-axis,
while Eq. 5.126b is more apropriate in the discussion of the boundedness of the solution
when the y -dimension of the plate extends to infmity.
For circular plates, the solution may assume the form of Fourier series as in Eqs.
5.93, but now wm(r) is governed by the ordinary differential equation

(v:.+~4)wm(r)= q~r)

(5.128)

The complete solution to the equation is given by Eq. 5.97, where wpm and Wem are
particular and complementary solutions for Eq. 5.128. For each m, Wem can be expressed
in terms of Kelvin functions (Section 1.10); thus

We",(r) =A",ber",g +Bmbeimg +Cmkermg +Dmkeim~r

(5.129)

Note that the following deflection function

-2:O~ keio~r

(5.130)

is a particular solution for the problem of a plate on an elastic foundation subjected to


a concentrated force P at the origin, as in Fig. 5.12, since it has the characteristic
singularity of Eq. 5.72 as r tends to zero. The same process of reasoning can be used
in the search for a particular solution to the problem of a concentrated moment. Specifically for m = 1, Eq. 5.129 can be rewritten as

we1 (r) =A1ber1g + B1bei1g + C1(ker1g +keil~r) + Dl(kerl~r -keil~r)

(5.131)

The term' within the last pair of parentheses in the equation above has the correct
characteristic singularity (Eq. 5.85) as r tends to zero. Thus a particular solution for

183

Bending of Elastic Thin Plates

the deflection function of the problem of a moment Ml about the y-axis and concentrated
at the origin of the reference axes is

-..J2Ml

41tD~ (kerl;r -

ke'lg)

(5.132)

Computing Kelvin functions is complicated and time consuming, especially for high
order functions. Karasudhi, Jou and Tansirikongkol (1976) avoided this type of analytical
functions by employing, for each harmonic m, central finite differences with respect to
r directly in the field equations and all pertinent boundary conditions.
5.10 EXERCISE PROBLEMS
5.10.1

The infinitely long plate shown in Fig. 5.8 has properties in domain 1 where
o ~ y < 00 different from those in domain 2 where - 0 0 < y SO. Find the solution
if the plate is simply supported along both long edges x =0 and a, and the
load is a point force P concentrated at x d and y 0, and 0 < d < a. Hint:
Adopt the Levy-Nadai solution (Eq. 5.57) for both domains. The (complementary) solutions are modifications of Eq. 5.62b taking into consideration
the boundedness of the solutions as y ~ 00 and y ~ - 0 0 for domains 1 and
2, respectively. The four pertinent boundary conditions at y 0, where the
two domains join, are

WI (x , 0) =WZ(x, 0),

Owl(X,y)
:\
uy

M)ll(X,O) =M)lz(x, 0),

)1=0

Owz(x,y)
:\'
cjy
)1=0

V)lz(x,O)- V)ll(x,O)=f(x),

where f(x) is the Dirac-delta distribution (Section 1.9) of the force P along
the x -axis, i.e.

f(x) =PS(x -d).


Note that if the two domains have the same properties, we can take advantage
of symmetry and fmd the solution for domain 1 only, and use only two
conditions at y 0, i.e.

OW1(X,y)
:\
uy
5.10.2

)1=0

=0

V ( O)=_f(x)
x,
2 .

')11

If the plate in Problem 5.10.1 is subjected to a concentrated moment Ml about


an axis parallel to the y-axis, the solution can be obtained as in the previous
problem by changing f(x) into

f(x) = Lim{MI [S(x -d)-a(x -d


8-+0

-e)]} = Mlf'>(x -d).

184

5.10.3

Foundations of Solid Mechanics

If the concentrated moment Ml in the preceding problem is about the x-axis


instead, the pertinent boundary conditions are
W1(x, 0) = WZ(x, 0),

dw1(X,y)

ay

M)ll(X,O)-M)lZ(x,O) =m(x),

)1=0

dwz(X,y)

=:.
uy

'

)1=0

V)ll(X, 0) =V)lZ(x, 0),

where m(x) is the Dirac-delta distribution of Ml along the x-axis, i.e.


m(x)

=Ml,(x - d).

5.10.4

If the y-spans of the plate in Problem 5.10.1 are fmite as in Fig. 5.11, we
can use solutions given by Eq. 5.62a or b. In addition to the four boundary
conditions at y =0, two more boundary conditions must be prescribed along
the edge y =b and two more on y =-c. If both domains are identical, take
advantage of symmetry and find the solution of domain 1 only.

5.10.5

Solve Problem 5.10.4 if the plate is subjected to concentrated moments.

5.10.6

A rectangular plate simply supported along all edges as shown in Fig. 5.10,
is subjected to a concentrated force P at x =c, y =d. Find the solution and
note the existence of corner forces. Hint: Adopt the Navier solution (Eq. 5.53)
and take the loading function as
q(x,y) =Po(x -c)o(y -d).

5.10.7

If the plate in Problem 5.10.6 is subjected to concentrated moments instead,


fmd solutions. Hint: For the case of a concentrated moment Ml about an axis
parallel to the y-axis, express the load function as
q(x,y)

=M1o(1)(X -c)o(y

-d).

5.10.8

For the plate as shown in Fig. 5.14, find Michell's solution for a concentrated
moment Ml about the x-axis. Hint:
Ml dw o
w(r, a) = 81tD f3a aa'
where Wo is as defined in Eq. 5.80.

5.10.9

Find a series solution for the circular plate shown in Fig. 5.14. Hint: Consider
two domains; domain 1 where 0:::; r :::; f3a and domain 2 where f3a :::; r :::; a.
Adopt the cosine series solution (Eqs. 5.93) for both domains, but the
boundedness of the solution at r =0 must be incorporated in domain 1. The
four pertinent conditions at the interface r = f3a, between the domains, are

Bending of Elastic Thin Plates

185

Mrl (J3a,8) =M,.z(Pa, 8), Vrl (Pa,8)- Vr2(Pa,8) =1(8),


Here 1(8) is the Dirac-delta distribution of the concentrated force P along the
circumference of radius Pa, i.e.
1(8)

=~ a(8).

In addition, two more boundary conditions must be prescribed at the plate


edge, i.e. Eqs. 5.74 for the clamped edge or Eqs. 5.87 for the simply supported
edge.

5.10.10 If the plate in the previous problem is subjected to a concentrated moment


Ml about the x-axis instead, fmd the solution. Hint: Adopt the sine series
solution, and write the load distribution in the fann
M
1(8) =- p2~2 a(1)(8),
then proceed in the same manner as in the previous problem.
5.10.11

If the concentrated moment Ml is about an axis parallel to the y-axis, what


is the solution? Hint: Adopt a cosine series solution. The conditions at the
interface between the two domains are
w l (Pa,8)=w 2(Pa,8),

OwI(r,8)

ar

Ow2(r, 8)
=
ar
r .. J!a

r=J!a

'

where

5.10.12 A circular plate is simply supported along its edge and subjected to a linearly
varying load
q(r,8)

=prcos8la,

where p is a constant and a is the plate radius. Find the solution. Hint: Table
5.1 gives a particular solution as
w,(r,8)

=prs cos 8/(192Da).

Find a complementary solution of the fann


weer ,8) =wl(r) cos 8,

Foundations of Solid Mechanics

186

where w1(r) is as defmed in Eq. 5.99b, but with C1 and Dl zero to make the
solution bounded at r = O. The remaining two arbitrary constants can be
determined from the boundary conditions at r = a, i.e.
w(a; a) =0, Mr(a, a) =O.
Answer:
pr(a 2- r2) [a 2(7 + v) - (3 + v)r~
a
w(r, a)
192(3+v)Da
cos.
5.10.13

Solve the preceding problem, if the plate outer edge is free, and the plate
region 0 :s; r :s; b (b < a) is a rigid core and fixed in space.

5.10.14

Suppose that the plate in Problem 5.10.13 has its outer edge simply supported
and the rigid core is rotated by a moment Ml about the y-axis through a small
angle 4. Find the relationship between Ml and 4. Hint: The (complementary)
solution takes the same form as in Problem 5.10.12, but wl(r) is given by Eq.
5.99b. Boundary conditions along the circumference of the core, r = b, are
w(b,a)=-b4cosa, awr,a)

=-cIcosa.

r=b

The relationship between Ml and 4 can be obtained by considering the moment


equilibrium of the core, i.e.
Ml =b

L
2Ic

[bVr (b,a)-M r (b,a)] cos ada.

5.10.15

Solve the preceding problem when the outer edge is flXed. Then specialize
both problems for the cases of a concentrated moment Ml at the center of the
plate, to prove Eqs. 5.90 and 5.92.

5.10.16

An annular plate is segmented along two radii in such a way that the straight
edges of the plate subtend an angle a at the plate center. Formulate the solution
when the plate is subjected to the load q(r,a) and the straight edges of the
plate are simply supported. Hint: Set the coordinate system so that one straight
edge is at a = 0 and the other at a = a, and one circular edge is at r = a and
the other at r =b. Adopt a 'polar' Levy-Nadai solution, i.e.
m1ta
w(r,a)= l: w",(r)sin-,
",=1

which satisfies the boundary conditions along the straight edges exactly.
Expand q(r,a) in a similar sine series. Determine arbitrary constants by means
of two boundary conditions along each circular edge.
5.10.17

Find the solution for the same annular plate of Section 5.7 when the inner

187

Bending of Elastic Thin Plates

edge is fixed instead. Hint: The column reaction distribution in Eq. 5.100a
should be modified to read
/(e)

111t(a 2 - b 2)qo
2kaa
'

where 11 is the fraction of the total load to be carried by the columns. Arbitrary
constants involved are to be determined from the boundary conditions;
w(b,e)=O,

dw~~,e) Ir=b

=0,

Mr(a, e) 0, Vr(a, e) -/(e).


The last unknown constant 11 can be evaluated from w(a,O)
Kosherick, Dundurs and Lee (1968).

=0.

Reference:

5.10.18

Solve the same annular plate of Section 5.7 when the supports are two concentric rings of equally spaced columns, situated in between and concentrically
with the boundary circles. The number of columns in each ring is k. Each
pair of opposite columns in different rings subtends the same angle 2a at the
center of the boundary circles.

5.10.19

Find the solution to the problem of an annular plate on elastic foundation and
supporting a circular ring of k equally spaced columns. The ring and the
boundary circles are concentric. The radius of the ring is c. Hint: Set the
reference coordinate system as in the example of Section 5.7. The column
force P on the plate can be put as a distribution in the form
/(e)=PI(2kac),

=0,

e;j _a<e<2;j +a,

j=O, 1,2, "',k)

(elsewhere).

The solution can assume the following series


w(r,e)

=m=O,,t,2k,
L ,,' wcm(r) cosme,

where wcm(r) is as given in Eq. 5.129. Boundary conditions are free conditions
along both circular edges, and four more conditions along the column ring.
The latter are

w1(c, e) =w2(c, e),

dw 1(r,e)

M r1 (c, a) = M r2(c, a),

r=c

=aW2(r,e)
a'
r

r=c

V r1 (c, a) - Vric, a) = /(e),

in which subscripts 1 and 2 denote respective domains; domain 1 where


b ~r ~ c, and domain 2 where c ~r ~a.

Foundations of Solid Mechanics

188

5.10.20

Study the bending by a unifonnly distributed load of a very long rectangular


flat slab supported on three parallel rows of equidistant columns as shown in
Fig. 5.17. It is assumed that the columns offer no restraint against rotation of
the slab and that the reaction of each column is a line load unifonnly distributed over a small interval in the same direction as the longitudinal column
lines. Reference: Karasudhi, Alam and Lee (1973).

2d

c
b

Longitudinal
column lines

xl

T~IT

line of
symmetry

~a

Fig. 5.17 Long flat slab on three rows of columns.


5.10.21

Fonnulate a rational analysis for a rectangular plate with eccentric stiffeners


as shown in Fig. 5.18. Then approximate the governing equations to obtain a
single fourth order differential equation similar to that for bending of an
ordinary plate, Eq. 5.34. Reference: Hovichitr, Karasudhi, Nishino and Lee
(1977).

Jr' --~~~~~ :~~ne_~~_~I~~:~~~"x


h

//

"

J",q(XTY)

,
,,

"

I.

WI---f--J-- Ribs

by

Fig. 5.18 Plate with eccentric stiffening ribs.

189

Bending oj Elastic Thin Plates

5.10.22

Since Reissner (1948, 1949) fonnulated a small deflection theory for bending
of isotropic sandwich plates, there has been a rapid development in the theory
and its application to the design of structural elements. A detailed literature
survey up to 1956 was made by Habip. Kao (1965, 1969, 1970, 1973) treated
problems of circular sandwich plates. Folie (1970) solved the bending of
orthotropic sandwich plates by means of a numerical integration scheme.
Discussions on bending and buckling of sandwich systems were made by
Plantema (1966). Karasudhi, Ng and Lee (1977) studied axisymmetric bending
of isotropic sandwich plates on elastic foundations as illustrated in Fig. 5.19.
The main assumptions in the analysis of bending of an elastic sandwich plate
are that in the core there are only transverse shear stress components unifonnly
distributed across the core thickness, and in the facings there are only in-plane
stress components unifonnly distributed across respective thicknesses. Solve
Problem 5.10.19 for isotropic sandwich plates.
Top Facing

t, t::~====~~~::::::~
h

Core

Bottom Facing

t2

~-E:-;:;g;'" Elastic Foundation

z
Fig. 5.19 Sandwich plate on elastic foundation.
5.10.23

Find an approximate solution to the problem of unifonnly loaded rectangular


orthotropic plates supported at the corners. Reference: Lee, Karasudhi, Zakeria
and Chan (1971).

5.10.24

Solve Problem 5.10.16 if the plate is an orthotropic sandwich plate. Reference:


Karasudhi, Jou and Tansirikongkol (1976).

190

Foundations

Table S.l

of Solid Mechanics

Complementary and Particular Solutions for Elastostatic Bending of Thin


Isotropic Plates.

Complementary solutions, ~w

=0 .

dwldr

MrlD

MelD

MrelD

QrlD

Qe lD

r2

2r

-2(l+v)

-2(1+v)

logr

r-1

(1-v)r-2

-(l-v)r-2

r 2logr

r(l +2Iogr)

-2(1 +v)logr
-3-v

-2(1 +v)logr
-1-3v

-r 4

-1

11r

r3

rlogr

CJwlCJr
(cos e, sin e)

1
-r2

3r 2

logr + 1

MrlD
(cose,sine)

-2(3+v)r

_ (1 +v)
r

MelD
(cos e, sin e)

2(I-v)
r3

-2(1 +3v)r

_ (1 +v)
r

MrelD
(sin e, -cos e)

2(1-v)
r3

2(I-v)r

(l-v)
r

QrlD
(cos e, sin e)

-8

2
r2

Qe lD
(sine,-cose)

-r2

VrlD
(cos e, sin e)

2(I-v)
r4

-2(3+v)

(3-v)
r2

VelD
(sine,-cose)

2(S-v)

(l+v)
r2

(cose,sine)

2(I-v)
r3

6(1-v)
r4

",-1

r lJl +3

r 3- 1J1

(l-v)m(m -1)(m -2)


r3-",

(cosme, sin me)

VelD

(sin me, -cosme)

(m + l)m[(I-v)m -4]
r 1-",

4(m + l)mr",-1

-4(m + l)mr"'-1

(l-v)m(m + l)r'"

+(I-v) (m + 2)]r'"

(m + 1)[-4

+(I-v)(m +2)]r'"

(l-m)m[4+(I-v)m]
r",+1

4(1- m)mr ...... - 1

4(I-m)mr ...... - 1

(l-v)m(I-m)r ......

(m-l)[4
-(1-v)(2-m)]r ......

+(I-v)(2-m)V ....

(m-l)[4v

(-m+2)r-+1

r-+z

r",+3

r 1- 1J1

r",+1

(l-v)m(m + l)(m +2) (m + l)m[(I-v)m +4] (l-m)m[4-(I-v)m]

(l-v)mz(m + 1)

(l-v)m z(m -1)

(sinme,-cosme)

VrlD

-(I-v)m(m + l)r...... - z

(l-v)m(m + l)r--z

(l-v)m(m -1)r"'-z

(l-v)m(m -1)r"'-z

(m +2)r"'+1

-mr-- 1
-(m+ 1)[4v

r"'+z

r-

-(l-v)m(m+ l)r--z

-1,0, 1).

- (1- v)m(m -1)r"'-z

mr

r'"

Qe lD

(cosme, sinme)

QrID

(sinme,-cosme)

MrelD

(cosme, sin me)

MelD

(cos me, sin me)

MrlD

(cosme, sinme)

owlor

(cosme, sin me)

Complementary solutions, ~w =0, (m

\0

--

;;

"tJ

~.

s~

Foundations of Solid Mechanics

192

Particular solutions, V4w =qlD ;/: O.


qID

dwldr

r4
-64

r3

rS
225

MrID

MaID

MralD

QrID

QalD

16

(3+v)r2
16

(1 +3v)r 2
16

--r2

r4
45

(4+v)r3
45

(1 +4v)r3
45

--3

r2

qlD
(cosme, sinme)

1
(m ;/: 0, 2, 4)

r
(m ;/: 0, 3, 5)

(cosme, sinme)

r4
64-20m 2+m 4

rS
225-34m 2+m 4

awldr
(cosme, sinme)

4r 3
64-20m2+m4

5r 4
225-34m2+m4

MJD
(cosme, sinme)

(vm 2 - 4v - 12)r2
64-20m2+m4

(vm 2 - 5v - 20)r3
225 -34m 2+m 4

MaiD
(cosme, sinme)

(m 2-12v-4)r 2
64-20m2+m4

(m 2- 20v - 5)r3
225-34m2+m4

MralD
(sinme, -cosme)

3m(l-v)r 2
64-20m2+m4

4m(1-v)r 3
225-34m2+m4

QrID
(cosme, sinme)

-2r
2
m -4

-3r
2

Qa lD
(sin me, -cosme)

-mr
m 2-4

-mr2
m 2-9

VrID
(cosme, sinme)

[(5 -3v)m 2-32]r


64-20m2+m4

[(7 -4v)m 2-75]r2


225-34m2+m4

ValD
(sinme, -cosme)

m(22-6v-m 2)r
64-20m2+m4

m(37 -12v-m 2)r2


225-34m2+m4

m -9

Bending of Elastic Thin Plates

193

Particular solutions, V4w =qlD


qlD
(cos 29, sin 29)

y!:

O.

qlD
(cos49,sin49)

(cos 29, sin 29)

r410gr
48

(cos49,sin49)

dwli:)r
(cos 29, sin 29)

(410gr + 1)r3
48

dwlor
(cos 49, sin 49)

1
_r 410gr
96
(410gr + 1)r3
96

MrlD
(cos 29, sin 29)

(7 +v+ 1210gr)r2
48

96Mr1D
(cos 49,sin49)

[12(1-v)logr
+7+vJr2

MelD
(cos 29, sin 29)

(1 +7v+ 12vlogr)r2
48

96Me/D
(cos 49, sin 49)

[12(v-1)logr
+1 +7vJr2

Mre lD
(sin 29,-cos 29)

(1-v)(310gr + 1)r2
24

Mre lD
(sin 49,-cos 49)

QrlD
(cos 29, sin 29)

(7 + 610gr)r
12

QrlD
(cos 49, sin 49)

-6r

Qe/D
(sin 49,-cos 49)

--r3

VrlD
(cos49,sin49)

[v-3(1-v)logrJr
6

24Ve/D
(sin 49,-cos 49)

[6(v-1)logr
+Sv-l3]r

Qe/D
(sin 29,-cos 29)
12VrlD
(cos 29, sin 29)
24Ve/D
(sin 29, -cos 29)

(2 + 310gr)r
6
-[3(1 +v)logr
+6+vJr
[6(3-v)logr
+13-SvJr

(1-v)(310gr + l)r2
24

Foundations of Solid Mechanics

194

Particular solutions, V4w =qlD


qlD
(cos 39, sin 39)

'#

o.

qlD
(cos 59, sin 59)

rSlogr

(cos 39, sin 39)

96

(cos 59, sin 59)

160

Ow/ar

(51ogr + 1)r4
96

Ow/ar
(cos 59, sin 59)

(51ogr + 1)r4
160

(cos 39, sin 39)

96M,ID
(cos 39, sin 39)

96MeiD
(cos 39, sin 39)

M,e lD
(sin 39,-cos 39)

Q,ID
(cos 39, sin 39)

QelD
(sin 39, - cos 39)

96V,ID
(cos 39, sin 39)

32VeiD
(sin 39, -cos 39)

-[4(5 -v)logr

+9+v]r3
[4(1-5v)logr

-1-9v]r 3

(1-v)(41ogr + 1)r3
32
(23 + 24 log r )r2
48
(5 + 81ogr)r 2
16
-[12(1 +3v)logr

+9v+37]r 2
[4(7 - 3v)logr

+17 -7v]r 2

160M,ID
(cos 59, sin 59)

160MeiD
(cos 59, sin 59)

M,e lD
(sin 59, - cos 59)

Q,ID
(cos 59, sin 59)

_rSlogr

[20(1-v)logr

+9+v]r3
[20(v-1)logr

+1 +9v]r 3
(1-v)(41ogr + 1)r3
32

-3r

16

QelD

5r2

(sin 59, -cos 59)

16

32V,ID
(cos 59, sin 59)

32VeiD
(sin 59,-cos 59)

[20(v -l)logr

+1 +5v]r 2
[12(v-1)logr
+7v-17Jr2

Bending of Elastic Thin Plates

195

Particular solutions, Vlw =qlD ~ O.

!L
D

r26}

r~{~i
smO

rQogr1{~~i
S

!L

. 20
sm

576

~-2

{=46}

(12Iogr+5)r 4Iogr ---2304


sin 40

r"
(n

(n

t'26}

(6logr -12)r4Iogr ____

. 0
sm

t'46}

logr ---sin 40

logr ---sin 20

or -4)

-r2

r4

r"+4

r2(logri

(n + 2)2 (n +4)2

qlD
(cosmO, sinmO)

(cosmO, sinmO)

r"

r"+4

~m

-2 or -1)
r m- 2

(m

~O

or -1)

r m- 4

(mO or 1)
logr

(m 2 or 4)

(logr)2
S

[(n +2)2-m~ [(n +4)2-m~


rm+2logr
Sm(m +1)
rmlogr
Sm(m -1)

[12(m2-S)+(m2-4)(m2-16)logr]r4
(m2_4)2(m2_16)2

CHAPTER VI
ELASTOSTATICS WITH DISPLACEMENTS AS
UNKNOWNS

6.1 FIELD EQUATIONS FOR PLANE PROBLEMS


The Cartesian field equations for the elastostatics of an isotropic solid in plane
strain may be obtained as a special case of Eq. 3.66, namely

J1(~U;+1~2Vekk';)+X;=O'

(i,k=1,2)

(6.1)

where
(6.2a)
(6.2b)

Due to Eq. 4.8b, we have

Substituting Eq. 6.3 into Eq. 6.1 yields field equations for plane problems as

~~ +

J1(

K:

1 ekk,;) +X; =0

(6.4)

The equation above holds for plane stress as well as plane strain problems depending
on which definition of K in Eqs. 4.8 is used.
Similarly, we can obtain freld equations in tenos of cylindrical displacement
components by specializing from Eqs. 3.67; thus

u,.

aUe]

2
2
J1 [ K_1ar(ekk)+~U'-r2- r2aa +X,=O

au,. Ue]

2 1
2
J1 [ K_1rd9(ekk)+~Ue+ r2d9 - r2 +Xe=O

(6.5a)
(6.5b)

197

Elastostatics with Displacements as Unknowns

where
(6.6a)

1[

=;: ara (ruT) + aUeJ


ao
torsion-free defonnation, Ue
ea

For axisymmetric
dependence on O. Thus Eqs. 6.5 reduce to

(6.6b)
must vanish, and there is no

J1:~ ~{![;! (rUT)]} +XT=O

(6.7)

the general solution of which can be obtained readily by successive integration with
respect to r. In fact, its complementary solution is a linear combination of r and r-l
Problems 6.9.1 and 6.9.2 should serve as good examples of this approach.
For torsion about the z-axis, U, must vanish, and all functions are independent of
O. The non-trivial field equation is

d[1 d

Il dr ;: dr (rUe>J + Xe =0

(6.8)

which is similar to Eq. 6.7. Problem 6.9.3 should serve as a good example of this
approach.
A solid medium is said to be in an antiplane state (of shear) with respect to the
xy-plane if U =v =0, W =w(x,y). There is just one field equation
J1~w +Z =0
which has any plane harmonic function as a complementary solution, i.e. ~w

(6.9)

=O.

6.2. SOLUTION SCHEME FOR LARGE PLANES


In this section we shall consider the elastostatics of an infinitely extended plane.
This may be an infinite plane (-00 <x, y < 00), a half plane (-00 <x < 00,0 S Y < 00), or
a layer infinitely extended in the x -direction (-00 < X < 00) but finite and constant in the
y-direction.
The field equations for in-plane (plane stress or plane strain) problems, Eq. 6.4,
in the absence of the body forces, i.e. with

Xl =Xz =0

can be expressed in explicit symbols as

~U + lC-l a/Ex.: + e,,) =0

(6.10)
(6. 11a)
(6.11b)

Foundations of Solid Mechanics

198

We seek solutions to these equations in the fonn of Fourier integrals (Sneddon,


1951)
11~

.
.

u(~,y)e"'~d~

(6. 12a)

11~
v(x,y) =v(~,y)e"'~d~

(6. 12b)

u(x,y) =-

1t_

1t_

in which a tilde (-) denotes a Fourier transform. Note that the Riemann-Lebesgue lemma
(Whittaker and Watson, 1965) ensures the boundedness of the solution as Ix I~ 00.
Substituting Eqs. 6.12 into Eqs. 6.11, we find

J2 ~ du

(_1:2 - . dV) =0

(6. 13a)

d 2 ( -i~-+dU dZV) =0
~v+~+-

(6.13b)

2
I:
~u+-+-- ~U-l~-

dy2

K-l

dy

dy2

K-l

dy

dy2

Seeking u(~,y) and v(~,y) in the fonn


wefmd
or

a=-;, -;,

~, ~.

Because of the double roots above, the functions u(~,y) and v(~,y) must be linear
combinations of the functions
e~, ye~, e~ and ye~.
Equations 6.13 show that if
u(~,y)

=(A +By)e~ + (C +Dy)e~

(6. 14a)

then
(6.14b)
The constants A, B, C and D must be detennined from four boundary conditions at the
two boundaries, say at y =YI and y =YI + h, if h is the thickness of the layer.
When the loading is symmetrical with respect to the y-axis, we seek the solution
in the fonn

=~ r~ u(~,y)sin~d~

(6. 15a)

2i~ v(~,y)cos~d~
v(x,y) =-

(6.15b)

u(x,y)

1tJo

1t

where

Elastostatics with Displacements as Unknowns

199

l-

u(~,y) = u(x,y)sin~dx

(6. 16a)

=(A + By)e-9 +(G +Dy)e't.Y


v(~,y)=

l- v(x,y)cos~dx

=[A +B(K~-l + y)]e-9 -

(6.16b)
(6. 17a)

[G +D(-K~-l +y)]e't.Y

(6. 17b)

The unified Hooke's law, Eqs. 4.6, gives the stress components

2l2l--

cr... (x,y)=-

1t 0

crxy(x,y)=-

1t 0

cr.. (~,y)cos~d~

(6.18a)

't(~,y)sin~d~

(6.18b)
(6.18c)

where
(6. 19a)
(6. 19b)
(6. 19c)
Consider a plane being composed of N parallel homogeneous layers, with differing
elastic moduli. The solution for layer n will involve four arbitrary constants; A", B", G"
and D". Altogether, there will be 4N arbitrary constants to be determined from boundary
conditions at the two surfaces of the whole plane and at the N - 1 interfaces between
layers. Considering the interface between layer n -1(Y,,_2 ~ Y ~ Y,,-l) and layer
n(Y"-l ~ y ~ y,,). If there are no applied tractions at the interface, u, v, crY)' and crxy must
be continuous, i.e.

=U"_l(~'Y)
v,,(~,y) =V"_l(~'Y)

(6.20b)

cr,,(~,y) =cr"_l(~'Y)

(6.2Oc)

t,,(~,y) =t"_l(~'Y)

(6.2Od)

u,,(~,y)

(6.20a)

Here the subscripts n -1 and n refer to the respective layers, and y is identical to Y,,-l.
If tractions are applied to the interface between layers n - 1 and n, which mayor may

200

Foundations of Solid Mechanics

not have different elastic moduli, then the requisite conditions are

=all_l(~'Y)

(6.21a)

VII(~'Y) =V"_l(~'Y)

(6.21b)

a,,(~,y)

o-II(~'Y)+ y (~,y) =o-lI_1(~'Y)

(6.21c)
(6.21d)

Here

is the Fourier sine transform of the applied traction in the x-direction, i.e.
(6.22a)

and, similarly,
(6.22b)

At the top surface of the first layer (n 1, Yo :s; Y :s; Yl) which is also the top surface of
the whole plane, the boundary conditions will involve

al(~'Yo) or il(~'Yo)

(6.23a)
(6.23b)

However, if this particular layer is an overlying half plane, the arbitrary constants
associated with exponentials of negative Y must be set to zero to assure the boundedness
of the solution as Y -+ Yo -00, i.e. Eqs. 6.23 must be replaced by

Al =BI =0

(6.24a,b)

At the lower surface of the last layer (n = N, YN -I :s; Y :s; YN) which is also the bottom
of the whole plane, the boundary conditions will involve

aN(~'YN) or iN(~'YN)

(6.25a)
(6.25b)

However, if this particular layer is an underlying half plane, the arbitrary constants
associated with exponentials of positive Y must be set to zero to assure the boundedness
of the solutions as Y -+ YN =00, i.e. Eqs. 6.25 must be replaced by
CN=DN=O

(6.26a,b)

If the loading is antisymmetrical with respect to the y-axis, the displacement and
the stress functions should take the form

2i-

u(x,y) =-

1t 0

a(~,y ) cos I;xd~

(6.27a)

201

Elastostatics with Displacements as Unknowns

V(X,y)=~

r- v(~,y)sinl;xd~

1tJo

21-2121--

(J.,,(X,y)=-

(J%(~,y)sinl;xd~

(6.27c)

(J",(x,Y) =-

t(~,y)cosl;xd~

(6.27d)

(Jyy(X,y)=-

(J(~,y)sinl;xd~

(6.27e)

1t 0

1t 0

1t 0

where

(6.27b)

u(~,y) =(A +By)e--9 +(C +Dy)e9

(6.28a)

v(~,y) =-[A +B(lC~-l + y)]e--9 + [C +D(-lC~-l + y)]e 9

(6.28b)

(6.28c)

(J%(~,y)

~ [-(lC+ 1)~ + (3 -lC) dy


dV]
=lC-l

-'t(~,y)=~ (dU)
dy +~v

(6.28d)

(j(~,y) =lC~ 1[(lC+ 1) :; - (3 -lC)~U]

(6.28e)

In addition, the Fourier transfonDs of the applied tractions are

X(~,y) =L-X(X,y)cosl;xdx

(6.29a)

y(~,y) =L-Y(X,Y)Sinl;xdx

(6.29b)

The relevant boundary conditions for this case are the same as in the previous case, i.e.
Eqs. 6.20 or 6.21, 6.23 or 6.24, and 6.25 or 6.26.
For antiplane problems where loading is symmetrical with respect to the y-axis,
the solution to the field equation (Eq. 6.9) for Z = 0 takes the fOnD

21-

W(X,y) =-

1t 0

and the only stresses are


(Jyz(X, y)

w(~,y)cosl;xd~

2L2L--

=-1t

(J.. (X,y)=-

1t 0

(6.30a)

t(~, y) cos l;xd~

(6.30b)

't%(~,y)sinl;xd~

(6.3Oc)

202

Foundations of Solid Mechanics

where
w(~,y)=Ae-9 +Bef,y
-

dw

t(/;,y) =J.L dy
i%(~, y)

(6.31a)
(6.31b)

=-J.L~w

(6.31c)

At the interface between the layers n - 1 and n, boundary conditions in the absence of
any applied traction are
(6.32a)
WII(~'Y) WII_I(~'Y)

=
ill(~'Y) =ill_l(~'Y)

(6.32b)

If a traction in the z-direction is applied at this interface, Eqs. 6.32 have to be replaced
by
(6.33a)
WII(~'Y) =WII_I(~'Y)

ill(~'Y) +2 (~,y) =ill_l(~'Y)

(6.33b)

where
(6.34)
in which 2(x,y) is the applied traction. At the top surface of the ftrst layer (n
boundary conditions are in

=1),

(6.35)
However, if this particular layer is an overlying half plane, the equation above must be
replaced by
Al =0
(6.36)
At the lower surface of the last layer (n

=N),

boundary conditions are in

WN(~'YN) or iN(~'YN)

(6.37)

However, if this particular layer is an underlying half plane, the equation above must
be replaced by
(6.38)
For antiplane problems in which the loading is antisymmetrical with respect to the
y -axis, the displacement and stress functions take the form

2l~ w(~,y)singd~
w(x,y) =1t 0

cryz(x,y)=-2l~-t(~,y)singd~
1t 0

(6.39a)
(6.39b)

Elastostatics with Displacements as Unknowns

203

(6.39c)
where
w(~,y) =Ae-9 +Be~

dw

(6.40a)

't(~,y) = J.1 dy

(6.40b)

iz(~'Y)=~w

(6.4Oc)

In addition, the Fourier transfonn of the applied traction is

i (~,y) =LooZ(x,Y)Sinl;xdx

(6.41)

The relevant boundary conditions for this case are the same as before, i.e. Eqs. 6.32 or
6.33, 6.35 or 6.36, and 6.37 or 6.38.
Technically, the solution scheme described so far is complete for any multilayered
planes. However, the exact closed form solution is only possible for a homogeneous
half plane subjected to simple loading types. For more complicated cases, approximate
solutions may be approached in similar manners as in the three dimensional cases which
are discussed in Sections 6.5, 6.6 and 6.7.
6.3 SOLUTION SCHEME FOR LARGE SPACES
This section follows the work of Muki (1960) to develop a method applicable to
the elastostatics of an isotropic solid domain which is infinitely extended radially
(0 s:; r < 00). The domain may be an infinite space (-00 < Z < 00), a half space (0 s:; z < 00),
or an infmitely large plate with a finite constant thickness (zo s:; z S ZN)' In cylindrical
coordinates (r, z), the equations of equilibrium, in the absence of the body force, are
given by Eqs. 3.67 with
X,=Xe=X.=O
(6.42)

e,

It may be verified by direct substitution that these statical field equations are satisfied
by the displacements

ifCl 2d'
=---+-,
drdz r de

if4>

Ug=- rdedz

d'
-2 dr

u = 2(1 _ v)V24> _ ifCl

dz 2
provided that the displacement potentials CI and ' satisfy

(6.43a)
(6.43b)
(6.43c)

Foundations of Solid Mechanics

204

(6.44a,b)

In these equations, v is the Poisson's ratio, VZ the Laplace operator (Eq. 3.69b), and ~
the biharmonic operator, i.e. ~ VZVZ. For the solution to be singlevalued with respect
to a, <I(r,a,z) and '(r,a,z) can be written in Fourier series of a with 21t as the period,
i.e.

<I(r,a,z)=
'(r,a,z)=

:-

: [wm(r,z)cosma-<Im(r,z)sinmEJ]

(6.4Sa)

['m(r,z)sinma+'m(r,z)cosmEJ]

(6.45b)

m=O

m=O

We will assume that wm ='I'm =0, and develop solutions corresponding to <lm and 'I'm.
If needed, the solution corresponding to Wm and 'I'm may be obtained by making the
replacements
'I'm by 'I'm'

Wm by <lm'

cosma by - sinma,

and sinma by cosma.

Substituting Eqs. 6.45 into Eqs. 6.44, we see that <lm and 'I'm satisfy
v:,<Im(r,z)

=0

(6.46a)

V!.'m(r,z)

=0

(6.46b)

where
(6.47)
Consider the case that a homogeneous material layer is unbounded in the radial
dimension but bounded by two infmite planes normal to the z-axis, say Z,,_l S; z S; z". If
the solution to Eqs. 6.46 is to be nonsingular for the whole range of r (0 S; r < 00), then
it must take the form
wm(r,z)
'm(r,z)

1=1- ~'m(~,z)Jm(~r)d~
= ~cim(~,z)Jm(g)d~

(6.48a)
(6.48b)

where J m( ) is a Bessel function of the ftrst kind (see Section 1.10), and the upper
tilde C) denotes a Hankel transform (Sneddon, 1951), i.e.

1-

<im(~'z) =
'm(~'z)=

L-

r<lm(r,z)Jm(g)dr

(6.49a)

r'm(r,z)Jm(g)dr

(6.49b)

Substituting Eqs. 6.48 into Eq. 6.46, and using the Bessel equation (Eq. 1.111), we fmd

205

Elastostatics with Displacements as Unknowns

that

&", and qt'" must satisfy the ordinary differential equations

(:2-~2Jd>",=O

(6.S0a)

(:2-~2!",=O

(6.S0b)

The general solutions of these equations are


4",(~,z)
=(A", +B",z)e9 +(C", +D",z)e~
-

'",(~,z)-E",e +F",e

(6.51a)
(6.S1b)

Here A", to F", are arbitrary constants to be determined from the Hankel transform of
the given boundary conditions on the two bounding infmite planes z =Zll_l and z = ZII'
The displacement and stress components can be put in terms of &", and qt'" as follows:
1 ..
u, ="2",;o[U"'+l(r,z)-V"'_l(r,z)] cosm9
(6.52a)
1 ..

Ue= "2",;o[U",+l(r,z)+ V"'_l(r,z)] sinm9

(6.52b)
(6.S2c)

(6.53a)

(6.53b)
(6.53c)

(6.S3d)

Foundations of Solid Mechanics

206

(6.53e)

(6.531)
where

L-(d!m +2'1'm)~2Jm+l(~r)d~,
Vm_1(r,z) =L-( d!m -2'1'm)~2Jm_l(g)d~.

Um+1(r,z) =

(6.54a)
(6.54b)

At a point on each of the boundary planes (where z =Z,,_l or z,,), there are three boundary
conditions, i.e. involving the prescription of either

and u, or 0,.
Before applying these boundary conditions, we fmd it convenient to rewrite Eqs. 6.52a,
6.52b, 6.53d and 6.53e so that the mth equation involves only Jm(g); thus
Urm
--=

cosma

l-[
0

~=
sinma

r-[m dci>mJm(g)
-2'1' dJm(g)]~2d~
dz
g
m d(g)

[d ci>m

= r-{

zrm

-l-{-[

Jo m V dz2 + (1

-V
cosm a
0

):2- ]

v)':I cl>m

(6.55a)
(6.55b)

Jo

09zm

sinma

dci>mdJm(g)
- Jm(g)]):2d):
+2m' - - ':I ':I
dz d(g)
m g

Jm(g) _ d'l'mdJm(~r)}):2d):
g
dz d(~r) ':I ':I

d2ci>m
):2- ]dJm(g)
d'l'mJm(~r)}):2d):
2 +(1-V)':IcI>m
d():) +m dz
):
':I ':I
dz
':Ir
':Ir

(6.56a)

(6.56b)

For axisymmetric torsion-free defonnation, Ua must vanish, all functions must be


independent of a, so that cI> = cl>o, ' = O.
For torsion about the z-axis, Ur and u, must vanish, all functions must be independent of a, and the only non-trivial equilibrium equation is
in which

Vi is defined by Eq.

=0
with m = 1. Thus Ua(r, z)
V~Ua(r , z)

6.47

(6.57)
assumes the solution in

Elastostatics with Displacements as Unknowns

207

the fonn
(6.58)
where A and C are arbitrary constants to be detennined from boundary conditions. The
non-vanishing stress components are
crez = J.L

L- ~z(Aeg Ce....g)Jl(~r)d~
L- ~z(Aeg +Ce....g)Jz(g)d~
-

crr9 =-J.L

(6.59a)
(6.59b)

6.4 HOMOGENEOUS HALF PLANES AND HALF SPACES


The flrst example to be discussed in this section is the problem of a homogeneous
isotropic elastic half plane -00 < x < 00, 0 ~ y < 00 subjected to a nonnal indentation by
a rigid plate of flnite width 2a. The contact area between the rigid plate and the half
plane is assumed to be smooth. The Fourier transfonns of displacement components
given in Eqs. 6.16 and 6.17 are
a(~,y)=(A+By)e-{y
(6.60a)
(6.60b)
and the boundary conditions are
V(x,O)

=vo,

cryy(x, 0) =0,
crxy(x, 0) =0,

~a)

(6.61a)

Ix I > a)

(6.61b)

<00)

(6.61c)

(Ixl

(O~lxl

where Vo is the prescribed unifonn indentation. The last equation above gives t(~, 0) = 0
or, because of Eqs. 6.60 and 6.19b,
2!;A + (lC-1)B = O.
On the other hand, Eq. 6.19c gives
2!;A + (lC+ 1)B =

cr(~,O).

J.L
Solving the last two equations above yields
A = (lC-1)cr(~,0)
4~J.L

B = cr(~,O)

(6.62a)

(6.62b)
2J.L
Using these results in the remaining boundary conditions, Eqs. 6.61a and b, we get

Foundations of Solid Mechanics

208

respectively

.. 21tj.l.vo
So G(;, 0);-1 cos l;xd; =--1'
K+

So" a(;,O)cosl;x~ =0,

(O~x ~a)

(6.63a)

(a <x <00)

(6.63b)

Equations 6.63 are called dual integral equations; they may be solved by schemes
described by Sneddon (1951 and 1966). Note thatEq. 6.63a states that v(x, 0) =constant,
for Ixl ~ a. However, since any static half plane problem has a rigid body displacement
solution, the actual value of the constant cannot be found. So we replace this equation
by

dv(x, 0) 0
dx
in the same region, i.e.
(O~x ~a)

The general solution of Eqs. 6.63b and 6.64 is


a(;, 0) =Crlo(~)

(6.64)
(6.65)

where 10 is the Bessel function of the flrSt kind and of zero order, and Co is a constant
to be determined in tenns of the total applied force Po. Substituting Eq. 6.65 into Eq.
6.18c for y =0 yields the contact normal stress
Gyy(x, 0) = (2
1t

a -x

(OSx<a)

2)112'

=0,

(6.66a)

(a <x <00)

(6.66b)
Note that the contact normal stress is singular at the edges of the rigid plate. The total
applied force Po is

or
(6.67)
The results are identical to those obtained by Sadowsky (1928).
The second example to be discussed is an antiplane problem of a homogeneous
isotropic elastic half plane. The solid region is denoted by -<XI < X < 00, 0 ~ y < 00. A
traction is applied on the half plane surface such that
dw(x,O) =k,

(Ixl

~a)

(6.68a)

G,.(x, 0) =0,

(Ixl >a)

(6.68b)

dx

Elastostatics with Displacements as Unknowns

209

Since the loading is antisymmetric with respect to the y-axis, the solution is given by
Eqs. 6.39 to 6.41. Since the solution must be bounded as y tends to infmity, B in those
equations must be set to zero, thus
w(!;,y) =Ae-9>,
~(!;, 0) = -J.t!;A .

Using these results in boundary conditions, Eqs. 6.68, we get


r~

Jo

1tkll
~(!;,0)cos1;xd!;=-2'

So ~ ~(!;, 0) sin 1;xd!; = 0,

<a)

(6.69a)

(a <x <00)

(6.69b)

(0

~x

These dual integral equations have the solution


- ):
_ 1tkaJ.l.ll(~)
t(':I'O)2

(6.70)

which when substituted into Eq. 6.39b for y = 0 yields the contact shear stress
O',.(X,o)=(

-J.1kx

(O~x<a)

2112'

a -x)

(6.71 a)

=0,
(a <x <00)
(6.71b)
Note that the contact shear stress is singular at x = a.
The third and last example to be discussed in this section is a three dimensional
problem of a homogeneous isotropic elastic half space. The solid domain is denoted
by 0 ~ r < 00, 0 ~ z < 00. A normal indentation is made on the surface of the half space
by means of a smooth rigid circular plate of radius a. Thus boundary conditions are
u.(r,e,O)=w o'

e, 0) = 0,
O'.,(r, e, 0) =0,
O'.. (r,

(O~r ~a)

(6.72a)

(a<r<oo)

(6.72b)

(O~r<oo)

(6.72c)

where Wo is the prescribed uniform indentation. This is an axisymmetric torsion-free


problem, thus U9 must vanish, and all functions are independent of e. Its solution can
be obtained from Eqs. 6.52 to 6.56 by setting m = 0 and q, WI = O. Since the solution
must be bounded as z tends to infmity, Eq. 6.51a simplifies to
ci>(!;,z) = (C +Dz)e-9

(6.73)

The last boundary condition, Eq. 6.72c, gives


!;C-2vD =0

On the other hand, Eq. 6.53c gives

(6.74a)

Foundations of Solid Mechanics

210

~3C+(1-2v)~2D = o-u(~,O)

(6.74b)
2J,1
where o-u(~, 0) is the Hankel transfonn of O'u(r, a, 0). Solving the last two equations
above yields
vo-u(~,O)
(6.75a)

~3

o-.. (~,O)

(6.75b)

2~2

Using these results in the remaining boundary conditions, Eqs. 6.72a and b, we get
respectively
J,1wo
(OSr Sa)
(6.76a)
Jo 0'.. (~,0)Jo(g)d~=-1_V'

r--

1-

o-.. (~, O)~o(g )d~ = 0,

(r >a)

(6.76b)

The solution of these dual integral equations is


-

O'u(~' 0) =

2J,1wosin I;a

(6.77)

1t(1- v)~

which when substituted into Eqs. 6.53c for z = 0 yields the nonnal contact stress
2J,1wo
)( 2 2)112'
1t -v a -r

O'u(r,a,O)=- (1
= 0,

(OSr <a)

(6.78a)

(r > a)

(6.78b)

Again note that the nonnal contact stress is singular at the edge of the rigid plate. The
total applied force Po to produce the indentation Wo is
P o=-

fJ

21t7O'u(r, a, O)dr

or
(6.79)
The results are identical to those obtained by Sneddon (1951). Note that for the half
space it is possible to obtain the 'stiffness equation', Eq. 6.79, linking Po and woo It
was impossible to fmd such a result for the half plane because it had an arbitrary
rigid-body displacement.

Elastostatics with Displacements as Unknowns

211

6.5 CONCENTRA1ED FORCE INSIDE A HALF SPACE


Mindlin (1936) was the ftrst to study the problem of a force at a point in the
interior of a homogeneous elastic half space. The solution was obtained by the method
of synthesis and superposition. It requires ingeneous guesses of proper potentials.
Mindlin (1953) reconsidered the same problem and solved it by means of Papkovich
potentials. The same potentials were applied by Rongved (1955) to fmd the solution for
the force in the interior of two joined semi-inftnite solids. In fact, these solutions can
be obtained by using the formulation described in Section 6.3. The present section
follows the work by Chan, Karasudhi and Lee (1974) who solved the problem of a
force at a point in the interior of a layered half space, as shown in Fig. 6.1. Chan et
al considered four cases, relating to a concentrated force applied either vertically or
horizontally in the interior of the upper layer or in the underlying half space. Only the
case of a horizontal force in the upper layer will be presented here to illustrate the use
of the formulation described in Section 6.3.
The force P acting in the x-direction at a depth z = z' below the free surface
produces a traction on the horizontal plane at that depth given by
_PO(r-E)
(
0)
) q( r
2
,E--+

(6.80)

1tr

where o(r) is the Dirac-delta function described in Section 1.9, and can now be represented by a Hankel transform
(6.81a)

x
Z'

Domain I

IIp V,
Domain 2
1l2= Il,
v2 = v,
Domain 3
1l3' V3

Fig. 6.1 A concentrated force in a layered half space.

Upper
layer

Lower
layer

Foundations of Solid Mechanics

212

or, due to Eq. 1.112c,


(6.81b)
where
{j(S)

r~

=Jo

r q (r )J o(sr)dr

=21t

(6.82)

Imagine a horizontal plane passing through the point of application of the force, and
dividing the material system into three domains; 0 < z < z' for domain 1, z' < z < h for
domain 2, and h < z < 00 for domain 3. Domains 1 and 2 have the same elastic properties.
For the boundedness of the solution as z tends to infmity, coefficients associated with
e9 in Eqs. 6.51 must be set to zero for the domain 3, i.e. Am3 = Bm3 = Em3 = O. The free
boundary conditions at the plane z =0 are
(6.83a,b,c)
0' 1(r, e, 0) = 0, 0',,1 (r, e, 0) = 0, 0'.91 (r, e, 0) = 0
The continuity conditions at the plane z =z' are
u r1 (r,e,z)

=urz(r,e,z)

(6.83d)
(6.83e)
(6.83f)
(6.83g)

0'.91(r,

e,z) - O'zQ2(r, e,z) =-q(r) sin e

O'zr1(r, e,z) - O'zrir, e,z) =q(r)cos

(6.83h)
(6.83i)

The interface continuity conditions at the plane z =h are


urir,e,h)

=ur3 (r,e,h)

(6.83j)
(6.83k)

u.z(r, e,h)

=u. 3(r, e,h)

(6.831)
(6.83m)
(6.83n)
(6.830)

The only nonhomogeneous conditions, Eqs. 6.83h and 6.83i, indicate that the solution
involves only the harmonic term m = 1. For simplicity, the subscript m will be omitted.
Substituting Eqs. 6.52 to 6.56 into Eqs. 6.83 yields a set of fifteen equations for the

Elastostatics with Displacements as Unknowns

213

fifteen coefficients, Al to FlJ A2 to F2 and C3,

D3

and F3 Introducing the dimensionless

terms

a=~, ~=zlh, p=rlh, /3=z'lh


we obtain the displacement components in the dimensionless forms
J.I1h
I
P cos au".(p, a,~) =:2 [u,,(p,~) + V,,(p, ~)]

(6.84a,b,c,d)
(6.85a)
(6.85b)

J.11h

p cos U.,,(p,

r1"

a,.,)r =Jo

w,,(a,~)]J1(pa)da

w,,(a.,~)+ D(a)

(6.85c)

where the subscript n denotes the domain number 1, 2 or 3, and

u,,(p,~) =l-{[U:(a,~)+ V:(a.,~)] +[u~~~)~) +V;i:;)]}J(pa)da

(6.86a)

V,,(p,~) = roo{[_u:(a,~) + V:(a., ~)] + [-u,,(a, 0 + v,,(a'~)]}Jo(pa)da

(6.86b)

Jo

D(a)

H(a)

D(a) =1- (a +b +4b<i)e-2a.+abe-4a

(6.87a)

1- J.1o
H(a) =l---e-2a.
1 + J.1o

(6. 87b)

(3 - 4v3) - J.1o(3 - 4v1)

a
b

=__(I_-_J.1o_)_
1 + J.1o(3 - 4v 1)

(6.88a)
(6.88b)
(6.88c)

Forn=1 and 2,

U:(a,~)= 16~0 [41 +(/3-~)a]e-l~-~la


V"(a., r)

"

...

=~e-l~-~Ia
41t

(6.89a)
(6. 89b)
(6.89c)

Foundations of Solid Mechanics

214

,,(a.,~) = 16~J[-(4'YO'Y2 + 1) +'Yl(/3 + ~)a.- 2/3~a.1e--@+Qa


+[k(a +bi)+b'Yl(2- /3- ~)a.+ 2b(1- /3)(1- ~)a.2Je-<2+~+Qa
+ab['Yl + (/3-

~)a.]e-<4-J3+Qa+{_~(a +b) + [a~ +b(8'Yo'Yl

+ 2- /3)] a.- 2b'Yl(1- /3+ ~)a.2 +4b(1- /3)~a.3}


+ab['Yl - (/3-

~)a.]e-<4+~-Qa+{-~(a +b) + [a/3 + 2b(4'Yo'Y2

+ 1) - bQa.- 2b'Yl(1 + /3 + ab [(4'YO'Y2 + 1) +'Yl(/3 +

+ b'Yl(2 - /3 -

V (a.

",

e-<2-~+Qa

~)a.2 + 4b /3(1- ~)a.3} e-<2+~-Qa

~)a.+ 2/3~a.1e-<4-J3-Qa +[-k(a + bi)

~)a. - 2b(1- /3)(1- ~)a.2Je-<2-~-Qa)

~) =~{e--@+Qa+ 1-1lo [e-<2-~+Qa+e-<2+~-Qa+e-<2-~-Oj}

w,,(a.,~) =

4n

1 + Ilo

16~J[4'YO'Y2 - 'Yl(/3 - ~)a. - 2/3~a.1 e--@+Qa


+[k(a -bi) +b'Yl(/3- ~)a.+ 2b(1- /3)(1- ~)a.2Je-<2+~+Qa
+ab(/3-

~)a.e-<4-~+Qa+{ ~(a - b) + [-b(/3+ 8'Yo'Y~ +aQa.

+ 2b'Yl(1- /3 +ab(/3-

~)a.2 + 4b(1- /3)~a.3}

e-<2-I3+Qa

~)a.e-<4+J3-Qa+[ ~(a -b)+ (b~ -a/3+ 8'Yo'Y~a.

+ 2b'Yl(1- /3 -

~)a.2 -4b/3(1- ~)a.3J

e-<2+J3-Qa

(6.90a)

(6.90b)

Elastostatics with Displacements as Unknowns

215

+ ab [4'Yo'Yz + 'Yt (J~ - ~)a - 2/3~a1 e-(4-fH)a

+[~(a -bit)-b'Yt(/3- ~)a+2b(1- /3)(1- ~)azJe-{Z-P-Qa)

(6.9Oc)

and for n =3,


U:(a,~) = V:(a,~) = w:(a,~) = 0

(6.91a,b,c)

1(

U,,(a,~) = 81t {-(c +d'Yt'Y3) + 2['Yt(c -d +d~)+df3'YJa

- 4/3(c - d + d~)a1e-{~+P)a + {(ac - bd'Yt'Y3)


+ 2bd['Yt(1-~) +'Y3(1- /3)]a+4bd(1- /3)(1- ~)a1e-{2+~+P)a
+ {(-C'Yt + d'Y3) + 2[c(1- /3) -d(1- ~)]a}e-{~-P)a
+ {(aC'Yt - bd'Y3) + 2[ac/3+bd('Yt'Y3 + 1- ~)]a
+ 4bd ['Yt (1 -

V (a ~) =
",

Ilil

~) + 'Y3f3l a Z+ 8bd /3( 1 - ~)a1 e-(Z+ ~ - P)a)

(6.92a)

[e-{~+P)a+e-{~-P>j

(6.92b)

21t( 1 + Ilil)

w,,(a,~) = 8~({-C +d'Yt'Y3 + 2['Yt(c -d +d~) -df3'YJa


-4/3(c -d +d~)a1e-{~+P)a+ {(ac -bd'Yt'Y3)
+ 2bd['Yt(1- ~)-'Y3(1- /3)]a+ 4bd(1- /3)(1- ~)a1e-{2+~+P)a
+ {(-C'Yt +d'Y3) + 2[c(1- /3) -d(1- ~)]a}e-{~-P)a
+ {(ac'Yt -bd'Y3) + 2[ac/3-bd('Yt'Y3 -1 + ~)]a
+

4bd['Yt(1-~) -'Y3f3l a Z+ 8bd/3(1- ~)a1e-{Z+~-P)a)

In Eqs. 6.89 to 6.92,


'Yo=1-v t, 'Yt=3-4v t, 'Yz=1-2v t, 'Y3=3-4v3

(6.92c)
(6.93a,b,c,d)
(6.93e,f)

Note that the solution functions take the same form for n = 1 and 2. Thus the
solution is given in the form of a two-domain problem. The integrals given by U:(a,~),

Foundations of Solid Mechanics

216

V:(a,~) and w:(a,~) correspond to the singularity of Mindlin's solution (1936), while
the integrals given by U,.(a,~), V,.(a,~) and w,.(a,~) are non singular. For a homogeneous half space, D(a) and H(a) (Eqs. 6.85c, 8.86 and 8.87) are equal to unity, and
the solution is Mindlin's solution (Problem 6.9.5 of the present chapter). To avoid
laborious numerical integration in Eqs. 6.85c and 6.86, the reciprocals of D(a) and H(a)
for the layered half space will be approximated in such a way that the integrals assume
the standard forms.
Since D (a) is positive for all values of a from zero to infInity and asymptotic to
unity as a approaches infmity, it is reasonable to approximate lID (a) by a series of
exponential functions of the form
1
(6.94)
D(a) == 1 +R(a)

where
R(a) =

k
I

: k.e

j=1 J

a+b-ab
l-a-b+ab

-p.a
J

tk.
j=2 J

(6.95)
(6.96)

s is a positive integer, kj are constants, and Pj are real numbers greater than zero. The
approximate function proposed in Eq. 6.94 is smooth and continuous and asymptotically
approaches exact values as a -+ 00 and at a = O. For a given set of s and Pj we may
fInd the constants kj for j = 2 to s by defIning
(a) = I-D(a)[1 +R(a)]

(6.97)

LOO[(a)]2da = minimum

(6.98)

and choosing the kj so that

Following the same procedure used in approximating lID(a), lIH(a) is approximated


by
1'
H(a) == 1 +R (a)

(6.99)

where

(6.100)

, I-JlG 6' ,
k=--:k.
I
2JlG j=2 J

(6.101)

Since the approximations of lID(a) and lIH(a) yield exact values at a=O, the integration of shear traction in the x-direction on a horizontal plane is equal to zero when
the plane is above the point of application of the force P, and equal to P when the

217

Elastostatics with Displacements as Unknowns

plane is below. The approximate solutions for each layer given by Eqs. 6.85 can be
expressed in the general fonn
1=1 +1"
(6.102)
where
(6.103a)
(6.103b)
where the indices I and n are either zero or positive integers, the coefficients e" and
e", are point functions and Ii is a positive point function. Thus the solution involves
inftnite integrals of a standard fonn
G(m,n,p)=

L-

a"e-paJ ... (pa)da,

(p

~O)

(6.104)

which can be evaluated in closed fonns. The integral I has a singularity at the point
of application of the concentrated force, corresponding to Mindlin's solution; the integral
r is non-singular. Thus the proposed solution is in a closed fonn. In the limiting case
of a homogeneous half space, J,lQ = 1, VI =V3 and kj =k; =0 for every j, and the integral
r simpliftes into the non-singular part of the Mindlin's solution.
Chan et al (1974) solved other plane strain and plane stress problems in which
the applied force is a line force unifonnly distributed on the line parallel to the y-axis
and through the point x =0, y =0, Z =e. The solution can be obtained from that for
a concentrated force P, by performing an appropriate integration with respect to y from
y =-00 to y =00.
6.6 LOAD TRANSFER PROBLEMS
Apirathvorakij and Karasudhi (1980), Niumpradit and Karasudhi (1981), Karasudhi, Rajapakse and Hwang (1984) and Karasudhi, Rajapakse and Liyanage (1984)
have solved several problems of long cylindrical elastic bars partially embedded in
elastic half spaces. The problem of torsion of a long bar embedded in a layered half
space, which has been solved by Karasudhi, Rajapakse and Hwang (1984), will be
presented here to illustrate a solution technique. Following the scheme proposed by
Muki and Sternberg (1970) for the treatment of an axial load transfer problem, the
system of the present study depicted in Fig. 6.2 is decomposed into an extended half
space and a ftctitious bar as shown in Figs. 6.3(a) and (b). Based on an appropriate
compatibility condition between the two latter systems the solution is derived from a
Fredholm integral equation of the second kind (Sneddon, 1966). The torque transfer
and the torque-twist relationship can be obtained numerically for various slenderness
ratios of the bar and ratios of shear moduli.

Foundations of Solid Mechanics

218

It should be recalled that the nontrivial displacement and stress components in a


torsion problem such as this are given by Eqs. 6.58 and 6.59. For reasons of convenience,
it is appropriate to non-dimensionalize all functions and variables with length dimensions
by a, which is the radius of the embedded bar.

20

-.--~----------~~----,-----~--~x

Fig. 6.2 Geometry of bar and embedding medium in torsion problem.


Before solving the main problem, one may wonder what is the factor of stress
singularity at the base perimeter of the bar. It has been found by Luco (1976), in the
problem of torsion of a rigid cylindrical bar partially embedded in a layered elastic half
space, as shown in Fig. 6.2, that the shear stresses at the base perimeter of the bar are
in the following fonns
2 i-O.5

O'Qz(r,h)-r(1-r)

r ()21k-0.5
J

O're(r,h)1 1 -

(6.105a)
(6.105b)

where k is a stress singularity factor in the range O:OS; k < 0.5 and h is the nondimensionalized length of the bar. For the current problem of an elastic bar, the factor k can
be taken as a function of Pl and ~ where

Pl =1,\ ~ =III
IJ.2

J.l3

(6.106)

in which III is the shear modulus of the upper layer, IJ.2 of the bar, and J.l3 of the
underlying half space. The method developed by Williams (1952) can be applied to
estimate k(Pl' PJ by considering an analogous antiplane system, as shown in Fig. 6.4,
consisting of two quarter planes with shear moduli III and IJ.2 and a half plane with a

219

Elastostatics with Displacements as Unknowns

t-~ (0)
r---~--~--~~r-----------r----X
II

Reglon
.0

II .I ,

:~
--~I"'I Uppert (z)
Layer;
I:

: L-I
II

ro----l'-

I
r'----~

r"'-- *

Lower ....
Layer;
fL3

fL

T (h)

(a) Extended half space B.

(b) Fictitious bar B .

Fig. 6.3 Composition of the torsion problem.

Fig. 6.4 Geometry of analogous system for estimating stress singularity factor.
shear modulus J.13. The equilibrium equation is given by Eq. 6.9 with zero body force,
i.e. VZw O. The solution for each constituent plane for small r can be put in the form

wj(r ,0) =r'Y[Aj sin(..,.a)+Bjcos(..,.a)]

(6.107)

'Y=k +0.5

(6.108)

where

Foundations of Solid Mechanics

220

i denotes a domain 1, 2 or 3 in the material system (Fig. 6.4), and Ai and Bi are arbitrary
constants. It should be noted that while displacements wi(r, 8) are bounded as r tends
to zero, stresses G8ri (r, 8) and Glri(r, 8) are singular. Continuity conditions appropriate
for the present antiplane case are
wl(r,O) = w2(r,0),
G8rI (r,O) = G8r2(r,0)
(6.109a,b)
(6.109c,d)
(6.10ge,f)
Substituting Eq. 6.107 into Eqs. 6.109 leads to a set of six homogeneous equations, i.e.
an eigenvalue problem in which k(PI, Pz) is the smallest positive real eigenvalue and
less than 0.5. In general, a trial-and-error process has to be adopted. For some values
of PI and Pz, the stress singularity does not even exist For the case of an infmitely
rigid bar (PI =0), it can be found that
k(O, Pz)

=7t- arcsin[f3z/(1 + Pz>],


1

=6'

(0 S; Pz < 1)

(6. 110a)

(Pz~ 1)

(6. 110b)

For a homogeneous half space (Pz 1) and an elastic bar, the stress singularity factor
is found to be the same as that of an infmitely rigid bar embedded in a layered half
space with Pz ~ 1, i.e.
(6.111)

..---i--..------ ...----------t--- r

}----__4i_

Linearly Varying
Shear Traction

Fig. 6.5 Unit torque in the interior of layered half space.

Elastostatics with Displacements as Unknowns

221

Now, it is necessary to consider the case of a shear traction equivalent to a unit


torque acting over a circular area of radius a in the interior of the half space at z = z'
as shown in Fig. 6.5. This auxiliary solution can be found by considering the half space
to be divided into three domains, namely, domain 1 (O:S; z :s; z), domain 2 (z' :s; z :s; h)
and domain 3 (h :s; z < 00) as shown. The solution for each domain is given by Eq. 6.58,
with two constants of integration All and CII The subscript n is used to denote a domain
number. The constant A3 must vanish in order to guarantee the boundedness of the
solution as z tends to infmity. Thus there are five constants; AI' Cl, A 2 , C2 and C3
The boundary and continuity conditions are:

a 9zl (r, 0) =0

(6.1l2a)
(6.112b)

u.u(r,h) = Um(r,h)

(6.1l2c)
(6.1l2d)

,2r

a. 91 (r ,z) - a. 92(r ,z)

or

=--,
1t

(r < 1)

(6.1l2e)

=0,

(r>l)

(6.1l2f)

,21-

a. 91 (r ,z) - a. 92(r,z) =--

1t 0

JI(~r)J2(~)d~,

(O:S; r < 00)

(6.1l2g)

Equations 6.1l2a to d and g can be used to determine the five constants of integration
AI' Cl, A 2, C2 and C 3 However, to incorporate the notion of the stress singularity of

the stress field in the form of Eqs. 6.105 in the main problem, the applied shear traction
in the fundamental solution for z' = h should be, in place of the linear Eqs. 6.1l2e and
f, in the form
2
2.\:-0.5
a 9zl (r ,h) - a 9z3(r ,h) = --(lI(k)r(l- r ) , ( r < 1)
(6.1l3a)
1t

=0,

(r> 1)

(6.1l3b)

or

where

1
(ll(k) ='2(k +O.5)(k + 1.5)

(6.1l4a)

~(k) = 2k+O.5(lI(k)r(k + 0.5)

(6.1l4b)

Foundations of Solid Mechanics

222

in which r is a gamma function. In addition, other boundary conditions for the fundamental solution in case z' h are

O'a.l(r ,0) =0
Uel(r,h)

=Um(r,h)

(6. 115a)
(6.115b)

Equations 6.113c and 6.115a and b can be used to determine the three constants of
integration AI' C1 and C3
The problem of axially symmetric torsion of an elastic cylindrical bar partially
embedded in a layered elastic half space is depicted in Fig. 6.2. In addition to those
previously defined, the following symbols are introduced; To the torque applied at the
top end of the bar which is flush with the surface of the half space; and x, y and z
Cartesian components of position vector x; thus X3 is identical to z. The top layer (r > 1,
0< z < h) and the underlying half space (z > h) are perfectly bonded on the contact
surface (r> 1, z h). The cylinder is bonded to the half space in the area (r < 1, z h)
and along the surface of the cylinder (r 1, 0 S; z S; h). Following the approximate
scheme similarly used by Muki and Sternberg (1970) in the elastostatic axial load
transfer, the system in Fig. 6.2 is decomposed into two systems; an extended half space
B as shown in Fig. 6.3(a) and a fictitious bar B. as shown in Fig. 6.3(b) with a shear
moduli J.1. equal to the difference between the shear moduli of the real bar J.12 and the
half space J.Ll' i.e.
(6.116)

The extended half space is subjected to a distributed torque t(z) which is exerted
by B. on B at X3 =z in a region D in place of the bar. In addition, B is also subjected
to end torques To - T.(O) and T.(h) applied at the terminal cross sections as shown in
Fig. 6.3(a). The bond torque t(z) and the torque To - T.(O) are assumed to be distributed
linearly in the form of Eqs. 6.112e over their respective cross sections A,(O < z < h) and
A o, while the torque T.(h) in the singular form of Eq. 6.113a over the cross section All'
Conversely, the bond torque and end torques are exerted by the extended half space B
on the fictitious bar B., which may be treated as a one-dimensional elastic continuum,
for which the equation of equilibrium is
dT.(z)

t(z)=-~

(6.117)

The torque-twist relation of the fictitious bar is


1',

(z)

=1tJ.1. ~.(z )

(6.118)
dz
where T.(z) and cIl.(z) are fictitious torque and angle of twist of the bar respectively.
The governing integral equation can be derived using the following compatibility
condition between B. and the edge of the region D:

Elastostaties with Displacements as Unknowns

dt1>.(z)

~=

223

dua(1-,z)

dz

(6.119)

(OS;z S;h)

Here u.a(r, z) is the displacement in the 9 -direction at a point x(r, 9, z) in domain B, and
1- denotes a value infmitesimally less than unity. Muki and Sternberg (1970) originally
demanded compatibility between B. and the corresponding average over a cross section
of region D. Karasudhi, Rajapakse and Hwang (1984) found that this was not sufficiently accurate. The function u.a(r, z) can be expressed in the form

So" ~(r, z;z )t(z)dz'

u.a(r, z) = [To - T.(O)] uf1]'(r, z ;0) + T.(h )uf1]'(r, z;h) +

(6.120)

where ~(r,z;z) is the auxiliary solution, i.e. the displacement in the 9-direction at a
point x(r, 9, z) due to a unit torque applied at a depth z' (Fig. 6.5). Substitution of Eqs.
6.117, 6.118 and 6.120 in Eq. 6.119 results in,
2T.(z)

- - = [To-T.(O)]t1>T.(Z,O) + T.(h)t1>T.(z,h)-

i"

0
,dT(z) ,
t1>T.(Z,Z ) - - , dz

dz

(6.121)

where t1>~.(z,z) is an influence function defined as


o
,d~(1-,z;z)
t1>T.(Z,Z) =
dZ

(6.122)

It should be noted that this influence function is smooth and continuous everywhere
except at z = z', where the magnitude of the discontinuity according to Eq. 6.112e is
equal to 211tlll for 0 < z < h. Incorporation of the stress singularity at the end of the
bar, t1>~.(z,z) when z = z' = h requires appropriate handling in the numerical solution
scheme.
Integrating the integrals in Eq. 6.121 by parts, while taking proper account of the
discontinuity, we find
2T.(Z)(1
1)
- -+-

i"

,dt1>~.(Z,z),

(6.123)
T.(z)
:.'
dz =Tot1>T.(Z,O)
1t
j.1. III
0
uZ
Equation 6.123 is a Fredholm integral equation of the second kind governing the distribution of T.(z) along the fictitious bar.
The real bar torque T(z) can be obtained by combining the fictitious torque T.(z)
with the corresponding area integral of the shear stress cre. in the region D of B, i.e.
T(z)=T.(z)+f A, rcre.(r,z)dA,

(OS;z S;h)

(6.124)

The angle of twist of the real bar taken as equal to that of the fictitious bar, thus
t1>.(z)= ue(1,z), (OS;zS;h)
(6.125)
Using Eq. 6.120 and performing appropriate integrations in Eqs. 6.124 and 6.125, we
obtain
T(z)

o
=To'te.(z,

0) +

i'"
o

T.(z)

'

<fte.(z,z)
dZ

dz

(6.126)

Foundations of Solid Mechanics

224

~.(z)

o
=TO~T(Z
, 0) +

i"'~T(Z,Z)
o

T.(z)

'

dZ'

dz

(6.127)

where 't~(z,z') and ~~(z,z') are influence functions dermed as

'r

'

o Z ) =JA. r(JQrT(r, z;z )dA


'tQr(z,

(6.128)

~~(z,z) =Uer(1,z;z)

(6.129)

Here (JQrT(r, z;z ') is a shear stress in the auxiliary solution depicted in Fig. 6.5.
The Fredholm integral equation for T.(z) has an integrable logarithmic singularity
at z z', and is solved numerically. We use a scheme which is similar to that employed
by Hopkins and Hamming (1957) and Lee and Rogers (1963) in the solution of Volterra
integral equations. Divide the interval of integration into n equal partitions and denote
the mesh points by Zj; j 1,2, ... , n + 1, with Zl 0 and Zn+l h. Equation 6.123
becomes

2T.(zj) ( 1

1)

- - -+- - ,1: -2 [T.(z) + T.(Zj+l)] [~r.(Zj,Zj+1)-~r.(Zj,z)] =TO~T.(Zi'O)


1t

J.1.

III

II

J =1

(6.130)

which is a set of n + 1 simultaneous equations, the solution of which gives the fictitious
torque T.(z) at all mesh points.
In order to find the stress singularity at the base perimeter of the bar, we note that
~~(h,h) contains a singular term of the form

So" ~o.s-"Jl(g)Jk+l,S(~)d~-r(1-r'J.)"-o.s,

(r

=n

(6.131)

Thus there exists in Eq. 6.130 an indeterminate term


T.(h )~~(h, h).

As Z tends to h, we speculate that T.(h) approaches zero as


T.(z)

=[ 1-(~JTS+",

(~-+ 1-)

(6.132)

In the numerical solution to Eq. 6.130, the numerical value of 1- in Eq. 6.131 is
increased towards unity gradually until T.(h) is sufficiently close to zero and the results
become stable. Moreover, the physical interpretation that the shear stress along the
mantle of the cylindrical bar is proportional to the bond torque, i.e.
(Jre(l,z)-t(z),
together with Eqs. 6.117 and 6.132 can help us to confirm the singular form of (Jrs(l,z)
as Eq. 6.105b.
Numerical results by the present approach were compared with those by Luco
(1976), who had been more rigorous but restrictive to the case of rigid bars (PI 0).
The agreement between the two approaches was good and became better as the bar
length increased. The significance of the singular stress field condition at the base

225

Elastostatics with Displacements as Unknowns

perimeter of the bar was found to be small, especially when the main concern was on
the relationship between the applied torque To and the angle of twist <1>.(0). This type
of situation is more pronounced in the problem of a flexible bar.
6.7 INFINITE ELEMENTS FOR MULTILAYERED HALF SPACES
Analytical methods such as that described in the preceding section become
impractical if not impossible when extended to problems of multilayered half spaces.
In order to bypass such difficulties, some research workers employ finite element
methods, i.e. truncating the semi-infmite domain into a large but finite field which is
then discretized into finite elements. However, such methods are computationally
inefficient As an example, Muki and Dong (1980) found that a finite element field of
at least 750 degrees of freedom is necessary to obtain accurate results for a layered half
space subjected to a uniform normal pressure on the surface of the half space.
The present section follows Rajapakse and Karasudhi (1985) who have shown that
the computational efficiency can be achieved by properly using infinite elements in
conjunction with finite elements. The whole material domain is separated into a near
field and a far field as shown in Fig. 6.6. The near field, consisting of the partially
embedded bar and a finite region of the half space around it, is discretized into conventional finite elements. The far field covering the rest of the half space is discretized

Embedded

Embedding
Multilayered
Half - Space

Bar; fLb' 'Ub

Fig. 6.6 Discretization of multilayered half space.

Foundations of Solid Mechanics

226

into infmite elements. Near the surface of the half space are horizontal infmite elements
(HE), while the rest of the far field are occupied by radiating infmite elements (RE).
Every infmite element has point nodes only on the interface between the near field and
the far field. An obvious major advantage of this approach is that nonlinearity effects
can be easily incorporated in the near field finite elements while all infmite elements
always remain linearly elastic. We now describe the development of each type of
infmite element.

Fig. 6.7 Layered half space.


Rajapakse and Karasudhi (1985) based their development of infmite elements on
the analytical solution of a layered half space, shown in Fig. 6.7, subjected to four types
of loading on its surface. The loading types are shown in Fig. 6.8. The orders of
magnitude of the displacement components at a large distance from the loads are shown
in Table 6.1; in terms of r for the top layer, and in terms of R (spherical radial
coordinate) for the underlying half space. Experience with the tedious algebra involved
with layered half spaces tells us that an analytical derivation of far field displacement
functions for a multilayered half space like Fig. 6.6 is rather impractical. However, we
should expect that the far field displacement functions for horizontal infmite elements
should be qualitatively the same as those for the layer in the table, and those for radiating
infmite elements the same as those for the underlying half space in the table. It is
important to note that the origin of the coordinate R should be at the intersection
between the axis of symmetry and the last layer-half space interface.

227

Elastostatics with Displacements as Unknowns

Table 6.1
Types of

Top layer

Displacements

loading

or half space

ur

Us

Torsion
(Fig.6.8a)
Vertical
(Fig. 6.8b)
Horizontal
(Fig.6.8c)
Moment
(Fig.6.8d)

Layer
Half space
Layer
Half space
Layer
Half space
Layer
Half space

r-2 +0(r-3)

Uz

R- 2 +O(R-3)

R- 1 +0(R-2)

cos 9[r-I + O(r -2)]

sin 9[r-I + O(r -2)]

cos 9[r-I + O(r -2)]

cos
+
cos 9[r-2 + O(r -3)]

sin
+
sin9[r-2 +O(r-3)]
sin 9[R-2 + 0(R- 3)]

cos 9[R- 1 + 0(R- 2)]


cos 9[r-2 + O(r -3)]

r-I +0(r-2)

9[R- 1

0(R- 2)]

cos 9[R-2 + 0(R- 3)]

9[R- 1

r-I +0(r-2)
R- 1 +0(R-2)
0(R- 2)]

cos 9[R-2 + 0(R- 3)]

Note: 0 ( ) =order of truncation error.

(a) Torsional loading.

(c) Horizontal loading.

_l

4/71"

1/71"
I
I

Y'
(b) Vertical loading.
Fig. 6.8 Types of loading.

+).--.---

(d) Moment loading.

Foundations of Solid Mechanics

228

Any infInite element in the rz-plane can be contracted into a fInite element in a
new ~11 plane by employing a transfonnation which is singular at ~ =+1. The coordinate
mapping for a horizontal infInite element is
r

=~2(1- ~rlLj(11)rj

(6. 133a)

z =LL(11)Z.
j
J
J

(6. 133b)

and for a radiating infInite element

=~2(1- ~rlLj(11)rj

(6. 134a)

z =~2(1- ~rlLj(11)Zj

(6. 134b)

where Lj (11) is the Lagrange polynomial for node j. In short, a fInite element is obtained
by a singular contraction of infmite elements as depicted in Fig. 6.9.

-1

-1

0
+1
+1

Fig. 6.9 Finite elements by singular contraction.


The relationship between displacement components inside an infmite element and
their nodal values has the usual fonn

Ur(~' 11) =~N; (~, 11)u:

(6. 135a)

(6. 135b)

229

Elastostatics with Displacements as Unknowns

u.(~, T\) =~Nj(~, T\)u:

(6. 135c)

in which a superscript or a subscript j denotes the node j, and NJ, Nj9 and Nj are
displacement interpolation functions. Being consistent with the analytical far field
displacements, these displacement interpolation functions can be expanded, for horizontal loading as,

N;(~, T\) =Nj9(~, T\) =Nj(~, T\) =4(l-~)Lj(T\)

(6.136)

and for moment loading as,

N;(~, T\) =Nj9(~, T\) =Nj(~, T\) =~(l-~)2Lj(T\)

(6.137)

For vertical loading, N:(~, T\) does not exist and NJ(~, T\) and Nj(~, T\) take the form
given by Eq. 6.136. For torsional loading, only Nj9(~, T\) exists and is given by Eq.
6.137. All elements developed are compatible in the sense that there is continuity
between the relevant ftrst derivatives of Nj within the elements. The stiffnesses of
these elements involve double integrals (with respect to ~ and T\), which can be
accurately evaluated by standard Gauss quadrature formula. Another major advantage
of this algorithm is that we do not need any special integration scheme and can use any
standard finite element subroutines, since eventually we are dealing with finite elements
(not infinite elements).
6.8 SATURATED LARGE SPACES
The Navier equations and Darcy's law for quasi-statics of porous isotropic elastic
solids completely saturated with fluids have been derived in Section 3.6.7 for Cartesian
system; see Eqs 3.123 and 3.124. For cylindrical coordinates, the Navier equations are
given by Eqs. 3.67 with the body force components there being replaced by the excess
pore pressure gradients. Replace

OPI

Xr by or

(6. 138a)

X b !OPI
9 Y r oa

(6. 138b)

oPI
oz

X. by -

(6. 138c)

Darcy's law, Eq. 3.124 remains invariant. To uncouple these four field equations.
Schiffman and Fungaroli (1965) introduced three displacement potentials, so that
(J2c1l 20'
Ox
ur =-oroz +; 09 +z or
(6. 139a)

230

Foundations of Solid Mechanics

+:.ax

u =_ (f<l> _2 d\}l
9
rdedZ
dr r de

ax

(f<l>

u =--+z--x

dZ 2
dZ
PI= ('A.+2)J.);z

(V2<1-2)J.~

(6. 139b)

(6. 139c)

(6. 139d)

It can be verified that the four field equations can be transfonned into the following
three uncoupled equations

C~<I>=V\~)

(6. 140a)

V2\}1=0

(6. 140b)

V2X=0

(6. 140c)

If the material domain is infinitely extended radially (0;5; r < 00), the solution
scheme for solids described in Section 6.3 can be adapted for the present purpose to
solve Eqs. 6.140. However, it is now necessary to use Laplace transfonns with respect
to the time t. The solution scheme starts by expanding <1>, \}I and X in Fourier series,
thus

<I>(r,e,z,t) = L <l>m(r,z,t)cosme
m=O

(6.141a)

'(r,e,z,t) = L \}Im(r,z,t)sinme
m=O

(6.141b)

x(r,e,z,t)= L Xm(r,z,t)cosme
m=O

(6.141c)

Eventually, the solution to Eqs. 6.140 can be written in the fonn

<i>m(~'z,p) =Ame-9 + Bm e9z +Cme4'Z + Dme"fl

(6. 142a)
(6. 142b)
(6. 142c)

where a super tilde (-) denotes a Hankel-Laplace transfonn, ~ is the Hankel transfonn
parameter versus r, p the Laplace transfonn parameter versus t, Am to Hm are functions
of ~ and p, and
(6.143)
Functions Am to Hm can be detennined from well-posed boundary conditions similar

Elastostatics with Displacements as Unknowns

231

to Eqs. 3.127 already derived for the Cartesian coordinate system. The boundary
conditions are prescribed on each boundary where z is constant and involve the prescription of either
(6. 144a)
U, or 0z,
(6. 144b)
U, or 0 ..

and p! or

apt
kaz

(6.144c)
(6.144d)

Apirathvorakij and Karasudhi (1980), and Niumpradit and Karasudhi (1981) applied this
theory to load transfer problems in a homogeneous soil domain, and found that the
inverse Laplace transform formula given by Schapery, Eq. 1.164, is accurate enough.
Later, Karasudhi, Karunasena and Puswewala (1987) proposed an infinite element
algorithm similar to that described in Section 6.7 for ideal solids. Most recently,
Karasudhi and Prechaverakul (1987) presented an analytical prediction of land subsidence due to water pumping from deep wells. A development of an infmite element
algorithm for such subsidence in multilayered soil media was carried out by Karasudhi
and Alvappillai (1988).
6.9 EXERCISE PROBLEMS
6.9.1

Use Eq. 6.7 to solve Problems 4.7.1 to 4.7.3.

6.9.2

A long mine tunnel of radius a is cut in deep rock. Before the tunnel is cut
the rock is subjected to a uniform pressure p. Considering the rock to be an
infmite, homogeneous, isotropic elastic medium, determine the inward radial
displacement at the tunnel surface due to the excavation.

6.9.3

Use Eq. 6.8 to solve Problem 4.7.4.

6.9.4

An infmite composite plate made of two elastic materials of distinct properties


is supported by a smooth rigid horizontal base. The thickness of the top and
bottom elastic layers are hI and ~ respectively. The top surface of the plate
is loaded by a uniform normal pressure q and the top elastic layer is subjected
to a uniform body force
in the downward vertical direction. Find displacements and stresses.
Hint: With reference to the axis z pointing downwards, only Uz is nonzero
in this problem. The only Navier equation for the top layer can be shown to
be

to

232

6.9.5

Foundations of Solid Mechanics

Using the fonnulation of Section 6.3, fmd Mindlin's solution which is the
solution to the problem of a force at a point in the interior of a homogeneous
isotropic elastic half space.
Answer: With reference to the coordinate system as shown in Fig. 6.10, the
pertinent displacement components are; for the case of the force P in the
z -direction,

r---------.--.---------r--~x

z
Fig. 6.10 Mindlin's solution.
Pr

[z -c (3-4v)(z -c) 4(1-v)(1-2v) 6cz(z +C)]


R (R
R23
2 2+ Z +c ) + 5 ,R2

[3-4V 8(1-vi-(3-4v) (z-ci


--+
+-Rl
R2
Rl

u, =161t~(l-v) --3-+
Rl
u

161tJl(1-v)

(3 - 4v)(z + c i - 2cz 6cz(z + C)2]


3
+ R25
R2
and for the force P in the x -direction,

u=

2cZ(1 -3X-

[3-4V 1 x 2 (3-4v)x2
----+-+-+
+161t~(l-v)
Rl
R2 Rl
Ri
Ri
P

'

2)

Ri

4(1-V)(1-2V)(1
x2
]
+ R 2+z +c
R2(R2+z +c) ,

233

Elastostatics with Displacements as Unknowns

Pxy
[1 3-4v 6cz 4(1-V)(1-2V)]
v=161tJl(1-V) Rl+ Ri - Ri, - Rz(Rz+z+c)z ,
w

where

6.9.6

6.9.7

Px

=161tJl(1-v)

[z -c (3-4v)(z -c) 6cz(z +c) 4(1-V)(1-2V)]


-3-+
+ Rz(R z+z+c ) '
Rl
Rz3
R z5

A homogeneous isotropic elastic half plane -00 <x < 00, 0 ~ y < 00, is subjected
to the following tractions on its surface y =0 in the region -a < x < a: (a) a
uniformly distributed normal force, (b) a uniformly distributed shear force, and
(c) a normal force distributed as a linear homogeneous function of x. Find the
stresses.
With the same geometry as in the preceding problem, the displacement in the

x -direction is prescribed uniformly in the region -a < x < a, while the normal
traction is assumed to vanish everywhere on the half plane surface. Find the
distribution of the contract shear stress in terms of the applied shear force Po.
Answer:
eJ..,(X, 0) =

1t(a -x)

=0,
6.9.8

(Ixl <a)
(Ixl >a).

With the same geometry as in the preceding problem, a uniform rocking (rotation about the z axis) is prescribed in the region -a <x < a, while shear
traction is assumed to vanish everywhere on the half plane surface. Find the
distribution of the contact norinal stress in terms of the prescribed angle of
rocking a.
Answer:

=0,
6.9.9

Po
z z 112'

(ixi >a).

A homogeneous isotropic elastic half space 0 ~ r < 00, 0 ~ z < 00, is subjected
to the following tractions on its surface where z =0 in the region 0 ~ r < a: (a)
a uniformly distributed normal force, (b) a uniformly distributed shear force in
the x-direction, and (c) a normal force distributed as a linear homogeneous
function of x. Find the displacements.

Foundations 0/ Solid Mechanics

234

6.9.10

With the same geometry as in the preceding problem, the displacement in the
x-direction is prescribed unifonnly in the region 0 S; r < a, while the nonna!
traction and the shear traction in the y-direction are assumed to vanish
everywhere on the half space surface. Moreover, the shear traction in the
x -direction vanishes outside the loaded area. For zero Poisson's ratio, fmd the
distribution of the contact shear stress in tenns of the applied shear force Po,
and the relationship between Po and the prescribed displacement Uo.
Answer:
Po
<Ju(r, 8, 0)

2 112'

27ta(r -a)

=0,

6.9.11

(OS;r<a)

(r> a),

With the same geometry as in the preceding problem, a unifonn angle of twist
(rotation about the z axis) is prescribed in the region 0 S; r < a. Find the
distribution of the contract shear stress in tenns of the prescribed angle of twist
ex, and the relationship between the applied torque T and a.
Answer: This is a pure torsion problem.
<J.e(r ,0)

-4~
=1t(2
2)112'
a -r

=0,

(0 S; r < a)
(r>a),

CHAPTER VII
LINEAR VISCOELASTICITY

7.1 LINEAR ELASTICITY AND NEWTONIAN VISCOSITY


Measurements of load-displacement relationships show that the theory of linear
elasticity holds for most metals under small strains. However, the simple fact that the
vibrations of metal systems decay even in vacuo indicates that there is some energy
dissipation. Hooke's law is insufficient, and we need to consider others. This need
has become more evident since the development of polymeric materials. These materials
possess a capacity to store some and dissipate the rest of the energy input. While the
theory of elasticity accounts for materials which are capable of storing energy without
any dissipation, the theory of Newtonian viscous fluids governs those which dissipate
the whole energy input. Deformed elastic bodies return to a natural or undeformed
state upon removal of applied loads. Viscous fluids, however, possess no capacity at
all for deformational recovery. In an elastic solid, stress is related directly to deformation; in a viscous fluid, stress depends (except for its hydrostatic component) upon
the rate of deformation. Hookean elastic solids and Newtonian viscous fluids represent
the ends of a spectrum. Between these extremes there are materials which incorporate
a blend of both linear elastic and (Newtonian) viscous characteristics. Appropriately,
such materials are named linearly viscoelastic.
Linear elastic solids retain linearity between load and deformation, and the linear
relationship is independent of time. Any current state of deformation of a linear elastic
solid can be determined completely without the knowledge of the entire loading history.
In other words, a linear elastic solid remembers only the undeformed state of its body.
Other solids do not behave in such a simple way. Their current state of deformation
cannot be determined completely unless the entire loading history is known. Linearly
viscoelastic materials remember the past, but their memory belongs to a relatively simple
class. For such materials, linearity between load and deformation is maintained, but
the linear relationship depends on time t. The principle of superposition still holds.
Thus if El(t) and ~(t) are deformations due to loads al(t) and aZ<t) respectively, then
the defo~ation due to the load Alal(t) + Azaz(t) is
E(t)

=AlEl(t) +Az~(t)

(7.1)

Foundations of Solid Mechanics

236

(b) Viscous dashpot.

(a) Linear spring.

Fig. 7.1 Basic elements in mechanical models.


where Al and A z are constants.
Constitutive laws for linear viscoelasticity could be introduced in a simple way
through the use of one-dimensional mechanical models made of massless linear springs
and viscous dashpots. The springs characterize linear elasticity, while thedashpots
incorporate Newtonian viscosity. See Fig. 7.1. In a spring with a spring constant G,
the stress a is related to its strain by
a=G
(7.2)
The analogous equation for the dashpot is
(7.3)
a=l1
where 11 is a viscosity constant and =dldt. By direct integration, we can rewrite Eq.
7.3 in the form

1'

1
(t) =-

11

(7.4)

a(s)ds

0-

in which the integral lower limit 0- denotes quiescent initial conditions, Eq. 1.123.

(a) Maxwell.

(b) Kelvin.

Fig. 7.2 Two-parameter models.

Linear Viscoelasticity

237

Fig. 7.3 Standard linear solid model.


The Maxwell model of viscoelasticity is the combination of a spring and dashpot
in series as shown in Fig. 7.2(a). The Kelvin or Voigt model is the parallel assembly
shown in Fig. 7.2(b). For many materials, even these two-parameter models have been
found to be inadequate. A three-parameter model constructed from one spring and one
Maxwell unit in parallel and known as the standard linear solid model is shown in Fig.
7.3. The generalized Kelvin model consists of one Maxwell unit and a sequence of
Kelvin units in series as depicted by Fig. 7.4. On the other hand, an assembly of one
Kelvin unit and a sequence of Maxwell units in parallel as shown in Fig. 7.5 is called
a generalized Maxwell model.

Fig. 7.4 Generalized Kelvin model.

Q-----<r"-----Q----- --------

6-----6------6-----Q----------

Fig. 7.5 Generalized Maxwell model.


7.2 CREEP AND RELAXATION
If the stress cr is kept constant in any mechanical models presented in the previous
section, the strain e tends to increase with time, i.e. there is creep. If the strain e is

Foundations of Solid Mechanics

238

kept constant, the stress cr tends to decrease with time, i.e. there is relaxation. The
creep compliance J(t) is defmed to be the strain e(t) when the stress cr(t) is equal to
the Heaviside step function H(t). The relaxation modulus J.1(t) is defined to be the
stress cr(t) when the strain e(t) is equal to the Heaviside step function H(t). For a given
material, J(t) and J.1(t) are related to each other.

<r

Aa;,..----

L~(S) +

,I

O~--~--~~------~

t2 13

ds

I
,,

,I

0"01-----1---+-1---- - -

t,

ds

..,----------

,-------i-----I I

AOj

ao-(s)

s+ds

(b) Arbitrary stress history.

(a) Stepped stress history.

Fig. 7.6 Derivation of convolution integral.


Because the superposition principle holds, the strain due to the stepped stress in
Fig. 7.6(a) is
(7.5)
The arbitrary stress history cr(t) of Fig. 7.6(b) may be analyzed as an infmity of step
loadings, each of magnitude dcr, so that the strain e(t) is

e(t)=cr(O)1(t)+ ('J(t-s)dcr(s)
o+

(7.6)

where dcr(s) [dcr(s )Ids] ds, and cr(O+) is the value of cr(t) at t slightly greater than zero,
i.e. the initial value of cr(t). Note that all functions involved are zero for t < 0 (Eq.
1.123). Due to the relationship between the Heaviside step function H(t) and the
Dirac-delta function aCt), Eq. 1.106, Eq. 7.6 can be rewritten as a single integral with
the lower limit infmitesimally less than zero, i.e. 0-, so that
e(t)

=l>(t-s)dcr(s)

in which cr(s) is considered as


cr(s)

(7.7)

=crc(s)H(s),

where crc(s) is a continuous function of time. In other words crc(t) is not only equal to
cr(t) for t > 0 but also continuous at t O. Figure 7.7 should help to clarify this concept

239

Linear Viscoelasticity

further. The same equation (Eq. 7.7) also holds for other discontinuities of O'(t).
Suppose there is a discontinuity AO'o at time to. The stress function in Eq. 7.7 should
be considered as
(7.8)
O'(s) =O'c(s)H (s) + AO'J/ (s - to>

--I7(t)
----a:(t)

_ _ _1...-_ _ _ _- - ' -_ _ _ _ _

Fig. 7.7 Continuous O'c(t) in comparison with discontinuous O'(t).


where O'c(s) is as shown in Fig. 7.7, i.e. it is continuous everywhere and parallel to the
O'(s) curve (dO'clds =dO'lds for 0 < s S; to and to S; s S; t). Substituting Eq. 7.8 into Eq.
7.7 we fmd
(t) = O'(O)l(t)+AO'oI(t -to)+ (' J(t -s)dO'c(s)
Jo+

(7.9)

which can be shown to be identical to that obtained directly by the superposition


principle. Readers with a mathematical background will recognize Eq. 7.7 as a Stieltjes
integral. Moreover, it can be shown that Eq. 7.7 still holds even when the stress function
is singular but integrable like the function (5(t - to). In other words, the superposition
principle and Eq. 7.7 should give the same result,
(t)

d.J(t-to)

d(

[for O'(t) =(5(t - to)]

),

t-to

(7.10)

The integral in Eq. 7.7 is often referred to as a convolution integral or hereditary


integral since the 'effect' at any time is seen to depend upon the entire 'cause' history.
The commutative law for convolution integrals, Eq. 1.l38a, permits the interchange of
J and 0' in this integral, i.e.
(7.11)
e(t) = J*dO' = O'*d.J
in which the following notation is introduced

v 1*dv2 =

v 1(t -s)dv2(s)

(7.12)

Foundations of Solid Mechanics

240

Equation 7.11 is also known as the creep law. Its Laplace transfonn is
e(p)=pi(p)a(p)
(7.13)
When 8(t) =H(t), o(t) is (by definition) equal to the relaxation modulus J.1(t). Equation
7.13 gives the relationship between the creep compliance and the relaxation modulus
in the fonn

p 2i(p)fj,(p) 1
or, due to Eqs. 1.124a, 1.132 and 1.135,
J*dJ.1= J.1*dJ =H(t)
or, due to Eqs. 1.124i and 1.135,

J*J.1= J.1*J =tH(t)

(7. 14a)
(7.14b)
(7.14c)

Following similar arguments we can express the relaxation law in the fonn
o(t) =J.1*dE =8*dJ.1
(7. 15a,b)
the Laplace transfonn of which is
(7. 15c)
a(p) =p fj,(p )e(p )
This yields the same relationship between relaxation modulus and creep compliance
given by Eqs. 7.14.
7.3 COMPLIANCE AND MODULUS OF MECHANICAL MODELS
For the linear spring model shown in Fig. 7.1(a), the stress-strain relationship is
given by Eq. 7.2, the Laplace transfonn of which is
a(p) =Ge(p)
(7.16a)
Comparing this equation with Eqs. 7.13 and 7.15c we fmd
-1

i(p)=L

(7.16b)
(7.16c)

Thus the creep compliance and the relaxation modulus for this type of element are
J(t) H(t)lG
(7.16d)

=
J.1(t) =GH(t)

(7.16e)
For the viscous dashpot model shown in Fig. 7.1(b), the stress-strain rate equation
is Eq. 7.3 and we fmd
(7. 17a)
a(p) 1'Ip(P)

-2

P
J(p)=1'1

fj,(P)

=1'1

(7.17b)
(7. 17c)

241

Linear Viscoelasticity

J(t) =-H(t)

11

(7. 17d)

=11l>(t)

(7. 17e)
For the Maxwell model shown in Fig. 7.2(a), the Laplace transform of the
stress-strain relationship in the spring takes the same form as Eq. 7.16a, i.e.
crG(p)= Gea(P)
(7.18a)
Jl(t)

while that in the dashpot takes the same form as Eq. 7.17a, i.e.
crll (p) =11P~(P)

(7.18b)

The stress O'(t) must be the same in the spring as in the dashpot, i.e.
cr(p) = crG(p) = crll (p)

(7. 18c)

The strain e(t) must be the sum of those in the spring and the dashpot, i.e.
e(p)=ea(P)+~(P)

(7. 18d)

Substituting Eqs. 7.18a and b into Eq. 7.18d, and using Eq. 7.18c, we fmd the Laplace
transform of the stress-strain relationship for a Maxwell model to be

e(p)

= G1 [1 + (P'tf1,J O'(p )

(7. 18e)

where
't =11/G

(7.180

Comparing Eq. 7.18e with Eqs. 7.13 and 7.1Sc, we find


j(p) =![p-l + p-2't-1

(7.18g)

=G(p +'t-1t

(7. 18h)
Thus the creep compliance and the relaxation modulus of a Maxwell model are
ii(P)

J(t)

=~(1+t't-l)H(t)

(7.18i)

=Ge-tl"'H(t)

(7.18j)
For the Kelvin or Voigt model shown in Fig. 7.2(b), the Laplace transforms of the
stress-strain relationship in the spring and the dashpot are the same as Eqs. 7.18a and
b, respectively. The strain e(t) must be the same in the spring as in the dashpot, i.e.
e(p) = eG(P) = ~(P)
(7.19a)
Jl(t)

The stress O'(t) must be the sum of those in the spring and the dashpot, i.e.
cr(p) = crG(p) + crll (p)

(7.19b)

Substituting Eqs. 7.18a and b into Eq. 7.19b, and using Eq. 7.19a, we find the Laplace
transform of the stress-strain relationship for a Kelvin model to be
cr(p) = G(1 + p't)e(p)
(7. 19c)

242

Foundations of Solid Mechanics

Proceeding as before, we find


J(p) =1.fp-l_ (p
G

+ "C-1fl

(7.19d)

=G(p-l+"C)

(7.1ge)

J(t) = ~(1-e-t/~)H(t)

(7.190

Jl(t) = G[H(t) + "CS(t)]

(7. 199)

Ji(p)

For the standard linear solid model shown in Fig. 7.3, we fmd
(7.20a)
(7.20b)
e(p) =f.o(P)

=e1(P)

(7.2Oc)

cr(p) = cro(P) + cr1(p)

(7.2Od)

Substituting Eqs. 7.20a and b into Eq. 7.2Od, and using Eq. 7.20c, we fmd
(7.20e)
Thus we have
Go+(GO+G1)"CIP
Jl(P) =
p(1 + "CIP)

(7.200

(7.20g)

J(p)= p 2Ji(p)

(7.20h)
(7.20i)
For a generalized Kelvin model, the governing equations are
1
1,Eo(P) = Go [1 + (P"Cof J (Jo(P)

(7.21a)

cri(p) = Gi(1 + P"C)Ei(P),

(i

=1 to N)

(7.21b)

cr(p) =cro(P) =cri(p),

(i = 1 to N)

(7.21c)

Linear Viscoelasticity

243

(7.21d)

)+ f
J(P)=[~(I+_1)+ f
E(P)=[~(l+_l
Go
Go

(1+ P 'CJ-1 ]a(P)


G;

'CoP

;=1

'CoP

(1+P'C J-1
;=1
G;

]p_

-1
Jl(P) =p 2J(p)
J(t)

(7.21e)

(7.21f)
(7.21g)

=[~+..!..+ f (l-e-t/~i)]H(t)

(7.21h)
Go 110 ;=1
G;
To obtain an analytical expression for Jl(t), j!(p) must ftrst be written as the quotient
of two polynomial functions of p. Then the inverse Laplace transform can be obtained
by means of the Heaviside expansion formula (Spiegel, 1965). Otherwise, one may
solve Eq. 7.14b or c numerically, or adopt one of the approximate methods for inverse
Laplace transforms described in Section 1.12.
For a generalized Maxwell model, the governing equations are
ao(P) =Go(1 + P'Co)'Eo(P)
(7.22a)
(i = 1 to N)

(7.22b)

(i = 1 to N)

(7.22c)
(7.22d)
(7.22e)
(7.22f)
(7.22g)
(7.22h)

In this case J (t) is not simple, but may still be obtained using the methods described
above.

Foundations of Solid Mechanics

244

Note that the presence of a dashpot in parallel with other units in a mechanical
model introduces a singular (but integrable) function, i.e. a(t) into the relaxation
modulus. For this reason, many researchers have used a generalized Maxwell model
without such a 'free' dashpot. The presence of a dashpot in series with other units in
a mechanical model introduces a term which tends to infinity with t into the creep
compliance. Note also that the attachment of H(t) to every nonsingular term makes
such expression valid even for t < 0, and underlines the axiom of nonretroactivity, i.e.
the effect at the present time is due to causes in the past and not in the future.
If the generalized Kelvin model is considered as typical for the behavior in creep,
then a general form of the creep compliance is

J(t)=[J, +

~o +'I'(t)JH(t)

(7.23a)

in which J, is known as the glassy creep compliance, and 'I'(t) the creep junction. J,
is a constant equal to the initial value of J(t), i.e.

J, =J(O~

(7.23b)

and 'I'(t) is a monotonically increasing function with a zero initial value and a constant
asymptote as t -+ 00, as shown in Fig. 7.8. The stress-strain relationship for this case
is given more simply by the creep law (Eq. 7.11) than by the relaxation law (Eqs. 7.15).

----~-----------------t

Fig. 7.8 Creep function.

--~--------------~~t

Fig. 7.9 Relaxation function.

On the other hand, if the generalized Maxwell model is considered as typical for
the behavior in relaxation, then a general form of the relaxation modulus is
J.1(t) =[GR +cl>(t)]H(t) +T\oa(t)

(7.24a)

where GR is termed the rubbery modulus, and cl>(t) the relaxation function. As depicted
in Fig. 7.9, cl>(t) monotonically decreases from an initial value and tends to zero as
t -+ 00. The initial value of cl>(t) is

Unear Viscoelasticity

245

(7.24b)
where G, is the glassy modulus.

G, and G R are the initial and final values of Jl(t), i.e.


G, =Jl(O"1

(7.24c)
(7.24d)

The stress-strain relationship for this case is given more simply by the relaxation law
(Eqs. 7.15) than by the creep law (Eq. 7.11). Substituting Eq. 7.24a into Eq. 7.15a we
fmd
<J(t) =110 l~ 8(t -s)d(s) +

i)G +cjl(t -s)]d(s).


R

Recalling properties of the Dirac-delta function and the argument used to put Eq. 7.6
as identical to Eq. 7.7, we can rewrite the equation above showing more explicitly the
discontinuity of e(t) at t =0 as
<J(t)

=11o(I)(t) + e(o"1Jl(t) + Jr'+ Jl(t o

s )de(s),

(t

> 0)

(7.24e)

where (I)(t) =de(t)/dt.


A special case of Eq. 7.23a is Eq. 7.19f of the Kelvin model for which we may
write
(7.25)
t is called the retardation time. Analogous to this is the relaxation time t which is the
time for Jl(t) in the Maxwell model to reduce to e-1 of its original value (see Eq. 7.18j
which is a special case of Eq. 7.22h), i.e.
Jl(t +t)

=e- Jl(t)
1

(7.26)

7.4 DIFFERENTIAL EQUATIONS FOR STRESS-STRAIN RELATIONSHIP


It should be possible to relate <J(t) to e(t) by means of a differential equation,
rather an integral one. Ifinitial conditions (at t 0"1 are prescribed properly, the solution
to that differential equation must agree with the creep law given by Eq. 7.11 as well
as with the relaxation law, Eqs. 7.15. A general fonn of such differential equation
would be
(7.27a)
P(D)<J(t) =Q(D)e(t)

where P(D) and Q(D) are differential operators, i.e. they are polynomial functions of
D which denotes differentiation with respect to time t,
D

=!

The Laplace transfonn of Eq. 7.27a is


P(p )cr(p) =Q(P )e(p)

(7.27b)
(7.28a)

Foundations of Solid Mechanics

246

Comparing Eq. 7.28a with the creep law (Eq. 7.11) and the relaxation law (Eqs. 7.15),
we find
P(P)=pQ(P}J(P)
(7.28b)
and
Q(P) =pP(p )fi(p)
(7.28c)
Naturally, the two equations above are consistent with the relationship between J(p)
and fi(p) as given in Eq. 7.14a.
The first four theorems to follow are related to the generalized Maxwell model.
Theorem 7.1. The relaxation function of the generalized Maxwell model satisfies
the following differential equation
P(D)Il(t) =aOG R

(7.29a)

where
N

P(D)=.L ap'

(7.29b)

.=0

Here ai(i =0 to N) are constants and N is the number of the Maxwell units in the model.
Proof of Theorem 7.1. Using Eqs. 7.22h and 7.29b we can put P(D)G(t) for t > 0
in the form
P(D)Il(t) =aoG o+

Jl Gie-I/'{~/-1laka~J

(7.29c)

where Go is identical to GR (Eq. 7.24d), and


-1
Gi
a ='t. =-

rti

Equations 7.29a and c are identical provided at (k

(7.29d)

=0 to N)

satisfy the equation

L (-1laka~ =0
k=O

(7.2ge)

Note that ai are the roots of the following polynomial equation


N

II(a-a.)=O

(7.29f)

II(a-a i) =(a-a 1)(a-a,J ..... (a-a N )

(7.29g)

i=l

where
N

i=1

It is possible to set the equation


(7.29h)
L (-1) aka = II(a-a;)
k=O
;=1
then determine each ak by equating the coefficients of ak. Values of at obtained in
such a way will satisfy Eq. 7.2ge. Thus the theorem is proved. Lastly, it is interesting
to note that
NkkN

Linear Viscoelasticity

247

P(D)J,t(t) = aOG R =

,,=0

(7.29i)

a"J,t(")(oj

where a superscript in parentheses denotes a time derivative, e.g. ( i") = d"( )/dt".
Theorem 7.2. If cr(t) and e(t) are continuous and N times continuously differentiable for t > 0, then the differential equation for the relationship between cr(t) and
e(t) is
(7.30a)
P(D)cr(t) = Q(D)e(t), (t>O)
where P(D) is as defined in Eq. 7.29b, and Q(D) is another differential operator one
order higher than P (D),
N+l

Q(D)= ,L bp'

(7.30b)

.=0

The relations among the coefficients of these operators are


N

b o = L a"J,t(")(O+)
,,=0

b r 1'\ oa r- 1 +

(7.30c)

L a"J,t(,,-r)(Oj,

(r

Il=r

=1 to N)

(7.30d)

bN + 1 =1'\ oaN

(7 .30e)

where 1'\0 is the constant of the 'free' dashpot (Fig. 7.5). Moreover, it should be proved
that the initial conditions are
N

N+1

r=k

r=k

L arcr(r-k)(Oj+1'\oak _1E(0+) = L bre(r-k)(Oj,

(k

=1

to N)

(7.30f)

Before proving this theorem, we state the lemma


NN

r-k

L B r- k ,,=,
L A"C,,_r =r=k
L Ar ,,=0
L B"Cr _k_",
r=k

(k

=1 to N)

(7.30g)

which may be proved by induction.


Proof of Theorem 7.2. The relaxation law is given by Eq. 7.15a or 7.24e, the
relaxation modulus by Eq. 7.22h or 7.24a, and coefficients ar by Theorem 7.1. Using
the formula for a derivative under an integral sign, Eq. 1.137, we fmd the time derivative
of Eq. 7.24e to be
cr(l)(t) =1'\ oe(2)(t) + e(OjJ,t(l)(t) + J,t(Oje(l)(t) + (I J,t(l)(t - s)de(s),

Jo+

(t > 0).

Using mathematical induction we can show that a time derivative of higher order can
be written

+ (I J,t(k)(t-S)de(s),

Jo

(t > 0,

k = 1 to N)

(7.30h)

Foundations of Solid Mechanics

248

Due to the equation above, we can write for t > 0,


P (D )a(t)

=110P (D )(I)(t) + e(O)P (D )~(t) + Jr'p (D )~(t - s )de(s)


o+

or, due to Eqs. 7.29i and 7.30g,


P(D)a(t)

=11oP(D)e(I)(t) +e(O)1:
a,,~(")(O)
,,=0

which is the same as Eq. 7.30a provided that Eqs. 7.3Oc, d and e hold. To prove initial
conditions, Eq. 7.3Of, we may start by writing Eq. 7.30h for t =0+, i.e.

I;
dl;)(O+) =11o(k+l)(O) + 1:

~(I;-")(O)e(")(O), (k = 1 to N).
,,=0
Replacing k by r - k, multiplying the result by art then taking a summation from r =k
to N, we have
,-1;

l: a,d,-I;)(O) =110 l: a,e(,-k+l)(O) + l: a, l: ~(r-I;-")(O)e(")(O), (k


,=1;
,=1;
,=1; ,,=0
N

=1 to N).

Making use of the lemma, Eq. 7.30g, we can put this equation in the form

1: a,d,-I;)(O) =110 1: a,e(,-k+l)(O) + l: e(,-I;)(O)l:


N

r=k

r=k

r=k

A=r

a,,~("-')(O),

(k

=1 to N)

(k

=1

=,;I11oa,-1 + ,,;, a,,~("-')(o)]e(,-I;)(O+)


+1loaN e(N -I; + 1)(O)-11oal;_Ie(O),

to N),

which is the same as Eq. 7.3Of due to Eqs. 7.3Od and e. Thus Theorem 7.2 is completely
proved.

Linear Viscoelasticity

249

Theorem 7.3. If the relationship between cr(t) and e(t) as given by Eq. 7.30a holds
( for t > 0), and the initial conditions are as in Eq. 7.30f, then the following relationship
holds for t ~ 0
(7.31a)
P(D)cr(t) =Q(D)e(t)
where
cr(t) =cr(t)H(t) +11oe(O~o(t)
e(t)

(7.31b)

=e(t)H(t)

(7.31c)

and
(7.31d)
Proof of Theorem 7.3. Due to Eq. 1.11Od, we can write
dk
k-l
-k [cr(t)H(t)] =d.k>(t)H(t) + I, cr<k-1-II)(0+)O(II)(t),
dt
11=0
Consequently,
P(D)[cr(t)H(t)]

=aocr(t)H(t) + Jl a k [ cr(k)(t)H(t) +

(k ~ 1)

(7.31e)

:t: cr(k-l-r)(o~o(r)(t)J.

or, due to Eq. 7.29b and the lemma, Eq. 7.30g,


P(D)[cr(t)H(t)]

= [P(D)cr(t)]H(t) +

I, O(k-l)(t) I, ardr-k)(o+)
k=l
r=k

(7.310

Similarly, we can write


Q(D)e(t)

=Q(D) [e(t)H(t)]
=[Q(D)e(t)]H(t) +

N+l N+l

I, O(k-l)(t) I, bre(r-k)(O~

k=l

r=k

(7.31g)
Due to Eqs. 7.31b, we have
N+l
P(D)cr(t) =P(D) [cr(t)H(t)] +11oe(O~ I, ak_1o(k-l)(t)
k=l

(7.31h)

Due to Eqs. 7.30a, 7.30, 7.31d, 7.31, 7.31g and 7.31h, Eq. 7.31a is proved. The
Laplace transform of Eq. 7.31a is given by Eq. 7.28a, i.e.
P (p )cr(p ) =Q (p )e(p )

(7.3li)

where
N

P(P)= I,arpr
r=O

(7.31j)

Foundations of Solid Mechanics

250

Q(P)=

N+l

l: brpr

(7.31k)

r=O

Theorem 7.4. If the relationship between O'(t) and (t) as given by Eq. 7.31a
holds, then the same relationship can be put in a convolution integral fonn as Eqs. 7.15,
the relaxation modulus of which is the same as that of the generalized Maxwell model,
i.e.
(7.32a)
J.I.(t) J.I.(t)H(t) + 11 oS(t)

where

bN +1

110=aN
The function J.I.(t) can be detennined from the differential equation
P(D)J.I.(t)

=bo'

(t > 0)

(7.32b)

(7.32c)

and initial conditions


J.I.(O~

=-aN1 [bN-11oaN - J

(7.32d)

J.I.(II)(O~ =J.-[b
N_II - "i:,l aN-lI+kJ.l.(II)(0~-11oaN-II-1]'
aN
k=O

=1 to N -1)

(n

(7.32e)

Proof of Theorem 7.4. Just as in Eq. 7.3lf, P(D)J.I.(t) for any time t can be put
in the fonn
P(D)[J.I.(t)H(t)] =[P(D)J.I.(t)]H(t) +

k=l

r=k

l: S(k-l)(t) l: arJ.l.(r-k)(O~

N-l

=[P(D)J.I.(t)]H(t) + l: ffk)(t)
k=O

l:

arJ.l.(r-k-l)(O~,

r=k+l

or, due to Eqs. 7.32a and 7.32c,


P(D)J.I.(t)

=bcll(t) + :;: S(k)(t)LJ+l arJ.l(r-k-l)(O~ + 11oak] + 11oa NS(N)(t)

(7.32t)

The Laplace transform of the equation above is


P(p )ii(P)

=bo+ Nfpi f arJ.l.(r-k-l)(O~ + 11oak] + 110aNpN


p k=O L.=k+l

(7.32g)

To establish Eq. 7.15a or b is the same as to establish its Laplace transfonn, i.e. Eq.
7.15c. Comparing Eq. 7.3li with Eq. 7.15c, we can see that the equation to be proved
is
(7.32h)
pp(p)ii(P) Q(P)

or, due to Eqs. 7.31j, 7.31k, 7.32b and 7.32g,


N

bk+l = rJ+l[arJ.l.(r-k-l)(O~+11oaJ,

(k =0 to N -1),

Linear Viscoelasticity

251

or
(7.32i)
which is identical to Eq. 7.3Od in Theorem 7.2 as should be expected. For k =N, Eq.
7.32i becomes identical to Eq. 7.32d; thus the remaining task is to show that Eq. 7.32i
for k = 1 to N - 1 is the same as Eq. 7.32e. With the introduction of a new symbol
n =N -k, Eq. 7.32i becomes
N-l

bN_" = L [arJ.!(ru-N)(O)+11oaN_,,_a +aNJ.!(")(O+),


r=N-"

or
1[
II(")(O~=bN
- -~I a lI(rU-N)(o~_'I"1 a

r-)

or, putting k

aN

N-"

r=N-"

rr-

10 N-,,-l ,

(n

=1

to N -1),

=r + n - N,

J.!(")(O) =~[bN_" aN

"i aN_n+kJ.!(k)(O+) -110 aN-"-I]'


l

k=O

(n

=1

to N - 1),

which is identical to Eq. 7.32e. Thus Theorem 7.4 is proved completely.


Being analogous to the preceding four theorems, the following four theorems are
related to the generalized Kelvin model which is described by the creep law more easily
than by the relaxation law.
Theorem 7.5. The creep compliance of the generalized Kelvin model satisfies the
differential equation
b
(7.33a)
Q(I)(D)J(t) =11:' (t > 0)
where
(7.33b)
Q(D)=

L bD'

(7.33c)

;=0 '

The b; (i =0 to N) are constants and N is the number of the Kelvin units in the model.
In addition 110 is the constant of the 'free' dashpot (Fig. 7.4).
Proof of Theorem 7.5. In view of Eqs. 7.21h and 7.33c, we can write for t > 0,
N

-1/~i[

Q(D)J(t)=bo;~oG/ 11> 11:t-;~1 eG; k~0(-1lbk(l

(7.33d)

where
(X.

I,

Consequently,

= 't-:-l = G./'I"1.
I'll

(7.33e)

Foundations of Solid Mechanics

252

(7.33f)
Equations 7.33a and 7.33f are identical provided bl: (k =0 to N) satisfy the equation
I: I:

1: (-1) bl:CLi =0
N

(7.33g)

1:=0

Such values of bl: are possible and can be obtained by the technique used in the proof
of Theorem 7.1. Thus Theorem 7.5 is proved. It is interesting to note that
Q(D)J(t)

1 b bo
=bo 1: -+-+-t,
N

(t > 0)

i=oGi Tlo Tlo

(7.33h)

and
(7.33i)
(7.33j)

Theorem 7.6. If cr(t) and e(t) are continuous and N times continuously differentiable for t > 0, then the differential equation for the relationship between cr(t) and
(t) is
P (D )cr(t) =Q (1)(D )e(t), (t > 0)
(7.34a)
where Q(D) is as defined in Eq. 7.33c and P(D) is another differential operator one
order higher than Q (D),
N+l
. N+1
di
(7.34b)
P(D) = 1: aD' = 1: a-.
;=0 I
;=0 'dt'
The relations among the coefficients of these operators are
bo
ao=Tlo
a'+1

= 1: b~(I1-')(O),
11='

(r

=0

(7.34c)
(7.34d)

to N)

The initial conditions can be shown to be


N

1: b (r-I:)(O) = 1: a

r=k '

r=k ,+1

a<r-I:)(O),

(k

=0

to N)

(7.34e)

In this case, the creep law Eq. 7.11 holds, the creep compliance is given by Eq. 7.21h
or Eq. 7.23a, coefficients b, are given by Theorem 7.5, and the proof follows the lines
adopted for Theorem 7.2.
Theorem 7.7. If the relationship between cr(t) and (t) given in Eq. 7.34a holds
(for t > 0), and the initial conditions are as in Eq. 7.34e, then the same relationship

Linear Viscoelasticity

holds for t

~ 0,

253

i.e.
P(D)o(t)

where

=Q(l)(D)(t)

(7.35a)

o(t) =o(t)H(t)

(7.35b)

=(t)H(t)

(7.35c)

(t)

The proof follows that for Theorem 7.3. The Laplace transform of Eq. 7.35a is
P (p )cr(p ) =p Q (p )e(p )
(7.35d)
where
N+l

P(P) = l: a,p'
,=0

(7.35e)

(7.35f)
Q(P)= l:b,p'
,=0
Theorem 7.8. If the relationship between o(t) and (t) given in Eq. 7.35a holds,
then the same relationship can be put in a convolution integral form Eq. 7.11. The
creep compliance is the same as that of the generalized Kelvin model, i.e. Eq. 7.21h or
Eq. 7.23a for which
bo
(7.36a)
Tlo=-

ao

The function I(t) can be determined from the differential equation


Q(l)(D )I(t) = be/Tlo,

(t > 0)

(7.36b)

and initial conditions


1(0)= aN+l
bN
1(1&)(0) =b1 [aN -1&+ 1 N

f bN_1&+J(k)(o)l,~

1&

k=O

(7.36c)
(n

=1 to N)

(7.36d)

The proof follows that for Theorem 7.4.


7.5 STEADY STATE HARMONIC OSCILLATION
If the stress o(t) is specified as a steady state harmonic function of time in the

form
(7.37a)
then the starting transient strain 'dies out' with time and only a steady state harmonic
strain is left, i.e.
(t)

=eoel(O>t-3>

(7.37b)

Foundations of Solid Mechanics

254

Without loss of generality, 0 0 , 0, m and a can be taken as positive real constants. More
specifically, 0 0 and 0 are the amplitudes of stress and strain respectively, m the angular
frequency, and a the phase angle. The stress-strain relationship may be put in the fann
o(t) = J.L(m)(t)

(7.38a)

or
e(t)=JO(m)o(t)
(7.38b)
The quantities J.L(m) and JO(m) are known as the complex modulus and the complex
compliance, respectively; they are complex functions of the frequency m. From these
equations we deduce
0 i5
J.L(m)=-e

(7.39a)

1
J (m)= J.L0(m)

(7.39b)

tana=~(m)

(7.39c)

J.Ll(m) .

(7.39d)
+"'~(m) + ~(m) = IJ.L(m) I
0
where J.Ll(m) and ~(m) are the real and imaginary parts of J.L(m), respectively, i.e.
00=

J.L(m) = J.Ll(m)+ill2(m)

(7.40)

Moreover, substituting Eqs. 7.37 into the relaxation law, Eq. 7.1Sb, yields for large t,

oi

mt

= oei(mt-a)i' e-iCD.rdJ.L(s), (t ~ 00),


0-

or

or, due to Eq. 7.39a,


J.L(m) = i~ e-lmrdJ.L(s)

(7.41)

Alternatively, substituting Eqs. 7.37 into Eq. 7.1Sa yields


ooeimt = oe-i8imi' J.L(t -s)elmrds, (t
0-

~ 00),

or, due to the commutative law for convolution integrals (Eq. 1.138a),

oi

mt

=oe-lliimi' elm('-8)J.L(s)ds, (t ~ 00).


0-

255

Linear Viscoelasticity

Alternatively

or, due to Eq. 7.39a,

Jl*(ro) =iroJl.(ro)

(7.42)

where Jl.(ro) is the Fourier transfonn of Jl(t), i.e.


Jl.(ro) =i~ e-imrJl(s)ds

(7.43)

If the relationship between the Fourier and Laplace transfonns as given by Eq. 1.143
holds, i.e.
(7.44)
Jl.(ro) = ,l(iro)
we can obtain the complex modulus Jl*(ro) from the Laplace transfonn of the relaxation
modulus as follows
Jl*(ro)

=iro,l(iro)

(7.45)

The inverse of Eq. 7.43 is

2i~ ~(ro)cosrotdro
Jl(t) =-

(7.46a)

2i~ ~(ro)sinrotdO)
Jl(t)=-

(7.46b)

1t

0-

or
1t 0-

where ~(ro) and ).I!(ro) are the Fourier cosine transfonn and Fourier sine transfonn,
respectively, of Jl(t), i.e.

~(ro) = i~ Jl(s)cosrosds

= 9HJl.(ro)]

(7.47a)

~(ro) = i~ Jl(s)sinrosds

= -S[Jl.(ro)]

(7.47b)

and 9t and S denote real part and imaginary part, respectively. Later, Eqs. 7.46 will
be rewritten for the generalized Maxwell model in a more convenient fonn to obtain
Jl(t) from Jl*(ro).
Similarly, we can obtain the following equations by means of the creep law:
(7.48a)
J*(ro) =iroJ*(ro)
where

(7.48b)

Foundations of Solid Mechanics

256

(7.48c)

If
J*(ro) =i(iro)

(7.48d)

J*(ro) = imi(iro)

(7.48e)

then
The inverse of Eq. 7.48c is

2121-

J(t) = -

J:(ro)cosrotdro

(7.49a)

J(t)=-

J!(ro)sinrotdro

(7.49b)

1t 0-

or

1t 0-

where
J:(ro) = 1~ J(s )cosrosds = 9t[J.(ro)]

J:(ro)

(7.49c)

=1~ J(s)sinrosds =-5 [J*(ro)]

(7.49d)

Later, Eqs. 7.49a and b will be rewritten for the generalized Kelvin model in a more
convenient form to obtain J(t) from J*(ro).
For the generalized Maxwell model with relaxation function given by Eq. 7.24a,
Eq. 7.41 becomes
j.1*(ro) = 1~ e-IcIlr[GR +~(s)]S(s)ds +

1:e-icBr~+1'1o 1~

e-IcIlrS(l)(s)ds,

or

or, integrating by parts,


j.1*(ro) = GR +~(O~-~(O~+iro

e-icBr~(s)ds +1'1oiro
Jro+

=GR + iro Jr-e-icBr~(s)ds + 1'1oiro


o+

(7.50)

Substituting Eqs. 7.24 into Eq. 7.43 yields j.1.(ro) in the form
j.1.(ro) = GR[1tS( ro) + ~J +

lro

or

Jro+-e-io>.r~(s)ds + 1'10

(7.51)

257

Linear Viscoelasticity

or, due to the fact (Eq. 1.11Ob) that c.oO(oo) =0,


ioof,.l..(OO) = GR +ioo

r- e-i.,.,<!>(s)ds +1'\oioo

Jo+

(7.52)

Comparing the equation above with Eq. 7.50, we can see that Eq. 7.42 is verified.
However, the relationship between f,.l..(OO) and ~(ioo) (Eq. 7.44) must be modified, due
to the term GRH(t) in the function fl(t), into the following (see Eq. 1.144a)
f,.l..(OO) = ~(ioo) + GR 1to(oo)

(7.53)

Multiplying the equation above by ioo and observing the fact that 000(00) = 0 we find
iOOIl.(OO) = ioo~(ioo)
(7.54)
Hence Eq. 7.45 still holds, though Eq. 7.44 does not The real and imaginary parts of
Eq. 7.50 are, respectively,
(7.55a)

flz(oo) =1'\000 +00

<!>(s)cosrosds
Jro
+

(7.55b)

Since
fll(O) = GR

(7.56)

the inverse Fourier transforms of Eqs. 7.55a and b give, respectively,

+~i-[ fll(OO): fll(O) ]sinootdOO}H(t) +1'\oo(t)

(7.57a)

={fll(O)+~i-[ flz(OO~1'\oOO]cosootdOO}H(t) +1'\oo(t)

(7.57b)

fl(t) = {fll(O)
fl(t)

The infmite integrals involve only absolutely integrable functions. While Eq. 7.50 gives
a way of finding fl(OO) from fl(t), Eqs. 7.57a and b give a way of finding fl(t) from
fl(OO). Explicitly, the complex modulus of the generalized Maxwell model can be
expressed as
(7.58)
For the generalized Kelvin model, the creep compliance are given in Eq. 7.23a,
or
(7.59)
where

258

Foundations of Solid Mechanics

(7.60a)

J; =J,+'I'(oo)
cjlC(t) ='I'(t)-'I'(oo)

(7.60b)
Note cjlC(t) monotonically approaches zero as t tends to infmity. Substituting Eq. 7.59
into Eq. 7.48a yields
J'(ro)=

i0-

e-lmr[J;+cjlC(s)]acs)ds+

i0+

1
e-lmrdcjlc+-

i-

110 0-

sacs)ds+-1

i-

110 0-

e-ilDrH(s)ds,

or, due to the properties of aCt) and the Fourier transform of H(t) (Eq. 1.l44a),
J'(ro) =J; +cjlC(O) +

r-e-lmrdcjlc +.![1ta(ro)
+~J,
110
1ro

Jo+

or, integrating by parts,


J'(ro) =J; +iro

r- e-ilDrdcjlC(s)ds +.![1ta(ro)
+~J
110
1ro

Jo+

(7.61)

On the other hand, the Fourier transform of J(t) is


J.(ro)

=i~ e-iGvJ(s)ds

or, due to Eqs. 1.144,


J.(ro) =J;[1ta(ro) +

i~J +

i:

e-ilDrcjlC(s)ds +

~J1tia<l)(ro)- ~].

or
imT.(ro) =J;[1tima(ro) + 1] +iro

r- e-lrD.rcjlC(s)ds +.![-1tma(l)(ro)+~J,
110
1ro

J~

or, due to the fact (Eqs. 1.110) that ma(ro) = 0 and ma(l)(ro) = -aero),
imT.(ro) =J; +irol-e-ilDrcjlC(s)ds +.![1ta(ro)+~]
(7.62)
~
110
1ro
Comparing the equation above with Eq. 7.61, we can see that Eq. 7.48b is verified.
However, Eq. 7.48d must be modified, due to the term (JR+t/11o)H(t) in the function
J(t), so that it reads
J.(ro) =i(iro) +J;1ta(ro) + 1tia(l)(ro)

(7.63)

110

Multiplying the equation above by iro and observing that ro3(ro) = 0 and roa(l)(ro) = -aero),
wefmd
',_,r (ro) = 1W1
',_,'i('1ro)+1ta(ro)
--

1W1.

110

(7.64)

259

Linear Viscoelasticity

Hence Eq. 7.48e must be modified to


. ,. T('lro) +
nB(
J "()
ro = lu.v
-ro)
-

(7.65)
1'\0
It should be noted that Eq. 7.39b can be used to obtain correct ~(ro) from a known
J"(ro), but in reverse the same equation cannot yield the term B(ro) in obtaining J"(ro)
from a known ~(ro). This complication does not arise if there is no dashpot in series
with other units in the model, i.e. when 1'\0 ~ 00. The real and imaginary parts of Eq.
7.61 are, respectively,
(7.66a)
(7.66b)
Because
the inverse Fourier transforms of Eqs. 7.66a and b give, respectively,
J(t) = {J1(0+)
J(t)

+~i-[ J 1(ro) -J1(0~ -nB(ro)/1'\o] sin rotdro + ~JH(t)

={J1(O)+~i-[ J2(ro)+~1'\oror1] cosrotdro+ ~JH(t)

(7.68a)

(7.68b)

Again the infmite integrals involve only absolutely integrable functions. While Eq. 7.61
provides a way of finding J"(ro) from J(t), Eqs. 7.68a and b give a way of finding J(t)
from J"(ro). Explicitly, the complex compliance of the generalized Kelvin model can
be expressed as

r n~~) + J.1 i;[1'\/t;(1 + ro2i;)f1 -i{(1'\oror1+ ;;1 roi;[1'\;(1 + ro2i;)f1} (7.69)

J(ro) = (Go 1+

Next, let us consider the dissipation of energy in steady state harmonic oscillation.
For a one-dimensional mechanical model, the work done in the time interval from to to
any t is

IlU =

9t[o(s)]9t(de)

(7.70)

Substituting Eqs. 7.37 into Eq. 7.70, in view of Eq. 7.39a, we find

IlU =~ro,~"(ro)

el(CDS-a)sin(ros -B)ds]

For the time interval equivalent to one cycle, the equation above becomes

(7.71)

Foundations of Solid Mechanics

260

or, due to Eq. 7.40,

~U =

1tfQJlz

(co) =1t~Jlz(co)
11l*(co)1 2

(7.72)

which is the amount of energy dissipation per cycle. Due to Eqs. 7.39c and d, this
equation can be rewritten
(7.73)
DefIning Umax as
(7.74a)

1 Il*(co)1 ,..2
=-1
Co
2

(7.74b)

we fInd the specific loss to be


~U 2 . s:
= 1tSlDu

Umax

(7.75)

If S is small, the energy dissipation per cycle and the specifIc loss assume the
approximate forms
~U = ~I 1l*(co)1 tanS

(7.76a)

~U =2manS

(7.76b)

Umax

Appropriately, Jlz(co) is called the loss modulus, and S the loss angle. For elastic models,
these two quantities are zero, and there is no energy dissipation.
7.6 THERMORHEOLOGICALLY SIMPLE SOLIDS
For a wide variety of materials known as thermorheologically simple solids, the
mechanical properties depend on the temperature in a simple analytical fashion. This
temperature dependence was fIrst proposed by Leadermann (1943) and Ferry (1970),
and applied by Schwarz! and Staverman (1952). The present derivation follows the
work of Morland and Lee (1960). For this type of materials, the temperature dependence
of the relaxation modulus is as shown in Fig. 7.10, in which
F (log t) =Il(t)
(7.77)

Linear Viscoelasticity

261

F(log t)

'-T-++--

log A (T)

O~------+-r---~~-------Iogt

f'~;[f+df
log

tA(T
10g(t+dt)A(T

Fig. 7.10 Relaxation modulus of thennorheologically simple solids.


Note in the figure that the curve for a temperature T has a constant 'shift' logA(T)
away from that for the reference temperature To, i.e.
F(logt)

=Fo[logt + logA(t)]
=Fo[logtA(t)] ,

or
~(t)=~@

(7.78)

where the subscript (0) denotes the reference temperature, and

1; =tA(T)

(7.79)

which is known as the pseudo or reduced time. Naturally A(To) is equal to unity. It
will be shown later that the creep compliance at an elevated temperature can be obtained
from that at the reference temperature in the same manner as Eq. 7.78, i.e.
~(t)

= ~(t)I,=tA(T)

(7.80a)
(7.80b)

In the process of getting ~(t) and l(t) by Eqs. 7.80, one should note the following
Heaviside step function, the Dirac-delta function and its derivatives
H[tA(T)] =H(t)
(7.81a)
o[tA(T)]

= o(t)

A(T)

(7.81b)

Foundations of Solid Mechanics

262

(7.81c)
In an isothennal process, T is kept independent of t, so the stress-strain relationship
assumes the usual fonn, i.e. the creep law or the relaxation law, Eq. 7.7 or 7.15a
respectively, while temperature enters the equations only as a parameter.
For isothermal harmonic vibration, Eqs. 7.37a and b become respectively
CJ(t) =CJilJ)~

(7.82a)

(t) =foeia:~

(7.82b)

where I; is given in Eq. 7.79, and 00 is the pseudo or reduced frequency, i.e.

00

00

(7.83)

=A(T)

The complex amplitudes CJo and Eo are related by the equation


CJo

=[J.11 (00) + ill2(00.)] Eo

(7.84a)

or
(7.84b)
where J.11(OO) and Il2(OO) are real and imaginary parts of the complex modulus J.1*(oo) at
the reference temperature To, and J 1(oo) and J 2(oo) are those of the complex compliance
J(OO). The dissipation of energy per cycle, Eq. 7.72, is

aU = xl EoI 21l2(OO) =-1t1

CJOI2J 2(oo)

(7.85a,b)

Figure 7.10 also shows that if the time taken for J.1 to decrease from some value
J.1 to J.1- is dt, then the time taken for Jlo to decrease the same amount is dl; =A(T)dt.
If I; corresponds to varying temperature T(t), then
J.1(t)

=A@ =Jlo(t)1 ,=~

A(T)dt < dl; < [A(T) +

dA:)~t) dtJdt

(7.86a)
(7.86b)

Neglecting the (dt)2 tenn in the equation above, we get


dl;=A(T)dt

(7.87a)

dl;=A(T)
dt

(7.87b)

or

Equations 7.87 hold for both isothennal and non-isothennal cases. Integrating Eq. 7.87b
and using the initial condition 1;(0) =0 we find
I;(t)

fA

[T(P)]dp

(7.88)

263

Linear Viscoelasticity

Applying Leibnitz's rule for differentiating an integral, Eq. 1.137, to this equation we
recover Eq. 7.87b. Putting t=t-s in Eq. 7.88 we find

~(t-s)= r-sA(T)dP.
This integral is the shaded area in Fig. 7.11, in which the identity
q=p+s
should be noted. Using q instead of p as the integrating variable we fmd

~(t -

s)

it

A (T)dq

=fA(T)dq- LSA(T)dq,
or
~(t -s)

=~(t) -~(s)

(7.89)

=J(~ - ~)

(7.90)

Thus if f(t) =J@, then we can write


f(t - s)
where ~ is as defined by Eq. 7.88, and

~ =~(s) =LS A (T)dp

(7.91)

The relationship between the relaxation modulus and the creep compliance, Eq. 7.14a,
is

j (P) =[p2~(p)r

(7.92a)

J o(P) =fp2~(p )]-1

(7.92b)

A (T)

Fig. 7.11 Graphical representation of ~(t - s).

Foundations of SoUd Mechanics

264

where

j (P)

and j o(p) are Laplace transfonns of J(E;) and lo(t). respectively. Note that

we can. without loss of generality. use the same Laplace transfonn parameter p against
both ~ and t. Due to Eq. 7.86a.
Thus Eqs. 7.92 give
(7.93)
Equations 7.86a and 7.93 are identical to Eqs. 7.80a arid b. respectively. Taking the
initial temperature. i.e. T(O) =To. we have A [T(O)] =A (To) =1. So the Heaviside step
function and the Dirac-delta function assume identical fonns with the argument t or ~.
i.e.
5(t)=5@
(7.94a.b)
H(t)=H@.
but the derivatives of the Dirac-delta function follow the following recurrence formula

=1.2.3......)

(7 .94c)

5<m)(E;) = A;T)![5<m-l)@]. (m = 1.2.3...... )

(7.94d)

a<m)(t)

=A (T) :~ [5<m -1)(t)]

(m

or. reversely.

As an example. both Eqs. 7.94c and d give 5(1)(t) =A(T)5<1)@. Incorporating the
temperature effects. the stress-strain relationship. Eq. 7.7 or 7.15a. can be written
(t) =a(t) + f/(t-S)dO'(S)
a(t)

=l~ ~(t - s )d[(s) - a(s)]

(7.95a)
(7.95b)

where a(t) is the strain due to the temperature change in a stress-free state. As
functions of the pseudo time ~. A(~) and J (E;) are given by Eqs. 7.86a and 7.93
respectively. while the others can be introduced as the following

a(~) =a(t)I,.,@

(7.96a)

(~) =e(t)I,=,@

(7.96b)

a(E;) = a(t)I,=,@

(7.96c)

and their time differentials take the form

df(s) =dJ(O =dJ(~) d~


d~

Substituting Eqs. 7.96 into Eqs. 7.95. and using Eq. 7.90. we obtain

(7.96d)

Linear Viscoelasticity

265

e(~) =a(~) + i~ J(~- Odcr(~)

cr(~) =i~ ~~ ~)d[(~) - a(~)]

(7.97a)
(7.97b)

The fmal results are then obtained as functions of t by simply noting that
/(t) =J(~I ~=~,)

(7.98)

and incorporating the special identities given in Eqs. 7.94.


7.7 THREE-DIMENSIONAL THEORY
The constitutive law for three-dimensional continua may be generalized from that
already described for one-dimensional mechanical models. We state it in the differential
fann
(7.99a)
P jj/d(D )CJ/d = Qjj/d(D )e/d
or the corresponding integral fonn

CJjj(t) = Cjj/d*de/d = e/d*dCjj/d

(7.99b)

The convolution integral symbol are as defined in Eq. 7.12. The latter fann (Eq. 7.99b)
is more convenient for describing various aspects of the theory, and we will use it
exclusively. As in linear elasticity, Cjj/d has the symmetric property given in Eq. 3.14
and is positive defmite at any value of t. Unless specified otherwise the discussion
which follows is restricted to isothennal conditions and thennorheologically simple
solids; the temperature enters only as a parameter, as in Eq. 7.79.
In the material domain V, the equation of motion and the strain-displacement
relationship are the same as those for any solids under small defannation, i.e. respectively
(7.100)

P. ..

"

1
2 ',J

=-(u.. +u .. )
10'

(7.101)

For isotropic materials, the stress-strain relationship, Eq. 7.99b, can be reduced to

CJjj(t)

=ajjA.*dekk + 2Jl*dv =ajjekk *dA. + 2v*dJl

(7.102)

which is analogous to Eq. 3.32 in linear elasticity. The introduction of the stress deviator
Sjj (Eq. 2.34) and the strain deviator ejj (Eq. 2.47) makes it possible to state separate
relationships for shear and bulk:
(7.103a)

Foundations of Solid Mechanics

266

(7.103b)
Here Jl(t) and K(t) are relaxation moduli for shear and bulk respectively, J(t) and B(t)
are their respective creep compliances. The relationship between A.(t), Jl(t) and K(t) is
the same as in elasticity, i.e. K (t) =')...(t) + 2Jl(t)/3. The relationship between a relaxation
modulus and its corresponding compliance is given in Eqs. 7.14. Equations 7.103 are
analogous to Eqs. 3.34 and 3.35 in isotropic elasticity.
For uniaxial stress in the xl-direction (only the stress G u is non-zero), Young's
modulus E(t) and Poisson's ratio v(t) can be introduced through the relationship
Gu(t) =E*du
(7.I04a)
and
(7.104b)

Ezz(t) = ~3(t) =-v*du

or, in Laplace transforms,


(7.104c)
and
(7.104d)
We can obtain the following relationship between Laplace transforms of material
functions as:

~(3x'+2~) 9jit
x,+~
~+3K

(7.105a,b)

X,
3K-2~
v=2p(X,+~) 2p(~+3K)

(7.105c,d)

Table 7.1
x,,~

X.

X.

E,v

~,K

Epv

3K-2~

2~pv

(1 + pv) (1- 2pv)

1-2pv

Jl

Jl

Jl

Jl

E
2(1 +pv)

X, +-Jl
23

E
3(1-2pv)

9K~
3K+~

~(3x'+2~)
x,+~

X,
2p{X,+u)

--

Jl,v

v-

3K-2~

2p(3K +u)

2~(1+pv)

3(1-2pv)
2~(1+pv)

267

Linear Viscoelasticity

A number of useful relations between the Laplace transforms of these isotropic relaxation
moduli are shown in Table 7.1, which is the analogue of Table 3.1 in elasticity theory.
Combining Eqs. 7.100, 7.101 and 7.102 we obtain Navier equations for viscoelasticity, namely

ifu

II*duI,D.. + (A. + ,...


")*duA,..
.., + X.I = p_'
at2

(7.106)

,...

which is the analogue of Eq. 3.66 in isotropic elasticity.


For quasi-static problems, the inertial term ifu/at 2 is neglected, so that the
equations of equilibrium become
0'IJ'}
... +x.=o
(7.107)
I
In terms of stress components, the compatibility conditions, Eq. 2.62, for isotropic
materials are
....*dJ +0'AI"
....*M+Sija.""tf.
..Y *dX+(X'.}
. .+XJ,'
. .)*dJ =0
(7.108)
0'v,""

where
L1(t)

=2[31(t) +B(t)]/9

(7.109a)

(7.I09b)
X,(p)=i(3i -2B)/(Y +4B)
Equation 7.108 can be called the Beltrami-Michell compatibility conditions for viscoelasticity. The energy identity is defined as
I(t) = Iv Xj*dujdV +

Is Xj*dujdS

(7.11Oa)

where S is the surface completely enclosing the domain V, and ll; is the surface force
(Eq. 2.6). This can be written
I(t) =

JvrO'IJ.. *de;.dV
v

(7.11Ob)
(7.11Oc)

or, for isotropic materials,


I(t) = Iv[(2~*deii)*deii + (K*dekk)*de".JdV

(7.111a)
(7.111b)

The Maxwell-Betti reciprocal theorem linking two states 1 and 2 can be written

X~l)*du~)dV + Is rl)*du~2)dS = X~)*du~l)dV + Is X?)*du~l)dS


"

"

"

'J

268

Foundations of Solid Mechanics

=Jvrd~)*de~~)dV
IJ
IJ

(7.112)

Equation 7.112 holds for anisotropy as well as isotropy. The volume change, i.e. Iv EkkdV,
can be evaluated by using Eq. 7.103b:

rEkkdV =.!.3JvrO'kk*dBdV

Jv

=j i[(XkO'q),j - xkO'q) *dBdV.


If B(t) is assumed to be spatially uniform, then Gauss's divergence theorem applied to
the fIrst term on the right-hand side yields

EkkdV

=j[L xknpq*dBdS - iXkO'q,/dBdV].

Using the stress vector formula (Eq. 2.6) and the equation of equilibrium (Eq. 7.107),
we obtain
(7.113)
This equation expresses the volume change of a homogeneous domain in terms of
integrals of the prescribed body and surface forces.
7.8 QUASI-STATIC SOLUTION BY SEPARATION OF VARIABLES
Following Alfrey (1944) for incompressible solids (v = 112) and Tsien (1950) for
a constant Poisson's ratio, i.e.
v(t) =vH(t)
(7. 114a)
we obtain separable solutions for the case in which the body force, boundary tractions
and prescribed boundary displacements have the form
Xj(Xj,t)=Xjo(Xj)X(t),

(in V)

(7. 114b)

Xj(xj,t)=~(Xj)X(t),

(on Sa)

(7.114c)

Uj(Xj,t)

=~(xJu(t),

(on S..)

(7. 114d)

where Sri and S.. are parts of the surface S. We assume a separable solution
uj(xj, t)

=ujo(xj)u(t)

(7. 115a)

Linear Viscoelasticity

269

(7. 115b)
Note that the displacement time function is the same as that prescribed on the surface
Su, and the stress time function is the same as that of the surface force on the surface
Sa and the same as that of the body force in the domain V.
In addition, the spatial functions in the equations correspond to the elastostatic
problem with unit shear modulus, i.e. they satisfy the equilibrium equations
crJiJ(Xi) + Xio(x;) =0

(7. 116a)

the strain-displacement relationship

="21 [ui,ix;)
+ Uj,i(X;)]

Eij(X;)

(7.116b)

and the stress-strain relationship (for 1.1 = 1)

2v

00

crij(xj) =(1- 2v) ijEkk(X;) + 2Eij(xj)

(7. 116c)

If the stresses are the basic unknowns in the solution scheme, then we must incorporate
the Beltrami-Michell compatibility conditions (Eq. 3.77)

1 o ,,(x,)+-u.Y
vs:ok k(X,)+X,.(x.)+X
cr kk(X.)+-cr
.. (x.)
I.},

l+v kk ,'J'
I-v i.f~",
',J'
!o"

=0

(7. 116d)

The boundary conditions are


(7.116e)
Ujo(X;) =~(x;),

( on Su)

(7.1160

The viscoelastic stress-strain relationship, Eq. 7.102 or 7.103, gives


X(t)

= Il*du

(7. 117a)

u(t) =J*dX
(7. 117b)
The displacements and stresses given in Eqs. 7.115 can be shown to satisfy all the
quasi-static viscoelastic conditions if the time functions for displacements and forces
are related by one of Eqs. 7.117. Note that the boundary surfaces must not move, i.e.
Sa and Su must not change with time.

7.9 STEADY STATE HARMONIC SOLUTION SCHEME


If the boundary conditions and body forces are specified as steady state harmonic
functions of time, the solution will have the same form

(7.l1Sa)
(7. l1Sb)
(7. 11Sc)

Foundations of Solid Mechanics

270

This solution is a special separable solution. The steady state harmonic stress-strain
relations for three-dimensional isotropic continua may be written
-'>
1::'1*
i) + 2J,L*(ro)ij(X
i)
uij(xJ
=uij'"
(ro)a(X

(7.119)

or, for shear and bulk separately,

=2J,L*(ro)e;(xi)
(J~(xJ =3K*(ro)e~(xi)
s;(x i)

(7. 120a)

(7. 120b)

The stress-strain relationship has exactly the same form as in elasticity, except that it
involves complex quantities. This means that we may solve the elasticity equations to
fmd the spatial parts Ui(xi)' ~(xi) and ~(Xi)' and then replace J,L and K in the elasticity
solution by J,L*(ro) and K"(ro) respectively.
Gottenberg and Christensen (1964) studied the torsional oscillation of a hollow
circular cylinder. They included inertial effects, and found analytical relationship
between the applied torque and the angle of twist. They used this relationship, together
with experimental measurements of the applied torque and the angular response, to
determine the complex shear modulus.
Note that the general treatment by separation of variables given in the previous
section can be applied only when inertial terms are neglected. By contrast, the solution
scheme for steady state harmonic oscillation discussed in the present section can include
inertial terms with no difficulty.
7.10 INTEGRAL TRANSFORM METHODS AND THEIR LIMITATIONS
The Laplace transforms of the equations of quasi-static viscoelasticity are identical
to those of linear elasto-quasi-statics. In particular, for the isotropic case (Table 7.2),
the Laplace transformed viscoelastic solution is obtained directly from the corresponding
elastic solution by replacing A., J.L, K, E and v with p'f...(p), p~(p), pK(P), pE(P) and
pv(P), respectively. The actual solution is then realized by inverting the transformed
solution. An entirely similar procedure also applies in the anisotropic case. If
appropriate, we may use Fourier transforms instead of Laplace transforms. The association of the viscoelastic solution with the elastic solution is called the elastic
viscoelastic correspondence principle, or the elastic-viscoelastic analogy.
We can include inertial terms in any integral transform method, but then (Table
7.3) the correspondence is between the transformed elastic solution and the transformed
viscoelastic solution. This correspondence is much less useful than that for the
quasi-static case, and there is in general no advantage in constructing the dynamicviscoelastic solution by this means. The dynamic-viscoelastic solution can be
approached directly by using an appropriate integral transform. The inversion of a
transformed dynamic solution is naturally much more difficult than that in the quasi-

Linear Viscoelasticity

271

Table 7.2
Elasto-quasi-statics

Equilibrium:
CJjjj+Xj=O
Strain-displacement relationship:
1
.. =-(u .. +u .. )
u 2 ',Il,'

Laplace transfonned
viscoelasto-quasi-statics

1(~
~)
uI,J+u
J,I

IJ.. =2

Stress-strain relationship:
sij=2J..1.eij

sij=2p!liij

o-a = 3pK'ea
Boundary conditions:
CJijnj =X j on So
Uj=Uj

o-ijnj =Xj on So

onS,.

static case. Again note that, in any integral transfonn method, the boundary surfaces
So and S,. must not change with time.
Gottenberg and Christensen (1964) applied the elastic-viscoelastic correspondence
principle to the quasi-static torsional oscillation of a hollow circular cylinder. Naturally,
the solution was found to consist of transient and steady-state parts. The transient part
is an exponentially decaying function of time.
In the book of Christensen (1971), there is a problem of a long cylinder being
subjected to an internal pressure, but restrained by an elastic outer case. Inertial effects
are included and the problem was solved by means of the Laplace transfonn.
Axisymmetry of the problem and the incompressibility of the material make the
mathematics tractable. Christensen and Schreiner (1965) studied the quasi-static
problem for compressible materials. The Laplace transfonn was applied to all conditions
involved. They also investigated a way of including the effect of a boundary which is
changing due to burning or erosion of the material.
Lockett (1961) studied the pressurization of a spherical cavity in an infmite space
using the Fourier transfonn. Inertial effects were included. They approximated the step
function pressure history by the difference between two sine integrals of difference
frequencies. This enabled them to use an inversion integral over a finite frequency.
range, rather than the usual infmite frequency range.

Foundations of Solid Mechanics

272

Table 7.3
Laplace transfonned
elastodynamics

Laplace transfonned
viscoelastodynamics

Equation of motion:
2
Ofij +Xj = pp u j
~

Strain-displacement relationship:
~)
e- IJ.. =-21(~u'.1+u
I.'

1(~
~)
e=IJ
2 u'.1+u
I.'

Stress-strain relationship:
iij =2j.liij

Sij=2pjieij

au = 3Keu

au =3pKeu

Boundary conditions:

aijnj =)( on Sa

iij

aijnj =)( on Sa

=u- on S,.
j

There are practical problems which involve variable boundary. One example is
the defonnation of a viscoelastic beam or a viscoelastic half space by a curved rigid
indentor. As the indentor is pressed into the viscoelastic domain, there are some parts
of the boundary of the domain which at flrst are traction free, but later they must
confonn to the geometry of the indentor in the contact region. For problems of this
class, the use of integral methods fails, and there is no generally available alternative.
Such problems are to be expected to be more difflcult than those with non-moving
boundaries. Some problems of this class have been solved by particular methods.
In his book, Christensen (1971) studied the simplest problem of this type, i.e. the
defonnation of a unifonn viscoelastic beam by a curve rigid indentor. It should be
possible to extend his procedure to dropping indentors, elastic beams with viscoelastic
coatings, culp indentors, viscoelastic plates, and viscoelastic beams and plates including
shear defonnation effects. This would comprise a generalization of the elastic contact
analysis given by Essenburg (1962).
The problem of the indentation of a viscoelastic half space by a rigid spherical
indentor was studied by Lee and Radok (1960) and Hunter (1960) and later by Graham
(1965). Christensen (1971) took the outline of the solution approach from Graham
(1965). More recently Graham (1967) extended his procedure to include the case where

273

Linear Viscoelasticity

the contact area has several relative maxima rather than just a single one. Ting (1968)
has also treated this more general case. Graham (1968) and Ting (1969) have outlined
restricted classes of viscoelastic contact problems which do admit a direct application
of the elastic viscoelastic correspondence principle.
A similar moving boundary problem, but even more complicated, is that posed by
a cylinder rolling over a viscoelastic half space. Here, again the elastic viscoelastic
correspondence principle does not apply, and other techniques of solution must be
developed. Such studies have been given by Hunter (1961) and Morland (1962, 1967).
See also Kalker (1990).
7.11 THREE-DIMENSIONAL THERMOVISCOELASTICITY
For thermorheologically simple solids, the stress-strain relationship for tbreedimensional domains with time varying temperature can be obtained by generalizing
that for one-dimensional models described in Section 7.6:
(7.121)
Here ~ and ~ are as defined in Eqs. 7.88 and 7.91, and av(t) is the strain tensor in the
stress-free state due to the temperature change from the initial temperature To to the
current temperature T(t). The isotropic form of Eq. 7.121 is
q(~)

Okk@

r~A

=2 Jo- Jl(~ -

~)de' q(~)

=3 i~ K(~- ~)d[Ekk(~) - akk(~)]

(7. 122a)
(7. 122b)

Specializing Eqs. 7.122 for the uniaxial stress case (Eqs. 7.104 with varying temperature)
we find

i~ E(~ - ~)d[El1(~) - a(~)]

(7. 123a)

~@ =~3@ =a@ - i~ v(~ - ~)d[El1 (~) - a(~)]

(7. 123b)

011@ =

where aCt) is the stress-free normal strain component in any direction due to the temperature change. The convolution integrals with respect to the pseudo time ~ in the
equations above seem to suggest that the Laplace transform might conveniently be used
to solve quasi-static boundary value problems. However, in terms of ~, the equilibrium
equations become
(7.124)

274

Foundations of Solid Mechanics

(7.125)
The presence of the tenns ~.j means that we cannot apply the Laplace transfonn with
respect to ~, nor construct a correspondence between elastic solutions and viscoelastic
solutions. When the temperature is spatially unifonn but time dependent, such tenns
~.j vanish, and then we can apply integral transfonn methods with respect to ~.
Another situation which is amenable to the use of integral transfonn methods
occurs when the temperature distribution T(xj,t) is known, whether from experimental
observations or by some analytical means. In such a case, the pseudo time ~ and the
stress free dilatation a are known functions of the coordinate Xj as well as the time
variable t. Muki and Sternberg (1961) solved a problem of this type using the Laplace
transfonn. The problem is that of an infinite isotropic viscoelastic slab subjected to
heating conditions which produce a known temperature variation through the thickness.
Lee and Rogers (1963) pointed out that this problem may be solved without using the
Laplace transfonn, and the problem was further discussed by Sternberg (1964). A
related but more difficult problem, involving thennal stresses in a cylinder, was studied
by Lockett and Morland (1967).
In a more general situation of coupled thermoviscoelasticity, the temperature upon
which the mechanical properties depend is an unknown function of Xj and t. These
problems are necessarily nonlinear, and must be expected to be very difficult to solve
analytically.
7.12 PROBLEMS

With permission of the publisher, Academic Press, Problems 7.12.4 to 7.12.6 and
7.12.10 are materials adapted/rom the book by Christensen (1971).
7.12.1 Reciprocal Theorem for Harmonic Oscillation
For steady-state harmonic oscillation, the Maxwell-Betti reciprocal theorem can be
written

Iv Xg)uJ~)dV + Is X<O~)uJ~)dS
=Iv X~)uJ:)dV + Is X!)uJ:)dS
= r[-pro2uJ:)~)+(J~P~~]dV
Jv
""

(7.126)

275

Linear Viscoelasticity

where a subscript (0) denotes the amplitude of a steady-state harmonic function, and a
superscript in parentheses denotes a loading system as adopted in Section 3.4.

7.12.2 Vibration of a bar with viscoelastic support


A viscoelastic slender bar of length L, cross sectional area A and mass density p
is supported viscoelastically and loaded axially as shown in Fig. 7.12. The bar material
has a constant Poisson's ratio but behaves like a standard linear solid in shear. The
whole system is in steady-state harmonic vibration due to the applied force pOcosoot.
Equations 7.1ge and 7.45 give the complex modulus of the support as

I
L
_
_
_
~
I
'
~

_ x----.
.r-,._____
.......;-=--_~ -

1-

--I

po cos wt

Fig. 7.12
GO(oo) =G(1 +irot)

(7.127)
where t=11/G. With the same symbols as in Fig. 7.3, the complex Young's modulus
of the bar is obtained as

2(1 +v)[Go+(Go+Gt)irottl
E (00) =
(1 . )
(7.128)
+lrot t
where v is the constant Poisson's ratio of the bar. In the bar, the pertinent stress-strain
relationship is

~(x) =Eo(oo) du~x)

(7.129)

and the equation of motion is

d~(x)
2 0
--=-prou (x)
(7.130)
dx
where a superscript (0) denotes the amplitude of a steady-state harmonic function.
Substituting Eq. 7.129 into Eq. 7.130 yields
<fuo(x) + duo(x) =0
dx 2

where

(7.131)

276

Foundations of Solid Mechanics

n---- W-'JrP
E*(ro)

(7. 132a)

=~IE{ro)1 rZ(COS~+iSin~)

(7. 132b)

Ez(ro)

tany=- E1(ro)

(7. 132c)

=9t[E*(ro)]
Ez(ro) =5 [E*(ro)]

(7. 132d)

E1(ro)

(7. 132e)

The solution to Eq. 7.131 takes the fann


uO(x) = pO[C1(ro) cos Ox + Cz(ro) sin Ox]

(7.133)

where C1(ro) and Cz(ro) are to be determined from the boundary conditions
A(J~(O) = G*(ro)uo(O),

A~(L) = pO

(7. 134a,b)

The results are

C1(ro) = [G (ro) cos nL -AnE (ro)sinnL]


G*(ro)C1(ro)
Cz(ro) = AnE*(ro)

(7. 135a)

(7. 135b)

Since the applied force is


pO cos rot = 9t(pOe imt )

(7.136)

the displacement of the end of the bar is


9t[uo(L)e imt ]

Due to Eq. 7.133, we can write


(7.137)
where
1J.*(ro) = [C1(ro) cos nL + Cz{ro) sin nLrl

(7.138)

Thus the loss angle (Eq. 7.39c) is


l) = arctan [ lJ.z(ro)]

1J.1(ro)

(7. 139a)

and the energy loss per cycle (Eq. 7.72) is


1t(p o)zlJ.z(ro)
llU=----

11J.*(ro)lz

where

(7. 139b)

277

Linear Viscoelasticity

(7. 140a,b)
7.12.3 Indentation on a Viscoelastic Half Space

H a normal indentation is made on the surface of a homogeneous isotropic elastic


half space by a smooth rigid circular plate, the stress at the contact between the rigid
plate and the half space is given by Eq. 6.78a. The half space is 0 S; r, Z < 00; the half
space surface is Z 0; the radius of the rigid plate is a; the contact surface is Z 0 and
S; r S; a; and Wo is the prescribed magnitude of the statical indentation. For quasi-statics
and a homogeneous isotropic viscoelastic half space with a constant Poisson's ratio (Eq.
7.114a), we suppose that the prescribed indentation is
u.(r,9,0; t) =woU(t), (0 S; r S; a)
(7.141)

The contact normal stress can be obtained by separation of variables (Section 7.8) as
2woX(t)
0'.. (r,9,0;t) =
112'
(OS;r<a)
(7.142)
n(1-v) (az-rz)
where X(t) is as given by Eq. 7.117a.
Alternatively, we can obtain the result by the elastic-viscoelastic correspondence
principle as
(7.143)
which is the same as Eq. 7.142.
7.12.4 Torsional Oscillation of a Hollow Cylinder

In this problem, we consider the torsional vibration of a hollow cylinder under


steady-state harmonic conditions. As in any axisymmetric torsion problems, the only
nonzero displacement is Us. and the relevant equation of motion for an elastic cylinder
is

J(fUe

~drz -

Ue 1 dUe (fUe)
(fUe
rZ+;:-a,:-+ dZz =P dtZ

(7.144)

where Ue is the displacement in 9-direction, p the mass density, and r, 9 and Z the
cylindrical coordinates. The nonvanishing stresses are O'ez and O'r9 which are related to
Ue through the equations

dUe

0: =Jlez
dZ

(7. 145a)
(7.145b)

278

Foundations of Solid Mechanics

The function Ue may be assumed in the fonn


Us =r4>(z )elmt

(7.146)

where 4l(z) is the amplitude of the angle of twist. Substituting Eq. 7.146 into Eq. 7.144,
wefmd
4>(z)=A

Sin~ +BCOS~

(7.147)

where
(7.148)

h is the length of the cylinder, and A and B are arbitrary constants. Note that O"r9 is
identically zero. The remaining nonzero stress O"er may be integrated over the crosssection of the cylinder to obtain the total twisting moment
M(

_1t(b4-a4)poi'h(A

z,t ) -

20

Oz -B . Oz
cos h
sm h

rLimt

(7.149)

where a and b are the inner and outer radii, respectively. If the end z =h is fixed
while the end z =0 has the applied torque M elmt, we can find the values of A and B
and further fmd that the ratio of the applied torque and the angle of twist at the end
z =0 is given by
M
1t(b 4 -a 4)pro2hcotO
(7.150)
4>(0) =
20
Replacing Il in the equation above by Il(ro) yields the required steady state harmonic
viscoelastic solution as
M
1t(b 4 -a4)poi'hcotO
(7.151)
4>(0) =
20
where
O =

roh~ Il(ro)
p

(7.152)

If we know Il*(ro), then Eq. 7.151 allows us to determine the angle of twist at z =0 for
a specified value of the applied torque. Alternatively, we may measure the applied
torque and the angular response and determine the complex shear modulus from Eq.
7.151, as did Gottenberg and Christensen (1964).

7.12.5 Quasi-Static Torsional Oscillation 0/ a Hollow Cylinder


We consider the same Problem 7.12.4, but now include transient starting effects.
While the cylinder end at z =h is fixed, the other end at z =0 has a prescribed angle
of rotation given

279

Linear Viscoelasticity
(t <0)

q,(O,t)=o,

q,(O,t)=ksinrot,

(O<t<oo)

(7.153a)
(7.153b)

Here k is the given amplitude of oscillation, and 00 is the angular frequency taken to
be sufficiently small that inertial terms may be neglected. The corresponding elastostatic
relationship between the twisting moment M (0, t) and the angle of twist q,(0, t) is given
by
(7.154)
or, due to Eq. 7.153b,
M(O,t)

=1t(b

-a;::"

sin rot

(7.155)

Applying the elastic-viscoelastic correspondence principle, i.e. replacing Il by p ~(P ),


M(O,t) by M(O,p) and q,(O,t) by <I>(O,p) in the equation above, we fmd the Laplace
transform of the viscoelastic torque to be
M(O p)

=1t(b 4_a4)p~(p )kro


2h (p2 + 00 2)

(7.156)

Assume the shear relaxation modulus in the form


Il(t)=[GO+Jl

Gie-t/~iJH(t)

(7.157)

Its Laplace transform can be put in the form (Eq. 7.22f)


_[

GiP] p-l

1l(P)= GO+I.--_1
i=lp +'ti

(7.158)

or, with a common denominator,


(7.159)
in which A(P) is an Nth degree polynomial in p. Substituting Eq. 7.159 into Eq. 7.156
and inverting the results by means of the Heaviside expansion formula (Spiegel, 1965),
wefmd

(7.160)

280

Foundations of Solid Mechanics

The ftrSt two tenns in this equation represent the steady state response, while the last
tenn describes the transient response. Gottenberg and Christensen (1964) used this
equation to predict the starting transient of the specimen, for which the complex shear
modulus was detennined in the preceding section.

7.12.6 Dynamic Response of an Incompressible Cylinder


Consider the problem of a cylinder subjected to an internal pressure and restrained
from expanding freely by an elastic outer case. The problem is axisymmetric, the
cylinder length is sufficiently large, and the cylinder material is incompressible. Thus
the zero dilatation condition can be written

au,.(r, t) u,(r, t)
ar

+-r-=O

(7.161)

where u,(r,t) is the radial displacement in the cylindrical coordinate (r,a,z) system.
The solution of this equation is

u (r t)= e(t)

(7.162)

"r

Hooke's law for the present problem is given by

au, II

(7. 163a)

a =2"-+"
t"'ar 3

u, II

ee =211-+t'" r
3

0:

(7. 163b)

II

a ... ="3

(7. 163c)

where II(r) is the ftrSt stress invariant, i.e. II = a" + aee+ a..,. Applying the correspondence principle to ~e elastic solution in Eqs. 7.163, we fmd the Laplace transfonns
of the viscoelastic stresses as
-(

)_

a" r,p -

2p~(P)C(P)
r2

il(r,p)

(7. 164a)

-(
)_2p~)C(P) il(r,p)
aeer,p r2
+ 3

(7. 164b)

_
il(r,p)
a..,(r,p) = - 3 -

(7.164c)

The single non-trivial equation of motion (in the r-direction) has the fonn

Oa" a" - aee

-+

or

ifu,.
-pat2

(7.165)

Linear Viscoelasticity

281

or, in the fonn of its Laplace transfonn,

dO-" 0-" - 0-00

-+
dr

r'

2_

=pp U

Substituting Eqs. 7.164 into this equation, we obtain


di 1(r,p) pp 2C(P)
3dr
r
or, integrating,
i 1(r,p) =3plogrp 2C(p) + 315(P)

(7.166)

(7.167)

(7.168)

Substituting Eq. 7.168 into Eqs. 7.164 we find


2pj!c
2CJ"(r,p) =--2-+ pp Clogr+D

(7. 169a)

2pj!c
2 CJoo (r,p)=-2-+ PP Clogr+D

(7. 169b)

o-..(r,p) =pp 2C logr +15

(7. 169c)

The boundary conditions of this problem are


CJ"(a,t)=-q(t)

(7. 170a)

CJ"(b,t) =-ku,(b,t)

(7. 170b)

where q(t) is the pressure prescribed inside the cylinder, and k is the stiffness of the
outer case restraining the cylinder. For a restraining case with a small thickness h,
Ech
k =b2(I-v~)
(7.171)
where Ec and Vc are the Young's modulus and Poisson's ratio of the case. We express
the relaxation function in shear as
A(P)
(7.172)
p~(P)= B(P)
where A (P) and B (P) are polynomials in p. The satisfaction of the Laplace transfonn
of the boundary conditions (Eqs. 7.170) detennines C(P) and 15 (P) as
C(P)

15(P)

pq(p)B(p)
F(P)

(7.173a)

pq(P)E(P)
F(P)

(7. 173b)

where
3
a pkB (P )
1 1)
F(P)=-2pA(P) ( a2- b2 +pp B(P)log"b
b

(7. 174a)

282

Foundations of Solid Mechanics

E(P)= A(P) -ppzB(P)logb _ kB(P)

bZ

(7. 174b)

If, for example, the applied pressure


q(t) =qJl(t)

(7.175)

qo
q~(P) =-

(7.176)

then its Laplace transfonn is


p

The denominator F (P) is a polynomial in p and can be written

(7.177)
-az> ...(p -am)
Here c is a real constant and aj are roots of F (P) =0; they can be determined by standard
F(P)= c(p -a 1)(p

computer programs. Substituting Eqs. 7.173 into Eqs. 7.169 and inverting the results,
by using the Heaviside expansion fonnula (Spiegel, 1965), we fmd
CJ"(r,t) =-qo

,f. Fj(t)[-2r zA (a) +Bj(r)lj

1=1

(7.178a)
(7. 178b)

CJ,.(r,t) =-qo ,l: Fj(t)Bj(r)


1=1

(7. 178c)

where
(7.179a)
Bj(r) =paJlogrB(a)+E(a)

(7. 179b)

7.12.7 Isothermal Harmonic Vibration

Some materials have real and imaginary parts of the complex compliance of the
fonn
(7. 180a)
(7. 180b)
where kl and kz are constants, and n is approximately 0.21. Moreover, the following
'shift factor' A(T) has been observed in a vicinity of a certain reference temperature

"oCT-TO>
A (T) = 10
= exp[2.303bo(T - TO>]

(7.181)

Unear Viscoelasticity

283

where bo is a constant and T is the temperature being kept constant all the time.
According to Eq. 7.83, the pseudo or reduced frequency becomes
00

=ooexp[-2.303bo(T - To)]

(7.182)

The real and imaginary parts of the complex compliance at the constant temperature T

are
11(00) =k1(00"f"

(7. 183a)

1 2(00) = ~(oo"f"

(7. 183b)

The dissipation of energy per cycle, Eq. 7.85b, is

flU =-n:1 (JoI2~oo-1l exp[2.303nbo{T - To)]

(7.184)

7.12.8 Isothermal Effects on Stretched String


Consider a viscoelastic string stretched between two fIxed points. The initial
tensile stress is crD, the properties of the string are uniform, the string material behaves
as a Maxwell solid in shear but as a Hookean solid in bulk, and the temperature T is
constant in space and time. The 'shift factor' A(T) is given as
A(T)

=eb,p-Tol

(7.185)

where bo is a known positive constant. Assume quasi-statics.


At the reference temperature To, the Laplace transforms of the bulk modulus Ko(t)
and the shear modulus Jlo(t) are given, respectively, by Eqs. 7.16c and 7.18h as

Ko(P) =Kp-l

(7. 186a)
(7. 186b)

Here K is the elastic spring constant in bulk, and t is given by Eq. 7.18f in which G
and 11 are, respectively, the elastic spring constant and the dashpot coeffIcient in the
Maxwell model as depicted in Fig. 7.2(a). With the help of Table 7.1, we can write
the Laplace transfonns of the Young's modulus and Poisson's ratio, respectively, at the
reference temperature as

Eo(P)=C(P +r-1fl

(7. 187a)

Vo(P) = (2pr1 +D(p +r-1f l

(7.187b)

where

9KG
3G
D=
3K+G'
2(3K +G)
3K+G
r= 3K t

(7.188a,b)
(7. 188c)

284

Foundations of Solid Mechanics

The inverse Laplace transfonns of Eqs. 7.187 give


Eo(t) =Ce-ttrH(t)
vo(t) =G+De-ttr}(t)

(7.189a)
(7. 189b)

The Young's modulus and Poisson's ratio at the temperature T, due to Eqs. 7.80 and
7.81, are
(7. 190a)
v(t) =v@ =

G+De-{/r}(~)

(7. 190b)

where ~=tA(T) as in Eq. 7.79 and H(~)=H(t) as in Eq. 7.81a.


Since the string has fixed ends, the longitudinal normal strain e(t) is a constant
function of t, i.e.
(7.191a)
or
(7.191b)
Here eO is the initial longitudinal normal strain, and can be related to the initial tensile
stress cf by means of Eq. 7.6, i.e.

(50

e=--

E(O+)

where E(O+) is the initial Young's modulus. Due to Eqs. 7.190,


1
E(Oi=C, v(Oi="2+D

(7.192)

(7. 193a,b)

The initial value of a function can be obtained from its Laplace transform by means of
Eq. 1.139. Substituting Eqs. 7.190a and 7.191a into Eq. 7.104a we find the longitudinal
normal stress
(7.194)
For the lateral nonnal strains, Eqs. 7.104b and 7.190b should be employed to obtain
&n(t) =

~3(t) = {~+De-tA(T)/r}oH(t) =-ev(t)

(7.195)

7.12.9 Varying Temperature Effects on a Stretched String

Consider the same viscoelastic string as in the preceding section, but allow the
temperature T to vary linearly with time t, i.e.
T(t) =To + aot
(7.196)
where a o is a known positive constant, and To is the initial as well as the reference

285

Linear Viscoelasticity

temperature. The thermal coefficient (normal strain per unit temperature change) is
known as c.
The Young's modulus and Poisson's ratio at any temperature T are given by Eqs.
7.190a and b respectively, but the pseudo or reduced time as given by Eq. 7.88, due to
Eqs. 7.185 and 7.196, is
~

=exp(aolv) -1

(7.197)

acPo

Hence
log(aobo~ + 1)

t=---

(7.198)

Equations 7.189 and 7.190 give the initial values


Eo(O+) = E (0+) = c

(7. 199a)

aobo

1
vo(Oj = v(Oj ="2 + D

(7. 199b)

The normal strain due to the temperature change in a stress-free state is


a(t)

=c(T - To)

(7.200a)

a(t)

=aoctH(t)

(7.200b)

or, due to Eq. 7.196,


or, due to Eqs. 7.94, 7.96 and 7.198,
J:

a(t) =a(':I)
A

C log(aobo~

bo

+ 1)

J:

H(':I)

(7.200c)

Substituting Eqs. 7.190a, 7.191b and 7.200c into Eq. 7.123a we fmd the longitudinal normal stress
r~

a(t)

=eOE@-Ccaoe-f"r Jo

e~/r

aobo~ + 1 d~
(7.201)

where
(7.202a,b)
Ql Q2
Q3
f(Q)=logQ +T!+ 2. 2! + 3. 3! + ...

(7.202c)

Substituting Eqs. 7.190b, 7.191b and 7.200c into Eq. 7.123b we fmd the lateral
normal strains
{c logQ Dce-{a+~/f)
}
En(t)=~3(t)=a(t)-eov(~)+ ----u;;;-+
bo
If(aQ)-f(a)] H@
A

(7.203)

Foundations of Solid Mechanics

286

Functions of ~ in Eqs. 7.201 and 7.203 can be express as functions of t by recalling


Eqs. 7.94 and 7.98, i.e.
Q-l
(7.204a)
~=-

e-

oJJo

E(~ = E(t) = C exp

Q
oJJrl'
}

(7.204b)

Q H(t)
oJJrl'
]

(7.204c)

e-

v@=v(t)= [12+ Dexp


where

(t)

Q=e"rN
7.12.10 Heating

0/ an Infinite

(7.205)

Slab

The problem is that of an infInite homogeneous isotropic viscoelastic slab of


thickness 2a being subjected to heating conditions which produce a known temperature
variation through the thickness. Rectangular coordinates are used such that the stress
free boundaries are at z =a. The displacements are zero except u.(z,t), the strains
are zero except ..,(z,t), and the nonzero stresses are O'...,(z,t), O'(z,t) and O'.. (z,t) with
O'...,(z,t)=O'(z,t)
(7.206)
The nontrivial equation of equilibrium is
oO'.. (z,t)
oz
=0

(7.207)

The pertinent boundary conditions are


O'.. (-a,t)=O

(7.208a,b)

Thus the only possible solution for O'.. (z,t) is trivial, i.e.
O'.. (z,t) =0

(7.209)

O'.. (a,t)=O,

With these conditions, the Laplace transforms with respect to


respectively

l; of Eqs. 7.122 give,


(7.210a)
(7.210b)

in which P is the transform parameter, and a(z,t) is the only nonzero strain due to the
temperature change in a stress-free state. Eliminating ~ from the pair of equations
above we obtain
(7.211)

Linear Viscoelasticity

287

where
(7.212)
4~+3K
The inverse Laplace transform of Eq. 7.211 is
~

o;a(z,~)

r~,q

=Oyy(z,~) =- Jo- .) (~- ~)da(z,~)

(7.213)

The integration above can be carried out fIrst with respect to the pseudo time ~, then
the fInal result as a function of the physical time t can be obtained by means of Eq.
7.98. Alternatively, Eq. 7.213 can be expressed in terms of t, so that
o;a(z,t) = Oyy(z,t) =-

S(t-s)da(z,s)

(7.214)

When the temperature T is a known function of z and t, a(z,t) can be derived, ~ can
be found from Eq. 7.88, @ from Eq. 7.212, S(t) from Eq. 7.98, and thence the stresses
from Eq. 7.214. A more general treatment of this problem was given by Muki and
Sternberg (1961) with detailed numerical results.

CHAPTER VIII
WAVE PROPAGATION

8.1 TERMINOLOGY IN WAVE PROPAGATION


When a portion of a defonnable medium is disturbed, it is defonned; the deformation disturbs the neighboring parts of the medium, and so propagates the disturbance
through the medium. Such a disturbance is tenned a wave and its progress through the
medium is called wave propagation. As the wave propagates it carries energy in the
fonn of kinetic and potential energies. The transmission of these energies is not by
any bulk motion of the medium, but is passed on from one particle to the next. In
other words, waves in defonnable bodies are characterized by the transport of energy
through motions of particles about an equilibrium position. Defonnability and inertia
of a medium are essential for the occurrence of waves. If the medium were not
defonnable, any part of the medium would immediately experience a disturbance upon
application of a disturbance to a certain part of the medium, thus the response of the
medium to the disturbance would be immediate, not gradual as in wave propagation.
Similarly, if the medium had no inertia there would be no delay in the displacement of
particles, and the transmission of the disturbance from one particle to another would be
instantaneous. When a wave propagates through a three-dimensional medium, we can
at a certain instant of time draw a surface through all points undergoing an identical
disturbance. As time goes on, such a surface, which is called a wavefront, moves along.
The nonnals to the wavefront, deflning the direction of wave propagation, are called

rays.
In one-dimensional problem a wave may be deflned by a function t(x - ct), where
a function of the spatial coordinate x and the time t, represents a physical quantity
such as a displacement, a velocity or a stress component. Figure 8.1 shows two identical
fonns of the function t at two instants, t and t + Ilt for a sufflciently small M , thus
t(x -ct) =t(x +llx -ct -CM),
hence
x -ct =x +llx -ct-cM,
or
c =llxIM.

t,

289

Wave Propagation

!(x-ct)

o~----------------------------

Fig. 8.1 An outgoing simple wave.


Appropriately, c is termed the phase velocity, andf(x -ct) a simple wave function since
it propagates though the medium without change of shape and magnitude. Iff is the
displacement of the particle at position x and time t, then df/dt is the particle velocity.
This is not the same as the phase velocity. An observer at the origin 0 will 'see' that
the wave f(x - ct) shown in Fig 8.1 is moving away from him/her with the speed c.
Such a wave is called outgoing. An incoming simple wave is described by a function
f(x + ct). If the shape of a wave does not travel, it is called a standing wave. A
standing wave can be obtained by superposition of an outgoing simple wave and an
incoming simple wave of the same characteristics, f(x - ct) + f(x + ct), e.g.
2sinroxcosrot sinro(x - ct) + sinro(x + ct).
Some travelling waves are not 'simple' but attenuate. Figure 8.2 shows the wave with

f(x, t) =e--J,.(X - ct),

where a is a positive real constant. The plots of the function f at two separate instants,
t and t + t!J, show that the function is unchanged in shape but decreases in magnitude
as time increases.

Ax = cAt
Envelope

-ax

o~--------------------------~

Fig. 8.2 An outgoing attenuating wave.

290

Foundations of Solid Mechanics

The three-dimensional wave equation is Eq. 3.111 or 3.112. In one dimension


this reduces to the simple wave equation.
"iff =c-2 "iff
(S.l)
ax 2
at 2
In oder to fmd the general solution to this equation we introduce two new variables
y=x-ct, z=x+ct
(S.2a,b)
so that
(S.2c,d)
x =(y + z)/2, t =(z - Y )/(2c )
Now

l(a la)

- - -aax
- + -aat
- - - ---ay -axay atay -2 ax

cat

(S.2e)

and
(S.2f)
Thus

(~-:2~[z)=(! -~:r)(;~ +~~)=4a~'z =0

(S.2g)

which has the general solution


(S.3)
f(x,t) =f(y) + f(z) =f(x -ct)+ f(x +ct)
Consider the problem of an elastic slender bar under axial excitation. Ifthe xl-axis
runs along the bar, the equation of motion of the bar in the absence of any body force
is specialized from Eq. 2.10 to be
"ifu1
O"U,l =P at2 '
Since

O"u

=EU1,l (Eqs. 2.43a and 3.36a), this is


"ifUl P"ifUl
ax; ="E at2

(S.4)

Thus, the axial wave speed is


Co

=VE/p

(S.5)

Consider a one-dimensional plane harmonic wave


(S.6)
u(x,t) =A exp[i(kx - rot)]
where k and ro are real constants independent of the spatial variable x and the time t.
The term A is known as the amplitude, and ro as the angular frequency. The wave has
phase velocity c given by
ro
c=(S.7)
k

291

Wave Propagation

Figure 8.3(a) shows a plot of the wave at a particular time, while Fig. 8.3(b) shows
that at a particular position. Ifx increases by 21C1k or t by 21C1ro, the argument kx - rot
changes by 21t so that the exponential function returns to its original value. This shows
that the wave has a wavelength A and a period T given by
~(U ),S'(u)

~(U),S'(U)

t"\ ~----h ~

~\JV'x

I-A-l

(a) t =constant

"

~T---j
~ /\. C'.t
~rszr=p
(b) x

=constant.

Fig. 8.3 Wavelength and period of a harmonic wave.


(8.8a,b)
A =21t T =21t
k'
ro
The quantity k (= 21C1A) which counts the number of wavelength in 21t, is termed the
wavenumber. A system is said to be dispersive, if its wavenumber k depends on the
frequency ro, or, in other words, its phase velocity c depends on the wavelength A.
Let us consider the superposition of two plane harmonic waves of the same
amplitude but slightly different wavenumbers and frequencies. The resultant of the
superposed waves is
u +u' =A{exp[i(kx -rot)] +exp[i(k'x -ro't)]}

1
i
= 2A eXP2[(k +k')x - (ro+ro')t] cos 2 [(k -k')x - (ro- ro')t]

(8.9)

which has the characteristics of 'beats'. The exponential factor represents a carrier
wave whose wavenumber and frequency are equal to the means of the superposed waves.
The cosine factor, which varies slowly when the difference in characteristics of the two
- - Cg

MOdula_ti~n)

Carrier
wave

---c

Fig. 8.4 An illustration of a wave group.

292

Foundations o[ Solid Mechanics

waves is small, represents the modulation and may be regarded as a varying amplitude.
The situation is depicted in Fig. 8.4. The 'wave group' ends wherever the cosine
becomes zero. The velocity of the advance of these points is called the group velocity
with the value

ro-uJ

c,= k-k"
For long groups (or slow beats), the expression above may be rewritten with sufficient
accuracy as

dro

(8. lOa)

c
, =dk

de
=c+k

(8. lOb)

de
=c-A
dA

(8.1Oc)

dk

and another useful relation is


1
c,

1 ro de

=;;- cZdro

(8.1Od)

In general c, < c, and thus dcldk < O. The group velocity is the same as the velocity of
energy transport. This is defmed as the ratio of the time average of the power per
cross section and the time average of the total energy per unit length of the wave. Some
verification of this point may be seen in the book by Achenbach (1973). If the system
is nondispersive, ro is not a function of k (hence A as well), and c, becomes identical
to c.
If the only non-zero displacement component is Ul and it is a function of Xl only,
then the governing equation for an isotropic elastic domain can be obtained by specializing Eq. 3.66 into
azUl

azUl

ax? =)'+2jJ. atZ

(8.11)

which is a special case of the three-dimensional pressure wave equation, Eq. 3.111. The
equation above implies that the motion is in the same direction as the propagation,
therefore the wave is called a longitudinal wave, compressive wave, pressure wave or
a P-wave (P standing for pressure or primary), and the wave speed is

=~)'+2jJ.
P

(8.12)

For incompressible solids, cp becomes infinity following the value of A..


For the same type of domain but the only non-zero displacement component
being a function of X z only, the governing equation is

Ul

293

Wave Propagation

(S.13)
which is a special case of the three-dimensional shear wave equation, Eq. 3.112. The
equation above implies that the motion is normal to the direction of the propagation,
therefore the wave is called a transverse wave, rotational wave, shear wave, or an
S-wave (S standing for shear or secondary), and the wave speed is
Cs

=-.J~p

(S.14a)

=-V A.:2~CP

(S.14b)

-2V
c
~
2(1-v)

(S.l4c)

Thus c. is always smaller than Cpo


If in such a domain the dilatation is zero or uniform, i.e.
eu=O

(S.lSa)

=0

(S.lSb)

or
eU,i

then the governing equation, Eq. 3.66, becomes


2V2

C.

at

_ (fUi
2

ui -

(S.16)

Comparing this equation with Eq. 3.112, we can say that it is an S-wave equation
governing each displacement component individually. Recall that the radial spherical
coordinate R is related to the Cartesian Xi as

R 2 =XiX i

(S.17)

The derivative of the equation above with respect to Xk is

aR

2R aXk = 2Xik'
or

or, by squaring,

or, due to Eq. S.17,


(S.lS)

294

Foundations oj Solid Mechanics

Taking derivative of the equation above with respect to R we fmd


oR ~(OR)_ (fR_O
OXk oR OXk - ox; -

(8.19)

If Ui is independent of any angular spherical coordinates, then


OUioR
Ui,k =oR OXk'
and

(8.20)
or, due to Eqs. 8.18 and 8.19,
(fUi
Ui,kk =OR2
Substituting Eq. 8.21 into Eq. 8.16 we obtain the simple wave equation
(fUi
_2(fui
OR2 =c. ot2

(8.21)

(8.22)

The wave is called an equivoluminal wave, and has a spherical wavefront and a wave
speed equal to Ca. Note a consequence that any wave equation is transfonnable into a
simple wave equation if the spherical radial coordinate R is the only spatial variable
involved.
If the defonnation in such domain is irrotational, i.e.
(8.23)
ClliA; =0
or, by defmition of the rotation tensor (Eq. 2.48),
1
"2 (Ui,k - Uk,i) =0,
then the tensor Ui,k is symmetric, i.e.
(8.24)

Accordingly,
or
kk,i = Ui,kk
Substituting this equation into Eq. 3.66 we find
2 2

(fUi

cp V Ui = ot2

(8.25)

(8.26)

295

Wave Propagation

Comparing this equation with Eq. 3.111, we can say that it is a P-wave equation
governing each displacement component individually. In the same way as obtaining
Eq. 8.22, we can change this equation into the simple wave equation
a"Ui

-2 a"Ui

oR2 =cp ot2

(8.27)

Therefore the wave is called an i"otational wave or a dilatational wave, has a spherical
wavefront, and the wave speed is Cpo
When each displacement component assumes the form
ui =dJVclnl - ct)
(8.28)
where di and ~ are unit vectors deftning the directions of motion and propagation,
respectively, the wave is a plane wave since the quantity Xini = constant describes a
plane with ni as its unit normal vector. The wave is propagating outward with the phase
velocity c. Introduce a new scalar
(8.29)
Y =xlnl-ct
Accordingly,
(8.30a)
(S.30b)
(8.3Oc)

or

(8.3Od)

(S.30e)
If
ni=di

then Eqs. 8.30b and c become identical, i.e.


or, by definition of the rotation tensor (Eq. 2.48),

(S.31)

Foundations of Solid Mechanics

296

(8.32)

COli =0

meaning that when the directions of motion and propagation are the same, the wave is
'irrotational'. On the other hand, if
nA, 0
(8.33)

then Eqs. 8.3Od and e give, respectively,


H:
H:,i

=0

(8.34a)

=0

(8.34b)

meaning that the wave becomes 'equivoluminal' when the directions of motion and
propagation are orthogonal.
In the following three sections (Sections 8.2, 8.3 and 8.4) we shall derive a theory
for the propagation of waves in continua. The general conditions for wavefronts,
dynamic and kinematic jumps, and rays in any type of solids will be considered ftrst.
The derivation of a wavefront from a preceding wavefront will be made. Nonstationary
and stationary surfaces of discontinuity will be deftned. Then inhomogeneous isotropic
elastic media will be considered. Pressure and shear waves will be distinguished. The
relationship between a velocity jump at a wavefront and that at a preceding wavefront
will be derived, and subsequently specialized for homogeneous isotropic elastic solids.
Then the problem of reflection and transmission of waves from an arbitrary smooth
interface of two media will be considered. Finally, we shall treat the reflection of waves
from the bounding surface of a body for various types of boundary conditions. The
derivation (in Sections 8.2, 8.3 and 8.4) is based upon the work by Keller (1964).
8.2 WAVEFRONT AND JUMPS
Let V be an arbitrary domain with a piecewise boundary S in the four-dimensional
x,,-space in which the position vector components are

{~={;}

(8.35.)

and let a four-dimensional vector be defmed as

{'tn.}

=1::: )=1 ::: )


'ti3

(Ji3

'ti4

-PVi

(8.35b)

Note that the subscript v runs from 1 to 4, i from 1 to 3, and Vi == au/dt. The equations
of motion (Eq. 2.10) can be written

297

Wave Propagation
tIV,V =-x.I

(8.36)

in which the comma-subscript for a derivative and Einstein summation conventions are
employed. Integrating the equation above over the hyperdomain V and applying Gauss's
divergence theorem (Eq. 1.88), we get

Is tjy~dS =- LXidV

(8.37)

where ~ is the outward unit normal to the hypersurface boundary S. This equation is
one of the most general forms of the equations of motion.
Consider two portions of V, i.e. VI and V2 which are so chosen that SI_2 is their
common boundary (see Fig. 8.5). The disturbance is taken as being in VI only. If ~
is the unit vector normal to SI-2 and outwards from VI' then -{. is the unit vector
normal to SI_2 and outwards from V2 For domains VI and V2, Eq. 8.37 gives

rtjy~dS + ~~
r tjy~dS =- J~rXidV

(8.38a)

rtjy~dS - ~~
r tjy~dS =- J~rXidV

(8.38b)

Noting that tjy is discontinuous on the hyperwavefront SI-2 and that the integrals in Eq.
8.37 are just summation of those for VI and V2 , the sum of Eqs. 8.38 gives

Jsr [tjy]~dS =0
1_ Z

where [tiv] denotes the jump in tjy across

SI_2'

(8.39a)

i.e.

[tjy] =t: - t~

(8.39b)

The supenninus C) and superplus (+) signs denote the quantity in the disturbed domain

Fig. 8.5

Hypersurface

SI-2

separating disturbed hyperdomain VI

from undisturbed hyperdomain V2

298

Foundations of Solid Mechanics

behind the wavefront, and that in the undisturbed domain in front of the wavefront,
respectively. Since Eq. 8.39a must hold for an arbitrary section of Sl-Z, we can conclude
that
[tivl~ =0
at all points of the surface of discontinuity. Equation 8.35b shows that this equation
is equivalent to
[crii]~j = [PVi]~
(8.40)
where i ,j = 1 to 3. If the hypersurface of the discontinuity is represented by
<1>(xi , t) =0

(8.41)

then we have (Eq. 1.99)

(8.42)

where

<1>.,=

a<1>

at'

There can be no jump in displacements even at the hyperwavefront

Sl-Z

and the

same can be said for a tangential derivative, i.e.

[~:i] =0

(8.43)

where s is an arc length of a space curve tangent to


aUi

as

Sl-Z'

ax"

-=u -

Now,
(8.44)

,.vas

and, as discussed in Section 1.8,

ax"

as

(8.45)

-=N

where Nv is a unit vector tangent to s thus also to


that

Sl-2'

Equations 8.43 to 8.45 show


(8.46)

which implies that the vector [Ui.v] in the four-dimensional space is normal to any N v
Thus it must be normal to ~, i.e.
[U.I,V ]

=N.r

-,:~,

where <X; is a scalar (in the four-dimensional space). The equation above can be rewritten
more explicitly as

299

Wave Propagation

But Ui.t is the velocity Vi' so that


2 -1/2

lUi) = (li(cI>.kcl>.k + cI>,,)

cI>.i

(S,47a)

2 -112

(S,47b)
[Vi] = (li(cI>.kcl>.k + cI
cI>.t
where i, j, k = 1 to 3. Detennining <l; from Eq. S,47b and substituting the result in
Eq. S,47a we fmd that, if cI>.t ::F- 0, then
[u . .] = [vi]cI>.i
'.J
cI>.t

(S,4S)

Suppose that we have solved Eq. S,41 for t as a function of Xi' Then we may
write
(S,49a)

cI>(Xi, t) == 'I'(x;) - t
So that the surface of discontinuity is given by

(S,49b)
'I'(xi) =t
The surface 'I'(Xi) = constant (in the physical xi-space) is a wavefront across which the
stress may be discontinuous at an appropriate instant of time. The function 'I'(Xi) is
known as the wavefront function. Accordingly, we can write

cI>.i ='I'.i

(S.50a)

'"
'Y.t

(S.50b)

=-1

d'l' = 1
(S.SOc)
dt
Figure S.6 shows wavefronts at times t and t + I1t spaced apart by a distance &, i.e.

Wavefront at t + M

3
Fig. S.6 Wavefronts at times t and t + 11t.

Foundations of Solid Mechanics

300

& =ellt

= elltni
dxi
-=en
dt
1

(8.51a)
(8.51b)
(8.51c)
(8.51d)

where e is the wave speed, and ni the unit vector nonnal to the wavefront, i.e.

'JI,i
-v'JI,j'JI,j

n.=_~
1

(8.52)

Equation 8.51a shows that


d

-1

-=e ds
dt

(8.53)

Equation 8.49b shows that


or

L\'JI =llt.
If we divide this equation by .!\xi' we get
'JI,i =L\t/.!\xi

(8.54a)

'JI,i 1
L\t =.!\xi

(8.54b)

or,

Multiplying this equation by A (Xi + .!\xi) - A (x;), we get

.dA=A.
,I
"',I dt

(8.55a)

or
dA A ,_"'.llf.
___
1'_,"
dt -

llf

.'Iftl.

't',J

But, by the chain rule of differentiation


dA
dxi
dt=A,idt'
or, due to Eq. 8.51d,
dA
-=enA
..
dt
1
,I

Equating this equation to Eq. 8.55b we find


A ,IT,l
.llf.
eniA, i = - - '
'JI,j'JI,j

(8.55b)

301

Wave Propagation

Since the equation above holds for any A (Xi), we can cancel A.i from both sides of it,
and obtain

'I'.i
'I'.i'l'.i

cni =--,
or, due to Eq. 8.52,
-2

'I'.i'l'.j =c

(8.56)

This equation, known as the Eiconal equation in geometrical optics, is a nonlinear frrst
order partial differential equation, the solutions of which yield possible wavefronts.
Using this equation in Eqs. 8.52, 8.54a and 8.55b, we fmd
(8.57a)
ni =C'l'.i
l1t

-1

-=c
n
Axi
I

dA

-=c2...A
.
dt
'1',1.'

(8.57b)
(8.57c)

or
dA

-=cnA.
dt
1.1

(8.57d)

or
dA

dxi

-=A.dt
.Idt
Incorporating the results above in Eqs. 8.40 and 8.48, we get, respectively
[<Jij]nj =-c[pvi ]

lUi) =-C- 1[vi]nj

(8.57e)

(8.58a)
(8.58b)

Equation 8.58a is known as the dynamic jump condition, and characterizes the
momentum conservation across the wavefront, i.e. the jump in surface traction is
proportional to the jump in momentum. Equation 8.58b is known as the kinematic jump
condition.
Differentiating Eq. 8.57a with respect to t, we get
dn i

dt

or, due to Eq. 8.51d,

de
d'l'.i
= dt'l'i+Cd/

Foundations of Solid Mechanics

302

or, due to Eq. 8.56,


dni
dt

_Ide
dt'

-=e -n-e

(8.59a)

"

or, due to Eq. 8.53,

or, due to Eq. 1.91a,

KN.
e=e -Ide
-n-e
dt
fl

(8.59b)

where K and Ni are curvature and unit nonnal, respectively, of the ray. When e is
constant, Eqs. 8.59 give a constant ni and zero curvature meaning that the ray is straight.
Differentiating Eq. 8.57a with respect to Xi we find
or, due to Eq. 8.59b,
-2 de

e"'
.. = n','. - e -dt
T,"

By Gauss's divergence theorem (Eq. 1.87), we may write the tenn

(8.60)
ni,i

as

n . . =Lim! nB.dS

',' v-+ovJs"

(8.61)

in which V is the volume of the space containing a 'bundle' of rays, S the surface

Fig. 8.7 Same bundle of rays nonnal to wavefronts at times t and t +.1t.

303

Wave Propagation

completely enclosing V, and Bi the unit vector normal to S and outward from V. The
situation is depicted more clearly in Fig. 8.7. Note that the scalar product niBj on the
two ends and curved surface of S is
njB j =-1, on S(t)
(8.62a)
= 1, on
= 0,

S(t+.1t)

elsewhere

(8.62b)
(8.62c)

in which S(t) denotes the area of the wavefront at the time t. Incorporating Eqs. 8.62,
we can rewrite Eq. 8.61 as

n. .
','

=Lim S(t +.1t) -S(t)


v -+0

=Lim S(t +.1t) -S(t)


tV-+o

S(t)c.1t

c-l dS (t)

(8.63)

=S(t)dt

The relationship between small areas of wavefronts at times to and t can be put as

S(t) =J(to, t)S(to)

(8.64)

where J(to,t) is the Jacobian of the mapping of areas of wavefronts normal to the same
group of rays, i.e.
;jknj(aX/aa) (axJap)
J (to, t) =
0
0
0
(8.65)
;jknj (aXj laa) (axklap)
where the superscript (0) denotes a quantity at the time to, and a and p are the parameters
of any wavefront. A point on the wavefront at the time t is located by the position
vector xj(a, P). In fact, we have already described this topic by Eqs. 1.100 and 1.101.
Substituting Eq. 8.64 into Eq. 8.63 yields

n=c -IJ- ldJ


'l'
dt
Consequently, Eq. 8.60 becomes
C"',ii

(8.66)

de)

=c-1(J-ldJ
dt - c -1 dt

(8.67a)

d 10g(Jlc)
dt

(8.67b)

=c

-1

Moreover, due to Eqs. 8.57c and 8.67b, we can show that

c 2 (A
A

.) =dlog(AJlc)
,j
dt

"',J

(8.67c)

Foundations oj Solid Mechanics

304

If all conditions are known at a wavefront for the time t, a subsequent wavefront
for t+M can be located easily by Eq. 8.51c, and rays normal to the subsequent
wavefront can be detennined by solving Eq. 8.59a. Such a solution exists wherever
the wavefronts are smooth. Besides its location, we may be interested also in magnitudes
of various jumps at the subsequent wavefront. Now we shall fonnulate equations
governing such magnitudes. We know that the displacement jump is always zero, i.e.
[ui ] =0
(8.68)

The derivative of any jump can be derived as follows

'Xi

'Xi

-a [{(Xi' t)] =-a [{(Xi' '1')]

=[a~i f(xi,

'I')r -[a~i 'I')r

=if +/

.)+ - if,J +/,'T,}


llf .)-

,/

'If

f(x i,

,IT,}

= [f)+ [f.,]'I',i

(8.69)

or, due to Eq. 8.57a,

a~. [((xi,t)] =[f)+ [f.,]nic-

(8.70)

Differentiating Eq. 8.68 with respect to xi' and using Eq. 8.69, we find
[ui)

=-[ui,,1'1',i =-[vi]'I',i =-Vi'l',i

(8.71)

where Vi denotes the velocity jump as a continuously differentiable function of Xi' i.e.
Vi(Xi ) =[Vi (Xi' '1')]

(8.72)

The kinematic jump obtained earlier as Eq. 8.58b is identical to Eq. 8.71. The derivative
of the latter equation with respect to Xk gives
and in particular this implies

[ui,ik] =-[Vi)'I',k - (Vi'l'),k

(8.73a)

[U.I, kk] =-[v.Ik]'"


k - (V.",
t,
' "k)k

(8.73b)

[Uk , ki] =-[Vk"k]'" i - (Vk'", k) ,I.

(8.73c)

Differentiating Eq. 8.72 with respect to Xi we find


Vi,i =[Vi) + [vi,,1'1',i

or
[Vi,,]'I',i

=Vi,i -

[Vi)

(8.74a)
(8.74b)

Multiplying the equation above by 'I',i and incorporating the Eiconal equation, we obtain
[Vi,,] =C2(Vi,i'l',i - [vi)'I')
(8.75)

305

Wave Propagation

Due to Eq. 8.74b, it can be shown that

[Vi)'I',i'l',j =Vi,j'l',i'l',j - [Vi,,]'I',j'l',i'l',j'


and

-[Vk,k]'I',i'l',i =-VU'l',i'l',i + [vk,t]'I',k'l',i'l',i'

Combining the last two equations, we get

=Vi,j'l',i'l',j - Vk,k'l',i'l',i
8.74a yields for i =j,
[v .. ]\1(
=V . . - [v . . ]
,J

[Vi)'I',i'l',j Substituting Eq. 8.75 into Eq.

[Vk,k1'l',i'l',i

c 2(V-ltl.llf
.llf . 't' ,I 'Y

.llf .)

I,} T,I T"

1,1

','

(8.76)
(8.77)

With the new definition of the velocity jump, we can write the dynamic and
kinematic jumps (Eqs. 8.58), respectively, as
(8.78a)
[CJij]nj =-cpVi

lUi) =-c-1Vinj

(8.78b)

and the dilatation jump is


(8.78c)
The derivation so far is based on a nonstationary surface of discontinuity, i.e.
eIl,t:: O. If eIl,t =0, the surface is fixed in space, i.e. stationary, then Eq. 8.47b becomes
Vi=O
(8.79)
and, accordingly, the jump conditions Eqs. 8.78a and 8.47a become

[CJij]nj =0

(8.80a)

lUi) =CJ.inj

(8.80b)

and
(8.80c)
for some

CJ.i :: O.

8.3 VELOCITY JUMPS IN ISOTROPIC ELASTIC DOMAINS


While Section 8.2 is independent of any particular constitutive laws, the present
section is restricted to isotropic domains which may be inhomogeneous but continuous,
i.e. the mass density p and the elastic constants A. and /.l are continuous and diferentiable
functions of Xi' Due to the kinematic jump conditions, we can write the cOIistitutive
relationship at the wavefront in the form
(8.81a)
[CJij] =-AOijVknkc-1- /.lVinjc- 1- /.lVjnic- 1
or, due to the dynamic jump condition,

[(c 2 - c;)Oij - (c: - c;)nin) Vj =0

(8.81b)

Foundations of Solid Mechanics

306

where cp and c. are as defined in Eqs. 8.12 and 8.14, respectively. The equation above
is an eigenequation with c 2 and Vj being eigenvalues and eigenvectors, respectively.
There are three eigensolutions. The first solution can be easily identified by observation,
i.e.
(8.82a)

an
V.=_J
J
c
p

=a 'I' .

(8.82b)

.J

where a is an arbitrary scalar. The second solution should have the eigenvector orthogonal to the first one (Eq. 8.82b), i.e.
(8.83a)
Vj'l',j =0
which makes Eq. 8.81b yield the eigenvalue as
(8.83b)
The third eigenvalue can be shown to be a duplicate of the second one. In case of Eqs.
8.82, the velocity jump is normal to the wavefronts, thus it is called a longitudinal or
compressive discontinuity and the speed cp is just that of P-waves, Eq. 8.12. In case
of Eqs. 8.83, the velocity jump is tangential to the surface of discontinuity, thus it is
called a transverse or shear discontinuity and the speed c. is just that of the S-waves
(Eq. 8.14). Moreover, the latter case also has a zero jump in dilatation (Eq. 8.78c)
which is another common feature in the classical theory of elastic S-wave propagation.
The equations of motion, Eq. 2.14, may be written in terms of displacements for
inhomogeneous isotropic elastic solids in the form
Vi,l =( CP- c. Uk,lei + c. Ui,kk +
2

2)

A,i
!l,k (
)
P
Uk,k + P Ui,k + Uk,i

(8.84)

Substituting Eqs. 8.71, 8.73 and 8.75 into the equation above yields
C2(Vi,j'l',j - [vi)'I')+ (c: - c;) [[Vk,k)'I',i + (Vk'l',k)
+c;[[Vi kl'l' k + (Vi'l' k)
"

1 + A,i Vk'l' k + !l,k


"k'
P
, P

(Vi 'I',k + Vk'l',i) =0

For P-waves, the problem is characterized by Eqs. 8.82, thus


Vi,j = a,j'l',i + a'l',ii
Setting
we find

(8.85)

(8.86)

c =c; in Eq. 8.85 and multiplying the result by 'I',i' then incorporating Eq. 8.77
2

Vk,k + C:(Vk'l',k),i'l',i + c;([Vi,k)'I',k - [Vk,k)'I'';)'I',i + c;[(Vi'l',k),k


-(Vk'l',k)'I',i +

~i (Vk'l',k)'I',i + ~k (Vi'l',k'l',i + Vk'l',i'l',;) =O.

Substituting Eq. 8.76 into the equation above, then incorporating Eq. 8.86, we get after
an extensive manipulation

Wave Propagation

307

a
'I',k a,k + 2p (P'l',k),k =0

(8.87)

Substituting Eqs. 8.57c and 8.67c into the equation above we fmd
da adlog(pJlc p ) 0
dt +
2dt
The solution to the equation above is

(pc~0 )"2

a =a o pCp r 1

(8.88)

(8.89)

where the superscript (0) denotes a quantity at an initial time to, and the terms without
any superscripts belong to a subsequent time t. Substituting Eq. 8.89 into Eq. 8.82b,
we can get the velocity jump at a subsequent wavefront as

0)112

V. =n. ( v?v?r1 pCp


, '"
pCp
For S-waves, the problem is characterized by Eqs. 8.83, thus
0

(8.90)

Vk,i'l',k =-Vk'l',~
Multiplying the equation above by 'I',i we find
1
Vk,i'l',k'l',i =-"2 Vk('I',i'l'),k'
or, due to the Eiconal equation, Eq. 8.56,

Vk,i'l',k'l',i =VkC;3 C.,k

(8.91)

2
Il,k
C.Vi
,''1',' + (2
cp - C.2) [VU]'!'i + C.2
(Vi'!')
k +-(Vi'l'
P "k + Vk'l' ,.) =0

(8.92)

In Eq. 8.85, setting

c =c; and incorporating Eq. 8.83a we fmd


2

,,

"

,,,

Multiplying the equation above by 'I',i' we can solve for [Vk,k] as

[Vk,k] = (C;: C;)(2C;Vi,i'l',i'l',i +

C;2~kVkJ

or, due to Eq. 8.91,

[Vk,k] = (C;: C;)(2C;l C.,kVk +

C;2~kVk)

(8.93)

Substituting Eq. 8.93 into Eq. 8.92, we can get


1
C.,k
'I' kVi k+-21.1 (Il'l' k) Yi--Vk'l'
i =0
c

(8.94)

Substituting Eqs. 8.57a, 8.57c and 8.67c into the equation above we fmd
dVi Vid 10g(J.Ll Ic.)
dt + 2dt
C.,kVkni =0

(8.95)

,,,,

Foundations of Solid Mechanics

308

In order to satisfy Eq. 8.83a, a general fonn of Vi is


Vi

=aNi + bBi

(8.96)

where Ni and Bi are unit normal and unit binonnal to the ray, respectively, and a and

b are scalars. Substituting Eq. 8.96 into Eq. 8.95, then incorporating Eq. 8.59b and the
Frenet-Serret fonnulas, Eqs 1.91 and 1.92, we get

d"( (d 10gULJ/c,) +lC't


. )
0
"(=

-+

dt
2dt
where 't is the torsion of the ray, and
"(=a +ib

(8.97)

(8.98)

in which i is the imaginary quantity, i.e. = -1. Note that


The solution to Eq. 8.97 can be shown to be

1"(I

0 _1)112

"(=111 ( c.~J

c~~

where
a=ao-

ei/j

and

1Vi 1 are identical.


(8.99)

(8.100)

c,'tdt'

Substituting Eq. 8.99 into Eq. 8.98 we find the velocity jump at a subsequent wavefront
for propagating shear waves, Eq. 8.96, to be
o oc.~0 -I)112

Vi =(V k Vk -o-J

or, smce

~=

pc"

(Ni cos a+ Bi sm a)

(8.101a)

(Nicosa+Bisina)

(8.101b)

c,~

00 )112

Vi= (V:V: P c' rl


pc,

In fact, a O is also defmed by the equation above with t =to, i.e.,

tanaO VioB io
V~N~

(8.102)

JJ

For a homogeneous domain, the rays are straight, thus a wavefront at a time t can
be constructed by advancing from the wavefront at a time to a distance c (t - to) along
a ray. Such construction of wavefronts, known as the Huyghen's principle, is illustrated
in Fig. 8.8, in which r l and rz are measured along rays, and ro along radii perpendicular
to the 'axis of revolution'. Essential geometric relations involved are

rg =r: sincjl'

(8.103a)

ro

(8.103b)

=rz sin cjl'

r l - r~ = rz - r: = c(t - to)

(8.103c)

The terms r l and rz are also known as the principal radii of curvatures of wavefronts.

309

Wave Propagation

8"-,

\ ::0 Wavefront
Wavefront at

at t

to

'Axis of revolution' for


wavefront infinitesimal areas

Fig. 8.8 Construction of initial and subsequent wavefronts for


homogeneous medium.
Corresponding inftnitesimal areas on the initial and subsequent wavefronts are generated
by 'revolving' arcs A Bo and AB about the axis of revolution through a small angle de,
thus they can be put, respectively, as
(8.104a)
S(to) =7~7~sincll'dcll'de
Set)

=7172sincll'dcII'de

(8.104b)

Hence the Jacobian of area transformation (Eq. 8.64) becomes


7172
J(to,t)="""""iJij
7172

(8.105)

Moreover, it can be stated for the homogeneous case that pO = p, c~ = cp ' c~ = c., njo = nj,
Njo=Nj, Bjo=Bj, t=O and BO=B, thus Eqs 8.90 and 8.101 give

V.
7172
V.V=-V
II
7172 I I

(8.106)

This shows that the magnitude of a fteld variable at the wavefront decays as R-1 for
spherical waves (with radius R) and as 7- 112 for cylindrical waves (with radius 7).
8.4

REFLECfION AND TRANSMISSION AT INTERFACES AND


BOUNDARIES

In Sections 8.2 and 8.3, we have considered the propagation of waves in continuous

Foundations of Solid Mechanics

310

media. We shall now study the situation at an interface between two media with different properties. Let an interface S separating media A and B be represented by
Xi =Xi(a,~) where a and ~ are the parameters of the surface S. We assume that the
surface S is smooth, thus ax/iJa and ax/af3 are continuous in a and ~, and
Euk(aX/aa) (axkla~) O. At each point on S, the unique unit normal to S is given by

*"

Euk(aX/aa)

r. =

(axkla~)

[u,.,.ewiaxmlaa) (ax"la~) (axolaa) (axpfa~)] 112

In fact,

(S.107)

ax/aa and ax/a~ are two unique tangents to S. The positive direction of Ti is

chosen as pointing from medium A into medium B.


Any 'incident' signal in medium A is a linear combination of compressive and
shear signals. Thus we specify the incident family of wavefronts as \jIO)(Xi) =t, where
(Eq. S.56)

(S.lOS)
The unit normal to a wavefront, in the direction of propagation, of the signal is (Eq.

S.57a)

ni(O) =co~~)

(S.109)

The magnitude of the signal is specified by VlO)(Xi), the jump in particle velocity across
a wavefront. For an incident S-wave we must have
(OlT ,(0) - 0
(S.110a)
ni'v
CO=c
i -,
sA
while for an incident P-wave we have
ijkn;U>V!O) = 0,

Co

= CpA

(S.110b)

Here CsA and CpA are, respectively, the speeds of shear and compressive waves in medium

A.

When the incident wavefront or rays meet the surface S, we can no longer be
assured that only the original shear or compressive wave is present in medium A, so
must allow both types of waves to be present in both media. If we denote the wavefronts
of these waves by
'j!V)(xJ

then the wave functions

'jIv) and

.. ..{V>.. ..{V) _

=t,

(v

=1,2,3,4)

(S.111)

unit normals to the fronts nlv ) are determined by


-2

(v) _

.. ..{V)

(-

1 2 3 4)

'f',i'f',i-cv,n;-cv'l',i'v-",

(S.112)

The type of each of these waves is specified by requiring that


(I), ,(I)

~jl;nj 'VI;

=0

(S.113a)

(S.113b)
(S.113c)

Wave Propagation

311

nr>vr) = 0

c4 = CaB,

(8. 113d)

Thus V(l) and V(3) are the magnitudes of compressive signals in media A and B,
respectively, while 02) and V(4) are the magnitudes of shear signals in those respective
media. The waves for v = 1 to 2 in medium A are called reflected, and the waves for
v = 3 and 4 in medium B called transmitted.
At each instant of time the reflected and transmitted wavefronts emanate from the
intersection of the incident wavefront and S. We must then have, on such intersection
which is assumed sufficiently small,

'!I") =",,),

(v

=1,2,3,4)

and the derivatives of every 'If with respect to a. or

(8. 114a)

p must be equal, i.e .

.. .<~)aXi =, ..<?)aXi
'1'.' aa. '1'.' aa.

(8. 114b)

=...<?) axi

(8.114c)

.. .<~) aXi

ap

'1'"

'1'"

ap

Equations 8.114b and 8.114c are equivalent, respectively, to


(

ax

c"ni - cOni) aa. =0,


(0)

("A

(0)
("A aXi
(c"ni - Coni) ap =O.

Thus the vector in parentheses in the equations above must be in the direction Ti and
so there exist scalars a" such that

c"n~O)+a T.
n~,,)= co'
" .,

(v -1 2 3 4)

-",

(8.115)

The last equation implies that the five wave normals ni("), v = 0, 1, 2, 3, 4 and the
interface normal Ti are in the same plane, known as the plane of incidence. Introduce
a unit vector Ni , tangent to S in the plane of incidence, in such a way that the angle
eo between nlO) and Ti can be put in the form
ni(OTi = cos eo

(8. 116a)
(8. 116b)

with
1t

o:::;e o :::;'2
Similarly the angle ell between

(8.116c)

nl") and Ti can be put in the form


ni("Ti =cose"

(8. 117a)

n~")N = sine

(8. 117b)

..

"

Foundations 0/ Solid Mechanics

312

Taking the scalar product of each equation in Eq. 8.115 with Nj and Tj , we find that
(8.118a)
sin(60 -6V>
a =---y
sin 60
Since all Cy are greater than zero, Eq. 8.118a implies 0::;; 6y
the media in which each wave belongs, we can deduce

(8.118b)
::;;

1t. Further considering

~::;; 61,62 ::;;1t

(8. 119a)

1t
0::;; 63,6 4 ::;;'2

(8. 119b)

These angles and the relevant normal vectors are illustrated in Fig. 8.9. Equations 8.115
and 8.118 determine the normals nly ), in terms of the given incident wave normal nj(O),
and the surface normal Tj

N.
I

Medium A

Medium 8

Fig. 8.9 Plane of incidence.


Since CO=CI or c2 , then by Eq. 8.118a 6 1 or 62 =1t-60 meaning that the reflected
ray of the same type as the incident ray suffers 'equal-angle' reflection. If for any
Cy (v = 1,2,3,4) and incident angle 60 ,
(8.120)
then Eq. 8.118a implies that there can be wave (reflected or transmitted) corresponding
to that value of v. For example, if the incident signal is an S-wave then no reflected
P-wave is possible for any incident angle 60 if
1t
< 60 ::;;'2
(8.121)

Wave Propagation

313

Medium A

Medium B

Fig. 8.10 Plane of incidence being partitioned, by five wavefronts


interfaces S+ and S_, into seven 'angular' sectors S".

\jIv)

and

in which the critical angle E> is defined by


CsA

sinE>=-,

(8.122)

cpA

It is clear that a reflected shear wave is possible for all angles of incidence of a
compressive signal. The various possibilities for transmitted waves can be deduced
easily by the reader.
Let P be a point at (XiO,tO) where S and the wavefronts meet. A neighborhood of
P is divided into seven 'angular' sectors S" by the five wavefronts, and interface S
which may be considered as two parts, S+ and S_, corresponding to t > to and t < to,
respectively (see Fig. 8.10). In the interior of each such sector the deformation and
stress fields are continuous. Denote by vi") and dij) the limits of Vi and C5ij as (Xi' t) ~ P
while remaining interior to the sector S". Then in obvious notations;

(8.123)

V~+)=V~6)_V~O)

Summing the above we fmd the following identity


V~O) + V~l) + V~2) + V~+)
IIII

=V~3) + V~4) + V~-)


III

(8.124)

Foundations of Solid Mechanics

314

But a velocity jump at a stationary surface must vanish (Eq. 8.79), i.e.
V~+)
, V~-)
, 0

= =

(8.125)

hence the identity becomes


Z

Vj(Y)

y=o

= L Vj(Y)
4

(8.126)

y=3

In the same manner, it can be shown using Eq. 8.80a that


Z

y=o

[cr~~1T. = L [cr~~1T.
"

y=3

"

(8.127)

Note that [dijl] in the equation above can be put in terms of v;v> due to the constitutive
relationship at the wavefronts, Eq. 8.81a.
In addition to the orthonormals T j and N j , we may recall the binormal (Eq. 1.90),

Bj =Eij"T/V"
(8.128)
which is tangent to S but normal to the plane of incidence. Now all wave normals can
be represented in terms of these unit base vectors as
nj(Y) =cos eyTj + sin eflj + (O)Bj

(8.129)

Using this representation in Eqs. 8.113, we find that the most general form for the
velocity jumps at various wavefronts is

v?) =(Xl coselT j + (Xl sinelNj

(8.130a)

V j(2) = ~ sin eZT j - ~ cos e.JVj + PzHj

(8. 130b)

VP)

=~ cos e3Tj + ~ sin epj

(8.130c)

V?) = (X4 sin e 4Tj - (X4 cos e/Vj + P4Bj

(8.13Od)

and for the incident signal (Eqs. 8.110), which is assumed known, is
Vj(O) =a.ocoseOTj

+ a.osin epj'

(co =cpA)

(8.130e)

or
(8.1300
At this stage one may note that the task is to determine six unknowns (Xl> {lz, ~,
(X4' Pz and P4 from six linear equations, Eqs. 8.126 and 8.127. The task can be organized
by rewriting Eqs. 8.129 and 8.130 as
(8.131)
VjM =bylTj + by.JVj

+ bv/J j

(8.132)

in which aYj and bYj are as defmed in Table 8.1. Accordingly and due to Eq.8.81a, we
can write the following
(8.133)
where

Wave Propagation

315

Table S.l
O-vj

j = 1

j=2

j=3

cos eo

sin eo

1
2
3
4

cose l
cose2
cose3
cose4

sinel
sine2
sine3
sine4

0
0
0
0

byj

j = 1

j=2

j=3

Cy

~coseo

~sineo

CpA

~sineo

Po

CsA

CpA

~sine2

-<Xocoseo
(Xl sinel
--<l:z cos e 2

~COSe3

~sine3

CpB

~sine4

~COSe4

P4

C.B

v=

o (P-wave)
o (S-wave)
v=

1
2
3
4

(Xl cose l

P2

CsA

(S.134a)
(S.134b)
(S.134c)
in which it should be understood that A= AA and Il = IlA for v = 0, 1,2; and A=AB and
Il =IlB for v = 3,4. Equations S.126 and S.127 become, respectively
b lj + b2j - b3j - b4j =-boj
(S.135a)

Tlj + T2j - T3j - T4j =-Toj


Since b13 = b33 = 0 and ay 3= 0 for all v, Eqs. S.135a and S.135b for j
respective forms as
b23 - b43 =-b03

T23 -T43 =-T03

(S.135b)

=3 assume simple
(S.136a)
(S.136b)

while other four equations remain as

blj+b2j-b3j-b4j=-boj,

U=1,2)

(S.136c)

T lj + T2j - T3j - T4j =-Toj ,

U = 1,2)

(S.136d)

Foundations oj Solid Mechanics

316

Equations 8.136a and 8.136b yield the values of ~ and f34' which exist only for an
incident S-wave, as
A. _ A. [(~A/co)cos60-(~B/C4)cos6.J
....2- ....0
L\
A.

....4 = f30

[(~A/co) cos 60- (~A/cz) cos 6:J

L\

(8.137a)
(8. 137b)

where
~B

~A

L\E-cos6 --cos6
C4
4 C2
2

(8.138)

In the meantime, Eqs. 8.136c and 8.136d form a set of four linear equations for
determining <Xl to <X4 as
sin 62
-cos 63
cos 6 1
sin 6 1
-cos62
-sin63
- PACI cos 262 -PAc2sin262 PB C3 cos 264
~A 26
--sm
Cl
1

~B 263
-sm
C3

Here Kv are given for an incident P-wave (co = Cb 60= 1t - 6 1) as


K2=-sin6l

and for an incident S-wave (co = C2, 60 = 1t -

K4 =_~A sin 261


Cl
6z) as

(8. 139a,b)
(8. 139c,d)

(8. 140a,b)
(8. 14Oc,d)
K3 =-PAc2sin262 , K4=-PAC2cos262
These results agree formally with the appropriate special cases in the book by
Ewing, Jardetzky and Press (1957) where the case of plane stress waves incident on
plane interfaces is treated.
We shall now consider reflection at boundaries. The bounding surface of a medium
can be considered as a special case of an interface. Now, however, there are no
transmitted signals and specific conditions are to be imposed on the displacements and/or
stresses at the surface. These conditions serve to replace the jumps specified on an
interface.
If all three displacement components are specified at the boundary,
vj(O)

vP) = prescribed quantity,

then Eqs. 8.123 give


(8.141)

317

Wave Propagation

It follows that for an incident P-wave with


cos(92- 9 0)
(Xl = <Xo cos(92+ 9 0)'

= Cl and 90 = 1t - 9 1,

sin 290

<Xz = <Xo cos(92+ 90)'

Similarly for an incident shear wave, with


sin 290
(Xl = -<Xo cos(9 l + 9 0)'

Co

Co

(8. 142a)

~2 = 0

= C2 and 9 0 = 1t - 9 2, we fmd that

cos(9 l - 9 0)
<Xz = <Xo cos(9 l + 90)'

~2 = -~o

(8. 142b)

If all three stress vector components are specified at the boundary,

(j~Tj = ~JTj =prescribed quantity.


then
2

v~o ldij1Tj =O.


It follows that, for an incident P-wave,
(X

ci

(C;COS2292+ sin 292sin 290)


1 -<Xo Cl2cos2292-C22
. 290 '
sm 292sm
2C1C2COS 292sin 290

<Xz=<Xo Cl2cos2292-C22'
. 290 '
sm 292sm
(8. 143a)

~=O

and for an incident shear wave,


)
2C1C2COS 290 sin 290
al=-<Xo( 2 2
2'
.
,
Cl cos 290 - C2 sm29 l sm290

_ (-c;
<Xz - <Xo

ci

cos2290 + sin 291sin 290)


. 290 '
Cl2cos 2290 - C22'
sm 291sm
(8.143b)

If the normal stress component and two tangential displacement components are
specified on the boundary,

d:TjTj = cr:TjTj = prescribed quantity,


vj(OlNj = v?lNj = prescribed quantity,
v~OlB.
,II,

then

=v~lB. =prescribed quantity

Foundations of Solid Mechanics

318

It follows that for an incident P-wave,


(Xl

=-<Xo,

~=o,

Pz=O

(8. 144a)

Pz = -/30

(8. 144b)

and for an incident shear wave,

<<It = 0,

~ = 0,

Finally, if the nonnal displacement component and the two tangential stress
components are specified,
(Ok.
(3k.
bed
.
Vi "1 i
Vi "1 i prescn
quantity,

= =
cf:>r/Vi =cfJ>r/Vi =prescribed quantity,
ct:>rIJi =cfJ>rIJi =prescribed quantity,

then

It follows that for an incident P-wave,

<<It =0,

~=O,

Pz=O

(8.14Sa)

Pz = -/30

(8. 14Sb)

and for an incident shear signal,

<<It = 0,

~ = -<Xo,

8.5 WAYES IN ISOTROPIC VISCOELASTIC MEDIA


The main purpose of this section is to modify the subject matters of Sections 8.3
and 8.4 to account for isotropic viscoelastic media. For a material point that has been
disturbed for a duration t, the viscoelastic constitutive law, Eq. 7.102, can be written
explicitly as
0' ..

II

=8..AekA;0 + 2~0 + 8 ..
II

II 0+

A(t - S )dkA;(s ) + 2

0+

Il(t - s )dev(s )

(8.146)

in which the subscript (0) denotes t O. Implicitly, it has been assumed in Eq. 8.146
that the functions A(t) and Il(t) are not singular at t O. Thus this equation does not
hold if the constitutive relationship is characterized by a mechanical model which
contains a 'free' dashpot in parallel with other mechanical elements. In other words,
the case which is characterized by a generalized Maxwell model (Fig. 7.5) with '110 0
is not applicable to Eq. 8.146 nor the derivation to follow. For a point at a wavefront,
t in Eq. 8.146 is zero since the material there has been disturbed for a zero duration.
Then the same equation gives a relation among jumps as follows
[O'ji] 8jiA,,[kA;] + 21lo[ev]
(8.147)

or, due to Eqs. 8.78b and c,

319

Wave Propagation

[<Jij] -5ijAuV"n"C- 1- IloVinj C- 1- IloVj nic-1

(8.148)

which is the same as Eq. 8.81a if initial values of Lame's functions are used there.
Thus like isotropic elastic solids, there can be two types of waves propagating in isotropic viscoelastic solids, i.e. the P-waves characterized by Eqs. 8.82 and S-waves by
Eqs. 8.83, while the wave speeds are
C __ ~~+21lo

C.

(8.149)

=...JJ.1oIp

(8.150)

Recalling the Leibnitz's rule, Eq. 1.137, the time derivative ofEq. 8.146 is obtained
as

(8.151)
where the prime () denotes a derivative with respect to t. At a wavefront, t in Eq.
8.151 is zero, so the same equation gives a relation among jumps as
[<Jij"l =~ij~[Ekk] + 2~[~) + ~ijAu[Ekk"l + 21lo[Eij,l]

The derivative of Eq. 8.147 with respect to


of jumps (Eq. 8.70), can be put as

Xm ,

(8.152)

in view of the formula for derivatives

[<Jij,m] + [<Jij"lnmc- 1 ~ijAu[Ekk,m] + ~ijAu[Ekk"lnmc-1 + ~ijAu,m[Ekk]

+21lo[Eij,m] + 21lo[Eij,,]n.,c -1 + 2Ilo,.,[Eij]

(8.153)

Substituting Eq. 8.152 into Eq. 8.153 we find


[<Jij,m]

=~ijAu[Ekk,m] + 21lo[Eij,.,] + ~ijAu,.,[Ekk] + 2Ilo,m[~j]


- ~ij~[Ekk]nmC -1 _ 2J.L'[Eij]nmc- 1

(8.154)

Setting m equal to j in the last equation yields


[<J",J
...]

='''''0
'} [E.....] + 21ruL(P ...] + '''0"
'} .[E....] + 211 .. [E.. ]
Mo,'

~J.J

Mo

M,}

'J

(8.155)
Due to the equation of motion, <Jij,j =PVi", and the strain definition, Ev =(Ui,j + uj,i)/2,
the equation above can be rewritten as
[Vi ,]
,

=(C 2 p

C;)[U" ti] + C;[Ui ] + Au,i


]+ Ilo,,,
P [U""",
P lUi + u. ,.]
L

"M>

,A

A,

(8.156)

Foundations of Solid Mechanics

320

Substituting Eqs. 8.71, 8.73 and 8.75 into this equation we find
CZ(ViJ"'.i - [Vi)"') + (C: - C;) [[Vk,k1"',i + (Vk"',k)

+C;[[Vi,k1'l',k + (Vi"',k),J + ~i VN,k + ~k (Vi"',k + Vk"'.;>


(8.157)
The resemblance of the last two equations with those for elastic solids (Eqs. 8.84 and
8.85) should be noted.
Proceeding in the same manner as already used to obtain subsequent jump magnitudes for elastic solids (Eqs. 8.90 and 8.101), we can get
Vi

=V; exp( ~it Odt')

(8.158)

in which vt is the velocity jump for elastic solids (Eq. 8.90 or Eq. 8.101b depending
on the wave type), and
n

=~+2~
Ao+2~'

=-,
~

for P-waves

(8. 159a)

for S-waves

(8. 159b)

Normally, n is negative, hence the decay of the velocity jump as the propagation
proceeds is more than that in elastic solids by an exponential factor. For a homogeneous
domain, it can be shown that

TITz
.. 0
VV.
=-exp[n(t
- to)] v,I IV.
II
TITz

(8.160)

in which the superscript (0) denotes a quantity at the time to. and T .. Tz are principal
radii of wavefronts as depicted in Fig. 8.8 and defmed in Eqs. 8.103.
For the reflection and transmission at interfaces and boundaries, the derivation in
Section 8.4 is also applicable to isotropic viscoelastic solids if initial (t =0) values of
A. and J.1 are used instead in all equations involved.
8.6 IN-PLANE HARMONIC SURFACE WAVES

In this section, we shall consider the harmonic vibration of isotropic elastic planes
of the same geometry as Section 6.2 where statics of planes with ~ < x < 00 was
considered. Knowing that the solution of a plane stress problem with a Poisson's ratio
equal to v/(1- v) can be obtained from its corresponding plane strain problem with a
Poisson's ratio v, we shall derive in this section only for plane strain cases. The
equations of motion, in terms of displacements and in the absence of the real body

Wave Propagation

321

force, can be specialized from Eq. 6.1 by taking Xj there as the D' Alembert's force.
The resulting equations in explicit symbols are
ifu
ifu
ifv
ifu
(A. + 21l) ox 2 + Il oy2 + (A. + Il) oxoy =p ot 2
(8.161a)
ifv

ifv

ifu

ifv

(A. + 21l) oy2 + IlOX2 + (A. + Il) oxoy =p ot 2

(8.161b)

where p is the mass density. For a 'discrete' harmonic wave which is outgoing in the
x -direction parallel to the surfaces and the stratification of the plane, the displacement
functions can be put in the form
u(x,y,t) =u(y)el(OlI-kz)
(8. 162a)
v(x,y, t) = v(y )ei(OlI-kz)

(8. 162b)
where ro and k are angular frequency and wavenumber, respectively. The derivation
to follow is mainly concerned with the determination of the wavenumber while there
is no applied traction in the plane. In other words, we are looking for what is commonly
known as a harmonic surface wave of Rayleigh type (J. W. Strutt, Lord Rayleigh, 1887).
Substituting Eqs. 8.162 into Eqs. 8.161 we find
_ 112 u
11;

+~ d 2u _i11(~-~)dV =-u
11~ dy2

11; 11~ dy

(8.163a)
(8. 163b)

Here a constant a of length dimension is used to nondimensionalize the independent


variables x and y; 11 is the dimensionless surface wavenumber; 11p and 11. are dimensionless frequencies of P and S-waves, respectively, i.e.
aro
aro
11 =ak, 11p =-, 11 = (8. 164a,b,c)
cp
c.

We call 11p and 11. dimensionless wavenumbers of body waves. The solution to Eqs.
8.163 takes the form
u(y)=AeUY ,

v(y)=Beay

(8. 165a,b)

Substituting Eqs. 8.165 into Eqs. 8.163 we find


(8.166)
A non-trivial solution of this system of equations exists if the determinant vanishes,
leading to the condition
where

Foundations of Solid Mechanics

322

(8.167a,b)
Thus
(8.168)
For each a, the eigenvector involved can be detennined from Eq. 8.166, and Eqs. 8.165
become
u(y)=Ae

"""-y
P

"'-y

+Ce

-a y a y

+Be ' +De'

iap ("",,-y
",-y) i1'\ ( -a JI
a JI)
v(Y)=--lAe P -Ce P --lBe ' -De'
1'\
a.

(8. 169a)
(8. 169b)

where A, B, C and D are arbitrary constants. If 1'\p < 1'\ < 1'\., it is more convenient to
write the equations above as
-pJl
a JI
u(y) =Ae
+Ce P +B cosa.y +D sina.y
(8. 170a)
ia(
)i1'\
v(y) =--\Ae-PJl _CeI1>JI -=-(D cosa.y -B sina.y)
1'\
a.

(8. 170b)

where
(8.171)
If 1'\ < 1'\p < 1'\., the solution fonn should be modified further for the sake of convenience
as
(8. 172a)
u (y) =A cos (lpy + C sin (lpy + B cos (laY + D sin (l.y
iap
i1'\
v(y) =-(C cosapy -A sinapy)-=-(D cosa.y -B sina.y)
1'\
a.

(8. 172b)

where
(8.173)
The arbitrary constants A, B, C and D in the various solutions are to be detennined
from boundary conditions in the same manner as described for statics in Section 6.2.
However, the present case is an eigenvalue problem, since we are investigating the
possible existence of harmonic vibration without any applied traction. For a plane of
N layers each having a finite dimension in the y-direction, the fonnulation leads to a
set of 4N linear homogeneous equations in 4N unknown arbitrary constants; Al to AN'
BI to BN, C1 to CN, and DI to D N. The eigenvalue 1'\ of this problem is obtained in the
usual way, i.e. by setting the detenninant of the set of equations to zero, leading to the
following equation
(8.174)
Roots of the equation above are commonly known as dimensionless Rayleigh wavenumbers and will be denoted by 1'\R.

323

M'avePropaganon

The simplest Rayleigh wave problem is probably the case of a homogeneous half
plane with -00 < x < 00, 0 S Y < 00. The most convenient solution fann is given by Eqs.
S.169, in which C and D must vanish for boundedness of the solution of the wave as
y tends to infinity. The boundary conditions for this case are
(S.175)
G,,(x,O) =0, G..,(x,O) =0
which leads to the following set of homogeneous equations
(S.176)
Consequently, the famous Rayleigh wave equation for a homogeneous isotropic elastic
half plane is obtained as
FR(Tl) == (2Tl2_Tl~)2 -4Tl2a p (Tl)a.(Tl) =0
(S.177)
The equation above can be further nondimensionalized by introducing
(S.17Sa)

Tlp c. _ fl=2V
y= Tl. = cp = -" 2<1=V)

(S.17Sb)

1.15
1.14 r-....
1.13
1.12
I. I I

11

0.05
0.1
0.15
0.2
0.25
0.3
113
0.4
0.45
0.5

(1.10

~1.09

"-

1.08

1.07

1.06

1.05
1.04

0.1

0.2
Poisson

0.3
S

Ratio

'?RI,?s
1.144123
1.131613
1.119688
1.108378
1.0977
1.087664
1.078269
1.072357
1.061351
1.053786
1.046778

0.4

tJ

Fig. S.ll Rayleigh wavenumber for homogeneous half plane.

'"'"
0.5

Foundations of Solid Mechanics

324

where 1lR is the root of Eq. 8.177, i.e. the dimensionless Rayleigh wavenumber for a
homogeneous isotropic elastic half plane. Substituting Eqs. 8.178 into Eq. 8.177 we
find
(8.179)
Due to the absence of the frequency ro in Eq. 8.179, we can say that the Rayleigh wave
in this case is nondispersive. Rationalizing Eq. 8.179 by squaring we find

16(1-i)s6 - 8(3 - 2i)S4 + 8s 2 -1 =0


(8.180)
which is a cubic equation in S2, with three roots. However, only one of these satisfies

the Oliginal equation, Eq. 8.179, while the other two values are spurious, arising from
the squaring. Moreover, since the wave must not be incoming as y tends to infinity,
the correct value of S2 must be a positive real quantity and greater than unity, so that

1lp < 11. < 1lR

(8.181a)

cR < C. < cp

(8.181b)

where CR is the Rayleigh wave velocity =aro/1lR' Results for various values of the
Poisson's ratio v are presented in Fig. 8.11.
Next to be considered is the occurrence of another special type of Rayleigh waves.
The material domain in this case consists of two different half planes, perfectly bonded
together as shown in Fig. 8.12. The displacement functions given for a general case
should be modified to suit the present case as
(8.182a)
(8. 182b)

I----rx
o
y

Fig. 8.12 Two different half planes perfectly bonded together.

325

Wave Propagation

(8.182c)
(8. 182d)
in which the subscript 1 denotes the material domain where --00 < Y S 0, and the subscript
2 for 0 S Y < co. The boundary conditions are

u1(0)
O'YJll(X,O)

=Uz(O),

=O'YJl2(X, 0),

v1(0) =vZ<O)

O'..,l(X,O)

(8.183a,b)

=0'..,2(X, 0)

(8. 183c,d)

Accordingly, the Rayleigh wave equation can be obtained in the form


FR (11) == 11

2 2 R 2
R 2 2
2
2 R, 2 2
(11 - 1-'111.1- I-'211.J +11 (lpl(lp2(l.1(l.2 - (lpl(lal(11 - pz11.J

(8. 184a)

Or
S2(S2 -

/31/33 -

-rS> (S2 - /33)(S2 -1)] 112


~)] 112 (S2 _ f3z}2 _ [(S2 _ rS> (S2 _ 1)] 112 (S2 - /31~)2

/3i +S2[(S2 - /33y')(S2

_[(S2 _ ~y.) (S2 -

+/31/32/33{[(S2 - /33y')(S2 -1)] 112 + [(S2

-rS> (S2 _ ~)] 112} =0

(8.184b)

in which the dimensionless quantities defined by Eqs. 8.164 and 8.178b have been used
in the process, and
(8. 185a,b)
(8.185c,d,e)
Observing the absence of the frequency in Eq. 8.184b, we can say that this particular
surface wave is nondispersive. Stoneley (1924) made a thorough investigation on the
subject, thus gained the recognition in having it named as the Stoneley wave. Koppe
(1948) solved Eq. 8.184b numerically and concluded that
11.1,11.2 < 11R

< 11Ri

(8.186)

where 11Ri is the greater Rayleigh wavenumber among those obtained separately (by
solving Eq. 8.177) for each of the constituent homogeneous half planes. Other analyses
were given by Love (1911, 1926), Sezawa and Kanai (1939), Scholte (1947) and
Cagniard (1962). There can be at most one Stoneley wave. While Love showed that
this wave could exist under the stringent condition that the S-wave velocities of the two
constituent half planes were nearly equal, Cagniard found a more general condition for
such existence. Without loss of generality, that condition (by Cagniard) can be restricted

Foundations of Solid Mechanics

326

to the case where i3:J < 1, then put in the following fonn
-(/33 + /34 - 2i + (/34 - 2)2[(1-/33i)(I-/33)] lIZ + /33/3i(1-Yz)(I-/33)]11Z < 0

(8.187)

Equation 8.187 should serve as a quick check for the existence, while Eq. 8.186 as a
good guide for quick computation of the Stoneley wavenumber from Eq. 8.184b.
8.7 ANTIPLANE HARMONIC SURFACE WAVES
Now consider the possible existence of harmonic surface waves in antiplane
problems; these are commonly known as Love waves. We suppose the plane is infInitely
extended in the x-direction; displacement components are independent of z; u and v
vanish; waves propagate in x-direction; and the equation of motion is

C;V2W =~~

(8.188)

This equation has Eq. 6.9 as its statical counterpart, and has a solution of the fonn
w(x,y ,t)

=w(y

)ei(0lI-1JX)

(8.189)

where 11 is the dimensionless wavenumber as defmed in Eq. 8.164a, and x and y have
been nondimensionalized by a constant a of length dimension. Proceeding in the same
manner as the in-plane problem in the preceding section, we fInd
w(y)=Be-a,)/ +Dea.,)/

(8.190)

where a.(11) is as defIned in Eq. 8.167b. If 11 < 11.. it is more convenient to rewrite Eq.
8.190 as
w(y)

=B cos a.y + D sin a.y

(8.191)

where a.(11) is as defmed in Eq. 8.171. In either fonn, the constants B and D are to
be detennined from boundary conditions, as in Section 8.6. For a plane of N layers
each having a fInite dimension in the y-direction, the fonnulation leads to a set of 2N
homogeneous equations in 2N unknown arbitrary constants; Bl to BN , and Dl to DN
Setting the detenninant of that set of equations to zero, we fInd the Love wave equation

FL (11) =0

(8.192)

Roots of the equation above are commonly known as dimensionless Love wave numbers
and will be denoted by 11L'
It can be shown that there is no Love wave in a homogeneous half plane, nor in
a full plane consisting of two different half planes as shown in Fig. 8.12. The simplest
Love wave problem is probably the case of a layered half plane as shown in Fig. 8.13.
Employing a subscript 1 for the upper layer where -h/a S y SO, and subscript 2 for the
underlying half plane where 0 S Y < 00; we fmd that the solution assumes the following
fonn
(8. 193a)

327

Wave Propagation

Fig. S.13 Layered half plane.


(S.193b)
The boundary conditions are
wl(O)

=w2(0),

O'yrl(X, 0)

O'yrl(x,-hla)

=O'yr2(X,0)

=0

(S.194a,b)
(S.194c)

The first boundary condition, Eq. S.194a, readily gives


(S.195a)

B2=Bl

Then the remaining boundary conditions lead to

1lz<l.2

fll<l.l

a81 sin(asih1a) a. cos(adhla)

J{Bl} {O}

Dl

=0

(S.195b)

Consequently, the Love equation for this case is

1lz<l.2
fll<l.l

FL (11) == tan(<lsih1a)--=-=0

(S.196)

Since the equation above cannot be nondimensionalized further to be completely


independent of both 1181 and 11.2' its roots are dependent upon the frequency co, thus
surface waves in this case are dispersive. Equation S.196 can have several roots, but
all of them must be real and positive and fall within the range

11.2 < 11L < 11.1' or c. l < cL < c. 2

(S.197a,b)

where CL is the Love wave velocity =aco/11L' This shows that there is no Love wave
in a layered half plane if the shear wave velocity in the upper layer is greater than that

Foundations of Solid Mechanics

328

60
55
50
45

40

35

'7L 30
h/a = I

25
20
15

~2/CSI = 1.297
u.

'2

'1

=2.159
3 rd. mode

10

2nd. mOdi
1st. mode

/1-1-

0.4 0.8 1.2 1.6

'7S2
(a) First six modes of vibration.

9~~----------~----~1

h/a = I
= 1.297

'I'-

CS2 /C SI

P-2 /P- 1 = 2.159

II

-3
-6
-9+-~.-.-.-.-.-.-.-.-~.-.-.-~

10

10.4

10.8

11.2

11.6

12

12.4 12.8

'7
(b) Example plotting of FL (1\) v.S 1\.
Fig. 8.14 Love wavenumbers for a layered half plane.

329

Wave Propagation

in the underlying half plane. In fact. the Love wave function FL(TI) in this case is
discontinuous and singular, if
ao1h1a = n1tl2, (n = 1,3,5, ... ),
or

[FL(TI)=oo, n=1,3,5, ... ]

TI=[TI!1-(n2':JJI2,

(8.198)

Conditions stipulated by Eqs. 8.197a and 8.198 should serve as good guides for quick
computation of values of TIL from Eq. 8.196. An example problem has been solved,
and its Love wavenumbers of the fIrst six modes of vibration are presented in Fig. 8.14.
The higher the mode, the closer the derivative of FL(TI) with respect to TI tends to
infInity. Extensive quantitative information on the phase and group velocities of this
type of Love wave can be seen in the book by Ewing, Jardetzky and Press (1957).
8.8 VIBRATION OF MULTILAYERED SPACES
In this section, we shall consider three-dimensional problems of harmonic vibration
of isotropic elastic spaces of the same geometry as Section 6.3 where statics of spaces
with 0 ~ r < 00 is considered. The displacement functions of the present case take the
form
u,(r, a,z ,t)
Us(r,a,z,t)

=u,(r ,a,z)eitol
=Us(r,a,z)e lliJl

u,(r,a,z,t) = u.(r,a,z)e

(8.199a)
(8. 199b)
(8. 199c)

lliJl

The equations of motion, in terms of the displacements and in the absence of real body
force, are given by Eqs. 3.67, namely

(A. + Il) or (kk)

1a

(A. + Il);:- oa (kk)

+ ~V U, -

oeUe) =-ro2pu,

u, 2 0
r2 - r2

..r

VZ 2 au, Us)
2
+ ~ Us + r2 oe - r2 =-ro PUs

(I\. + Il) OZ (kk)


'1

+ Ilv2u, =-ro2PUr

(8.200a)
(8.200b)
(8.200c)

where kk(r,a,Z) is the shape function of the dilatation, i.e. Eq. 2.72. Harkrider (1964)
introduced potentials ~, 'I' and X so that
-1

a u,(r,a,z)
-1

a Us(r, a,z)

OZ'P lOx

= or +oroz +;:-oa
1 ~ 1 OZ'P

Ox

=;:- oa +;:- ozoe - or

(8.201a)
(8.201b)

330

Foundations of Solid Mechanics

~ V2
(fqI
'1'+ az 2

-I

a u.(r,e,z) =az-

(8.201c)

and so transformed Eqs. 8.200 into


V2<I> = -1'\!<I>

(8.202a)

V2'P=-1'\~

(8.202b)

V 2X=-1'\!X

(8.202c)

here a constant a of length dimension has been used to nondimensionalize the spatial
variables r and z, and 1'\p and 1'\. are dimensionless frequencies or body wavenumbers
as defined in Eqs. 8.164b and c. Using Hooke's law, we can express the stress components in terms of these potentials as
()2<I> a3'P 2a'P 21'\! -1'\~ )
(8.203a)
cr (r,e,z) = 21. az2 + az3 +1'\. az+
2
<I>

2 a2<I> 2 a 3'P 1'\~ a'P ()2X)


---+---+----r azae r az 2ae r ae araz

(8.203b)

()2<I>
a3'P
2a'P 1 ()2X )
crrz(r,e,z)=~2azar +2az2ar +1'\. ar +;:-azae

(8.203c)

a (r
9z

e z) ={
,

.I

The other three stress components can also be expressed in terms of these potentials.
As in the elastostatics of layered spaces in Chapter VI, the potentials can be expressed
in the form of Fourier series with respect to e and Hankel transforms with respect to
r as
(8.204a)
'P(r,e,z)=Lcosme
m

x(r,e,z) = Lsinme
m

Jor-'(z,1'\)Jm(T\f)d1'\

(8.204b)

Jor- X(z, 1'\)Jm(1'\r)d1'\

(8.204c)

Substituting Eqs. 8.204 into Eqs. 8.201 and 8.203 we find

Jor- uAr,Z,1'\)d1'\

(8.205a)

r- u (r,z, 1'\)d1'\

(8.205b)

r- u.(r,Z,1'\)d1'\

(8.205c)

Jor- cr.z(r,z, 1'\)d1'\

(8.205d)

a-1ur(r,e,z)=Lcosme
m

a-1ue(r,e,z) = Lsinme

Jo

a-1u.(r,e,z) = Lcosme

Jo

cr.. (r,e,z) = Lcosme


m

Wave Propagation

331

O'o.(r, e, z) = Lsinme
m

r- o-o.(r, z, 1'\)d1'\

Jo

(S.205e)
(S.205f)

where

~
[ d lm(1'\r)
Jm(1'\r)]
u,(r,z,1'\)=1'\ UR(Z,1'\) d(1'\r) +mvL(z,1'\)--;V:-

(S.206a)

~
[
Jm(1'\r)
dlm(1'\r)]
u a(r,z,1'\) =1'\ -mu R(z,1'\)--;V:--V L(Z,1'\) d(1'\r)

(S.206b)

u.(r, z, 1'\) = WR(Z, 1'\)Jm(1'\r)

(S.206c)

o-zz(r,z, 1'\) = O'R(Z, 1'\)Jm(1'\r)

(S.206d)

[
Jm(1'\r)
dlm(nr)]
O'o.(r, z, 1'\) = 1'\ -m'tiz, 1'\)--;V:-- 'tL(Z, 1'\) d(1'\r)

(S.206e)

[ d lm(1'\r)
Jm(nr)]
O',..(r,z,1'\) =1'\ 'tR(Z,1'\) d(nr) +m'tL(z,1'\)--;V:-

(S.206f)

- d'P
uR(z,1'\)=cI>+&

(S.207a)

deb d2qt
2WR(Z,1'\) = dz + dz2 + 1'\.'

(S.207b)

in which

d 2eb d 3'P
2d'P
2 2 eb]
O'R(Z, 1'\) = 2~ dz2 + dz3 +1'\. &+ (21'\p -1'\')2

2-)

(S.207c)

deb
d2qt
'tR(z,1'\)=~2&+2 dz2 +1'\.'

(S.207d)

vL(z,1'\)=X

(S.20Sa)

dX

'tL(Z, 1'\) = Il dz

(S.20Sb)

The equations governing eb(z, 1'\), 'P(z, 1'\) and X(z, 1'\) can be obtained by substituting
Eqs. S.204 into Eqs. S.202, i.e.
(S.209a)
(S.209b)

332

Foundations of Solid Mechanics


2-

~~ = <X~(11)X

(8.209c)

where <Xp(11) and <X.(11) are as defmed in Eqs. 8.167. The solution to Eqs. 8.209, which
are ordinary differential equations, is simple, i.e.

cb = A~e-0,.- + B~e~-

(8.210a)

'it=A.pe-iX,z +B.pea,z

(8.210b)

X=-\e-iX, + Bxea,z

(8.21Oc)

where A~, B~, Av , Bv , -\ and Bx are arbitrary constants. If 11p < 11 < 11., it is more
convenient to write Eqs. 8.210b and c as

'it = Avcosa.z +Bvsina.z

(8.211a)

X= Axcosa.z +Bx sina.z

(8.211b)

where a.(11) is as defmed in Eq. 8.171. If 11 < 11p < 11., Eq. 8.21Oa should be modified
also for the sake of convenience as
cb=A~cosapz +B~sinapz

(8.212)

where ai11) is as defmed in Eq. 8.173. The arbitrary constants are to be determined
from boundary conditions. For a space of N layers each having a finite dimension in
the z-direction, the formulation leads to a set of 6N linear equations in 6N unknown
arbitrary constants; A~l to A~N' B~l to B~N' AV1 to AVN , BV1 to BvN , -\1 to AxN, and BX1
to B xN As an example, consider the interface at z = Z,,_l between the layers n -1 and
n. If there are no any applied tractions at this interface, the relevant conditions are the
continuity of
Noting, in Eqs. 8.206, that
dJm(nr)

Jm(nr)
and - d(nr)
11 r
are independent of each other we see that these conditions can be expressed (for z = Zn-1)
as

(8.213)

(8.214)
in which the subscripts n -1 and n denote the respective layers. Deliberately, the last

333

Wave Propagation

two equations are put separately to show that ~ and ' appear together in a set of four
equations, while X appears separately in another set of two equations. Deserving notice
is that these equations are independent of Fourier series hannonic m. Only the R-set
of equations, which is for ~ and ', exists in the axisymmetric (torsion free) case where
Ue =0 and all functions are independent of e.
For torsion about the z-axis, Ur and U. must vanish, all functions are independent
of e, and the only non-trivial equation of motion is
~Ug(r,z)=--11!Ug(r,z)

in which ~ is defmed by Eq. 6.47 with m


a-IUg(r,z)

=1.

(S.21S)

Thus Ue(r, z) assumes the fonn

=fO ue(z, TJ)JI(TJr)dTJ

(S.216)

and we can write, according to Hooke's law,


O'a.(r, z) =

L-

aa.(z, TJ)J1 (TJr)dl1

(S.217)

where
(S.21S)
Now, if we change symbols as follows: ue into X( == VL), and aa. into 'tL' we can write
boundary conditions in the identical fonn as the L-set of equations of the general case.
In other words, only the L-set of equations exists in this 'pure torsion' case. Substituting
Eq. S.216 into Eq. S.21S we obtain an ordinary differential equation of the second order
with respect to z, with the solution

=TJ(Be -,s +DeB,Z)

(S.219)

=TJ(B cos a.z +D sin a.z)

(S.220)

u e(z,l1)

or
Ue(z, TJ)

where B and D are arbitrary constants to be determined from boundary conditions.


For a sufficiently large r, the approximate formulas for Bessel functions (Eqs.
1.121) can be used to reduce the amount of computation, since Eqs. S.206a,b,e and f
are simplified into
ur(r,Z,TJ) =TJu R (Z,TJ)Jm_l(TJr),
(r -+00)
(S.221a)
ue(r,z,l1)=--11VL(Z,TJ)Jm_l(TJr),

(r -+00)

(S.221b)

aa.(r,z,TJ)=-TJ't L(z,l1)Jm_l(W),

(r -+00)

(S.221c)

a .. (r,z,TJ)=TJ't R (z,l1)Jm_l(TJr),

(r -+00)

(S.221d)

These equations show the separability of boundary conditions into R and L sets of
equations. Moreover, we can treat the problem as two separate problems; one being
associated with potentials ~ and ', and another with X only.

Foundations of Solid Mechanics

334

When we are investigating the possible existence of surface waves, we seek to


determine discrete values of 11. In other words, we investigate the possible existence
of the harmonic vibration when there is no applied traction on the surfaces nor on any
interfaces between layers. The R -set of 4N linear homogeneous equations gives the
Rayleigh wave equation identical to that for the in-plane case, i.e. Eq. 8.174. On the
other hand, the L-set of 2N linear homogeneous equations gives the Love wave equation
identical to that for the antiplane case, i.e. Eq. 8.192. In general, Rayleigh waves and
Love waves are dispersive, but their wavenumbers must be real, as concluded by Ewing,
Jardetzky and Press (1957). The solution corresponding to a surface wavenumber 11
can be constructed simply by removing the integration symbol in all expressions derived
so far, and replacing Jm(11T )dl1 in those expressions by
Jm(l1r)

or Ym(11T)

or H2)(l1r)

or H~)(l1r).

Which function of r should be selected is determined by the requirement that each


surface wave must be attenuating and outgoing as r tends to infInity. Thus shape
functions for far fIeld Rayleigh waves and Love waves assume the forms of
H~)(l1Rr)

and H~)(l1Lr),

respectively. It should be noted that the effects of the Love waves on 'axisymmetric'
functions (associated with potentials <I> and 'P), e.g.

u" u., 0'.. and 0',.


are negligible for r ~ 00. The same can be said for the Rayleigh waves for the 'pure
torsion' functions (associated with potential X), e.g.
lie

and

O'er

The complete solution to problems of forced vibration consists of the surface wave
functions already discussed and also the Cauchy's principal values (Spiegel, 1963) of
the infmite integrals, Eqs. 8.204. The integrands of these infmite integrals are naturally
singular at the values of the surface wavenumbers. Unfortunately, none of the existing
solutions for multilayered spaces are in such a form that a detailed investigation of the
associated infmite integrals is possible. Jardetzky (1953) made an attempt to answer
some of the important questions associated with these integrals without obtaining a
complete solution to the problem. But the problems considered by Jardetzky were
axisymmetric, so only Rayleigh waves were discussed. Later, Press, Harkrider and
Seafeldt (1961) developed, for multilayered half spaces, a computer algorithm to
determine the roots of Rayleigh and Love wave equations, Eqs. 8.174 and 8.192,
respectively. The number and values of these roots play an important role in the solution
approach, since they are poles of the integrands when evaluating the infmite integrals
by means of appropriate contours in a complex plane. Furthermore, such a process
involves branch line integrals. Jardetzky ventured to make a general conclusion that,
in all cases of wave propagation from a point source in a half space formed by parallel
layers displaying different elastic properties, all expected branch line integrals vanished

335

Wave Propagation

except the one corresponding to the branch points (11, and 11.) of the underlying half
space. Thus each potential was obtained in the fonn of a sum of two parts, of which
the frrst part represented a discrete spectrum of surface waves and the second a continuous spectrum given in the fonn of a branch line integral. This important point (by
Jardetzky) will be scrutinized further.
Fonnally, the C.P.V. (Cauchy's principal value) of each infinite integral in Eqs.
8.204 can be obtained by means of appropriate contours in the plane of the complex
variable ~ (Churchill, 1960), i.e.
~=l1+i~.
These contours bypass branch points and poles on the real axis 11. Branch points are
at ~ = n. and n" while poles are at ~ = nR or nL' The C.P.V. obtained in such a
way contains continuous spectra of body waves and discrete surface waves. Each surface
wave is directly associated with the residues at a particular surface wavenumber (or two
opposite poles). However, these surface waves do not assume the outgoing nature for
large r. Thus a superposition upon such a surface wave by another surface wave of
the same wavenumber is needed to make that particular wave to be in the fonn of
H~)(l1Rr)

or H~)(l1Lr),

(for large r).

In summary, the complete solution to problems of forced vibration can be taken as a


summation of a body wave solution and a surface wave solution, and each of these
solutions individually satisfies all governing conditions outside the loaded domain. The
body wave solution consists of continuous spectra of waves, while the surface wave
solution consists of discrete waves. In addition, we can derive a unique relationship
among all functions of surface waves, of the same as well as of different wavenumbers,
but of the same wave type - (Rayleigh type or Love type). Various important conclusions made in this paragraph, on the composition of the complete solution and the
individual natures of body and surface wave solutions, are to be verified in the next
two sections where two simple cases are considered.
8.9 ASYMMETRIC VIBRATION OF A HOMOGENEOUS HALF SPACE
Consider the harmonic vibration of a homogeneous isotropic elastic half space
occupying the domain 0 ~ z < 00 and subjected to a unifonn shear traction on its surface
in a circular area of radius a, such that
(Ju(r,
0) =0, (0 ~ r < 00)
(8.222)

e,

J.l, (OSr<1)

(8.223a)

=0, (1 <r <00)

(8.223b)

(J... (r,e,O)=

Equations 8.223 can be expressed as the single equation


(J... (r ,e,O)

=J.l

l-

J 1(11)Jo(l1r )d11

(8.224)

Foundations of Solid Mechanics

336

or, due to a recurrence fonnula of Bessel functions, Eq. 1.112c,


CJ (r,
....

e, 0) =J.1

i0

J (11) [ dJl (l1 r ) + J 1(l1r )]dl1


1
d(l1r)
l1r

The associated cylindrical stress components are


CJ&.(r, e, 0) =-(J1/S(r, e, 0) sin e
CJn(r,

(8.225)
(8.226)

e, 0) =CJ1/S(r, e, 0) cos e

(8.227)

In view of Eqs. 8.205 to 8.208,8.222,8.223 and 8.225 to 8.227, we can set m 1, and
CJR (O,l1) 0
(8.228a)

'tR(O, l1) =J.111- 1Jl(11)

(8.228b)

=J.111- 1Jl(11)

(8.229)

'tL(O, 11)

Substituting Eqs. 8.207c and d into Eqs. 8.228, and using Eqs. 8.210a and b (with

Bib =B", =0 for boundedness of the solutions as z tends to infinity), we fmd


2
[ 2112 -11~ - 211 a.l {Aibl { 0 1
2112-11~

-2a p

AJ = 11- J (11)J
1

On the other hand, using Eqs. 8.208b and 8.21Oc (with Bx


solution as z tends to infmity), in Eq. 8.229, we get

(8.230)

=0 for boundedness of the

-.Ax =11- 1J 1(11)

(8.231)

Solving the preceding two equations yields


A _ 211a.(11)Jl(11)
F R (l1)

(8.232a)

A _ (2112-11~)Jl(11)
'" l1FR(l1)

(8.232b)

ib-

J 1(11)

(8.233)

~ =- 11a. (11)

where FR(l1) is the Rayleigh wave function identical to that of the in-plane problem,
i.e. Eq. 8.177. We can, consequently, say again that there exists a Rayleigh wave but
not a Love wave for the vibration of a homogeneous half space. Subsequently, we can
write Eqs. 8.204 in the fonn

. . . ( e)=

..,r,
'(

,z

cos

ei- 211 <X,(11)Jl(11)Jl(l1r )e-a,C1l)'d


F R(l1)

(2

2)1 (

)I ( )

-,(II)'

ei- 11 -11. 111


l11 r e
r, e)=
,Z
cos
F ()
o

l1R11

11

(8.234a)

d11

(8.234b)

Wave Propagation

337

. i-

J 1(Tl)11 (Tlr)e-a,(11).
(8.235)
(. )
dTl
o
Tl<X,\Tl
Now, our immediate task is to determine the Cauchy's principal value (C.P.V.) of each
infmite integral in the equations above. For that in Eq. 8.234a and r> 1, consider the
complex integrand
x(r,e,z) =-sme

2~(l6(~)11 (~)H~I)(~)e-a,(I;).
FR(~)
d~

(8.236)

and use the dual integral contours recommended by Lamb (1904) as depicted in Fig.
8.15, where ~ is the complex variable, i.e.
~ = Tl + i~
(8.237)
Consistent values of radical quantities (lp(~) and (l.(~) are also indicated on various parts
of the contours. The integrand in question vanishes on the integral paths where I~I -+ 00.
Thus, according to the residue theorem of complex variables, the contour C1 gives

r-2Tla.(Tl)11(Tl)H~I)(Tlr)e

-a,(Il)' dTl

Jo

= 1ti2TlR(l.(TlR)11(TlR)H~I)(TlRr)e-a,(IlR)'

F~(TlR)

FR(Tl)

. r-2i~(l.(i~)11 (i~)H~I)(ig)e-a,~'

+1 Jo

FR(i~

d~

- -- -

CI)

tI

...

c..

tI

CI)

tI

...

c..

tI

Fig. 8.15 Integral contours for 4>(r,e,z) and 'Y(r,e,z), Eqs. 8.234.

(8.238)

Foundadons of Solid Mechanics

338

where
(8.239)
On the other hand, the contour C2 gives

1-

-1'1, 211ail1)Jl (l1)H[l)(l1 r )e-,{tJ).

--------dl1-

l-l'lp211as(11)Jl (l1)H[l)(l1r )e

~~

-,(11)'

~~

dl1

r 0 211as(11)Jl (11)H[1)(11T)ea.,.(II). d _ 1ti2(-TlR)a.(--11R)J&-l1R)H[1)(-11Rr)e


- J-I'Ip
FR(l1)
11F'R(-11R)

-'(-I'IR)'

. r-2i~a6(i~)Jl(i~)H[1)(ig)ea.,.r~. ~
+1 Jo
F R(i~)
d..,

(8.240)

where
(8.241)
Due to the analytic continuation, Eq. 1.114e, we can have identities involving Hankel
functions of negative arguments as follows
H~)(-l1)

=-H~)(l1),

(m

=0,2,4, 6, ... )

(8.242a)

= H~)(l1),

(m

=1,3,5,7, ... )

(8.242b)

Subsequently, rewriting Eq. 8.240 to involve only positive limits of integration and
positive arguments of functions we fmd
r-211a..(11)Jl(11)H?)(l1r)e -,(11)< dl1 = _1ti211 Ra s (l1R)Jl(l1R)H?)(l1Rr)e

Jo

F R(l1)

-,(IIR)

F'R(l1R)

lip 211a

(11)J (l1)H(2)(l1 r ) [ea.,.(II). + e-,(II)J


1
1
dl1
F R (l1)

+ r II, 4(211 2 -l1~il1as(l1)Jl (l1)H?)(l1 r )e -,(11)' dl1

Jllp

F R(l1)F R(l1)

. r-2i~a..(iWl(i~[1)(ig)ea.,.(~
+1 Jo
FR(i~)
d~
Adding Eqs. 8.238 and 8.243 together, and using Eqs. 1.113a and b, we get

L-

211a.(11)Jl(11)Jl(l1r )e-,(11)'
--------dl1 =
o
~~

1t211Ra,(l1R)Jl(l1R)Yl

~~

(l1Rr)e -,(IIR)'

(8.243)

Wave Propagation

339
(TIp 211<X.(11)Jl('l"\)Hf)(l1r)COs[(11;

+ )0

-11 2 12z]

dl1

FR (l1)

(TIl 2(211 2

-11~)211<X.(11)Jl (11)H~2)(l1r)e-p(TI)z dl1

)Tlp

FR(l1)F R(l1)

(~2~<x.(i~)Jl(i~)H~1)(ig) COs[(~2 + l1;)II2Z]


- )0
FR(i~)
d~

(8.244)

It should be noted that every tenn on the right-hand side of the equation above, except
the fIrst one, characterizes a wave which is outgoing and/or attenuating as r increases.
Due to the presence of H~l)(ig), the last tenn of that equation decays exponentially.
To make the fIrst tenn on the right-hand side of Eq. 8.244 to represent an outgoing
wave for large r as well, we should superimpose upon it a surface wave function equal
to
1ti211 R<X.(l1R)Jl(l1R)Jl (l1Rr)e -P(TlR)z

(8.245)

F~(l1R)

Accordingly, we have for r > 1,


.m(
'V r,

e,Z )-_cos e{

) -P(TlR)z
i2
()J ( ) (2)(
1t l1R<X. TlR 1 l1R HI l1Rr e
- - - - - : - ,- - - - - FR(l1R)
(TIp 211 <x. (l1)Jl (l1)Hf)(l1r ) cOs[(l1;

+ )0
+

F R (l1)

-11 2rz]

dl1

(TIl 2(211 2-11;)211 <X.(11)Jl (l1)Hf)(w)e-p(TI)z dl1

)Tlp

FR(l1)F R(l1)

(~2~<Xs(i~)Jl(i~)H~1)(ig)cos[(~2+11;rzJ }
FR(i~)
d~

(8.246)

- )0

The expressions derived so far are not valid for a case of small values of r which is
our next task. To detennine the Cauchy's principal value (C.P.V.) of the infInite integral
in Eq. 8.234a for r < 1, the complex integrand to be used is

2~<x.(~)Hfl)(~)Jl (~)e-p(Qz

FR(~)

d~

(8.247)

but, for continuity of solution at r = 1, the same superimposing surface wave function,
Eq. 8.245, must be used. Proceeding in the same manner as having obtained Eq. 8.246
we can eventually show that this equation holds as well for r < 1, if in that equation
we replace

Foundations of Solid Mechanics

340

by Hf)(11)

J 1(11)

by

J 1(11 r )

J 1(i;)

by

H~l)(i;)

H~l)(i;r)

by

J 1(ig)

H?)(11 r )

(8.248)

expressions derived so far hints that they are not suitable for large values of z. For
large z, the integration in Eq. 8.234a can be truncated at 11, which is sufficiently larger
than 11 p ' and we have
.2
-a,.(I1)Z
.m.( 9 ) 9 l1 11<X..(11)Jl(11)Jl(11 r )e
d
(8.249)
'V r, ,z -cos
Jo
FR (11)
11

which properly contains spectra of outgoing and exponentially decaying waves as z


tends to infmity. At this point, we may conclude that any physical quantity given by
infmite integrals as follows
A( ) = r~!(11,z)Jm(11r)d
r,z
Jo FR (11)
11

(8.250)

is equal to, for small and intermediate values of z,


A(r,z)

=c.P.v.[ r~!(11,z)Jm(11
Jo FR(11)

r ) d111- 1ti!(11R',Z)Jm( 11R r)


~
F R(11R)

(8.251)

In addition, Eq. 8.251 implies that the C.P.V. of an infmite integral alone implicitly
contains discrete non-outgoing waves for large r, thus does not give the correct value
for a physical quantity.
At this stage, it should be quite obvious that the evaluation of \f(r, 9, z) defined
in Eq. 8.234b can follow the same procedures as used for cI>(r, 9, z). The results are
the following: for small to intermediate z,
\f(r, 9, z) =cos 9

{c.P.v.[ r~(2112 -11;)Jl(11)Jl(11 r )e


Jo

11FR(11)

-<1,(I1)Z

d11]

(8.252)

(8.253)
where 11 is sufficiently larger than 11. In detail, Eq. 8.252 can be written for r > 1 and
small to intermediate z as,

Wave Propagation

341

) -,(l1R)<
e, =COS e{ _m'(21'\R - 1'\.2).11(1'\~ )H(2)(
1 1'\Rr e
2

\{I(r, z)

()

COS 1'\.z
21'\~r

1'\RFR(1'\R)

'll1 p (21'\2 -1'\;).Il(1'\)Hf)(w) sm[(1'\; -1'\2)1I2Z


-1
~

1'\FR(1'\)

11, (21'\2 -1'\;).Il(1'\)Hf)(1'\r)[e-,(11)<

l1p

ea'(l1)<]d
1'\

----=-FR(1'\) F R(1'\)

21'\

. (_(2~2 +1'\;).Il(i~)H~l)(ig) Sm[(~2 +1'\;)lI2ZJ }


+1 Jo
~R(i~)
d~

(8.254)

The equation above also holds for r < 1 and small to intermediate z, if we replace
various terms in the equation as stipulated in Eq. 8.248 and replace r-1 (in the second
term of Eq. 8.254) by r. In fact, the second term of Eq. 8.254 is the residue of the
integral at ~ =0, and the singular integrands at the lower limit of integration in the third
and fifth terms of Eq. 8.254 nullify each other.
Now, there remains the task of determining x.(r,e,z) defined in Eq. 8.235, which
does not have any surface wave characteristics, hence the task is much simpler than for
<I(r, e, z) and \{I(r, e, z). The dual integral contours to be used are as depicted in

II)

t:I
I

'-"

II)

t:I

Fig. 8.16 integral contours for x.(r,e,z), Eq. 8.235.

Foundations of Solid Mechanics

342

Fig. 8.16. The results contain only the Cauchy's principal value (C.P.V.) of the infInite
integral involved, i.e. for small to intennediate z,

l}

x(r,a,z)=-sina{c.p.v[ r- J l(Tl)Jl(Tlr)e-,(Il)' d
Jo
Tla..(Tl)
TlJ

(8.255)

and for large z,


(8.256)
where Tl is suffIciently larger than Tl,. In detail, Eq. 8.255 can be written for r > 1 and
small to intennediate z as,
. a{ sin(Tl,z)

X(r, a,z) =-sm

2rTl.

LIl'Jl(Tl)H?)(Tlr)cos[(Tl~_Tl2)1I2zJ
Tla.(Tl)

r-Jl(i~)H~I)(ig)cos[(~2+Tl~)lflzJ

+ Jo

~a.(i~)

d~

dTl
(8.257)

The equation above also holds for r < 1 and small to intennediate z, if we replace
various tenns in the equation as stipulated in Eq. 8.248 and replacing r-1 (in the fIrst
tenn of Eq. 8.257) by r. Again, it should be noted that the singular integrands at the
lower limit of integration in the second and third tenns of Eq. 8.257 nullify each other.
In summary, the complete solution to this problem can be expressed as a sum of
a surface wave solution and a body wave solution, i.e.

=c'II(r, a,z) +4>b(r, a,z)


'P(r,a,z) =o/(r,a,z)+ (r,a,z)

(8.258b)

x(r, a, z) = t (r, a, z) + x b(r, a, z)

(8.259)

CI>(r, a,z)

(8.258a)

in which the superscripts s and b denote a surface wave and a body wave, respectively.
Explicitly, the surface wave solution functions, for r> 1 and z not large, are
c'II(r, a, z) =

'P'(r, a,z)

t(r,a,z)=o

21tiTlRa,(TlR)Jl(TlR)H?)(TlRr)e-P(IlR)' cos a

F~(TlR)

1ti(2Tl~ - Tl~)Jl (TlR )H?)(TlRr)e -,CIlR)r cos a

TlRFR(TlR)

(8.260a)

(8.260b)
(8.261)

Comparing Eqs. 8.258a,b and 8.259 with Eqs. 8.246, 8.254 and 8.257, respectively, and
noting Eqs. 8.260 and 8.261, we can identify explicitly the body wave solution functions
4>b(r,a,z), (r,a,z) and xb(r,a,z).
With a brief reflection on the results obtained in this section, one should be able to

Wave Propagation

343

appreciate some conclusions made in the last paragraph of the preceding section for the
general case.
8.10 AXISYMMETRIC TORSION OFA LAYERED HALF SPACE
Consider the harmonic torsional vibration of a layered half space consisting of an
upper layer (1) where -hla S z SO, and an underlying half space (2) where 0 S z < 00.
There will be Love waves provided that the inequality Eq. 8.197b holds. The half space
is subjected to a shear traction on its top surface in a circular area of radius a such that
CJ.9l (r,-hla) =J.l.lr, (r < 1)
(8.262a)

=0,
(r>l)
Equations 8.262 can be written as the single equation (Eq. 6.1l2g),

(8.262b)

(0 S r < 00)

(8.263)

CJ,9l(r ,-hla) = J.l.l

L-

J 2(11)Jl(11 r )dr\.

Relevant boundary conditions for the present case are


Usl (r,0) = Um(r, 0), CJ.91 (r,0) = CJ.f1l,<r, 0)
= J.l.l

CJ.9l (r,-hla)

L-

(8.264a,b)
(8.264c)

J 2(11)Jl(11T)d11

The solution is chosen to take the fann of Eq. 8.220 for the upper layer, and Eq. 8.219
with D2 = 0 for the underlying half space, i.e.
Usl(r,z) =

L-

11(Bl cos adZ +Dlsinad Z)Jl(11T)d11

Um(r,z)

L-

(8.265)
(8.266)

11B2e-a,2'Jl(11r)d11

The frrst boundary condition, Eq. 8.264a, readily gives


B2=Bl

(8.267a)

then the remaining boundary conditions can be put as


[

J,l.za.2

J.l.lasl

ad sin(aslhla) ad cos(adhla)

J{Bl}
Dl

= J2(11)/11

(8.267b)

This equation can be solved for Bl and D l Substituting the results in Eqs. 8.265 and
8.266 yields
(8.268a)
(8.268b)

344

Foundations of Solid Mechanics

-I
Fig. 8.17 Integral contours for Ue(r,z) in Eqs. 8.268.
cosas 1z
A (TJ,z) == - - = -
a sl cos(as 1h/a)

(8.269a)

(8.269b)
e~.r

AzC11, z) ==-----c=-a s1 cos(as1 h /a )

(8.269c)

The Cauchy's principal values of infmite integrals in Eqs. 8.268 can be detennined
as in the previous section using dual integral contours as shown in Fig. 8.17. For
simplicity, only one surface wavenumber is explicitly shown in the figure; i.e. one pole
at ~ = 11L lying in between branch points at ~ = 11s1 and 11s2' and another pole -11L in
between branch points -11.1 and -11.2. For the present case, we find that for r > 1,
Uel(r,z)

~ 1ti[A(11L, z) + C(11L,Z)]J 2(11L)H?>(11Lr)

=- ~

F L (11d

(8.270a)

345

Wave Propagation

Ue2(r,z) = _l:1tiA(T1L,z~z(l1L)H?>(l1Lr) + T1a A4(11,z)Jz(11)Hf>(l1 r )d11


FL(l1L)
Jo
+i i- A4(i~,z)Jz(i~)H:1>(ig)d~,
Here

(Z not large)

(S.270b)

l: denotes summation of all surface waves, and


F'
_[~~)] _dPL~)1
L(l1L) - l1-l1L _ - dl1
T1-T1

A ( z)
311,
A4(11,z)

(S.271a)
T1=T1

A(l1,z)+C(l1,Z) A(l1,z)-C(l1,Z)
2FL(l1)
2F dl1)

(S.271b)

Az(l1,Z) Az(l1,-z)
2F dl1)

(S.271c)

= 2FL(l1)

in which
(S.272)
The higher the surface wave mode, the less significant it is, due to the closer to infInity
its F~(1lL) is. It should be noted that Eqs. S.270 do not contain any integral with respect
to 11 in the range from 11.z to l1d' For r> 1 and large z, Eq. S.270b does not hold, so
Usz(r,z) should be calculated anew by truncating the integration in Eq. S.26Sb at 11',
which is suffIciently larger than 11.z, i.e.

Ue2(r,z)=c.p.v{iT1Az(l1'Z~z(~l(l1r) d~

(S.273)

The equation above holds for r:S 1 also as long as z is sufficiently large. For
small and intermediate z and r < 1, we can show that Eqs. S.270 hold, provided that
we replace
by Hf>(l1)
'z(l1)
Hf>(l1r ) by
(l1 r )
by H~l>(i~)
'z(i~)
H:1>(ig) by 'l(ig)

'1

(S.274)

and add to Eqs. S.270a and b, respectively, the following terms

[ A(O,Z)+C(O,Z) + A(O,Z)-C(O,Z)]
2F L(O)
r
2FL(O)
and

(S.27Sa)

Foundations of Solid Mechanics

346

(8.27Sb)
which are residues of integrals at the pole ~ =O.
The complete solution to this problem can be put as a sum of a surface wave
solution and a body wave solution, i.e.

=U;l(r,z) + u;l(r,z)
Um(r,z) =~(r,z)+~(r,z)

(8.276a)

Uel(r ,z)

(8.276b)

The surface wave functions


U;l(r,z)

and ~(r,z),

for r > 1 and z not large, are the ftrst terms on the right-hand side of Eqs. 8.270a and
b, respectively.
8.11 TOTAL SOLUTION TO VIBRATION OF HALF PLANES
For in-plane problems of harmonic vibration of isotropic elastic half planes, any
physical quantity is in the form of an inftnite integral as follows
(8.277a)

A(x y)= r-/(rt,y)cosrtx dn


FR(rt)
'I

Jo

or
(8.277b)
where FR(rt) is the Rayleigh wave function, Eq. 8.174. For antiplane problems, any
physical quantity takes the same forms as above with FR(rt) being replaced by the Love
wave function FL(rt), Eq. 8.192. Any surface wavenumber, rtR or TtL, if it exists, must
be real positive and greater than the shear wavenumber (rt.) of the underlying half plane,
and greater than that of the overlying half plane for a multilayered full plane. For the
underlying half plane, /(rt, y) in Eqs. 8.277 is exponentially decaying, i.e.
/(rt, y)

=h. (rt)e-<J('l)Y

(8.278)

where a(rt) is a radical function for the underlying half plane given by Eq. 8.167a or
b. For the overlying half plane, a(rt) should be replaced by -a(rt).
After going through the same arguments as carried out for three dimensional
problems in the preceding sections, one should be able to arrive at the following conclusions: Any physical quantity given by Eq. 8.277a is, for small and intermediate
values of y,
A(x,y)

=c.P.v.[ r-/(rt,y)cosrtx drt1-l: 1ti/(rtR,;)cosrt~


Jo

FR(rt)

FR(rtR)

(8.279)

Wave Propagation

347

where L denotes summation of all Rayleigh waves involved in the problem being
considered, and F~(1'\R) is the derivative of the Rayleigh wave function at a Rayleigh
wavenumber 1'\R' For sufficiently large values of y in the underlying half plane, the
integration in Eq. 8.277a can be truncated at 1'\., i.e.
A(x,y)

=c.p.V.[il1f(1'\'~~(~~S1'\X d1'\]

(8.280)

where 1'\. is sufficiently larger than 1'\p or 1'\. depending on which is involved in a(1'\)
in Eq. 8.278. The Cauchy principal value (C.P.V.) in Eq. 8.279 can be put in a more
explicit form by integrating a complex integrand along appropriate contours. For x> 1,
the complex integrand to be used is
.~

f(~,y)el d~
FR(~)

(8.281)

where 9t(~) =1'\. For a simple case such as a homogeneous half plane, the dual contours
depicted in Figs. 8.15 and 8.16 can be used. However, for a more complicated case
such as a layered or multilayered half plane, we follow the procedure: formulate the
surface wave function and its derivatives, and Eq. 8.277a symbolically; evaluate the
surface wavenumbers, the derivative of the surface wave function at each and every
surface wavenumber; evaluate Eqs. 8.279 and 8.280 numerically.
All these schemes are applicable to any physical quantity given by an infmite sine
integral of the form Eq. 8.277b.
8.12 VffiRATION OF VISCOELASTIC HALF SPACES
The solution to the vibration of an isotropic viscoelastic solid can be obtained
from that of the corresponding elastic one by simply replacing the Lame's constants A.
and Il by the complex moduli A.(ro) and Il(ro), respectively. These moduli have positive
real parts and positive imaginary parts. So the real and imaginary parts of the wave
velocities, cp and c. given by Eqs. 8.12 and 8.14 respectively, are also positive. The
dimensionless body wavenumbers, ~p and ~ (~= 1'\ + i~) determined from Eqs. 8.164b
and c, will have positive real parts but negative imaginary parts. The surface Rayleigh
and Love wavenumbers, ~ and ~ determined from Eqs. 8.174 and 8.192 respectively,
will have the same signs as the body wavenumbers.
Then any physical quantity in the vibration of a multilayered viscoelastic half plane
can be evaluated with the same procedures as those described for the elastic one in the
preceding section. However, each physical quantity in the present case is given solely
by the C.P.V. of its infmite integral. With reference to the dual contours as in Fig.
8.15 (or 8.16 or 8.17), branch points and poles (if any) will not be on the contours like
the elastic case. The 'positive' branch points and poles (~p' ~, ~ or ~) are in the
fourth quadrant of the complex plane, while the 'negative' branch points and poles (-~p,

348

Foundations of Solid Mechanics

or -~) are in the second quadrant inside the contour C2 The residue at each
pertinent pole, -~p or -~, has an outgoing wave fonn as x tends to infInity. Hence,
the C.P.V. ofan infInite integral representing a physical quantity already contains
(though not explicitly) discrete outgoing surface wave functions, and it is not necessary
to superimpose as the second term on the right-hand side of Eq. 8.279.
This scheme is applicable to three-dimensional viscoelastic cases.

-~, -~

8.13 INFINITE ELEMENTS FOR A HOMOGENEOUS HALF SPACE


Analytical methods such as that described in Sections 8.9 and 8.10 become
impractical if not impossible when extended to load transfer problems and/or multilayered half spaces. Several numerical methods have been developed to bypass such
difficulties. At this stage of development, the most useful numerical algorithm seems
to be the one that involves infmite elements for the far field and finite elements for the
near field, as shown in Fig. 8.18(a). At present this algorithm is restricted to a
homogeneous isotropic elastic half space. Research on algorithms for layered and
multilayered half spaces is in progress. The present section describes infInite elements
which have been developed by Rajapakse and Karasudhi (1986) and Karasudhi (1986)
for load transfer problems in a homogeneous isotropic elastic half space.
The whole material domain is separated into a near field and a far field as shown
in Fig. 8.18. The near field, consisting of the partially embedded bar and a fInite region
of the half space around it, is discretized into conventional fInite elements. The far
field covering the rest of the half space is descretized into infmite elements. Near the
surface of the half space are horizontal infInite elements (HE), while the remainder of
the far field is occupied by radiating infmite elements (RE). Every infInite element has
point nodes only on the interface between the near field and the far field. Displacement
functions of an elastodynamic infInite element should decrease with distance, and
represent outgoing waves. A logical way to obtain such displacement functions is to
investigate analytically the far field behavior of the half space.
Axisymmetric torsional oscillation of a homogeneous half space is one of the
simplest wave propagation problems, since there is only one body wave and a single
displacement component Us governed by Eq. 8.215. For sufficiently large r, the solution
is
(8.282)
Instead, if the spherical Navier equations (Eqs. 3.68) are considered as the governing
equations, then for sufficiently large R
Us-h~)(T\.R)

(8.283)

A verification of Eqs. 8.282 and 8.283 can be achieved with the help of the approximate
formulas for Bessel functions for large argument, Eqs. 1.121. These solutions are rather

349

Wave Propagation

Embedded Bar

--I

~ro -1
I

/--201

z
(a) Geometry.

z
R
(b) Infinite elements, HE and RE.
Fig. 8.18 Discretization of material domains.
special and 'particular'. While they satisfy the equations of motion and the attenuating
and outgoing wave requirements, they cannot satisfy the boundary conditions at the half
space surface, since each is a function of a radial coordinate (r or R) only. Weighing
advantages and disadvantages of these solutions against those of continuous spectra of

Foundations of Solid Mechanics

350

waves (see Eqs. 8.270 and 8.273), Karasudhi et al favored the fonner. They used Eq.
8.282 for horizontal infInite elements and Eq. 8.283 for radial infInite elements. These
equations imply that cylindrical wavefronts exist in a horizontal infInite element, and
spherical wavefronts in a radiating infInite element. The waves in these respective
infmite elements attenuate as r-l12 and R-l , i.e. for a horizontal infmite element,

U;-(T).7r

l12 -itJ,r

(8.284)

u!-(llftrl e-iT),R

(8.285)

and for a radiating infmite element,


The superscript b denotes the body wave solution, which is also the complete solution
for the axisymmetric torsion of a homogeneous half space. In order to enable simple
analytical evaluation of infmite integrals of impedance and mass matrices of infmite
elements, the displacement function for a horizontal infmite element was taken in the
fonn

ue-e-<l+i)ll,'

(8.286a)

ue-e-<1+i)ll,R

(8.286b)

and for a radiating infmite element,


b

In general vibration of a homogeneous half space, each displacement component


consists of a body wave part and a surface wave part. Following the scheme adopted
for torsional vibration, we fInd the body wave displacement functions, for a horizontal
infmite element,
b -<l+i)llp'
u,-e

(8.287a)

b -<1+i)ll,'
ue-e

(8.287b)

b -<l+i)ll,'
u.-e

(S.287c)

b -<l+i)llpR
uR-e

(8.288a)

b -<l+i)ll,R
u,-e

(8.288b)

b -<l+i)ll,R
ue-e

(8.288c)

and for a radiating infmite element,

Analytically, the surface wave displacement components can be obtained as

_pz

u"" --Am(e

A (

-qz

dH!;)(ll Rr)

-ale ) d(llRr)

-pz

-qz)H!;)(llRr)
llRr

Uem =m me-ale

(8.289a)

(8.289b)

Wave Propagation

351

(8.289c)
where the superscript s denotes the surface wave, m the hannonic in the Fourier series
with respect to Am an arbitrary constant, T\R the Rayleigh surface wavenumber of the
homogeneous half space, and
(8.290a,b)
p =a.(T\R)' q =a/T\R)

e,

a l =2pqia/iT\R)'

a z =2T\iia p (iT\R)

(8.290c,d)

The radical quantities in the equations above are as defined in Eqs. 8.167, 8.171 and
8.173. It should be noted that the origin in the coordinate system for Eqs. 8.289 is on
the surface of the half space. Equations 8.289 may be assumed to take the following
approximate form
-- iAmiZe
b ( ) -{1+il11Rr
(8.291a)
urm
bUem

(Z) e-{1+i)l1R r

-A
b ()
Uzm
m zze

(8.291b)

-{1+il11Rr

(8.291c)

-a1e-<[Z)

(8.292a)

where
bl(z)

=(e-

bz(z)

=T\~lp(e-PZ - aze

PZ

-1/Z )

(8.292b)

Corresponding spherical displacement components for this surface wave can be obtained
by a standard coordinate transformation as
u;m

=Am[ib l(z) sin <I> + bz{z) cos <1>] e-{I +il11Rr

(8.293a)
(8.293b)
(8.293c)

Assuming that the displacement components in Eqs. 8.291 and 8.293 are independent
of each other, we obtain surface wave displacement components for a horizontal infinite
element as
(8.294a)

b1(z )e-{1+il11Rr

(8.294b)

b z(z)e -{1+il11Rr

(8.294c)

Uem UZm -

and for a radiating infinite element as


u;m-[ibl(z) sin <I> + bz{z) cos <I>]e-{1+il11Rr

(8.295a)

352

Foundations of Solid Mechanics

(8.295b)
~_bl(z)e-{l+i)IJRr

(8.295c)

Total displacement components for a horizontal infmite element can be written

u"" =A;mu;", + A:mu:m

(8.296a)

uem =A!..u!.. +A~~

(8.296b)
(8.296c)

and for a radiating infmite element


(8.297a)
(8.297b)
(8.297c)
Here
are arbitrary constants, and
b
b
urm
to u.m ,

U;m to U!..

U;", to U:m, and u;m to

u~

are given by Eqs. 8.287,8.288,8.294 and 8.295, respectively. Considering each infmite
element to have two nodes as shown in Fig. 8.18(b), we can express the unknown
constants,

A;m to A:m or A;m to A~,


in terms of the displacement components and coordinates of the nodes.
8.14 EXERCISE PROBLEMS
8.14.1

Show detailed derivation for velocity jumps in isotropic elastic domains, from
Eq. 8.84 to Eq. 8.102.

8.14.2

Show detailed derivation for waves in isotropic viscoelastic media of Section


8.5.

8.14.3

A homogeneous isotropic elastic half plane -00 < x < 00, 0 S; y < 00, is subjected
to the following time harmonic tractions on its surface y =0, -a <x <a: (a)
a uniformly distributed normal force, (b) a normal force distributed in such a
way that the displacement component in the y-direction on the contact area
is uniform in the statical case (Eq. 6.66a). Find the relationship between the

Wave Propagation

353

applied force and the contact normal displacement, and compare the results
with those by Quinlan (1953).
8.14.4

With the same geometry and time harmonic functions as in the preceding
problem, the displacement in the y-direction is prescribed uniformly in the
region -a < x < a, while the shear traction is assumed to vanish everywhere
on the half plane surface. Moreover, the normal traction vanishes outside the
loaded area. Find the distribution of the contact normal stress in terms of the
prescribed indention, and the relationship between the applied force and the
indentation. Reference: Karasudhi, Keer and Lee (1968).

8.14.5

Solve the shear counterpart of the preceding problem, i.e. when the displacement in the x-direction is prescribed uniformly in the region -a <x < a,
while the normal traction is assumed to vanish everywhere on the half plane
surface.

8.14.6

Solve the rocking counterpart of Problem 8.14.4, i.e. when a uniform


rocking (rotation about the z-axis) is prescribed in the region -a < x < a, while
the shear traction is assumed to vanish everywhere on the half plane surface.

8.14.7

A homogeneous isotropic elastic half space 0 S; r < 00 ,OS; z < 00, is subjected
to the following time harmonic tractions on its surface z =0, 0 S; r < a: (a)
a uniformly distributed normal force, (b) a normal force distributed in such a
way that the displacement component in the z -direction on the contact area
is uniform in the statical case (Eq. 6.78a). Find the relationship between the
applied force and the contact normal displacement, and compare the results
with those by Quinlan (1953) and Sung (1953).

8.14.8

With the same geometry and time harmonic functions as in the preceding
problem, the displacement in the z-direction is prescribed uniformly in the
region 0 S; r < a, while the shear tractions in the x and y-directions are
assumed to vanish everywhere on the half space surface. Moreover, the
normal traction vanishes outside the loaded area. Find the distribution of the
contact normal stress in terms of the prescribed displacement, and the relationship between the applied force and the prescribed displacement. In other
words, this is a three dimensional counterpart of Problem 8.14.4. Reference:
Robertson (1966).

8.14.9

Formulate an analytical scheme for solution to general vibrations under few


loading types of two homogeneous elastic half spaces (and half planes) perfectly bonded together.

354

Foundations of Solid Mechanics

8.14.10 Fonnulate an analytical scheme for solution to axisymmetric torsion and


antiplane vibrations under few loading types of a layered elastic half space.
8.14.11

Formulate symbolically a solution scheme for axisymmetric (torsion free) and


in-plane vibrations under few loading types of a layered elastic half space,
then outline a practical algorithm to obtain results numerically.

8.14.12 Employ a few simple vibration cases of a homogeneous isotropic viscoelastic


half space (or half plane) to verify various points made in Section 8.12.

CHAPTER IX
PLASTICITY

9.1 FACTS FROM SIMPLE TESTS


If a long metal rod is pulled in a testing machine, the normal stress in the direction
of the applied load (a) may be plotted against the normal strain in the direction of the
elongation (e) as shown in Fig. 9.1, in which OA * denotes the elastic range, AB the
ideal plastic range, BD the nonlinear work-hardening range, and DE 'necking' range.
There is a small unstable region A *A, but it will be ignored in this study, since no well
accepted theories have been proposed as yet for unstable materials. In the ideal plastic
range AB, the strain e increases while the stress a remains constant. There is a nonlinear
rise of the curve beyond the point B until the ultimate load at the point D. A test

OL-______________~F~~________________~

~p~==Jt
Fig. 9.1 Typical stress-strain curve in simple tension tests of elastic-ideal
plastic-work hardening metals.

Foundations of Solid Mechanics

356

specimen may 'neck' at a certain strain, so that its cross-sectional area is reduced in a
small region. When necking occurs under continued elongation, the load reaches a
maximum (ultimate load) and then decreases, although the actual average stress at the
necked cross-section continues to increase. The specimen metal flows beyond the point
D and the specimen breaks at the point E. For mild steel, the flat yield region AB is
large due to many microscopic slips along slip planes of the material crystal. For many
other solids, the stress-strain curve may not have such a flat region (Fig. 9.2), or even
not have an elastic range (Fig. 9.3). Materials like cast iron or any rock materials,
which allow very little plastic deformation before reaching the breaking point, are called
brittle materials. However, rocks tend to become ductile when subjected to large
hydrostatic pressure (large negative mean stress au). This point of view was discovered
by von Karman (1911) in his experiments with marbles.
Upon unloading at any stage in the defonnation, the strain is reduced along an
elastic path and reloading retraces the unloading path with minor deviation, and then
further plastic deformation is produced when approximately the previously applied stress
is exceeded. For better understanding of this point, we may refer to the complete
unloading-reloading cycle CFC in Figs. 9.1 to 9.3. For all practical purposes, the
unloading path CF and the reloading path FC are identical straight lines; parallel to the
elastic line OA for Figs. 9.1 and 9.2, and parallel to the tangent at zero stress for Fig.
9.3. Thus we can see that, at any stage of defonnation such as C in the figures, the
total strain e can be taken as

A, B = Elastic limit

o
E

= Ultimate load
= Breaking point

OL-__~~F__~________________________~_

~T~

e:

Fig. 9.2 Typical stress-strain curve in simple tension test of elastic-work


hardening materials.

357

Plasticity

co

o = Ultimate
= Breaking

load
pOint

Fig. 9.3 Typical stress-strain curve in simple tension test of work


hardening materials.
e=e(P)+e(-)

(9.1)

where e(P) is the plastic strain and non-recoverable; while e(-) is the elastic strain and
can be computed from the stress at that state by the Hooke's law. If we accept this
fact (that the line CF is parallel to the elastic line OA), Eq. 9.1 holds also for a state
C1 below its yielding state C (Fig. 9.4); and the plastic strain does not change with the
position Cl. However, such a state like Cz which is above its yielding state C is

inadmissible.

Tests of specimens subjected to simple compression and simple shear give diagrams similar to those of Figs. 9.1 to 9.4. Simple experiments on specimens under
combined stresses have been mostly carried out on thin-walled circular cylinders;
subjected to a combination of an axial tension, twisting moment about the cylinder axis,
and an internal or external pressure. It is found that the shear stress, among other things,
is by far the major cause of yielding. Moreover, experiments by Bridgman (1923) have
shown that hydrostatic pressures of the order of the yield stress have practically no
influence on yielding of metals. If a tensile test or shear test is run at atmospheric
pressure in the standard manner, the stress-strain curve recorded will hardly differ in
the small strain range from those obtained in immersion of the whole apparatus in a
pressurized chamber. The major effect of the hydrostatic pressure is to increase greatly
the ductility of the material, i.e. to permit much larger deformation before fracture.
Conversely, very small changes in material density are found when a metal is subjected
to repeated plastic deformation, indicating that the plastic volume change ~) is small.

Foundations of Solid Mechanics

358

C2 ( inadmissible)
Ideal
plastic

Parallel to OA

Fig. 9.4 No change in plastic strain below yielding, and inadmissible


above-yielding state.
9.2

BASIC ASSUMPTIONS AND COMMON CHARACfERISTICS OF


VARIOUS THEORIES

There are several theories of plasticity, but their common basic assumption, based
on results of simple tests reviewed in the preceding section, can be listed for the general
three-dimensional case as the following: the strain tensor can be expressed as the sum
of the plastic strain tensor E;~) and the elastic strain tensor ~), i.e.
-

(P)

(e)

(9.2)

Eij-Eij +Eij

By its meaning, ~j> is related to the stress tensor


C (e)
CJij = ijklEkl

CJij

by Hooke's law, i.e.

(9.3)

Following Eq. 9.2, we may write the dilatation and the strain deviation tensor,
respectively, as
(P)
(94)
IOu-IOu
+Eu(e)

(P)
(e)
eij-eij
+eij

(9 5)

With the assumption that the plastic volume change is negligible, we have
E~)=O

and Eq. 9.4 becomes

(9.6)

359

Plasticity

(9.7)
Equations 9.3 and 9.7 imply that Hooke's law holds for the mean stress and the mean
strain at all times, provided the material is isotropic, i.e.
0;;

=3K;;

(9.8)

where K is the bulk modulus. However, Eq. 9.8 does not exist for the general case of
anisotropic materials. By the definition of the strain deviator, Eq. 2.47, we can write

ejj(P) -- ejj(P) - .!.3 elk(P)\:Ujj

(99)

Substituting Eq. 9.6 into Eq. 9.9 we find

er)=fft)

eV)

(9.10)

which allows us to use the notation eW) in places where we should write
to indicate
a strain deviation.
A common essential characteristics of all theories of plasticity is that there exists

a yield junction
(9.11)

which depends on the state of stress and strain and on the history of loading denoted
by the work-hardening parameter 1C. The parameter K is a function of the plastic strain
tensor eW). The equation f =0 represents a closed surface in the nine-dimensional stress
space, with axes 0jj (i,j 1,2,3). When a state of stress can be represented by a point
on the yield surface f =0, change in plastic deformation occurs. For a state of stress
with f < 0, the material is elastic (governed by Eq. 9.3), and there is no change in plastic
defonnation, i.e.
iI ..W)
""'ii _ '(P)
(9.12)
(it=e;; =0, f<O

The state of stress where f> 0 has no meaning, i.e. it is inadmissible. The states f =0,
f<O andf>O correspond respectively to points C, C1 and Cz in Fig. 9.4, which is an
illustration of one-dimensional stress.
At this stage, it is appropriate to consider the meaning of loading and unloading
in a plastic state, where the yield condition governs, i.e.
f(C1jj,fft), K) =0

(9.13)

The time derivative of the yield function is

J=

of

a. +
v

OOjj

where

of

O~)

~)+ of K

OK

(9.14)

Kis the time rate of 1C, which is a function of eW). Thus

OK

'(P)

K=~)e;j

(9.15)

Foundations of Solid Mechanics

360

Obviously, 1 = 0 and < 0 at a time t would imply 1 < 0 the next instant t + dt. Such
a change leads to an elastic state and corresponds to unloading. However, for such a
state, Eq. 9.12 applies, i.e. e~)=O. Hence, due to Eq. 9.15, Kmust also vanish. Thus,
for unloading, 1 =0 at its onset, and

. a/

I, <0

1 = - 0 '..

aO'ij

(9.16)

Otherwise, loading occurs, if


- 0 ' ..

al .
aO'ij !J >0, 1=0

(9.17)

al .
aO'ij IJ =0, 1=0

(9.18)

or neutral loading, if
- 0 ' ..

The yield function 1 is sometime called a loading junction because of its direct association with these loading criteria. Before making a geometric interpretation of these
criteria (and for reasons to be discussed later), we have to choose a sign convention in
writing Eq. 9.13 such that allaO'ij be directed along the outer normal to 'hypersurface'
1 = O. Then, for a state of stress on the yield surface; loading, unloading, or neutral
loading takes place, according to whether the stress increment vector is directed outward,
inward, or along a tangent to the yield surface. Since loading from one plastic state
must lead to another plastic state, 1 remains zero and so does its time rate

. _ a/ al .(p) al aK .(p)
1 = aO'ij O'ij + ae;~)eij + aKa~)fij =0

(9.19)

For this reason, Eq. 9.13 is called consistency condition by Prager (1948).
For ideal plastic solids, 1 is not affected by ~), thus during loading from a plastic
state, Eq. 9.19 gives

al .
aO' .. O'ij =0, 1= 0

(9.20)

IJ

A comparison of this equation with Eqs. 9.17 and 9.18 shows that loading from a plastic
state for ideal plastic solids is always neutral.
A material is said to be isotropic if the yield function 1 depends only on the
invariants (with respect to orientation of the reference coordinate system) of stress, strain
and strain history. If1 is an isotropic function of the stress alone, the theory is called
an isotropic stress theory for ideal plastic solids. In such theories
1=/(11,12'/3)
(9.21)
where Ii>

12

and 13 are the three invariants of stress tensor, i.e.


11

=0'0 =0'1 + 0'2 + 0'3

(9.22)

12

=21 (112 -

(9.23)

O'ijO'ji)

=0'10'2 + 0'20'3 + 0'30'1

361

Plasticity

(9.24)
in which 0"1> 0"2 and 0"3 are principal stresses.
As mentioned in Section 9.1, it has been found by Bridgman (1923) that the plastic
defonnation of metals is essentially independent of hydrostatic pressure. The result has
been verified by Crossland (1954) and many others. If this finding is accepted, the
yield condition will be independent of II and can be written as
1= 1(12)13)
(9.25)
where 12 and 13 are the second and third deviatoric stress invariants (Eqs. 2.39 to 2.41),
i.e.
(9.26a)

1
1
2
13 ="3SiiSjA,s"i =/3 -"3ll2+

2/1

(9.26b)

or, written as functions of differences in principal stresses, they are

1
2
2
:q
12 = 6{[(O"I - 0"3) - (0"2 - 0"3)] + (0"1 - 0"3) + (0"2 -0"3) J

(9.27)

1
3
3
13 = 81 ([2(0"1 - 0"3) - (0"2 - 0"3)] + [2(0"2 - 0"3) - (0"1 - 0"3)]

-[(0"1 - 0"3) + (0"2 - 0"3)]1

(9.28)

The yield surface 1(12,/3) 0 can be represented by a cylinder in a three dimensional


space with 0"1' 0"2' 0"3 as coordinate axes. The axis of the cylinder is equally inclined
to the coordinate axes, i.e. perpendicular to the so-called x-plane where 0"1 + 0"2 + 0"3 =O.
If 13 does not appear in I, the yield surface 1(/z) =0 in the principal stress space will
be a circular cylinder, i.e. its projection on the x-plane is a circle. If 13 appears in I,
then the cross-section of the yield surface 1(12)13) =0 on the x-plane will not be circular.
A yield surface in the three dimensional principal stress space is shown in Fig. 9.5,
while projection of two yield surfaces on the x-plane are shown in Fig. 9.6. Due to
Eqs. 9.27 and 9.28, a two-dimensional plot of the surface 1(12,/3) 0 with 0"1 - 0"3 and
0"2 - 0"3 as coordinate axes is also possible, as shown for example in Fig. 9.7.

9.3 VARIOUS YIELD FUNCTIONS

With permission 01 the Publisher, Prentice-Hall, a significant amount 01 material


in this section was adapted from the book by Fung (1965 J.
H. Tresca (1868) stipulated that the maximum shear stress, during the plastic flow,
must have the constant value k, the yield stress in simple shear. Thus the Tresca's

Foundations of Solid Mechanics

362

yield function is
(9.29)
where

0"1

is the maximum principal stress and

0"3

is the minimum principal stress, i.e.

0"1 ~ 0"2 ~ 0"3'

However, f in the form of Eq. 9.29 is not analytic; since it violates the rule that the
labelling of the principal axes 1, 2, 3 should not affect the form of the yield function.
To conform with this rule, we note that during plastic flow one of the differences
10"1 - 0"21 ,I 0"2 - 0"31 ,I 0"3 - 0"11 has the value 2k. Hence, we may put f in a form symmetrical with respect to principal stresses as
f= [(0"1-O"i-4e] [(O"z-0"3)z-4e] [(0"3-O"i-4e]

(9.30)

or, due to Reuss, as


f= 4Ji -27J; - 36eJi

+ 96k4JZ -64k 6

(9.31)

When the direction and relative magnitudes of the principal stresses are known by
symmetry consideration or by intuition, it is possible to use the simple form of Eq. 9.29.
Otherwise, we are obliged to use the general form of Eq. 9.31.
In 1913, Richard von Mises assumed that the yield function f is independent of
J 3 and takes the simple form
(9.32)
It can be shown that the yield criteria of von Mises and Tresca give the same
yield stress in simple shear; but, in simple tension, the yield stress by von Mises' criterion is smaller than that predicted by Tresca's, by a factor -f3/2. For no other type
of stresses is the discrepancy between the predictions of the yield stresses by von Mises
and by Tresca's criteria as large as that in simple tension (von Mises, 1913).
As illustrative examples of the yield surface f = 0 when f is not affected by
hydrostatic pressure, the von Mises' criterion is plotted in the principal stress space as
shown in Fig. 9.5, the von Mises' and Tresca's yield surfaces projected on 1t-plane in
Fig. 9.6, and the two-dimensional plotting of these two yield surfaces in Fig. 9.7.
A more general form of f, in isotropic stress theory for ideal plastic solids which
is unaffected by hydrostatic pressure, is
f=F(JZ,J3)-e

(9.33)

A simple example is
F

=Ji -

CJ;

(9.34)

A good correlation with Osgood's experimental results (1947) is obtained by taking


C =2.25.
To reveal and preserve initial anisotropy, we may use
(9.35)

Plasticity

363

Here Dij/d is composed of a set of 21 independent constants, due to the symmetry


conditions
Dij/d =D/dij =Dji/d

=Dijlk'

If incompressibility is assumed, the number of independent constants is reduced to 15.

When a metal rod is subjected to repeated tension-compression tests, the stressstrain curve sometimes appears as in Fig. 9.8. The tension stroke and the compression
stroke are dissimilar. This is referred to as the Bauschinger effect (J. Bauschinger,
1886). Von Mises' and Tresca's yield functions discussed in the preceding paragraphs
imply the initial isotropy, and the equality of tensile and compressive yield stresses at
all stages of the deformation, i.e. no Bauschinger effect.
To include Bauschinger effect but preserve isotropy, some yield functions have
been suggested for work hardening solids (by including ;<y~ as follows.
Axis is the line
0:2

Y OJ ="2=

'\

0'3

von Mises
Tresca

-------OJ
Fig. 9.5 A yield surface in the principal stress space.

Fig.9.6 Projection of yield surfaces on


the 1t-plane.

von Mises ; J2 =const.

Tresca;
maximum shear= const.

Fig. 9.7 Yield surface plotted on the


plane of (cr1 -cr3), (crZ -cr3).

Fig. 9.8 Bauschinger effect in a simple


tension-compression test.

Foundations of Solid Mechanics

364

/=F(JJ-msijr:$)-k",

(m=aconstant)

(9.36)

/=F(JZ,J3 )-msijr:$)-k"

(9.37)

/= F(JZ,!3) - H(Jz,J3)sijr:$) _kz

(9.38)

/=F(Jz,J3) - [P(Jz,J3)Sij + Q(Jz,J3 )tUJr:$)-k"

(9.39)

where tij is deviator of the square of the stress deviator, i.e.

2
OJ3
tij =su,S1r,i -3 Jl'ij =aa.

(9.40)

/ =Dijld(sij - mr:$~ (Sid - me~~ - k"

(9.41)

1/

Finally, the expression

yields the Bauschinger effect, and does not preserve the initial anisotropy during
defonnation.
It should be noted that each of these suggested fonns uses a single analytic function
to represent the entire yield surface. H a yield surface is composed of piecewise smooth
surfaces which meet to fonn comers, it would be convenient to use a separate expression
for each of these piecewise smooth surfaces. This concept was generalized by Koiter
(1953, 1960) and will be discussed in Section 9.4.
9.4 HARDENING AND FLOW RULES

With permission 0/ the publisher, Prentice-Hall, a significant amount 0/ material


in this section was adapted from the book by Fung (1965).
Work hardening in a simple tension experiment means that the stress is a monotonically increasing function of increasing strain, as illustrated by points C and C' on
the stress-strain curve in Fig. 9.9. In other words, when the stress 0' increases by a
small amount dO', the plastic state C (0', e) is transferred to another plastic state
C'(O'+dO',e+OO), and
(9.42)
dO'OO> 0
On removing the stress dO', the transfer will follow the elastic path C'F', and the system
will arrive at the elastic state C"(O',e+OO(P~. It is obvious from the figure that
dO'OO(P) > 0

(9.43)

which shows that only the elastic part of the strain, 00('), is recovered upon the completion of a loading cycle. The concept has been generalized by Drucker (1951) and
can be stated as: Suppose, at a point in the material domain with a state of stress O'ij
and strain Eu, an external agency increases the stress by dO'ij and produces a small strain
increment du. On the removal of the stress dO'ij, an elastic strain ~) is recovered.
Then, the material is said to be work hardening or strain hardening if the following

365

Plasticity

U'

c'

+ dU'

de:

<e)

Fig. 9.9 lllustration of work-hardening by a simple tension test.


two conditions hold true.
dcrijdEv > 0,
dcrij(dEv -~ == dcrij~) > 0,

upon loading

(9.44)

upon load removal

(9.45)

If we accept Drucker's defInition of strain hardening, we will say that there is

ideal plastic deformation when the equality sign prevails in Eq. 9.45, i.e.
dcrij~)=O

(9.46)

and the yield function is independent of ~). The differentials dcrij and ~) must be
interpreted as in Fig. 9.9.
In addition to having the yield function f to furnish a criterion when yielding
occurs, we need further information on the increment or rate of deformation in order
to complete our description of the material behavior. In other words, we need to derive
the plastic flow rule. Equation 9.20 may be recalled for this purpose and rewritten as

af

-dcr = 0
acrij 'I

(9.47)

A comparison of Eqs. 9.46 and 9.47 leads to

~)=A. af

acrij

(9.48)

366

Foundations of Solid Mechanics

where A. is an arbitrary constant of proportionality.


written in terms of the rate of strain as

Alternatively, Eq. 9.48 can be

f!t) =A ()aji
al

(9.49)

The sign of A. or A is restricted by the condition that plastic flow must involve
dissipation of energy, i.e.
(9.50)
Equation 9.48, the rule we are seeking for the plastic flow of ideal plastic solids, was
frrst given by von Mises (1928), and is known as the theory 01 the plastic potential.
An interesting geometric interpretation of Eq. 9.49 has been given by Prager
(1948). It is that the plastic strain rate ij') is a normal vector to the surface 1=0 in
the nine-dimensional stress space with eJji(i,j = 1,2,3) as coordinates.
It should be noted that a direct application of the flow rule in a one-dimensional
stress experiment could be difficult, since the plastic flow will continue when the stress
reaches its yielding limit, with no possibility of increasing the stress beyond that limit
To deduce some useful information from an experiment, it should be a combined-stress
system, such as a thin-walled circular cylinder subjected to a twisting moment and to
a tension or internal pressure. Such arrangement gives non-trivial changes of deJji and
deu, to which the preceding derivation applies.
Koiter (1953, 1960) generalized the flow rule by allowing the yield criterion to
be specified by several yield functions,
/"(eJji)' /z(eJji)'

... , !,.(eJji)

(9.51)

We have an elastic state if all these functions are negative. At a plastic state, at least
one yield function vanished. It is inadmissible to have any yield functions greater than
zero. If functions J,. = ... = I". = 0, whereas all other !'S are negative, the Koiter's
generalization of the flow rule (Eq. 9.48) is
'(P)

a/".

aJ,.

ij =A":\+ ... +A,,.:\


ueJji
ueJji

(9.52)

where A", ... , A". are arbitrary nonnegative constants of proportionality. Thus, the basic
concept leads at once to a general incremental stress-strain relationship in ideal plasticity.
For von Mises' yield function, Eq. 9.32, the flow rule, Eq. 9.49, gives
'(P)

ij =Asji

(9.53)

For Tresca's yield condition, the multi-yield functions may be chosen as


/,.=eJZ -eJ3 -2k,
/z=eJ3 -eJl -2k,
A=eJ1 -eJZ -2k,
(9.54)
For example, if /,. =0, while all other !' s are negative, then
'(P)-O
1 ,

;'CjJ)-A

-z -

l'

;'CjJ)--1.

"'3

:av

367

Plasticity

where Al > 0, and ft, Ez, ~ are the principal strains, the directions of which coincide
with those of the principal stresses under Tresca's condition of isotropy.
Following Drucker's definition of work hardening (or strain hardening) materials,
we can prove the following:
(1) The yield (or loading) surface must be convex.
(2) The plastic strain increment must be a vector normal to the yield surface at
a smooth point on the surface, and it must lie between adjacent normals to the yield
surface at a corner of the surface.
(3) The plastic strain rate must be a linear function of the stress rate.
In a short statement; the above three points are often referred to as convexity, normality,
and linearity, respectively. The statement is often interpreted as the requirement for
stable materials. The yield function and the yield surface assume the general fOnDS,
Eqs. 9.11 and 9.13, respectively. Moreover, the yield surface should be taken as a
hypersurjace (in the nine-dimensional stress space). The proof of these points can be
found in Naghdi (1960). Only a simple proof of linearity would be presented here, i.e.
the flow rule will be derived.
The normality of the plastic strain rate at a smooth point of the yield surface
implies that

acrii

ft=A af

(9.55)

where A is a function, in general, of stress, stress rate, strain, and strain history. Since
is associated with a positive work done during loading, we can show that A must
be nonnegative. Substituting Eq. 9.55 into Eq. 9.19, we can solve for A. Eventually,
Eq. 9.55 becomes

ij')

(9.56)
where

a = -[(a~ + ;~a~;)a::r

(9.57)

Equation 9.56 proves the linearity. These results were first obtained by Prager (1948)
and Drucker (1959).
For a yield surface being composed of a number (say n) of smooth surfaces!, 's
which meet to form corners, Koiter's generalization (1953, 1960) gives the plastic strain
rates as follows
(9.58)
where

Cp =0; if

t, < 0,

or

at, .
-a
cri cri < 0

(9.59a)

Foundations of Solid Mechanics

368

(9.59b)
and h/s are positive functions of stress, strain, and strain history. Conditions in Eqs.
9.59 specify, of course, whether yielding and/or loading occur as governed by Eqs. 9.16
to 9.18.
We may now wonder what possible fonns the plastic strain ~) can have in the
yield function f(Jij'~)' lC). Laws governing this aspect are called hardening rules. If
they are independent of the hydrostatic pressure, the yield surfaces in the principal stress
space are cylinders of infinite length perpendicular to the 1t-plane (Jl + (J2 + (J3 0, and
characterized by their cross-section on this plane. We know also that these crosssectional curves are closed, convex and piecewise smooth, but they could change in
size and shape as plastic defonnation proceeds. Several proposed hardening rules
characterized by the 1t-plane cross-sectional curves are illustrated in Fig. 9.10. In Fig.
9.10(a) Tresca's initial yield surface is shown as a regular hexagon on the 1t-plane. In
Fig. 9.1O(b) is shown the so-called isotropic hardening, which assumes a unifonn
expansion of the initial yield surface. Isotropic hardening, though conveniently used in
analysis, has little experimental support. Figure 9.1O(c) illustrates Prager's kinematic
hardening. The initial yield surface translates in the 1t-plane without rotation and
without change in size. None of the flow rules deduced from Drucker's hypothesis
is violated, the Bauscbinger effect can be represented very simply, and the development
of anisotropy due to plastic defonnation appears most naturally. With some variation
in the model, Prager (1954) was able to represent various models of plasticity: rigid
perfectly plastic, rigid work-hardening, elastic, etc. Hodge (1956) pointed out that the
concept of kinematic hardening must be applied in nine-dimensional stress space.
Modified kinematic hardening by Ziegler (1959) is illustrated in Fig. 9.10(d). In Fig.
9.1O(e) the plastic defonnation causes a linear segment to move. In Fig. 9.1O(t) the
plane loading surface changes with plastic loading in some independent manner. More
complicated hardening rules have been proposed from time to time, e.g. Hodge (1957)
has extended the kinematic hardening to include an expansion of the yield surface
simultaneously with its translation.
In contrast to the incremental theories ofplasticity discussed so far, there is another
group called deformation theories of plasticity. The difference lies in the flow rule.
The characterizing feature of the defonnation theory is the unique relationship between
the plastic strain and the stress deviator in the plastic state; there is no such uniqueness
in the incremental theories. The plastic strain as a unique function of the stress deviator
given in Prager's fonn is

F$)= PSij + Qtij'

where tij is as defmed in Eq. 9.40. For isotropic materials P and Q are naturally
functions of the invariants J 2 and J 3 It can be shown that Prager's fonn is the most

Plasticity

369

o
(0)

o
( c)

(e)

(0

Fig. 9.10 Several hardening rules: (a) initial yield condition (Tresca), (b)
isotropic hardening, (c) kinematic hardening (Prager), (d) kinematic
hardening (Ziegler's modification), (e) independently acting plane
loading surfaces, (f) interdependent plane loading surfaces. (From
Naghdi, 1960).
general fonn under the assumption that the plastic strain can be expressed in powers of
sij. Different choices of the yield function f(sij,rft), K), functions P and Q, and the
hardening parameter K lead to different theories of plasticity. Defonnation theories were
developed by Hencky (1924) and Nadai (1950, 1963) and their followers. Handelman,
Lin and Prager (1947) pointed out some conceptual difficulties of defonnation theories.
In some problems, a defonnation theory yields the same answer as the corresponding
incremental theory. Such is the case where in the loading process the ratio of the stress
components
eJ11 : eJ22 : eJ33 : eJ12 : eJ23 : eJ31
is kept constant at all times. For such problems the incremental flow rule can be
integrated to yield the defonnation rule. For a much wider class of loading, Budiansky
(1959) and Kliushnikov (1958) independently showed that the defonnation rule can be
derived from the incremental rule provided that 'comers' develop in the loading surface
as plastic defonnation proceeds.
9.5 INCREMENTAL FORMULATION FOR ISOTROPIC HARDENING
The work increment per unit volume in the transfer from the plastic state of stress
eJij and strain fv to another plastic state of stress eJij + deJij and strain ij + de;j (see the

Foundations 0/ Solid Mechanics

370

illustration of one-dimensional stress in Fig. 9.9) is


(9.60)

dW=O'iidEv

which may be split into elastic and plastic parts


dW =dW() + dW(P)
where

(9.61)

dW()

=O'ii~)

(9.62)

dW(P)

=O'iid4") =siid4")

(9.63)

In writing the last equation, we have used the condition of plastic incompressibility, Eq.
9.6.
For isotropic stress theory of materials obeying work-hardening hypothesis, the
yield function takes the fann
(9.64)
/=/(J,.,J 3, W(P~

where W(P) is the total plastic work, i.e.

w<P) = dw<P) = siid4")

(9.65)

As the plastic flow proceeds, the yield surface / (in the nine-dimensional stress space)
enlarges, as measured in the direction of its nonnal, by a distance
d/=

a/ dO'..

OO'ii

(9.66)

II

or, because of Eq. 9.55,


(9.67)

d/=Ad4")dO'ii
H the Prandtl-Reuss equation (Prandtl, 1924 and Reuss, 1930) holds, i.e.
d4")=XSii

(9.68)

d/=TdJ,.

(9.69)

Eq. 9.67 can be rewritten as


H T is a function of stress invariants, Eq. 9.69 shows that the yield function expands
unifannly, i.e. there is an isotropic hardening.
Introducing the equivalent or effective stress
as
= (31,.)112 = (3s iis/2)112
(9.70)

cr

cr

and, in a similar fashion, the equivalent or effective plastic strain increment dE}P) as

cJ?) =(~)d4")/3)112
we can have, due to Eq. 9.68,
d-(P)

=2Xcr/3

Furthennore, the plastic work increment dW(P) in Eq. 9.63 can be rewritten as

(9.71)

(9.72)

Plasticity

371

=2'1.<"0)2/3

dW(P)

(9.73)

=crd\P)

(9.74)
With the assumption of the independence of f from J 3 and the existence of a
universal plastic stress-strain curve relating cr and e(P), the total plastic work W(P)
becomes a single-valued function of ?), and Eq. 9.64 is simplified into

f=fT.cr,H(?~]
(9.75)
which is known as the strain-hardening hypothesis for isotropic strain-hardening
materials. Equations 9.56 and 9.57 for the present case can be written, respectively as

A~(P) -IT
_ A(af
-

aa

uc, ..

IJ

)2 acr

acr

aaij aa/cI

da/cI

(9.76)

(9.77)

where
,
H

Using Eq. 9.70, we get

dH

=de_(P)

(9.78)

acr 3s"",
aa"", ="2 cr

(9.79)

Furthermore, dividing Eq. 9.72 by Eq. 9.68 we obtain


a?) 2 cr
Thus Eq. 9.77 becomes

6= {

af

(9.80)

,af)-l

(9.81)

-H-=

aH aa

For materials obeying the Mises' yield condition,

f=cr-H
Thus Eqs. 9.81 and 9.76 become, respectively,
1

6=H'

(P) _ 9sijs/cIds/cI
de - ---=--IJ
H'(2a)2

f may be taken as
(9.82)
(9.83)
(9.84)

Due to Eq. 9.70, s/cIds/cI can be written as


(9.85)

Foundations of Solid Mechanics

372

Substituting Eq. 9.85 into Eq. 9.84 we fmd


3Siido

(P)

(9.86)

~ =2H'o

Numerical values of H' can be obtained easily as the slope of the stress versus plastic
strain curve in uniaxial tension.
As expressed in Eq. 9.3, the stress 0ii is related to the elastic strain ~) according
to Hooke's law, at all times. The same condition is applicable to their increments, i.e.
for isotropic materials,
(9.87)
where E, J.1 and v are Young's modulus, shear modulus, and Poisson's ratio, respectively.
However, the total strain increment is the sum of elastic and plastic increments, i.e.
dev=~)+~)

(9.88)

Substituting Eqs. 9.86 and 9.87 into Eq. 9.88 we fmd the total strain increment tensor
in terms of the stress increment tensor

1 - 2v
dsii 3siido
dev="3EdOkkSii+ 2J.1 + 2H'o

(9.89)

Multiplying Eq. 9.89 by sii' and using Eqs. 9.70 and 9.85, we find
Sk/dek/

--;;{1

1)

= OdvlH' + 3J.1

or
do

Sk/dek/

(9.90)

H'o= (Oil+~3,..
Substituting Eq. 9.90 into Eqs. 9.86 and 9.89 leads to, respectively,
(P) _

SiiSk/dek/

(9.91)

- 2(0)2 1 +~
3,..

(9.92)
Alternatively, the relationship between the stress increment tensor and the elastic
strain increment tensor can be put in the fann of the inverse of Eq. 9.87 as

ii

E
...I~(O)I: 2" rt ,,(o)
3(1 _ 2v) uc.kk uii + fAU<'ii

(9.93)

or, in view of Eqs. 9.7, 9.10 and 9.88,

(P)

dOii = 3(1 _ 2V)dekkSii + 2J.1deii - 2~

(9.94)

Plasticity

373

Substituting Eq. 9.91 into Eq. 9.94 we fmd the relationship between the stress increment
tensor and the total strain increment tensor
E

3J.1 s(iSk/OOk/

- v

(0')

dO'(i = 3(1 2) de".".S(i + 2J.1de(i -

2'

1 +!!.

(9.95)

311

Equation 9.92 or its inverse, Eq. 9.95, can be conveniently used in the numerical analysis
of boundary value problems, in which the loads (and displacements) are increased by
small amounts at each step of the analysis.
For ideal plastic solids, the yield function is independent of the plastic strain tensor,
thus Eq. 9.78 gives
H'=O
(9.96)
and, with the von Mises yield condition, the yield surface is
(j-...J3k =0
(9.97)
or
0' O'y
(9.98)

where k and O'y are yield stresses in simple shear and simple tension, respectively.
Substituting Eqs. 9.96 and 9.98 into Eqs. 9.91, 9.92 and 9.95 leads to, respectively,

~) =3s(isk/~

(9.99)

2cfv

1 - 2v

ds(i

3s(isk/ook/

(9.100)

deu=~dO'_S(i+ 2J.1 + 2cfv


E

dO'(i =3(1 _ 2v) de".".S(i + 2J.1de(i -

3JlS"sk/ook/

(9.101)

Equation 9.100 or its inverse, Eq. 9.101, can be conveniently used in the numerical
analysis of boundary value problems of ideal plastic solids.
For rigid-strain hardening, the elastic strain tensor is assumed to be negligible in
comparison with the plastic strain tensor, thus E and J.1 tend to infmity,

U =it)

(9.102)

and, due to Eq. 9.92,


(9.103)
In addition, Eq. 9.63 states that the scalar quantity Sk/~ in the equation above denotes
the plastic work increment, which is exactly equal to (jd?) as in Eq. 9.74. With the
help of the defmition of the equivalent strain increment in Eq. 9.71, we have
Sk/OOk/

=~~OOk/OOk/)1I2

(9.104)

374

Foundations of Solid Mechanics

Substituting Eq. 9.104 into Eq. 9.103 we obtain the stress-strain increment relationship
in the form
(9.105)
in which 0 is a function of the stress deviation tensor as in Eq. 9.70. The same ~e
of relationship for rigid-plastic solids can be obtained by simply setting 0 = cry = "'3 k
in Eq. 9.105.
9.6 VISCOPLASTICITY
Constitutive models of plasticity discussed so far can describe the plastic behavior
of materials, but cannot be used to deal with rate sensitive yield materials due to their
fundamental assumption of time independence. To overcome this difficulty, Bingham
(1922) introduced the concept of a viscoplastic model through a stress-strain relation in
a uniaxial context. Since then, many viscoplastic laws have been suggested. According
to these laws, the viscous properties of these materials introduce a time dependence of
the states of stress and strain, and the plastic properties make these states depend on
the loading paths. Thus different results will be obtained for different loading paths
and different durations of the process. Most existing constitutive laws of viscoplasticity
can be described by stress-strain incremental relationships based, to a certain extent, on
the elements of plasticity theory, i.e. the yield criterion, the flow rule and the hardening
rule. An extensive review of these was made by Hill (1950).
Many metals when subjected to uniaxial tension or compression have been found
to exhibit several rate effects. Among these the yield stress sensitivity to the rate of
straining is the most pronounced one. As early as in 1909, Ludwik: proposed an
empirical constitutive relation to describe such metals under tension as
cr=crylog(e(vp%iVP)
(9.106)
with e(vp) denoting the viscoplastic strain rate, the difference between the total strain
rate and the elastic strain rate, i.e.
e(vp)=e-e(e)
(9.107)
The elastic strain rate obeys Hooke's law, i.e. (for the uniaxial tension case)

e(e)=~
E

(9.108)

In addition, cry is the yield stress in uniaxial tension, and ~vp) is a material constant.
Obviously, Eq. 9.106 fails to describe the material behavior near the origin of the (cr,e(vP)
plane. Later Prandtl (1928) proposed the hyperbolic sine speed law; this has the
advantage of being valid at the origin and admissible in the compression region; it is
e{vp) = EiVP) sinh ~
cry

(9.109)

375

Plasticity

A general relation was proposed by Malvern (1951) in which the viscoplastic strain rate
is assumed to be a function of the current stress and strain, i.e.
e(vp) =f(cr, e)
(9.110)
Many forms have been proposed for the function f(cr,e). For example, Gilman (1960)
assumed
f(cr,e) =EEoexp(-A/cr),
where A and Eo are material constants.
Sokolovsky (1948) in his attempt to separate dynamic effects from strain hardening, introduced a linear viscoplastic law in which the viscoplastic strain rate is a
function of the dynamic overstress above the fIrst yield point. Mathematically, it can
be written as
(9.111)
Perzyna (1963) likewise proposed the following relation for the viscoplastic strain rate

e(vp)={

:y

-1)

(9.112)

While all these constitutive laws provide useful explicit relations for computing
viscoplastic strain rates, their applications generally come after the material stress state
exceeds a specifIc yield condition. The total strain rate can then be split into an elastic
part e(e) governed by Hooke's law and a viscoplastic part e(vp), given by one of these
laws.
It should be mentioned that there exist several one-dimensional constitutive laws
of viscoplasticity in the literature. An excellent survey made by Cormeau (1976) is
recommended to all who are interested in the historical development.
Following the derivation adopted in the one dimensional cases, several models of
three dimensional viscoplasticity have been developed by Perzyna (1966), Philips and
Wu (1973), Nicholson and Philips (1974), and others. SimplifIed versions of Perzyna's
model have been proposed by Zienkiewicz and Cormeau (1974), and Hughes and Taylor
(1978). A brief review of Perzyna's model and these two modifIed models for isotropic
solids will now be presented.
Perzyna's constitutive laws for elasto-viscoplastic materials under multi-axial
deformation assume the existence of an initial yield surface beyond which viscoplastic
flow will occur. The rate of increase of viscoplastic strain components is taken as a
function of the extent to which the stress state exceeds the static yield criterion, which
can be expressed by
F(crm,]2,J3)
f(cr ..) =
1
(9.113)
cry

I,

where f denotes the static yield surface, and

crm =3"cru

(9.114)

Foundations of Solid Mechanics

376

The sign of the yield functional F is taken such that 1 is negative for all elastic states,
positive for viscoplastic states, and zero for the incipient plastic states.
The total strain rate v is resolved into elastic ~) and viscoplastic Pii parts as

U =Eli + Pii

(0)

(9.115)

where
(9.116)
(9.117)
in which ).1, K and 'Y denote shear modulus, bulk modulus and a material dependent
fluidity parameter, respectively; the symbol < cl>(f) > defmes a switch-on/off operator
such that
(9. 118a)
< cl>(f) > = cl>(f), if I> 0
=0,
if I~O
(9. 118b)
For many practical cases cl>(f) is assumed as 1" with n being a positive integer depending
on the material under considemtion. Larger values of n characterize slower 'creep'
rates, but different values of n give the same final plastic solution. Using the chain
rule of differentiation to evaluate oFlocrii' we can transform Eq. 9.117 into

I of of
OF)
Pii='Y<cl>(f ( 3 ocr", aii+aJ/ii+aJ/ii

(9.119)

where tii is as defmed in Eq. 9.40. Equation 9.115 together with Eqs. 9.116 and 9.119
define the constitutive relations of Perzyna's model.
For von Mises materials, Zienkiewicz and Cormeau (1974) simplified Perzyna's
model with the following viscoplastic strain rates

Pii='Y<cl

01

ocrii

(9.120)

where

<cl=(tJ", if 1>0
=0,

if I~O

(9.121a)
(9.121b)

1 =f3];--(Jy

(9.122)

~ ~3 21312 2~ 2~l

(9.123)

Note that Pii is a symmetric second order tensor, and thus can be expressed in the fonn
of six distinct elements in a column matrix as
{P} =(1311

where the superscript T denotes the transpose of a matrix.


Hughes and Taylor (1978), on the other hand, proposed

377

Plasticity

{P} =3y< (I {s}/Gy


where the individual tenns are given by
{s} =[Sl1 s'J:}.

S33

SlZ

SZ3

(9.124)
(9.125)

S3J T

<(P>=/", if />0

(9. 126a)

=0, if /~O

(9. 126b)

and
(9.127)
The basic characteristic of all these models is that when a stress state at a point
falls outside the yield surface, they pennit viscoplastic 'creep' at that point (Fig. 9.11)
to occur, so the stresses are redistributed and relaxed till the stress state falls back on
to the yield surface. Further, when the stress state at no point exceeds the yield surface,
the solution by these models will correspond to the elastoplastic solution of the given
problem.
In both strain rate models described above, n equal to one was used~ It should
be note that the relations proposed by Zienkiewicz and Conneau (1974) have been well
accepted for the whole path of the viscoplastic 'creep', while those introduced by Hughes
and Taylor (1978) may not reflect the true behavior of such stress relaxation. However,
both models should give the same steady state or fmal solution.

A : />0 J Stress State after Instantaneous

Loading

B : / =0, Elasto - plastic State

Path of Viscoplastic I Creep

~~------------------~U2

~I
Fig. 9.11 Viscoplastic 'creep' on to three-dimensional von Mises yield surface.

CHAPTER X
FINITE DEFORMATION

10.1 DIFFERENT DESCRIPTIONS OF CHANGING CONFIGURATION


So far, we have restricted our analysis of defonnation to problems involving small
displacements only. Now we shall remove this restriction and consider finite displacements. The analysis involves the so called geometrical nonlinearity. Indicial
notation rules using subscripts to denote Cartesian components of tensors will be
followed from the present section until Section 10.6.
In Fig. 10.1, a particle at pO located by Xi in the original configuration is displaced
to a new position P located by Yi in the defonned configuration. The mapping from
Xi to Yi is one-to-one (Section 1.13), i.e. a particle originally at pO is displaced to a
single position P only. In other words, the current location Yi of the particle which

ORIGINAL

DEFORMED

nI

Fig. 10.1 Original and defonned configurations in the same ordinary


physical space.

379

Finite Deformation

occupied the point Xi at time t =0 characterizes the mapping of the initial configuration
into the current configuration (at time t) of the fonn
(10.1)
Yi =Yi(Xj , t)
In reverse, the mapping is in the fonn
Xi

=xi(Yj, t)

(10.2)

These two mappings are unique inverses of one another.


Coordinates Xi and Yi are nonnally called Lagrangian coordinates and Eulerian
coordinates, respectively. A description which uses Xi and t as independent variables
is called a material or Lagrangian description, while Yi and t are used as independent
variables in what is called a spatial or Eulerian description. Truesdell (1952) found
that the Lagrangian coordinates were introduced by L. Euler in 1762, and the Eulerian
coordinates by J. Ie R. D' Alembert in 1752!
As shown in Fig. 10.1, the displacement from pO to P is denoted by 14, i.e.
Yi =xi + ui

(10.3)

In addition, note that the vector dxi defining the infmitesimalline element pOQo in the
original configuration becomes dYi' defining the element PQ in the defonned configuration, thus

(10.4)
or, due to Eq. 10.3,
(10.5)
and
aYi

aUi

'Xj

'Xj

-a =Oij+-a

(10.6)

To be more complete it should be added that the displacement of point QO is Ui +dui ,


where
aUi

du=-dx
I

aXj

All vectors involved in the transfonnation of pOQo into PQ are illustrated in Fig. 10.2.
In the preceding paragraph, Yi and Ui are regarded as functions of Xi and time t.
Conversely, when Xi and Ui are regarded as functions of Yi and t,
aXi

dx =-dy.
I

aYj

(10.7)

(10.8)

Foundations of Solid Mechanics

380

dxj =8.. - dUj


dYj
"dYj

(10.9)

For a function A(yj,t), its partial derivative with respect to Xj is

dA dYk

-a
A(yj,t)=:\-a
!Xj
cJYk !Xj

(10.10)

If A in this equation is taken as Xj, we get the identity

8.. = dxjdYk
" dYkdXj

(10.11)

Similarly, we can show that


(10.12)

Fig. 10.2 One-to-one mapping of an infmitesimalline element between


two configurations.
10.2 MATERIAL DERIVATIVE AND CONSERVATION OF MASS

In fact Xj is truly independent of t, but Yj is not so. For a function A (Yj' t), its

material derivative (L. Euler, 1770 and J. L. Lagrange, 1783) is


d
dA dA
dtA(Yj,t)=-at+dYi Vj

(10.13)

dYj
v=
I
dt

(10.14)

where Vj is the velocity, i.e.

or, due to Eqs. 10.3 and 10.13


~

v=
I
dt

(10.15)

Finite Deformation

381

(10.16)
The acceleration <X; is the material derivative of Vi and can be obtained according to
the rule of Eq. 10.13 as
(10.17)
The ftrst and the second tenns in this equation are called the local and the convective
parts of the acceleration, respectively.
To lay a groundwork for the material derivative of the volume differential dV,
two very close points P and Q in the defonned conftguration at time t are considered.
Let Yi and Yi + dYi denote the position vectors for P and Q, respectively as shown in
Fig. 10.3. At time t + At (where At is sufftciently small), these two points assume,
respectively, the positions P' and Q'. Let the position vectors for P' and Q' be Zi and
Zi + dzi, respectively. According to the concept of one-to-one mapping, we can state
that
aZi
dz.=-dy.
(10.18)
aYj J
If Vi is the velocity at P, then the position vector for P' can be put in the form
Zi =Yi + viAt
(10.19)
Substituting the last equation into Eq. 10.18 yields

av

dzi =dYi + aYj dYjAt

(10.20)

To be more complete, it should be added that the velocity at Q is Vi + dVi' where

3
Fig. 10.3 Transfonnation of element dYi at time t into element dzi at
time t+At.

Foundations of Solid Mechanics

382

Ov
dv.=-'dy.
, iJYj J

(10.21)

All vectors involved in the transfonnation of the line element PQ into the line element
P'Q' are illustrated in Fig. 10.3. Next, consider a volume differential dV at P, which
we may choose for convenience to be a rectangular element as shown in Fig. 10.4, i.e.
dV =dy1dYzdY3
(10.22)
where dYI' dyz and dY3 are mutually orthogonal line elements characterized, respectively,
by the following vectors
s:
dYi(Z)-d
- YZuiZ

At time t

(10.23)

At time t+M

3
Fig. 10.4 Defonnation of a volume differential dV at time t into dV'
at time t + l1t.
Substituting Eqs. 10.23 into Eq. 10.20, we fmd the vectors characterizing the same line
elements at time t + l1t
(I)

dz,

(2)

dzi

(3)

dzi

s:
Ovi
=dYlu'l
+-dYIl1t
'iJYI
s:
Ovi
=dYZUiZ
+:;--dYzl1t
UYz

(10.24)

Ovi
l1t
=dY3Us:i3 +:;--dY3
UY3

In general, dz?), dz?) and dzf3) are no longer mutually orthogonal, but rather fonn a
parallelepiped as shown in Fig. 10.4. The volume of the latter can be expressed,
according to the law of products of vectors (Eqs. 1.5, 1.6, 1.31 and 1.32), in the fonn

383

Finite Deformation

dV' =eu"dzi(l)dzj(2)dzr)

(10.25)

where V" is the pennutation tensor defmed in Eq. 1.33. On the other hand, the basic
defmition of the material derivative of dV is
d(dV)

=Lim dV' - dV

(10.26)
Ilt
Substituting Eq. 10.25 into Eq. 10.26, using Eqs. 10.22 and 10.24, and neglecting the
remaining tenns of Ilt we obtain

dt

Ar-+O

d(dV) =av i dV
dt
0Yi
For the material derivative of an integral in the fann

(10.27)

L A (Yj,t)dV,

where V is the domain of the continuum at time t, the operations of differentiation and
integration may be interchanged, since the differentiation is with respect to a defmite
portion of the continuum (i.e. a specific mass system). Therefore

L A (Yj,t)dV

=L ! [A (Yj,t)dV]

which, upon carrying out the differentiation and using Eq. 10.27, results in

rr

d
[dA(yj,t)
av i ]
dt Jv A(Yj,t)dV =Jv
dt +A(Yj,t)OYi dV

(10.28)

Applying Gauss's divergence theorem (Eq. 1.87) into Eq. 10.28 we fmd

d
[dA(yj,t) oA(Yj,t)
dt Jv A (Yj,t)dV = Jv
dt
0Yi Vi dV + Js A (Yj,t)vinidS

(10.29)

where S is the surface enclosing V and ni is the unit outward. nannal vector to S. With
the help of Eq. 10.13, we can rewrite Eq. 10.29 in the fann

d
oA(Yj,t)
dtJvA(Yj,t)dV= Jv ot dV+ JsA(Yj,t)vinidS
Without the assumption of the spatial differentiability of A (Xj,t) and
derivative above can be considered as

~l A (xj,t)dV --.Lnn a[LA(Xj,t)dV]


Ilt
dt v

Ar-+O

(10.30)
Vi'

the material

(10.31)

The numerator on the right-hand side of the equation above can be considered as
consisting of two tenns; one over the region that V(t) and V(t + Ilt) share in common,
i.e.

oA IltdV
Jv ot
'

Foundations 0/ Solid Mechanics

384

and another the volume swept by particles occupying an element of area dS, i.e.

A (xj , t )vjAtnjdS .

Substituting these two terms into Eq. 10.31, we can readily obtain Eq. 10.30.
If A(Yj,t) in Eqs. 10.28 to 10.30 is taken as a unit scalar quantity, we have the
change of the material volume with respect to time, i.e. the material derivative of the
volume, in the forms

-d
dt

iy i

dV=

-'dV
yaYj

dv

(10.32)

vjnjdS

(10.33)

Note that Eq. 10.32 is equivalent to Eq. 10.27. If A in Eq. 10.30 is replaced by the
material mass density p, then the left-hand side of that equation must vanish because
of the conservation of mass, i.e.

rap dV + Jsrpv.n.dS
=0
Jyat
JJ

(10.34)

Applying the same technique to Eq. 10.28, but dropping the integration symbols there
(since the equation holds for any V), we fmd
dp
dvj
-+p-=O
dt
aYj

(10.35)

or, due to Eq. 10.13,


(10.36)
Equations 10.34 to 10.36 are called the equations of continuity (J. Ie R. D' Alembert,
1752 and L. Euler, 1757). The integral form (Eq. 10.34) is useful when the differentiability of PVj cannot be assumed.
In the one-to-one correspondence between the original and the deformed
configurations as depicted in Fig. 10.1, another basic form of the conservation of mass
is

pOdVO

=pdV

(10.37)

where pO and VO are the mass density and the material domain at the original state. For
convenience, we may choose dVo as a volume of an infmitesimal rectangular element
of the form
(10.38)
where dxh dx2 and dx3 are mutually orthogonal line elements in the original configuration, and characterized respectively by the following vectors

385

Finite Deformation

(10.39)

dx?) =dx30i3
Substituting Eqs. 10.39 into Eq. 10.4 yields, respectively, the vectors characterizing the
same line elements in the deformed configuration as follows
d ( l ) -dy
-'dx
Yi - dXl 1
(10.40)

dy
d (3)-_'dx
Yi - dX3 3
In general, the line elements dyP), dy?) and dy?) are not mutually orthogonal. In other
words, an original rectangular element of volume dVo transforms into an element of a
parallelepiped shape of volume dV. The latter can be put in the form
dV =f\jkdYP)dyy)dy~3)
(10.41)
Due to Eqs. 10.38 and 10.40, Eqs. 10.41 and 10.37 become, respectively
dydydy
dV=fvk-'-'-"dVO
dXl dXz dX3

(10.42)

or
o

dYidYjdYk

=ijk dXl dXz dX/

(10.43)

Before proceeding further, we should recall the following properties of the determinant
(Eqs. 1.46 and 1.48)
1

dY, 1 =f\'k dYi dYj dYk


dXs
v dXl dXz dX3

1 dX, 1 = ..
11M

dy.

dXi dXj dXk


!}kdYldYmdY"

(10.44)
(10.45)

Substituting Eq. 10.44 into Eq. 10.43 we fmd

pO =1 dYi 1
p dXj

(10.46)

Foundations of Solid Mechanics

386

However, Eq. 1.45 which is another property of determinants implies that

I~:; II ~~: 1=1 ~:;~ I,


which, due to Eq. 10.12, becomes

I~; II ~~: I=I8vI =1

(10.47)

Consequently, Eq. 10.46 can be put in an alternative form as follows

:0=1 ~~: I

(10.48)

Equations 10.46 and 10.48 relate the mass density ratio to a unique transformation from
one configuration to another as described in Section 1.13. Moreover, since the mass
densities are naturally positive definite, the transformation is admissible and proper.
10.3 STRESS TENSORS IN DIFFERENT DESCRIPTIONS
The equation of equilibrium, Eq. 2.8, obtained in Chapter IT corresponds to a
deformed configuration and is simply expressed in terms of the Eulerian stress tensor
crv. This is a natural physical concept. However, if strains are referred to the original
configuration, it would be more convenient in the course of analysis to define stresses
with respect to the same configuration.
As has been established in Eq. 2.6, a stress vector Xi on a small area dS in a
strained continuum is related to the Eulerian stress tensor crv by
XidS

=crjinjdS

(10.49)

where ~ is the unit outer normal to dS. In the undeformed state, imagine the force
XidS acting on the area dSo which is the original of dS. Simultaneously, a stress vector
in the same direction as Xi' and a stress tensor <fv may be introduced in a relationship
similar to Eq. 10.49, i.e.

X?

-;;()

.-0
XidS =ujinjdS

(10.5 0)

where nio is the unit outer normal to dSo. Known as the Lagrangian stress tensor or
the first Piola-Kirchhoff stress tensor (G. Piola, 1833 and G. Kirchhoff, 1852), <fv is a
stress component in the xrdirection on a plane (in the unstrained continuum) perpendicular to the Xi-axis. Equating the last two equations we find

crjinjdS

=cr;njOdSO

(10.51)

In order to fmd the relationship between crv and <fv, we must know the relation between
njdS and njOdSo. Let dYi and 8Yi be the sides of the small parallellogram dS. According
to the law of a cross product of two vectors (Eqs. 1.6 and 1.~2), we can write
(10.52)
nidS =Evkdyj8Yk

387

Finite Deformation

Another equation similar to Eq. 10.52 can be written for the original state as follows
niodSo = Eijldxjax l

(10.53)

where dx i and ax i are the originals of dYi and BYi' respectively. Therefore, by Eq. 10.7,

aXj aXl

n dS = P. .. L---dy/By
"~ay/aYm

Multiplying the last equation by


get

(10.54)

ax/ay,. and making use of Eqs. 10.45 and 10.48, we


(10.55)

or, due to Eq. 10.52,


aXiO
p
-a
n i dS =(jn,.dS
y,.
p

(10.56)

Substituting Eq. 10.56 into Eq. 10.51 we fmd

poaxm
cr--n
,. p aYj m dS = crndS
,. J
or, (since njo is nonzero),

pO aXj

cr=--cr .
,. p aYm ""

(10.57)

Equation 10.57 shows that the Lagrangian stress tensor ~ is not symmetric, though the
Eulerian stress tensor crij is symmetric, as will be shown formally later. Thus it will
be inconvenient to use ~ in a stress-strain law which is normally in terms of symmetric
stress and strain tensors. For this reason, l:ij known as the second Piola-Kirchhoff stress
tensor or simply the Kirchhoff stress tensor is introduced as
(10.58)
or, due to Eq. 10.57,
(10.59)
From the last equation we can see that l:ij is a symmetric tensor, i.e. l:ij = l:ft. Making
use of the identities in Eqs. 10.12, we can obtain the inverses to Eqs. 10.57 to 10.59
as the following
(1O.60a)

aYi
cr=-l:
,.
aXm Jm

(l0.60b)

Foundations oj Solid Mechanics

388

p dYidYj

aft

Due to Eqs. 10.6 and 10.9, I.ft and

=pOdXI dX", I./m

aft

(l0.60c)

can be expressed in tenns of displacements

I.ft = pO[aft - (Sj", ~Ui + Sa ~Uj _ ~Ui :Uj )a/m]


p
YI
y", OYIOY",

(10.61a)

(10.61b)
At this stage, it is appropriate to consider the directions of the stress vector
components on each side of the infmitesimal parallelepipeds. As shown in Fig. 10.5,
a stress vector a~i acting on the plane (in the original configuration) nonnal to Xl -direction has Cartesian components ~l' ~2 and ~3' Alternatively, the same stress vector
due to Eq. 1O.60b, is
o
dYi
dYi
dYi
ali =I.u -a +I.12 -a +I.13 -a
!Xl

!X2

!X3

(10.62)

It should be noted that each vector on the right-hand side of Eq. 10.62 is parallel to a
particular edge of the infmitesimal parallelpiped in the defonned configuration. For
example, the vector I.lldy/dxl is parallel to dyP> (defined in Eq. 10040), since both
vectors are parallel to the same vector dy;ldxl' In fact, an original edge in the Xl-direction
and with a unit length becomes a vector dy;ldxI in the defonnedconfiguration. Similarly,
we can say that the shear components ~2dy;ldX2 and I.13dy;ldx3 are parallel to dy?)

ORIGINAL

DEFORMED

Fig. 10.5 Decomposition of stress vector ~i into three components parallel


to edges of each parallelepiped.

389

Finite Deformation

and dyP), respectively. Thus the stress vector on the side originally perpendicular to
xI-axis can be decomposed into three components as shown in Fig. 10.5. Repeating
the same discussion for other sides of the parallelpiped, we can show that "Lnay/axz is
parallel to dy?), ~ay/aX3 to dyP), ~Iay/axi to dy;(I), ~3ay/aX] to dyP), ~Iay/axi to
dyP), and ~zay/axz to dy?).
Such knowledge of the directions of the stress vector components is useful in
interpreting the phenomena physically and in making rational assumptions in real
structural systems. As an example, taking the case in which the original dimension of
the structure in the x3-direction is relatively small, we may assume that ~3' ~I and ~2
are zero, since that gives the same effect as the usual assumption that stress vector
components in the dyp>-direction should vanish.
Instead, we choose the current volume dV as a volume of an infInitesimal rectangular element of the form
(10.63)
where dYI' dY2 and dY3 are mutually orthogonal line elements in the deformed configuration, and characterized respectively by the following vectors
(lO.64a)
(lO.64b)
(I0.64c)

Substituting Eqs. 10.64 into Eq. 10.7 yields, respectively, the vectors characterizing the
same line elements in the original configuration as follows
ax;
=-dYI
aYI

(1O.65a)

(2)
ax;
dx =-dY2
I
ayz

(1O.65b)

(I)

dx
I

(3)

dx;

ax;

=-a
dY3
Y3

(1O.65c)

Two more stress tensors may be introduced as


00

ay;
aXk

0" .. =-O"k

(10.66)

o ay; aYj
Eji =
0"1m

(10.67)

JI

and

ct:

-a-a
Xm
Xl

Note that
is not symmetric but l:.~ is. Making use of Eq. 10.12, we can obtain the
inverse of Eqs. 10.66 and 10.67 as the following

Foundations of Solid Mechanics

390

(Tt2

DEFORMED

ORIGINAL

Fig. 10.6 Decomposition of stress vector (Ju into

three

components parallel

to edges of each parallelepiped.


(10.68)
(10.69)
It should be interesting to see the direction of each component of a stress vector acting
on a plane in the deformed configuration which is normal to a Cartesian coordinate
axis. As an example, Fig. 10.6 depicts the stress vector (Jli acting on the plane (in the
deformed configuration) normal to the Yl -direction. This stress vector has Cartesian
components (J11' (J12 and (J13' but due to Eq. 10.68,
(10.70)
of which each vector on the right-hand side is parallel to a particular edge of the
infmitesimal parallelepiped in the original configuration, e.g. a:taX/aYl is parallel to
dxP) (given in Eq. 1O.65a).
While (Jij and cfv are physical by their direct association with the stress vector Xi;
I:ij' cfij and I:~ are pseudo except when we restrict ourselves to small deformation, for
which we shall show that
(Jij

=(J~ =I:ij =cr': =I:~,

(for small deformation)

For an arbitrary function A(Yi), we have, because of Eq. 10.6,

aA

aAaYj

-=--

(10.71)

Finite Deformation

391

(10.72)
in which we may neglect the tenn

aA aA
-a
=:;-,
!Xj UYj

au/aXj,

and thus arrive at the result

(for small defonnatton)

(10.73)

Using Eq. 10.73, we can readily obtain Eq. 10.71 from any of the Eqs. 10.58 to 10.60
and 10.66 to 10.69.
lOA EQUATIONS OF MOTION IN DIFFERENT DESCRIPTIONS
In Fig. 10.1, a continuum at time t occupies a domain V with a boundary S in
the defonned state. The body is subjected to a body force Xj per unit mass and a
surface traction Xj per unit area of S. Newton's second law of motion states that the
rate of change of the linear momentum is equal to the applied force, i.e.
d
(10.74)
dt (pvjdV) =dFj
where dFj is the resultant force acting on dV. Note the identity
EukVjVk =0

(10.75)

(since the tenn represents a cross product of two identical vectors), and the rate of
change of the moment of momentum
d
d
(10.76)
dt (E;jkYjpvkdV) =E;jkYj dt (pvkdV) + EukVjvkPdV
Substituting Eqs. 10.74 and 10.75 into Eq. 10.76 we find
d
dt (ijkYjpvkdV) =dTj

(10.77)

where dTj is the resultant torque acting on dV, i.e.

dTj =ijkyjdFk
Writing Eqs. 10.74 and 10.77 for V as a whole, we have, respectively,

(10.78)

!IvPVjdV=Fj

(10.79)

(10.80)

Iv ijkYjpvkdV =Tj

where

Is :KjdS
T j =Iv ijkPYrkdV + Is E;jkYrkdS

F j =Iv pXjdV +

(10.81)
(10.82)

392

Foundations of Solid Mechanics

Equations 10.79 and 10.80 are Euler's first and second laws of motion, respectively,
already described in Section 2.4. To derive the equations of motion in terms of the
Eulerian stress tensor aij in a more formal way using the frrst law, we substitute Eq.
10.81 into Eq. 10.79 and use Eqs. 10.30 and 10.49 to obtain

i [ at
v

a(PV;)
---pX.
dV=

Iss(a .. -pvjv.)n.dS
J

JI

(10.83)

If aji and pVjVj on the right-hand side of the equation above are assumed differentiable
with respect to Yj' we can apply Gauss's divergence theorem and obtain

i[ a~;j)

- PX]dV

i[

= ~~ a(~~Vj) JdV.

Since this equation holds for an arbitrary V, the integrands on both sides must be equal,
i.e.
(10.84)
or, due to Eq. 10.13,

aaji

~+pXj=p

uYj

dVj
rap a
~
dt +Vj -:\+:;-(pVj) .
ut

uYj

However, the last term in this equation must vanish due to an equation of continuity,
i.e. Eq. 10.36. Thus
aaji
dVj
aYj + pXj =P dt
(10.85)
This is the Eulerian equation of motion in the differential form, while Eq. 10.83 is that
in an integral form. Alternatively, when the spatial differentiability of p and Vj can be
assumed, substituting Eq. 10.81 into Eq. 10.79 and making use of Eq. 10.28 yield

r[d(PVj)

av j

Jv ~+PVjaYj -pXj dV= Js ajinjdS,

or

(10.86)

If aji is also differentiable with respect to Yj' the equation above can be reduced to Eq.
10.85.
On the other hand, if Euler's second law of motion is used, we will obtain

r [

JVijk PVjVk-Yj

(aa/k
dVk]
ay, +pXk-P"dt dV= JVilka/kdV,

393

Finite Deformation

from which, due to Eqs. 10.75 and 10.85, we can obtain


Eilka/k =0.

The last equation simply means the symmetry of the Eulerian stress tensor, i.e. aij =aji.
To adopt Eq. 10.85 as a governing equation of related boundary value problems
makes the task difficult, since the variable Yi itself is unknown, and both the velocity
and the acceleration defmed in Eqs. 10.16 and 10.17 contain nonlinear terms. The
problems become simpler when the deformation is considered so small that: these
nonlinear terms may be neglected; there is no need to distinguish Xi from Yi; the basic
unknown in the problem is Ui'
In the Lagrangian description, the functions involved have Xi and t as independent
variables, thus the velocity and the acceleration are

dUi(Xj,t)
vi(xj,t) = dt

(10.87)

dvi ifUi
Cli(Xj,t)=-at= dt 2

(10.88)

The linear momentum of the system is

Iv

Ii = pvidV =fyO pOvidVo

(10.89)

Its material derivative in the Lagrangian description is


dli
dt =

J.

yOp

odvi
-at dV

(10.90)

Noting the fact that the same body force Xi per unit mass acts in both dV and dVo, and
that the same surface force (Eq. 10.50) acts on both dS and dSo, we can change the
resultant force acting on the system (Eq. 10.81) into

J. pOX;dVo + f. a~njOdSO

Fi =

SO

yO

(10.91)

Equating Eq. 10.90 with Eq. 10.91 we fmd

=f. a~.n?dSO
J. [podvidt - pOXJdVO
'
yO

soP}

(10.92)

If the spatial differentiability of the stress field can be assumed, we can transform the
surface integral in Eq. 10.92 into a volume integral by Gauss's divergence theorem, and
finally get the equation of motion for an arbitrary domain VO as

ddfi

odvi

-+
P X =P dXj
,
dt

(10.93)

Alternatively, we can obtain Eq. 10.93 and the symmetry of l:ij by considering the
equilibrium of the infmitesimal parallelepiped after deformation, in the same way that

Foundations of Solid Mechanics

394

we do for curvilinear coordinates. The equations of motion can be expressed in terms


of the Kirchhoff stress tensor by substituting Eq. 1O.60b into the last equation:

a (aYi

aXj aXm:Ejm

0
odvi
+ P Xi =P at

(10.94)

or, due to Eq. 10.6,


(10.95)
10.5 FINITE STRAIN TENSORS
In Fig. 10.1, the square of the line element connecting two neighboring points P
and Q in the displaced configuration is dYidYi, while that in the original configuration
is dxidxi. The difference between these two squares, because of Eq. 10.4, is
(10.96)

or, because of Eq. 10.7, is


(10.97)
where
(10.98)
(10.99)
The tensor Eij now known as the Green's strain tensor was introduced by O. Green in
1841 and by B. de St.Venant in 1844. The tensor fv now known as the Almansi's
strain tensor was introduced by A. L. Cauchy in 1827 for small strains, and by Almansi
(1911) and Hamel (1912) for finite strains. In other words, Eij is a tensor in the
Lagrangian description, while E;j is one in the Eulerian description. Both tensors are
symmetric, since
Eij

=Eft,

Eij

=Eft

(10.100)

hence there is at least one set of principal axes for each of them. Another important
consequence of Eqs. 10.96 and 10.97 is that: a necessary and sufficient condition that
a deformation of a body be a rigid-body motion (consisting only of translation and
rotation without changing distances between particles) is that all the components of the
strain tensor Eij or E;j be zero throughout the body.
With the use of Eqs. 10.3 and 10.6, the strain tensors can be expressed in terms
of the displacement Ui as
(10.101)

395

Finite Deformation

e.. =!(auj + aUj _ aUk aUk)


IJ

2 aYj

ayj ayj

aYj

(10.102)

Alternatively, we can write Eij and F.jj as


(10.103)
(10.104)
where
(10.105)
(10.106)
when the displacement Uj is regarded as a function of the Lagrangian coordinates Xj,
and
(10.107)
(10.108)
when the displacement Uj is regarded as a function of the Eulerian coordinates Yj.
The strain energy function U, per unit mass, can be defined as for small deformation (Eqs. 3.1 to 3.7). In the Lagrangian description, the following can be derived
pOdU =cr~d(au/ax)

(10.109)

which is analogous to Eq. 3.4. Substituting into the equation above Eq. 10.3 for Uj and
Eq. 1O.60b for cr~, we get

396

Foundations of Solid Mechanics

=~.

!jdYidYi )
2,\dXm dXj

:fill

=~.

!jdYidYi_3.)
2,\dXm dXj :fill

(10.110)

:fill

Replacing the term in the parentheses in the equation above by Ejm as defmed in Eq.
10.98, we get
(10.111)
which is equivalent to Eq. 3.7 for small deformation. If the material is assumed to be
elastic, i.e. when the stress components are unique and homogeneous functions of the
strain components, there exists the strain energy function U as a scalar homogeneous
function of the strain components, i.e.
U =U(Eij)
(10.112)
Accordingly, its total differential is

dU
dU=-dE ..

dnv "

(10.113)

In Eqs. 10.111 and 10.113, each of the nine strain components are considered as an
independent variable in the function U. Comparing these two equations, we have

odU

,t.. =p (10.114)
dEij
"
which is equivalent to Eq. 3.10 for small deformation.
Thus we may conclude that a rational form of the constitutive law is the relaand the Green's strain Eij.
tionship between the Kirchhoff stress
The small strain tensor as obtained in Chapter II can be specialized from Eq.
10.102 by restricting ourselves to small derivatives of displacements, i.e. dU/dYj 1,
and consequently neglecting their products in that equation. Eventually, we get

.tv

Eij =21 (dU'


dY~ + dU')
d~ , (for small deformation)

(10.115)

and Eqs. 10.102 and 10.107 become identical. With the additional help of Eq. 10.73,
we can conclude that there is no need to distinguish Eij from Ejj, Ev or ~ in small
deformation.
10.6 REFORMED LAGRANGIAN DESCRIPTION
If a transformation function

(10.116)
or, reversely,
(10.117)

397

Finite Deformation

is established properly, an undefonned continuum element can be refonned into an


element with a simpler shape and more convenient dimensions. So the new description,
using Zi and t as independent variables, may be called a reformed Lagrangian
description. The reformed Lagrangian coordinates Zi are sometimes called nonnal
coordinates; they are independent of t just as the Xi' The one-to-one correspondence
with the Eulerian description may be written
(10.118)
or, reversely,
Yi

=Yi(Zj' t)

(10.119)

Similar to Eqs. 10.10 to 10.12, we can derive the following


d

oA dXk

OZi

!Xk oZi

::IA(xj,t)=-a: I

(10. 120a)

(10. 120b)

(l0.120c)

(l0.12Od)
dXi dZk

OZi dX k

OZk OXj

= ij =OXk OZj

0Yi OZk

OZi OYk

(10. 120e)

- - = ( ) .. = - OZk oYj

IJ

0Yk OZj

(10. 120f)

(10. 120g,h)
Denoting the refonned Lagrangian description by the superscript r, and adopting the
same procedures as used in the preceding sections to derive various conditions in the
correspondence between two different descriptions, we can get the one-to-one mapping
of a vector defining an infmitesimal line element as
(l0.12Ia,b)
or, reversely,
OZi

OZi

dz. =-dx =-dy.


I

dXj

oYj

(1O.121c,d)

Foundations of SoUd Mechanics

398

The one-to-one mapping of a volume and the conservation of mass are

~: =:~: =1 ~:; 1=1 ~~: r


l

p' = dV =1 oYj 1=1 OZj I-I


p dV'

OZj

oYj

(10.122a,b,c)
(1O.122d,e,f)

The one-to-one mapping of a surface is

oz.
po
- ' n~dS' =-n~dSo
OXj'
p.1

OZj

(1O.123a)

(1O.123b)

-ndS =-ndS
oYj ,
p' 1
The equality of force vectors on two corresponding surfaces gives
=r

~
=XjdS
XjdS
=XjdS

(l0.124a,b)

The stress vector formulas are

Xj =vjinj' Xj =ajinj ,

00-

=r-,.~

Xj=ajinj

(10. I 25a,b,c)

The reformed Langrangian stress tensor is

cf.. =p' OZj a O

pOox",

JI

(10. 126a)

l1li

p' OZj
cf..=--a.

pay",

JI

(l0.126b)

l1li

The reformed Kirchhoff stress tensor is


OZj
p' OZj OZj
:Eij =0Yl <fa = 0Yl oY", a ml

p'oz.' _oz.
'IJ. =__
1 E
IJ
po OXI ax", ml

(10. 127a,b)
(10. 127c)

The equation of motion is

Oafi.
.Ovj
-+pX.=pOZj
,
at
or

J.

(OYj ~
OZj OZ'" ~j",

(l0.128a)

Ovj

+ P Xj =p at
r

(l0.128b)

The strain energy function U is

p'dU=EqdEij
where Eij is the reformed Green's strain tensor, i.e.

(10.129)

Finite Deformation

399

E~. =!(dYkdYk _ dXkdXk)


"

2 dZj dZj

dZj dZj

dXkdXm
=---Ekm
OZj OZj

(10. BOa)
(10. BOb)

Thus another rational fonn of the constitutive law is the relationship between the
refonned Kirchhoff stress tensor ~ and the refonned Green's strain Eij. For linear
elastic solids, Hooke's law applies, and relationships in the Lagrangian and refonned
Lagrangian descriptions take the following fonns
(1O.131a)
(10.131b)
Making use of Eqs. 10. 120e, 10. 127c, 1O.130b and 10.131, we get the relationship
between the stiffness tensors in these two different descriptions as

r pr dZj dZj OZk dZ1


Coo = - - - - - C
"Id
pO dXm
dXi ldX
dXp 11I1IOp
o

(l0.132a)

az-az- dZ az-Cmnop

(l0.132b)

pO dXj dXj OXk dXl

Cijld =pr

rn

II

An eight-noded three dimensional element, which may not be rectangular, can be


transfonned into a rectangular element with 2 x 2 x 2 dimensions by means of the following transfonnation
x
I

=a=l
L N(a)(z.)x~a)
J

(10.133)

where the superscript a in parentheses denotes a node number, and N(a)(zj) is an


interpolation function taking the fonn
N(a)(z) =4(1 + ZlZ~a~ (1 + Z2Z~a~ (1 + z3zia~

(10.134)

In the equation above, zla) is the position vector of the node a in the refonned Lagrangian
description, and can be expressed as
z(a)--1
for a = 1,4,5,8
(l0.135a)
1 -,
= 1, for a = 2,3,6,7

(10. 135b)

z(a)=_l
2
, for a = 1,2,5,6

(10. 136a)

= 1, for a = 3,4,7,8

(10.136b)

z(a)=-l
3
, for a = 1,2,3,4

(10. 137a)

= 1, for a = 5,6,7,8

(l0.137b)

Foundations of Solid Mechanics

400

Such a refonned eight-noded three-dimensional element can be put as in Fig. 1O.7(a).


It can be said that each component of the position vector Xi inside the element can be
obtained by interpolation among its nodal values using Eq. 10.133. The same interpolation function can be used for the displacement vector u; also, i.e.
g

ui = I. N(a)(z)u?)

(10.138)

a=l

To avoid some difficulties involved in the treatment of finite rotations in large


defonnation analysis of shells, Kanok-Nukulchai, Taylor and Hughes (1981) assigned
nodes 1 to 4 to the shell midsurface and nodes 5 to 8 to the shell top surface, and
proposed the following interpolation function

N(a)(z) =f(zj' z}",

=z,J(Zj' z}"~,

(a) Three dimensional element.

for a

=1 to

for a

=5

(10. 139a)

to 8

(10. 139b)

(b) Shell element.

Fig. 10.7 Refonned eight-noded continuum elements.


where
(10.140)
Moreover, xla ) and u?) for a =5 to 8 were assigned relative quantities with respect to
nodes a - 4; e.g. while X~2) is Xl of the node 2, X~6) is equal to Xl of the node 6 minus
that of the node 2. It should be noted that z~a) and z~a) are as given by Eqs. 10.135 and
10.136 respectively, while Eqs. 10.137 are replaced by the condition 0 ~ Z3 ~ 1. Figure
10.7(b) depicts such a refonned eight-noded shell element.

Finite Deformation

401

10.7 STRAIN TENSORS IN CURVll.INEAR COORDINATES


Notations and rules developed in Sections 1.14 and 1.15 will be followed. Figure
10.8 depicts the geometry of an infmitesimal parallelepiped before and after deformation.
A point pO originally located by x moves to a new position P located by y, while
another point QO located by x+dx moves to a position Q located by y+dy. In terms
of curvilinear coordinates, the differential vector dx is as defmed in Eq. 1.173, and the
differential vector dy can be put in a similar form, i.e.
dy = ay del + ay de2 + ay de3 ae l
ae 2
ae3

(10.141)

dy=GAdeA

(10.142)

or
where G Ais the covariant base vector after deformation, i.e.

G _ ay =
A- aeA-y,A

(10.143)

Note the direction of each G A in Fig. 10.8. In fact, all identities derived in Sections
1.14 and 1.15 in terms of g and x also hold when g and x are changed into G and y,
respectively. For examples; the square of the length of the element PQ is
DEFORMED

ORIGINAL

Curve

Fig. 10.8 Original and deformed configurations of an infmitesimal parallelepiped.

402

Foundations of Solid Mechanics

(10.144)
where G'AjJ. is the covariant metric tensor after deformation, the volume of the infmitesimal parallelepiped after deformation is

dV =-{(jd81d82d83
where G =det[G~, and the derivative of Gil with respect to 8 v is

(10.145)

G~V={~rGA
in which

{~}

(10.146)

is the Christoffel symbol of the second kind after deformation.

The difference between Eqs. 10.144 and 1.177 is


dy . dy - dx . dx =(G'AjJ. - g~d8Ad81l

(10.147)

Comparing the last equation with Eq. 10.96, we define the strain tensor in curvilinear
coordinates as follows
(10.148)
As shown in Fig. 10.8 the displacement vector u(8\ 8 2, 83) is related to the position
vectors x and y as
(10.149)
y=x+u
The contravariant and covariant components of u have been defmed in Eqs. 1.199 and
1.202, respectively, i.e.
u = VAgA= V~A

(10. 150a,b)

Substituting Eq. 10.149 into Eq. 10.143, and using Eqs. 1.174, 1.200 and 1O.150a, we
fmd

GA =(~+ v;~g.,

(10.151)

Due to the last equation, Eq. 10.148 becomes

_.!.(
1C
1C
1C P)
2 gA1Cv;1l + gKJ1V;A + g1CpV;AV;1l

'AjJ. -

(10.152)

or, making use of Eq. 1.209,


(10.153)
10.8 EQun..mRIUM EQUATIONS AND STRESS TENSORS IN CURVILINEAR
COORDINATES
Consider the equilibrium of an infmitesimal parallelepiped at a point P in the
deformed configuration shown in Fig. 10.9. Forces involved are the body force X (per

Finite Deformation

403

unit mass) and tractions on the six infmitesimal surfaces. Since the area of the surface
e A= constant in the original configuration is, due to Eq. 1.193, proportional to de f1de lC,
it is logical to put the resultant force vector corresponding to this surface in the form
(10.154)
dTA= tA~ def1delC, (A. :;!: J.! :;!: x:)
where g is as defined in Eqs. 1.183b to d. Naturally, the resultant force vector on the
side corresponding to the surface e A+ deA= constant in the original configuration is
a(dTA)
dTA+--A-de\ (A. not summed).
ae
All resultant force vectors involved are depicted on appropriate sides of the parallelepiped, following the sign convention of stresses (positive if in tension), in Fig. 10.9.
Accordingly, the equilibrium of this parallelepiped can be put in the form
a(dTl) del + a(dT2) de2 + a(dT3) de3 + OXdV o = 0
ae l
ae2
ae3
p

(10.155)

Substituting Eqs. 1.186 and 10.154 into the last equation we fmd

(~tlC).lC+pOX~ =0
The vector

(10.156)

may be expressed in the form


tA='t~G

f1

Fig. 10.9 Deformed configuration of an infmitesimal parallelepiped being


subjected to forces.

(10.157)

Foundations of Solid Mechanics

404

A consideration of the moment equilibrium of the infmitesimal parallelepiped leads to


the condition
3

I. dT'" x G",da'" =0,

"'=1
or, due to Eqs. 10.154 and 10.157,

't~Gf1 X G", = 0,

or
1

2('t~-~)Gf1xG", =0.
Since the vector Gf1 x G", is nonzero for J.1::. A, the equation above gives
't~=~
(10.158)
i.e. the stress tensor ~ is symmetric.
The body force X may be expressed in terms of its contravariant components
before deformation as follows

X =X~g",

(10.159)

Substituting Eqs. 10.157 and 10.159 into Eq. 10.156, and using Eqs. 1.194 and 10.151,
we find
(10.160)

The equation above holds for dynamic equilibrium as well if X is considered as the
sum of the actual body force vector and the D' Alembert's force vector, which is equal
to the negative of the acceleration vector.
If 't~f1gf1 is the stress vector on the surface where a 1 constant in the original
configuration, the resultant force dT 1 there, because of Eq. 1.193, becomes

dT 1 ='t~f1gll"g gllda2da3

(10.161)

Comparing the last equation with Eq. 10.154 for A = 1, we get


(10.162)

or, for any A,

t'" ='t~gll' (A not summed)

(10.163)

On the other hand, Eqs. 10.154 and 10.157 give


dT 1 = 'tlf1GIl{ida2da3

(10.164)

Equating Eq. 10.161 with Eq. 10.164, we get

1 ('tlf1G \
'tlf1g =__
o f1 -{ill
'"
or

(10.165)

Finite De/ormation

405
'tllLg
OIL

12g + 't13g =_1_ ('tIlG + 't12G + 't13G )


='tllg
01 + 't02
03-{ill
I
2
3

(10. 166a,b)

or, on the surface where SA. =constant,


.Au

.. I

A.3

'0' gIL = 'to gl + 'to g2+'t0 g3 =

1 (A.lG
uG2+'tA.3G3),
_ru.
't
I +'t
-Vg U
(A. not asummed)

(10. 167a,b)
Substituting Eq. 10.151 into Eq. 10.165 then considering the vectorial components of
the result in g}, g2 and g3 directions yield, respectively,
(10. 168a)
(10. 168b)
(10.168c)

DEFORMED

Curve

Fig. 10.10 Decomposition of stress vector 't~lLglL into three components


parallel to edges of each parallelepiped.

Foundations 0/ Solid Mechanics

406

Thus we may write a general relationship between ~ and .(I' as the following

~ =}gu<-"(~ + v;~,

(A. not summed)

(10.169)

It should be noted that while .(I' carries a meaning similar to the Cartesian Kirchhoff
stress tensor 11;, ~ carries a meaning similar to the Cartesian Lagrangian stress tensor
ag. Moreover, Eqs. 10.167 imply that the stress vector ~gJ1 has three vectorial components in a 1, a 2 and a 3 directions in the original configuration, but in Gl> G2 and G3
directions in the deformed configuration. Figure 10.10 depicts, as an example, such
vectorial decompositions of the stress vector 't~J1gJ1' If the original dimension of the
structure in the a 3-direction is relatively small, we may assume that .-r3, .-r1 and .-r2 are
zero, i.e. the same assumption as of the plane stress conditions with respect to the
surface where a 3 =constant.
10.9 PHYSICAL COMPONENTS OF VECTORS AND TENSORS
When the a-system is a local rectangular Cartesian coordinate system, mathematical conditions involved are very much simplified as Eqs. 1.211 to 1.213 show, and
the physical meaning of the quantities involved can be comprehended more easily. Most
obviously, a contravariant or a covariant component of a vector such as in Eqs. 10.150
and 10.159 becomes indeed a Cartesian component of that vector. The strain tensor
given in Eq. 10.153 carries the same physical meaning as Eq. 10.101, since the former
can be reduced to the form

~ =~(V~J1 +VJ1,A +V"'AV",~,

(for Cartesian system)

(10.170)

Next, rewriting Eq. 10.154 under the special conditions mentioned above for 1..= 1, as
an example, we have
(10.171)
dT1 =t 1da2da\ (for Cartesian system)
which shows that t 1 has the same physical meaning as cr:i in Eq. 10.62. With an
observation that GA has the same physical meaning as 'OY/dX1 for the global Cartesian
coordinate system, we can also conclude that the stress tensor.(l' in Eq. 10.157 carries
the same meaning as ~ii in Eq. 10.60b. Further, we can write the equation of equilibrium,
Eq. 10.160, in the form

['t"J1(a~ + v.~]." + pOX~ =0, (for Cartesian system)

(10.172)

which clearly has the same physical meaning as Eq. 10.95.


In general, components of a tensor derived for a curvilinear coordinate system are
not mutually orthogonal, and do not have ordinary physical dimensions. Still, it is
expedient to formulate some physical problems in terms of such non-physical components, then derive the corresponding physical components at a later stage. The derivation

Finite Deformation

407

for physical components from the non-physical ones referred to the a-system can start
with setting up a local rectangular Cartesian coordinates system (z-system) at the
material point pO as shown in Fig. 10.11. Denoting the unit vector in the direction of
z~-axis by j~, and noting the identities

o( ) o( )oz"
oa). = oz" oa).
o( ) o( )oa"
oz). oa" oz).

--=----

(10.173a)
(10.173b)

we have
(10. 174a)

. ax 09"
=oz). =oz).~
oz). ..
oa" =J).' gIL

J).

(l0.174b)
(10.174c)
(10. 174d)

9 1 Curve

Fig. 10.11 Global Cartesian coordinates Xi, curvilinear coordinates a\ and


local Cartesian coordinates Zi of material point pO.

408

Foundations of Solid Mechanics

In addition, the displacement vector u can be written as


Ao

(10.175)

u=u JA
Equating this equation with Eq. 10.150, and using Eq. 1O.174a, we get
A dZ A

u =-v ll
dell

(10.176)

Similarly, for the body force vector X, we can show that


XA= dZ AXIl
dell g

(10.177)

where X: has been defined in Eq. 10.159, while X A is the Cartesian component in
zA-direction, i.e.
(10.178)
Being consistent with the derivations of strain tensors of Eqs. 10.98 and 10.148,
the physical strain component defmed with respect to the z-system is

1[ dY dY

E,.,. =2

ax ax]

dZ A dZ Il - dZ A dZ Il

Using Eqs. 10.148, 1O.173b and 1O.174c, we can rewrite the last equation as
del< de p
P. - - . - A .
~-dZA dZIl~

(10.179)

(10.180)

or, conversely,
(10.181)
In fact, ~ is a local Cartesian component of the Green's strain tensor.
For the relationship between stress tensors of the two different curvilinear coordinate systems, the derivation is based on the equilibrium of the system shown in Fig.
10.12 where the infmitesimal tetrahedron PRST denotes the deformed configuration of
pORoSoTo in the original state. In addition to the surface forces on this infmitesimal
tetrahedron depicted in Fig. 10.9, there is another force XdS~ on the surface RST, where
dS~ is the area of the surface ROSoTo. The condition of equilibrium of the forces acting
on the tetrahedron can be put in the form
XdS~=(dTl+dT2+dT3)/2

(10.182)

The body (actual and D' Alembert's) forces do not enter into this equation, because they
are of higher order of smallness than the surface forces. Substituting Eq. 10.154 into
the last equation, and using Eqs. 1.188, we find
X = (gl<. n;,)tl<

(10.183)

where ~ is the unit vector normal to the surface ROSoro. On the other hand, if we
write X with respect to the z -system, i.e. for an infmitesimal tetrahedron formed by

409

Finite De/ormation

'dY2 and'd
Jl'dY 1, J2
J3 Y3, we get
X=('h:~SA

where SA is introduced in a similar form as

(10.184)

was in Eq. 10.157, i.e.

SA=r.'4L dy
(10.185)
az ll
Since X, being a physical quantity, does not depend on the choice of coordinate systems,
we can equate Eq. 10.184 with Eq. 10.183 and get
(jA' ~)SA= (gil' ~)f

(10.186)

By taking the direction ~ coincident with a zlt-axis and substituting Eqs. 10.157 and
10.185 into Eq. 10.186, we have

V~

::p =

(gil .

jJ~

!A

(10.187)

or, due to Eqs. 10.173b and 10.174d,


r.'4L =az Aaz ll'tlCp
aeltaep

(10. 188a)

or, conversely,
(10.188b)

ORIGINAL

Curve

DEFORMED

T_----"

Fig. 10.12 Original and deformed configurations of an infmitesimal tetrahedron


being subjected to forces.

Foundations of Solid Mechanics

410

In fact, 'J:,'J..jL is a local Cartesian component of the Kirchhoff stress tensor. Next, we
will find the physical components when the a-system is orthogonal, for which some
prevailing conditions are listed in Eqs. 1.210. Setting a local rectangular Cartesian
coordinate system (z-system) to have the same orientation as the a-system, i.e. j1' jz
and j3 are in the same directions as gl' gz and g3' respectively, we can show that
g" =...{fuj"

(10. 189a)

az" __ r::- B"


aa l1 -"'Igl1l1 11

(l0.189b)

ae" =_1_ 8"

(l0.189c)

az l1

w;:;:

11

Consequently, physical components in directions of orthogonal curvilinear coordinates


can be specialized from Eqs. 10.176, 10.177, 10.180 and 10.188 as

u"=...{fuv"

(10. 190a)

X "...{fu"
= guXg

(l0.190b)

~ =-Vg""g11l1 ~

(10. 190c)

=-Vg""g1111 -(11

(10. 19Od)

'J:,'J..jL

10.10 BOUNDARY CONDIDONS AND CONSTITUTIVE RELATIONSHIP


IN CURVILINEAR COORDINATES
At the point pO in the original configuration, let 8u be a virtual displacement with
contravariant components as follows
(l0.191a)
dS~ a small surface passing through that point, n~ the unit outward vector normal to

dS~, and

X the

stress vector on dS~. Substituting Eq. 10.163 into Eq. 10.183 we find
-

VJ1

X =X g gl1

(l0.191b)

where X~ is a contravariant component of X and has an explicit form as the following


3

~ = I. (g,,' n~)'t~

(10.192)

(X. 8u)dS~ =[(~gI1J8V1CJdS~

(10.193)

,,=1
The virtual work due to X through 8u is, because of Eqs. 1.178 and 10.191,

Thus the boundary conditions on the boundary dS~ involve the prescription of either

Finite Deformation

411

X:g"lC or VlC
(10.194)
A
If dS~ is a surface where 9 = constant, n:, is in the same direction as fl, then due to
Eqs. 1.192c, 10.192 and 10.194, boundary conditions become either in
~g"lC or VlC, (on surface where 9 A=constant)
(10.195)
At the instant that the position vector y for the point pO(91, 92, 93) in Fig. 10.8 is
increased by dy, such vector for the surface where 9A+ d9A= constant is increased by
dy + o~:~) d9\

(A. not summed)

Thus the corresponding increase of the strain energy for the infmitesimal parallelepiped
at the point pO (Figs. 10.8 and 10.9) is

pOdUdVo =

{[dTA+ o(dr) d9Al . (dY+ o(dy) d9A)-dr. dY}


A=l
o9 A J
o~
+ pOX. dydVO

(10.196)
A
where U is the strain energy per unit mass. Neglecting the (d9 )2 terms in the equation
above yields

o(dTA) d9A+POXdvol . dy+


dTA. o(dY)d9A
A=l o9 A
A=l
09 A
or, due to the equilibrium given by Eq. 10.155.

pOdUdVO=[

pOdUdVo =

(10.197)

f dr o(dy)
d9\
o9

A=l

or
(10.198)
Substituting Eqs. 10.143 and 10.154, in view of Eqs. 1.186 and 10.157, into the equation
above, we get
pOdU =

<"G,,' dGA

(10.199)

or, due to the stress symmetry given by Eq. 10.158,

=2'1"J.a
' .d(G GJ

P dU

= 2<"d(GA. G" - gA . gJ,


or, due to Eqs. 1.178 and 10.148,
(10.200)
which is the same as Eq. 10.111 and equivalent to Eq. 3.7 for small defonnation. Thus

Foundations of Solid Mechanic.

412

we may conclude that a rational form of the constitutive law is the relationship betweeJ
the non-physical stress -(- and the non-physical strain ~.
Following the same scheme as adopted to derive Eqs. 3.11 to 3.15, we obtain the
stress-strain relationship for linear elastic solids as
-(-=D')..pDfJ~

(10.201

and that for the local Cartesian (z~) system as


(10.202
r!"'=C~~
Recalling Eqs. 10.180 and 10.188a, which relate the Cartesian strain and stress tensor:
with the non-physical ones, we can get the relationship between stiffness tensors 0
different coordinate systems as the following

aelLaea ae~
az" az 3 az az~

D')..pDfJ =C')&l; ae~

(10.203

or
(10.204:
10.11 COMPATIBILITY CONDmONS
The Riemann-Christoffel tensor R~ (Spiegel, 1959) is dermed in the deforme(
configuration as

R~ =~:JG
_~~r +{ lG{A.}G _1 lC}G{ A.1
ae
ae ~
4tv
lC

ILVGl

1CV

lCcJ

(10.205:

and its associated fourth order tensor commonly known as the Riemann-Christoffe,
covariant curvature tensor is
R'AjJ.VGl

=G~.,.R:VGl

(10.206:

Only 6 of the 81 components of R'AjJ.VGl are distinct, due to the following symmetr)
conditions
(10.207:
R'AjJ.VGl = RVGl'AjJ. = -R~VGl = -R'AfuJN = -R'AlsN1L
The 6 distinct covariant components of the curvature tensor may be chosen as
R 2323 R 3131 R 1212 R l231 R2312 R 3123
(10.208:
If we take the strain components ~ as arbitrary continuous functions of coordinates of points in the material domain, the deformation of the material body would b~
insured of continuity only if
(10.209:

It is easy to prove that these compatibility conditions are necessary. but difficult to prov

Finite DejormtJtion

413

that they are sufficient; the latter will not be given here. The proof of sufficiency is
based on the fact that the covariant differentiation in a Euclidean space is commutative,
thus the Riemann-Christoffel tensor and curvature tensor are zero in the Eulidean space
where the positions of points in the material body are defmed after defonnation.
Without loss of generality in the application into the compatibility conditions, we
can put the curvilinear coordinates identical to the reference Lagrangian (global Cartesian) coordinates, i.e.

a1 =X a2 =x2' a3 =~
1,

(10.210)

Accordingly, the covariant metric tensor becomes


Gij = (oy/oxj ) (oy/ox)

(10.211)

or, due to y=x+u,


(10.212)
where Eij is the Green's strain tensor. Also, the contravariant metric tensor is given by
..
1 dG
G" = 2G dE .. '

"
1 dG
= 4GdEv'

(i = j)

(10.213a)

(i :f:. j)

(10.213b)

where
(1O.214a)

G =det[G~

1 +2Ell

2E12

2E13

2E12
2E13

1+2Ezz

2~

2~

1+2~3

(10.214b)

= (1 + 2Ell ) (1 + 2Ezz) (1 + ~3) + 16El~EI3

va

-4(1 +2Ell)~-4(1 +2EzJE~3-4(1 +2~3)E~2

(10.214c)

In fact,
as indicated by Eq. 10.145 is the ratio between the new and the original
volumes; thus, in view of Eqs. 10.37 and 10.46, we can write

..JG =IdY/dXjl
dUI
dXl
dUz
dXl
dU]
dXl

1+=

(10.215)

dUI
dx2
1+ dUz
dX2
dU]
dX2

dUI
dx3
dUz
dX]
dU]
1+dX3

(10.216)

Foundations of Solid Mechanics

414

Two compatibility conditions given by Novozhilov (1961) in an explicit fonn are


listed below, while the remaining four conditions can be obtained from these two by
cyclic pennutation of numerals 1, 2, 3:

G3J a~3 (2 aE aEn) _a~3 (aE13 + aEz3 _aEIZ]


13 _

Laxz aXI aX3 aXI axz aXI aX3


+ GJ2 aEIz(aE12 + aEl3 _aEz3) + aEn aEn _2 aE aElz _aEnaEn]
L aXI aX3 axz aXI aXI aX3 axz aX3 aXI aX3
+ G J 2 aE (aElz + aEl3 _aEz3) + aEn a~3 _2 aE aEl3 _aE a~3]
L aXI aX3 axz aXI aXI
ax z
aX3 axz axz
aXI
+GzJ aEzz (2 aEl3 _aEn) + a~3 (2 aE12 _aEn) _aEzz a~3
LaX 3 aXI aX3 axz aXI ax z aXI aXI
(aE13 + aEz3 _aEIZ) (aElz + aEz3 _aE13] ;
1. axz aXI aX3 aX3
aXI axz
ll

l3

OZE lz a~n OZEzz


2-----aXlaXZ axi
ax;

ll

ll

Finite De/ormation

415

-2 aEzz (aE + a~ _aE12] ;


aX2

aXI

(10.217)

etc.

13

aXI

aX3

If the strain components are sufficiently small compared with unity and can be
neglected, then the difference between the covariant and the contravariant components
of the Riemann-Christoffel tensor will be slight and consist only of terms of the same
order of magnitude as the strain components. In order to incorporate this condition, we
must omit all terms in Eqs. 10.217 which have strain components as factors but not
their derivatives, since the latters may be considerably greater than strain components
themselves. With this modification, we obtain the six distinct compatibility conditions
as follows

+ (aEz3+ aE13 _aE12)2 _aEzz (2 aE12 _aEu)_ aEU(2 aE12 _aEzz ).


aXI

aX2

aX3

aX2

aXI

aX2

aXI

aX2

aXl'

!fEu a (aE12 aE aEz3)


13

aX2aX3 - aXI

aX3

+ aX2

aX I

= aEu aEll + aEzz (aE12 + aEz3 _aE13 ) + a~3 (aE13 + a~ _aE12)


aX2 aX3

aXI

aX:!

aXI

aX2

aXI

a.xz

aXI

aX3

_aEll (aEI3 + OEI2_ a~)_ aEzz(2aEI2_ aEll)_ a~3(2aEI3_ aEll ).


aXI

aX2 aX3

aXI

aX3

aXI

aX2 aX2

aXI

aX3'

etc.

(10.218)

Finally, if we assume that not only strains but their derivatives are small compared
with unity, we can omit all product terms in the previous equations whereupon Eq.
is obtained.

2.62

Foundations of Solid Mechanics

416

In Chapter IV, we have illustrated the use of the compatibility conditions in the
solution approach, with strains as basic unknowns, to the problems of elastic planes
under the small defonnation theory. Such an approach under the ftnite defonnation
theory is not so practical, because the compatibility conditions become too complicated.
10.12 PROBLEMS

10.12.1 Constitutive Law in Eulerian Description


In the Eulerian description, if the strain energy function U is defmed as per unit
mass, it can be shown with the same procedure as deriving Eq. 3.4 that

pdU

{aYjau.)

(J ..

-'

!I

(10.219)
where eij is as defmed in Eq. 10.107. Substituting Eq. 10.48 into the equation above
yields an identity which is similar to Eq. 10.111 of the Lagrangian description, i.e.

pOdU
where

(Jijdfv

_I

1-1dU*
<lEv= -ax",
ay"

(10.220)
(10.221)

Note that Ev defmed in Eq. 10.221 is a symmetric tensor of the second order. If (Jij
components are unique and homogeneous functions of the components of the 'new'
strain tensor Ev, we can have the strain energy function as a unique and homogeneous
function of such strain tensor components. Thus there can be a rational constitutive
law in the Eulerian description relating the Eulerian stress tensor (Jij with the strain
tensor Ev deftned in Eq. 10.221. However, such a law is rather complicated due to the
nature of Eq. 10.221 and the type of description.

10.12.2 Maxwell-Betti Reciprocal Theorem for Finite Deformation


Adopting superscripts 1 and 2 in parentheses to denote respective systems as in
Section 3.4, we can write for a linear elastic solid undergoing a fmite deformation the
following
(10.222)

(10.223)

Finite Deformation

417

If the generalized Hooke's law, Eq. 3.11, holds relating Eli with Eli' the equation above

becomes

pO

r X;(l)u;(2)dVo+f. ~l)u?)dSo= r CjmklE~)Oay;(1)[oYl2) -~ilJdVO

JyO

JyO

SO

iXm

OXj

(10.224)

Interchanging the superscripts 1 and 2 in the last equation, we get

Pol

yO

X~)u~l)dVo+f. ~)u~l)dSo=1

SO

yO

C. E(2)Oy;(2)[oYl1)-~.JdVO
jmld 1<1

a,v""ma,v

""j

(10.225)

Though the stiffness tensor Cli1<l is symmetric as in Eq. 3.14, Eqs. 10.224 and 10.225
are not identical. Thus the Maxwell-Betti reciprocal theorem does not exist in fmite
deformation.

REFERENCES
The article by Truesdell and Toupin (1960) contains an extensive list of references from
the years 1678 to 1960. The works cited in this book and published in the twentied
century are listed alphabetically below:
1.
2.
3.

4.
5.
6.

7.
8.
9.
10.
11.
12.
13.
14.

15.

16.

Abramowitz, M., and Stegun, I.A. (eds.), Handbook of Mathematical Functions,


Dover Publications, New York, 1965, pp. 355-494.
Achenbach, J.D., Wave Propagation in Elastic Solids, North-Holland, 1973, pp.
73, 74, 76, 217 and 218.
Agabein, M.E., Lee, SL., and Dundurs, J., 'Bending of an Annular Slabs Supported on Columns', J. Franklin Inst., Vol. 284, No.5, November 1967, pp.
300-307.
Alfrey, T., 'Non-homogeneous Stresses in Viscoelastic Media', Quart. Appl.
Math, Vol. 2, No.2, 1944, pp. 113-119.
Almansi, E., 'Sulle deformazione fmite dei solidi elastici isotropi, 1', Rend. Accad.
Naz. Lincei, (Ser. 5), Vol. 20, 1911, pp. 705-714.
Apirathvorakij, V., and Karasudhi, P., 'Quasi-static Bending of a Cylindrical
Elastic Bar Partially Embedded in a Saturated Elastic Half-Space', Int. J. Solids
Structures, Vol. 16, No.7, 1980, pp. 625-644.
Bellman, R, Kalaba, RE., and Locken, J., Numerical Inversion of the Laplace
Transform, American Elsevier, 1966.
Bingham, E.C., Fluidity and Plasticity, McGraw-Hill, New York, 1922.
Biot, M.A., 'General Theory of Three Dimensional Consolidation', J. Appl. Phys.,
Vol. 12, 1941, pp. 155-164.
Boresi, A.P., and Lynn, P.P., Elasticity in Engineering Mechanics, Prentic-Hall,
Englewood Cliffs, New Jersey, 1974, pp. 310-316.
Bridgman, P.W., 'The Compressibility of Thirty Metals as a Function of Pressure
and Temperature', Proc. Am. Acad. Art. Sci., Vol. 58, 1923, pp. 163-242.
Budiansky, B., 'A Reassessment of Deformation Theories of Plasticity', J. Appl.
Mech., Vol. 26, 1959, pp. 259-264.
Cagniard, L., Reflection and Refraction of Progressive Seismic Waves, E.A. Flinn
and C.H. Dix (translated and revised), McGraw-Hill, New York, 1962, pp. 47-49.
Cesaro, E., 'Sulle formole del Volterra, fondamentali nella teoria delle ditorsioni
etastiche', Rendiconto dell' scienze fisiche e matematiche (Societa reale di
Napoli), Vol. 12, 1906, pp. 311-321.
Chan, K.S., Karasudhi, P., and Lee, S.L., 'Force at a Point in the Interior of a
Layered Elastic Half Space', Int. J. Solids Structures, Vol. 10, No. 11, 1974, pp.
1179-1199.
Christensen, RM., Theory of Viscolasticity-An Introduction, Academic Press,
New York, 1971, pp. 43-51, 63-72, 105-107.
419

420

Foundations of Solid Mechanics

17.

Christensen, R.M., amd Schreiner, R.N., 'Response to Pressurization of a


Viscoelastic Cylinder with an Eroding Internal Boundary', AIAA J., Vol. 3, 1965,
pp. 1451-1455.
Churchill, R.V., Complex Variables and Applications, 2nd ed. McGraw- Hill,
New York, 1960, pp. 34-43, 57, 80-86, 120, 154-173, and 179.
Conneau, I., Viscoplasticity and Plasticity in the Finite Element Method, Ph.D.
Dissertation, Univ. ColI. Swansea, 1976.
Cost, T.L., 'Approximate Laplace Transfonn Inversions in Vicoelastic Stress
Analysis', AIAA Journal, Vol. 2, No. 12, 1964, pp. 2157-2166.
Crossland, B., 'The Effect of Fluid Pressure on the Shear Properties of Metals',
Proc. Inst. Mech. Engr., Vol. 169, 1954, pp. 935-944.
Drucker, D.C., 'A More Fundamental Approach to Plastic Stress Strain Relation',
Proc. 1st. U.S. Natl. Congress Appl. Mech., Chicago, 1951, pp. 487-491.
Drucker, D.C., 'A Definition of Stable Inelastic Material', J. Appl. Mech., Vol.
26, 1959, pp. 101-106.
Essenburg, E 'On Surface Constraints in Plate Problems', J. Appl. Mech., Vol.
29, 1962, pp. 340-344.
Ewing, W.M., Jardetzky, W.S., and Press, E, Elastic Waves in Layered Media,
McGraw-Hill, New York, 1957, pp. 24-31, 136, 211-214.
Ferry, J.D., Viscoelastic Properties of Polymers, 2nd. ed., Wiley, New York,
1970.
Folie, C.M., 'Bending of Clamped Orthotropic Sandwich Plates', J. Eng. Mech.
Div., ASCE, Vol. 96, No. EM3, Proc. Paper 7340, June 1970, pp. 243-265.
Frazer, R.A., Duncan, W.J., and Collar, A.R., Elementary Matrices, Cambridge
University Press, 1938.
Fung, Y.C., Foundations of Solid Mechanics, Prentice-Hall, Englewood Cliffs,
1965, pp. 142, 143, 150-152, 476 and 477.
Galerkin, B., 'Contribution a la solution generale du probleme de la theorie de
l'elastcite dans Ie cas trois dimensions', Comtes Rendus, Vol. 190, 1930, pp.
1047-1048; Comtes rendus (Doklady) acado sci. U.R.S.S., Ser. A, Vol. 14, 1930,
p.353.
Gilman, J.J., 'Physical Nature of Plastic Flow and Fracture', in Plasticity, E.H.
Lee, and P.S. Symonds (eds.), Proc. 2nd. Symp. Naval Structural Mechanics,
Providence, 1960, pp. 44-49.
Gottenberg, W.G., and Christensen, R.M., 'An Experiment for Detennination of
the Mechanical Property in Shear for a Linear Isotropic Viscoelastic Solid', Int.
J. Eng. Sci., Vol. 2, 1964, pp. 45-57.
Graham, G.A.C., 'The Contact Problem in the Linear Theory of Viscoelasticity',
Int. J. Eng. Sci., Vol. 3, 1965, pp. 27-46.
Graham, G.A.C., 'The Contact Problem in the Linear Theory of Viscoelasticity
When the Time Dependent Contact Area Has Any Number of Maxima and

18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.

31.

32.

33.
34.

References

35.

36.
37.
38.
39.

40.

41.
42.
43.
44.
45.

46.

421

Minima', Int. J. Eng. Sci., Vol. 5, 1967, pp. 495-514.


Graham, G.A.C., 'The Correspondence Principle of Linear Viscoelasticity Theory
for Mixed Boundary Value Problems Involving Time-Dependent Boundary
Regions', Quart. Appl. Math., Vol. 26, 1968, pp. 167-174.
Haar, D. ter, 'An Easy Approximate Method of Determining the Relaxation
Spectrum of a Viscoelastic Material', J. Polymer Sci., Vol. 6, 1951, p. 247.
Habip, L.M., 'A Survey of Modern Developments in the Analysis of Sandwich
Structures', Appl. Mech. Review, Vol. 18, No.2, February 1965.
Hamel, G., Elementare Mechanik, Leipzig and Berlin, 1912.
Handelman, G.H., Lin, C.C., and Prager, W., 'On the Mechanical Behavior of
Metals in the Strain-Hardening Range', Quart. Appl. Math., Vol. 4, 1947, pp.
397-407.
Harkrider, D.G., 'Surface Waves in Multilayered Elastic Media, I. Rayleigh and
Love Waves from Buried Sources in a Multilayered Elastic Half-Space', Bull.
Seismological Soc. America, Vol. 54, No.2, April 1964, pp. 627-679.
Hencky, H. 'Zur Theorie plastischer Deformationen und der hierdurch hervorgerufenen Nachspannungen', Z. angew. Math. Mech., Vol. 4, 1924, pp. 323-334.
Hill, R, The Mathematical Theory of Plasticity, Oxford: Clarenden Press, 1950.
Hodge, P.G., Jr., 'Piecewise Linear Plasticity', Proc. 9th. Int. Congress Appl.
Mech., Brussel, 1956, and Vol. 8, 1957, pp. 65-72.
Hopkins, I.L., and Hamming, RW., 'On Creep and Relaxations', J. Appl. Phys.,
Vol. 28, 1957, pp. 906-909.
Hovichitr, I., Karasudhi, P., Nishino, F., and Lee, S.L., 'A Rational Analysis of
Plates with Eccentric Stiffeners', PERIOD/CA, Int. Assoc. Bridge Struct. Eng.,
Proceedings, p-9n7, 1977.
Hsu, H.P., Fourier Analysis, Simon and Schuster, New York, 1970, pp. 107 and

119.
47.

48.
49.

50.

51.

Hughes, T.IR, and Taylor, RL., 'Unconditionally Stable Alogorithms for


Quasi-Static ElastoNiscoplastic Solid Finite Element Analysis', Computers and
Structures, Vol. 8, 1978, pp. 169-173.
Hunter, S.C., 'The Hertz Problem for a Rigid Spherical Indenter and a Visco-elastic Half-Space', J. Mech. Phys. Solids, Vol. 8, 1960, pp. 219-234.
Hunter, S.C., 'Tentative Equations for the Propagation of Stress, Strain and
Temperature Fields in Viscoelastic Solids', J. Mech. Phys. Solids, Vol. 9, 1961,
pp. 39-51.
Jardetzky, W.S., 'Period Equation for an n-Layered Half Space and Some Related
Questions', Columbia Univ. Lamont Geols. Obs. Tech. Rept. Seismology, No. 29,
1953.
Kalker, II, Threedimensional Elastic Bodies in Rolling Contact, Kluwer, Dordrecht, the Netherlands, 1990.

422

Foundations of Solid Mechanics

52.

Kanok-Nukulchai, W., Taylor, RL., and Hughes, T.J.R., 'A Large-Defonnation


Fonnulation for Shell Analysis by the Finite Element Method', Computers and
Structures, Vol. 13, 1981. pp. 19-27.
Kao, J.S., 'Bending of Circular Sandwich Plates', J. Eng. Mech. Div., ASCE,
Vol. 91, No. EM1, Proc Paper 4449, August 1965, pp. 168-171.
Kao, J.S., 'Circular Sandwich Plates under Eccentric Loads', J. Eng. Mech. Div.,
ASCE, Vol. 95, No. EM1, Proc. Paper 6413, February 1969, pp. 235-245.
Kao, J.S., 'Bending of Circular Sandwich Plates due to Asymmetric Temperature
Distribution', AIAA J., Vol. 8, No.5, May 1970, pp. 951-954.
Kao, J.S., 'Differential Equations for Asymmetric Bending of Circular Sandwich
Plates', AIAA J., Vol. 11, No.4, April 1973, pp. 573-575.
Karasudhi, P., 'An Infinite Element Algorithm for Elastodynamic Load Transfer
Problems', Computational Mechanics' 86 - Theory and Applications, Proc. Int.
Con! Computational Mech., G. Yagawa, and S.N. Atluri (eds.), Springer-Verlag,
Tokyo, May 1986, pp. IX73-78.
Karasudhi, P., Alam, K.M.A., and Lee, S.L., 'Flat Slabs on Three Rows of
Columns', Israel J. Tech., Vol. 11, No.3, 1973, pp. 149-156.
Karasudhi, P., and Alvappillai, A., 'An Infinite Element Algorithm for Predicting
Land Subsidence', Computational Mechanics' 88-Theory and Applications, Proc.
Int. Con! Computational Engg. Sci., S.N. Atluri and G. Yagawa (eds.),
Springer-Verlag, Atlanta, Georgia, April 1988, pp. 63.i.l to 4.
Karasudhi, P., Jou, T.S., and Tansirikongkol, V., 'Curved Box Girder Bridges as
Sandwich Plates', J. Struct. Eng., Vol. 4, No.2, 1976, pp. 63-69.
Karasudhi, P., Karunasena, W.M., and Puswewala, V.G.A., 'An Infinite Element
Algorithm for Multilayered Saturated Half Spaces', Proc. US-Korea Joint
Seminar/Workshop on Critical Engineering Systems, Seoul, May 1987, Vol. 1,
pp. 207-216.
Karasudhi, P., Keer, L.M., and Lee, S.L., 'Vibratory Motion of a Body on an
Elastic Half Plane', J. Appl. Mech., Vol. 90, 1968, pp. 697-705.
Karasudhi, P., Ng, H.K., and Lee, S.L., 'Axisymmetric Bending of Sandwich
Plates on Elastic Foundation', Eng. J. Singapore, Vol. 4, No.1, 1977, pp. 17-26.
Karasudhi, P., and Prechaverakul, S., 'A Theoretical Prediction of Land Subsidence due to Water Loss from Aquifers', Proc. US-Asia Con! Engineering for
Mitigating Natural Hazards Damage, P. Karasudhi, P. Nutalaya, and A.N.L. Chiu
(eds.), Bangkok, December 1987, pp. C8-1 to C8-12.
Karasudhi, P., Rajapakse, RK.N.D., and Hwang, B.Y., 'Torsion of a Long
Cylindrical Elastic Bar Partially Embedded in a Layered Elastic Half-Space', Int.
J. Solids Structures, Vol. 20, No.1, 1984, pp. 1-11.
Karasudhi, P., Rajapakse, RK.N.D., and Liyanage, K.K., 'A Reconsideration of
Elastostatic Load Transfer Problems Involving a Half Space', Trans. Canadian
Soc. Mechanical Eng., Vol. 1, No.4, 1984, pp. 218-226.

53.
54.
55.
56
57.

58.
59.

60.
61.

62.
63.
64.

65.

66.

References

67.
68.
69.
70.

71.
72.
73.

74.
75.
76.
77.

78.
79.

80.

81.

82.
83.

423

Karman, T. von, 'Festigkeitsversuche unter allseitigem Druck', Z. VDI, Vol. 55,


1911, pp. 1749-1757.
Keller, H.B., 'Propagation of Stress Discontinuities in Inhomogeneous Media',
SIAM Review, Vol. 6, No.4, October 1964, pp. 356-382.
Kliushnikov, V.D., 'On Plasticity Laws for Work Hardening Materials', J. Appl.
Math. Mech. (translated from Prikl. Mat. i. Mekh.), Vol. 22, 1958, pp. 129-160.
Koiter, W.T., 'Stress-Strain Relations, Uniqueness and Variational Theorems for
Elastic-Plastic Materials with a Singular Yield Surface', Quart. Appl. Math., Vol.
11, 1953, pp. 350-354.
Koiter, W.T., 'General Theorem for Elastic-Plastic Solids', in Progress in Solid
Mechanics, LN. Sneddon, and R. Hill (eds.), North-Holland, 1960, Chapter 4.
Koppe, H., 'Uber Rayleigh-Wellen an der OberfUiche zweier Medien', Z. angew.
Math. Mech., Vol. 28, 1948, pp. 355-360.
Kosherick, H.J., Dunduti, J., and Lee, S.L., 'Annular Slabs in Circular Core
Buildings', J. Structural Div., ASCE, Vol. 94, No. ST2, February 1968, pp.
471-483.
Lamb, H., 'On the Propagation of Tremors over the Surface of an Elastic Solid',
Phil. Trans. Roy. Soc. (London) A, Vol. 203, 1904, pp. 1-42.
Leadermann, H., 'Elastic and Creep Properties of Filamentous Materials and Other
High Polymers', Textile Foundation, Washington, D.C., 1943, p.175.
Lee, E.H., and Radok, J.R.M., 'The Contact Problem for Viscoelastic Bodies', J.
Appl. Mech., Vol. 27, 1960, pp. 438-444.
Lee, E.H., and Rogers, T.G., 'Solution of Viscoelastic Stress Analysis Problems
Using Measured Creep or Relaxation Functions', J. Appl. Mech., Vol. 30, 1963,
pp. 127-133.
Lee, S.L., and Ballesteros, P., 'Uniformly Loaded Rectangular Plate Supported
at the Comers', Int. J. Mech. Sci., Vol. 2, No.3, 1960, pp. 206-211.
Lee, S.L., Karasudhi, P., Zakeria, M., and Chan, K.S., 'Uniformly Loaded Or
thotropic Plates Supported at the Comers', Civil Eng. Trans., Int. Engrs., Australia, Vol. CE13, No.2, October 1971, pp. 101-106.
Lockett, F.J., 'Interpretation of Mathematical Solutions in Viscoelastic Theory
Illustrated by a Dynamic Spherical Cavity Problem', J. Mech. Phys. Solids, Vol.
9, 1961, p. 215.
Lockett, F.J., and Morland, L.W., 'Thermal Stresses in a Viscoelastic, Thin
Walled Tube with Temperature Dependent Properties', Int. J. Eng. Sci., Vol. 5,
1967, pp. 879-898.
Love, A.E.H., Some Problems of Geodynamics, Cambridge University Press,
London, 1911, 1926.
Luco, J.E., 'Torsion of a Cylinder Embedded in an Elastic Half Space', J. Appl.
Mech., Vol. 43, No.3, 1976, pp. 419-423.

424

Foundations of Solid Mechanics

84.

Ludwik, P., 'Uber den Einfluss der Deformationgeschwindigkeit bei bleibenden


Deformationen mit besonderer Berucksichtigung der Nachwirkungserscheinungen', Phys. Zeitschrift, Vol. 10, 1909, pp. 411-417.
Malvern, L.E., 'The Propagation of Longitudinal Waves of Plastic Deformation
in a Bar of Material Exhibiting a Strain-Rate Effect', J. Appl. Mech., Vol. 18,
1951, pp. 203-208.
Marcus, H., Die Theorie elastischer Gewebe, Springer, Berlin, 1932, p. 173.
Mase, G.E., Continuum Mechanics, Schaum's Outline Series, McGraw-Hill, New
York, 1970, pp. 59-61, 64, 67, 68, 74-76, 98, 99, 103, 105-107 and 109.
Michell, I.H., Proc. London Math. Soc., Vol. 31, 1900, p. 100.
Michell, I.H., Proc. London Math. Soc., Vol. 34, 1902, p. 223.
Mindlin, R.D., 'Force at a Point in the Interior of a Semi-Infinite Solid', Physics,
Vol. 7, 1936, pp. 195-202.
Mindlin, R.D., 'Force at a Point in the Interior of a Semi-Infmite Solid', Proc.
1st. Midwest. Con/. Solid Mech., Univ. Ill., Urbana, 1953, pp. 56-59.
Mises, R. von, 'Mechanik der festen Karper im plastisch deformablen Zustand',
Gottinger Nachrichten, Math.-Phys., Kl. 1913, pp. 582-592.
Mises, R. von, 'Mechanik der plastischen Formanderung von Kristallen', Z.
angew. Math. Mech., Vol. 8, 1928, pp. 161-185.
Morland, L.W., 'A Plane Problem of Rolling Contact in Linear Viscoelasticity
Theory', J. Appl. Mech., Vol 29, 1962, pp. 345-352.
Morland, L.W., 'Exact Solutions for Rolling Contact between Viscoelastic Cylinders', Quart. J. Mech. Appl. Math., Vol. 20, 1967, pp. 73-106.
Morland, L.W., and Lee, E.H., 'Stress Analysis for Linear Viscoelastic Materials
with Temperature Variation', Trans. Soc. Rheology, Vol. IV, 1960, pp. 233-263.
Muki, R., 'Asymmetric Problems of the Theory of Elasticity for a Semi-Infmite
Solid and a Thick Plate', Progress in Solid Mechanics, Vol. 1, LN. Sneddon and
R. Hill (eds.), North-Holland, Amsterdam, 1960, pp. 399-439.
Muki, R., and Dong, S.B., 'Elastostatic Far Field Behavior in a Layered Half
Space under Surface Pressure', J. Appl. Mech., Vol. 47, 1980, pp. 69-90.
Muki, R., and Sternberg, E., 'On Transient Thermal Stresses in Viscoelastic
Materials with Temperature-Dependent Properties', J. Appl. Mech., Vol. 28, 1961,
pp. 193-207.
Muki, R., and Sternberg, E., 'Elastostatic Load Transfer to a Half Space from a
Partially Embedded Axially Loaded Rod', Int. J. Solids Structures, Vol. 6, No.
1, 1970, pp. 69-90.
Nadai, A., Forschungsarb., Nos. 170 and 171, Berlin, 1915.
Nadai, A., Z. angew. Math. Mech., Vol. 2, 1922, p. 1.
Nadai, A., Elastische Platten, Berlin, 1925.
Nadai, A., Theory of Flow and Fracture of Solids, McGraw-Hill, New York; Vol.
1, 1950 and Vol. 2, 1963.

85.

86.
87.
88.
89.
90.
91.
92.
93.
94.
95.
96.
97.

98.
99.

100.

101.
102.
103.
104.

References

425

105. Naghdi, P.M., 'Stress-Strain Relations in Plasticity and Thennoplasticity', E.H.


Lee and P.S. Symond (eds.), Plasticity, 1960, pp. 121-169.
106. Neuber, H., 'Ein Neuer Ansatz zur LOsung raumlicher Probleme der Elastizitatstheorie', Z. angew. Math. Mech., Vol.14, 1934, pp. 203-212.
107. Nicholson, D.W., and Philips, A., 'On the Structure of the Theory of Viscoplasticity', Int. 1. Solids Structures, Vol. 10, 1974, pp. 149-165.
108. Niumpradit, B., and Karasudhi, P., 'Load Transfer from an Elastic Pile to a
Saturated Elastic Soil', Int. 1. Num. Anal. Methods Geomech., Vol 5, No.2,
April-June 1981, pp. 115-138.
109. Novozhilov, V.V., Theory of Elasticity, (translated by J.K. Lusher), Pergamon
Press, New York, 1961, pp. 59-61.
110. Osgood, W.R., 'Combined-Stress Tests on 24 ST Aluminum Alloy Tubes', 1.
Appl. Mech., Vol. 69, 1947, pp. 147-153.
111. Papkovich, P.F., 'Solution generale des equations differentielles fondamentales
d'elasticite, exprimee par trois fonctions harmoniques', Comptes Rendus, Vol.
195, 1932, pp. 513-515; 'Expressions generale des compos antes des tensions, ne
renfennant comme fonctions arbitraires que des fonctions harmoniques', Comptes
Rendus, Vol. 195, 1932, pp. 754-756.
112. Perzyna, P., 'The Constitutive Equations for Rate - Sensitive Plastic Materials',
Quart. Appl. Math., Vol. 20, 1963, pp. 321-332.
113. Perzyna, P., 'Fundamental Problems in Viscoplasticity', in Recent Adv. Appl.
Mech., Academic Press, New York, 1966.
114. Philips, A., and Wu, H.C., 'A Theory of Viscoplasticity', Int. 1. Solids Structures,
Vol. 9, 1973, pp. 15-30.
115 Plantema, F.J., Sandwich Construction, John Wiley and Sons, New York, 1966.
116. Prager, W., 'The Stress-Strain Laws of the Mathematical Theory of Plasticity',
1. Appl. Mech., Vol. 15, 1948, pp. 226-233, and Vol. 15, 1949, pp. 215-218.
117. Prager, W., 'The Theory of Plasticity: A Survey of Recent Achievements', Proc.
Inst. Mech. Engrs., Vol. 169, 1954, pp. 41-57.
118. Prandtl, L., 'Spannungsverteilung in plastischen Korper', Proc. 1st. Int. Congress
Appl. Mech., Delft, 1924, pp. 43-54.
119. Prandtl, L., 'Ein Gedankenmodell zur kinetischen Theorie der festen Korper', Z.
angew. Math. Mech., Vol. 8, 1928, pp. 85-106.
120. Press, F., Harkrider, D., and Seafeldt, C.A., 'A Fast, Convenient Program for
Computation of Surface-Wave Dispersion Curves in Multilayered Media', Bull.
Seismological Soc. America, Vol. 51, No.4, October 1961, pp. 495-502.
121. Quinlan, P.M., 'The Elastic Theory of Soil Dynamics', ASTM Special Technical
Publication No. 156, Symposium on Dynamic Testing of Soils, Atlanta City, N.J.,
July 1953, pp. 3-34.
122. Rajapakse, R.K.N.D., and Karasudhi, P., 'Elastostatic Infinite Elements for
Layered Half Spaces', 1. Eng. Mech., ASCE, Vol. 111, No.9, September 1985,

426

Foundations of Solid Mechanics

pp. 1144-1158.
123. Rajapakse, R.K.N.D., and Karasudhi, P., 'An Efficient Elastodynamic Infinite
Element', Int. J. Solids Structures, Vol. 22, No.6, 1986, pp. 643-657.
124. Reissner, E., 'Finite Deflection of Sandwich Plates', J. Aeronaut. Sci., Vol. 15,
1948, pp. 435-440.
125. Reissner, E., 'Small Bending and Stretching of Sandwich Shells', NACA TN 1832,
1949.
126. Reuss, E., 'Beriicksichtigung der elastischen Fonnilnderungen in der Plastizitatstheorie', Z. angew. Math. Mech., Vol. 10, 1930, pp. 266-274.
127. Robertson, I.A., 'Forced Vertical Vibration of a Rigid Circular Disc on a
Semi-Infmite Elastic Solid', Proc. Camb. Phil. Soc., Vol. 62, 1966, pp.547-553.
128. Rongved, L. 'Force at a Point in the Interior of One of the Two Joined SemiInfinite Solids', Proc. 2nd. Midwest. Con/. Solid Mech., Lafayette, Ind., 1955.
129. Sadowsky, M., 'Zweidimensionale Probleme der Elastizitlitstheone', Z. angew.
Math. Mech., Vol. 8, No.2, Apri11928, pp. 107-121.
130. Schapery, R.A., 'Approximate Methods of Transform Inversions in Viscoelastic
Stress Analysis', Proc. 4th. U.S. National Congress Appl. Mech., Vol. 2, 1962,
pp. 1075-1085.
131. Schiffman, R.L., and Fungaroli, A.A., 'Consolidation due to Tangential Loads',
Proc. 6th. Int. Con/. Soil Mech., Montreal, Vo1. 2, 1965, pp. 188-192.
132. Scholte, J.O., 'The Range of Existence of Rayleigh and Stoneley Waves', Monthly
Notices Roy. Astron. Soc.: Geophys. Suppl., Vo1. 5, 1947, pp. 120-126.
133. Schwarzl, F., and Stavennan, A.J., 'Time-Temperature Dependence of Linear
Viscoelastic Behavior', J. Appl. Phys., Vo1. 23, 1952, p. 838.
134. Scordelis, A.C., Pister, K.S., and Lin, T.Y., J. Amer. Concr. Inst., Vo1. 28, 1956,
p.241.
135. Sezawa, K., and Kanai, K., 'The Range of Possible Existence of Stoneley Waves
and Some Related Problems', Bull. Eanhquake Research Inst. (Tokyo), Vol. 17,
1939, pp. 1-8.
136. Sneddon, I.N., Fourier Transforms, McGraw-Hill, New York, 1951, pp. 5, 6, 17,
18,27,53,60,62,65,275,423,424,434,456,460,487,489 and 502.
137. Sneddon, I.N., Mixed Boundary Value Problems in Potential Theory, NorthHolland, 1966, pp. 30,40-46,64,80-133, 136, 140, 147, 176,213,222,229 and
256.
138. Sokolnikoff, I.S., Mathematical Theory of Elasticity, 2nd ed. McGraw-Hill, New
York, 1956, pp. 25-29.
139. Sokolovsky, V.V., 'The Propagation of Elastic-Viscoplastic Waves in Bars', Prik.
Mat. Mekh., Vo1. 12, 1948, pp. 251-280.
140. Spiegel, M.R., Vector Analysis, Schaum's Outline Series, McGraw-Hill, New
York, 1959, pp. 206, 207 and 212.
141. Spiegel, M.R. Advanced Calculus, Schaum's Outline Series, McGraw-Hill, New

References

427

York, 1963, pp. 263 and 272.


142. Spiegel, M.R., Laplace Transforms, Schaum's Outline Series, McGraw-Hill, New
York, 1965, pp. 46, 47, 61, 62, 73, 74, 201, 202, 207 and 210.
143. Spiegel, M.R., Advanced Mathematics, Schaum's Outline Series, McGraw-Hill,
New York, 1971, pp. 211, 219, 349, 365, 376, 377, 383, 384 and 387.
144. Stakgold, I., Boundary Value Problems of Mathematical Physics, Vol. 1, Macmillan, New York, 1967, pp. 6, 7, 18-27, 33, and 273.
145. Sternberg, E., 'On the Analysis of Thermal Stresses in Viscoelastic Solids', Proc.
3rd. Symp. Nav. Structural Mech., Macmillan, New York, 1964, p. 348.
146. Stoneley, R., 'Elastic Waves at the Surface of Separation of Two Solids', Proc.
Roy. Soc. (London), A, Vol. 106, 1924, pp. 416-428.
147. Sung, T.Y., 'Vibrations in Semi-Immite Solids due to Periodic Surface Loading',
ASTM Special Technical Publication No. 156, Symposium on Dynamic Testing
of Soils, Atlanta City, N.J., July 1953, pp. 35-68.
148. Suriyamongkol, S., 'Bending of Polygonal Plates', Japan-Thai Civil Engineering
Con/., Proc. Recent Advance Struct. Eng., Chulalongkom Univ., Bangkok,
Thailand, March 1985, pp. 85-108.
149. Swanson, S.R., 'Approximate Laplace Transform Inversion in Dynamic Viscoelasticity', J. Appl. Mech., Vol. 47, December 1980, pp. 769-774.
150. Timoshenko, S.P., and Goodier, J.N., Theory of Elasticity, 3rd ed. McGraw-Hill,
New York, 1970, pp. 122, 124, 146-149.
151. Timoshenko, S., and Woinowsky-Krieger, S., Theory of Plates and Shells,
McGraw-Hill, New York, 1959, pp. 118-120, 126-135, 139, 140, 143, 161, 187,
191, 192, 194-196,202,204, 206, 207, 210-212, 215, 216, 218-220.
152. Ting, T.C.T., 'Contact Problems in the Linear Theory of Viscoelasticity', J. Appl.
Mech;, Vol. 35, 1968, pp. 248-254.
153. Ting, T.C.T., 'A Mixed Boundary Value Problem in Viscoelasticity with TimeDependent Boundary Regions', in Developments in Mechanics, H.J. Weis, D.F.
Young, W.F. Riley, and T.R. Rogge (eds.), Vol. 5, Iowa State Univ. Press, Ames,
Iowa, 1969, p. 591.
154. Truesdell, C., 'The Mechnical Foundations of Elasticity and Fluid Dynamics', J.
Rational Mech. Anal., Vol. 1, 1952, pp. 125-300; corrected reprint in The
Mechanical Foundations of Elasticity and Fluid Dynamics, Gordon & Breach,
New York, 1966.
155. Truesdell, C., and Toupin, R.A., 'The Classical Field Theories', Encyclopedia of
Physics, S. Fliigge (ed.), Vol. III/1, Springer-Verlag, Berlin, 1960, pp. 226-793.
156. Tsien, H.S., 'A Generalization of Alfrey's Theorem for Viscoelastic Media',
Quart. Appl. Math., Vol. 8, 1950, p. 104.
157. Vijakkhana, C., Karasudhi, P., and Lee, S.L., 'Comer Supported Equilateral
Triangular Plates', Int. J. Mech. Sci., Vol. 15, 1973, pp. 123-128.
158. Wang, C.-T., Applied Elasticity, McGraw-Hill, New York, 1953, pp. 22 and 23.

428

Foundations of Solid Mechanics

159. Watson, G.N., A Treatise on the Theory 0/ Bessel Functions, Cambridge University Press, 1966.
160. Whittaker, E.T., and Watson, G,N., A Course on Modern Analysis, Cambridge
University Press, New York, 4th edition reprinted, 1965, p. 172.
161. Widder, n.v., The Laplace Trans/orm, Princeton University Press, Princeton,
N.J., 1946.
162. Williams, M.L., 'Stress Singularities Resulting from Various Boundary Conditions in Angular Corners of Plates in Extension', J. Appl. Mech., Vol. 19, 1952,
pp. 526-528.
163. Wylie, C.R., Advanced Engineering Mathematics, McGraw-Hill Kogakusha,
Tokyo, 4th. Edition, 1975, pp. 512, 513, 515 and 528.
164. Ziegler, H., 'A Modification of Prager's Hardening Rules', Quart. Appl. Math.,
Vol. 17, 1959, pp. 55,..65.
165. Zienkiewicz, O.C., and Cormeau, I.C., 'Viscoplasticity-Plasticity and Creep in
Elastic Solids, A Unified Numerical Solution Approach', Int. J. Num. Methods
Eng., Vol. 8, 1974, pp. 821-845.

AUTHOR INDEX
E

Essenburg, F., 272


Euler, L., 52, 379, 380
Ewing, N.M., 316,329,334

Abramowitz, M, 25, 28
Achenbach,I.D., 107,292
Agabein, M.E., 178
Airy, G.B., 114
Alam, K.M.A., 188
Alfrey, T., 33, 268
Almansi, E., 394
Alvappillai, A., 231
Apirathvorakij, V., 217, 231

F
Ferry, I.D., 260
Folie, C.M., 189
Frazer, R.A., 16
Fung, Y.C., 361, 364
Fungaroli, A.A., 229

Ballesteros, P., 178


Bauschinger, I., 363
Bellman, R., 34
Beltrami, E., 103
Betti, E., 100
Bingham, E.C., 374
Biot, M.A., 108
Boresi, A.P., 132
Bridgman, P.W., 357, 361
Budiansky, B., 369

Galerkin, B., 105


Gilman, 1.1., 375
Goodier, I.N., 132
Gottenberg, W.G., 270,271,278
Graham, G.A.C., 272, 273
Green, G., 394
H

Haar, D. ter, 33
Habip, L.M., 189
Hamel, G., 394
Hamming, R.W., 224
Handelman, G.H., 369
Harkrider, D.G., 329, 334
Hencky, H., 369
Hill, R., 374
Hodge, P.G., Ir., 368
Hooke, R., 88
Hopkins, IL., 224
Hovichitr, 1., 188
Hsu, H.P., 31
Hughes, T.I.R., 375, 376, 377, 400
Hunter, S.C., 272, 273
Hwang, B.Y., 217, 223

Cagniard, L., 325


Cauchy, AL., 86,394
Cesaro, E., 64
Chan, K.S., 189, 211, 217
Christensen, R.M., 270, 271, 272, 274, 278, 280
Churchill, R.V., 26, 117, 335
Collar, A.R., 16
Cormeau, 1., 375, 376, 377
Cost, T.L., 32
Crossland, B., 361
D

D' Alembert, J.le R., 51, 379


Dong, S.B., 225
Drucker, D.C., 364, 367
Duncan, W.J., 16
Dundurs, I., 178, 187

J
Iardetzky, W.S., 316, 329, 334
Iou, T.S., 183, 189
429

Foundations of Solid Mechanics

430
K

Kalaba, R.E., 34
Kalker, J.J., 273
Kanai, K., 325
Kanok-Nukulchai, W., 400
Kao, J.S., 189
Karasudhi, P., 180, 183, 188, 189,211,217,223,
225, 226, 231, 348, 353
Kannan, T. von, 356
Karunasena, W.M., 231
Keer, L.M., 353
Keller, H.B., 296
Kirchhoff, G., 159, 386
Kliushnikov, V.D., 369
Koiter, W.T., 364, 366, 367
Koppe, H., 325
Kosherick, HJ., 178, 187

Moo, R., 203, 217, 222, 223, 225, 274, 287


N

Nadai, A., 61, 168, 180, 369


Naghdi, P.M., 367, 369
Navier, CL.M.H., WI, 167
Neuber, H., 106
Newton, I., 51
Ng, H.K., 189
Nicholson, D.W., 375
Nishino, P., 188
Niumpradit, B., 217, 231
Novozhilov, V.V., 414

o
Osgood, W.R., 362

Lagrange, J.L., 380


Lamb, H., 337
Lame, G., 94
Leadennann, H., 260
Lee, E.H., 224, 260, 272, 274
Lee, S.L., 178, 180, 187, 188, 189, 211, 353
Levy, M., 168
Lin, C.C., 369
Lin, T.Y., 180
Liyanage, K.K., 217
Lockett, P.J., 271, 274
Lockett, J., 34
Love, A.E.H., 325
Luco, J.E., 218, 224
Ludwik, P., 374
Lynn, P.P., 132
M

Malvern, L.E., 375


Marcus, H., 180
Mase, G.E., 76
Maxwell, J.C., 100
Michell, J.H., 103, 171
Mindlin, RD., 211
Mises, R von, 61,362,366
Morland, L.W., 260, 273, 274

Papkovich, P.F., 106


Perzyna, P., 375
Philips, A., 375
Piola, G., 386
Pister, K.S., 180
Plantema, P.J., 189
Poisson, S.D., 94
Prager, W., 360, 366, 367, 368, 369
Prandtl, L., 370, 374
Prechaverakul, S., 231
Press, F., 316, 329, 334
Puswewala, V.G.A., 231
Q

Quinlan, P.M., 353


R

Radok, J.RM., 272


Rajapakse, R.K.N.D., 217, 223, 225, 226, 348
Rayleigh, Lord (J.W. Strutt), 321
Reissner, E., 189
Reuss, E., 362, 370
Robertson, LA., 353

Author Index

431

Rogers, T.G., 224, 274


Rongved, L., 211

s
Sadowsky, M., 208
Saint Venant, B. de, 394
Schapery, R.A., 33
Schiffman, R.L., 229
Scholte, J.G., 325
Schreiner, RN., 271
Schwarzl, F., 260
Scordelis, A.C., 180
Seafeldt, C.A., 334
Sezawa, K., 325
Sneddon, LN., 198, 204, 208, 217
Sokolnikoff, LS., 64, 106
Sokolovsky, V.V., 375
Spiegel, M.R., 14,32,243,279,282,334,412
Stakgold, I., 24
Staverman, AJ., 260
Stegun, LA., 25, 28
Sternberg, E., 217,222,223,274,287
Stoneley, R., 325
Sung, T.Y., 353
Suriyamongkol, S., 180
Swanson, S.R., 34

T
Tansirikongkol, V., 183, 189
Taylor, R.L., 375, 376, 377, 400

Timoshenko, S.P., 132, 168


Ting, T.C.T., 273
Toupin, R.A., 419
Tresca, H., 361
Truesdell, C., 379, 419
Tsien, H.S., 268

V
Vijakkhana, C., 180

w
Wang, C.-T., 76
Watson, GN., 25, 198
Whittaker, E.T., 198
Widder, D.V., 33
Williams, M.L., 218
Woinowsky-Krieger, S., 168
Wu, H.C., 375
Wylie, C.R., 180
Y
Young, T., 94
Z
Zakeria, M., 189
Ziegler, H., 368
Zienkiewicz, O.C., 375, 376, 377

SUBJECT INDEX
Cauchy's principal value. 334. 335. 347
Cauchy-Riemann conditions. 117. 120
Characteristic
equation. 13. 16
value. 13
vector. 13
Christoffel symbol. 41. 43. 402
of the second kind. 41. 43. 402
Coefficient of
consolidation. 109
permeability. 109
Cofactor. 11.40
Collocation method. 33. 34
Column matrix. 10
Comma-subscript convention. 18.41.297
Compatibility conditions. 64. 82. 100. 102. 113.

Admissible transformation. 36. 37. 386


Airy function. 116. 127. 133, 135. 136. 137
Almansi's strain tensor. 394
Analytic continuation. 26
Angular frequency. 254. 290. 321
An~otropy.

90.268. 362. 364. 368

Anticlastic bending. 166. 181


Antiplane problems. 197.201.202.208.218.326.
346.354

Atrenuatingwave, 289. 339


B

Base vector. 4, 38. 41. 43


Bauscbinger effect, 363. 364. 368
Beltrami-Michell compatibility conditions. 103.

267. 269. 412-416

Complementary solution. 104. 121. 164. 186. 197


Complete quietude. 28
Complex
compliance. 254. 259. 262. 282. 283
modulus, 254. 257. 262. 275. 278. 347
plane. 334. 347
Components. 1.4
Conjugate. 119
harmonic functions. 117. 120
Conservation of
energy. 87
mass. 384. 398
Conservative
force. 103, 105. 113. 115. 125
force porentials. 103. 114
Cons~tency condition. 360
Constitutive relationship. 99. 305. 318. 396. 416
Continuum mechanics. 1. 18
Contraction. 5. 9. 38
Contravariance. 6. 38
Contravariant
base vector. 40.41
component, 41. 402. 406. 415
metric tensor. 40
tensors. 38
Convexity. 367
Convolution integral. 29. 239. 273
Coordinate
curve. 37. 39

267.269

Bending moment. 159


Bessel functions. 25. 204. 333
Biharmonic
equation. 103. 116
function, 103. 117. 121
operator. 103, 114. 164. 169. 204
Binormal. 19
Body
force. 47. 69. 86. 391. 393
wave solution. 335. 342
wavenumber. 321. 347
waves. 321
Boundary
conditions. 99. 109. 410
value problem. 99. 168. 273. 373. 393
Branch points. 32, 335. 347
Bulk modulus, 94. 266. 283. 359. 376
C

Canonical form. 15
Carrier wave. 291
Cartesian
components, 4. 19.378.388.406.408.410
coordinares. 36
coordinare system. 4. 7. 17.406
tensor. 1. 7. 9
433

Foundations of Solid Mechanics

434
surface, 37, 40
systems, 4, 6, 43, 406, 412
Comer force, 161, 162, 167
Covariance, 6, 38
Covariant
base vector, 37,41,43,401
component, 42, 402, 406, 415
derivative, 42
metric tensor, 38, 402
partial differentiation, 43, 44
tensors, 38
Cramer's rule, 13
Creep, 237, 377
compliance, 238, 240, 257, 263
function, 244
law, 240, 244, 245, 255
Cross product, 3
Curvature, 20, 155, 302
Curvilinear
coordinates, 36,401,410
coordinate system, 6, 38,43,67,406
Cylindrical
bending, 165
coordinates, 67, 114, 203
rotation components, 74
wavefront, 309, 350
D

D' Alembert's force, 51, 52, 87, 321, 404, 408


Darcy's law, 109, 229
Dashpot, 236, 240, 244
Deflection, 154
Deformation, 1,47,61,378,394
theories of plasticity, 368
Determinant, 11, 36, 385
Deviatoric stress, 60, 361
Differential geometry, 19
Dilatation, 63, 69, 71, 329,
Dilatational wave, 295
Dirac-delta function, 23, 183, 184, 185, 211, 238,
261,264
Direction cosines, 5
Dispersive wave, 291, 324, 325, 327, 334
Displacement, 2, 61, 379
interpolation functions, 229, 400
potentials, 104, 106,203,229,329
Divergence theorem, 18, 45, 53, 86, 99, 297, 302,
392

Dot product, 3,40


Dual integral equations, 208, 209
Dummy index, 6, 18
Dynamic jump condition, 301, 305
E

Effective
plastic strain increment, 370
stress, 370
Eiconal equation, 301, 304, 307
Eigenvalue, 12, 13, 16, 55, 56, 306, 322
problem, 12, 13, 16, 322
Eigenvector, 13, 14, 15, 45, 56, 306, 322
Einstein summation convention, 5, 38, 297
Elastic
range, 355
solid, 86, 88, 93
strrun, 357, 358, 364, 372, 374
Elasticity, 86, 235
Elastic-viscoelastic
analogy, 270
correspondence principle, 270, 279
Elastostatics, 111, 196
Energy dissipation per cycle, 260, 262, 283
Energy identity, 267
Engineering shear strrun, 62, 90
Equations of
continuity, 384, 392
motion,49,53,69,296, 391, 392, 393
Equilibrium, 49, 86, 87, 102,402
equations, 159, 269
Equivalent
plastic strain increment, 370
stress, 370
Equivoluminal wave, 294
Euclidean space, I, 4, 11, 39, 413
Euler's laws of motion, 52, 53, 392
Eulerian
coordinates, 379, 395
description, 379, 394, 397, 416
stress tensor, 386, 387
Excess pore pressure, 108
gradient, 109, 229

F
Factor of stress singularity, 218
Far field, 122, 225, 334, 348
Field equations, 99, 164, 196, 203, 229

Subject Index

435

Final-value theorem, 30
Finite
deformation, 47, 378, 416, 417
elements, 225, 228, 348
First
order tensor, 6
Piola-Kirchhoff stress tensor, 386
Flexural rigidity, 162
Flow rule, 365, 366
Fourier transform, 30, 198, 255, 259, 270
Fredholm integral equation, 217, 223
Free index, 6, 8, 18
Frenet-Serret formula, 20, 46, 308
G

Galerkin's vector, 105


Gamma function, 27, 29, 32
Gauss quadrature formula, 229
Gauss's divergence theorem, 18,53,86,297,302,
383,393

Generalized
coordinates, 15
Kelvin model, 237
Maxwell model, 237
Generic letter, 5
Geometrical nonlinearity, 378
Gibbs notation, 2, 38, 44
Glassy
creep compliance, 244
modulus, 245
Gradient, 23, 109, 229
Gram-Schmidt orthogonalization process, 180
Green's
strain tensor, 394, 398, 408
theorem, 18, 157
Group velocity, 292
H

Half
plane, 121, 140, 144, 145, 197, 218, 233, 323,
324, 326, 327, 346, 352, 353, 354
space, 203, 207, 209, 217, 225,232,233,272,
277, 334,343, 347,348,353,354
Hankel
function, 25,28, 338
transform, 204, 230, 330
Hardening rules, 368, 374

Harmonic
equation, 103
function, 103, 104, 117
operator, 113
surface wave of Rayleigh type, 321
Heaviside
expansion formula, 243, 279, 282
step function, 24, 238, 261, 264
Hooke's law, 88, 357, 359, 399, 417
Hookean elastic solid, 235, 283
Hoop stress, 124, 141
Huyghen's principle, 308
Hydrostatic
pressure, 95, 356, 357, 361, 368
stress, 60
Hyperdomain, 297
Hype~urface,

297, 360, 367


I

Ideal plastic
deformation, 365
range, 355
solids, 360, 362, 366, 373
Identity matrix, 10, 12, 13
In-plane problems, 73, 189, 196, 320
Incompressible material, 95, 280
Incremental
formulation for isotropic hardening, 369
stress-strain relationship, 366
theories of plasticity, 368
Indicial notation, 5,44, 378
Infinite
elements, 225, 348
plane, 121, 142, 143, 197

space, 203
Initial-value theorem, 30
Inner product, 9
Interpolation function, 229, 399
Invariant, 54, 55, 60, 102
Inverse
Laplace transform, 32,231, 243,284, 287
matrix, 12
Irrotational wave, 295
Isochoric deformation, 83
Isothermal harmonic vibration, 262, 282
Isotropic
elastic solids, 93, 292
hardening, 368, 370

Subject Index

436
materials, 93, 372
plane problems, 112, 113
~ilierny, 360, 370

J
Jacobian, 36, 303, 309
Jump, 297, 301, 304
K

Kelvin
functions, 27, 182
model,237
Kinematic
hardening, 368
jump condition, 301, 305
Kinetic energy, 87,288
Kirchhoff
shear forces, 159
stress tensor, 387
Kronecker
delta, 8,44
symbol, 13, 18, 40, 43
L

Lagrange polynomial, 228


Lagrangian
coonlinates, 379, 397
description, 379, 393, 396
stress tensor, 386, 398
Lame's
constants, 94, 347
strain potential, 105
Laplace
equation, 103
expansion, 11
operator, 103, 113, 115
ttansform, 28,230,240
Laplacian, 18, 121
Laws of ttansformation, I, 8, 17
Leibnitz's rule for differentiating an integral, 30,
263
Lerch's uniqueness ilieorem, 30
Levy-Nadai solution, 168, 181, 183, 186
Linear
elastic solid, 88, 235
momentum, 52, 391, 393
spring model, 236

ttansformation, 15
viscoelasticity, 235
Linearity, 29, 235, 367
Load ttansfer, 217, 231, 348
Loading function, 360
Loss
angle, 260, 276
modulus, 260
Love
wave equation, 326, 327
wave velocity, 327
wavenumbers, 326
waves, 326, 334
M
Mass density, 51, 384, 386
Material
derivative, 380
description, 379
Maxwell model, 237, 318
Maxwell-Betti reciprocal ilieorem, 100, 267, 416
Metric
form, 39
tensor, 38
Michell's solution, 172, 184
Mindlin's solution, 216, 232
Mises yield condition, 362, 366, 371, 373
Modulation, 292
Modulus of foundation, 181
Moment
equilibrium, 51, 404
of momentum, 52, 391

NaturaI state, 86, 89, 235


Navier
equations, 101,229,267,348
solution, 167, 181
Near field, 225, 348
Neutral loading, 360
Newton's second law of motion, 51, 52, 391
Newtonian viscous fluids, 235
Nonstationary surface of discontinuity, 305
Normal
coordinates, 15, 397
strain, 62
stress, 48

Subject Index

437

vector, 23, 312, 366, 367


Normality,367

o
Octahedral
plane, 60, 80
shear stress, 60, 61
One-to-one
correspondence, 35, 384, 397
mapping, 34, 378, 397, 398
Orthogonal
curvilinear coordinate system, 38, 410
eigenvectors, 14, 56
Orthogonality condition, 8
Orthogonalization process, 180
Orthonormality conditions, 8
Orthotropic elastic solids, 91, 111
Outer product, 9
Outgoing wave, 289
p

P-wave, 292, 306


P-wavenumber, 321
Papkovich potentials, 211
Particular solution, 104, 121, 164
Period, 176, 204, 291
Permutation tensor, 9, 44, 383
Phase
angle, 254
velocity, 289, 295
Physical components of tensors, 406
Piola-Kirchhoff stress tensors, 386, 387
Plane
biharmonic equation, 116
biharmonic operator, 114
harmonic function, 117, 197
harmonic wave, 290, 291
of incidence, 311
strain, 83, 112, 196,217, 320
stress,96, Ill, 162, 197,217,316,320
wave, 295
Plastic
flow rule, 365
strain, 357, 358
strain tensor, 358
work increment, 370, 373
Plate

bending, 154
on elastic foundation, 180, 187
Poisson's ratio, 94, 96
Polar coordinates, 115
Poles, 32, 334, 335, 347
Polynomial approximation, 28
Porosity, 109
Position vector, 17,37,381,399,402
Positive definite, IS, 91, 107, 265, 386
Potential, 103, 104, lOS, 113, 114,203,229,329
energy, 86, 288
Prandtl-Reuss equation, 370
Prefactor matrix, 10
Pressure wave speed, 106, 292
Principal
axis, 55
direction, 55
normal,19
plane, 55
stress, 55
stress space, 361
Principle of virtual work, 98, 156
Proper transformation, 36, 37, 386
Pseudo
frequency, 262, 283
time, 261, 273, 283, 285, 287
Pulse, 24
Pure torsion bending, 166, 181
x-plane, 361, 362, 368

Q
Quadratic form, 15
Quarter planes, 218
Quasi-static problems, 34, 108,229,267,270,278
Quiescent initial conditions, 236
R
Radius
of curvature, 20
of torsion, 20
Rayleigh
wave velocity, 324
wavenumbers, 322
Rectangular
Cartesian coordinate system, 4, 7, 410
coordinate system, 4, 7, 39

438

Subject Index

Reduced
firequency, 262, 283
time, 261, 283, 285
Reflected wave, 311
Reformed
Green's strain tensor, 398
Kirchhoff stress tensor, 398
Lagrangian coordinates, 397
Lagrangian description, 396
Lagrangian stress tensor, 398
Relaxation, 238, 244, 377
function, 244
law, 240, 244, 245, 251, 262
modulus, 238
time, 245
Residue, 335, 341, 346, 348
theorem, 337
Retardation_time, 245
Riemann-Christoffel
covariant curvature tensor, 66, 412
tensor, 412
Riemann-Lebesgue lemma, 198
Riemannian space, 39
Rigid-body
displacement, 64, 69, 72, 112, 116
motion, 394
rotation, 74
translation, 74
Rigid
-plastic solids, 374
-strain-hardening, 373
Rotation
tensor, 63
vector, 63
Rotational wave, 293
Rubbery modulus, 244

s
S-wave, 293, 306, 307
S-wavenumber, 321
Sandwich plate, 189
Saturated porous elastic media, 108, 229
Scalar product, 3
Schapery's approximation, 33, 231
Second
order tensors, 6
Piola-Kirchhoff stress tensor, 387
Separation of variables, 268, 270

Shear
discontinuity, 306
modulus, 94, 97
strain,62
stress, 48, 388
stress resultant, 158
wave, 106,293
wave speed, 106,293
Shift factor, 261, 283
Shifting property, 29
Signum, 31
Simple
closed curve, 18, 45, 156
wave equation, 290, 294, 295
wave function, 289
Singular
contraction, 228
matrix,12
Skew coordinates, 43
Small deformation, 47, 52, 390, 396
Space, I, 4, 11, 39, 413
curve, 19, 20, 46
Spatial description, 379
Specific loss, 260
Spherical
Bessel function, 27
coordinates,67,226,293
Hankel function, 28
rotation components, 74
stress, 60
waveftunt, 294, 295,309,350
Standard linear solid model, 237
Standing wave, 289
Static equilibrium, 51
Stationary shear stresses, 57, 58
Steady
state harmonic function, 253, 269
tensor field, 17
Stieltjes integral, 239
Stiffness tensor, 90, 399
Stirling's formula, 32
Stoneley wave, 325
Strain
deviator, 63,94
energy, 86, 98, 411
energy function, 87, 107, 395, 398, 416
hardening, 364
hardening hypothesis, 371
tensor, 61, 394, 398, 402,408,413,416

439

Subject Index

Stress
component, 48, 49, 386, 388, 390
concentrntion factor, 124
deviator, 60, 94
-free state, 264, 273, 285, 286
invariants, 54, 60
potential, 104
resultant, 156, 159
space, 371, 366, 370
symmetry, 51, 54, 393
tensor, 48, 386, 387, 389, 398,404,410
vector, 50, 386, 388, 390, 398, 406, 410
Summation convention, 5, 10, 38, 297
Supplemented shear forces, 159
Surface
trnction, 47,48, 50, 391
waves, 320, 321, 325, 334,342, 345, 347, 350
wave solution, 335, 342, 346
Symbolic notation, 2, 9, 17, 38

T
Tangent, 19,46
Tensor
equation, 1, 5, 38
field, 17
Theory of plastic potential, 366
Thermorheologically simple solids, 260, 273
Torsion, 20, 308
Total differential, 22, 87, 396
Traction,47,48,50,403
Transformation
law, 1,8
of coordinates, 34
Translation property, 29
Transmitted wave, 311
Transpose matrix, 10
Transverse
discontinuity, 306
wave, 293
Tresca's yield condition, 361, 366, 368
Twisting moment, 159

u
Uniaxial stress, 94, 266, 273, 372
Unified Hooke's law, 112
Uniqueness theorem, 30
Unit eigenvector, 14

Universal stress-strain curve, 371


Unloading, 88, 360

v
Variational symbol, 98, 156
Vector
addition, 2
components, 6
equations, 2
field, 17, 18
product, 3
subtrnction, 3
triad, 4
Velocity jump, 304, 305
Virtual
displacement, 410
quantity, 98, 156
work 98, 156,410
Viscoelasticity, 235
Viscoplastic
model,374
strain rate, 374
Viscoplasticity, 374
Viscous dashpot model, 236, 240
Voigt model, 237, 241

w
Wave, 288
equation, 106, 290, 292, 293, 295
group, 292
propagation, 288
Wavefront, 288, 294, 295, 297, 299, 310, 350
function, 299
Wavelength,291
Wavenumber, 291, 321, 322, 325, 326
Widder's general inversion formula, 33
Work hardening, 355, 364, 367
hypothesis, 370
parameter, 359
range, 355

y
Yield
condition, 359, 361, 366, 369, 371, 373, 374
function, 359, 365, 370, 376
Young's modulus, 94, 96

Mechanics
From 1990, books on the subject of mechanics will be published under two series.

SOLID MECHANICS AND ITS APPLICATIONS


Series Editor: G.M.L. Gladwell
Aims and Scope of the Series
The fundamental questions arising in mechanics are: Why?, How?, and How much? The aim of this
series is to provide lucid accounts written by authoritative researchers giving vision and insight in
answering these questions on the subject of mechanics as it relates to solids. The scope of the series
covers the entire spectrum of solid mechanics. Thus it includes the foundation of mechanics;
variational fonnulations; computational mechanics; statics, kinematics and dynamics of rigid and
elastic bodies; vibrations of solids and structures; dynamical systems and chaos; the theories of
elasticity, plasticity and viscoelasticity; composite materials; rods, beams, shells and membranes;
structural control and stability; soils, rocks and geomechanics; fracture; tribology; experimental
mechanics; biomechanics and machine design.

1.
2.
3.

R.T. Haftka, Z. Gilrdal and M.P. Kamat: Elements of Structural Optimization. 2nd rev.ed.,
1990
ISBN 0-7923-0608-2
J.J. Kalker: Three-Dimensional Elastic Bodies in Rolling Contact. 1990
ISBN 0-7923-0712-7
ISBN 0-7923-0772-0
P. Karasudhi: Foundations of Solid Mechanics.

FLUID MECHANICS AND ITS APPLICATIONS


Series Editor: R. Moreau
Aims and Scope of the Series
The purpose of this series is to focus on subjects in which fluid mechanics plays a fundamental
role. As well as the more traditional applications of aeronautics, hydraulics, heat and mass transfer
etc., books will be published dealing with topics which are currently in a state of rapid development, such as turbulence, suspensions and multiphase fluids, super and hypersonic flows and
numerical modelling techniques. It is a widely held view that it is the interdisciplinary subjects that
will receive intense scientific attention, bringing them to the forefront of technological advancement Fluids have the ability to transport matter and its properties as well as transmit force,
therefore fluid mechanics is a subject that is particularly open to cross fertilisation with other
sciences and disciplines of engineering. The subject of fluid mechanics will be highly relevant in
domains such as chemical, metallurgical, biological and ecological engineering. This series is
particularly open to such new multidisciplinary domains.

1.
2.
3.
4.

M. Lesieur: Turbulence in Fluids. 2nd rev. ed., 1990


ISBN 0-7923-0645-7
O. Metais and M. Lesieur (eds.): Turbulence and Coherent Structures. 1991
ISBN 0-7923-0646-5
R. Moreau: Magnetohydrodynamics. 1990
ISBN 0-7923-0937-5
ISBN 0-7923-1020-9
E. Cousto1s (ed.): Turbulence Control by Passive Means. 1990

Kluwer Academic Publishers - Dordrecht I Boston I London

Mechanics
From 1990, books on the subject of mechanics will be published under two series:
FLUID MECHANICS AND ITS APPLICATIONS
Series Editor: R.I. Moreau
SOLID MECHANICS AND ITS APPLICATIONS
Series Editor: G.M.L. Gladwell
Prior to 1990, the books listed below were published in the respective series indicated below.
MECHANICS: DYNAMICAL SYSTEMS
Editors: L. Meirovitch and G.lE. Oravas
1. E.H. Dowell: Aeroelasticity of Plates and Shells. 1975
ISBN 90-286-0404-9
2. D.G.B. Edelen: Lagrangian Mechanics of Nonconservative Nonholonomic Systems.
1977
ISBN 90-286-0077-9
3. J.L. Junkins: An Introduction to Optimal Estimation of Dynamical Systems. 1978
ISBN 90-286-0067-1
4. E.H. Dowell (ed.), H.C. Curtiss Jr., R.H. Scanlan and F. Sisto: A Modern Course in
Aeroelasticity.
Revised and enlarged edition see under Volume 11
5. L. Meirovitch: Computational Methods in Structural Dynamics. 1980
ISBN 90-286-0580-0
6. B. Skalmierski and A. Tylikowski: Stochastic Processes in Dynamics. Revised and
enlarged translation. 1982
ISBN 90-247-2686-7
7. P.C. MUller and W.O. Schiehlen: Linear Vibrations. A Theoretical Treatment of MultiISBN 90-247-2983-1
degree-of-freedom Vibrating Systems. 1985
8. Gh. Buzdugan, E. MiMilescu and M. Rade: Vibration Measurement. 1986
ISBN 90-247-3111-9
9. G.M.L. Gladwell: Inverse Problems in Vibration. 1987
ISBN 90-247-3408-8
10. G.I. Schueller and M. Shinozuka: Stochastic Methods in Structural Dynamics. 1987
ISBN 90-247-3611-0
11. E.H. Dowell (ed.), H.C. Curtiss Jr., R.H. Scanlan and F. Sisto: A Modern Course in
Aeroelasticity. Second revised and enlarged edition (of Volume 4). 1989
ISBN Hb 0-7923-0062-9; Ph 0-7923-0185-4
12. W. Szempliriska-Stupnicka: The Behavior of Nonlinear Vibrating Systems. Volume I:
Fundamental Concepts and Methods: Applications to Single~Degree-of-Freedom
Systems. 1990
ISBN 0-7923-0368-7
13. W. Szemplidska-Stupnicka: The Behavior of Nonlinear Vibrating Systems. Volume II:
Advanced Concepts and Applications to Multi-Degree-of-Freedom Systems. 1990
ISBN 0-7923-0369-5
Set ISBN (Vols. 12-13) 0-7923-0370-9
MECHANICS OF STRUCTURAL SYSTEMS
Editors: J.S. Przemieniecki and G.lE. Oravas
1. L. FrYba: Vibration of Solids and Structures under Moving Loads. 1970
ISBN 90-01-32420-2
2. K. Marguerre and K. Wi>lfel: Mechanics of Vibration. 1979
ISBN 90-286-0086-8

Mechanics
3. E.B. Magrab: Vibrations of Elastic Structural Members. 1979
ISBN 90-286-0207-0
4. R.T. Haftka and M.P. Kamat: Elements of Structural Optimization. 1985
Revised and enlarged edition see under Solid Mechanics and Its Applications, Volume 1

5. 1.R. Vinson and R.L. Sierakowski: The Behavior of Structures Composed of Composite
Materials. 1986
ISBN Hb 90-247-3125-9; Pb 90-247-3578-5
6. B.E. Gatewood: Virtual Principles in Aircraft Structures. Volume 1: Analysis. 1989
ISBN 90-247-3754-0
7. B.B. Gatewood: Virtual Principles in Aircraft Structures. Volume 2: Design, Plates,
Finite Elements. 1989
ISBN 90-247-3755-9
Set (Gatewood 1 + 2) ISBN 90-247-3753-2
MECHANICS OF ELASTIC AND INELASTIC SOLIDS
Editors: S. Nemat-Nasser and G.lE. Oravas
1. G.M.L. Gladwell: Contact Problems in the Classical Theory of Elasticity. 1980
ISBN Hb 90-286-0440-5; Pb 90-286-0760-9
2. G. Wempner: Mechanics of Solids with Applications to Thin Bodies. 1981
ISBN 9O-286-0880-X
3. T. Mura: Micromechanics of Defects in Solids. 2nd revised edition, 1987
ISBN 90-247-3343-X
4. R.G. Payton: Elastic Wave Propagation in Transversely Isotropic Media. 1983
ISBN 90-247-2843-6
5. S. Nemat-Nasser, H. Abe and S. Hirakawa (eds.): Hydraulic Fracturing and GeotherISBN 90-247-2855-X
mal Energy. 1983
6. S. Nemat-Nasser, R.I. Asaro and G.A. Hegemier (eds.): Theoretical Foundation for
Large-scale Computations of Nonlinear Material Behavior. 1984 ISBN 90-247-3092-9
7. N. Cristescu: Rock Rheology. 1988
ISBN 90-247-3660-9
8. G.I.N. Rozvany: Structural Design via Optimality Criteria. The Prager Approach to
Structural Optimization. 1989
ISBN 90-247-3613-7

MECHANICS OF SURFACE STRUCTURES


Editors: W.A. Nash and G.lE. Oravas
1. P. Seide: Small Elastic Deformations of Thin Shells. 1975
ISBN 90-286-0064-7
2. V. Panc: Theories of Elastic Plates. 1975
ISBN 9O-286-0104-X
3. 1.L. Nowinski: Theory ofThermoelasticity with Applications. 1978
ISBN 90-286-0457-X
4. S. Lukasiewicz: Local Loads in Plates and Shells. 1979
ISBN 90-286-0047-7
5. C. Fii't: Statics, Formfinding and Dynamics of Air-supported Membrane Structures.
1983
ISBN 90-247-2672-7
6. Y. Kai-yuan (ed.): Progress in Applied Mechanics. The Chien Wei-zang Anniversary
Volume. 1987
ISBN 90-247-3249-2
7. R. Negrutiu: Elastic Analysis of Slab Structures. 1987
ISBN 90-247-3367-7
8. J.R. Vinson: The Behavior of Thin Walled Structures. Beams, Plates, and Shells. 1988
ISBN Hb 90-247-3663-3; Pb 90-247-3664-1

Mechanics
MECHANICS OF FLUIDS AND TRANSPORT PROCESSES
Editors: R.J. Moreau and G.}E. Oravas

1. J. Happel and H. Brenner: Low Reynolds Number Hydrodynamics. With Special


Applications to Particular Media. 1983
ISBN Hb 90-01-37115-9; Pb 90-247-2877-0
ISBN 90-247-2687-5
2. S. Zahorski: Mechanics of Viscoelastic Fluids. 1982
3. J.A. Sparenberg: Elements ofHydrodynamics Propulsion. 1984 ISBN 90-247-2871-1
4. B.K. Shivamoggi: Theoretical Fluid Dynamics. 1984
ISBN 90-247-2999-8
5. R. Timman, A.J. Hermans and G.C. Hsiao: Water Waves and Ship Hydrodynamics. An
Introduction. 1985
ISBN 90-247-3218-2
6. M. Lesieur: Turbulence in Fluids. Stochastic and Numerical Modelling. 1987
ISBN 90-247-3470-3
7. L.A. Lliboutry: Very Slow Flows of Solids. Basics of Modeling in Geodynamics and
Glaciology. 1987
ISBN 90-247-3482-7
8. B.K. Shivamoggi: Introduction to Nonlinear Fluid-Plasma Waves. 1988
ISBN 90-247-3662-5
9. v. Bojarevi~s, Va. Freibergs, E.I. Shilova and E.V. Shcherbinin: Electrically Induced
Vortical Flows. 1989
ISBN 90-247-3712-5
10. J. Lielpeteris and R. Moreau (eds.): Liquid Metal Magnetohydrodynamics. 1989
ISBN 0-7923-0344-X
MECHANICS OF ELASTIC STABILITY
Editors: H. Leipholz and G.}E. Oravas

1. H. Leipholz: Theory of Elasticity. 1974


ISBN 90-286-0193-7
2. L. Librescu: Elastostatics and Kinetics of Aniosotropic and Heterogeneous Shell-type
Structures. 1975
ISBN 90-286-0035-3
3. C.L. Dym: Stability Theory and Its Applications to Structural Mechanics. 1974
ISBN 90-286-0094-9
4. K. Huseyin: Nonlinear Theory of Elastic Stability. 1975
ISBN 90-286-0344-1
5. H. Leipholz: Direct Variational Methods and Eigenvalue Problems in Engineering.
1977
ISBN 90-286-0106-6
6. K. Huseyin: Vibrations and Stability of Multiple Parameter Systems. 1978
ISBN 90-286-0136-8
ISBN 90-286-0050-7
7. H. Leipholz: Stability of Elastic Systems. 1980
8. V.V. Bolotin: Random Vibrations of Elastic Systems. 1984
ISBN 90-247-2981-5
ISBN 90-247-3099-6
9. D. Bushnell: Computerized Buckling Analysis of Shells. 1985
10. L.M. Kachanov: Introduction to Continuum Damage Mechanics. 1986
ISBN 90-247-3319-7
11. H.H.E. Leipholz and M. Abdel-Rohman: Control of Structures. 1986
ISBN 90-247-3321-9
12. H.E. Lindberg and A.L. Florence: Dynamic Pulse Buckling. Theory and Experiment.
1987
ISBN 90-247-3566-1
13. A. Gajewski and M. Zyczkowski: Optimal Structural Design under Stability Constraints.1988
ISBN 90-247-3612-9

Mechanics
MECHANICS: ANALYSIS
Editors: V.J. Mizel and G.1E. Oravas
l. M.A. Krasnoselskii, P.P. Zabreiko, E.I. Pustylnik and P.E. Sbolevskii: Integral
Operators in Spaces of Summable Functions. 1976
ISBN 90-286-0294-1
2. V.V. Ivanov: The Theory of Approximate Methods and Their Application to the
ISBN 90-286-0036-1
Numerical Solution of Singular Integral Equations. 1976
3. A. Kufner, O. John and S. Puclk: Function Spaces. 1977
ISBN 90-286-0015-9
4. S.G. Mikhlin: Approximation on a Rectangular Grid. With Application to Finite
ISBN 90-286-0008-6
Element Methods and Other Problems. 1979
5. D.G.B. Edelen: Isovector Methods for Equations of Balance. With Programs for
Computer Assistance in Operator Calculations and an Exposition of Practical Topics of
the Exterior Calculus. 1980
ISBN 90-286-0420-0
6. R.S. Anderssen, F.R. de Hoog and M.A. Lukas (eds.): The Application and Numerical
ISBN 90-286-0450-2
Solution of Integral Equations. 1980
7. R.Z. Has'minskiI: Stochastic Stability of Differential Equations. 1980
ISBN 90-286-0100-7
8. A.1. Vol'pert and S.1. Hudjaev: Analysis in Classes of Discontinuous Functions and
Equations of Mathematical Physics. 1985
ISBN 90-247-3109-7
9. A. Georgescu: Hydrodynamic Stability Theory. 1985
ISBN 90-247-3120-8
10. W. Noll: Finite-dimensional Spaces. Algebra, Geometry and Analysis. Volume I. 1987
ISBN Hb 90-247-3581-5; Pb 90-247-3582-3

MECHANICS: COMPUTATIONAL MECHANICS


Editors: M. Stem and G.1E. Oravas
1. T.A. Cruse: Boundary Element Analysis in Computational Fracture Mechanics. 1988
ISBN 90-247-3614-5
MECHANICS: GENESIS AND METHOD
Editor: G.1E. Oravas
l. P.-M.-M. Duhem: The Evolution of Mechanics. 1980

ISBN 90-286-0688-2

MECHANICS OF CONTINUA
Editors: W.O. Williams and G.1E. Oravas
1. C.-C. Wang and C. Truesdell: Introduction to Rational Elasticity. 1973
ISBN 90-01-93710-1
ISBN 90-286-0515-0
2. P.J. Chen: Selected Topics in Wave Propagation. 1976
ISBN 90-286-0007-8
3. P. Villaggio: Qualitative Methods in Elasticity. 1977

Mechanics
MECHANICS OF FRACTURE
Editors: G.C. Sih

1. G.C. Sib (ed.): Methods of Analysis and Solutions of Crack Problems. 1973
ISBN 90-01-79860-8
2. M.K. Kassir and G.C. Sib (eds.): Three-dimensional Crack Problems. A New Solution
of Crack Solutions in Three-dimensional Elasticity. 1975
ISBN 90-286-0414-6
3. G.C. Sib (ed.): Plates and Shells with Cracks. 1977
ISBN 90-286-0146-5
4. G.C. Sib (ed.): Elastodynamic Crack Problems. 1977
ISBN 90-286-0156-2
5. G.C. Sih (ed.): Stress Analysis of Notch Problems. Stress Solutions to a Variety of
Notch Geometries used in Engineering Design. 1978
ISBN 90-286-0166-X
6. G.C. Sih 3nd E.P. Chen (eds.): Cracks in Composite Materials. A Compilation of Stress
Solutions for Composite System with Cracks. 1981
ISBN 90-247-2559-3
7. G.C. Sih (ed.): Experimental Evaluation of Stress Concentration and Intensity Factors.
Useful Methods and Solutions to Experimentalists in Fracture Mechanics. 1981
ISBN 90-247-2558-5
MECHANICS OF PLASTIC SOLIDS
Editors: J. Schroeder and G.JE. Oravas

1. A. Sawczuk (ed.): Foundations of Plasticity. 1973


ISBN 90-01-77570-5
ISBN 90-286-0233-X
2. A. Sawczuk (ed.): Problems of Plasticity. 1974
3. W. Szczephlski: Introduction to the Mechanics of Plastic Forming of Metals. 1979
ISBN 90-286-0126-0
4. D.A. Gokhfeld and O.F. Chemiavsky: Limit Analysis of Structures at Thermal Cycling.
1980
ISBN 90-286-0455-3
ISBN 90-247-2777-4
5. N. Cristescu and I. Suliciu: Viscoplasticity. 1982

Kluwer Academic Publishers - Dordrecht / Boston / London

Você também pode gostar