You are on page 1of 302

Reinhold Kienzler .

George Herrmann

Mechanics in Material Space


Springer-Verlag
Berlin ONLINE L1BRARY
Heidelberg Engineering
GmbH http://www.springer.de/engine/
Reinhold Kienzler · George Herrmann

Mechanics
in Material Space
with Applications to Defect
and Fracture Mechanics

, Springer
Prof. Dr. Ing. Reinhold Kienzler
Universitlit Bremen
Postfach 330440
D-28334 Bremen
Germany

Prof. em. Dr. se. teehn. George Herrmann


Stanford University
Ortstrasse 7
CH-7270 Davos Platz
Switzerland

Ubrary of Congress Cataloging-in-Publication Data applied for

Kienzler, Reinhold
Mechanics in material space: with applications to defect and fracture mechanics / Reinhold Kienzler ;
George Herrmann. - Berlin; Heidelberg ; New York; Barcelona ; Hong Kong ; London ; Milan ; Paris ;

ISBN 978-3-642-63121-4 ISBN 978-3-642-57010-0 (eBook)


DOI 10.1007/978-3-642-57010-0
This work is subject to copyright. AII rights are reserved, whether the whole or part of the material is
concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitations, broadca-
sting, reproduction on microfilm or in any other way, and storage in data banks. Duplication of this
publication or parts thereof is permitted only under the provisions of the German Copyright Law of
September 9, 1965, in its current version, and permission for use must always be obtained from Sprin-
ger-VerIag. Violations are liable for prosecution under the German Copyright Law.

© Springer-VerIag Berlin Heidelberg 2000


Softcover reprint ofthe hardcover 18t edition 2000

The use of general descriptive names, registered names, trademarks, etc. in this publication does not
imply, even in the absence of a specific statement, that such names are exempt from the relevant pro-
tective laws and reguIations and therefore free general use.

Typesetting: cameraready by author


Cover design: MEDIO GmbH, Berlin
Printed on acid free paper SPIN: 10740995 6213020 543210
Preface

The title of the present text is somewhat unfamiliar and needs to be elucidated first,
before describing briefly the aims and scope of this work.

Almost half-a-century ago, the late great 1. D. Eshelby advanced the notion of a
force on a defect or a singularity in a stressed solid. This notion is radically
different from the usual Galilean or Newtonian concept of a force, which engineers
of various disciplines encounter daily in their work, concerned perhaps with the
determination of flight trajectories of launched satellites, or with stress, deformation
and structural stability analysis of high-rise buildings. The Eshelby-type force is
always to be understood as a relative change of the total energy of a given system
with respect to some quantity which alters the configuration of that system. The
latter quantity might be the displacement of a foreign or missing atom in a lattice,
the change in location of a dislocation, the change in size or shape of a crack,
cavity or inclusion or the change in location of a phase boundary in a material. All
such changes of configuration of certain objects occur within the material in which
they find themselves, by contrast to changes in the configuation of a bridge under
some Newtonian loadings, which occur in what might be called the physical space
of our surroundings, in which the bridge finds itself and in which Newton's laws are
valid.

Thus the term space, whether physical or material, is used here not in a strictly
mathematical sense, as possessing a certain metric and possibly other properties, but
in essentially a descriptive meaning.

The realization that it is helpful, in fact most desirable, to distinguish between


configurational, Eshelbian forces on one hand, and the common, usual Newtonian
forces on the other, may lead naturally to the terminology of material forces on one
hand, and physical forces on the other. And, proceeding along this line of thought,
it is intriguing to investigate the possibility of constructing the edifice of Mechanics
in Material Space in parallel (or in analogy) with the well-established classical
mechanics, which now, for reasons of symmetry and aesthetics, shall be referred to
as Mechanics in Physical Space.

Indeed, the quantities and problems commonly examined in Mechanics in Physical


Space find, to a large extent, their counterpart in Mechanics in Material Space as
detailed in the Introduction. As an example, the physical Cauchy stress tensor
corresponds to the so-called material momentum tensor, or the Eshelby tensor,
VI Preface

related to configurational forces. But there are aspects for which the parallelism is
incomplete.

For example, the quantities needed in Mechanics in Material Space can be


determined only after some related problem has been solved in Mechanics in
Physical Space. This aspect might lead to the belief that Mechanics in Material
Space is entirely superfluous. We obviously do not share this point of view, just as
we do not think, even though it is possible, to conceive of classical mechanics
without the introduction of the concept of force, as has been tried in the past. It
might be added that most recently attempts have been made to establish the
framework of Mechanics in Material Space completely independently of Mechanics
in Physical Space, as will be mentioned in the Introduction.

The objective of this book is to present, in as complete a manner as possible, a


Mechanics in Material Space (or Configurational Mechanics) and exhibit its
usefulness in Defect and Fracture Mechanics.

Even though the various topics discussed are essentially mathematical in nature, the
presentation is kept at a level which should be accessible to any engineer, physicist,
material scientist or interested student with only some knowledge of calculus and
elements of partial differential equations. The mathematical foundations of the
conservation and balance laws discussed in the text rest on concepts of Lie group
theory, symmetry and invariance, but a knowledge of these topics is not required in
order to derive a benefit from our presentation. Even familiarity with the calculus of
variations is not a prerequisite. In addition, all topics, with one exception in Chapter
7, are presented in a linear formulation. This restriction makes the text absorbable
by engineers, at the obvious expense of generality, which we sacrifice in favor of
readability.

The contents of the book may be briefly summarized as follows:


The Introduction attempts to circumscribe the essential substance of the book and
the spirit in which this substance is to be presented, drawing specific parallels
between the Mechanics in Physical Space and the Mechanics in Material Space. As
already mentioned, conservation and balance laws on both sides of this parallelism
form the basis for developments in subsequent chapters, and this strictly
mathematical topic is presented in Chapter I without explicit recourse to group
theory.

The elements of the linear theory of elastictiy are recalled in Chapter 2, and
conservation and balance laws in differential and integral form are established.
Chapter 3 is devoted to the discussion of the properties of the Eshelby tensor,
whereas the notions of energy-release rates, crack-extension forces and stress-
intensity factors, which form the essence of linear elastic fracture mechanics, are
introduced in Chapter 4.

Elastostatics in material space is extended to inhomogeneous elastostatics in Chapter


5 and to elastodynamics in Chapter 6.
Preface VII

The treatment of dissipative systems of various kinds is taken up in Chapter 7,


while Chapter 8 deals with interacting fields, where the elastic field is coupled in
one instance to the electric field (piezoelectricity, dielectrics) and in another
instance to the thermal field (thermoelasticity). The latter, based on an analogy, may
also be interpreted as a fluid- saturated porous medium.

It is a remarkable circumstance that some elements of fracture mechanics, such as


energy-release rates and stress-intensity factors, might be developed noton the basis
of continuum theories, but on the basis of the much older and simpler theories of
strength-of-materials. Bars with cracks in tension-compression, shafts in torsion and
beams and cylinders (pipes) in bending are dealt with in Chapter 9, while plates and
shells are considered in the final Chapter 10.

The principal parts of this book were written during the authors' several stays of
several weeks each at the Mathematisches Forschungsinstitut Oberwolfach,
Germany, under a program called "Research in Pairs" invited by the Director of the
Institute, Professor Dr. Matthias Kreck. The authors are greatly indebted to Prof.
Kreck and to the Institute for having provided such generous hospitality.

The authors are also grateful for the partial support provided both by the U. S. Air
Force Office of Scientific Research through a grant to Stanford University and by
the University of Bremen.

The book could have never taken the form it has without the benefit of discussions
and collaboration with numerous former graduate students from Stanford University,
in particular Dr. N. Chien, Dr. 1. Eischen, Dr. Y. Pak and Dr. H. Sosa. Especially
Dr. T. Honein contributed many novel ideas which have found their entry into the
book.

Greatly appreciated were also the fruitful comments offered by many colleagues, in
particular Prof. D. Barnett (Stanford University), Prof. H. Gao (Stanford University),
Prof. D. Gross (Technical University of Darmstadt) and Prof. J. Willis (University
of Cambridge).

The preparation of the manuscript was skillfully and faithfully carried out by
B. Neumeister-Schulze and the responsibility for drawing the Figures rested with
M. Grabendorff and S. Stiihrmann. Their accurate work is gratefully acknowledged.

Finally, we would like to express our thanks to the publisher, and in particular to
Dr. Merkle, for excellent cooperation.
Contents

Preface v

Introduction 1

1 Mathematical Preliminaries 13

1.1 General Remarks 13


1.2 What is a Conservation Law? 14
1.3 Trivial Conservation Laws 17
1.4 System with a Lagrangian; Noether's Method 19
1.5 System without a Lagrangian; Neutral-Action Method 42
1.6 Discussion 48

2 Linear Theory of Elasticity 51

2.1 General Remarks 51


2.2 Elements of Linear Elasticity 51
2.3 Conservation Laws of Linear Elastostatics 62
2.4 Alternative Derivations of Conservation Laws 72

3 Properties of the Eshelby Tensor 81

3.1 General Remarks 81


3.2 Physical Interpretation of the Components of the Eshelby Tensor 82
3.3 Invariants, Principal Values, Principal Directions and
Extremal Values of the Eshelby Tensor 86

4 Linear Elasticity with Defects 95

4.1 General Remarks 95


4.2 Path-Independent Integrals and Energy-Release Rates 96
4.3 Example: Hole-Dislocation Interaction 100
4.4 Path-Independent Integrals of Fracture Mechanics 107
x Contents

5 Inhomogeneous Elastostatics 121

5.1 General Remarks 121


5.2 Symmetry Transfonnations 122
5.3 The Homogeneous Case 124
5.4 The Inhomogeneous Case 124
5.5 Relation to Stress-Intensity Factors 127
5.6 Examples 128

6 Elastodynamics 133

6.1 General Remarks 133


6.2 Time t as an Additional Independent Variable 134
6.3 Convolution in Time 143
6.4 Domain-Independent Integrals 146
6.5 Energy-Release Rates 151
6.6 Wave Motion 156

7 Dissipative Systems 161

7.1 General Remarks 161


7.2 Diffusion Equation 161
7.3 Non-Linear Wave Equation 163
7.4 Viscoelasticity 165

8 Coupled Fields 173

8.1 General Remarks 173


8.2 Piezoelectricity 174
8.3 Thennoelasticity 179
8.4 Mechanics of a Porous Medium 192

9 Bars, Shafts and Beams 195

9.1 General Remarks 195


9.2 Elements of Strength-of-Materials 196
9.3 Balance and Conservation Laws for Bars and Shafts 200
9.4 Balance and Conservation Laws for Beams 203
9.5 Energy-Release Rates and Stress-Intensity Factors 205
9.6 Examples 211
Contents XI

10 Plates and Shells 219

IO.I General Remarks 2 I9


10.2 Plate Theories 220
10.3 Conservation Laws for Elastostatics of Mindlin Plates 224
10.4 Reduction to the Classical Theory 229
10.5 Conservation Laws for Shells 233

Appendix A 239

Conservation Laws for Inhomogeneous Bars under


Arbitrary Axial Loading 239

Appendix B 245

B. I Elastodynamics of Inhomogeneous Bernoulli-Euler Beams 245


B.2 Reduction to Statics 252

Appendix C 259

C.I Elastodynamics of Inhomogeneous Mindlin Plates 259


C.2 Reduction to Statics 269

References 273

Symbol Index 281

Author Index 287

Subject Index 291


Introduction

Classical mechanics, as we know it, is the body of knowledge which is concerned


with equilibrium and motion of objects which possess mass and which are placed in
the usual Euclidian physical space. Depending on the circumstances and goals of the
mathematical modeling of real, observable phenomena, these massive bodies may be
idealized as mass particles, rigid extended bodies or deformable materials obeying
a variety of constitutive relations. The physical space in which the motion, or rest
(equilibrium) as a special case, takes place is usually considered to be isotropic,
homogeneous, not changing in time and admitting inertial reference frames which
permit the application of Newton's equations of motion. The ultimate goal of
mechanics consists in, on one hand, "explaining" observed phenomena and/or, more
importantly, predicting future behavior of mechanical systems based on current,
experimental information.

This body of knowledge, as regards its basic notions and concepts, has been highly
developed over the centuries, thanks to the ingenious advances of Galileo, Newton,
Euler, Lagrange, Hamilton and others. An account of the features of different
formulations of mechanics by these "fathers", builders and developers of the house
of mechanics, in a modem mathematical setting, has been given by V. I. Arnold
(1989). Today's activities concentrate not so much on physical fundamentals per se,
as on an increasingly realistic description of constitutive behavior of materials, on
more refined experimental techniques and on advancing analytical and
computational methods which would improve the predictive capabilities of
mechanics.

Progress along this line of aCtIVity has naturally led to the recogmtlOn that
materials, on some scale, cannot be regarded as perfect continua, but rather contain
a variety of numerous defects, such as microcracks, inclusions, vacancies, voids,
dislocations, etc. One feature of such defects is that they can move within the body
as a result of load application or thermal influences, through several mechanisms,
such as mass transport through diffusion, void or crack nucleation and growth,
grain boundary sliding, dislocation motion, etc.

Thus the necessity arises to be concerned with objects (defects) which may have
mass (e. g., inclusions) or may have no mass (e. g., voids) and which may move
within the material in which they find themselves (whether or not the material as a
whole moves with respect to the physical space in which it resides).
2 Introduction

To characterize these objects one is led to the introduction of the concept of a


"force" acting on a "defect" in the material, by analogy to the concept of the force
acting, (through contact or at a distance), on a material body in physical space. This
initial first step can lead to the construction of a whole edifice of knowledge, in
some sense dual to the usual Mechanics in Physical Space (MiPS), which may be
called, for want of a better name, Mechanics in Material Space, or Mechanics in
Materials. However, the term "Mechanics in Materials" is too close to the term
"Mechanics of Materials" which traditionally designates the description of
mechanical behavior of materials and possibly also theories of strength-of-materials.
For this reason it appears preferable to employ the term Mechanics in Material
Space (MiMS), even though the word "Space" is understood here not in a strictly
mathematical sense (possessing, e. g., a well-defined metric and possibly other
properties) but rather a location. But this space can be, and frequently is, both
anisotropic and nonhomogeneous. As it turns out, a far-reaching duality exists
between the usual Newtonian forces in physical space as they act on masses, on the
one hand, and the material (non-Newtonian) forces in material space, as they act on
objects (defects), which mayor may not have mass, on the other hand. As outlined
briefly below, a significant amount of effort has been spent during the past decades
on furthering our knowledge of the mechanics of defects, but this knowledge lays
dispersed in numerous diverse journals and conference proceedings, and in fact the
term "Mechanics in Material Space" has been introduced into the literature only
recently by one of the authors of the present monograph.

The principal purpose of this text, then, is to pull together some of the current
knowledge, to present in a coherent manner the main precepts of MiMS, to
emphasize the duality to MiPS and to outline some of the technologically important
applications of this field.

The beginnings of Mechanics in Material Space, even though he never used the
term, go back to 1951, when J. D. Eshelby (1951) in a classic paper, introduced the
notion of a force on an elastic singularity (or defect). He defined it as the negative
gradient of the total energy of the body with respect to the position of the defect in
the material (not in physical space). This force, which later was sometimes called a
material force, a driving force, a configurational force, a quasi-force, a
thermodynamic force (or affinity), a non-Newtonian force, is quite different from
the usual Newtonian force, exerted through contact (surface forces or tractions) or
at a distance (body forces) by bodies other than the one under consideration. This
difference gives rise to a far reaching duality between the two sets of forces, as will
be outlined below.

An example of a material force is the crack-extension force of fracture mechanics,


characterizing stress-intensity factors at a crack tip in materials. Another example is
the material force (actually material traction) acting on an element of a phase
boundary or on the surface of a stressed solid, even though that surface may be free
of physical surface tractions (contact forces).

The material force acting at a crack tip is frequently evaluated by means of the so-
called J integral. It is a path-independent integral, i. e., its value is independent of
the curve (in plane problems) surrounding the crack tip, provided the region
Introduction 3

enclosed by the curve is elastically homogeneous and does not contain any other
defects. The value of the J integral is equal to the negative gradient of the total
energy with respect to the position of the crack tip, called the energy-release rate,
and is related to the stress-intensity factor at the crack tip. The J integral has been
successfully used as a crack parameter in fracture mechanics of brittle (or "small-
scale yielding") solids. Its path independence may be viewed as an aspect of general
conservation laws, which play a central role in mechanics and physics.

It is a remarkable fact, that this important quantity has been introduced into the
literature in a somewhat ad hoc fashion. It first appeared in a work by Eshelby on
forces on defects in elastic solids already mentioned (Eshelby, 1951) in a three-
dimensional setting, but not related to fracture mechanics. Then it was introduced
again by Sanders in a brief note (Sanders, 1960), but did not take off until Rice
published his well-known paper (Rice, 1968). Related work by Cherepanov (1967)
should also be mentioned. Prior to this, a paper by Gunther appeared in an obscure
journal (Gunther, 1962) in which he showed how to derive conservation laws in
elastostatics, (of which the J integral is a particular case), based on Noether's first
theorem (Noether, 1918), but did not apply them to defect mechanics, not being
aware of Eshelby's work. Gunther's conservation laws, little known at the time, were
rediscovered independently by Knowles & Sternberg (1972), who showed that the
J integral is one of the quantities which came to be known as the j, Land M
integrals (Budiansky & Rice, 1973).

It turns out that a systematic derivation of conservation laws for a mechanical


system possessing a Lagrangian function, i. e., a potential, can be based on
Noether's first theorem (Noether, 1918) on invariant variational principles, which
in tum is based on Lie's theory of continuous groups (Lie, 1912). Even though
group theory underlies Noether's theorem, this branch of mathematics will not be
dealt with explicitly in the present work.

Another important quantity which has been introduced by Eshelby into mechanics,
and more specifically into elasticity, is what he called first (Eshelby, 1951) the
Maxwell tensor of elasticity, by analogy with electrostatics, and later the energy-
momentum tensor. In the present pages, we shall refer to this quantity as the
material momentum, or, more appropriately, the Eshelby tensor as has been done by
various contemporary authors, e. g., Maugin (1993). This author (Maugin, 1995)
introduces even the concept of Eshelbian mechanics whose main ingredient is the
Eshelby tensor. As we shall see, this tensor enters, in particular, the J integral and
plays in material space a role analogous to the Cauchy stress tensor in ordinary
elasticity (in physical space). Since in a typical problem of Mechanics in Material
Space, the deformation is accompanied by change of configuration of the body (e.
g., crack advance or inclusion displacement), one could introduce the term
Configurational Mechanics which would be an alternate term to MiMS.

Under numerous circumstances, the mechanical system under consideration cannot


be modeled omitting energy dissipation and, as a consequence, a Lagrangian
function cannot be constructed in a straightforward manner. This means that
equations of motion can be written down, but they are not the Euler-Lagrange
equations of a variational problem. Thus, application of Noether's theorem to
4 Introduction

establish conservation laws is no longer possible and other procedures need to be


developed.

Recently, Honein et al. (1991) succeeded in fonnulating an alternate procedure


which has been tenned the "Neutral Action" method, as will be discussed in
Chapter 1 and applied to several problems of continuum mechanics in the following
chapters.

To emphasize the parallelism (or correspondence) between the (usual) mechanics in


physical space and the (not so common) mechanics in material space, we have
prepared a Table which stresses this duality by juxtaposition and which is in
essence self-explanatory.

In both physical space and material space the concepts of force are introduced as
negative gradients of the total energy. Further quantities and concepts which can be
introduced both in physical and material spaces concern tractions, momentum (or
stress), application of conservation laws, both in local and integral form, as well as
with and without source terms. Free-body diagrams can be sketched and such
typical notions of mechanics as stability of equilibrium and trajectories of moving
objects can be discussed both in physical as well as in material space. It may be
mentioned that the parallelism between material and physical space even leads in
one instance to some new findings in the mechanics in physical space, as will be
discussed in Chapter 6.

It should be pointed out that the material presented in this Table i!' of necessity
sketchy and merely serves the purpose of making the reader better acquainted with
the parallelism (and also some differences) between the mechanics in physical and
material space, i. e., the general goal of this text.

The principal difference is surely this: In order to evaluate any of the quantities
defmed in material space (because they are related to the total energy of the
mechanical system under consideration) it is always necessary to solve a boundary-
value problem (of the usual kind) first in physical space. Stated differently, it is
never possible to move in the other direction, i. e., by starting with some quantities
in material space, evaluate conversely those in physical space. For this reason some
authors prefer not to use the concept of material space at all, but simply discuss
energy changes (or energy-release rates) as they occur when some quantity
charaterizing the material (such as a phase boundary, void size or crack length) has
changed. Yet, it seemed to us that despite this absolute lack of necessity, the edifice
of mechanics in material space possesses a sufficiently strong foundation and a few
solid comer stones, which makes the whole enterprise of its construction a
worthwhile endeavor.

In this connection it should be mentioned that Gurtin (1995, 1996) and Gurtin &
Podiu-Guidugli (1998) have made an attempt to introduce the Eshelby tensor and
configurational (material) forces directly, without reference to any physical stress or
deformation. This was accomplished by postulating a general configurational force
balance, taking an appropriate form of the second law of thennodynamics and
requiring invariance under reparametrization, the latter being equivalent to
Introduction 5

invariance of a certain expression for rate of work. These configurational forces,


whose importance is not in question, are thus viewed as basic primitive objects.

A final remark concerns the fact that some controversy exists regarding whether
forces arising in certain systems should be viewed as Newtonian (physical) or
configurational (material). Specifically, investigating the force on a disclination in
a nematic liquid crystal, Eshelby (1980) claimed that it has the character of a
Newtonian force, a view supported by Ericksen (1995). The opposite stance is taken
by Kroner (1993), who makes a strong case for his position that it is a material,
configurational force.
6 Introduction

TABLE

Mechanics

in physical space in material space


Space is usually assumed to be Space may be Riemannian (shells),
Euclidian, homogeneous, isotropic not homogeneous, anisotropic and
and not changing in time possibly changing in time (aging
materials, creep)

Forces
(on discrete objects)

Example ample
, .1

r
m A I
• •
T
x x
p

Particle of mass m in gravity field Elastic plate of given dimensions and


(acceleration g) above a reference material properties subjected to
plane P at distance x tension T. Contains a "defect" A
(inclusion or cavity) at distance x
from the lower edge

Total (potential) energy II Total energy II depends on several


quantities a j , one of them being x

II = mgx II = II (a j • x)

Force F acting on mass m Force J acting on defect A


dII J = _ dII
F=--=-mg
dx dx
F is called "physical" or "Newtonian" J is called "material" force. Also
force called: quasi-force, driving force,
configurational force, thermodynamic
force, non-Newtonian force, force on
a defect, inhomogeneity force
Introduction 7

Force , Traction
(in continua)
G
n n

XJ

BodyB BodyB

Physical traction! Material traction 9


Physical force acting across a unit Material force "acting" across a unit
area with normal ni area with normal nj , i. e., the negative
change of total energy if the unit area
is displaced a unit in xi -direction

Stress (Tij (or "physical momentum") Material traction Vi introduced via


introduced through the Cauchy material stress b;j (or "material
relation via physical traction !i (cf. momentum"), i. e., the Eshelby tensor
Section 2.2) (cf. Section 3.3)

Gn J· = bI).. n.J

Total Energy Density H (Hamiltonian)

Cauchy stress (Tij ("physical Material stress (or "material


momentum") introduced as derivative momentum", i. e., Eshelby tensor),
of Hamiltonian associated with introduced as total derivative of
translation in physical space (cf. Hamiltonian associated with
equation 1.68) translation in material space (cf.
equation 1.69)
oH oH
bij == H8ij - uk,' - -
~ .. J ~k,i
J,I
8 Introduction

Balance of Linear Momentum in Integral Fonn


with Source Tenns
(by divergence theorem)

S,A S,A

f b..dV = fb.. dA
)1.) )'
n.)
B

~ I
S

<li dA ~J,
('~fidV)
B

q; is physical force density due to the j; is material force density due to


presence of continuous body forces continuous inhomogeneities in
material space

Conservation of Linear Momentum in Integral Fonn


without Source Tenns

F., =0 J., =0
i j =0
(physical homogeneity) (material homogeneity)
Introduction 9

Balance of Linear Momentum in Local Form


with Source Terms

Body forces qj disturb the divergence- Material inhomogeneities jj disturb ~he


free character of the Cauchy tensor divergence-free character of the
Eshelby tensor

Conservation of Linear Momentum in Local Form


without Source Terms
a ...
jI,)
= 0 b...
}',}
= 0

Cauchy tensor is divergence-free Eshelby tensor is divergence-free


in the absence of body forces in the absence of inhomogeneities

Conservation of Angular Momentum


Cauchy stress is symmetric Eshelby tensor is not symmetric

a..I) = a)1..

(as a consequence of the balance (because rotation couples physical and


of angular momentum) material space)

Free-Body Diagram

AI A'
2
2

Beam under distributed loading Crack in a remote homogeneous but


arbitrarY stress field
10 Introduction

Stability of Equilibrium
of Discrete Objects

}---llo-f-- X,
c

Mass particle m on a saddle-shaped Hole A in an elastic material in a


surface in a field of gravity g (acting plane stress field produced by
in negative X3 -direction) dislocations B and C

Total energy II = mgxixl , x2); Material forces acting on A


at Xl = X2 = 0 it is
all all
F =0 J =- - =0
l=O
- -
x
J
' J OxJ '

all all
F
Ox2 =0 J = - -
Ox2 =0
= - -
2 2

because of equilibrium because of symmetry

Stability of equilibrium Calculation shows that

2
unstable unstable a ll < 0
1 '
Ox l

stable stable alll> o.


Ox;

Equilibrium at a saddle point Equilibrium at a saddle point


Introduction 11

Trajectorie of motion

Plane trajectory of a mass particle in Plane trajectory of an inclusion A near


field of gravity g with assigned initial a cavity of radius a in inhomogeneous
velocity Vo and determined on the stress field produced by remote
basis of Newton's equations of physical tractions. In the absence of a
motion. (Physical force F not "true" equation of motion, material
tangential to trajectory, but rather in force J is assumed to be tangential to
the direction of the acceleration) the trajectory
1

Mathematical Preliminaries

1.1 General Remarks

The first Chapter of this text is an essentially mathematical exposition concerning


establishment of conservation laws, as they will be employed in all subsequent
chapters, which deal with phenomenological models of physical systems.

In constructing conservation laws, two different methodologies may be followed. If


a Lagrangian function is available for the system of interest, the classical Noether's
method, possibly with the extension of Bessel-Hagen, might be applied. This
method, which is based on Lie's theory of continuous groups and invariance under
group action in jet spaces, is presented in a manner which does not require any
knowledge of these more advanced mathematical concepts.

If a Lagrangian function is not available, and the system of interest is given only by
some set of partial differential equations, the recently developed Neutral Action
(NA) method might be used to advantage. It will be derived also in quite an
elementary fashion as regards the mathematics employed. It is of interest to point
out that the NA method to construct conservation laws might be applied also to
systems possessing a Lagrangian. In that case, it leads to the same results which
would be obtained on the basis of Noether's method, together with the Bessel-Hagen
extension.

The essential difference between the two methods lies in the following: In Noether's
method, the so-called characteristic of a conservation law is determined by the
chosen transformations, while in the NA method this quantity is to be found from
certain invariance requirements.

In several theories of mechanical systems, such as elasticity or beam theory, a third,


simple method of establishing conservation laws might be used, which actually is
Noether's method in disguise. The governing strain-energy density Wand a vector
Xi W, might be subjected to the differential operations of grad, div and curl with
respect to material coordinates xj which correspond to the transformations of
translation, self-similar expansion and rotation, respectively.
14 I Mathematical Preliminaries

1.2 What is a Conservation Law?

Consider a system involving m independent variables xj (i = 1, 2, ..., m) and Jl


dependent variables va (a = 1, 2, ..., Jl) and let it be governed by a set of q
differential equations represented by

(1.1)

where ~ is a set of differential operators acting on xj , va and derivatives of va


with respect to Xj •

If a set of m associated functions Pj (xj , v" ' V"j' ... ), (j = 1, 2, ..., mY, satisfies the
equation

dP2 dP
+ ... +-.!!!.. = dP = P .. = 0
i
+- (i = 1, 2, ..., m) (1.2)
dx2 dx m dx 1,1
i

for all solutions of (1.1), then such a differential form is called a conservation law
(or a divergence expression).

Here dldx j is the total differential operator, for which the comma symbol is also
used.

For example, if

then

By contrast, aJaxj is here the symbol for partial differentiation with respect to the
variable indicated, implying that in this operation all other forms of variables,
whether independent, dependent or partial derivatives of dependent with respect to
the independent ones, remain fixed.

Taking the function P2 indicated above,


1.2 What is a Conservation Law? 15

dP2 = OP2 OV4


but
dx3 OV4 &3

In equation (1.2), the summation convention over repeated indices has been
introduced. The range of summation of latin indices is 1 through m, that of greek
indices is 1 through Jl.

The general conservation law (1.2) may also be written in symbolic notation

div P = 0, (1.3)

where P is the m-component vector Pj (j = 1,2, ..., m).

If m = 1, we are dealing with a set of Jl ordinary differential equations involving


a single independent variable x and the conservation law is

dP = 0 or P = const. (1.4)
dx

and provides a first integral. The number Jl of dependent variables is not


restricted. Occasionally, instead of d/dx we use the abbreviation

de) =(y.
dx
If we are concerned with a dynamical system in three-dimensional space referred to
Cartesian coordinates Xj (i = 1,2,3) and time t, such that m = 4, the conservation
law (1.3) reads

(1.5)

It indicates that the time rate of change of a density P, is balanced by the


divergence of the associated fluxes (or currents) Pj , P2 and P3 •

For example, if P, is the mass density p and there are no mass fluxes across the
unit volume occupied by the mass p , then dp/dt = 0 and p = const. If P, is the
energy density of a unit volume (or a unit mass), then the conservation law (1.5)
states that the time rate of change of the energy density dP, /dt is balanced by the
divergence of the fluxes of energy out of the unit volume dP j /dx j + dP 2 /dx2 +
dP3 /dx3 , i. e., energy is conserved, which is a statement of the first law of
thermodynamics. It is, however, not always possible to provide an immediate
physical interpretation for all conservation laws derived in the following chapters.

A conservation law in differential form (1.5) may also be written in integral form.
We first consider the divergence theorem in three-dimensional space
16 I Mathematical Preliminaries

{dJ; ;;
j dx j
dV = j/; n j dA (i = 1, 2,3) , (1.6)

B S

where J: are "well-behaved" functions defined in a body B with the infinitesimal


volume element dV = dxl dx2 dx3 , which is enclosed by a surface S with area
element dA and unit outward normal vector n (see Figure 1.1).

II

Fig. 1.1: Body B with volume V, surrounding surface S with area A and unit outward
normal vector n

If only the components PI' P2 and P3 are transformed in (1.5), the conservation law
reads in integral form

(1.7)

Relation (1.5) is now expressed with respect to a finite volume V, rather than a
infinitesimal volume dV, i. e., conservation laws may be formulated either in the
local (strong) or global (weak) form. In two spatial dimensions the divergence
theorem is the integral over a plane surface S in terms of an integral along a
(closed) line (or path) , r (arc length s) , which encloses the area (see Figure 1.2),
i. e.,

(i=I.2). (1.8)

Again, it is presumed that the vector point function J: is sufficiently well-behaved


(no singularities, jumps, etc.)
1.3 Trivial Conservation Laws 17

Fig. 1.2: Plane surface S with area A, surrounding curve r with arc length s and unit outward
normal vector n

1.3 Trivial Conservation Laws

As pointed out by Olver (1993), there exist two kinds of "trivial" conservation laws.
In the first kind, each component of Pj vanishes identically upon the solution of the
system. In the second possible kind of triviality, the conservation law holds for all
functions va' regardless of whether the governing differential equations (1.1) are
satisfied or not. As an example of this first kind of triviality consider the 2-
component vector P with components

p] = (cJv]l - (cJv2l ,
Ox] Ox2

in two dimensions (m = 2) with two dependent variables v j = v j (x j , x 2) and


v2 = v2 (Xl> x2) (Jl = 2). The conservation law

is trivial, if one of the governing field equations is assumed to be given by

On inspection, P j and P2 vanish for all solutions of the field equation given above.
18 1 Mathematical Preliminaries

As an example of the second kind of triviality it may be recalled that the curl of a
vector point function f is known to be trivially divergence-free, i. e.,

div curl f = O.

In this circumstance one speaks of a null divergence and the associated conservation
law does not depend on any particular set of differential equations, which, in the
above example, may govern the vector field.

A specific case of the above identity in 2 dimensions (m = 2) and a single


dependent variable v = v (xJ, x2) ()4 = 1) is obtained if we set

0\;
P =--
2 Ox
J

Then

dPJ oPl o2V


-
dxl Oxl Ox2
o(~)
Ox2

dP2 OP2 o2V


-
dx2 o (~) OxJ 0x2
Oxl

dPl dP2
and + - =0,
dxl dx2

regardless of the differential equations which might govern v = v (xJ, x 2), such as,
for instance,

02V + o2V = 0 (Laplace's, or harmonic equation)


Ox12 Ox}

or (biharmonic equation).

The usefulness of trivial conservation laws is restricted, whereas that of non-trivial


conservation laws is quite diverse and they have important applications, both
mathematical and physical, as ·summarized, for example, by Olver (1993). They are
employed in the discussion of global existence theorems, stability of solutions,
conditions for shock waves, scattering theory, uniqueness theorems, qualitative
information, development of algorithms in computational mechanics, phase
1.4 Systems with a Lagrangian; Noether's Method 19

transfonnations and others. Our concern here is only the application of conseIVation
laws to defect and fracture mechanics within the context of elasticity, strength-of-
materials, and related coupled fields, such as thermoelasticity and piezoelectricity.

1.4 Systems with a Lagrangian; Noether's Method

The methodology for establishing conseIVation laws, as mentioned, is different


depending upon whether the system considered is Lagrangian or not.

The Lagrangian function

A Lagrangian system possesses a Lagrangian function L which depends, in general,


on the independent variables Xi (i = I, 2, ... , mY, the dependent variables va (a = I,
2, ..., r1J), the first derivatives va,l. = O\Ja / dtI and possibly higher derivatives

(1.9)

In this work we shall not consider the dependence of L on second and higher
derivatives, except in the special case m = 1, It == 1, where second derivatives
will be included.

Such a Lagrangian function may be looked upon as a potential and, for example
in elastostatics (cf. Chapter 2), may be identified as the negative of the strain (or
internal) energy density W (per unit of volume) of the system

L == - W (u i ) (1.10)

for a body which is homogeneous and

L == - W (Xi' Ui ,;) , (1.11)

if the properties of the body are not constant, but rather functions of the
independent variables (coordinates) Xi' In (1.10) and (1.11), uij is the
displacement gradient.

In certain dynamic problems, the Lagrangian L might be the so-called kinetic


potential

L==T-W, (1.12)
20 I Mathematical Preliminaries

where T is the kinetic energy.

Systems with no dissipation usually possess a Lagrangian and, in a quasi-static


process, the work of external forces (body forces and surface tractions) is balanced
by changes in the internal energy.

The Euler-Lagrange equation

If a Lagrangian for a general system is postulated, the equations of equilibrium (or


motion) are already predetermined and they are derived as follows:

One considers the action integral A which is the integral of L over an arbitrary
n
domain in the space of independent variables Xi

(1.13)

Next one applies Hamilton's principle by performing "the variation" 8A of A, i. e.,


by considering different values of the functional A for different sets of functions
v",(x), with the subsidiary conditions that the domain n remains fIxed during
variations. Thus, among all possible functions V",(X) , each of which supplies a
number A, the solution function is distinguished by the feature, that the associated
A is stationary, i. e., ~A = O. To perform the variation 8A of A, the rules of the
calculus of variations are to be applied, originally developed by Euler (1744), and
presented in numerous texts, e. g., Gelfand & Fomin (1963). One needs to know
these rules to perform the variation, just as one needs to know the rules of
differential calculus to perform differentiation. To follow the subsequent sections
of the present text, however, no knowledge of the calculus of variations is
required.

Let us consider fIrst the special case m = 1, Jl = 1, then

L = L (x , v, v..J (1.14)

and
b

A / L (X,' v, vox) dx, (1.15)

where a and b designate some arbitrary points on the x-axis. Without knowledge of
the specifIc form of L, i. e., its specifIc dependence on x, v, v,xl we can calculate the
variation ~A of A by varying under the integral sign v and v,x' according to
1.4 Systems with a Lagrangian; Noether's Method 21

v -t v I- ~v ,
(1.16)
V,x -t V,x I- 8v,x

but doing nothing to x. The function 8v is the so-called variatIOn of v and is


arbitrary, except that it vanishes at the end points a and b of the interval under
consideration (admissible functions), and that it is small in comparison to v (see
Figure 1.3). Analogous features are valid for 8v,x.

-+- -----.! --'- x


a b

Fig. 1.3: Function v and its variation bv

The variation of the action integral results in


b b

v I- 8v, V
.f
+
I
8v.,/ dx - v, v.x) dx. (1.17)

a n

The integrand L at x, v I- 8v, Vo< I- 8v,x will now be expanded in a Taylor series
about the value at x, v, v.• , just as the value of a sufficiently smooth function of
one variable f (x) at x + a can be expressed in terms of its value at x and its
derivatives at x, namely

f(x + a) = f (x) + a df (x) ./-


(/2

I! d.x 2!

Since 8v and 8v,x are assumed to be small, the Taylor series is truncated after the
linear tenn in 8v and 8v,x, i. e., terms involving (8v f, 8v 8v.•. (8v.xi and higher order
products, abbreviated by 0 (8 2), are neglected.

Equation (1.17), then changes to

I (L
b

elL elL
(x, v, v.. ) + .- 8v + -- 8v
0\1 0\1......
a
22 I Mathematical Preliminaries

J(~ SV + :: Sv,)dx. (1.18)

Next we wish that the second term be also a variation 8v and not the variation 8v.x .
This wish can be fulfilled by knowing that the operation of variation 8 and
differention d( )/dx = ( ),X may be interchanged and by considering the
differentiation of a product

+-
oL 8v
O\J ,x

b
,x
or

J b(OL)
,x
oL
O\J 8v,x dx = O\J 8v
,xa
/b j
=
d (oL)
dx O\J
,x
8v dx +
jb oL
O\J
,x
8v,x dx.
a a a

Since 8v = 0, by requirement, at a and b, we have

b b

/
oL 8v dx
O\J
,x
,x / ~ (~r ) 8v dx
a a

and therefore
b

8A / (~ 8v (1.19)

For oA = 0 and 8v arbitrary the consequence is

oL _!!:..- (oL )= o. (1.20)


O\J dx O\J,x

This is called the Euler-Lagrange equation of the variational problem M = 0,


which is identical to the equilibrium equation of the system. Note that d/dx is the
total differential operator

d 0 (1.21)
dx eX

The operator
1.4 Systems with a Lagrangian; Noether's Method 23

E()= -a -d- - a) () (1.22)


/ LV dx OV,x

is referred to as the Euler operator (which is always acting on the Lagrangian L).
Thus the Euler-Lagrange equation above may be written for short

E (L) = 0 . (1.23)

If L is a function also of v.xx , then

E (L) =aLov _.!!:..-


dx
(aL
ov,x
)+ ~
dx 2
(aL
ov.u:
)= 0 . (1.24)

In the most general case X j ( i = 1, 2, ... , m) and va (a = 1, 2, ..., Jl) and L


depending on Xi' Va and all derivatives of v" with respect to X j up to order J, the
Euler operator is

E = E (- D) ~,
a
J OV J a
(1.25)
J

where the multi-index notation introduced by Olver (1993) is employed. Here vJa
represents all possible p-th order partial derivatives of va

ap V
a
(1.26)

with J = Uf' i2' ..., i ,,) an unordered p-tuple of integers, 1 :::;; i p :::;; m indicating
which derivatives are being taken, and the number of J, # J = p, indicating how
many derivatives are being taken. The differential operator ( - D)J is defined by

(1.27)

where the DJ are all possible p-th order total derivatives

(1.28)

Transformation of dependent and independent variables

We return to the special case m = 1, Jl = 1, and consider again

L=L(x,v,v.x)
24 I Mathematical Preliminaries

and

a
We now do something which is usually not done within a standard course of study
in engineering science or applied and engineering mechanics, namely we subject the
Lagrangian L to an infinitesimal transformation of both the independent and
dependent variables, i. e., we pass from the usual, unstarred quantities x, v to starred
quantities x', v* according to the prescription

X -t x' = x + £ ((x, v) , (1.29 a)

v -t v ' = v +£ cP (x, v) . (1.29 b)

Here the single constant parameter £ is supposed to be small in the sense, that the
ensuing terms without £ and linear in £ will be retained, while terms with £2 and
with higher powers will be omitted. The functions (and cP, with the arguments
indicated, are completely arbitrary. The justification for carrying out this
transfonnation is for the moment obscure and will become clear only when the
result is considered. But note the difference between the transformation (1.16),
which took us from L to the associated differential equations, and the
transfonnation (1.29).

The transformed A shall be called A ' and reads


b'

A' (1.30)

We next wish to express all the staned quantities in tenns of the original,
unstarred ones. This means, in particular, that the transformed domain a ' - b ' will
be expressed in tenns of the original domain a - b and the differential dx ' can be
expressed as (dx '/dx) dx . Using the transformation prescription

x' = x + £ ((x, v) ,

it is readily seen that the derivative dx '/dx is

dx' d(
- =1 +£ (1.31)
dx dx

Thus

/L f
b

dx'
A' + £ ((x, v), v + £ cP (x, v), ? ) - dx.
dx
a
1.4 Systems with a Lagrangian; Noether's Method 25

The question mark ? indicates the dependence of L on dv dx but expressed in # / #,

terms of v, (, t/J and e, which we do not know yet. We proceed to find it as follows

dv' dv' dx dv' 1


--- ----
dx' dx dx' dx dx'
dx

dv '/dx is again readily found from the relation

v' == v + et/J (x,v)

as
dv • dv + e dt/J == v + E dt/J .
dx dx dx .x dx

We also need

1 1
dx'I+E d (
dx dx

We recall that, by binominal expansion

(1 + yr 1 ==1 - y+O(y2).

Thus, finally,

dx'
'f
-dv == V
,.x
+ E -dt/J ~t1 - E -d
dx dx
(J + 0 (E 2)

(1.32)

and A ' becomes


b

A' + E( (x, v), V + Et/J (x, v), v,.x +

+ e (dt/J _ v d ( )/ dx ' dx . (1.33)


dx ,.x dx/j d.x

The integrand L at x " v ' and dv •/dx ' will now be, again, expanded in a Taylor
series about the value at x, v, voX'

The result is first, with (1.31)


26 I Malhematical Preliminaries

j
b

A. (L (x, v, v) +

+
filL(
e -
~
+ ilL
-c/J
ov
+ - ilL
OV,x
/dc/J
-
dx
- V
,x
-
dx,
2 1 1
d(Jj+ 0 (e); f + ed(J
- dx
dx

and finally, omitting tenns in e 1 and higher,

A • -=A +e +-
aL
OV,X
a

/II ~
b

-= A + f + c/J : + (dx
a
(1.34)

As a side remark, the operator (0/ eX + c/Ja/ ov is referred to as the infinitesimal


generator w (of a Lie group in the space x, v)

(1.35)

and the operator

prell w -= w + /dc/J _ v d(J~ (1.36)


r -dx ,x dx ov
.r

is referred to as the first prolongation prll) w of the group into the jet bundle space
x. v, v,x.

These designations belong to the theory of continuous Lie groups, whose knowledge
is not essential for us and which we simply use here. But to explore this
background, the reader is referred to, e. g., Olver (1993), where for w, the letter v
is used, and vice versa.

Thus we can write


b

A • -= A + e j/pr(]) w + 1xJL dx. (1.37)

a
The first term in the integrand above describes the change of L to L' in the
domain a - b and is a differential operator, while the second tenn is a factor
which describes the change of domain a - h to a . - b '.
1.4 Systems with a Lagrangian; Noether's Method 27

In the integrand of equation (1.34), we rewrite the 3 tenns involving d/dx as


follows:

deP _ v d ( ) ilL + L d( ==
/ dx ,x dx OV.. dx

_ ( (ilL + v ilL + v ilL ) , (1.38)


ex ,x ov ,xt ov
..
where (1.21) and the product rule jg.< == (jg)., - f, g was used. It is noted that the
tenns (v.<xilL/ov.. cancel. Finally we can rewrite (1.34) with (1.38) as
b

A • == A + E /(~ leP :-
..
+ ( (L - V,x : - )
,x
J+

(1.39)

It is noted that ilL/ ov - d/dx (elL! OV.x) is the Euler-Lagrange expression E(L)
(1.20, 1.23)

(1.40)

The factor (eP - (v...) is referred to as the characteristic Q

(1.41)

The bracket to which the total differential operator d / dx is applied shall be


referred to as the current (or flux) P

ilL
P==eP-+( (,L -ilL
-v ) (1.42 a)
ov,x ov,)( .x

(1.42 b)
28 I Mathematical Preliminaries

Thus, in "short-hand" notation we can write

I:
b

A • =A + E / + Q E (L )j dx. (1.43)

a
Along solutions v(x), we must have E(L) = O. Then only dP/dx is left in the
integrand which can be readily integrated. In two or more dimensions we would
apply the divergence theorem.

We then have

(1.44)

Here ph is the value of P at band P" is the value of P at a.

We note that if

dP = 0 -t P = const. or ph = P", (1.45)


dx

i. e., if a conservation law exists, then A • = A. This expresses the converse part
of Noether's first theorem (1918), which, in essence, states that if the (action)
integral A is invariant with respect to the action of an infinitesimal continuous
one-parameter Lie group, then there exists a conservation law.

Before proceeding further, let us summarize the results obtained so far and
introduce some further notions of group theory. We are interested in
transformations leaving the action integral invariant, because Noether's theorem
then guarantees the existence of a conservation law. Such transformations are
called variational symmetries. More specifically, if the transformation functions ,;
and c/J are functions of the independent and dependent variables only, as indicated
in (1.29), the symmetries generated are called geometric or point symmetries. On
the other hand, if ,; and c/J also depend on derivatives of the dependent variables
up to an arbitrary order, we speak of generalized or Lie-Backlund symmetries. (It
might also be relevant to note that if the transformations are functions of the
independent variables, the dependent variables and first derivatives of the
dependent variables, then they are termed contact symmetries.)

The infinitesimal transformation group (E is the infinitesimal parameter) given by


(1.29)

x =X +E';(X,V),

v = v + E c/J (x, v)
defines a variational point symmetry of the Lagrangian function L (x, v, v), if the
transformation leaves the action integral A invariant, i. e.,

A' =A. (1.46)


1.4 Systems with a Lagrangian; Noether's Method 29

From (1.37), the invariance condition is given by

(pr(l) w + ix) L = 0, (1.47)

or written out at length (cf. 1.34)

( ~ + f/J ~ + (df/J - v d()~ + d(JL = O. (1.48)


/ eXt ov dx ,x dx ov,x dx
From equations (1.47) or (1.48) the transformation functions (and f/J can be
determined once a specific Lagrangian L is given. An example of this procedure
is given further below.

In turn, the characteristic Q (cf. 1.41) is determined by the transfonnation chosen,


and the conserved current P can be constructed immediately from equation (1.42),
whereas the conservation law itself is given by (1.45).

Remark: Noethcr established also a second theorem and even though no use of it will be made
here. its brief statement is desirable for reasons of completeness. This theorem is concerned with a
family of transformations depending on r (r i; J) arbitrary functions (of independent variables).
rather than on a single (constant) parameter e in the first theorem. and states that there exist r
relations between the ensuing Euler-Lagrange equations and their derivatives. In examining
Hilbert's assertion concerning the breakdown of energy theorems in general relativity, Noether
(1918) showed that the energy relations "break down" if and only if the action integral remains
invariant under a translation group of independent variables which itself is a subgroup of
transformations involving functions (rather than constant parameters). In a footnote, Noether points
out that in c1a~sical (nonrelativistic) mechanics there exist no transformation functions which
would leave the action integral invariant. Olver (1993) terms those systems which do admit
arbitrary functions leaving the action integral invariant "abnormal systems". However, the
condition of normality is satisfied by most physically important systems of differential equations.
Thus. there appears to be no need for us to be concerned with Noether's second theorem.

One dependent and two independent variables

In a next step, we treat Lagrangians which depend (still) on one independent


variable (m = 1) but two dependent variables (p = 2). For example, the
Timoshenko beam theory is governed by such a Lagrangian.

The above development is easily extended to this case. The Lagrangian is now

(1.49)

The transformation to starred quantities reads

X' =X +E ((X, v) ,

(1.50)
30 I Mathematical Preliminaries

The infinitesimal generator w is

and the first prolongation pr(l) w

d d
prflJ w ~w + ( cf>1 v df) 0 + ( cf>2 v df) 0
dx - 1,x dx ~
1,x dx - 2,x dx ~4x

~ w + (dcf>a _ v d f)~ (a ~ 1, 2), (1.52)


dx a,x dx oV
a,x

where the total differential operator d( ) / dx is

d
+ V +V (1.53)
dx a,x dJ a,.u eN
a ",x

The invariance condition A' ~ A is now

!f ~ dt
+ cf> ~
"eN
"
+ (dcf>a - v df)~
dx tl,x dx oV
~
+ dfJL
dx
~ O. (1.54)

Replacing the partial differentiation dUdX by the total differentiation dLldx with
the help of (1.53), adding and subtracting the term cf>" d/dr (dUdV,,) and using the
product rule for the fourth tenn

leads, after some straightforward manipulations, to

-d
dx
f
cf>
a oV
- + f ~L - -
-oL
a,x
oL
oV
- v . ) }+
a,.>
,r
(l •.

(1.55)

The term in the second square bracket is the goveming Euler-Lagrange expression
E" (L), the term cf>" - (v"., is identified as the associated characteristic Q" and the
first square bracket is the current P, hence
1.4 Systems with a Lagrangian; Noether's Method 31

dP
-+Q
dx a
Ea (L)=O. (1.56)

Along solutions Ea (L) vanishes, leaving the current P constant

P =cP (/L - ilL V)=COflS£,


a ovilL +
IJ ov ayf (a = 1.2). (1.57)
a,x a.x

General case

We consider now the general case with an arbitrary number of independent


variables m and an arbitrary number of dependent variables fl, with L still
depending on the first order derivatives only

L = L (Xi. Va' V,,) (i = 1.2, .... m), (1.58)

(a ,J.2, ...• jI).

The Lagrangian is defined over an arbitrary domain .n. The one-parameter group of
(point) transformations is defined as

(1.59)
Va = Va + EcP" (Xj • V/l)'

We omit intermediate manipulations analogous to the simpler cases discussed in


detail above. The change of the action integral A due to this transformation then
follows to be

A . .= A + E / (pr(l) w + f;.)L d.n (1.60 a)

~A " jrP,,; 'Q. E. (L) )dn. (1.60 b)

The infinitesimal generator wand its first prolongation are, respectively, given by

(1.61 a)

pr(l) w =w + (d cP a ~v. dS.)~. (1.61 b)


dx.
f
a,J dx·
I
ovcr./.
32 1 Mathematical Preliminaries

The transfonnation functions (; and cP" are detennined from the invariance
condition (1.60 a) with A • = A as

Ipr(l) W +
(J ".)L 1,1
= 0, (1.62)

which leads to an overdetennined system of partial differential equations for these


functions. Once (; and cP" are known, the characteristics Q" and the conserved
cuments P j are given as

(1.63)

p.
I
= cP a OV.
elL
a,l
+"
j
ILS. - v . elL).
I' 1J a,j ov .'
i'l,1
(1.64)

(Sjj is the Kronecker symbol with Sjj = 1 if i = j and Sij = 0 for i "# j.)

Along solutions, all the Jl Euler-Lagrange expressions should vanish

(1.65)

From A* = A, (1.60 b) and (1.65) the desired conservation law follows to be

dP j
= div P = 0, (1.66)
dx j

with P j given by (1.64). Application of the divergence theorem leads to

.. do.=f.n.dcl!l+ [r=p R =0, (1.67)


f 1,1 I I

r = j
r

{} 00

where an is the surface bounding 0. and R r is the r-th of the p singularities within
0., i. e., enclosed by an.

Physical intetpretation

Already at this stage it is possible, and probably even desirable, to identify some
quantities introduced above in physical terms. In elastostatics, where v" are
displacements and x j , e. g., Cartesian coordinates, the expression
elL
a.,a (1.68)
OV.
a,1
1.4 Systems with a Lagrangian; Noether's Method 33

shall be called the physical momentum tensor or Cauchy stress tensor and

boo'J = v
oL U (1.69)
0

aJ
-

O\l 0
-

IJ
a,1
the material momentum tensor or Eshelby tensor. Thus (1.64) reads

(1.70)

If we recall

and take as transformation constant (physical) translations, i. e., cPa C" = const.
and (j = 0, the physical momentum is conserved

Pi; = 0 => Uin.i = O. (1.71)

On the other hand, if we take constant coordinate transformations, or material


translations, i. e., cPa = 0 and 0
= cj = COflst., the material momentum is
conserved

=> (1.72)

In general, cP" and 0 have to be detennined from the condition (1.62).

Besse/-Hagen extension

An important extension of Noether's first theorem has been proposed by Bessel-


Hagen (1921) who weakened the requirement

A' =A

and showed that it is sufficient to demand

A' =A + II?. 1,1 dO or A* =A +E f·· 1,1 dO (1.73)


o 0
in order to obtain conservation laws. Here Kj is an arbitrary vector point function,
whose divergence enters the integrand and the underlying symmetries are called
divergence symmetries, cf. Olver (1993).

Indeed, instead of

div P = Pi; = 0
34 1 Mathematical Preliminaries

we now have

div (P + K) = ( Pi + K); = 0. (1.74)

The Bessel-Hagen extension will be applied to an example further below.

Second derivatives in the Lagrangian

In this text, we will have occasion to deal with Lagrangians which depend not
only on first derivatives, but also on the second derivative (e. g., elementary beam
theory) at least in the case m = 1, Ji = 1.

Thus we consider a Lagrangian L

L = L (x, v, v.<' v.,) (1.75)

and a point transformation

x' = x + £ ~ (x, v) ,

v' = v + £ tP (x, v) ,

which leads to
dv'
L'=L'(x~ v~
dx'
and

We already know what x', v' , dv '/dx' and dx' are in terms of unstarred
quantities and we need to know additionally, d 2 v '/dx ,2. By the chain rule we have

2
d v' d (dv') d (dv,) dx
dx ,2 = dx' dx' = dx dx, dx"

which results with (1.32) in

=!!- Iv + £ (d tP _ v d()jdx.
dx !',x dx,x dx dx '

Recall that
1.4 Systems with a Lagrangian; Noether's Method 35

dx d(
- =1 -E-
dx' dx

and we have

=v,xx

=v,xx

Again, by Taylor expansion we find


b

A' =A + E

(1.76)

The term in square brackets is the second prolongation prl2) w of the infinitesimal
generator w = (d/dX + </Jd/dV into the jet bundle space x, v, v.. ' v.u' With (1.36),
prl2) w is given by

pr(2) w pr(1) w + (-"'- - - -


d 2A, d2 ( -2 d( v )~. (1.77)
=
dx2 dx 2
V
,x dx ,xx ov,xx
Thus we can write b

A'=A+E / !pr(2) w + c:: ) L dx (1.78)

Proceeding similarly as above, this expression may be recast by repeated use of


the product rule into the form
b

(1.79)

a
with the characteristic Q and the Euler-Lagrange equation E(L) as given before
by (1.41) and (1.24), respectively,

Q = </J - (v..'

E(L) = elL _ d ( : ) + d 22 (~L ).


ov dx .. dr. __
36 I Mathematical Preliminaries

The conserved current is now changed to

P = <P ilL _v ilL ) -


ov ,;<
+ '"
('IL ,;< ov,;<

+ d<p ilL _ <p .!!- (ilL ) . (1.80)


dx ov.xx dx eN.xx

The terms in P above can also be rearranged in a different way in terms of Q as

P = (L +Q filL _.!!- (ilL )}+ dQ ilL. (1.81)


eN
~
dx cN ~u
dx cN
,~

It may be mentioned, that the apparatus of group theory is completely developed,


such that the general case of arbitrary numbers of independent and dependent
variables may be found in textbooks on the subject, e. g., Olver (1993). If the
Lagrangian depends on derivatives of the dependent variables up to an arbitrary
order giving rise to higher prolongations, the governing equations are also found
in the literature.

Example

The question arises as to what the most general functions '" = '" (x, v) and
<p = <p (x, v) might look like, but it does tum out that to answer this question a
specific form of the Lagrangian must be introduced, e. g., examples of specific
problems have to be considered.

We pick as a simple example, the elastic, homogeneous bar of constant


compressional stiffness EA with the Lagrangian L

L = l EA U
12
• (1.82)
2

In the product EA, E is Young's modulus and A the cross-sectional area. The
displacement is given by u and u l = U,. = duldx is its first derivative, i. e., the
strain.

Obviously, the Lagrangian is given as

(1.83)
1.4 Systems with a Lagrangian; Noether's Method 37

and the associated Euler-Lagrange equation (cf. 1.22 and 1.23) consists of a single
tenn

~ (::) = ~ (EAu 1 = EAu // = 0,

where u // = d 2u/dx 2. Hence

U // = O. (1.84)

We are interested in nontrivial conservation laws only, i. e., laws which are valid
only along the solution u // = O.

The conserved current P is given by (cf. e. g., 1.42 b)

(1.85)

where Q is the characteristic given by (1.41)

(1.86)

or, upon substitution

P = - !.- (EAu /2 + cPEAu /. (1.87)


2

We now wish to determine the most general ,; and cP, such that the conservation law
dP /dx = 0 still prevails. Thus ,; and cP have to be detennined from the invariance
condition (1.48) which leads with (1.82) to

EA (d cP u / -
dx
!.-
2dx
d,; It /2)= O. (1.88)

The same condition may be obtained directly from the condition dP /dx = 0,
together with the equation of equilibrium (1.84).

We consider first the geometric transformation (symmetry)

,; = ((x, u),
(1.89)
cP = cP (x, u).

Substitution into the above equation leads to


38 I Mathematical Preliminaries

=u l at> +U 12 (Oc/> _ i O')+U I3 (_ i 0')=0.


ax m.t 2ax 2~t

Since (and cP, by definition, do not depend on u I, the above equation is satisfied
only if the factors of each "form" of u 1 vanish separately.

Thus

at> =0 - cP = cP (u),
ax
o(
- =0 - ( = ( (x) ,
Ju

j ~ ~ C2l _
1
cP=-Cu+C
2 2
1 J( = 0 _ = 4

2 ax o(
- =C
ax 2

where the arrow ~ indicates "leads to" and the C;'s are arbitrary constants.

Now (and cP are completely determined and the current P is

P = (C 2 X + C/) ( - i2 EAu 12) - C, ( -


-
i2 EAuu I) + C4 EAu I. (1.90)

We next introduce the designations already employed when discussing elastic bars
in Kienzler (1993).

-i EAu 12 =B is the material force, (1.91)


2

-i EAuu 1 = H is the material scalar moment (virial), (1.92)


2

EAu 1 =N is the physical force. (1.93)

The current P is then

P = C/ B + C2 (Bx - H) + C4 N = COl/st. (1.94)

and there follow three conservation laws

= 0, (1.95 a)
1.4 Systems with a Lagrangian; Noether's Method 39

(Bx -HY = 0, (1.95 b)

N' = O. (1.95 c)

Let us now consider, still with geometric symmetries, the extension of Bessel-Hagen
mentioned earlier. We introduce a function K = K (x, U, U /) and the conserved
current is now augmented by K

P +K = canst, (1.96)

which, by differentiation, (cf. 1.74) and (1.88) becomes

(1.97)

We notice immediately that K cannot depend on 1/ , because oK/aui will


introduce u II as a factor and the product has to vanish. Ordered by the "forms"
of u~ equation (1.97) can be written as

+U12 (e*P
a;; L2 o()+
c1l

+ U /3 (_ L o()+ _1_ oK = O.
2 OU EA d.x

From the last term above we conclude K = K (ll) and from the next to the last
( = ax).

The second term implies

04> -l o( = 0
c1l 2 d.x

and we conclude

4> = uiJ (x) + 12 (x),


( = 2/ j~ (x) dx + C/.

The first tenn in the conservation law implies

1
--
oK
EA ou
40 I Mathematical Preliminaries

or, by substitution

It follows that

11/ = const. = C3

and

12/ = const. = C5

or, by substitution

1 oK
- -
EA ou
By integration

and

Thus

(1.98 a)

(1.98 b)

K = - 21 C 3 EAu 2 - C.s EAu + C6 EA. (1.98 c)

The constant C6 does not contribute anything of interest.

The conserved current is

P + K = (C I + C2 x + C3 xl) (- i EAu /2) + (1.99)


2
1
+ ( - C2 U + C.I xu + C4 + C,. x) EA u / +
2

+ C3 (- i EALl) - C5 EAll .
2
1.4 Systems with a Lagrangian; Noether's Method 41

We introduce the additional notation

R - !.... EAu 2 (1.100)


2 '

v = EAu. (1.101)

Thus

P+ K = C/ B + C 2 (Bx - H) + C] (B~ - 2Hx + R) +

+ C4 N + C5 (Nx - V), (1.102)

which leads to the five independent conservation laws

B/ = 0, (1.103 a)

(Bx - Hy = 0, (1.103 b)

(B~ - 2Hx + R)! =0 , (1.103 c)

N/ = 0, (1.103 d)

(Nx - V)! = o. (1.103 e)

We observe that the employment of the Bessel-Hagen extension was indeed


fruitful, since the number of conservation laws has increased from three to five.
Their physical significance will be discllssed in Chapter 9.

We next explore whether consideration of generalized symmetries might similarly


bring an enrichment of conservation laws and we consider

, = ,(x, u, u),

cf> = cf> (x, u, u). (1.104)

This is the most general form of transformation one can think of, for the
dependence on u// in ,and cf> might be eliminated by the governing differential
equation.

Introduction of this into dPldx = 0 results in

and we note immediately that nothing is gained, because u// = 0 in our example.
Would the Lagrangian, in another example, depend on derivatives of u higher than
42 I Mathematical Preliminaries

the first , it is possible that further conservation laws could be established by


consideration of generalized transformations (symmetries).

1.5 Systems without a Lagrangian; Neutral Action Method

For systems without a Lagrangian (e. g., for some systems with dissipation) the
governing equations are no longer the Euler-Lagrange equations of a variational
problem, but rather a set of partial differential equations which has been obtained on
the basis of some balance laws (e. g., balance of linear momentum) and some
postulated constitutive relations (e. g., standard linear solid in viscoelasticity).

Until recently, no systematic procedure existed to derive conservation laws for such
systems, but a few years ago, a novel methodology was advanced, cf. T. Honein et
al. (1991).

Null Lagrangian

In order to develop this new procedure, a result of the calculus of variations is


required as discussed below. This calculus itself will, however, not be needed for
the application of the new method.

In addition to trivial conservation laws (as mentioned in Section 1.3), trivial


variational problems do also exist. They occur when the Euler operator (1.25) is
applied to a Lagrangian function which itself is a divergence.

Let gj = gj (X k, va' Va,k' .. .) and the Lagrangian L be given by


- dg;
L=-=g;,j'
dx.I
Then it follows

(1.105)

as will be shown below. The converse is also true. If the Euler-Lagrange equation
is satisfied identically, then l is a divergence, i. e.,

if E (l) = 0 then l = gi, i . (1.106)

Such a Lagrangian is called a "null Lagrangian".


1.S Systems without a Lagrangian; Neutral Action Method 43

Let us consider first a simple example in one dimension. The function g is taken as

g=21 v 2 ,
where v = v(x) is arbitrary.

The "divergence" of g is

dg = v v
dx ,.<'

such that L = dg/dx = L (v, v). The associated Euler operator is given in (1.22)
and the Euler-Lagrange equation (1.23) is identically satisfied by any v (x)
d
E(L)
=!~ -
dx
(~,.. )/(v v,..)

= VoX -
d
- v
dx
= O.

To throw additional light on this result, we consider the action integral of L , i. e.,
b b

A d"
-i2
dx
dx = g(h) - g(a) = -I
2
V2 I b
.
/ x=a
a a

Thus A does not depend on the explicit functional dependence g(x) inside the
interval a, b, but only on the values at the end points a and b. Hence any function
v = v(x) satisfying the boundary conditions 112 V (a) = g(a) and 112 v2 (b) = g(b)
will lead to the same value for A. The variation of v to v + 8v inside a, b does not
change A and, since 8v (a) = 8v (b) = 0 is always assumed, it follows

M =0 E(L) =0.

In two or three dimensions, the same argument follows from the divergence theorem
(1.6)

A = !L
B
dV =
)

B
I:g __dV=
tii,/ )

S
;;
tii
n , dA .

The value of the action integral depends only on the values of g along the boundary
S. Since the variations 8v" vanish along S, the action integral is not affected by the
variation of A. Thus

E" (L) .- 0

for any vector point function g; = g; (xk , v"' v,,_ k)' For a rigorous treatment of the
material above the reader is refelTed to Olver (1993).
44 I Mathematical Preliminaries

The procedure

Having introduced the concept of null Lagrangians, the basic idea of the new
procedure is as follows:

As we have seen in Section 1.4, a conservation law Pi,i = 0 is expressible in terms


of the Euler-Lagrange equations E" (L) as (cf. 1.60 b)

(1.107)

where the elements of the set Q" are the characteristics of the conservation law
which are fixed by the transformation functions (; and cP" as

Now, instead of the Euler-Lagrange equations EjL) = 0, we have a set of J1.


equations on it dependent variables v" (a = J. 2..... it) which we can abbreviate as

(1.108)

where the multi-index notation of Olver (1993) - already discussed above (cf. 1.26)
- is used. The possibility is not excluded that they may be Euler-Lagrange
equations.

Now, instead of the characteristics Q being specified by the transformation functions


(, and cPa , we determine Q" ' employing the symbol - J:, instead (in order to avoid
confusion), such that

f".:1" = Pi'; , (1.109)

where the functions f" are not considered to be pre-determined as Q" were.

Thus it follows from (1.109) that a requirement for the existence of conservation
laws is that

(1.110)

since E{3 (Pj ,;) = 0 as shown above. The sumJ:. Ll" implies that it is formally a null
Lagrangian whose action integral

A = Irdv
B
= 1f..:1
B
a dV
if = Ip
S
} n·} dA

has vanishing variation for any dependent variable v"' i. e., 8A = O. In other words,
in order to construct conservation laws for any system (whether or not it possesses
a Lagrangian), governed by a set of differential equations .:1" = 0, we try to
construct a product f".:1" whose action A does not change variationally, i. e., the
1.5 Systems without a Lagrangian; Neutral Action Method 45

action A behaves neutrally under its variation. Thus the name "Neutral Action"
(NA) method was given to this procedure (Chien, 1992).

Example

As an application of the NA method let us consider again the elastic bar, but start
not with the Lagrangian L, but rather with the differential equation ~ = u II = O. In
following the NA method we multiply the differential equation by f = f (x, u, U I )
and apply the Euler operator to the product fu II , and set it equal to zero, i. e.,

E (fu II ) = 0 (1.111)

recalling that (1.24)

E (fu II ) = 13 (fu II) _.!!:..- (13 (tu II ) ) + d2 (13 (fu II ) )

t7u dx Ju I dx 2 Ju II

The first two terms above vanish and the last may be evaluated in steps as

-
d dJ
--
dxdx

The only term which may be set equal to zero at this stage is the last, because it
contains a derivative higher than the highest appearing in the differential equation
and should be eliminated by substitution into this equation. Thus
46 I Mathematical Preliminaries

+ a2f u /2 = o.
au 2

Applying the same arguments as before, the factors of u //2 and u II must vanish
separately, io eo,

From the fIrst above, by integration, we find

f = h (x, u) U I + g (x, u)o

Substitution into the second yields

Since hand g do not depend on u I, we conclude

and fInally

(1.112)

Substitution back into d 2f/ dx 2 = 0 and full use, now, of the governing differential
equation U ll = 0 results in

f/ u / + f/ u + f/ + 2u / f/ 00

The factors of u, u I and the tenn without these must vanish separately, because h,
f2 andfJ are all functions of x only

f/ =0 ~h = 2"1 C j x + Co ,

f/ =0
1.5 Systems wilhout a Lagrangian; Neutral Action Method 47

[/ + 2[/ = 0 ---71:J = -
1
-C j
2'
r - -C
1
2
2X -
1
-Ct.
2

The labelling of constants has been chosen to coincide with those of the same
example, but treated via Noether's theorem in the previous Section.

Before proceeding we modify the product f~ as follows

[ 4 =[ U II = .!!:- (fu I) - U I df
dx dx
2
_ did
--([ll)
--
(d ) df
-fl l + l I -
dx dx dx· dx 2

and, since d2f/~ = 0, the current P is found to be

P = [u I _ d[ u.
dx
The complete f is

I Clr···
' -C,x - -C ) 1 1
f I
=u (-- J2 . 2 - 2·
f

1
I U (-C] x (1.113)
2
and its derivative is, because II II = 0, given by

d[ = u/(- C1x - ic, + iclx + Co) ~


dx . 2 - 2'
1
+ - l l C 1 + C.s .
2 .
It follows that P can be expressed as

- UU I ( - .l2 C x -
3
i2 C
2
+ Co) - II 2 i2 C
3
- llC j

- i l l 12 (C J + C, X + C.I x2) +
2

1
+ II / ( - C) u + C1 XII + C4 + c~ x) + (1.114)
2 - .
48 I Mathematical Preliminaries

Comparsion of the above expression for the current P shows that it is up to the
constant factor EA identical to the corresponding expression obtained by the
application of Noether's theorem together with the Bessel-Hagen extension (1.74).
The comparison of the two procedures which lead to the same result indicates that
the NA method appears to be more straightforward, since it deals only with one
function J, rather than with three, namely (. cf> and K.

1.6 Discussion

The establishment of conservation laws may be schematically illustrated by means


of a flow diagram, Fig. 1.4.

by
L

y~~:~her&
/ ~el-Hagen

E(L)=O } - - - - - - r - - - -
.. CL
1:1=0 ~NA

Fig. 1.4: Flow diagram pertaining to conservation laws

If a Lagrangian function L is available (the arrow to L from above is supposed to


indicate that L has been constructed based on some external, most likely physical,
considerations) then conservation laws CL can be established by the application of
Noether's theorem. A broader class of such laws is obtainable if Bessel-Hagen's
extension is used, as indicated by the arrow from L to CL.

On the other hand, based on L and applying the rules of the calculus of variations,
symbolized by S, the associated Euler-Lagrange differential equation indicated by
E(L) = 0 can also be derived, made graphic by the arrow from L to E(L) .= o.

The governing differential equations of a problem may, however, be not the result
of a variational procedure (in the absence of L), but could have been formulated on
the basis of some balance laws and some constitutive relations describing material
behavior. This will occur for most dissipative systems and the corresponding
equations are symbolized by a = 0, the arrow to this equation indicating their
establishment based on outside considerations.
1.6 Discussion 49

The arrow leading from the corner of differential equations (either E(L) = 0 or ~ =
0) to the CL corner and labeled NA symbolizes the fact, that the Neutral Action
method can be applied equally well to E(L) = 0 or ~ = 0 and leads to conservation
laws. If it is applied to E(L) = 0, then one obtains the same conservation laws as if
Noether's method, but together with the Bessel-Hagen extension, were applied to the
corresponding Lagrangian L.

It should be reemphasized in this context that for equations of the type ~ = 0, no


procedure existed for a systematic construction of conservation laws, until the NA
method was advanced. What was done was to concoct, by hook or by crook, an
artificial L (cf., e. g., the "mirror image" system used in Morse & Feshbach, 1953)
and then apply Noether's theorem, with or without the Bessel-Hagen extension.

But, as it turns out, even if a "natural" L exists and thus E(L) = 0 is available, the
NA method applied to E(L) = 0 leads to conservation laws in a more straight-
forward manner than application of Noether's theorem with the extension of Bessel-
Hagen. In the former (NA) method, as already mentioned, one deals with only one
set of functions to be determined, while in the latter, three sets of functions are to
be found, leading to identical results.
2

Linear Theory of Elasticity

2.1 General Remarks

The linear theory of elasticity, thanks to its relative simplicity and wide
applicability, constitutes the principal model of continuum mechanics which will be
employed in this text.

The essential elements of this classical theory are summarized in Section 2.2 and
specialized to plane and antiplane elastostatics. In Section 2.3, Noether's theorem is
used, with the extension of Bessel-Hagen, to delive conservation laws in three-
dimensional and plane elastostatics, leading to three such laws in physical space and
three laws in material space. Two alternative procedures to establish conservation
laws are presented in Section 2.4, namely the Neutral Action (NA) method and the
method based on the application of differential operators acting on the Lagrangian
density or its moment. Finally, a third alternative procedure, based on duality, is
briefly mentioned.

2.2 Elements of Linear Elastostatics

It is the objective of the present Section to briefly restate the basic equations of the
theory of linear elasticity and to introduce the notation used. For a more
comprehensive presentation of the subject, the reader is referred to textbooks on
elasticity, e. g., Timoshenko & Goodier (1970), Hahn (1985), Barber (1992), Gould
(1994).
52 2 Linear Theory of Elasticity

Kinematics

Consider a body B which is embedded in a Cartesian coordinate system X k in the


undeformed state. Imagine further that the body is deformed by (physical) forces
into the configuration B~ A material point C in the undeformed state at position X
finds itself as C/ at position x after deformation (Figure 2.1).

-71""=----------. X2

Fig. 2.1: Undefonned C and defonned C' position of a material point and displacement vector u

The difference between the positions in the deformed and undeformed state is
designated, as usual, by the displacement u. If the body is continuous before and
after deformation, u is a continuous vector field defined inside B with components

Ui = U i (xJ i,k = 1, 2. 3. (2.1)

and
cUi
U.
I,k
=-
dt (2.2)
k
is the displacement gradient. In the geometrically linearized theory of elasticity,
the symmetric part of ui,k is the strain tensor

E··
IJ
= -21 (u].I.. !- U . .),
I,J
(2.3)

whereas the antisymmetric part is the rotation tensor


1
(v.
lj
= -
2 (u.
J,I
- u .) ,
I.j
(2.4)

With the completely skew-symmetric, third-rank tensor of unity (the permutation


tensor or Levi-Civita tensor)
1 if ijk cyclic: 123, 231, 312
E
ijk
j
= - 1 if ijk anticyclic: 132, 321, 213
o otherwise,
(2.5)

the three independent quantities of (Vi) may be represented as components of the


rotation vector (J

(2.6)
2.2 Elements of Linear Elastostatics 53

As in the previous Chapter, the summation convention is implied. The relation


between the rotation tensor and the rotation vector is given by

(2.7)

The rotation tensor plays a minor role in the classical, linear theory of elasticity, but
it becomes an important quantity when discussing components of the Eshelby
tensor, as will be shown later.

From the six components of the strain tensor cij (2.3), three continuous components
of the displacement vector ui for a simply connected region can be determined by
integration, if and only if the equations of compatibility are satisfied, i. e., the
components of the incompatibility tensor Tfij vanish

(2.8)

For a multiply connected region, additional (integral) conditions have to be


satisfied. When studying dislocations or, more generally, defects in solids or
crystals (e. g., Kroner, 1958) the components of Tfij "# 0 become important.

Statics

External forces cause internal stresses. In order to introduce the concept of stress,
imagine that the body is divided into two parts. To maintain the left part (say) in
equilibrium, there must exist forces distributed over the surface of separation S.
Consider an element of area of this surface characterized by the outward drawn
normal unit vector n, as shown in Figure 2.2.

XI

Fig. 2.2: Surface of separation S with 1m element of area M (normal n) showing the resullant force
M and resullant moment liM exerted through contact by the other part of the body (not
shown) on the part of the body under consideration

The resultant force D.F and the resultant moment D.M are indicated. In the limit as
D.A ~ 0, D.F and D.M also approach zero, but the ratios D.F/AA and D.M/AA remain
finite
54 2 Linear Theory of Elasticity

. l!:.F dF
llm - = - = t (2.9)
dil-O M dA n'

. l!:.M dM = J!..
t1m - - (2.10)
dil-O M
dA n

These limits are called the traction vector ! and the couple stress vector It. In the
classical theory of elasticity, It is set equal to zero. In the so-called Cosserat-type
theory of elasticity, cf. e. g., Mindlin & Tiersten (1962), Mindlin (1967),
dM/dA = It "# O.

The traction ! depends on the orientation of the element of area and is, thus, a
vector function

! = t (n) . (2.11)

The components of the traction ! i are related to the corresponding components of


the normal vector nj through the components of the stress tensor Gji via the
celebrated Cauchy relation

t"
n
= G ..
I'
n.I , (2.12)

or, in other words, the stress tensor maps the normal vector into the traction
vector. This mapping is linear and homogeneous. The Cauchy stress tensor, or the
physical momentum, is symmetric in view of the equilibrium of moments
(It = 0, balance of angular momentum)

(2.13)

and is connected to the vector of body forces qi per unit of volume by the
equation of force equilibrium (balance of linear momentum)

(2.14)

The diagonal terms G ll ' G 22 and G.u are called normal components of stress or
normal stresses for short, the off-diagonal terms G l2 = G 2l , G 23 = G 32 and
G
31
= G n are called shear stresses.

Due to the symmetry of the stress tensor, three perpendicular directions of an


element of area always exist for which the shear stresses are zero. These principal
directions and principal stresses follow from the eigenvalue problem

(2.15)

The roots of the characteristic equation

(2.16)

are always real and are called the eigenvalues or principal stresses G, ~ G" ~ G m ·
2.2 Elements of Linear Elastostatics 55

The principal directions n" nil and nlll are evaluated from the linear, homogeneous
algebraic system of equations (2.15). The coefficients )/. )2' )3 of the characteristic
equation (2.16) are the three invariants of the stress tensor in three dimensions,
namely the trace, the square trace and the determinant, respectively. They are given
as
1
= a j; = -2 Eikm Eikll amn'

1
=-E
2 ikm
E
if"
akl anlfJ '

(2.17 a-c)

If the coordinate system chosen coincides with the principal directions, the fonnulae
above reduce to

(2.17 d)

(2.17 e)

(2.17 f)

Of course, it is possible to introduce also invariants of the strain tensor and the
rotation tensor by simply replacing a by E or UJ on the left and right hand sides,
respectively. It may be mentioned that the only non-vanishing invariant of the
rotation tensor UJij (2.4) is with (2.6) given by

(2.18)

Hooke's law

Assuming the body to be continuous, the generalized Hooke's law of the linear
theory of elasticity is given by the linear tensor relation (mapping)

(2.19)

The fourth-rank tensor of elasticity Cijkl satisfies the following symmetry relations

(2.20)

leading to 21 independent constants. If the material behavior is isotropic, c;j*/ is


determined by merely two constants, namely the Lame constants A and II
56 2 Linear Theory of Elasticity

(2.21)

Thus, equation (2.19) reads

(2.22 a)

For convenience, Hooke's law may be expressed via (2.3) also in terms of the
displacement gradient as

a IJ.. ~ JL (u. + u.) + A U


kk
S.. (2.22 b)
I,J J,I ,IJ

Inversion of (2.22 a) leads to


1 A Sij
(2.23)
Eij ~ 2JL aij - 2JL (3A + 2f..L) a kk ·

Alternatively, the Lame constants A and f1 may be replaced by Young's modulus E


and Poisson's ratio u, or by the bulk modulus K and the shear modulus G

A ~------ ~K
uE - -2 G (2.24 a)
(1 + v) (l - 2 u) 3 '

JLC
E ~ G. (2.24 b)
2 (1 + u)

For later use the so-called isotropy condition (Eshelby, 1975a) is given here
which is easily verified by (2.22). It is written either as

(2.25 a)

or, alternatively, with the permutation tensor, as

(2.25 b)

Strain energy, Lagrangian function

The strain energy density W per unit of volume is defined as

(2.26)

Since the material behavior is linear (2.19), equation (2.26) is readily integrated
leading to
1
W -
2 a·.IJ E·
lj
2.2 Elements of Linear Elastostatics 57

(2.27)

By differentiation with respect to Eij' the character of W as an elastic potential is


revealed immediately, leading to the constitutive relation

a.. = -,
oW (2.28)
IJ tk ..
Jl
which indicates an alternative manner to introduce stress. Due to the symmetry of
the stress and strain tensors, (2.28) may also be expressed in terms of the
displacement gradient (2.2) as

a .. =--.
oW (2.29)
IJ ou..
j,I

Thus, the strain energy may be interpreted as

W = W (xk , u j ) • (2.30)

W might depend on xk explicitly, as indicated above, if the material constants are


functions of X k' i. e., ~ and A (or E and u) vary from position to position (material
inhomogeneity).

Similarly to W (2.29), a potential of external forces V might be introduced, if the


force field is conservative, as

(2.31)

If the body forces result from dead loads, the potential of external forces is given
by

V = - qj Uj ' (2.32)

Also V = V (xk , u) might be an explicit function of x k if the forces per unit of


volume qj are functions of X k (physical inhomogeneity).

With (2.30) and (2.32), the Lagrangian function (already introduced in Chapter 1,
e. g., in equation 1.58) is

L = - (W + V) = L (x k, Ui ' ui ) • (2.33)

If the material is homogeneous, both physically and materially, L does not depend
explicitly on x k leading to

L = L (u i , uj ) • (2.34 a)

In the absence of body forces (qi = 0), L depends merely on the displacement
gradient
58 2 Linear Theory of Elasticity

L = L (u j ) • (2.34 b)

The action integral introduced in Chapter is defined as an integral of the


Lagrangian over a body B with volume V as

A, / L lx, ' U" U;) dV. (2.35)


B

Boundary conditions

The three sets of field equations, namely

equations of equilibrium (2.14)


1
kinematic relations t;..IJ = -2 (u·),1 + u·)
lJ. ,
(2.3)

Hooke's law (2.19)

form a set of 15 equations for the 15 unknowns U j , E jj and a jj • The aim of the
theory of elasticity is to solve these field equations for a given set of appropriate
boundary conditions and body forces (direct problem).

Consider a body B with surface S. The surface S might be divided into two parts,
a surface St at which the traction vector ,( is prescribed and a surface Su on
which the displacement vector is given ,t
s
ti = (ajj n)s, at SI' (2.36 a)

ut = (u.)s at Suo (2.36 b)


s 'u
It is assumed that every point of the surface belongs either to S, or to Su' but
mixed boundary conditions, e. g., boundaries on an elastic foundation

t; = tj (!:Ii) (2.36 c)
s s s

are also admissible. It can be proved that the field equations (2.14), (2.3) and
(2.19), together with the boundary conditions (2.36), admit a unique solution (cf.
Timoshenko & Goodier, 1970).

Navier - Lame equations

The set of field equations (2.14), (2.3) and (2.19) might be reduced by eliminating
the components of stress and strain. Starting with the equations of equilibrium
2.2 Elements of Linear Elastostatics 59

(2.14)

and replacing the stress components by the strain components via Hooke's law

(2.22)

and, finally, employing the kinematic relations

£ij = "21 (uji + ui ) (2.3)

leads to the Navier - Lame equations

jl Ujjj + (A. + jl) Ujjj + qj = 0 . (2.37)

With it, three partial differential equations are given to calculate three unknown
displacements. The method of solving these equations is usually based on the
Boussinesq-Papkovich-Neuber potentials (cf. Barber, 1992).

Plane elastostatics

Many problems of the theory of elasticity can be approximately treated as two-


dimensional. A problem is two-dimensional, if all field quantities and boundary
conditions depend on only two coordinates, e. g., X J and x2• In the plane-strain
approximation it is assumed that uJ and u2 are independent of x3 and u3 = O.
From (2.3) it follows

Hooke's law (2.22) yields

Therefore, the associated state of stress is not plane.

The plane-stress approximation starts from the assumption

From Hooke's law it follows that the associated state of strain is not plane, namely

£13 = £23 = 0,
U
- - (all + a22)·
E
60 2 Linear Theory of Elasticity

Various two-dimensional problems of the theory of linear elasticity are worked out
and discussed in detail (cf. Timoshenko & Goodier, 1970; Barber, 1992). Such
problems are formulated conveniently in terms of stress.

In plane elasticity, the field equations and boundary conditions are essentially the
same as in three dimensions, except that the range of the indices is restricted to two.
Hooke's law, however, is slightly different in the plane-strain and the plane-stress
formulations. The following representation is most convenient, because plane stress
and plane strain can be expressed in terms of one set of equations. With the shear
modulus G (cf. equation 2.24 b) and the material constant K given as
for plane strain
(2.38)
for plane stress,

Hooke's law turns out to be


3 - K
a ji = G (2Eji + -- E l>..), i, i, k = 1,2 (2.39 a)
, ' K - 1 kk I)

or, in inverted form


1
3) akk' l>,). (2.39 b)
E·· = -
I) 8G (4a..I) + (K -
,

The strain-energy density is then given by


1 3 - K
W = G (E·· E·· + - - - E.E··) (2.40)
IJ 'J 2 K _ 1 "JJ

such that the constitutive relations

a .. =--
oW
I) ok.
I)

recover Hooke's law (2.39a) for plane stress as well as for plane strain. The
associated Navier - Lame equations reduce from (2.37) to
K - 1
(K-l)u ... +2u ... + - - q . =0. (2.41)
I.» ),)1 G '
Remark: Alternatively, equations (2.22), (2.23), (2.27) and (2.37) may be used directly for plane
strain (indices ranging between 1 and 2). For plane stress. however, A. has to be replaced by

A.' = A. (1 - 2 u)/( 1 - u) ,

and in Eq. (2.23), the factor 3,,1 is to be replaced by 2X.

Antiplane Elasticity

Some problems may be approximated by a particularly simple form which is


2.2 Elements of Linear E1astostatics 61

referred to as antiplane shear. This state is defined in terms of the components of


displacement as follows

(2.42 a)

(2.42 b)

(2.42 c)

The kinematic relations are then


1
e13 = 2 ~1 '
(2.43 a)
1
e23 = 2 ~2 '
(2.43 b)

e 12 = ell = En = E33 = 0 (2.43 c)

with the condition of compatibility given as

E13,2 = E23,1 • (2.44)

Hooke's law reduces to

(2.45 a)

(2.45 b)

(2.45 c)

and the equation of equilibrium follows to be

(2.46)

If we eliminate the stresses from the equilibrium condition by Hooke's law and
replace, in turn, the strains by the displacement X, we can conclude that the
Navier-Lame equation of antiplane shear is a harmonic equation for X

X··
,II
= LlX = 0 , (2.47)

provided that the material is homogeneous, i. e., the shear modulus G is constant.
62 2 Linear Theory of Elasticity

2.3 Conservation Laws of Linear Elastostatics

Several possibilities exist to derive conservation laws, as discussed in Chapter 1. In


this Section, we will use the classical approach via Noether's theorem (with Bessel-
Hagen's extension), whereas alternative methods of derivation are discussed in
Section 2.4.

The governing Lagrangian L = L (xk • Uj • uj ) is given by (2.33). With (2.29) and


(2.31) it follows that

dL
(2.48)
oil..
),1

dL
- = +q .. (2.49)
ok I
I

For later use, the following abbreviation for the explicit derivative of L with
respect to xj is introduced as

dL d (W + V)
=ji' (2.50)
at., ati

Conservation laws for Lagrangians of the form L = L (xk • U j , uj ) are derived in


Chapter 1 following equation (1.58). Here V k is to be replaced by Uk' The Euler-
Lagrange equations (1.65)

Ek (L) = dL -!l-. dL = 0 , (2.51)


~k dx.I clt k,I·

upon substitution (2.48) and (2.49), are identified with the equations of
equilibrium (2.14)

a ,),,
... + q.) = 0 .

If the constitutive equations (2.48) would not have been used, but the partial
derivative aLI au.. I,)
would have been evaluated instead by use of (2.33), (2.27) and
(2.3) as

then the Navier-Lame equations (2.37) would folIow from (2.51).


2.3 Conservation Laws of Linear Elastostatics 63

The associated current is given by (1.64) as

aL aL 1, 2, 3 for 3D
p. cf> - (L 8.. -
I
=
"ou. ~l
+ (;
J
U . -)
Y"'J ou . ~, 1, 2 for 2D.

Since the range of latin indices i, j and greek indices a is the same, it is no longer
necessary to make a distinction between them and thus P j can be rewritten as

(2.52)

The Eshelby tensor of linear elastostatics is given by

(2.53)

Along solutions of the Euler-Lagrange equations Ek (L) = 0, the action integral


(2.35) is invariant with respect to infinitesimal transformations given by cf>j' 0, if
the current P j is divergence-free, i.e.,

p .. = dP; = O. (2.54)
1,1 dx.
I

From (2.54) or, alternatively, from the invariance condition (1.62), the unknown
functions cf>j = cf>j (xk• u) and 0 0
= (X k, u) can be determined, which we proceed
to evaluate. With
d (W + V) a (W + V) a(w + V)
Uk,j;
dx.I atj OUk,j (2.55)

and

d( a( a(
_J = _ J + _ J U
dx.
I
dx.I auk k,;

it follows from (2.52) - (2.55)

o = (Jcf>j + acf>j Uk') G .. + cf>. Goo. +


dx; auk ,I IJ J IJ,I

+ ( ac:
_J
at;
+
a". _ J Uk'
OUk ,I
) boo
I)
+
64 2 Linear Theory of Elasticity

= fOcP.at: +
0rP·
t.m~ uk,;
)
aij - cPj qj +

f at;
o( o( )
+ _ J + _ J Uk' b .. - ( j .. (2.56)
t.m ,I lj J J
k

This equation is a polynomial function in first derivatives uij with coefficients


which depend on Xl and u;. Rearranging (2.56) leads to

(- U.~) ',J
(2.57 a)

o(
+ _ J
Ou
f k
Uk'
,I
V (j..
'J
j +

+ _J o( f W (j.. - a' Uk' j + OilocP· fIt,.. a.. j


_J +
at. 'J Ik ,J ",I IJ
I k

o(
+ _ J
at.

jJV j (j ..
lj
+ 1- It;) (2.57 d)

+OrPat. rfa ..j+


j
'J j- Itk,;1 (2.57 e)

(2.57 f)

- ( j .. (2.57 g)
J J

Setting all the coefficients equal to zero leads to the requirement that the functions
o and cPj satisfy an over-determined system of linear differential equations.

From (2.57 a) it follows that 0 must not be a function of u j ' Thus

With it, part (2.57 b) is satisfied and comparison of (2.57 c 1) with (2.57 c2) yields
the conclusion that
OrPj
= hjk (xJ,
Ou k
i. e.,

From (2.57 e) it follows that the functions hjk (xtJ are actually constants and, due to
the symmetry of the stress tensor, the terms ahj laxi form a skew symmetric constant
matrix.
2.3 Conservalion Laws of Linear Elastostatics 65

Thus

Again, in view of (2.57 c 1) and (2.57 C 2), a0 lax; are also constants leading to

Proceeding further with the discussion of (2.57 c 1) and (2.57 c 2), we split the
constant matrices aij and {3;j into symmetric and antisymmetric parts and,
additionally, the symmetric parts into spheJical and deviatoric parts as follows

{3jJ.. = m 8··'i + {3./


J'
+ en £ nlJ'
..

with m, y, en, kn. {3ji~ a/ = const. and

{3f; = (3~ ) a~ = a~. , {3(; = a(; = O.

With this, it follows from (2.57 c 1) and (2.57 c z) after some rearrangements

mW(n-2+2y)+ (2.57 h)

(2.57 i)

+ a·J kl a·,)1 Uk',i - {3./a'k


)1 t
Uk'J = 0 •
(2.57 j)

Depending on the dimension of the problem, /I = 8;; is equal to 3 in three


dimensions, equal to 2 in two dimensions and equal to J in one dimension. The
first term (2.57 h) vanishes, if y is chosen to be
2 -n
y =-2-'

and the second term (2.57 i) vanishes if

en = kn

and if the isotropy condition (2.25 b) is applicable. The third term (2.57 j) vanishes
only,

if

This can be shown by inserting Hooke's law (2.22) into (2.57 j) and compaJing the
coefficients of every uij uk.! .
66 2 Linear Theory of Elasticity

The symmetry transfonnations are thus completely detennined to be

(2.58 a)

(2.58 b)

But the requirement Pi,j = 0 places additional constraints on the inhomogeneity qi


and ji of physical and material space, respectively, as can be seen from (2.57).

If cPj '#. 0 it follows from (2.57 f) that the physical space has to be
homogeneous, i. e., body forces are not admissible, qj = O.

If 0 '#. 0 it follows from (2.57 g) that the material space has to be


homogeneous, i. e., jj = O. In view of (2.50), (2.27) and (2.32) it follows
that the material constants A and JI (or E and u) are not allowed to be
functions of xj , but constant body forces qi are admissible.

If 0,i '#. 0 it follows from (2.57 d) additionally that body forces have
to vanish.

From (2.52) the current Pi can be represented with (2.58) as


2 -n
- p.I = (en Enk)· Uk + m -2-) u. + m "E" "X) + r· ) a..
k ')
.1.'
I)
+

(2.59)
+ (enEnk,X
) k +mx.) +k) b.
I)

As mentioned above the constants in , m, m n , Ij and kj may be chosen arbitrarily.


Upon setting in succession these constants, except one of them, equal to zero, the
following conservation laws are obtained in local fonn

rj '#. 0 : a IJ... ,I
= 0, (2.60 a)

mn'#. 0: Enkj (xk a;).i = 0, (2.60 b)

kj '#. 0 : b Ij,l
~ 0, (2.60 c)

en'#. 0: Enkj (xkb jj + Uk a;),i = 0, (2.60 d)


2 -n
m'#.O: (x.) b I).. + -2- u·) a).
I) ,I
= O. (2.60 e)

Equation (2,60 a) is the homogeneous equation of equilibrium and states that the
stress tensor a jj is divergence-free in the absence of body forces qj' The constants
rj describe rigid body translations in physical space. Equation (2.60 b) confinns
together with (2.60 a) the symmetry of the stress tensor a ij = aji' The constants
mn describe rigid body rotations in physical space. Due to material translations kj
it is shown from (2.60 c) that the Eshelby tensor is divergence-free in the absence
2.3 Conservation Laws of Linear Elastostatics 67

of inhomogeneity forces jj' Due to material rotations, the physical space and the
material space are coupled. Equation (2.60 d) leads with (2.60 a) and (2.60c) to a
symmetry relation

(2.61)

From the definition of the Eshelby tensor (2.53) it follows that this equation holds
only if the isotropy condition (2.25 b) is satisfied, i. e., the conservation
law (2.60 d) requires that the material behavior is isotropic.

With the strain-energy density


1
W = -2 a··IJ U .. ,
I,)

the conservation law (2.60 e) provides a connection between the trace of the
Eshelby tensor and W

hi; + (2 - fl) W = 0 . (2.62)

This equation is valid, if the material is linearly elastic, i. e., the strain energy
density has to be a homogeneous quadratic fonn in the displacement gradient.
The associated constant m describes a scale or similarity transfonnation, which
couples, again, physical space with material space.

Prior to discussing equation (2.59) further, it may be stated that the infinitesimal
group

(2.63 a)

(2.63 b)

allows for five independent conservation laws. The conservation laws (2.60) agree
with those obtained previously by GUnther (1962) and Knowles & Sternberg
(1972). As was shown in Chapter 1, the Bessel-Hagen extension provides a
helpful tool to increase the number of independent conservation laws. Because the
five conservation laws (2.60) are related to five independent transformations,
namely physical and material translation, physical and material rotation and (only)
material scaling, the question arises whether physical scaling would lead to an
additional conservation law, completing the parallelism of physical and material
transfonnations.

It turns out (Kienzler, 1993) that the transformation

0=0, (2.64 a)

(2.64 b)

and the function K i which will be used in the Bessel-Hagen extension


68 2 Linear Theory of Elasticity

-pE
K; u· (2.64 c)
i-en -i)u I

with
P = const.
3 in three dimensions
and 2 in plane stress
n
2(1 + u) in plane strain
1 in one dimension
leads for p -:;:. 0 to a conserved current
Eu;
(2.65)

in the absence of body forces. The conservation law


Eu.
p -:;:. 0 : ( xj G;j - 1 _ (n ~ 1) u ) ,i = 0 (2.66)

might be interpreted as scaling in physical space. It will be discussed in greater


detail further below.

It is worthwhile mentioning that the six conservation laws given above do not
constitute the only conservation laws derivable from Noether's theorem. As
Edelen (1981) already noted, the restricted notion of variational symmetry (here,
the geometric or point transformations (2.63) are used, but Noether's theorem is
also valid for generalized transformations, where 4Jj and ,;; may depend on
derivatives of uj up to an arbitrary order) precludes any claim of complete
classification of conservation laws. The complete classification of all first order
conservation laws (i. e., conservation laws involving X k , uj and u,) of linear,
homogeneous, isotropic elastostatics in the absence of body forces is given by
Olver (1984). This topic is not persued further here, because the additional
conservation laws seem to be of restricted importance.

We return to (2.59) and wish to lower the constraint concerning the homogeneity
of physical and material space, qj and Jj, respectively. But, when dealing with
material rotations en, we still assume that the material is isotropic and, when
dealing with material similarity m, the material is linearly elastic. The general
discussion of further source terms is given in Section 6.4, where, additionally,
dynamic effects are treated in the context of domain-independent integrals.

If body forces qj and inhomogeneity forces Jj are admitted, the local conservation
laws (2.60) and (2.66) become local balance laws in the form

rj -:;:. 0: G···
1),1 - qj' (2.67 a)

physical mn -:;:. 0: £"kj (xk G;Jj = - £"kjXk qj' (2.67 b)


E ll;
p -:;:. 0: (xj Gij
-

i - (n -1) u/ = - xj qj , (2.67 c)
2.3 Conservation Laws of Linear Elastostatics 69

kj #:0 : bij; - ij , (2.67 d)

material en #:0 : E n1j (xk b;j + Uk a;).; - En1j xkij , (2.67 e)

m #: 0: 2 --u
(xj b jj + - 2
n . a.. ) (2.67 t)
J IJ,j
-xjij'

Integrating over the volume V of a body B under consideration and applying the
divergence theorem leads, with surface S and area A, to balance laws in integral

r
form as follows

Fj = fij n i dA = - dV, (2.68 a)

physical M. ~ 1"9
S

S
x, aij n; dA
B

~- lok
B
j x, qj dV, (2.68 b)

V = fix.rJ a··
IJ
EU
1 - (n - 1) u
i
)n. dA
I
=-
j ;;~j q. dV,
J

S
B (2.68 c)

material

From (2.68), the significance of the balance laws and the correspondence of those
in physical and material space become obvious. Fj (2.68 a) are the resultant
(physical) forces and Mn (2.68 b) are the resultant (physical) moments ("vector
moments") acting en the surface S of the body B. They have to be in equilibrium
with the applied body forces. In the absence of body forces qj' the corresponding
conservation laws

(2.69 a)

Mn = 0 (2.69 b)

lead to the rather trivial statement that the forces applied to the surface S have to
be self-equilibrated. The quantity V (2.68 c) may be called "scalar moment" (or
"expanding moment", "Aiehmoment" in German, or "virial") and should not be
confused with the potential of external forces for which the same symbol is used.
70 2 Linear Theory of Elasticity

The scalar moment (r • q in symbolic notation) is to be contrasted with the usual


vector moment (r x q) as sketched in Figure 2.3.

' - " = = - - - - - - ' - - - - - - - - - - Xl

Fig. 2.3: Scalar moment V = T • q = rq, and vector moment M 3 = r X q = rqq> sketched
in two dimensions

The scalar moment is of much lesser significance than the vector moment, because
it does not enter the equations of equilibrium in physical space. Nevertheless, it has
been thoroughly discussed by Mobius (1837) and Schweins (1849). The
conservation law

v = 0 (2.69 c)

has been introduced quite recently (Kienzler & Kordisch, 1990). By contrast to Pj
and M., V is valid only in linear elastostatics.

Correspondingly, Jj (2.68 d) expresses the resulting (material) forces on the


inhomogeneities inside of S. If the material is homogeneous the right hand side
vanishes and a conservation law, first introduced by Eshelby (1951),

J.) = } .. n·I dA
I)
= 0 (2.69 d)

is established. From this conservation law the celebrated path-independent J


integral (Rice, 1968) and the associated energy-release rates of cracks are derived
which are widely used in fracture mechanics. From L. (2.68 e), the resulting
(material) moments on the inhomogeneities inside of S are evaluated. The
conservation law

Ln = fk (xkb Uk aij) n
j ij + i dA = 0 (2.69 e)

s
requires the material to be homogeneous and isotropic. If the material is not
isotropic, the isotropy condition (2.25) is not satisfied and a tenn
2.3 Conservation Laws of Linear Elastostatics 71

-Ink' J (a"", um. ,+Ja"kn u·J,m) dV


B

has to be added to the right hand side of (2.68 e).

Equation (2.68 f) represents the material virial or material scalar moment. Like the

Pi i
physical virial, the conservation law

M = b, ' 2 ; n u u;j n; dA = 0 (2.69 f)


s
is valid only in linear (though not necessarily isotropic) elasticity.

b.

e.

Fig. 2.4: Material transformations;


a. body with defect in the original configuration
b. defect translated by an infinitesimal amount kj
c. defect rotated by an infinitesimal amount f 3 with respect to 0
d. defect rotated by an infinitesimal amount f c with respect to C
e. defect expanded self-similarly with respect to D
f. defect expanded self-similarly with respect to 0
72 2 Linear Theory of Elasticity

The notations L n integral and M integral are due to Budiansky & Rice (1973), who
showed that their values are related to energy-release rates associated with cavity
rotation and self-similar expansion, respectively.

For clarity, the conservation and balance laws discussed above are rearranged in
Table 2.1, where they are given also in symbolic notation for completeness. In
addition, the three material transformations which yield conservation and balance
laws are depicted in Fig. 2.4.

In closing this Section, it may be mentioned that so-called dual conservation laws
have been established in elastostatics by Bui (1974) and Li (1988), (1992). The
authors apply Noether's theorem in conjunction with the principle of stationary
complementary potential energy. The integrals are expressible in tenns of rates of
decrease of complementary energy associated with cavity or crack translation,
rotation and scaling. By variational principles, upper and lower bounds of the
integrals ], Land M might be obtained. This topic, however, will not be discussed
further in the present text.

2.4 Alternative Derivations of Conservation Laws

In this Section, we discuss briefly several alternative methods to derive conservation


laws in elasticity.

Neutral Action method

The procedure called the Neutral Action (NA) method has already been introduced
in Chapter 1. It is of special value if the system is not governed by a Lagrangian,
but merely by a system of partial differential equations A" (1.108). This procedure
will be used extensively in Chapters 7 and 8 and in the Appendices A and B. Here
we shall apply it to homogeneous isotropic elasticity in the absence of body forces.
The system is described by the homogeneous Navier-Lame equations (2.37)

A.t
= rII u..·
IJJ
+ (A + rIJ) u·JJ' = 0 • (2.70)

The differential equations are to be multiplied by the characteristics /; of the


conservation law. We assume that

(2.71)
Balance law in local form In words Transformation Balance laws in integral form Conservation laws Restrictions

cf. ego (2.67) °ij,i = - qj Balance of physical Translation in Fj = 0 physical


rj *0 linear momentum physical space Fj = J uij n i dA if qj = 0 homogeneity
V' 'u=-q S

E llk} (x k aj,j = - E llk) X k qj Balance of physical Rotation in M" = 0 physical


mn *0 angular momentum physical space M" = J E niy X, uij n i dA if En'j X, % = 0 isotropy
V' • (r x u) = - r x q S

( cu,) _
xj uij - 1 _ (n _ 1) v .i - - x j qj Balance of physical Scaling in V=O physical
p*O expanding momentum physical space V = J (xj uij - Eu' ) n
i dA if xjqj = 0, homogeneity,
1 - (n - 1) v
V' • (r • U - fJu) = - r • q S OJ) = Cijk/ llk,/ linearity

bjj,i J
= - j Balance of material Translation in ~=O material
kj *0 linear momentum material space ~. = J bij n i dA if jj = 0 homogeneity
V"b=-j
S

Ellk) (xk hi} + Uk aj,i = - Ellie) xlcJj Balance of material Rotation in Ln = 0 material
en *0 angular momentum material space Ln = J E niy (x, bij + 1I,u;) n i dA if E niy x,j) = 0, q) = 0, isotropy; no
V' • (r x b + u x a) = - r xj S EIlk) (O,n;" Uk.1Il + ajm llm,J = 0 body forces

(b+
x; if
2 -n
-2~ lljOi" ),j= - xji). Balance of material Scaling in M= 0 linear elastic
m*O expanding momentum material space M = J (xj bij + 2 ; n 11) u;) n i dA if x j jj = 0, qj = 0, material; no
V' • (r • b + au' a) = - r • j S W = 1/2 uij 1I i./ body forces

Table 2.1 Balance and conservation laws of linear elasticity


74 2 Linear Theory of Elasticity

The unknown functions /; have to be determined from (1.110)

(2.72)

with the Euler operator Ek ( ) defined by (1.25). Since

(2.73)

it follows

Ek (f; 4) =( if) (2.74)


Uk,nm ,nm

Performing the differentiations indicated results in

Jl.fkjj + (A + Jl)fjjk = o. (2.75)

Comparison of (2.70) with (2.75) leads to the conclusion that any displacement
field Uk obeying the Navier-Lame equations (2.70) but belonging, possibly, to a
different boundary-value problem may serve as a characteristic for a conservation
law, i. e., we have an infinity of conserved currents parametrized by solutions Uk
(x) of the Navier-Lame equations.

The physical interpretation of this particular case will be given first, before we
proceed to discuss possible solutions of (2.75) in general. Denoting the original
solution of the boundary-value problem under consideration by
(1)
Ui (xj ) = U; (x)

and the characteristic /; = ui by


0)
U; (x) = U; (x) ,

equation (2.73) reads

(2.76)

From (1.109) it follows that /; .1; is a divergence

Integration of (2.76) by parts results in


(1) (1) (1) (2)
F
J;
4i = [(A U
m,m
8·iJ .. + Lt·J,t.»
+ rIJ (Lt t,J u.)
"J
. -

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


- [(AU m.m 8·'J + rIJ . . + u.))
(Lt I.) ),1
Lt.)
t 'J
. +
2.4 Alternative Derivations of Conservation Laws 75

0) 0) (1)
+ [J4 U ...
),11
+ (A +J4) u..
l,lj
.j U"
)

The last bracket vanishes due to (2.75) and Hooke's law (2.22 b) leads to the
conservation law

(2.77)

Integration over a body B with surface S and application of the divergence theorem
leads to the conclusion that the conservation law above may be written as

j h)uji u
0)
j nj
dA -
-
A2) (1)
uji u i n j dA ,
(2,78)

s s
which is just the Betti-Maxwell reciprocal theorem, see, e.g., Barber (1992).

Returning to (2.71) and (2.75), we assume now that f are functions of the
displacements Uno and the displacement gradients uno,n of the actual boundary-value
problem with field equations (2.70). Performing the total differentiation of the
functions f we arrive at an equation involving second and third order derivatives
of the displacement field. Since f themselves depend on derivatives up to the first
order only, the coefficients of higher derivatives have to vanish.

The results indicate that f are linear in displacements and displacement gradients
and we find, similarly to the one-dimensional case (see equation 1.112)

(2.79)

To arrive at the most general form of j; under the constraint condition (2.75) is a
straight-forward but cumbersome task. We, therefore, indicate only the
characteristics f leading to the conservation laws of the preceding section. It may
be mentioned that in addition to Olver's (1984) complete set of first order
conservation laws, i. e., conserved currents involving Ujj uno,n' Ujj uno' u ij and Uj ,
additional conserved currents are found involving lI j uj ,1/U1 which, however, will not
be presented here. We shall now write out only those characteristics, which lead
to the six conservation laws already derived above. If we choose

(2.80 a)

(2.80 b)

(2.80 c)

with kk • en. m. r j mn • p = const., we find, as above, by partial integration the


conserved currents given by (2.60) and (2.66), giving rise to six groups of
conservation laws (four vectorial and two scalar) indicated in (2.69 a - f). Let us
present the details for the special case (2.60 c), choosing kk "# 0 and en' m, r j ,
m n • p = O. With (2.70) the product f: Aj is given either in terms of displacements
76 2 Linear Theory of Elasticity

or by

in tenns of stresses. For convenience, the latter relation is treated further. It follows

(2.81)

Since the material behavior is homogeneous, the strain-energy density W is a


homogeneous function of the displacement gradient. Thus,
oW
W == W (u j ) ~ W,k == at .. Llj,jk' (2.82)
I,)

With the constitutive equation (2.29)

a·· ==--
aw
JI at..
I,)

it follows that

and, finally, with the Eshelby tensor (2.53), the desired conservation law results in

Uj •k anum == - bntlc •m == 0,

which is identical to (2.60 c).

Obviously, the characteristic /; plays the role of an integrating factor and the
conservation law is a "first integral" of the equations of motion.

Remark: The notion "first integral of the equation of motion" is better known from dynamics of
the one-dimensional motion of a point mass m. The equation of motion (Newton's second law) is
given by

m x = F,
where dots denote derivatives with respect 10 time. The potential of external forces V is defined by

F = _ 011'
ax
and the kinetic energy T is given by

T = !.- mi2 •
2
2.4 Alternative Derivations of Conservation Laws 77

If we assume that the mass and the force are independent of time, it follows that

t = mix,
V=-Fi.
Multiplying the equation of motion by the integrating factor i

mix =Fi

it follows immediately that

t = - v,
i. e., the first integral of the equation of motion provides the expression for the conservation of
energy

T + V = const.

Comparing the characteristic /; (2.79) and (2.80) with the characteristic of the
corresponding one-dimensional problem (LIB) one might ask the question, whether
characteristics (or conservation laws) quadratic in xk exist in three-dimensional
elasticity. In fact, such conservation laws do exist (see Olver, 1984) but only under
the unphysical condition

7Jl + 3;' = 0 (2.83 a)

or, with Poisson's ratio u

u =
7 (2.83 b)
8

Application of differential operators

A simple way to derive material balance laws (2.67 d - f) and material conservation
laws (2.60 c - e) in local form is to subject the Lagrangian density L (or its moment
Xi L) to familiar operations from vector calculus such as grad, curl and div. This
procedure was introduced by Golebiewska-Herrmann (1982) and extended by
Eischen & Herrmann (1987) and will be discussed in detail in Chapter 6. Here we
present merely the derivation of the conservation law (2.60 c) for a homogeneous
elastic material in the absence of body forces. We consider the gradient of the
strain-energy density W(u i ) , i. e., grad W c W.i' Refen'ing to (2.82) we have
dW
W·,I = -
du.- U
j,ki
),k

= a kj Uj,ik

Rearrangment leads, with Wi = Wk ;)ki' to


78 2 Linear Theory of Elasticity

(2.84)

Due to the absence of body forces, the right hand side vanishes and, again, we are
left with the conservation law (2.60 c). In a similar way the conservation laws
corresponding to material rotation and material scaling can be derived by
considering curl and div of the vector xjW. This will be outlined in detail in
Chapter 6.

Relationship between the three methods

Three methods of deriving conservation laws in elastostatics have been presented


above and it might be desirable to examine possible relations between them. We
consider only the law of conservation of material momentum which corresponds to
translation in material space.

The Noether transformation functions for this case are

cf>j = 0

leading to the characteristic

and the conservation law

bjj,j = 0 .

In the NA method the characteristic f~ is chosen as

which also corresponds to material translation and leads to the material conservation
law above.

Finally, taking the gradient of the strain-energy density W represents in fact the
operation of translation in material space and leads, as shown, to the same
conservation law. Thus, all three methods imply, each one in its own manner, a
translation in material space.

Similarly one can show that the conservation law based on rotation and self-similar
expansion (or scaling) in material space can be established either by Noether, the
NA method or by evaluating the curl and the div of the vector xjW, respectively.
2.4 Alternative Derivations of Conservation Laws 79

Duality

Finally, the employment of duality as a method to derive conservation laws may be


mentioned. By duality we mean here the correspondence between two processes of
continuum mechanics (here, deformation processes of elasticity), in which the
dependent variables x (current position of material points, physical coordinates) are
interchanged with the independent variables X (reference position of material points,
material coordinates), see Fig. 2.1. In this consideration it is not advisable to
introduce the displacements u as independent variables

Uj = xj - Xj'

since this field and its derivatives depend on both material and physical coordinates.
As pointed out by Rosel (1986), a dual process is not the same process run
backwards that one should consider but the following: there is another body (with
different constitutive equations), which has as a reference configuration the current
configuration of the original body, and deforms into the original configuration of the
latter.

Duality under "inverse defonnation" was discussed by Shield (1967). Chadwick


(1975) revealed the connection to the Eshelby tensor and showed that this tensor
and the Cauchy-stress tensor are dual quantities. By consequent use of physical and
material coordinates as dependent and independent variables, Golebiewska-
Herrmann (1981) established duality in elastodynamics.
3

Properties of the Eshelby Tensor

3.1 General Remarks

One peculiarity associated with the Eshelby tensor bjj (cf. equation 2.53) is that it is
used primarily as an integrand in global representation of path-independent integrals,
such as the j, Land M integrals. Regarding the tensor itself, we know essentially
only, for the time being, that it is not symmetric and that its divergence vanishes if
the material is locally homogeneous. In this Section, we will provide a physical
interpretation of its Cartesian components and show what its first and second
indices indicate.

Furthermore, we will discuss the quantities which are usually investigated when
dealing with tensors, such as invariants, principal values, principal directions, etc.
Surprisingly enough, this subject has been treated only very recently (Kienzler &
Herrmann, 1997).

As has been shown in the previous chapters, the Eshelby tensor bij given by

(3.1)

(in the absence of body forces, i. e., V = 0 ) represents a second rank tensor field,
i. e., a tensor-valued function of position. For completeness, it may be mentioned
that there exists another tensor of fourth rank with constant components, which is
also sometimes referred to as the Eshelby tensor. This latter quantity is useful in
studying inclusion problems, cf. e. g., Mura (1991), and its properties are well
known. Both Eshelby tensors have been discussed in the same paper by Herrmann
(1993). As before, we are concerned exclusively with the second rank tensor
(3.1).
82 3. Properties of the Eshelby Tensor

3.2 Physical Interpretation of the Components


of the Eshelby Tensor

For the sake of simplicity and clarity, the following considerations are developed in
two dimensions. The arguments, however, apply equally well in a three-dimensional
setting.

In order to provide a physical interpretation for the components of the Eshelby


tensor, we consider a body B of volume V. We focus attention on an arbitrary
differential volume dV of unit width and area dX I dX 2 in the undeformed state and,
additionally, on a neighbouring element which, also in the undeformed state, is
located at an infinitesimal distance A, along the XI -axis from the original one. The
body is now deformed by some surface tractions and the deformed state is
characterized by the stress field ui (x k ), the displacement field u i (xJ and the
strain-energy density W(xJ. The tot~l energy JI 101 of the body consits of the sum of
the work which the surface tractions did along surface displacements called JIO and
the total internal energy

If ~ /W(xJ dV ,
B
that is

The energies JI" and JI j do not need to be made specific.

The deformed state of the original element mentioned above will be called state ®
while the deformed state of the neighbouring element will be called state @.

Keeping in mind our goal, namely the interpretation of the tensor components bi) we
wish to calculate the change f:.JItot of the total energy JItot of the body, as the
original volume (or area) element dX I dX2 is displaced within the material from the
deformed state ® to the deformed state @. Even though we seek to calculate the
change f:.JItot for the whole body (global consideration), it turns out that we can
obtain the desired result by examining energy changes associated with the element
only (local considerations), without ever having to perform any integrations over the
whole volume V of the body.

For this purpose, we assume that the stress field on one hand and the displacement
field, as well as the strain-energy density on the other hand, are separated from each
other, as if they were independent fields. Figure 3.1 shows the element dX I dX 2 with
the stresses acting across its faces (no body forces are applied), and Figure 3.2
shows the "independent" displacements of the four comers of the element in state ®
(upper part) and state @ (lower part).
3.2 Physical Interpretation of the Components of the Eshelby Tensor 83

t O'zz + O'zz,z dxz

- - -..... Oil + Oil, zdxz


r - - - - - - - - - - - - - , - - -- - - -- - - - -- - - -- ---

Oiz + OiZ,1 dxl

f- "" + "",,<ix, <ix,

,r-----------------------------~-------------------
,
, Oil .........- - - ,
, ,
, ,
, ,
, ,
, ,
, ,
, ,
, ,
'" ... '
dxl
Fig. 3.1: Infmitesimal element in the (Xl' xz) - plane with stresses across its faces

State @

,
'..
..... .--------------1..~,: u,(xl + dxl> xzJ
: dx J

-' '-
,
, ,
,

Lx,
Xz
dx2 State

U,(X I + AI' xz) :,........-- ......,: u,(xl + dx l +A» xz)


dxl
Fig. 3.2: Infinitesimal elements in st:lte ® and state ®
84 3 Properties of the Eshelby Tensor

The stresses in Figure 3.1 are now considered as "applied" and thus constant in the
passage from ® to @. Such a transfonnation shall be called material translation
and is illustrated in Figure 3.3.

To calculate the energy change I::.II tot in this transformation we proceed


sequentially by considering that is occuring to the four forces CD, ~, @, ® of the
element dx. dX 2 (Fig. 3.3).

First we consider the face CD (actually a surface of unit width) and calculate the
energy change due to transfonning it into position XI +A I . Internal energy is
removed from the element in the amount

(3.2)

This change of internal energy is indicated in Figure 3.3 by the hatched area.

rT"T//7~~~---<!f,~~~~~..... = .. ~ ....
I
I--@
I
3 I
I
I

L,
X2 I

I. -_
J.
I AI _: 2
I
I

' .. ~'
dx]
Fig. 3.3: Transition from state ® (-) to slate @ (-----)
The work that the forces aJl dxl and all dx] had already done to get into the
defonnation state u, (XI' Xl)' U2 (XI' Xl) is changed by the translation AI' They are
now at a location characterized by the displacements u, (x, + A,. x 2). U2 (x, + A,. Xl)'
The additional work (it is negative) will increase the energy IId by an amount

(3.3)

A factor 1/2 does not appear, since the state of stress is, as mentioned, constant
in this transfonnation. The change of total energy II 10' by the material translation
of face CD by the amount A, is
3.2 Physical Interpretation of the Components of the Eshelby Tensor 85

and in view of equation (3.1)

(3.4)

This quantity blJ is indicated in Figure 3.4 by an arrow at face CD

The translation of face @ does not cause any change in the internal energy JIi
The change of external work, however, can be calculated leading to a change of
total energy as

(3.5)

The translation of face @ causes the following changes

Hence

(3.6)

Finally, the translation of face ® leads, analogously, to

(3.7)

The contributions of all four faces due to the transformation A/ are indicated in
Figure 3.4, together with the contributions due to a transfonnation A2 in x2 -
direction, calculated similarly.

If the element is not oriented along the XI' Xl - axis, the faces dX 2 and dX 1 should
be replaced by the projection of the element of area (of unit width) dA in x/ - and
x2 - direction, respectively, i. e., by n/ dA and nl dA, where n is the the unit
outward normal vector (see Figure 1.2).

The physical interpretation of the components b;j is thus given:

The component b;j of the Eshelby tensor represents the change of total
energy of a deformed body produced by a material unit translation ( Aj = 1)
in the x j - direction of an oriented unit element of area (n i dA = 1) projected
in the Xi - direction. This change of energy can be calculated as described
above.
86 3 Properties of the Eshelby Tensor

It now also becomes clear why bij is not a symmetric tensor. We have shown that
b12 is the work done by tractions ali n l in the displacement gradients u i ,2, while b21
is the work done by tractions a2i n2 in the displacement gradients Ui,j • In a general
deformation there is no reason why b12 should equal b2j •

t b22 + b22,2 dx2

~
- - - - I.. b2! + b21,2 dx2
.---------,-----------,--- - - - - - - - - - - - - - - - - --

!
4 b12 + b I2,! dx]

bu _
1
3 l-bU+b'U dx
'
b12 2

I
I
I
b2! ..

~
I

L,
I
X2 I
I
I
I
b22
I
1..- ~ I

dx]
Fig. 3.4: Components of the Eshelby tensor at an infinitesimal element

The "material equilibrium conditions" are read immediately from Figure 3.4 as

bji,j = 0 . (3.8)

They indicate that the change of the total energy of a body of volume V due to a
material translation of an element of area (or volume) is zero in a homogeneous
material.

3.3 Invariants, Principal Values, Principal Directions and


Extremal Values of the Eshelby Tensor

Invariants

Since b ij is a second-rank tensor, the usual three invariants exist in three dimensions,
namely the trace b I}, the square trace b 12 and the determinant b 13
3.3 Invariants, Principal Values, etc. of the Eshelby Tensor 87

(3.9 a)

(3.9 b)

hl3 = det (bj) • (3.9 c)

The three eigenvalues or principal values (rl' r", rJlJ) are known to be the roots of
the characteristic equation

(3.10)

Since the Eshelby tensor is not symmetric in general

(3.11)

two of the principal values might be complex and the principal directions are not
necessarily mutually perpendicular, nor necessarily real.

A closed form solution of (3.10) does not exist and, therefore, further discussion
of the general three-dimensional (3D) case is not possible.

Thus we restrict ourselves to two dimensions (2D), a situation that can be


investigated in detail. This restriction is not as serious as it seems, since we are
concerned primarily with rotations of coordinate systems, which is essentially a
plane problem. The only difference between 3D and 2D is that in 2D the axis of
rotation is made to coincide with the x3 axis.

The strain-energy density W in the linear theory of elasticity (cf. Chapter 2) is

1 (3.12)
W = -2 a··u·
IJ J.I·

Thus the components of the Eshelby tensor (3.1) are

b" =W - a" uJ.J - an u2.1

1 1 1 1
-
:2 a" uJ.J
-
:2 an u2.1 + :2 a 21 uJ.2 + -
2
an u 2.2' (3.13 a)

bn - a" uJ.2 - an u 2.2 ' (3.13 b)

b2J - a 21 uJ.J - an u 2.J ' (3.13 c)

bn =W - a 21 uJ.2 - an U2.2

1 1 1 1
+ :2 a" uJ.J + -2 aI'--.
U'I -2 a'l
- UI .-' -
2
an u 2.2· (3.13 d)

It follows immediately that the first invariant of bj ; vanishes identically


88 3 Properties of Ihe Eshelby Tensor

(3.14)

Thus the Eshelby tensor in two dimensions is purely deviatoric. The second
invariant is

(3.15)

We wish next to express the second invariant of the Eshelby tensor in terms of
commonly used invariants of the theory of elasticity. We select the invariants of
stress rather than strain, because boundary-value problems in two dimensions are
most effectively formulated in terms of stresses. The rotation tensor, however,
will necessarily also be involved, because bij depends on the components of the
displacement gradient rather than on those of the strain tensor. From the
mathematical point of view this fact is obvious; the fact itself, however, is rather
astonishing since the rotation plays quite an insignificant role in linear elasticity.

The rotation tensor (2.4) in 20 is characterized by one quantity, namely its


invariant UJ3 = UJ (cf. equation 2.18)
1
UJ = 2 (u2.J - ud . (3.16)

Replacing the components of the displacement gradient In (3.13) by the


components of strain and the rotation

(3.17 a)

(3.17 b)

(3.17 c)

(3.17 d)

and using Hooke's law in the fonn (2.39 b)

1 K-3 a,.,. 8.. 1


E·· = -
I) 2G (a..'J + --
4 1<1< iy

yields, after some algebraic manipulations

(3.18)

With (2.24 b) and (2.38) the constant E . is given by

E
for plane strain (3.19)

for plane stress.


3.3 Invariants, Principal Values, etc. of the Eshclhy Tensor 89

If bij is decomposed into its symmetric and antisymmetric parts, i. e.,

bij = 2"1 (bji + bi) + 2"1 (bji ~ bi) I' b(j + A bjj , (3.20)

then the first invariants of I' bjj and A bij of course vanish and the second invariants
hI' 12 and bA 12
are calculated to be

hSI2 = - _1_
(2£ ')2 r~I/ - 4 al2 J!I'laI/ + (£ 'UJ/J ' (3.21 a)

bA 12 = 1_
+ __ al/ (£ 'UJ)2 (3.21 b)
(2£ '/
and, as expected, one finds

(3.22)

Principal values

The characteristic equation of the 20 eigenvalue problem is given by

(3.23)

Since hII vanishes (3.14), the roots of (3.23) are with (3.18)

Y/J/ = +

U/
I
,.l a2 (3.24)

and the principal directions are given by

(3.25)

It is noticed that the two eigenvalues YI and YJI may be either real or imaginary.
Assume, e, g., a material point under a hydrostatic state of stress (all = an = a,
2
a l2 = 0; all = 2a, aI2 = a ), the eigenvalues are imaginary, Yl,ll = ± iUJa. On the
other hand, assume a material point under simple shear (all = a, an = ~ a, a l2 =
0; a II = 0, a 12 = ~ a 2 ), the eigenvalues are real, Yl,ll = ± UJa.

The occurence of complex eigenvalues and complex eigenvectors has been


thoroughly discussed in the context of qualitative behavior of ordinary differential
equations (e. g., Hubbard & West, 1995). Here, imaginary eigenvalues appear to
90 3 Properties of the Eshelby Tensor

indicate that cross-sections with vanishing energy change due to material translation
in tangential direction do not exist. We will comment on this point further below.

A detailed discussion of the eigenvalues (3.24) for a general state of stress is to be


found in Kienzler & Herrmann (1997).

Material traction vector

The basic Cauchy relation of the theory of elasticity (2.12)

relates the traction vector !i to the unit nonnal vector nj characterizing the
orientation of the element of area and thus defining the stress uji (physical
momentum). It will now be applied to the Eshelby tensor bji (material momentum).

The significant difference between the Cauchy relation and the one below is that we
define the material traction vector q i through the Eshelby tensor rather than vice
versa

q j = bjj nj • (3.26)

The components nj may be expressed in the usual way in tenns of an angle 1ft as
indicated in Figure 3.5.
Xz
G1 ---------- G
n In

Fig. 3.5: Maleri;ll traction vector Q on a cross-section with unit normal vector n

With n/ = cos 1ft and n2 = sin 1ft it follows

q/ = bi! nj = bll cos 1ft + b2/ sin 1ft, (3.27 a)

(3.27 b)
3.3 Invariants, Principal Values, etc. of the Eshelby Tensor 91

The usual transfonnation of the material traction vector with components 1 and 2 q q
to the components in the nonnal and tangential directions called bnn and bn•
respectively, are

l (b 12 + b2/) sin 2 rft


2
(3.28 a)
1 1 1
bnJ = - (b 12 - b2/) + - (b 12 + b21 ) cos 2 rft - - (b Jl - bn ) sin 2 rft
2 2 2
(3.28 b)

Extremal values

As for the Cauchy stress tensor, we wish to explore several special directions
associated with extremum values of bnn , bn" and the magnitude of the material
traction vector IGI , given by the angles 1ft" ' rft' and rftG , respectively.

The calculation leads to

(3.29 a)

with the associated quantities

(3.29 b)

(3.29 c)

'TT'
+- (3.29 d)
2

(3.30 a)

with the associated quantities

h,:' = 0 , (3.30 b)

'TT'
± (3.30 c)
4'
92 3 Properties of the Eshelby Tensor

extr
IGI = max Ib:,rl (3.31 a)

with the associated quantity

G I
rJlj,2 = rJlJ.2 . (3.31 b)

It is noticed that (in contradistinction to the properties of the Cauchy tensor) the
normal component b;:"r is associated with a non-vanishing tangential component
bna:, which is obviously due to b12 :f. bu. Further, it is noticed that the maximum
lr
absolute value of b:: , namely max Ib,~/rl is always larger (or equal in the special
1r
case b12 = b2/ ) than b::. and, in addition, the associated component bna:. is
vanishing (because b ll = - bn always). The maximum value of the magnitude of
the material traction vector IG l",'r turns out to be just equal to the maximum
value of b ::Ir.
., bnn'ex,r bas d b",tr . .
Th e quantities 111 an III are invarIant an d can be ca Icu Iated'
, Just as b / 2'
in terms of the rotation UJ and the stress invariants 0/1 and 0/2

b exl, ± _1_ II "I} - 4,,12 / I "I} + (£ 'UJ)2 / (3.32 a)


'III
2£'

b,':: =_1_,,1 (£'UJ),


2£' j
(3.32 b)

b exlr = b exlr +
bas (3.32 c)
"'
lUI lit .

Comparison of (3.32) with (3.21) leads to

b ex"
'III ± 1- bsI 2 '
(3.33)

b,':: l+bAl2 . (3.34)

Two values b;:"r I and b;:''' 2 exist, one is pOSItiVe, the other one negative but
equal in magnitude. Only one quantity bJ~s, however, exists. Its value might be
either positive or negative, depending on the sign of "I} and UJ (see 3.32 b).
These relations can be conveniently discussed in terms analogous to Mohr's circle
(e. g., Timoshenko & Goodier, 1970). Elimination of the transfonnation angle rJI
from (3.28) and use of (3.14) leads to

(3.35)

with
1
a =- (b}2 - b21 ) (I:J; = bll~s (3.36)
2
and

(3.37)
3.3 Invariants, Principal Values, etc. of the Eshelby Tensor 93

Thus in the (b nn , bllt ) Cartesian plane, any pair bl/II ' bllt is located on a circle with
radius s = 1- bS/ and center at a = 'l'2 (b 12 - b21 ) on the bnt axis (Figure 3.6).
2

All the quantities of interest, such as the individual components of bij , its extremal
and associated values, the traction vectors as well as all relevant angles, can now be
read off Figure 3.6. The eigenvalues Yw are located on the intersection points of the
circle with the abscissa. Therefore, real eigenvalues exist only if the radius s is not
smaller than a. If they exist, the corresponding cross- sections are, in general, not
mutually perpendicular. Further discussion and examples may be found in Kienzler
& Herrmann (1997).

bmin
-""nn--,---t-...;.----------::;;/4==----~L.....::::..L;l.L.C\__n'----____r--.bnn

maxlbntl
max
=IGl
=Ial +s

Fig. 3.6: Circle representation of the Eshelby tensor

In conclusion, it may be mentioned, that there always exists an orientation of an


element of area given by ljJ t whose translation in tangential direction leads to a
maximum energy release. Therefore, h,~\tr will be the important quantity when the
influence of boundaries and interfaces on energy changes are studied. Employing
the example of a stretched elastic finite sheet, GoJebiewska-Herrmann & Herrmann
(1983) have shown that the material traction can be interpreted as the rate of total
energy change as a function of material addition along the sheet edge. Consideration
of material tractions were also essential in studying the stability of plane boundaries
94 3 Properties of the Eshelby Tensor

in stressed bodies as discussed by Gao (1991 and 1994) and Freund (1995). Further,
in studying the optimum shape of inclusions associated, again, with the minimum of
total energy, material tractions had to be used by Schmidt & Gross (1995 and
1997).
4

Linear Elasticity with Defects

4.1 General Remarks

In this Chapter, we shall consider a linearly elastic body with defects in the fonn of,
e. g., dislocations, voids, inclusions or cracks, see Figure 4.1.

o
.. F

Fig. 4.1: Elastic body with defects

To evaluate material forces "acting" on such defects, conservation laws established


in Chapter 2 in global fonn will be applied in Section 4.2; this will serve the
purpose of calculating energy-release rates associated with translation, rotation and
self-similar expansion of a defect. An example of such a calculation is presented in
Section 4.3 which involves a circular hole in the presence of one or two
dislocations. Path-independent integrals of fracture mechanics are discussed in
Section 4.4. The universal singular fields around crack tips, crack-extension modes
and free-body diagrams in material space are also presented.
96 4 Linear Elasticity with Defects

4.2 Path-Independent Integrals and Energy-Release Rates

For a linearly elastic material, homogeneous in physical and material space, material
conservation laws in local fonn have been derived in Section 2.2 as

b..IJ , =0, (4.1 a)

= 0, (4.1 b)
2 -n
(xj bij + - 2 - uj Vi)'i = 0. (4.1 c)

Integration of these local conservation laws over a body B with volume V, surface
S of area A and unit outward normal vector n leads to the corresponding
conservation laws in integral or global f0I111 as indicated in equations (2.69) and
(2.68)

fijn;dA = 0, (4.2 a)

s
f'k j (xk bij + Uk Vi) n i dA =0, (4.2 b)

j ((x. b ..
S
I"') I)
+ 2 -n u. v). n dA
2 ) I I
= o. (4.2 c)

These relations for perfect materials are helpful in establishing path-independence


of the corresponding integrals in the presence of defects inside the body under
consideration. The right-hand sides of the integrals, in general, will no longer
vanish. We already introduced the notation J. Land M integral (2.68) as

=
j 1;. n V,) ,
dA , (4.3 a)

o f"j (x, b'j 'u, a;) n, dA (4.3 b)

M =
f
s

S
(x.) b.I) + --
2-n
2 u.) v.)
I)
n 1 dA (4.3 c)

Consider a body containing a defect that is represented in Figure (4.2) by a black


spot. This defect may be regarded as an inclusion, cavity, dislocation or another
material inhomogeneity.

For the surface shown in Figure 4.2, the homogeneous relations J i = 0 ,Ln = 0
and M = 0 still hold, because there is no defect inside the indicated surface
4.2 Path-Independent Integrals and Energy-Release Rates 97

fi )dA=O.

-
Fig. 4.2: Contour for path-independent integrals in the presence of a defect.

The contributions along the parallel parts of the surface cancel each other, because
the integrands are smooth, regular functions along 53 and 54 . By reversing the path
of integration along 52 we can conclude that the quantities lj' L n and M evaluated
along surfaces close to the defect (52) will be equal to those evaluated along
arbitrary surfaces far from the defect (Sl)' as long as the surfaces enclose the defect
and no other defects are included.

Having established path-independence of the integrals J, Land M, any convenient


path surrounding the defect can be chosen in evaluating them.

We would like now to establish the meaning of the J integral as an energy-release


rate for a material translation of a defect. Already in his pioneering work concerning
forces on elastic singularities, Eshelby (1951) showed that the value of a path-
independent integral surrounding a defect can be interpreted as the change in the
total energy associated with unit defect translation. He elaborated further on the
subject in later papers (Eshelby, 1970, 1975 a, 1975 b). This topic was also
considered by Freund (1972), Budiansky & Rice (1973), Buggisch et al. (1981),
Eischen & Herrmann (1987) and KienzleI' (1993) using arguments different from
those presented by Eshelby.

Fig. 4.3: Original body (left) and replica (right) with defect
98 4 Linear Elasticity with Defects

In this Section, we will follow the physical interpretation of Eshelby (1975 a). He
suggests the following thought experiment. The picture on the left of Figure 4.3
shows an elastic body containing a singularity which is enclosed by an arbitrary
surface S. The body is subjected to mechanical tractions t. We shall call this body
the original system. The picture on the right represents an exact replica of the
original body with Sand SI marked out. It is seen that SI is the surface S that has
suffered a vector shift - ~Ak in the undeformed state.

If we replace the material inside SI of the replica by the material inside S of the
original body in the deformed state, we obtain a configuration that differs from the
original one by a vector shift + ~ Ak of the defect. The passage from S to SI in the
defonned state is nothing else but a material translation that has been discussed in
detail in Section 3.2 for an infinitesimal element in a homogeneous body. For
perspicuity, we repeat some of the arguments and evaluate the change of energy of
the system due to a material translation of the defect step by step as follows: After
both bodies come to an equilibrium state under the mechanical loading, we cut out
the material enclosed by S in the original system and throw it away. Then we apply
equivalent surface tractions on the resulting hole to prevent the rest of the original
system from relaxing. We then cut out the material inside SI from the replica and
maintain, now, the tractions on the surface of the cut-out piece to prevent
relaxation. The internal energy II;; inside SI differs from the internal energy II}
inside S of the original system by the addition of energy in the crescent-shaped
region 1 and the subtraction of energy in the crescent-shaped region 2. This
difference can be calculated by expanding the internal energy about II},
dII/
II;; = II} - __s ~Ak + 0 (~Ak/
dx k

(4.4)

Here the divergence theorem is used only in the material region I and 2 where
there is no mathematical singularity. The strain-energy density W or its derivative
may be infinite in the region occupied by the defects, but this region is subtracted
out and thus does not contribute to the energy change.

So far no energy has been changed in the remaining material or the loading
mechanism of the original system. We shall now try to fit SI into the hole S of
the original system. The boundaries of Sand SI would coincide in the natural
state because one is a mere vector shift of another. But once the deformation has
taken place the displacements on SI will differ from those on S by

(4.5)
4.2 Path-Independent Integrals and Energy-Release Rates 99

In order to fit Sl into the hole, we have to supply a displacement of an amount


given by expression (4.5) on the surface of the hole. This requires work which will

f
increase the energy of the original system by the amount

~n D
=- u j ajj nj dA + 0 (~Ak)2

= '6 A,
S

fft
s
u" nj dA ,0 (6.1,)2. (4.6)

During this process, the surface tractions change according to

(4.7)

whose extra terms contribute an amount of order (~Aj to the energy change.

We can now fit SI into S and weld across the interface. Though the displacements
are now continuous across the interface, the stresses are not. They differ by an
order of ~Ak' As we relax these quantities, the displacements would be of order
~Ak and an amount of energy of order (8Ai is extracted, which can be ignored in
comparison with (4.4) and (4.6).

We have now finished the task of moving the defect in the original body by
+ ~Ak' The total energy change ~Irm = ~Jf + ~na during this procedure is

~nlot = ~Ak fa jj llj,k - W~jkJ n


j dA .
s
As outlined in the Introduction, we define the force on the defect as the negative
gradient of the total energy of the body B with respect to the change in position of

Jk = -
IlAk
[1m
.
-> 0
~ =
all. k
f
the defect within the body, this force can be expressed as
~ntot [W ~jk - aji lliJJ n j dA (4.8)
S

which is, because of (2.53) identical to (4.3 a). Thus the path-independent integral
i k gives the value of the total energy-release rate.

As we apply the displacements ~Ui (4.5) to S, the tractions across S change


according to (4.7). The extra term will depend on the way the forces exerted by
the loading mechanism vary as their points of application move. However, as the
extra term in (4.7) only alters (4.6) by a quantity (l>A k / , the external forces do
not enter the final result explicitly.

In a similar way it may be shown that the L II integral represents the change of
energy of the system due to an infinitesimal rotation (Oil of the defect with
respect to the origin of the coordinate system
tot
L o = - iJn
-- (4.9)
n iJwn
100 4 Linear Ela~licily with Defects

and the M integral characterizes the energy-release rate due to a self-similar


expansion

Xj~aXj

of the defect, such as a cavity or crack


OlllOI
if =-- (4.10)
oa
(cf. Budiansky & Rice, 1973; Eischen & Herrmann, 1987).

Change of reference point for the L n and M integrals leads to (Bakker, 1984)

M = MO - xp J;, (4.11 a)

(4.11 b)

where xp are the coordinates of the new, arbitrarily fixed point of reference.
Analogous equations for changing the reference point exist for a system of
physical forces as known in elementary statics.

4.3 Example: Hole-Dislocation Interaction

As an example of the application of path-independent integrals we consider the


interaction of a circular hole and an edge dislocation in plane elastostatics. Let the
edge dislocation lie at the origin of a Cartesian coordinate system (XI' x2) and have
Burgers vector components b l and b2 • The stress field produced by this dislocation
(cf. Volterra, 1907) is given as
(0) 2G 2 2
all (- bj x2 (3x} + x2 ) + b2 Xj (Xl - X;) ) ' (4.12 a)
4
r 7T (K + 1)

(0) 2G 2 2
an /+blx l (X} - x2 ) + b2 Xl (Xl + 3x;) ) ' (4.12 b)
r 4 7T(K+l)

(0) 2G /
an + b j Xl (XI2 - X22) + b2 X2 (Xl2 + X;) ) ' (4.12 c)
4
r 7T(K+l)

r = Ix} +x;. (4.12 d)

The shear modulus is G and K is given by equation (2.38). This state of stress is
altered due to the presence of a circular hole (radius 1"0) centered at a distance d + 1"0
from the origin on the XI -axis (Figure 4.4). KienzleI' & Duan (1987) derived a
4.3 Example: Hole-Dislocation Interaction 101

simple fonnula to obtain the hoop stresses at the boundary of a stress-free circular
hole based on the stress distribution that would exist along the boundary of the hole
in its absence. (It may be mentioned that the general background of this formula is
based on the involution correspondence in plane elastostatics for regions bounded
by a circle, as outlined by Honein & Herrmann (1988).)

Let a~) (ro ,(jJ) and a~~ (ro ' (jJ) be the stresses calculated from (4.12) along the
curve coinciding with the boundlU)' of the hole (r, (jJ polar coordinates with respect
to the center of the hole) and .J}O) the value of the first stress invariant at the center
of the prospective hole (here uI}O) :: (a~~ + ai~) / Xl :d f,O ).
X2 : °
The hoop stress a rprp ( (jJ) at the boundary of the hole due to the applied load (in this
case the eigen-stress field of the dislocation) follows from

( m\::
a rprp"-/ u
/(0)
1
+ 2 ('l.a (O)
rprp
(I'0'''-/
m\
--
m\)
arr(0) ( ro ' .,-/ . (4.13)

-----.JII'"----+--~----L!+---_Xl

..

Fig. 4.4: Interaction between circular hole and dislocation.

Introducing the dimensionless distance

r :: d/ro (4.14)

it follows from (4.13) with (4.12) after standard coordinate transfonnations and
suitable rearrangements (cf. Kienzler & Kordisch, 1990)

+ b
2
/ --
1
I
+ Y
- 4 (1 + y) sin
{2 (l
2
(jJ {I + (1 + y) cos (jJJ
+ y) (l + cos (jJ) + rY
J). (4.15)
102 4 Linear Elasticity with Defects

We wish to calculate the material force exerted by the dislocation on the cavity
(peach-Kohler force, Peach & Kohler, 1950) or, vice versa, the force exerted by the
cavity on the dislocation. It appears much simpler to evaluate the latter force by
means of a path-independent integral choosing the contour along the rim of the
hole. The only nonvanishing stress component is U rprp and the strain energy W is
given with (2.39) and (2.40) by
K+1 2 1 2
W = - - U pp (rp) = - UPf/J (rp).
16 G 2E'
As in the previous Chapter, the constant E ' is defined as

E'=~=j1~U2
1 + K
for plane strain
(4.16)
E for plane stress.
The integrals (4.3) reduce to

2...

U~9"
'0
J} / (rp) cos rp d rp, (4.17 a)
2E' 0
2...

U~p
'0
J2 / (rp) sin rp d rp. (4.17 b)
2E' 0
2 2...
NfJ '0 / a~f/J(rp)drp, (4.17 c)
2E' 0
2 2...

a~9"
'0
L'" =- (1 + r) / (rp) sin rp d rp. (4.17 d)
2E' 0

such that from the knowledge of U rprp( rp) (4.15) the integrals can readily be
evaluated.

The reference point for !vf has been chosen to be the center of the hole, while the
reference point for L J = L'" is the origin of the coordinate system. Land M for
an arbitrary point of reference are given by the transformation (4.11)

M =!vf -x/ J; • (4.18 a)

L = L'" - EJ'1·· x·'"


,
J) ' (4.18 b)

with xp and x/ the distance from the respective reference point. The evaluation of
the integrals (4.17) with (4.15) is cumbersome but straightforward. The results are

E' (b; ( 1 + 1 )+ bi 1 )
4'oTr (1 + r) I' r (2 + r) (l + r/ r (2 + r) ,
(4.19 a)
4.3 Example: Hole-Dislocation Interaction 103

E* b] b2
12 =+ (4.19 b)
4rO 7T (1 + y) (l + yyJ

MJ - '0 (1 + y) II , (4.19 c)

Lr + '0 (1 + y) 12 • (4.19 d)

The equivalence of the relations between MJ and II (4.19 c) on one hand and
between Lr and 1 2 (4.19 d) on the other is remarkable. By choosing x; = d +
'0 and x; = 0 in (4.18 b), it turns out that LO referred to the center of the circular
hole vanishes, i. e., the lines of action of the material forces II and 1 2 intersect
each other at the center of the hole.

In order to discuss the integrals in tenns of energy-release rates with respect to


the change of the position or the size of the hole relative to the dislocation, it is
stated that the total energy of our system depends - geometrically - only on two
parameters, y and rJ. With

b] = b cos rJ, (4.20 a)

b2 = b sin rJ, (4.20 b)

b = /b} + b; is the intensity of the inclined edge dislocation and rJ is the angle
of inclination (cf. Figure 4.4). Consider now an infinitesimal transfonnation of the
hole with respect to the coordinate system consisting of four parts: a translation ~ I
in the xrdirection, a translation ~2 in the x2-direction, a rotation UJ about the origin
and a self-similar "expansion" ~ (l' '0 '0.
The geometrical parameters change to

y ~ y + ~ 1/'0 - (l' (1 + y), (4.21 a)

rJ ~ rJ - ~2 /['0 (1 + y)J - UJ • (4.21 b)

The path-independent integrals are related to energy-release rates via (4.8), (4.9)
and (4.10) as
iJlI lot iJlI tot
iJy iJlI tot arJ 1 iJlI tot
1=---
] iJ8
- -- --- (4.22 a)
] iJy iJ8] iJrJ iJ8] '0 iJy

iJlI iJy
lot
iJlI tot arJ 1 iJlI tot
- -- = +----
iJy iJ8 2 arJ iJ8 2 '0
(l + y) iJrJ
(4.22 b)

iJlI tot all tot


M - -- = + (1 + y) (4.22 c)
iJ(l' iJy

(4.22 d)
104 4 Linear Elasticity with Defects

From (4.22 a) and (4.22 c) the relation (4.19 c) is rederived and (4.22 b) with
(4.22 d) confirm Eqn. (4.19 d). Integration of (4.22 a) and (4.22 d) leads to the total
energy II'o l of our system up to an insignificant constant of integration

II'OI =
7r
Gb
2
(K + 1)
fin /r (2 +r)
(l + rl
J- cos 2 ()
(l
1
+ rl
)+ C. (4.23)

Equation (4.23) is in agreement with a fonnula given by Dundurs & Mura (1964),
who studied the interaction between an edge dislocation and a circular inclusion.
They showed that under certain circumstances (stiffness ratio of inclusion and
matrix material, distance between inclusion and dislocation, strength of the
inclusion) there exists a position of equilibrium (1/ = 12 = 0) between inclusion and
dislocation. From equation (4.19 a) it is obvious that such an equilibrium position
will never exist (except when r tends to infinity) between a hole and a dislocation.
The force between the dislocation and the hole is always a force of attraction. Thus,
if the hole or the dislocation could move, the distance between them would
decrease.

We would like to consider next the shape of trajectories envisaging the possibility
that the cavity, by means of some mechanism (e. g., diffusion, cf. Stark, 1976),
could move towards the dislocation. Lacking equations of motion (of the type of
Newton's second law for mass points, where, in general, the force is not tangential
to the trajectory), it is usually assumed that the material force is tangential to the
trajectory, i. e., path of motion.

1,

Fig. 4.5: Interaction between circular hole and dislocation. polar coordinates

We assume that the dislocation is fixed at the origin of the coordinate system and
the angle () is (without loss of generality) set () = 7r /2. The center of a circular
cavity of radius ro is placed at an arbitrary point x,. x2 , thus at distance r = + /x/ xi
from the dislocation. By means of equation (4.19 a and b), the material forces 1,
4,3 Example: Hole-Dislocation Interaction 105

and J2 can be calculated. In tenns of a polar coordinate system T, qJ, see Figure 4.5,
and of the dimensionless distance p between dislocation and center of the cavity
T
P = - = (l + y), P > 1,
TO

the material forces in radial and tangential directions are given by

J T
= -
E' b
41r rop
2
I~ 1
- 1
+ Sin~ qJ J' (4.24 a)

2 s/'n 2 m
+ E •b ---L (4.24 b)
Jip
81r To P ~

The trajectories of possible motion of the cavity have to be determined from the
differential equation
dx2 = J2
dx 1 J
1
'

or, in polar coordinates


dp J
- = p -T
dqJ Jrp
If the distance between dislocation and hole is large, i. e., p Y> 1, the differential
equation can readily be integrated, leading to
sin qJo cos 2 qJ
P = Po - - - - (4.25)
cos 2 qJo sin qJ

Each trajectory is specified by the choice of <pI) and pI) . Some of the trajectories,
with the restriction p Y> 1, are sketched in Figure 4.6.

~-+---+---t----+---+-- Xl
To
1 2 3 4 5

Fig. 4.6: Trajectories of motion of a cavily in a slress field due 10 a dislocation

Cherepanov (1984) investigated a related problem concerning the interaction


between a spherical cavity and a dislocation in three dimensions. The fonnulae
106 4 Linear Elasticity with Defects

derived there appear to be quite similar to (4.24). The trajectories in his Figure 4
look, however, rather different.

Concluding this Section, we consider the stability of material equilibrium of a


circular hole in the stress field of two symmetric dislocations as depicted in Figure
4.7.

Since the theory is linear, the stress fields of the two dislocations may be
superimposed. The results for J/, J 2 or ~. J Ip' however, do not follow from
superposition, because they are quadratic forms in a rplp (4.17) and an interaction term
will occur. Due to the symmetry of the problem it can be concluded, however, that
(/ = (2 = 0 corresponds to an equilibrium position. The material force in the
X 2 -direction is zero and the material forces towards the dislocations in the
x/ -direction are equal but opposite. If the hole is shifted by a small amount along
the x/-axis to the right, say, the attracting force of the right dislocation becomes
larger than that of the left. The hole will thus move further to the right. Therefore,
the equilibrium position is unstable with respect to x/ .

~ ---~
- - - - - - - ' f - t......------+-------'--I~-----X J
b ;1 b

fJ fJ

Fig. 4.7: Interaction of a circular hole with two dislocations

If the hole is shifted, on the other hand, by a small amount in the x2-direction, the
xr components of the force exerted by the two dislocations are still equal and
opposite. The forces, however, are now inclined with respect to the x/ - axis, both
components in X2 - direction add and drive the hole back into the original position.
Thus, the equilibrium position is stable with respect to x 2• The total potential energy
given as function of the position of the hole n ((J, 0 (influence surface), therefore,
possesses a saddle point at (/ = (2 = 0 and this equilibrium position is overall
unstable.
4.4 Path-Independent Integrals of Fracture Mechanics 107

4.4 Path-Independent Integrals of Fracture Mechanics

The broadest application of path-independent integrals is to be found in fracture


mechanics, specifically it is Rice's J integral taken around a crack tip which serves
as an important tool in ascertaining structural integrity of various components. It is
the purpose of this Section to briefly summarize some of the most important
concepts of fracture mechanics within the context of the linear theory of elasticity
and its relation to the mechanics of material space outlined so far. Readers
interested in comprehensive treatment of fracture mechanics should consult one of
the numerous texts available, e. g., Kanninen & Popelar (1985), Freund (1993),
Gdoutos (1993), Anderson (1995) and Gross (1996). A detailed listing of the
classical literature on fracture mechanics may be found in Paris & Sih (1965).

Linear elasticity leads to stress singularities at a sharp crack tip and Williams (1957)
has shown that the asymptotic behavior of stresses, strains and displacements in the
vicinity of a crack tip is the same for every crack problem. In particular, the stress
distribution is given universally by

aij = 1_
1 I hjJ (((J)
(KJ + KJI hjJI (((J) + KJIl IijJ I( ((J)l ) (4.26)
2rrr

with plane polar coordinates r, ((J centered at the crack tip (Figure 4.8). The
universal dimensionless functions J;j depend only on the angular coordinate ((J and
will be given explicitly further below. Different crack problems are distinguished
by the values of K 1 , K II and Kill .

At the crack tip, the stresses are governed by a 1/11- singularity. The constants
K 1 , K II and Kill are measures for the intensity of the increase of stresses near the
crack tip. They depend on the geometry of the body under consideration, on the
crack configuration and linearly on the extel11al load. They are called stress-
intensity factors (cf. Irwin, 1958).

Fig. 4.8: Coordinates for the crack near-tip fields


108 4 Linear Elasticity with Defects

Irwin (1958) defined three basic kinds of crack-tip deformation - called modes -
which are distinguished by the relative displacements of the crack edges (see Figure
4.9)

mode I: opening mode (K/) ,

mode II: inplane sliding mode (KIJ) ,

mode III: antiplane sliding mode (Kill) .

II III
Fig. 4.9: Crack-opening modes

Generally, the modes of crack extension occur in combined form. This is referred
to as mixed-mode condition. Stress-intensity factors are tabulated for several
conditions of geometry and loading, e. g., Sih (1973), Tada et al. (1973).

The total energy II of the system decreases as the crack length increases. This
decrease of llper unit crack tip extension is called the energy-release rate (or crack-
extension force). For brittle materials, Irwin (1957, 1958) established the relation
between the energy-release rate W (already discussed in Section 4.1 in the context
of material forces) and stress-intensity factors
222
K[ + K II KIll
;f} = + - (4.27)
E' 2G
As already introduced in (4.16), E . is Young's modulus E for plane stress and
E/( J - u 2 ) for plane strain, G
= lh E/( 1+ u) is the shear modulus and u is Poisson's
ratio. In the remainder of this Section we restrict ourselves to plane problems
(Kill = 0).

Having recognized that the total energy II of the system decreases by !!:.ll = - rrJ as
the crack length increases, the question arises as to where this energy might be
transferred. The answer to this question is contained in the proposition of Griffith
(1920), who assumed that the energy released by crack advance is absorbed by the
newly created crack surfaces, thus allowing the formulation of an energy-balance
equation (first law of thermodynamics)

!!:.ll + !!:.f = 0 or '/; = !!:.f,

where f is the specific surface energy per unit crack advance.


4.4 Path-Independent Integrals of Fracture Mechanics 109

The stresses and displacements near the tip of a crack are given by (cf. e. g.,
Gdoutos, 1993)
K} m m 3m
a /I = - - cos x... (1 - sin x... sin _'I') +
,I27rr 2 2 2

+ ~ sin !/!... (- 2 - cos !/!... cos 3 rp) (4.28 a)


,I21Tr 2 2 2'

0 22 = ~ cos!/!... (l + sin!/!... sin 3 rp) +


,I21Tr 2 2 2
K
+ __
ll_ sin !/!... cos !/!... cos 3rp (4.28 b)
,I21Tr 2 2 2

K} .
0 12 =-- Sin !/!... cos !/!... cos 3rp +
,12m 2 2 2

+ --
K cos !/!...(l . sin !/!... sin 3 ({J)
Il
(4.28 c)
,I21Tr 2 2 2 '

4G til I r
21T
/K} (2K - 1) cos !/!... _ cos
2
3;)+

+ K ll (2K + 3) sin ~ + sin 3; )j , (4.29 a)

+ K u (- (2K - 3) cos ; . cos 3;)/ . (4.29 b)

The symbol K (cf. Equation 2.38) equals (3 - u) / (I + u) for plane stress and
3 - 4 u for plane strain.

For pure mode I, K l may be used directly as a failure parameter. It turns out that a
crack extends when a critical material-dependent value of K l is reached, called
fracture toughness K /c . The assessment of a crack under mixed-mode conditions is
more involved. A variety of failure criteria exist. Different hypotheses are compiled
in, e. g., Kienzler (1993).

Let us consider a plane crack of length 20 in an infinite body referred to an (Xl' X 2)-
system of coordinates and subjected to some arbitrary loading as shown in Figure
4.10. The crack edges are free of sUlface tractions.

Rice (1968) defined the] integral as

J = j(Wdx] - ti ui,} ds) (4.30)


r
110 4 Linear Elasticity with DefeclS

F
./

:--===::==--:----~~'Xl

a a

Fig. 4.10: Crack in the (Xl' xz)-plane

with the components of traction tj = U jj nj introduced in equation (2.12).


Comparing (4.3 a) and (2.53) with (4.30) it becomes obvious that I is identical to
the component II

(4.31)

of the vector quantity

(4.32)

We want now to evaluate the J integral given by (4.30). In choosing the path r of
integration (arc length s), it is sufficient to start at any point of the lower edge of
the crack and proceed counter-clockwise around the crack tip to any point of the
upper edge, as sketched in Figure 4.11.

a . .. a .'
r

Fig. 4.11: Path of integration of the J integral around a crack tip

Two features are special to Rice's I integral. The first is that the contour r is not
necessarily completely closed, the reason being that the integrand of (4.30)
vanishes along the crack. Indeed, along the crack edges both dx2 and tj vanish.
The second feature is that the defect, namely the crack, is not enclosed
completely, but only a crack tip. The reason for the I integral maintaining its
path-independence is that in the x/-direction there is only one concentrated
material force (crack-extension force) at each crack tip and no material forces
along the crack edges. This will be discussed further below.
4.4 Path-Independent Integrals of Fracture Mechanics 111

The integrals Ji given by (4.32) around the crack tip can be evaluated by
considering the singular stress and displacement fields given by (4.28) and (4.29)
following a circular contour with infinitesimally small radius r, see Figure 4.12.

--_=::::::==+-~+--_Xl

a a

Fig. 4.12: Circular path around a crack tip with infinitesimal radius r

The result is

2
K}2 + Kl1
J} (4.33)
E'
K} K
J2 - 2 - -l1. (4.34)
E'

It is seen that J} is precisely the energy release rate c,f) per unit crack-tip advance in
its own plane as given by equation (4.27). By contrast, the expression for J 2 has no
immediate physical interpretation and will also be discussed below. The discussion
will further elucidate, why the component J, is path independent (such that the
evaluation of J, in (4.32) along any arbitrary path will deliver (4.33)) and J 2 is path
dependent (such that J2 in (4.32) and (4.34) coincide only if the path is taken along
a contour with infinitesimally small radius).

We proceed next to calculate the path-independent integrals J, L 3 = Land M based


on a contour around the complete crack as indicated in Figure 4.13.

We introduce the notation J;' for the value of Ji around the right crack tip and J/
for the value of J j around the left crack tip, both taken in the counter-clockwise
sense. The value of Ji around the complete crack will be denoted by J;'o,. The
following result is obtained.

J il0l = J'
i +
JI
i
!,a r;Jdx
+ -0 Ii; l' (4.35)

(4.36)

(4.37)
112 4 Linear Elasticity with Defects

where UJ is the jump in the material traction of Gi (cf. Chapter 3) across the crack.

IN = jf(W8
l
.. - a.,.
I)
u.. ) n.)1,
)~ ~.I ) XI'
O' - . - a... u.~.I. ) n.)1,
!(w~ I))~ ) XI'
0-

(4.38)

.. a
. .. a .'
Fig. 4.13: Palh enclosing the crack complelcly

Since the crack faces are assumed to be free of applied tractions, (4.38) reduces
to

IN = 0, (4.39 a)

(4.39 b)

I:
With the abbreviations

[WI = a (W (Xl' 0') - W (X)' 0-)) dx I ' (4.40 a)

(4.40 b)

(4.35) - (4.37), can be rewritten as

J )IOI = Jf
1 +
J'
l' (4.41)

(4.42)

L = a (Jf - J}) + [HI, (4.43)

M = a (J/ - Jz') . (4.44)

If the objective of a stress analysis consists in detennining stress-intensity factors


KJ> KIJ in terms of, essentially, far-field quantities, evaluated numerically for a
complicated boundary-value problem, then the procedure might be as follows. We
evaluate i/
Ol
, Jt , M and L in any convenient fashion using far-field quantities.
l
4.4 Path-Independent Integrals of Fracture Mechanics 113

Further, we evaluate [WI and [HI along the crack edges. It appears that this last
calculation is not very sensitive with regard to mesh size in, e. g., a finite element
analysis (cf. Kienzler & Kordisch, 1990).

With (4.41) - (4.44) I/ and I; at the right crack tip can then be given as

I/ = 1/2 (It + Mia), (4.45)

t
I; = 112 /(It ~ [W]) + (L - [H])laJ ' (4.46)

and from these K/ and K J; at the right crack tip are found to be

K/ ~ (IE * I/ ~ E * I; + IE' I/ + E * I;) , (4.47)

(4.48)

Similarly, for the left crack tip we have

I/ = 1/2 (It- Mia), (4.49)


t
I} = 1/2 /(It ~ [WJ) ~ (L ~ [Hf)laJ (4.50)

leading to

K'I = -
1
2
(/~ E' I/ + E' J} , 1- E' J/ - E ' J}) , (4.51)

1
2
(1- E * I/ + E * J) ~ 1~ E 'J/ - E ' I}) . (4.52)

As a specific example we consider a uniform far-field applied stress aij


(homogeneous state of stress), cf. Golebiewska-Hemnann & Herrmann (1981).
We will calculate the stress-intensity factors K t and KJI from path-independent
integrals. In our example, the complete analytic solution is known and the stress-
intensity factors, obviously, would be calculated directly based on this solution,
without recourse to any integrals. In a complicated boundary-value problem,
which would have to be treated numerically, the procedure to obtain the K values
using path-independent integrals is more efficient. Employing the analytical
solution, the following expressions are obtained

[jd = 0, (4.53)
Xl
--------
r-- - (4.54)
la 2 X}
Integration leads to
114 4 Linear Elasticity with Defects

[Wj = 0, (4.55 a)

[HI (4.55 b)

and further

(4.56 a)

(4.56 b)

M (4.56 c)

With (4.45) and (4.46) we find

(4.57)

(4.58)

and by means of (4.49) and (4.50) we have

1/ = - 1/ = - I] , (4.59)

Ii = - I; = - 12 • (4.60)

Substitution into (4.47), (4.48), (4.51) and (4.52) results finally in

(4.61)

(4.62)

As is frequently done in statics and strength-of-materials, it might be helpful to


introduce the concept of the free-body diagram in material space. We consider the
crack to be our object and sketch all the material forces acting on it, see Figure
4.14.

Forces in the x/-direction are present only at the two crack tips. They are equal in
magnitude and opposite in direction. It is now graphically apparent why Rice's I
integral around a crack tip is path independent and that the path does not have to
be closed as already discussed. As regards material forces in the xrdirection, it is
seen that their distribution becomes singular at the crack tip and that their sum
vanishes, i. e., Itt
= 0, as already found. The path-dependence of 1 2 taken around
a crack tip is also obvious, because this resultant depends on the length of the
crack segment enclosed.
4.4 Path-Independent Integrals of Fracture Mechanics 115

Fig. 4.14: Free-body diagram in material space of a crack in a remote stress field

The expression 12 = - 2K[ K u / E • merely represents the strength of the singularity


at the crack tip in the x 2-direction. The L integral is the total vector moment of the
force distribution in the x 2-direction, taken with respect to the crack center. The M
integral, finally, represents the scalar moment (or virial) of the two material forces
1/ and I; with respect to the crack center.

If the remote loading is such that

(4.63)

then, as is seen from Equation (4.54), the jump /i 2l vanishes

(4.64)

and the free-body diagram is reduced, as shown in Figure 4.15.

: 1/
a a

Fig. 4.15: Free-body diagram of a crack in a remote stress field with a':J = a;;
Only under this special circumstance, does 12 also become path independent and
equation (4.43) reduces with (4.58) and (4.60) to

analogous to the relation, which follows from (4.44) with (4.57) and (4.59) as

M = 2a I[ .
116 4 Linear Elasticity with Defects

The latter relation is also applicable without the restriction (4.63) and has been
discussed by Freund (1972).

If the remote loading is such that

we see from equation (4.56 b) that the material vector moment L vanishes, because
the vector moment [HI due to the material traction [j21, given in equation (4.55 b),

2
{HI 47Ta
+ --
'"
a 22 '"
an
E'

is just balanced by the moment of the material forces J/ and J; , see equation
(4.43), (4.58) and (4.60).

So far we discussed the situation of an infinite plane elastic domain subjected to


constant remote stresses inducing a homogeneous state of stress in the absence of a
crack. If the crack is inserted into this stress field, we found that the total material
forces J/"! vanish. Since the material forces describe the change of total energy of
the system due to a unit translation of the whole crack in xr or x 2 -direction, this
result is obvious, because the energy of the system is the same wherever the
position of the crack is. If the state of stress is not homogeneous, either due to non-
uniform loadings or due to the presence of a boundary or interface, the situation is
changed. Generally J;' will not be equal to- J/ causing a resultant material force
J/"! "# 0, i. e., the energy of the system is changed. if the position (and orientation)
of the crack is varied.

It is tempting to apply the concepts of equilibrium of elementary statics to the free-


body diagram in material space. In elementary statics, free-body diagrams are used
to determine reaction forces from equilibrium conditions. If our free-body diagram
(cf. Figure 4.14) would be one in physical space, the object (e. g., a beam of length
2a) would be translating and rotating in accordance with the principle of linear and
angular momentum. To be held in equilibrium, a vector force and a vector moment
equal but opposite to J/,,! and L, respectively, would have to be applied
additionally. By contrast, in material space our object. the crack, will not change its
configuration, e. g., extend or move in some direction, unless the material forces are
sufficiently large and beyond some critical value.

Although Rice's J integral is defined as an energy-release rate per unit crack


extension, a finite value of J does not imply a change in configuration of the crack.
A material transformation (translation, rotation or self-similar expansion) has to be
understood in terms of virtual change in configuration. In physical space it is
common practice to apply the virtual-displacement theorem. Consider, e. g., a rigid
body subjected to some arbitrary loading as depicted in Figure 4.16. The body is
supported in such a manner, that any change in configuration is excluded.
4.4 Path-Independent Integrals of Fracture Mechanics 117

\\ 1/
'"
~
M3
~

k'
X3
4/111

Xl

Fig. 4.16: Rigid body. rigidly supported

By means of the virtual-displacement theorem it is possible to calculate the resultant


force exerted by the external loads across a part fJ.B of the body in an arbitrary,
say xrdirection. We apply a virtual translation Sj in thexrdirection to 6.8, as shown
in Figure 4.17, and calculate the virtual work fJ.A of the forces in the translation
Sj ej

Fig. 4.17: Virtual displacement applied to the part /::"8 of the body

M = }}i Fi • S1 e1

leading to the component of the resultant force in the xrdirection acting across 6.8
it" _ M
F1 --
Sl
Although no displacements Ui occur, it is possible to calculate the rate of energy
change, here external work, due to a virtual displacement. Similarly, it is possible
to calculate contributions to the change of energy, if a part of a crack is virtually
displaced. Consider Figure 4.15 and, in evaluating Ji,
say, by the line integral
(4.32), we detennine the change of energy of the system due to a virtual translation
of a part of the crack in the xrdirection as shown in Figure 4.18.
118 4 Linear Elasticity with Defecls

Fig. 4.18: Virtual translation of a part of the crock

.,.;;::=-------.
J!!J/c

v
r
I
I
!l I m
~:
o
Fig. 4.19: Block on a rough surface (left) and crack under tensile loading (right)

This does not imply that the crack becomes kinked or that the crack tip is displaced
in some directions for any value of the change of total energy. Similarly, in
calculating the J/ integral there is, in general, no crack extension in the x/-direction
involved. It is only if 1/ exceeds a critical, material-dependent value J/ e (the so-
called fracture toughness, to be determined experimentally) that crack extension will
occur (see Figure 4.19, right part).
4.4 Path-Independent Integrals of Fracture Mechanics 119

As an analogy one might consider a solid block on a rough base subjected to tensile
loading F, as depicted in Figure 4.19, left part. As long as the force F is smaller
than the limiting friction force between the block and the base, the block will
remain at rest. Only if the force F reaches the limiting friction force, or exceeds it,
will the block lose its position of equilibrium and start moving.
5

Inhomogeneous Elastostatics

5.1 General Remarks

During the last few decades, considerable efforts have been spent on the design,
manufacturing and analysis, including failure characterization, of a new class of
materials whose properties change continuously in a particular direction (see, e. g.,
Moore & Samdani, 1993; Erdogan, 1995 and recent conference proceedings, such
as Ilschner & Cherradi, 1995). The terms "functionally gradient materials", or
"functionally graded materials" (FGMs) became common usage. A skillful control
of the concentration profiles of the constituents of an FGM opens the door for a
wide range of possibilities to improve the quality and characteristics of a system.
For instance, use of FGMs allows one to solve critical problems of compatibility
and adhesion between very dissimilar materials. In like manner, risks originating
from thermomechanical stress concentrations at interfaces can be minimized. FGMs
are also employed to enhance the material resistance against corrosion and wear, as
well as in the design of new types of thermal barriers. It is also contemplated that
FGMs can be utilized for electrical insulation in reactor components and to fabricate
special coatings which would resist oxidation and other kinds of chemical reaction.

The material presented in this chapter should also be useful whenever the scale of
a fracture mechanics problem dictates the modeling of materials by smoothly
varying parameters. Some soil mechanics and geophysics applications offer an
example of such a situation. It is also recognized that, due to material diffusion, the
elastic moduli near the interface of two bonded materials are sometimes better
modeled by smooth, yet rapidly changing functions of the spatial coordinates.

In what follows, we shall not aim at presenting many possible (and probably
relevant) aspects of conservation laws for inhomogeneous materials, but rather
concentrate on establishing a path-independent J integral for several classes of
inhomogeneity for plane elastostatics. This might be useful in evaluating energy-
release rates for plane cracks in materials whose properties change in the direction
of possible crack advance.
122 5 Inhomogeneous Elastostatics

5.2 Symmetry Transfonnations

We begin by recalling that in terms of a Lagrangian function

(5.1)

the conserved currents (or fluxes) Pj leading to the conservation law

Pi'; = 0 (5.2)

are expressed quite generally as (1.64)

p. = <P dL + (; /LfJ .. - v . dL) (5.3)


I a ~. J" 9 ~ ~ .
cr.1 a,I

The functions 0
and <Pet which describe the infinitesimal transformation and which
were introduced in Chapter 1

(5.4 a)

(5.4 b)

with E a small parameter, are either to be determined from the conservation law
(5.2) or by the invariance condition (1.62).

In plane elastostatics the dependent variables v" = u" are the components of
displacement in Cartesian coordinates and a = 1, 2. The independent variables xj
are the material coordinates and j = 1, 2 also. Therefore, it is not necessary to
distinguish between greek and latin indices in the following.

We already have stated the explicit expression for L (i. e., the negative of the strain-
energy density) in plane elastostatics. Concentrating on plane strain conditions we
have (cf. the remark following equations 2.41)

W = 2"A E j; Ejj + /L Eij E jj (5.5)

with the strain Eij = ( Uj ,; + II j ) /2, or, in terms of derivatives of IIi

W = 21 (A + 2/L) (111.1
2
+
2
ltv) + 21 /L 2
(ltI,2 +
2
ltv) + (5.6)

where A and jj are Lame's constants. For plane stress, A may be replaced by

A' = A (l -2v)/(l - v)
5.2 Symmetry Transformations 123

with Poisson's ratio u.

We wish to be concerned with a smoothly inhomogeneous body whose properties


can be described by

(5.7 a)

(5.7 b)

The form of Pi given above, together with the requirement that Pi'; = 0, leads to
an overdetennined system of linear partial differential equations for ~ and rPi'

The general solution of this system is easily obtained and is given by (Honein &
Hernnann, 1997)

(5.8)

where the arbitrary constants A, Ao' B. Bo• C. Co. D and Do are related to Lame's
moduli A. and J.l through the equations

Having solved for ~ and rPi' we can immediately obtain conservation laws by
substituting equations (5.8) and (5.6) into (5.3), provided (5.9) are satisfied. Thus,
it is seen that equations (5.9) may be regarded as a condition that restricts the
class of inhomogeneous materials admitting conservation laws under the special
version of Noether's theorem considered here. For every material in this class, the
number of conservation laws corresponds to the number of independent
parameters A, Ao •...• Do. We note that the constants A, Ao and Bo do not appear
in equations (5.9) and that they correspond to transfonnations (variations) of the
dependent variables only. Conservation laws arising from such variations (~ = 0)
refer to conservation laws in physical space, in contradistinction to those obtained
when ~ :f:. 0 and which are tenned here conservation laws in material space.

Indeed, one may verify that, in this case, the conservation law corresponding to
A o (translation along the Ilraxis) is nothing but the first equilibrium equation
(conservation of linear momentum in the x/-direction). Similarly, the conservation
law associated with Bo (translation along the 1l2-axis) is the second equilibrium
equation (conservation of linear momentum in the x 2-direction). The conservation
of angular momentum is associated with the parameter A.
124 5 Inhomogeneous Elastostatics

In the sequel, (by putting A = A o = B o = 0) we will concentrate exclusively on


conservation laws in material space. We shall start by recovering known results for
a homogeneous body.

5.3 The Homogeneous Case

By equations (5.9), C must vanish, while B, D, Co and Do may be chosen arbitrarily.


Upon setting in succession these last constants, except one of them, equal to zero,
one can show, by substituting (5.8) and (5.6) into (5.3) and applying the divergence
theorem, that the resulting conservation laws lead to

(1) The L integral for B :#= 0,

(2) The M integral for D :#= 0 ,

(3) The 11 integral for Co :#= 0,

(4) The 1 2 integral for Do :#= 0.

These results agree with those obtained previously by GUnther (1962) and Knowles
& Sternberg (1972) and have already been discussed in Chapters 2 and 4.

It was also mentioned in Chapter 2, that the transformation (5.4) deals only with the
so-called geometric symmetries. In her original paper, Noether (1918) allowed for
the case where (; and 4>; depend not only on xj and Um ' but also on all derivatives of
un, of any order. On the other hand, Bessel-Hagen's (1921) fruitful extension
(discussed in Chapters 1 and 2) is not included in our considerations. Hence, any
claim of complete classifications of conservation laws is precluded. The use of a
stronger version of the theorem may yield a less restrictive condition than (5.9) and
thus a wider class of inhomogeneous materials admitting conservation laws.
However, this topic will not be pursued here further.

5.4 The Inhomogeneous Case

The general solution of equations (5.9) can easily be effected by means of the
method of characteristics. This, however, will not be discussed in detail, for we are
not looking for completeness at this point. Rather, our aim is to examine some
5.4 The Inhomogeneous Case 125

situations that allow an extension of the important I-integral to inhomogeneous


materials; therefore, symmetries related to translation, rather than rotation, will be
emphasized.

Thus we let B = 0, and also Do = 0 since it is associated with the translation in the
x2-direction and not in the xJ -direction. Subsequently, we examine two cases:

a: D = 0

In this instance, equations (5.9) are equivalent to


A. 1 = /-L,l
(5.10)
A. /-L
and
- 2C = Co J.L 1 / Jl , (5.11)

which requires, in view of equation (2.24), that Poisson's ratio u be independent


of x J and the shear modulus Ii be an exponential function in xJ' i. e.,

(5.12)

where Jlo is an arbitrary function of X 2 and a is an arbitrary constant. To arrive at


physically meaningful results, however, these must be chosen in a manner which
insures that Jl > 0 in the region of interest.

Now, the components of P can readily be calculated from (5.3). And it is not
difficult to show, by applying the divergence theorem, that the ensuing path-
independent integral is given by

Ie = p(wn J - uk,J tk) -1 aUk tkJ ds, (5.13)


r
where tk = G jk nj are the components of tractions along the boundary r of a
domain S and nJ is the fIrst component of the unit outward normal to r.

This result remains valid for plane stress (with a proper adjustment of the strain
energy density, as discussed above), and it reduces to Rice's I integral for a
homogeneous body, as can easily be seen upon setting a = O. On the other hand,
one may verify, by direct calculation, that the obvious three-dimensional version
of the integrand of (5.13) remains divergence-free, provided that Poisson's ratio is
independent of Xl and the shear modulus satisfies

(5.14)

where Jlo is an arbitrary function of the coordinates x2 ' Xl and a is an arbitrary


constant.
126 5 Inhomogeneous Elastostatics

b: D ~ 0, but II and J are independent of Xl

In this case, equations (5.9) can be readily integrated and the result requires that
Poisson's ratio be constant and the shear modulus have the form

II = C (XI + Co I D) ·2C iD , (5.15)

which represents a power-law dependence, C being an arbitrary constant.

For later convenience, we let

- 2C I D = a, (5.16 a)

Co = 1, (5.16 b)

D =m. (5.16 c)

Now, it is readily shown that the resulting path-independent integral emanating


from the associated conservation law is given by

J. " fiwn 1 -', u., •m (Wx, n, -', x, uk' - ~ uu, 'J Jd<, (5.[7)
r
which, in terms of the standard J and M integrals, and

(5.18)

can be written as

J. = J + m (M - aU) , (5.19)

where we recall that M is defined in two dimensions by

M " I(WX, n. - '. x, u.) d< . (5.20)

r
The shear modulus II may, in this case, be expressed as

(5.21)

and must be subjected to the condition II > 0 in the region of interest, 110 being a
constant.
5.5 Relation to Stress-Intensity Factors 127

It is interesting to note that the homogeneous case is now recovered not by


setting a 0, which leads to a linear combination of the J and M integrals, but
rather by putting m .c O.

Under the same restrictions on the material parameters, one may verifY, on one hand,
that the result (5.19) holds true under plane stress conditions and, on the other hand,
that the obvious three-dimensional version of the integrand of (5.19) remains

I
divergence-free, provided M is replaced by its three-dimensional analogue given by

M = (Wxk nk - tk Xi Uk,; - ~Uk tJ ds . (5.22)


I'

5.5 Relation to Stress-Intensity Factors

One of the major concerns pertaining to linear elastic fracture mechanics analysis is
the accurate prediction of stress-intensity factors at crack tips. In this endeavour, the
standard J integral has proved to be an extremely useful tool. Under combined
opening (mode I) and sliding (mode II) this integral, around a right crack tip, is
related to the stress-intensity factors K, for mode I and Kif for mode II, by the
following relation
1 2 2
J ~- - (K1 + KI/) , (5.23)
E'
as already has been discussed in Section 4.4. For plane stress, E' IS equal to
Young's modulus E, for plane strain, E' equals E/(1 - v 2).

In this Section, we will prove that the same relation holds true for the J e integrals
derived in the previous Section under the conditions of their respective validity.

The literature concerned with analytical determination of stress-intensity factors


for inhomogeneous materials with smoothly varying elastic moduli is quite sparse.
The work by Delale & Erdogan (1983) appears to be the first to consider the
problem of mode I deformation in the presence of a modulus variation in the
direction parallel to the crack line. Young's modulus was assumed to vary
exponentially while Poisson's ratio was constant. In that paper, the authors stated:
"Even though no systematic study of the problem appears to have been made, it is
reasonable to expect that in nonhomogeneous materials with continuous and
generally differentiable elastic moduli the nature of the stress singularity at a crack
tip would be identical to that of a homogeneous solid." A similar statement was
also made by Erdogan (1983).

The Delale-Erdogan conjecture was proved to be true by Eischen (1987). Using an


eigenfunction-expansion technique similar to that of Williams (1957), Eischen
considered a constant Poisson's ratio and a general functional form of the Young's
128 5 Inhomogeneous Elastostatics

modulus variation, and was able to show that an r -II stress and strain singularity
exists at the crack tip, r being the radial distance measured from that tip.

Moreover, the angular variation of the singular stress field and the associate
displacements round a crack tip in an inhomogeneous material were shown to be
exactly the same as the angular variation in a homogeneous material. Eischen noted
also that these conclusions are not altered if Poisson's ratio is allowed to vary
smoothly.

Armed with these results, it is now a simple matter to relate the Je integrals to stress-
intensity factors. Taking the path of integration to be a small circular loop of radius €
centered at the crack tip and noting that the stress and displacement fields have,
respectively, 0 (r - 'h) and 0 (r + 'h) singularity there, it is easily seen that the
contribution of the last term of the integrand of both (5.13) and (5.17) vanishes as €
tends to zero.

Also, by taking the origin to coincide with the crack tip in Case b, the contribution
due to the M integral under the same limiting procedure is also seen to vanish. Thus,
in the limit, we are left only with the terms that form the standard J integral and we
obtain the following result
1 2 2
J =- (K[ f K a) , (5.24)
e E'
which follows by recalling that the functional form of the local elastic field at the
crack tip is unaffected by the inhomogeneity.

Note that, in relation (5.24), E' is evaluated taking the crack-tip values of Young's
modulus and Poisson's ratio.

Being truly path-independent integrals in their respective region of validity, the


extended Je integral derived here can be used directly in numerical methods, such
as the finite element method. Stress-intensity factors for graded materials can be
calculated accurately based on remote fields, rather than on less accurate numerical
crack-tip fields. The details are not different in principle from those for calculating
the J integral for homogeneous materials.

5.6 Examples

In this Section, we will illustrate the use of the J e integrals derived in Section 5.4 by
applying them to two simple examples. Both of these involve the double cantilever
beam shown in Figure 5.1.
5.6 Examples 129

Each of the anns will be modeled as a built-in beam of unit width bent either by a
force FIJ or a couple M o applied at the end. We recall that the elementary beam
equations may be written for the coordinate system depicted in Fig. 5.1 as
z

h
-1---.x

Fig. 5.1: Double cantilever beam

dQ + q = 0, (5.25 a)
dx

dM _ Q ~ 0, (5.25 b)
dx

M
(5.25 c)
.. E(x) /'

(5.25 d)

where, with appropriate sign convention, q is the lateral load intensity, Q the shear
force, M the bending moment, w the deflection and I the area moment of inertia. The
dependence of Young's modulus E on the coordinate x has been noted explicitly, for
we will be concerned with inhomogeneous materials. Readers without knowledge of
beam theory are referred to Chapter 9.

The governing equations must be supplemented by boundary conditions. At the crack


tip, x .= - a , we have

w l-a ~ 0
(5.26)

-dw
dx
I II =-a
~W'

=-a
~ O. (5.27)

The boundary conditions at the left end, x - a - e, will be specified shortly.


130 5 Inhomogeneous ElastoSlalics

Example I

As our first example, we will assume that the specimen is subjected to the forces
F o only. In this case the boundary conditions at the end of the upper arm read

-d w
dx 2
2
L end
=0
'
(5.28)

!!..- /E(x) I ~)I / = F (5.29)


dx ( L dx 2 lend 0

We assume further that Young's modulus is an exponential function in x and we


write

E (x) = Eo e {3x , (5.30)

where {3 is a real constant and Eo is a positive number.

Under these conditions, the integration of the field equations for the upper ann is
easily carried out and leads to the following expressions for the deflection and its
derivative at the end x = - a - e
F e/3 a
W
L end
= - 0
{33 EO I
(2 - 2e/3 f + 2{3 e + (32 e2), (5.31)

F e f3a
=+ 0 (1 _e/3 f +(3 e). (5.32)
{32 Eo I

Now, we can evaluate the J e integral given by equation (5.13) along the contour
surrounding the whole specimen. Due to symmetry, the calculation needs to be
performed only for half of this contour, which is depicted by the dotted curve in
Figure 5.1. In view of the kinematic assumptions of the elementary beam theory,
the strain energy density vanishes at the left end and the following result is
obtained

(5.33)

which, on substituting equations (5.31) and (5.32), results in

f =
e2 F02 (5.34)
e E (- a) I
The stress-intensity factor K[ is now calculated, by applying formula (5.24), to be
eF
K =- o (5.35)
1 II
We note that K[ has the same dependence on the load as in the homogeneous
case. However, for a given end deflection, K[ depends on {3, a, etc. As mentioned
before, a beam of unit thickness is considered. For an arbitrary thickness, b say,
f e in (5.34) is to divided by band K[ (5.35) must be multiplied by 1/15 (cf.
Kienzler, 1993).
5.6 Examples 131

Example 2

For our second example, we assume that the beam is loaded by couples only. Thus,
the boundary conditions at the end of the upper arm read

d -
E(x) I -
2

d.x 2 end
w!. = M o' (5.36)

-d
d.x
(1E(x) 1dd.x-W)L 2

2 end
= O. (5.37)

In anticipation of applying equation (5.17), we choose, by letting a = 0, the


origin to coincide with the crack tip, and we assume that Young's modulus varies
according to

E (x) (5.38)

where Eo and m are arbitrary constants subjected to the requirement that E be


positive in the material body.

The field equations are now easily integrated and yield the following results for
the upper arm:

W
o
---
Eal
M
fm-x6 3
+ X )
2 '
2
(5.39)

M
(m x2 x) '
2
dw o
--- (5.40)
- +
d.x Eol
dZ W Mo
= - (mx + 1) (5.41)
d.x 2 Eo I

The longitudinal displacement It, which vanishes along the central line z = h12,
may be calculated by requiring that the shear strain E" vanishes and may be
written as
dw
It =- (z - h12) - . (5.42)
d.x
The only nonvanishing component of the stress field is a.u ' which is given by
Mo
0xx = - - (z - hI2). (5.43)
I
Having these formulae, an evaluation of the J integral given by equation (5.17)
f

may be carried out to yield the following result:


2
M 3
i) along the path: = h, - ( $ x $ 0: 0 (- -2 m 2 {1 + 3mf) .
2Eo I

ii) along the path x - f, 0 $ z $ h:


M
2

0
3m
(-2 2 (2 - 3mf + 1/
2Eo 1
132 5 Inhomogeneous Elastostatics

Due to symmetry, the sum of these two expressions results in half the contribution
to the J. -integral, which is then given by

J = -
M; (5.44)
" Eo I'
leading to the following equation for the stress-intensity factor
Mo
K = - (5.45)
I II
Again K/ has the same dependence on the cross-sectional moment as in the
homogeneous case. For arbitrary thickness, b say, J. and K/ have to be multiplied
by lib and 1/ Ib,
respectively, as in the previous example.

In both of these examples, a simple calculation leads to the energy-release rate rJ


of the whole specimen per unit crack extension. In the first example, rJ is equal to
J., whereas in the second example, rg turns out to be

rJ = -
M; (l - me),
EoI
and does not coincide with J. (equation 5.44), as would b~ the case for a
homogeneous body (m = 0).
6

Elastodynamics

6.1 General Remarks

The homogeneous force equilibrium equations (qi = 0) of elastostatics (cf. eq.


2.60 a)

aji,j = 0

are extended to elastodynamics by addition of the inertia term pU j

(6.1)

where p is the mass density and U j , as before, the displacement vector. Dots
indicate differentiation with respect to time.

In contradistinction to elastostatics, the physical stress a ji is now no longer


divergence-free, i. e., we now have to deal with a source term in the stress
formulation and an additional independent variable, the time I.

It appears that there exist three different possibilities in dealing with the inertia
term of elastodynamics in the context of conservation laws.

I. The number of independent variables is increased by one, namely the time I,


and conservation laws have now the form

Pj,j = 0, where (6.2)

and we find ourselves in the (XI' Xl' Xl' I)-domain.

2. One takes care of the time-dependence by means of an appropriate


convolution.

3. The inertia term is dealt with as a true source term and the conservation laws
of elastostatics become now balance laws. In global (i. e., integrated)
134 6 Elastodynamics

form this implies that the path integrals of e1astostatics have to be


augmented by domain integrals, which together will be domain
independent.

In this Chapter, we shall present and discuss sequentially the three possibilities,
together with illustrative examples whenever possible. The Chapter concludes with
a brief discussion of the basic relations of wave motion.

6.2 Time t as an Additional Independent Variable

Before treating the general three-dimensional problem, we discuss this procedure


first with the aid of a spatially one-dimensional example of an elastic bar
(Herrmann, 1980) of cross-section A, mass density p, mass per unit of length
m= pA, Young's Modulus E and axial stiffuess EA (cf. Chapter 9).
The Lagrangian density (kinetic potential) per unit of length is given by

L=T-W, (6.3)

where the kinetic energy T (density) and the strain-energy density Ware given as
1 _.2 1 - 2
T = -2 mu = -
2 mu ,1, (6.4 a)

W = - EAu .
1 /2 1 EAu 2 .
= - (6.4 b)
2 2'x
The Lagrangian L is obviously a function of u / and u
and may depend explicitly
on x and t if EA and m
are not constant in space and time

L = L (x, t, II /, li) , (6.5)

In order to apply the Euler-Lagrange equation (1.65)

Ell (L) c elL __ ~ (elL )


e7v{J dx j e7v(3,;
we have to identify the following quantities, with {3 = J (subscript omitted below)
and j x, t
.0.

-? u, (6.6 a)

(6.6 b)
6.2 Time t as an Additional Independent Variable 135

d( )
dx2
~ d~/ =( r' (6.6 c)

vf3,1 ~ u l, (6.6 d)

vf3,2 ~u, (6.6 e)


cJL
-- ~ cJL -EAu l = -N, (6.6 t)
Ovf3,l eJul

cJL
-- ~ cJL =mu =p. (6.6 g)
Ovf3,2 ou
Here N = EA u I is the axial force and p = mlt is called the canonical momentum
density (Morse & Feshbach, 1953) and corresponds to the momentum of a bar of
unit length moving with velocity It. The Euler-Lagrange equation is rewritten as

~ (cJL)+ ~ (cJL)= 0 (6.7)


dt ou dx mIl
and leads (m and EA not being functions of x and t) to the equation of motion

mu = EAu l1 , (6.8)

or, after division by m, to the wave equation


with a = ~
p
I¥,
m
(6.9)

a being the phase velocity.

We wish now to calculate the components of the material momentum tensor, here
truly the elastic energy momentum tensor, because one of the independent
variables is the time.

From the general equation for the current Pi (1.64),

PI = cPlJfJ Ov
cJL + (,.
j
!LB. -
( 'Ij
VIJ'
fJ,j
cJL)
Ov
f3,i Ai
we know, that - dLJdVf3 ,i must supply the components of physical momentum and
~jj - vf3j dLJdvf3 ,j the negative components of material momentum. We already
identified the components of physical momentum. Here it is a vector with
components Nand - p. We proceed now to calculate the components of material
momentum

b.. = u. -
cJL - LB ;
Ij ,j mI.,I
i J

with L given by

L = -1
2
r-· 2
mu - EA u 12) . (6.10)
136 6 Elastodynamics

The result is

bu : b =u l elL -L (6.11 a)
11 dll I
btl: b 12 = - Nli = - EAu Ili , (6.11 b)

b 21 = pu / = mUll
-. /
, (6.11 c)

btl : b22 . dL
= II -
dti
-
L = -
1;-
2
mu·2 ~
A12)
E'rlU . . (6.11 d)

The quantity b" is the material force per unit of length, also called wave stress
(cf. Morse & Feshbach, 1953), whereas bxt is the stress rate of work or rate of
energy flow along the bar, b,x is called the wave momentum and, finally, bit is the
total energy per unit of length H = T + W or the Hamiltonian per unit of length.
In this one-dimensional example the material force bu happens to coincide with
the expression for the Hamiltonian bit except for a minus sign, on the other hand,
btl and b ,x differ merely by the factor - ii.

Equation (6.2) yields, with the implications cjJfJ = cjJ = 0 and ~j = aj = const.

(a J = 1, a z = 0) , (6.12 a)

(6.12 b)

or after substitution (Herrmann, 1980)

1(NU I + Pli/ I + (im /J' 0, = (6.13 a)

1 1 '
- (Nti /) + 2 (Nll I + PliJ' = o. (6.13 b)

The first equation above states that the spatial rate of change of the material force
b u is balanced by the time rate of change of wave momentum btx • The second
equation states that the spatial rate at which stresses do work in the velocity li, btl
, is balanced by the time rate of change of the Hamiltonian bit, or in other words,
(6.13 b) corresponds to the equation of continuity for energy flow along the bar,
or the principle of energy conservation (first law of thermodynamics).

The integral (global) form of bjiJ = 0 leads to the line integral

I bji nj ds = 0.
r

To illustrate this form, let us consider a bar of length and a rectangular circuit in e
the x -t plane sketched in Figure 6.1, where 1" is some arbitrarily fixed time.
6.2 Time f as an Additional Independent Variable 137

t*

-+- -----L----..X
o
Fig. 6.1: Integration contour in the space-lime domain

The closed line integral can then be evaluated as follows:

e e
I
x

I pu
=

=X:
o
l
dx
I
/=, I
+-
2
o
(ptl + Nu I) dt
/
x =0
(6.14 a)
e

= t :
1
-
2
0
I (pu + Nu I) dx /
, =,'
=

t =e

I
x
"
- -
1
2
0
(pu + Nul) dx
/
, =0
+ /
0
Ntl dt
x
/ =0
(6.14 b)

The first term on the right-hand side of each equation above can be evaluated on
the basis of prescribed initial conditions. Similarly, the second terms on the right-
hand side can be evaluated on the basis of prescribed boundary conditions.

In the second equation (6.14 b) it can be ascertained that the second integral of
the right-hand side vanishes, since either N (free displacement) or (fixed u
displacement) are zero at the ends of a bar. This leads to the conclusion that the
total energy of the system at time zero is equal to that at any arbitrary time ( , i.
e., the Hamiltonian is a constant of the motion, independent of time
e
:t I (ptl + Nu I) dx = 0 , (6.15 a)
o
or, in other words, the total energy is conserved.

By means of modal analysis it can be shown that the wave momentum p = m u


and the displacement gradient u / are orthogonal functions. Therefore, the first
two integrals in (6.14 a) vanish, leading to the conclusion that, due to ( being
arbitrary, the material forces per unit of length b u at the two boundaries are equal

b xt ! =0
=bxx ! =e
(6.15 b)
138 6 Elastodynamics

provided the bar is homogeneous. In other words, the material forces acting across
the ends of the bar are in equilibrium with each other, because the resultant of the
distributed material forces along the bar vanishes.

Having the conservation laws already established, it might be intriguing to


determine, in a backward manner, the characteristicfemploying the NA method. We
know from Section 2.4 that the conservation laws are equal to the product of the
equation of motion (6.8)

4 = mil - EAu II

multiplied by the characteristic f, i. e., an integrating factor. Replacing Nand p by


EAu I and mu,
respectively, it follows from (6.12) and (6.13)

= U I (mil- £Au II) (6.16 a)

'.7 U 1 4,

-~ u (mil - £Au II) (6.16 b)

= u 4.
Hence, the characteristic f is found as a sum of the displacement gradients in space
and time

(6.17)

such that the relation

p. 1,1
= - (~.bl.
~J i/./ - ab---
} = - a-u- (mil -£Au
1),1 ).;
II) = -fi~ (6.18)

can be stated immediately.

It should be added that the characteristic Q (cf. equation 1.63) within the context
of Noether's approach, is given as

Q=4>-('u.
1 ,I

(6.19)

and, as before, Q = - f, see equation (1.107) and (1.109).

Let us recall that the local conservation laws above are related to the translation in
material space 0
= aj = const. and 4>/3 = 4> = O. The corresponding conservation
law
6.2 Time t as an Additional Independent Variable 139

related to translation in physical space cf> = a = const. and 0


= 0 is the equation of
motion (6.8). Of course, further conservation laws exist and will be discussed for
the general three-dimensional case in Section 6.4 and, in more detail, for the one-
dimensional case in Chapter 9, in the context of strength-of-materials theories.

Returning to the general domain (Xl' X2, X3, t) it might be instructive to represent the
complete energy momentum tensor in the form of a 4 x 4 matrix (cf. Morse &
Feshbach, 1953)

I
I
I
I
I
I (6.20)
b31 b32 b i3 I b3,
__________________ L _
I
I
bit .

The bjj above are the components of the Eshelby tensor (i, j, f3 = J, 2, 3)
aL (6.21 a)
b ij = + uf3,j ~ - LSij
{J"
=- (L~ij + a ik Uk) (6.21 b)

already discussed several times.

The vector R with components btl, b'2' b'3 is called the field- (or wave-)
momentum density and is defined by

R.I = b.tl = + pul,l.. It j . (6.22)

The vector S with components bIt' b2" b i , is called the field intensity defined by

(6.23)

and represents the work of stresses along the rates of displacements, or the
components of the energy-flow vector.

The element bit in the matrix above represents the total energy density H = T +
W, i. e., the Hamitonian.

The complete matrix is then written as


140 6 Elastodynamics

XI bll bl2 b13 : - u lj u j


I
x2 b21 b22 b23 I - U2j Uj
I (6.24)
x3 b31 b32 b33 : - u 3j uj

-------------------~----
I
puj,Iu j PUj.2Zl j PUj,3Zl j I H

To facilitate a direct comparison with the development in Morse & Feshbach (1953),
it should be mentioned that their matrix on p. 321 is the transpose of the matrix
given above. It also should be mentioned that the representation (6.24) is identical to
(6.21 a), provided, the indices i andj range from 1 to 4 (x4 = t), while the range of
f3 remains 1 to 3.
As discussed in Section 1.1, the terms bt; (i = 1, 2, 3, t) are densities and the terms
bji 0 = 1, 2, 3; i .~ 1, 2, 3, t) are fluxes and the conservation laws indicate that the
time rate of change of a density is balanced by the divergence of the associated
fluxes. The first three conservation laws (no summation over time t)

b.+b
jk,j tk,t
=0 j, k ~ 1, 2, 3 (6.25 a)

or with (6.22)

bj k .j· + R; = 0 (6.25 b)

express the balance between the net material force and the time rate of change of
the field-momentum density. The fourth equation

b.
jt,j
+b It,t =0 (6.26 a)

or with (6.23)

(6.26 b)

expresses the balance between the divergence of the energy-flow vector and the
time rate of change of the Hamiltonian, i. e., the principle of energy conservation.

Since the Eshelby tensor bij in material space corresponds to the Chauchy stress
tensor U ij in physical space, one might try to construct a 4 X 4 matrix (cf.
Herrmann & Kienzler, 1999), analogous to the matrix (6.24) by introducing the
negative physical momenta - PZl j as matrix elements utj such that the equations of
motion (6.1) are rewritten as

U·k·fU
j ,j tk,t
=0. (6.27)
6.2 Time t as an Additional Independent Variable 141

Since the stress tensor is symmetric it is tempting to introduce

(6.28)

Thus it seems that the physical momenta play the role of either densities or
fluxes. From

ajl,j + all,t =0 (6.29)

it becomes obvious that a" is equal to the negative of the mass density p.
Equation (6.29) is then rewritten as

- plij,).. - P =0 (6.30)

and, surprisingly, is identified as conservation of mass (cf. Lamb, 1932). Thus the
complete 4 x 4 matrix is given by

I
I
I
I
I
I
I
I
a 23 I
I
I
I (6.31)
I .
a 33 I - PU 3
---------------------f-----
• • . I
t - pUJ - PU2 - PU3 I - P
I
I

which might be called the mass stress tensor. As a, seemingly, novel result we
recognize that the stress equations of motion are intimately linked to conservation
of mass. This is in physical space. In material space, the balance of material forces
and field momentum is intimately linked to conservation of energy. The duality
between both matrices is rather striking. It provides an example in which a novel
result in the familiar physical space has been established due to a well-known
(Morse & Feschbach, 1953) result in the less familiar material space.

Similarly, the balance equations of angular momentum may be arranged in matrix


form in physical as well as in material space. If we replace the volume forces qj by
d'Alembert's inertia forces - pi; j in (2.67 b), we find in physical space

(6.32)

or, shorter with

(6.33 a)

and
l42 6 Elastodynamics

(6.33 b)

leading to the fonn

(6.34)

This (4 x 3) matrix fonned by (6.33) may be completed to a (4 x 4) matrix with


the help of the balance equation of scalar momentum or virial (2.67 c)

(~j uij - ] Ell;


- (n - 1)
) -
V,,;
(~.} Plt .
}
J' ~ 0, (6.35)

or, shorter with

(6.36 a)
] - (n - 1) V
and

(6.36 b)

resulting in

(6.37)

With the abbreviation J3 ~ £/[1 - (n - J) v}, the (4 x 4) matrix m u (i, j = J, 2, 3,


t) is written out as

XI

I
I
I
I
I
I
I
I
- - - - - - - - - - - - - - - - - - - - - - - - - - - - - _I _ _ _ _ _ _ _
I
XIPU 2 + X2Plt I I
I

k =], 2, 3. (6.38)

In material space, the corresponding matrix may be constructed with the


integrands of the Land M integrals, cf. equation (2.68 e) and (2.68 f),
respectively. If we introduce

e
I nnk
= E
k}
'l'
· (x b .. + Uk u.),
Y
(6.39 a)

e
1/
= x.} b I} - ( tu..} n-]
+ --
2 u.}) uI}.. (6.39 b)
6.3 Convolulion in Time 143

and additional densities ern and err (expressions, which will be given in Section 6.4),
(6.40 a)

(6.40 b)

the corresponding conservation laws in material space are given as

e. . + e111,1
111,1
= 0 '
(6.41 a)

(,,; + elt,l =0 . (6.41 b)

Again, the parallelism between physical and material space is remarkable


(Herrmann & Kienzler, 1999).

6.3 Convolution in Time

A path-independent integral for elastodynamics has been established by Gmtin


(1976). It is based on convolution in time, denoted by *. For any two functions
f(x,t) and g(x,t), convolution in time is defined as (cf., e. g., Flugge 1962)

f*g /f
o
(x, t - 1') g (x, 1') d1' = F (x, t). (6.42)

In the theory of mechanical vibrations, equation (6.42) is known as Duhamel's


integral, see, e. g., Weaver et al. (1990). In order to provide a better understanding
of what follows, some rules to treat convolution equations are briefly recalled. By
substitution of a new parameter A given by

A=t-1' (6.43)

into (6.42) it is easy to show that convolution in time is commutative

(6.44)

The integral (6.42) depends on two variables, namely x and t ; the upper limit of
integration depends also on t . Generally, differentiation of an integral depending
on a parameter with respect to this parameter is performed as (Flugge, 1962)
144 6 Elastodynamics

f3(y) f3(Y)

d
dy
/ h (~y) dx / OIl ~ y) dx ,

a(y) a(y)

+ d~~) h (f3(Y), Y) - d~~) h (a (y), y)

Applying this fonnula to (6.42) leads to

(f * g)" ~- r * g + f (x, 0) g (~ t) (6.45)

and

(f * g),x = .f,x * g + f * g,x . (6.46)

For the fonnulae given above, the variable x may also be interpreted as a vector x
with components Xk •

We now choose for f and g the displacement component uj = uj (x k• t) and assume


homogeneous initial conditions, i. e.,

(6.47)

Applying (6.45) once to uj * uj leads to

(U/ * ul'
:Y = Uj * u) . (6.48)

Applying (6.45) again leads to

(u)*}
u.r· c
ii}* u·
}}}
= It. * u. (6.49)

From equation (6.46), on the other hand, it follows


1
-2 (u.} * u),
} ,I = u.} * u·}. 1.. (6.50)

Thus, the following relation

= l
1
ii.} * u·
j.l
= (u.} * u.)··
j.I
= -2 lu}. * u.)··
( , ) ,i 2 (ii} * u)}.i (6.51 )

is valid. With the fundamental equations

(6.52)

and with (6.50) and (6.51) one can write


1
( -2 akJ' * U'k)
}'!,i ~'c'k} pq Up,q * uk'
}, I

= Ic
( , )kpq U p,q * u },I/
/
k
- akJ',k * u',j.l
6.3 Convolution in Time 145

f'lak]· * U·· /kJ j,I


- pii.
J
* uJ",i

= (a kj * Uj,i~ - i (u j * u~-) (6.53)

and by the usual divergence theorem

/11 2 ak·] * U·J. k n.I - ak]· n k * UJ.I + 12 p UJ * u·. n.JdA


J I

=/Ii (pu
s
j * uj + anj * Uj / I) fJ ik - a kj J
* uj,i nk dA = 0 (6.54)

valid in two or three dimensions.

In two dimensions one can write a path-independent integral! = !(t)

I = 12 r ]
/!a k. * Uk
J.
+ plioJ * li.Jdx
J 2 - fk.] n k * u.J. J ds J, k = 1,2
(6.55)
r r
for a contour r around a crack tip, Figure 6.2, just as the J integral of 20
elastostatics.

ds

Fig. 6.2: Contour of integration around a crack tip

For an isotropic body in plane strain or plane stress, the relation between !(t) and
Kit), the stress-intensity factor, will be

11 -~ v
plane stress,
I with E" = plane strain. (6.56)
2

Restrictions on the use of !(t) are discussed by Gurtin (1976).

It appears that convolution has not been employed extensively in establishing


conservation laws or in applications to fracture mechanics. Mention might be
made of the work of Francfort & Golebiewska-Hemnann in thermoelasticity
(1982).
146 6 Elastodynamics

It may be mentioned further that the path-independent integral (6.55) as given in the
real time domain has an equivalent expression in the Laplace transformed domain.
The convolution integrals are then replaced by products of Laplace transforms of the
corresponding functions, cf. Nilsson (1973, 1990) and Freund (1993).

6.4 Domain-Independent Integrals

A third manner to establish conservation laws of elastodynamics, while renouncing


the desire to have path-independent integrals, is to construct instead a sum of path
integrals and domain integrals which itself would be independent of the domain.
This point of view was followed by Eischen & Herrmann (1987) where, in addition,
a connection between these balance laws and energy-release rates for defect motions
was established. Not only inertia terms of the type pu i were included, but also

material inhomogeneities, thermal effects, anisotropy and body forces. In this


Section, thermal effects are not treated, but relegated to Chapter 8.

The equations of motion (6.1), augmented to include body forces qi' are given as

(6.57)

and Hooke's law is

(6.58)

The material is considered to be inhomogeneous (indicated by the dependence of


the components of the elasticity tensor on xp) and anisotropic. The symmetry
conditions (6.52) still hold

the isotropy condition (2.25), however, is not satisfied in general


111m
(0.
JIl
It m,l. +0.1t·
nJ j,m
);t:. O. (6.59)

The Lagrangian density is

(6.60)

with the kinetic energy T


1
T= - P ("'(:,) It; It;, (6.61)
2
6.4 Domain-Independent Integrals 147

the elastic strain energy W


1
W = -2')
c" 1d (xr I u·,},
U (6.62)
' kl
and the potential of external forces

(6.63)

By partial differentiation, the constitutive relations are recovered as

Goo
ow (6.64)
]I oil..
I,J

q. = ~ -
ov (6.65)
I Ott;
The desired results are obtained next by simple vector and tensor calculus
operations involving the Lagrangian, as outlined in Chapter 2.

Gradient

The first balance law is obtained by considering the gradient of the Lagrangian.
From Chapter 2 we know that this operation is identical to applying the
translational symmetry group to the action integral. The calculation involves the
following steps
elL elL elu.
eltt; + __ elL cluj elL
grad (L) = L ,k -----..!d + --+-
elu; Oxk Ott.
I,J
Ox k. Ott; Oxk Oxk

(6.66)

The "explicit" derivative of the Lagrangian is, as discussed In Chapter 1,


calculated according to

dL d . /,
iJx = -dx L(xP' u·I t u..IJ' u.)
I El j const, Ui,j = COILS/,
. =const,
Ui X
p
~ const for p ¢. k
k k

(6.67)

When the divergence of the stress tensor is introduced into equation (6.66), it
follows that
dL
Oxk
The equations of motion (6.57) are then substituted producing
148 6 Elastodynamics

Finally, with equation (6.21) it follows that


. • oL
bJ'k',J + ( pu 1 u'k)
I,
+ -dt =0, (6.68)
k
with "dLldxk given by equation (6.67) above.

This differential fonn represents balance laws valid for a linear, inhomogeneous,
anisotropic solid which is subjected to inertial and body force loadings. If
inhomogeneity and body force loadings are absent, the resulting expression is in
accord with equation (6.25) (see also Fletcher, 1976). An integral fonn of equation
(6.68) may be obtained upon application of the divergence theorem. If B is a
regular bounded body with volume V enclosed by a surface S of area A, whose
unit outward nonnal vector is n, it follows that

S
fk nj dA ,
B
/lr Pl', ""k)' + oL
dt k
JdV = 0 . (6,69)

Curl

The second balance law is obtained by considering the curl of the "Lagrangian
moment" Lx. That is,

curl (Lx) = Ekij (Lx),;

o (Lx)J ou /n.n
I
0 (Lx) Otl
=E" J __ /n,
klJ Otl dt
111 I
at In,n

= Ek1)" IIfpu 111


II nI,l,x} -a
nm
, x}
um.nl 4 q 111 U111.1,x) + oL
dt. xl,
}
I

where oLi OX; is the same as given in equation (6.67), provided that k is replaced by
i. As before, the pennutation symbol is denoted by E kij , If the divergence of the
stress tensor is introduced, it follows that

E kij I (Lx),; pllm U m'; xj f (a"m um,; x),,, -

- a"m, " um,; xj - al/l" Um,i xj,n qm um,; xj I = E kij : -


1
xj ' (6.70)

When the equations of motion (6.57) are substituted and the quantity E kij am; uj,m is
added and subtracted on the left side of equation (6.70), it follows that
6.4 Domain-Independent Integrals 149

If the divergence of the stress tensor, the equilibrium equations and the Eshelby
tensor (6.21) are then substituted, the desired result is obtained as

SIi;.., Ib ni x)' + ani u).) • n + Sk""I} !Plt m


(, . X -
U tn,') pit.I u.)"
) +

(6.71)

Equation (6.71) represents another balance law. The first three quantities in the
second parentheses account for material inertia, inhomogeneity and body forces,
respectively. The last quantity in parentheses vanishes when the material is
isotropic, see equation (2.25 b). If inhomogeneity, body force and anisotropy are
neglected, a rearrangement of indices leads to

/x k
Snk)' (-- b..'i + Uk a I)..),i + S fl k'") ("x k Plt
. In
U In.}. - Uk pu).)" O. (6.72)

The fluxes

and the densities

have already been introduced in (6.39 a) and (6.40 a), respectively. Thus, in short,
equation (6.72) is rewritten as

(6.73)

in accordance with equation (6.41) (no summation over time t).

Equation (6.71) appears in integral fonTI as

J!;k" c...) I)
/b . x·J
(I nl
+ a . u.)n dA
til) fl
+

,/"j !f B
pUm um,; 'J - P"; uj'
aL
+ - x + q. u. +
ax.)
I
J)
(6.74)

+ (amj Ui,m + ajm um,;)} dV = 0 .

Divergence

A third balance law is obtained by considering the divergence of the "Lagrangian


moment," i. e.,
150 6 Elaslodynamics

div (Lx) = (LXJ.k

J (Lxk) Jli i
Jll i efxk

JL
= pit; It i.k Xk - aji Iti.jk Xk + qi U;,k Xk + - Xk + nL. (6.75)
efxk
A parameter n has been introduced which is equal to 3 for three dimensions and 2
for two dimensions. This accounts for a term x k. k which arises from the operations
indicated in equation (6.75). If the divergence of the stress tensor is introduced it
follows that

(Lxk) k +(aji Iti,k xJj - aji.j U~k x k - aji It~k xk.j - qi It i.k x k
.. JL
= pu.1 UI, k x k + -eX x k + nL
k
The equilibrium equations are then substituted resulting in

I/rU;k
\) + a)1.. u·I. k.) xkj.
.)
= (pit 1 UI. k xkr + a·
)1
u.
I.)
(6.76)

It proves convenient to write equation (6.76) as

+ a·)1 U'k)
I,
xkj. = (pit., Uk
I,
XkJ' + (- L + pit; it) + (n - I) L +
.)
JL
f 2qi It; + - Xk .
efxk
The following two relationships are useful

L ,,1 I( pit; ul .. (aji il),j + qj il;/ '

- L + pit. Lt, c' f- Lt + ptit, it.)'" (t a .. it.) .


I'! 1 1 ]I 1,)
From this follows the desired result
. . n - 1 ) 7 r . n - 1 7"
/h)'k X k - a k (til k + - - ilk II + / .. Lt + pili /U;,k X k + tu;+ -2- u),1 +
1 ) 2 J.j I "

(6.77)

Again the fluxes

e x) h I) . n 1 u· ) a ..
1/
= ( til) + --
2 ) I)
and the density

ell . Lt + pili (ili.k Xk I tit i + n 2 1 It;)


6.5 Energy-Release Rates 151

have already been introduced in (6.39 b) and (6.40 b), respectively, such that, in
short, equation (6.77) is rewritten in accordance with (6.41) as

(t,; + ell,t = 0, (6.78)

if inhomogeneity and body force are neglected as before (no summation over time
t).

The integral form of equation (6.77) is

(6.79)

6.5 Energy-Release Rates

In this Section, expressions will be derived for the energy-release rates associated
with certain crack (defect) motions valid for two-dimensional fracture problems. An
extension to three dimensions is straightforward. Figure 6.3 depicts a crack located
in a two-dimensional elastic region. As discussed in Chapter 4 in the framework of
elastostatics, the energy released during three possible crack motions will be studied
in an elastodynamic setting.

Fig. 6.3: Body with crack and definition of integration paths


152 6 Elastodynamics

These motions are: (0 rate of rigid translation such that all points on the crack
surfaces move with v = vk e k , where vk are constants; (iO rate of rigid rotation about
the X J - axis such that points on the crack surface move with velocity v = vk ek = -
E Jkll XII UJ ek , where UJ is a positive constant; (iii) rate of self-similar expansion along
the crack axis such that points on the crack surface move with velocity v = V k ek =
ll'Xk e k , where ll' is a positive constant.

In the current dynamic setting it is not sufficient to consider merely the energy-
release rates, but the energy-rate balance equations have to be introduced. As already
mentioned in Section 4.4, such energy-balance equations in fracture mechanics can
be established only if the concept of surface energy, as introduced by Griffith
(1920), is taken into account.

The energy-rate balance condition which must hold during any of these crack
motions may be stated as follows (see Freund, 1972 and 1993)
• •i •a
p K+II +II +F, (6.80)

f
where

p lim n} it I ds, (6.81 a)


}I
rr-+ O

fi ..
fo

K = lim f d A = Iim - pit; It; dA , (6.81 b)


f, -+0 f, .... O 2

If
S(I) S(I)

IIi lim (6.81 c)


lim f d A
r,-+o r,-+o
2 a}/ l lI,}· dA ,
Sit) Sit)

II
a
~ lim
f, -+0
fdA - lim ; , " ,
r,-+o
dA. (6.81 d)

S(I) SIt)

The quantities P, K. IIi and II O are the rate of work of the tractions on f o , the total
kinetic energy, the total elastic strain energy and the total potential of the external
body forces in A, respectively. Recall, that T, Wand V are the densities per unit of
area of the kinetic energy, strain energy and potential of external forces,
respectively. The quantity F is the energy-absorption rate on both the crack
surfaces and the crack tips, measured as a change in energy per unit thickness per
unit time. For brevity, the paths f,t and f El are referred to collectively as f,. The
paths f, and f c move rigidly with the crack as it executes the motion described
above. Therefore, the position of the paths fc and f c is time dependent, and the
transport theorem must be used to evaluate K, jj' and ital jja. Consequently,
6.5 Energy-Release Rates 153

K = f.-+
lim
0 'j
!
!Pii; it; dA
f'
(6.82 a)

-Ii
S

lim (fa..Jl Lt I,J.. dA


f.-+O
2 a.JI u··
I,J
Vk n k. ds) , (6.82 b)

S r£ f fc

(6.82 c)

s f . If,

After some manipulations it follows that

F = lim
f.-+O
I((T + W + V) Vk Ilk (6.83)

f£ f fc

This expression for the energy-release rate will be split into two parts, one having
to do with the energy absorbed at the crack tips and another accounting for
energy absorbed along the crack surfaces

F = F,;p + Fsur , (6.84)

where

(6.85 a)

(6.85 b)

It can be shown that near the tip of an extending crack the field quantities obey
the "transport assumption" (Ehrlacher, 1981)

il( )=_vil()
a k eXk
Furthermore, observe that on r c , Ilk' = Ilk - and aji Ilj = O. After introducing the
Hamiltonian H = T + W + V, it follows that

Flip = lim ftHV k Ilk - aji Il j U;.k vkj ds, (6.86 a)

fji
f.-+O

Fsur = lim (H' - H-j Vk Ilk' ds. (6.86 b)


f.-+O
f,

The next step is to evaluate the expression for the crack-tip energy-release rate
given above for the three crack motions discussed earlier.
154 6 Elastodynamics

Case (i). vk = constant

(6.87)

The energy-release rate measured as an energy change per unit translation per unit
thickness is denoted 'r; J'
This energy-release rate can then be written in terms of
line integrals around the remote paths f ll and f c , and a domain integral over S by
applying the divergence theorem to equation (6.87). Introducing the symbol J k
(k -- 1, 2), it follows that

"kT ---
'f' j/ru'i:
!
!llU
Jk -a..
JI ll·kiI
I"
n·J dS I . j/i L!
·H ,

fa f,

-
j/ ..
!pll.I ll.I, k -
..
pll.1 ll.I, k + eX -J dA.
ilH
(6.88)
k
S

This expression may be thought of as an extension to the usual J integral found


for crack problems. Equation (6.88) accounts for inertia, body force and
inhomogeneity. For completeness, the expression for the explicit derivative of His

I 1
2 P,k ltll
p p
+-c
2 prs/,k II II
p,r 5,1
-qp,k II
p'

Case (ii). Vk - - E 3kll XII U)

.,
'f R (6.89)

The energy-release rate measured as an energy change per unit angular rotation
per unit thickness is denoted by r; R. Again, this quantitiy can be expressed in
terms of remote line integrals and a domain integral. Introducing the symbol L 3

L., l
'f, R = fr ~ 1)
/Hx)
L
t)un -- a
mn
l ln,l
· X] + a1m)
. ll ../' nIn ds +

,}d;' H / X, n; dJ
r
"1;,
) C,,:'l}
!pZt In . x - pit m
zi m,l) II
m,l
.X +
J
pllI· llJ' -

q. ll. + la
I.
mi u)'. III + a illl UIII .·
) + ilH x.JdA (6.90)
1 J eX. J
I
6.5 Energy-Release Rates 155

Case (iii). Vk = axk

(6.91)

The energy-release rate measured as an energy change per unit thickness is denoted
by r; E. This quantity can be expressed in terms of remote line integrals and a
domain integral. Introducing the symbol M

M r; E = j!HX k ljjk - aji It;,k xkj n j ds


fo

It is seen that the balance laws given by equations (6.69), (6.74) and (6.79) do not
lead to energy-release rates given by equations (6.87), (6.89) and (6.91). Indeed,
to arrive at expressious for energy-release rates, the Hamiltonian H instead of the
Lagrangian L needs to be used.

It is instructive to note that jf the operations grad (H), curl (Hx), and div (Hx)
had been considered, the resulting balance laws would have been in a form
readily associated with energy-release rates.

grad (H) :

(6.93)

curl (Hx) :

! 3ij f Hx lj
j;m - a/till u ,,' x)'
"
+ a / t.i lu.j
)
n /tI ds

-}ij /
f

PLl m Ltm,i xj
S

+ (ami Uj,m + a jm Um) + dH x.j dA = 0 . (6.94)


cit. )
I
156 6 Elastodynamics

div (Hx) :

The expressions presented in equation (6.93 - 6.95) could have been derived in an
alternate manner by eliminating

L ~ T - (W + V) (6.96 a)

in favor of

H ~ T + (W + V) (6.96 b)

in the balance laws shown earlier.

6.6 Wave Motion

In linear problems of dynamics of continua, waves are recognized by the existence


of elementary solutions of the form

(6.97)

where k i are the wave numbers, UJ is the frequency, A the amplitude and 0 the
phase. Substitution into the governing equation for 1J results in a relationship
between UJ and k.I

G (UJ,k) ~O. (6.98)

For a wave in a bar governed (cf. eq. 6.9) by the equation


IE
a~/-,
p
this relationship is linear

d .- a2 k 2 ~ 0 or UJ~ak.

For a wave in a beam (cf. Chapter 9) governed by the equation


6.6 Wave Motion 157

a = lEI
pA '
this relationship is nonlinear

if - cJ k 4 = 0 or

Systems, for which the (W, k )-relationship is nonlinear are referred to as


dispersive, and the relationship itself is called the dispersion relation.

The ratio w/k = c is referred to as the phase velocity, while d wldk = egis referred
to as the group verocity. In non-dispersive systems the two velocities coincide. It is
said further, that energy propagates with the group velocity.

Since the phase 0 is defined as

0=kx-w{ (6.99)

in a spatially one-dimensional problem, it follows

00
k w = -- (6.100 a, b)
Jt
and
dk iJw
-+-=0, (6.101)
Jt OX
which can be interpreted as a statement of conservation of waves. In two or three
dimensions, conservation of waves requires (consistency relations for the
existence of 0)

ok; + oW = 0 (6.102)
Jt OX;
and
iJk; iJkj = 0 . (6.103)
oxj OX;
We consider next a slowly varying wavetrain of the form

cf> = Re (Ae;8) = a cos (0 + '1/), (6.104)

where a = ,t4 j, '1/ = arg A are both slowly varying and apply this first to the
energy equation for a bar (cf. Section 6.2), cf> --+ u. Equation (6.13 b), after
devision by iii , is rewritten as

: (11/ 1cJ + It 12) + ~ (- a 2 ltlt I) = 0 , (6.105)

E EA
where a 2
P m
158 6 Elastodynamics

The first parenthesis above contains the energy density, and becomes approximately

wheras the second parenthesis contains the energy flux, and becomes approximately

The approximation consists in neglecting terms with a and iT.


We consider next average values over one period. The average of the sin 2 function
is J/2 and we have

;r, ~ p f cl k 2 )a 2, (6.106 a)

, ; / - 0' l cl wka 2 , (6.106 b)


2
where ;r and ./-are the averaged energy density and the averaged energy flux,
respectively.

Recalling that UJ ., ak we observe

(6.107)

which turns out to be a general result and holds even for dispersive waves, when
a a (k), and the average energy relation also holds in general
e7tK
_. ~ - e7 . ..
(a?')~ O. (6.108)
e7t dt
The averaged Lagrangian density for the bar may be written as

;1' " ~2 (d - (i k 2 .Ja 2 (6.109)

and in terms of derivatives of the phase 0

;1'
.'.
~. ;1' ( 0 " 0/ a) (6.110)

or, for a more general system

;1'
. ~ fJ'
- (- 0 ,t' 0 ,,' a) (6.111 )

Since fJ' does not depend on derivatives of a, the Euler-Lagrange equations, cf.
equation (1.65), for this system are
e7fi'
o (6.112)
e7a
and

(6.113)
6.6 Wave Motion 159

We can return to the variables UJ, k; and a and write the above equations as

-
cY =0
oa
and

(6.114)

The latter equation is referred to as the conservation of "wave action". It has to be


supplemented by the consistency relations (6.102) and (6.103).

In linear problems, g must always be of the form

(6.115)

where

(6.116)

is the dispersion relation. Thus d <:J. Ida = 0 is in fact this relation.

The energy-balance equation for a wave can immediately be obtained by


postulating the average Lagrangian

and considering the rate of change in time t


dg cY" 0[./'
- = - e + - e. (6.117)
dt ~
V>::J
00.oj oj

By partial integration and substitution of the Euler-Lagrange equation (6.113)


(conservation of wave action) for :;i, as well as recalling that 0 = - UJ and
e.oj = k.,J we obtain
~ ( UJ og) + !!:- (UJ 0[1' - [./) = 0 , (6.118)
dx. ok. dt OUJ
J J
which states that the time rate of change of the Hamiltonian is balanced by the
divergence of its fluxes (in physical space). We can also establish, in a similar
fashion, the "wave momentum" equation by calculating
dg cY. 0[./
- = - e
+ -
';.:>0
e... (6.119)
dx; ~ • V'O,j
olJ

Again, by partial integration and making use of the conservation of wave action,
equation (6.113), we obtain

~ Ik CJ[J! _ fl'S ) _!!:... Ik 0[/) = 0 (6.120)


dx. ('; Ok. ij dt (' i OUJ .
J J
160 6 Conservation Laws in Elastodynamics

Here kj off/ow is the wave-momentum density vector in the direction k.I with
magnitude energy density divided by the phase velocity, as can be shown (cf.
Whitham, 1974).

The formal appearance of the equations (6.120) and (6.118) suggests to introduce
the (4 x 4) Eshelby tensor for the (k j , co) formulation in analogy to (6.24). If we
introduce

(6.121 a)

off
b ti = - kj oW ' (6.121 b)

the wave-momentum equation is rewritten as

b}I,}
... +b.=O.
11,1
(6.122)

By the definitions

bjl = - W -,
off (6.123 a)
ok.}

o~
bit = W - - ~, (6.123 b)
ow
the energy equation follows to be

..
b}I,} + bIt, I = 0 (6.124)

(no summation over time t). An equivalent formulation involving the Cauchy
stress tensor analogously to (6.31) is not apparent.

One of the most important results obtained in this brief discussion of wave
motion, is that, in nonuniform media, where the material parameters are functions
of x j and t, neither the energy, equation (6.118), nor the wave momentum,
equation (6.120), are conserved (they have to be supplemented by source terms),
but wave action, equation (6.114), is always conserved. For a very complete
treatment of wave motion, reference is made to Whitham (1974).
7

Dissipative Systems

7.1 General Remarks

As has already been mentioned, the neutral action (NA) method appears to be the
most useful tool in constructing conservation laws (and path-independent integrals)
for mechanical systems with dissipation (Honein et aI., 1991). One does not have to
worry about the establishment of a Lagrangian, but can consider directly the
governing set of differential equations. The inherent time-dependence of dissipative
systems necessitates the establishment of conservation laws in the space-time
domain (see Chapter 6 and Section 8.3).

The mathematical modeling of deformable (elastic) systems cannot forgo occasional


inclusion of dissipative mechanisms in order to get closer to reality, such as
exemplified by viscoelasticity. Some processes, such as heat conduction, are
described by the diffusion equation. In this Chapter, we consider conservation laws
for this latter equation first, and proceed to examine the non-linear wave equation
and viscoelasticity in succession.

7.2 Diffusion Equation

The governing equation for the three-dimensional diffusion problem is known to be


(cf., e. g., Butkov, 1968)

a = U,I - (ku)j = 0 , (7.1)

where U = U (XI' Xl ' x3 ' t) is the dependent variable, which can be the density of
the diffusing substance or the temperature in heat conduction problems. As usual,
162 7 Dissipative Systems

Xi are the Cartesian coordinates and t is the time. Further, k k (x) IS the
coefficient of diffusion or the coefficient of thermal conductivity.

To establish conservation laws valid for this system by the NA method, we have
first to specify the dependence of the characteristic I which we take to be

I = I (Xi' t, u , u, ' u,,) , (7.2)

where i ranges obviously from 1 to 3 in three dimensions. The condition for the
existence of a conservation law of the usual form

(7.3)

requires the satisfaction of the equation

E (fA) = 0, (7.4)

where, as before, E ( ) is the Euler operator acting on the expression in


parentheses.

In solving this equation for I, it was found (Honein et aI., 1991, Chien, 1992) that
it has to be of the form

I = I (Xi' t) (7.5)

and that it must satisfy

(7.6)

With the characteristic that depends only on the independent variables, we already
know that we shall obtain a conservation law in physical space. For any I which
satisfies its governing equation (7.6), the associated conserved currents are

Pi = k (Ii U - lu) , (7.7 a)

Pt =Iu, (7.7 b)

leading to Pi.i + p t .t = 0.

In the special case of one dimension and for constant k,f (XI' t) has to satisfy

kill +1,=0 (7.8)

with the corresponding currents PI and P, being

(7.9 a)

P, =Iu. (7.9 b)
7.3 Non-Linear Wave Equation 163

In particular, one such f is found to be (Chien, 1992)

f = eAt jA COS (If Xl) + B sin (If J.


Xl) (7.10)

Here A, B and A, are arbitrary constants and the corresponding currents PI and P,
are

(7.11 a)

P, = eAt jA cos (If Xl )+ B sin (If x)Ju (7.11 b)

An interpretation of this conservation law in physical terms has not been found.

7.3 Non-Linear Wave Equation

The simplest non-linear wave equation (cf. Whitham, 1974) in one dimension,
which describes a special case of shock wave motion may be written as

U t + c(u) U,I = 0, (7.12)

where the propagation speed c(u) is taken to be the negative of the local
disturbance u = U (Xl' t) for shock wave motion.

Here we take

~ = U t - UU,/ = 0 (7.13)

and conservation laws have to be found from

For

f = f (Xl' t, U) (7.14)

we find

It - ulI = 0 (7.15)
164 7 Dissipative Systems

and a solution is

(7.16)

where n is any real number except - J, -2 and 0 and C is an arbitrary constant.

This leads to the corresponding conservation law

Po + p,., = 0

with

-C (_t_ ll " '2 +~1l"'1) (7.17 a)


n+2 n+J

P, = C (_t_ ll" '1 + Xl II ") . (7.17 b)


n + 1 n
A physical interpretation of this conservation law is again not obvious and has not
been given yet.

A two-dimensional problem can be treated similarly on the basis of the governing


differential equation

.1 = u" - uU I - UU,2 =0 (7.18)

with the solution

(7.19)

where n is any real number except - J, -2 and 0 and A is an arbitrary constant.

The corresponding conservation law for this f has the currents

PI = -A !_t_
I~ +2
,2 u" + ( Xl
2 (n + 1)
+
2 (n
X2
+ 1)
)u" '1/, (7.20 a)

P2 = -A !_t_ u " '2


n +2
+ (XI
2 (n + 1)
+ x2
2 (n + 1)
)u" '1/, (7.20 b)

P3 = A
! -t- u"
n + 1
'1 + (_I
X
2n
X )u" /
+ ~
2n
. (7.20 c)

Again, a physical interpretation of this conservation law is not apparent.


7.4 Viscoelasticity 165

7.4 Viscoelasticity

One-dimensional linear viscoelasticity

We consider one-dimensional linear viscoelasticity on the basis of the Kelvin-Voigt


model first (Chien et al., 1993 a). In the absence of body forces, the stress equation
of equilibrium is a J = 0 and the relation between stress a and derivatives of
displacement U is

a = E u,l + TJU,1' . (7.21)

Here E is Young's modulus and TJ the viscosity coefficient of a Kelvin-Voigt


element depicted in Figure 7.1.

Fig. 7.1: Kelvin-Voigt model

It is convenient to introduce a "potential" cP as

cP =Eu +TJU, (7.22)

and the governing equation a is then

a = cP,1l =0 . (7.23)

The characteristic f of the conservation law is taken to be

f = f (cP , cP,l' cP,,) (7.24)

and the condition for existence of conservation laws becomes

E (fa) = 3.-
Ot/J cP,11 - (-.!!L
Ot/J cP,11 ).1 - (~cP
etI> ,11 (II + f,11 = o. (7.25)
,1 ,I

Since the only unknown in the above equation is the characteristic f which
depends on cP, cPJ and cP) , it follows that all coefficients of second and higher
order derivatives of cP in this equation must be set equal to zero independently.
The resulting set of equations is as follows
166 7 Dissipative Systems

coefficient equation
02!
4>.11 4>./1 -2 =0, (7.26 a)
04>,1
2 02!
4>, It - =0, (7.26 b)
2
04>,1
02!
4>,1t 2 4>]=0, (7.26 c)
eX/> 04> 1
02! 02!
4>.11 2~- 4>,I + 4>] =0, (7.26 d)
04> 04> dP,l 04> 04>,]
02! 2
leaving 04>2 4>,] =0. (7.26 e)

After some mathematical manipulations, the solution of the above set of equations
is found to be

(7.27)

Since t and !2 are arbitrary functions of 4>.1 , it is permissible to rename them


as

t (4)) = - gil (<1>.1) , (7.28 a)

!2 (4).1) = hi (<!» , (7.28 b)

where ( Y denotes differentiation with respect to the argument. Defining and t


F in this way will allow the conservation laws delived below to have a simpler
appearance.

With

(7.29)

and.:l = 4>,ll it is now possible to construct divergence-free expressions,


Po + Pt.t = °,
out of the product j~. The resulting currents PI and Pt are

(7.30 a)

(7.30 b)

Since the stress, a = 4>.1 , of this system is constant with respect to Xl' the
conservation pertaining to only h (4),1) i:- 0,

(7.31)
7.4 Viscoelasticity 167

implies that any function of the constant stress is a constant with respect to XI ,
where Al (t) is an arbitrary function of time.

The conservation law corresponding to only C "I:- 0 can be rewritten as

(7.32)

where Air) is an arbitrary function in t . In tenns of stress and displacement,

(7.33)

This relation expresses the dependence between stress, displacement and velocity
which can be obtained by integrating the constitutive equation of the system and
taking into account the equilibirum equation.

In terms of stress, the currents that pertain to only g (4).1) "I:- 0 can be written as

P, " - g'lu) / u, dx , (7.34 a)

P, = g (a) , (7.34 b)

where g is any function of stress.

If one chooses g (a) = a 2 /(2 E) and splits the stress a into an elastic component
a, = Eu,I and a viscous component a,/ = 'T] u 1r , i. e.,

a = a, + a,., (7.35)

and, furthennore, introduces the strain rate as

E) = U,I, , (7.36)

then the conservation law P/,/ + P" = 0 can be written as


2
d ae 2
- (-) = aE - 'T](E) , (7.37)
dt 2£ ,I .1

which states that the time rate of change of the elastic strain-energy density a; /
2E
is equal to the rate of work aE,1 done by the tractions minus the energy dissipation
a,., E" = 'T] (Ei . This is a fonn of the statement of the principle of conservation (or
rather balance) of energy (first law of thennodynamics) for the mechanical system
at hand. By choosing a different function graY, other conservation laws might be
constructed.

Two-dimensional viscoelasticity

In two-dimensional viscoelasticity, the stress equations of equilibrium are (aj;,j = 0)


168 7 Dissipative Systems

UIJ.J + UU.2 = 0, (7.38 a)

Unl + U 22 .2 = 0. (7.38 b)

For linear viscoelasticity based on the Kelvin-Voigt model, the stress components
are related to the normal strains in the Xr and xr direction (EI/' E22 ) and the shear
strain (E/2) as well as to the strain rates cl/,1 ,C22,1 and E I2 ,I under plane strain
conditions as

(7.39 a)

Un = (A + 2J.t) E22 + A EI/ + (a + 213) c22,1 + a EI/,I' (7.39 b)

(7.39 c)

where A and J.t are the Lame constants, a and 13 the viscosity coefficients of the
system.

The two governing equations for two-dimensional linear viscoelasticity in terms


of displacements are

&1 = (A + 2J.t) u/,/I + JIU 122 + (A. + jt) U2.12 +


(7.40 a)
+ (a + 213) U/,lll + f3U/.221 + (a + 13) U2 ,/21 '

+ (a + 213) U2 .221 + f3u 2 .1JI + (a + 13) UU21 . (7.40 b)

As before, the condition for existence of conservation laws by the NA method is

(7.41)

Since &j contain third derivatives in Ui , in order to evaluate this condition of


existence, one needs to compute the third total derivative of fj. If one assumes
characteristics that depend on Xi ' t, ui and also on derivatives of Ui , calculating
the third total derivatives of fj would be viltually impossible in the absence of
advanced computing devices with large memory capacity. Due to this difficulty in
evaluating equations (7.41), the general solutions may be found heuristically.

By restricting the dependence of J; to Xi and t only, the condition for existence of


conservation law by the NA method, equation (7.41), requires that

1/,// + 11.22 = 0, (7.42 a)

12.11 + 12.22 = 0, (7.42 b)

(7.42 c)
7.4 Viscoelasticity 169

where A 3 (t) is an arbitrary function of t. TIle corresponding conseIVation law is PJ.J +


P2.2+ Pt,t = 0 with the currents

(7.43 a)

(7.43 b)

P, = 0, (7.43 c)

where hand 12 are functions satisfying equations (7.42).

If h ,12 are taken to be constant, the above conseIVation law expresses the
conditions of equilibrium for the system. Due to the absence of a time current P"
equations (7.43) provide path-independent integrals in material space which might
be useful in the numerical analysis of cracks and defects for two-dimensional
viscoelastic materials.

A conseIVation law that relates to the dissipation of elastic energy for two-
dimensional linear viscoelasticity can be constructed if one considers a special
case where the Lame constants (A, JI) and viscosity coefficients (a, 13) are related
by

Jla = Af3 ' (7.44)

which implies that

r = ~ =i (7.45)
13 a
One solution for existence of conseIVation laws in this special case is given by

(7.46 a)

fi = A [ru2,' + 1l2,,) , (7.46 b)

and the corresponding currents are

PJ = A [( r Il Lt + Il J ,II ) all + (r 1l2" + Il v ,) ad ' (7.47 a)

+ 13 [( r Uu + Il u i + (r Uv + uvi I + (7.47 c)
170 7 Dissipative Systems

+ 2a (y u" + u",1 (y u" + u,.,,1 + 2iJ (y u'" + uwl (y Uu + u,.,J I.


Denoting the elastic energy of the system as

we =W/,,=/l=o, (7.48)

where W is the strain energy density given by

(7.49)

the conservation law for this special case of JHX = Af3 can be written as

~ [we] = all e1l,t + a 22 e22" + 2a12 e12,,- (a + 2f3) (e;1" + e;2,,) -

(7.50)

which expresses the balance between dissipation of the elastic strain-energy density,
the rate of work done by tractions and the strain-energy density dissipated for this
special two-dimensional viscoelasticity problem,

The balance law for the special two-dimensional viscoelasticity problem where
fIa = A f3 , equation (7.50), can be verified to hold also in the general case without
placing any restraint on the Lame constants and viscosity coefficients. The
conservation law that would yield this result is found to have the currents

(7.51 a)

(7.51 b)

P, = - A /2 we + (a + 2f3) (ul./ UI./' + U 2 .2 U 2 ,2t) + a (u2.2 UlJ + U/,/' u2.2,) +

(7.51 c)

Within the framework of the Neutral Action method, equations (7.51) can be
obtained if one modifies the condition for existence of conservation laws as given
by equation (7.41) to

(7.52)

in which case the corresponding characteristics are

(7,53 a, b)

J~ = u2J ' (7.53 c, d)


7.4 Viscoelasticity 171

The C' integral

Fracture at high temperatures becomes a time-dependent process, since most


materials behave viscoelastically at elevated temperatures. The aim of this
Subsection is merely to introduce the analogue to the J integral, the so-called C'
integral, without providing the complete background of the theory. The interested
reader is refered to, e. g., Riedel (1987). Under steady-state creep conditions, it is
common to introduce a creep potential W' with the constitutive equation
OW' OW'
a .. = -~ =-- (7.54)
IJ {]i..
IJ
OttJ,I..
and the strain rates given by
. 1· .
E. = -
lJ 2 (uj,1.. + u.I,}.) (7.55)

(dots indicate differentiation with respect to time). As in elasticity, the


constitutive equation (7.54) does not require linearity of the material response.

The creep potential W' is supposed to be a function of the strain rates only, i. e.,

(7.56)

Therefore, the elastic response of the material is neglected in comparison to creep


effects, W' does not depend explicitely on space and time coordinates and the
constitutive law does not take history-dependent effects into account, i. e., the
state of stress a jj at a point is uniquely defined by the actual strain rate i j . at that
point. Thus, the analogy provided by Hoff (1954) is applicable. It states ~ that the
stress distribution in a body under a given load would be the same if the strain
tensor Eij is replaced by the strain rate tensor iij' It follows that the stress
distribution under static loading is time-independent in the steady-state creep
regime

(7.57)

If we neglect body forces qj in the following, the homogeneous equations of


equilibrium are

a'),f
... = 0 . (7.58)

The C' integral is derived by applying the vector operator grad to the creep
potential and performing rearrangements as described several times earlier in this
text

grad (W') = W:i'

oW' . OW'
=--u
Oi' j,ki
+--
lj,k c1x i
172 7 Dissipative Systems

OW'
=Ukj Uj,;k + - -
dx;
OW'
= (ukj' U")k - Uk'k +--
),1 • J. U"j,1
dx;
The last two tenns vanish due to (7.56) and (7.58) and the conservation law follows
to be

(W' l)ki - uk'


J
U,.)
j,r.
k =0 . (7.59)

The divergence theorem, finally, leads to the C ' integral as

C: ~ !IW' S• . u kj ti,,) nk dA . (7.60)

s
In two dimensions, the first component C/is identified with C •

C' = fwn un 1 - kj k uj,J ds ' (7.61)

r
In viscoelasticity, C' plays the same role in assessing defects in solids as J does
in elasticity and plasticity, cf., e. g., Riedel (1987), Kienzler (1993).

Final remark

In concluding this Chapter on conservation laws in dissipative systems, it should


be mentioned that a relationship has been recently established between the NA
method and the concept of symmetry, cf. Chien et al. (1996). Symmetry, by
definition, is a transformation of an object into itself which leaves the object
invariant. The symmetries of an object is the set of all transformations which
leave the object invariant. The transformation is accomplished by means of
transfonnation functions. In the present context, two types of symmetries might
be distinguished, namely variational symmetries and symmetries of differential
equations. In a variational synunetry, the object which is left invariant is the
action integral A discussed in Chapter 1 (integral of the Lagrangian density over
material space). In symmetries of differential equations, the solution space of the
differential equations is left invariant. The symmetries of a set of differential
equations is the set of all transformations which transform solutions of the system
into other solutions, cf. Bluman & Kumei (1989).

In the 1996 reference cited above, it was found that in the solution space of a
given system of differential equations, the condition for the existence of
conservation laws, as imposed by the NA method, is adjoint to the condition for
symmetries of these equations. For definition of adjoint operators see Olver
(1993). In the space of solutions, characteristics of conservation laws as
established by the NA method are adjointly related to the characteristics of the
symmetry of the governing equations of the system of interest.
8

Coupled Fields

8.1 General Remarks

We employ the designation "coupled fields" with reference to physical phenomena


characterized by the coupling of elastic stresses and deformations to some other
physical manifestations describable mathematically as fields. In particular, we shall
consider the coupling of the elastic to the electric field, as it occurs in piezoelectric
and dielectric materials and, in a separate Section, the coupling of the elastic to the
thermal field. These latter two fields, by analogy, peltain also to a well known
model for porous media. In dealing with these rather different fields, we propose to
adopt two different approaches, in order to lend emphasis once more to the different
methodologies available to establish conservation and balance laws in material space
and thus provide an analytic framework for the analysis of defects and fracture.

When considering piezoelectricity (and as a special case elastic dielectrics), we


recall first the standard fundamental relations and, having discussed various
potentials, we proceed immediately to evaluating the gradient of one of them and in
this manner arrive at the Eshelby tensor for piezoelectricity and the J integral for
this coupled field. Without dwelling upon the details, we discuss a particular
example and make reference to some recent work, in which crack-extension forces
(i. e., energy-release rates) have been investigated for piezoelectric materials.

In the following Section, we present and discuss thennoelasticity in an entirely


different manner. We restate the basic (linear) relations of time-independent
thermoelasticity in one and two dimensions and apply the Neutral Action (NA)
method to obtain a path-independent J integral. Time-dependent thermoelasticity is
considered next in one and two dimensions. Application of the NA method leads to
several conservation laws whose physical meaning, however, is not always apparent.

Finally, it is pointed out that all relations of this Section on thermoelasticity, might
be interpreted as those of a fluid-saturated porous medium, performing an
appropriate change of notation.
174 8 Coupled Fields

8.2 Piezoelectricity

Piezoelectric materials are extensively used in many branches of modern


technology, including aerospace, automotive, medical and electronic industries, in
order to convert mechanical energy into electrical energy and vice versa. Devices
such as transducers, sensors, actuators, capacitors and resonators are finding an ever
growing field of applications in these industries. Typical examples of piezoelectric
materials are furnished by crystals like quartz and Rochelle salt, ceramics like
barium titanate (BaTi0 3 ) or lead zirconate titanate (PZD, as well as polymers, such
as PVDF. A common characteristic of all such materials is that they are anisotropic
and that they allow electromechanical interactions to take place. This means that
they deform when subjected to an electric field and, conversely, they induce an
electric field when subjected to mechanical loads.

A very poor feature of piezoceramics is their brittleness, which implies a low


fracture toughness. Significant advances have been made in the processing and
manufacturing of these materials during the recent past, but mechanical failure, as
well as electrical degradation due to the presence and growth of various defects
such as cracks and holes are still of great concern. For a recent review of defect
problems in piezoelectric ceramics, reference should be made to Sosa (1995).

In this presentation, we shall consider the simplest mathematical model of a linear


piezoelectric solid. In additon to the basic quantities of elastostatics, which are
recalled to be the stress G jj, the strain Cjj = 0z (uj .i + u;) and the displacement U i , we
have to introduce merely the electric field vector Ej and the electric displacement
vector D j •

The electric field Ej is the negative gradient of the electric potential c/J

E./ = - c/J.,/ (8.1)

and Gauss' law of electrostatics, in the absence of electric charge, states that the
electric displacement D i is divergence-free

D /,/.. = 0 . (8.2)

In the absence of body forces, the stress field G ji is also divergence-free, just as in
the purely elastic case

G ... =
jI,}
0 . (8.3)

Thus it is noticed that the above field equations are still not linking the
elastostatic to the electric field and this coupling comes about only through the
linear constitutive relations, which can be written as (cf. Tiersten, 1969)
8.2 Piezoelectricity 175

(8.4)

(8.5)

where cijkJ are the familiar elastic constants (measured in a constant electric
field), e/;jj = e kji the piezoelectric constants and E ik = E ki the dielectric constants
(measured at constant strain). In the most general case of anisotropy (triclinic
crystals), the piezoelectric material is described by 21 + 18 + 6 = 45
independent constants.

Alternatively, the constitutive relations may be given in inverted form as (cf.


Sosa, 1995)

(8.6)

(8.7)

where Sijkl is the compliance tensor measured at zero electric displacement, gkij is
the piezoelectric tensor and f3ik is the dielectric impermeability tensor measured at
zero stress. It turns out that depending upon the problem at hand, either one or
the other fonn of the constitutive relations may prove to be more convenient, just
as in the purely elastostatic case.

Most analyses of piezoelectric materials have been carried out for transversely
isotropic solids, such as PZT and barium titanate mentioned above. They possess
the symmetry of a hexagonal crystal class 6 mm (cf., e. g., Federov, 1968) and
are characterized by five elastic, three piezoelectric and two dielectric constants,
that is a total of ten independent material constants with strong piezoelectric
coupling.

The strain-energy density stored in any linear elastic piezoelectric material can be
written as (e. g., Pak, 1992)
1 1
W = "2 a ij £ij + "2 Ek Dk (8.8)

Strain £ij and electric displacement D k together define the physical distortion of
the material. Therefore, £ij and Dk may be chosen as independent variables and aij
and Ek as dependent variables. To eliminate aij and Ek from (8.8), the constitutive
relations (8.4) and (8.5) have to be modified. Solving (8.5) for Ek and eliminating
it in (8.4) yields

(8.9)

(8.10)

with

(8.11)
176 8 Coupled Fields

(8.12)

€ik = rik' (8.13)

and the inverse rik of the dielectric tensor € ik defined by

rim € mk = (j ik • (8.14)

With (8.9) and (8.10) the strain-energy density is calculated to be

w = W (Eij' DJ = ~ C;j1d Eij E kl - e:j Dk E ij + ~ f ik Di Dk • (8.15)

It becomes clear immediately that W is the thennodynamic potential with respect


to the elastic strain and the electric displacement and the constitutive relations
oW oW (8.16)
Osij Otl),1..

Ek = - -
ow (8.17)
oD k
are in accord with (8.9) and (8.10).

For our purpose, however, it is more convenient to consider Eij and Ek as


independent variables and aij and Dk as dependent variables. Analogies may then
be drawn between the two sets of mechanical and electrical variables. Stress a ij
and electric displacement D k are governed by the field equations (8.3) and (8.2),
respectively, whereas strain E ij and electric field Ek are gradients of the
displacement vector U i and the electric potential 4>, respectively.

The corresponding thermodynamic potential follows from a Legendre-


transfonnation and is designated as electric enthalpy density (e. g. Suo, 1992)
1 1
'F = W - Dk Ek = - aij Eij - - Dk Ek • (8.18)
2 2
(For a general definition of Legendre transformations see Arnold, 1989).
Eliminating aij and D k from (8.18) by the constitutive relations (8.4) and (8.5)
leads to

(8.19)

The alternative fonnulation of the constitutive relations is then


o'F
a').. (8.20)
Ott ..
J.'

o'F
Dk = - (8.21)
oEK
The governing equations for the linear elastic dielectric material are contained in the
preceding equations (McMeeking, 1990). In dielectric materials no coupling exists
8.2 Piezoelectricity 177

between elastostatics and the electric field, i. e., ekij and gJcjj in (8.4) - (8.7) are made
to vanish (Toupin, 1960). Thus the strain-energy density Wand the electric
enthalpy 'P are equivalent up to the sign to the term V2 € ik Ei Ek •

Application of a variational principle would allow the establishment of boundary


conditions. For the purpose at hand they are not relevant and may be found, e. g.,
in Pak & Herrmann (1986 a) or Khutoryansky & Sosa (1995). They still might be
briefly stated as: on each boundary or interface either the elastic displacement ui or
the (physical) traction aji nj has to be specified (l/j being the unit normal to the
element of area of the boundary or interface) as well as either the jump in the
electric potential cP or the jump in the normal component of the electric
displacement D i • An alternative formulation is given in Sosa & Khutoryansky
(1996).

To derive the Eshelby tensor of piezoelecn'icity we proceed in like manner as in


elastostatics. We calculate the gradient of the potential 'P, perform some integrations
by parts and make use of field equations already stated.

The governing potential (8.19) in tenns of displacement gradients uij and electric
fields Ek

(8.22)

is restated, emphasizing the implicit dependence on the coordinates XI' but


omitting an explicit dependence on XI which would imply a inhomogeneous
material (cf. Section 6.4).

Differentiation with respect to Xk yields


iJ'P iJ'P
'P,k E. . (8.23)
ok . u·.I,jk
= --
I,}
+ --
iJE.
}
},k

Because of relations (8.20) and (8.21) this can also be written as

Since Ej,k = Ekj (because of Ej = - cP j and Ek = - cP.k)' partial integration yields

(8.24)

With (8.2), (8.3),

(8.25)

and rearrangement of (8.24) we are left with

(8.26)

and the Eshelby tensor of piezoelectricity bjk is thus obtained as


178 8 Coupled Fields

(8.27)

with

(8.28)

for a homogeneous material, i. e., no explicit dependence of the enthalpy density on


the coordinates in the material. For inhomogeneous materials, source terms are
present on the right-hand side of (8.28). The derivation, then, proceeds along the
same line as outlined in Section 6.4.

By the usual divergence theorem, we immediately obtain the path-independent Jk

I
integral for these materials, with the corresponding bjk , as

J, c hi'.i dA . (8.29)
s
For nonlinear forms of J k , reference is made to Pak & Herrmann (1996 a), where
it is also shown that the above J k is indeed the total energy-release rate for a
defect being translated in the xcdirection. Recent developments and possible
directions of further research concerning piezoelectric ceramics are discussed in
Gao et al. (1997).

As an example, a crack of length 2a embedded in an infinite lead zirconate


titanite material (PZT-5H) and subjected to remote tensions a~ normal to the
crack (mode I) and a remote electric field E:" was considered by Pak (1992) (see
Figure 8.1).

---,.._===;=:==_;----_x j
-=- E'"

.. a a
~'

Fig. 8.1: Crack orientation and loading

The loading direction and the poling direction coincide. For details of the
transversely isotropic material behavior and the relevant material constants,
reference is made to Pak (1992) where further literature on piezoelectric material
properties may be found. By use of an integral fonnalism, the energy-release rate,
i. e., the crack-extension force was obtained numerically for PZT-5H resulting in
8.3 Thennoelasticity 179

a OO2 a OO E OO E oo2
J = a'TT ( - + 2a - -) (8.30)
E· JE· {3* {3*
with

E' = 102 GPa ,


2
(3' = 10.4 . 107 V ,
N
a = 0.64 . 10 -2 .

The critical crack-extension force of the material under consideration is about 10.0
Nlm (cf. Pohanka & Smith,1988). In the absence of electrical loads, a crack of
length 20 mm will undergo unstable crack growth at an applied far-field load of
a; f¥ 4MPa. The applied far-field electrical loads are typically of the order of
E'" = ± 2.0 ·10 5 VIm. It turns out that the energy-release rate given by (8.30) which
predicts that an electric field applied in either direction inhibits cracking, contradicts
experimental results, cf. Park & Sun (1995). The critical fracture load of cracks
perpendicular to the poling direction in simple tension and three-point bending of
PZT- 4 specimen was lower for a positively applied electric field and higher for a
negatively applied electric field. Cracks parallel to the poling direction were not
significantly affected by the electric field.

An attempt to bring the theoretical predictions into qualitative agreement with such
experimental results was undertaken by Gao et al. (1997), They argued that
electrical yielding may have a significant effect on the crack-tip energy-release
rate. A drastically simplified piezoelectric model (involving only three material
constants by contrast to the customary ten) pennitted the calculation of local and
global energy-release rates in closed form, with the result that only local rates show
agreement with experiments. It was pointed out that additional studies with fully
anisotropic and nonlinear models are required to further clarify the fracture behavior
of piezoelectric solids under combined mechanical and electric loading.

8.3 Thermoelasticity

Thermoelasticity represents another example of a coupled field, making the elastic


and the thennal fields interdependent. By conu'ast to piezoelectlicity, however, time-
dependent (diffusional) thermoelasticity is a dissipative continuum and therefore
even time-independent thermoelasticity will be discussed on the basis of the Neutral
Action (NA) method (Chien & Hemnann, 1996). In the latter model the temperature
increment () is determined first from Laplace's equation (harmonic problem) and this
() plays then the role of a source tenn in the elasticity equation (cf. Section 6.4).
180 8 Coupled Fields

The linearized theory of coupled thermoelasticity is given by the equation of


equilibrium in tenns of displacements

rIJ u·',~ + (A + r
H IJ) ,e·l , {39
- 1 =0 (8.31)

and the equation of heat conduction

(8.32)

These equations, which may be found, e. g., in Biot (1956), describe the time
history of the defonnation and thennal fields. The indices i and k range over the
spatial variables (Xl' X2) in two dimensions, t is the time and it is recalled that
repeated indices imply summation. In the above equations the symbols have the
following meaning:

Uj is the component of displacement in xi-direction,

T is a constant reference temperature,

o = 0 (x j• t) is the increment of temperature above reference temperature T,

e = U jj is the dilatation,

A. • Ji are Lame's constants of elasticity,

{3 = (3A. + 2Ji) a"

at is the coefficient of thennal linear expansion,

k is the coefficient of heat conduction,

c is the specific heat per unit of volume.

The constitutive law for stress is given as

(8.33 a)

(8.33 b)

(8.33 c)

and the heat absorbed per unit volume is

h = cO + T{3e . (8.34)
8.3 Thennoelasticity 181

Time-independent thermoe/asticity

One-dimensional problem
We consider first time-independent thermoelasticity and a one-dimensional problem.
There are two governing equations ..:1] and ..:12

..:1] = (.Il + 2Jl) u' l1 - f30.] = 0, (8.35 a)

(8.35 b)

To obtain conservation laws for this system, the NA method requires that

Assuming the characteristics II and 12 to be linear functions of u, u.J ' and 0.] the
solution to the above equation is found to be (Chien, 1992; Chien & Herrmann,
°
1996)

(8.36 a)

+A 2 [(.Il +2Jl)u,1 -f30J +A 3 0./, (8.36 b)

where all A;'s are arbitrary constants.

Having solved for the characteristics of conservation laws, II and 12 , the


conserved current of the system (PI = constant) can be constructed directly from
the product of J; ..:1; and the results are as follows:

For only Ao ::F- 0 ,

p] = 21 (.Il + 2Jl)
2
Lt,] - f3 uo .J . (8.37)

With the elastic strain energy of the system we 112 (.Il + 2Jl) U,; this
conservation law can be rewritten as

PJ.J = u-:; - f3u.J 0.1 - f3 u0 .JI = 0. (8.38)

Since 0,11 = 0 by equation (8.35 b), this law implies that the spatial change of the

°. W;
elastic strain energy is equal to the product of the strain u. I and the
temperature gradient 1 scaled by the factor f3.

For only AI ::F- 0,

p] = (.Il + 2Jl) (OU,I - un,) - 1 f30 2 • (8.39)

This conservation law relates to a balance of displacement, strain, temperature,


and temperature gradient whose physical meaning is not apparent.
182 8 Coupled Fields

For only A2 "# 0 ,

P, = (.J + 2Ji.) u" OJ - f30 OJ = 0" a" . (8.40)

This conservation law implies that the product of the stress and the temperature
gradient is a constant. This result is expected since both the stress and the
temperature gradient are themselves constants by equation (8.35).

For only A3 "# 0 ,

1 2
P, (8.41)
= -
2 0.I
Since the temperature gradient OJ is a constant by equation (8.35 b), any function
of OJ will also be a constant. This conservation law is an example of such a
function. Note also that since the constant A3 does not appear in the characteristic
II, this balance law can be valid for systems with .:1, "# O.
Two-dimensional problem
We proceed next to a two-dimensional problem. The governing equations for a
two-dimensional elastic body under the influence of a time-independent
temperature field are given by equations (8.31) and (8.32) as

(8.42 a)

(8.42 b)

(8.42 c)

To construct conservation laws for this system by the NA method, we require that

Assuming the characteristics.t: to be linear functions of u, • uJ.J • uJ.2 • u2 ' u2,J ,


u2.2 ,0, OJ and 0.2 ' the solution to the above equations is found to be

11 = Ao
I (.J + 3Ji.) uJ.J + (A + 2Ji.) u2 .2 - f3 .J
A
+ 2j.L 0 /
+ j.L
+ A, Ji. u2,J +

(8.43 a)

h =
8.3 Thermoelasticity 183

(8.43 b)

Ao f31 JL2 U1 + A 1 -
f3 JL2
- '2
't + A
2
f3 I/.
r-
A + 2JL u +
2
A+JL A+JL A+JL

(8.43 c)

where all A;'s are arbitrary constants.

The corresponding conservation laws PJ./ + P2,2 = 0 constructed out of the product
/; 4; are as follows

For only Ao :t- 0 ,

(8.44 a)

For only Ai :t- 0 ,

PJ = (A + 3JL) jU 2,2 a 12 + A U 2,2 U 1,2J~ (A + 2JLf U 2,2 u1,2 +


184 8 Coupled Fields

For only A2 ::/;. 0,

For only A3 ::/;. 0,

+ _f3_1 f3 J.L (}2 + J.L (A + 2J.L) (It] ()] - () ltl,l)} , (8.47 a)


A+J.L 2 '

(8.47 b)
8.3 Thermoelasticity 185

For only A4~ 0,

(8.48 a)

(8.48 b)

For only As~ 0,

(8.49 a)

(8.49 b)

For only A6~ 0,

l(l -lrl
2 ,1 2 ,2' (8.50 a)

n,1 n,2 . (8.50 b)

For only A7 ~ 0,

n,1 n,2 ' (8.51 a)

ln
2 ,2 -lol
2
2,1 (8.51 b)

Conservation laws listed above are valid independently or in combination with


each other. For A6 , A7 ~ 0 , the corresponding conservation laws relate to
balances of the temperature gradients which arise from the harmonic nature of the
temperature field. Note also that since the constants A6 and A7 do not appear in
the characteristics J; and h, these balance laws can hold for systems where A J
and A2 ~ O.

For A 4 and As ~ 0 , the conservation laws express balances of stresses,


displacement gradients and temperature gradients whose physical meaning is not
apparent.

The laws corresponding to Ao ' AJ ' A2 , and A3 ~ 0 , which express various


balances of stresses, displacements, displacement gradients, temperature and
temperature gradients, can be combined in a special way which will yield balance
laws leading to modified J integrals for this thermoelasticity problem. In the
absence of thermal effects, these modified J integrals, derived below, will reduce
to the J integrals for elasticity.

Denoting the elastic strain energy for this two-dimensional problem as

(8.52 a)

with
1
, = -2 (uj,I.. + uI,J. .)
E,;i (8.52 b)
186 8 Coupled Fields

and setting Ao = A3 1 , the corresponding conservation law becomes


A + 3JL

P1=- W e + U 1,1 U11 + U2,112


U + l ~ ,,2 +
rr ~ fU
\'] ()
1 -
() U )
11'
2A+JL A+JL '
(8.53 a)

(8.53 b)

Also, by setting AJ = A2 - -1 - , the correspond'mg conservatton


. Iaw WI'II
read A + 3JL

U2,2 u12 + u],2 un + ~


A+JL
(u 2 () 2 - ()
'
Un) ,
,
(8.54 a)

P2 =- we + U2,2 U22 + U1,2 U12 + ~ A13; JL 0


2
+ Af3~JL (U 2 0,2 - 0 U Z /

(8~54 b)

Since we have div P = PJ ] + P22 = 0, and by the divergence theorem,

/ div P d A = / P; II; ds , (8.55)

s r
where dA is an area element on the surface S, and ds is a line element on the
boundary r of S with outward nonnal II; , we can rewrite equations (8.53) and
(8.54) as path-independent integrals of the form

- l2 -.-L
A + JL
1Jl) ~.
lj
- U'k Uk' +
J. ',1
~
A + JL
(0 U .. -
',J
u.
"
O)Jn ds = 0 ,
J

(8.56)

with i = 1 corresponding to equation (8.53) and i = 2 corresponding to equation


(8.54), except for the sign. The integrand of (8.56) represents the Eshelby tensor
of time-independent thennoelasticity and is the same for three-dimensional
problems, provided the range of indices i, J, k is extended to 1, 2, 3 (Kienzler,
1993). The Eshelby tensor might also be derived by applying the gradient to the
appropriate Lagrangian (cf. Eischen & Herrmann, 1987).

In the absence of a thermal field, setting 0 = 0 , it can be readily verified that


equation (8.56) is identical to the J integral for elasticity, It is also of interest to
note that this result, equation (8.56), is almost identical to that given by Gurtin
(1979), which he presented, however, without showing how it was obtained
(Gurtin's equation (2) contains a flaw concerning the sign of the J-l f3f( A + J-l) -
tenn).
8.3 Thennoelasticity 187

Time-dependent thermoelasticity

One-dimensional problem
The governing equations for one-dimensional thermoelasticity with a time-dependent
temperature field are given by equations (8.31) and (8.32) as

(8.57 a)

4 2 = kO Jl - cO., - Tf3u. 11 = 0 . (8.57 b)

To obtain conservation laws for this system, the NA method requires that

Assuming the characteristics 11 and 12 to be functions of x, t, U, u. 1 ' up 0, 0'1 and


0" , the solution to the above equation is found to be

h = - gl (t) I(T f3 + C (A f3+ 2fJ) )u. , - k 0 1


. j - -
C
f3
dg 1
-
dt
x all +

+ g2 (t) gO (all) + g3 (x, t) , (8.58 a)

h = gl (t) all + g4 (x, t) , (8.58 b)

where gl (t) and l (t) are arbitrary functions of t , l is an arbitrary function of


stress all' and g3 (x, t), l (x, t) are functions satisfying
3 4
(A + 2/L) g11
, - T f3 gIl
, = 0 , (8.59 a)
3 4 4
f3!J,1 + k g,ll + C g.1 = 0 . (8.59 b)

The corresponding conserved currents are given by

4 4 4
+ g (x, t) kO,1 -!J,1 kO + T f3 8.1 Lt, (8.60 a)

P, = l £gl (t) (alll - g4 (x, t) h . (8.60 b)


2 f3
For only l (t) # 0 , the corresponding conservation law can be rewritten as

(8.61)
188 8 Coupled Fields

where we is, as before, the elastic strain energy for this one-dimensional problem.
This conservation law expresses, how the time rate of change of the strain energy
is balanced by the strain rate, temperature and stress. Thus it is a form of the law of
conservation (or balance) of energy.

For the balance law corresponding to l (t) "# 0 , this law merely expresses the fact
that any function of the constant stress l (aJ/) is also a constant with respect to
x. Note also that as l (t) does not appear in the characteristic J;, this balance law
is valid for a system that does not satisfy.12 = 0, and is thus partially trivial.

For conservation laws corresponding to g3 (x, t) and l (x, t) "# 0, these laws are
balance laws in physical space which express various balances of stress,
displacement, temperature gradient and the heat absorbed per unit volume. In
particular, the balance law corresponding to only l (x, t) being a constant A is of
the form

(8.62 a)

P, = - Ah , (8.62 b)

which indicates that the time rate of change of heat absorbed per unit volume is
equal to the space rate of change of k times temperature gradient, as required by
the governing equation,.12 = 0 , of this system.

Two-dimensional problem
The governing equations for two-dimensional thermoelasticity with a time-
dependent temperature field are given by equations (8.31) and (8.32) as

(A. + 2Jl) U 111 + /-L U 1,22 + (A + /-L) u 2,12 - f3 8 1 = 0 , (8.63 a)

(8.63 b)

k (8,11 + 0,22) - cO" - T f3 (u 1,lt + U 2,2,) = 0 . (8.63 c)

To obtain conservation laws for this system, the NA method requires that

Assuming the characteristics fl' j; and h to be functions of Xl ' x2 ' t, Ul , UJ,J , u 12 '
u 1,,, u 2 ' U2,!, u 2,2' U2,t' 0, 8,1 ,0,2> and 8,1; withf, andJ; depending only linearly and
quadratically on uJ,J, u 12 ' u v , u 2.2' 8, the solution to the above equation is found
to be
8.3 Thennoelasticity 189

A! k 0 _(T{3
+ A- + 2JL ,1 A- + 2JL

+A!k O_(T{3 c
+ -) U 2f } + g 5 (Xl' X2, t) ,
A- + 2JL ,2 A- + 2JL {3 ,
(8,64 b)

h = A full
j" ,
+ un -
, A-
{3
+ 2JL
O}+ g6 (Xl' X2, t) , (8.64 c)

where land l are functions of XI' X2 and t satisfying:

(8,65 a)

(8.65 b)

l is an arbitrary function of t, A is an arbitrary constant, and l, g5, and l are


functions of XI' X2 and t satisfying:
4 4 6
(A- + 2JL) gn + (A- + JL) g,12
5
+ JLg,22 - T{3g,lf = 0, (8.66 a)
5 4 5 6
(A- + 2JL) g22 + (A- + JL) g,12 + JLg,n - T{3g2f = 0 , (8.66 b)
4 5 6 6 6
{3 (gl" + g2) + k (gn + g22) + cg f = 0 . (8.66 c)
" ,

Having solved for the characteristics, the corresponding currents can be


constructed from the product f .:1;, and the results are as follows.

For only l , g2 "# 0 ,


190 8 Coupled Fields

gl (Xl' Xl' t) { -1 fi A. + 2J.1-J (UI,I + ul ) - {39/


3
+

+ /Ll (Ul,l - U1)1 fi A. + 2/L) (U 1.1 + Ul .l ) - {39/} +

+ gl (Xl' Xl' t) {/L (A. .1- 2/L)2 (UI,I + U2,2)2 (U 2,I - UI) +

+ 3/L
1 3 (
UI,2 - U2,I
)3
+

+ {3 /L (U I,2 - u2) 9 f (A. + 2/L) (U l •I + ul,) - {39/}, (8.67 a)

P2 g2 (Xl' X2, t) { 1fi A. + 2/L) (u 1,1 + U2.2) - {39/


3
-

- /L2 (U 2,I - UI.2/ fi A. + 2/L) (U I. I + U2,2) - {30/} +

+ gI (Xl' X2, t) {/L (A. + 2/L)2 (UI,I + U2,2/ (U 2.1 - U


v) +

+ -1 /L3 (U I2 - u21 )3 +
3 . .

+ {3 /L 9 (U I,2 - u 2) f (A. + 2/L) (UI,l + u 2) - {39/}. (8.67 b)

For only l "!- 0 ,

PI = ~ (alli + ~(a12i + all f/L U2.2 + /L U2. l - /L UI,2/+

(8.68 a)
8.3 Thennoelasticity 191

As l, land g3 do not appear in the characteristic h , these conservation laws will


be valid for a system with displacements and a temperature field that do not satisfy
equation A3 = O. In this sense, these laws do not represent the full system of
interest, because they are partially trivial. However, given any elastic body under
the influence of a time-dependent temperature field, these conservation laws must
hold along the solution of the system.

Thus, despite this deficiency, the above laws might still be useful in the analysis of
the present system. Their physical significance is not apparent.

For only A :F- 0,

P1 = !k{30,1 - (T{32 + C (A + 2/1.)) Ul,t) all +

+ f'{30,2 - (T{32 + C (A + 211)) U2 ,t) an +


+ 2k{3/-L (0,1 U2,2 - 0,2 U1,2) , (8.69 a)

f'{30,2 - (T{32 + C (A + 211)) U2 ,t) a22 +


+ f'{3(),l - (T{32 + C (A + 2/-L)) u1 ,j an +
+ 2k{3/-L (0 2 U11 - 0 1 U2 ), (8.69 b)

C (A + 2/-L)
f A +
2
2/-L
(ul,l
2
+
2
U2,2) + AU 1,1 u2,2 +

+!!:.- (U}2 + 2U 12 U21 + Ui1) -


2' " '
-{30 (U 1,1 +U2,2))+T{32 /-L I~ (U J,2 +U 2,2/-

- 2u 1,1 U2,2)+ ~ {32 c0 2. (8.69 c)

Similar to the conservation law corresponding to gIrt) :F- 0 in the one-dimensional


problem, equation (8.61), this conservation law can be written in a form that
expresses the balance of dissipation of elastic strain energy, strain rates,
temperature and stresses as
dw e
- - = E..
dt 'j,t
aj'.. + {3 0 Ekk,
t
• (8.70)

This can also be written in the form of a line integral being equal to the time rate
of change of a volume integral as

J{
r
a··'j !4{3o.
~ ,j
- (T{32 + C (A + 211) u.j,I))n.I + OJ u}... }n.,

s
192 8 Coupled Fields

+ Tf32 JL Ii2 u I,J.. u I,J.. + i2 u "J.. uJ,I.. - U ..


1.1
u ..
j,J
j+ i2 f32 c02} dA. (8.71)
For only g4, g5, l :t; 0,

PI = g4 all + g5 a12 + g6 kO,j - u j fi A + 2JL) g,~ + Ag,; J-


- U2 !JL (g,~ g,~) J Tf3g,~ uj - g,~ kO, (8.72 a)

P2 g5 a22 + g4 a12 + g6 kO,2 - u2 fi A + 2JL) g,; + Ag,~ J-


- u j IJL (g,~ + g,~) J+ Tf3g,~ u 2 - g,~ kO, (8.72 b)

P, = -l h . (8.72 c)

This law is a balance law in physical space. If l, land l are taken to be


constants, it will reduce to the governing equations of the system in conserved
form.

8.4 Mechanics of a Porous Medium

M. A. Biot (1956) has shown that the mechanics of a fluid-saturated porous medium
and of thermoelastic continua are isomorphic, i. e., the governing equations of an
elastic porous solid whose pores are filled with a compressible viscous fluid are
analogous to those of thermoelasticity and may be obtained from each other by a
change of notation. The constitutive equations for mechanics of a porous medium as
given by Biot are

a..IJ + asIJ.. = N E .. +S eS IJ.. - ll' pS..lj (8.73)


IJ

Here p is the fluid pressure, while a is

a=-fp, (8.74)

where f is the porosity. N, S, and ll' are material constants.

Substitution into the equilibrium equation

(8.75)

results in

NU'I, kk + (N + - ,ll'p.
S) e., 1 1
= O. (8.76)
8.4 Mechanics of a Porous Medium 193

In addition, a relation between fluid pressure, dilatation of the solid e and fluid
content € was postulated as

a = Qe + R€ , (8.77)

where € = Ui,i' with U i being the absolute displacement of the fluid.

The increment of fluid content of a porous element is

( = [(€ - e), (8.78)

which may also be written as

(8.79 a)

with

..!- = [2 (8.79 b)
M R
Darcy's law (which corresponds to Fourier's law of heat conduction) expresses the
proportionality between the relative fluid flow a(
fat and the pressure gradient P j
and may be written as

k] div grad p = k] p "" = o( (8.80)


,J] at
or
1 (8.81)
k] p.ll - M P,t - ae. 1 = 0 ,

where k, is the permeability coefficient.

Comparing (8.76) with (8.31), it is seen that the fluid pressure p plays the role of
the temperature increment O.

Equations (8.31) and (8.32) of thermoelasticity correspond to equations (8.76) and


(8.81) for a fluid-saturated porous medium employing the following change in
notation

porous medium ~ thennoelastic medium

N ~ fl

S ~ A

a ~ f3
k
k, ~ -
T
194 8 Coupled Fields

1 c
M T
Thus, conservation laws obtained for all cases of thermoelasticity in Section 8.3
apply immediately to mechanics of a porous medium with the same change of
notation. Interpretations of conservation laws for a thermoelastic medium can also
be analogously applied to a porous medium.
9

Bars, Shafts and Beams

9.1 General Remarks

Theories which fall under the general phrase "Strength-of-Materials" are to be


understood as an approximate mathematical description of the static (and dynamic)
response to various loadings of elastically deformable bodies whose one dimension
is considerably larger than the other two. If such a body, considered straight, is
subjected to loads along its axis, it is called a bar in tension-compression, if it is
subjected to twisting moments, it is called a shaft in torsion, and if it is subjected to
bending moments and/or transverse loads, it is called a beam in bending. Curved
bodies, i. e., arches and frames, may also be considered.

Theories of strength-of-materials were developed primarily during the eighteenth


century by Bernoulli and Euler and thus preceded the establishment of the theory of
elasticity, by approximately a century, by Cauchy, Navier and Lame.

It is a remarkable circumstance that defects in elastic solids, considered theoretically


first within the context of the theory of elasticity, may, to a certain extent, be
analized also on the basis of strength-of-materials. It may be noted that these
investigations were carried out some fOlty years after those on the basis of the
theory of elasticity.

In this Chapter, we first summarize the elements of strength-of-materials for bars,


circular shafts and beams and proceed to discuss conservation and balance laws of
these bodies in physical and material space. Some of the attendant derivations,
particularly for inhomogeneous bodies, are rather cumbersome, and are relegated to
Appendices A and B.

In the final Section of this Chapter, the approximate calculation of energy-release


rates for certain crack geometries and the associated stress-intensity factors is
discussed and, whenever possible, compared with results available in the literature,
based on the theory of elasticity.
196 9 Bars, Shafts and Beams

9.2 Elements of Strength-of-Materials

Bars in tension-compression

We consider a unifonn elastic bar, of unspecified length e


and cross-section A
which may be subjected to end loads N' , tv' and distributed applied axial loads n,
measured per unit length of the bar, cf. Figure 9.1.

a.)
Ln(x)
N 1
.....

b.)
A

Fig. 9.1: a. Bar under tension-compression


b. Deformed infinitesimal element (cross-section A) of a bar

Equilibrium of forces in the x-direction requires

(9.1)

where N is the resultant internal axial force and primes indicate differentiation
with respect to the axial coordinate Xl = X.

Let u be the displacement of any cross-section, then u/ is the strain E

(9.2)

Hooke's law is expressed as

N = EA E, (9.3)

where E is Young's modulus of the material. The product EA is called the axial
stiffness. The bar may be inhomogeneous because either E(x) or A(x) or both may
vary along the axis.
9.2 Elements of Strength-of-Materials 197

The strain energy Wand the potential of the axial loads V, per unit length of the
bar, are given, respectively, by

W = l EA U /2 1 N
2
(9.4)
2 -lEA'
v = - nu. (9.5)

Shafts in torsion

The basic equations of torsion of circular shafts are given by equations which are
analogous to those of bars. As in Figure 9.2, the shaft may be subjected to end
torques T I and T r, as well to torque per unit of length t, applied along the shaft.
The equilibrium of an element of the shaft requires

(9.6)

where T is the resultant of distributed shear in each cross-section and may be


called internal torque.

a)
TI~

b)
\
\
tdx
••
I

T..- _I-
~ T+T'dx
,
I

I
I
I

,"
dx
Fig. 9.2: a. Circular shaft under torsion
b. Deformed infinitesimal element of a shaft

The kinematic relation may be expressed as

r=RlY. (9.7)

r
where tJ is the rotation of the cross-section. is the angle of twist per unit of length
or the shear strain and R is the radius of the cross-section. The elastic law can then
be written as

T = G Ip r. (9.8)
198 9 Bars, Shafts and Beams

where G is the shear modulus and f p the polar moment of inertia of the cross-
section. Again, the shaft may be inhomogeneous, because G(x) or fix) or both may
vary along the axis of the shaft. The strain energy W and the potential V of the
external loads, per unit of length, may be written, similarly as for bars, as

w = l2 GI fj12
p
(9.9)

V = - tlJ. (9.10)

Beams in bending

The classical, or elementary, or Bernoulli-Euler theory of beams is different from


the theories of bars and shafts. While still one-dimensional in the sense that all
quantities depend on only one independent variable, namely the axial coordinate x,
the deformed configuration described by the deflection w(x), lies in a plane, because
w(x) is nomal to x. The beam, see Figure 9.3, may be subjected to end moments M
and M, and transverse forces Q' and Q', transverse distributed loads q and
distributed moments m, both measured per unit of length.

As is seen from Figure 9.3, equilibrium of forces and moments requires,


respectively,

QI = - q, (9.11)

M 1 = Q - m, (9.12)

where Q is the internal shear force and M the internal bending moment.

The kinematical constraint requiring plane cross-sections to remain plane and


normal to the central axis implies

(9.13)

where IfF is the angle of rotation of the cross-section. The elastic relation expresses
proportionality between the curvature ~ and the bending moment M, the factor
being the bending stiffness El and I the area moment of inertia. Thus

M = El ~ or M = - El W/~ (9.14 a,b)

It is a feature of the elementary beam theory that the shear force Q is not related
through a form of the elastic law to some kinematic quantity, but is, rather, to be
determined from the equation expressing equilibrium of moments.

This is different in the more advanced Timoshenko beam theory, which includes the
effect of shear deformation, thus uncoupling IfF and w~ which implies that cross-
sections, while remaining still plane, are no longer normal to the deformed central
9.2 Elements of Strength-of-Materials 199

axis. The loss of the constraint equation is "compensated" by an elastic relation


between Q, Wi and tft. In our work, we shall not be concerned with the Timoshenko
beam theory so as not to unduly increase the volume of the text presented. It should
be mentioned, however, that some investigations of conservation laws in
Timoshenko beam theory have been carried out (cf. Sosa, 1986).

a.
~m(x)
000000
q(x)

b. mdx

~
qdx
......---fJ-,A
f-------+
--- ~J: ---
w
w+w'dx

dx
Fig. 9.3: a. Beam under bending
b. Infinitesimal beam element (cross-section A) in the undeformed and deformed
configuration

The strain energy per unit of length W in elementary beam theory may be written
as

W =i E1 tf/2 =i E1w // 2 =.!.. M


2
(9.15)
2 2 2 EI
and the potential of the externally applied loads per unit of length Vas

V= -qw -m tft= -qw +mw~ (9.16)


200 9 Bars, Shafts and Beams

Thus, the Lagrangian function L = ~ (W + V) = L(x, w, w ~ w /J depends on


derivatives of second order and describes a one-dimensional grade two material or,
more precisely , a one-dimensional Cosserat continuum with constrained rotation
(Eshelby, 1976 b).

9.3 Balance and Conservation Laws for Bars and Shafts

We propose to establish all balance and conservation laws of interest to us in a


unified and comprehensive fashion by considering the derivative of the product of
the coordinate x raised to the k-th power and the potential energy density W + V.
Thus, we evaluate the derivative of the expression i' (W + V) and attempt to
determine the value of k for which conservation laws exist. We obtain in a first step

(9.17)

We start with bars in tension-compression and assume that EA and n are not
constant along the bar. The analogous results for shafts in torsion will be given at
the end of this Section. Substitution of the expression for Wand V given in
equations (9.4) and (9.5), respectively, and of the equation of equilibrium (9.1),
together with partial integration where appropriate, leads to the intermediate result

fk (W+V-Nul}/= -x k (~(EAYu/2-nit)+kxk-l (W+V-EAu/ 2) .

Further similar manipulations on the right-hand side of the above equation yield the
final result

- Xk b - kx k-1h -1 k (k - 1)X k -2 r - ~ k (k - 1) (k - 2 )X k-3EAu 2.


(9.18)
Here the abbreviations are introduced as follows

material force B = W + V - Nu I, (9.19 a)

material virial H = -l EAu / u (9.19 b)


2 '
second-order virial R = -l2 EAu 2 ' (9.19 c)

material loading b = 1 (EAr u /2 - n / u, (9.19 d)


9.3 Balance and Conservation Laws for Bars and Shafts 201

material virial loading (9.19 e)

second-order viria1 loading (9.19 f)

For k = 0, i. e., when considering the gradient of the potential energy density
W +V, it is seen that the balance law

(9.20)

is obtained, which embodies the balance of linear material momentum B,


corresponding to the balance of physical momentum given by equation (9.1).

The source term b vanishes if the stiffness EA is uniform and the applied loading
along the bar n is constant, and in this case the conservation law results

B = const. (9.21)

For k = 1, i. e., when considering the (first) scalar moment (virial), the balance
law is obtained

(xB - HY = - xb - h (9.22)

The source term vanishes if EA is uniform and, in addition, the bar is load-free
along its length, n = O. The associated conservation law is then

(xB - H)/ = 0 ; xB ~ H = const . (9.23)

For k = 2, i. e., when considering the second scalar moment, we obtain the
balance law

(x 2 B - 2x H + RY = - x 2 b - 2xh - r . (9.24)

The source term on the right vanishes under the same (no additional) conditions
as in the previous case, i. e., (EAl = 0 and n = 0, and the conservation law reads

(x 2 B - 2x H + R)/ = 0; x 2 B - 2x H + R = const. (9.25)

It is seen from inspection of the last tenn on the right-hand side of equation
(9.18) that it can not be made to vanish for k ~ 3. Thus no further conservation
laws using the scheme proposed here seem to exist. The number of balance laws,
however, is obviously infinite.

The three conservation laws (9.21), (9.23) and (9.25) agree with those given in
Chapter 1, equations (1.103 a - c), respectively, where the bar with constant axial
stiffness under end loading (EA = const.. n = 0) was taken as an example to
illustrate the application of Noether's formalism and the Neutral Action method.

There exist several other methods of obtaining the balance and conservation laws
above which should be mentioned for completeness. One would consist in suitable
202 9 Bars, Shafts and Beams

interpretation of the physical and material balance laws of plane elastostatics


(Giinther, I962), another by means of the virtual-work theorem (Kienzler &
Herrmann, 1986 a), by a variational principle with varying boundaries (Kienzler,
1986) and finally, by duality considerations, interchanging dependent and
independent variables in the Lagrangian (Rosel, 1986).

If one is interested in establishing conservation laws for nonuniform bars (EAl"# 0


under arbitrary axial loading n "# 0, the Neutral Action method provides the required
efficient tool. Due to fairly extensive derivations, this development has been
relegated to Appendix A. Here, for completenes, we list some results. For
inhomogeneous bars without distributed loading, e. g., EA "# canst, n = O.

(EA B/ = 0, (9.26 a)

(EA/~B-H/=O, (9.26 b)

(2EA IL rj-t )dxB -21-t H , ~r ~O, (9.26 c)

(9.26 d)

(9.26 e)

(The quantity V = EAu is the so-called virial, c£ Sections 1.4 and 2.3, and should
not be confused with the potential of external forces for which the same symbol
is used.)

With the stiffness distribution E(x)A(x) given explicitely, the integrals above can
readily be evaluated and integration constants are of no consequence, cf.
Appendix A. For EA = canst. the above set of equations (9.26 a-c) obviously
reduces to the set (9.21), (9.23) and (9.25). Equations (9.26 d) and (9.26 e) are
physical conservation laws discussed already in Sections 1.4 and 2.3.

For a homogeneous bar, e. g., EA = EA o = const., under constant external


loading n = no = const. we find (cf. Appendix A)

B/ = 0, (9.27 a)

jJ.Bx - H - -1 -no- ~no x 3 + 3 N.x 2 -6 vx)j/ =0, (9.27 b)


4 EA o

1 no
IBx 2 -2Hx +R - - - - (no x 4 + 4 Nx 3 - 12 Vx 2 )j / =0, (9.27 c)
8 EA o

IN + no x / I =0, (9.27 d)
9.4 Balance and Conservation Laws for Beams 203

(9.27 e)

Again, the first three equations (9.27 a-c) express the zeroth, first and second
order consetvation of material force and the last two equations (9.27 d and e)
express the zeroth and first order consetvation of physical force. To emphasize,
for bars, the five consetvation laws can always be established for arbitrarily
specilled functions EA =E(x) A(x) and n =n(x). As will be shown in Appendix B,
this is not the case in beam theory.

All the above expressions for bars remain valid for shafts, if one introduces a
change of notation according to the tension-torsion analogy

n ~ t, (9.28 a)

N ~ T, (9.28 b)

EA ~ GIl" (9.28 c)

£ ~ y. (9.28 d)

u ~ b. (9.28 e)

9.4 Balance and Conservation Laws for Beams

Beams in bending shall be treated employing the same methodology as bars in


tension-compression and shafts in torsion. We begin again by considering the
derivative of .t(W + V) to obtain

which is clearly identical to the corresponding equation for bars and shafts (9.17).
It is only when substituting the specific expressions for Wand V, equations (9.15)
and (9.16), respectively. that differences begin to appear. Indeed,

Wi = 1 (EI l tfl2 + M tfI/ , (9.30)

VI = - qI W - qw I - m IIjI - m~. (9.31)

Since Q = M I + m from the moment equilibrium condition, Q/ may be written as

(9.32)
204 9 Bars, Shafts and Beams

and

M// = - q - mI. (9.33)

We can write in a fIrst step

fxk (w + V) JI
=x k (1 (ElY Tjl2 +MTjI/ -qw/ -q"w -mTjl-m/t/I)+

+ kx k -l (W + V) . (9.34 a)

Repeated partial integration of the term J! M rf/, together with suitable


substitutions, leads to

k k k
x MTjI/ = (x MTjll + (x QW)/ +x k (qw/ + mTjl)-

- kx k -l (M TjI + Qw /) . (9.34 b)

Further partial integrations and substitutions of the equilibrium equations and the
constitutive law (9.14) result finally, omitting intennediate steps,

fxk B _kx k-l H + ~ k (k-l)x k-2 RJ/ = -x k b _kx k-l h -

- ~ k (k-l )X k-2 r - k (k-l) (k-2 )X k-3EI (ifl + ~ Tjlw). (9.35)

Similarly as in bar theory, the following abbreviations have been introduced for
beams

material force B =W + V - M TjI - Qw I, (9.36 a)

material virial H = - "21 Mrjr - "23 Qw, (9.36 b)

second-order virial R = - 2EI~ - 3EITjlw, (9.36 c)

material loading b = -1 (EllTjl2 + q /w + m /rjr, (9.36 d)

material virial loading h = -5 qw +-


3 m ,I,
(9.36 e)
2 2 '/"
second-order virial loading r = 2 (EIl ifl - 3mw. (9.36 f)

It does not appear possible to generalize the above balance law further, because
the combination rf + 3/2 TjI w cannot be represented as the derivative of an
expression, as may be verified.

It is remarkable that the structure of the general balance equation above for beams
is the same as that for bars and shafts, except that the six abbreviations (9.36) differ
9.5 Energy-Release Rates and Stress-Intensity Factors 205

from (9.19) and the last source tenn in equation (9.35) differs from that of (9.18).

As for bars, we can now, in turn, consider k = 0, k = 1, and k = 2, realizing again


that for k ~ 3 no conservation laws would be derivable.

We obtain

k = 0 (9.37)

k = 1 : (xB -H/ = -xb -h, (9.38)

k = 2 :

(9.39)

The conservation law

B = const. (9.40)

is obtained for constant stiffness El-and constant loading q and m.

The conservation law

(xB -H/ = 0; xB - H = const . (9.41)

is obtained if, in addition to constant stiffness E1, the loading vanishes, q = m = O.

Finally, the conservation law

x 2B - 2xH + R = const. (9.42)

is obtained under the same conditions as the conservation law for k = 1.

The interpretation of the above balance and conservation laws follows the same
lines as for bars and shafts. The somewhat lengthy derivations for E1 "# const, q,
m "# 0 are relegated to Appendix B, where time-dependent defonnations are also
considered.

9.5 Energy-Release Rates and Stress-Intensity Factors

In this Section, we investigate bars, shafts and beams, with discontinuities in their
stiffness. Especially, energy-release rates are calculated due to the translation of
niscontinuous cross-sections. It turns out that remarkably simple formulae evolve to
206 9 Bars, Shafts and Beams

calculate stress-intensity factors for these one-dimensional structural components


with cracks. We start within the context of bar theory and extend the results,
subsequently, to shafts and beams by analogy considerations.

Consider a bar containing a jump in axial stiffness EA at an arbitrarily fixed


position (given by

EA (l) = cons!. for x < (.


E(x)A(x) = (9.43)
l EA (2) = const. for x > (

(see Figure 9.4, state @). The jump in stiffness EA may be due to a jump in E
(e. g., phase transition) or to a geometric change in the cross-sectional area A.

At the transition point x = (, we can distinguish between continuous and


discontinuous variables.

If it is assumed that the axial load n is smooth, it is obvious that n, N, and u are
continuous, while EA and ul are discontinuous. The expression for the material
force B, given by (9.19 a), might be rearranged using (9.2) - (9.5) as
2
_
B - _ 1 N - nu . (9.44)
2EA

...--X
~ ., EA(/)

,
,
EA (2)
S
, ~
I"
))
I-
« .1 p.

®
( EA(/) EA (2)
S
-
:,
'A-

Fig. 9.4: Bar with jump in axial stiffness

Therefore, it follows that the material force B is discontinuous. The jump term lBj
is easily calculated to be

[B] = B' - B - = -1 N2 (() [C] , (9.45)


1
where lCj is the jump in the compliance C
EA
1 1 (9.46)
[C] - - - - - -
EA (2) EA (1)
9.5 Energy-Release Rates and Stress-Intensity Factors 207

To provide a physical interpretation of [BJ, a bar (length f) is considered to be


composed of two sections, EA(l) (length () and EA(2) (length f - (), subjected (for
sake of simplicity) to end forces No. The potential Oi of the internal forces, i. e., the
strain energy of the system, is given by

0' .f =
t

Wdx = -1
2
IN; t

-
EA
dx
'
(9.47)

o 0

and, due to Clapeyron's theorem (e. g., Fung, 1965), the potential of the external
forces follows to be

(9.48)

Thus the complete potential energy 0 is given by

(9.49)

We wish now to calculate the change of energy when the cross-section x = (is
translated by a small amount A (Figure 9.4 state @). Using (9.47) and (9.49) the
result is

A = 0: o @ -- - 21 (N/
+
EA
( N/ (f -
EA (2) (I)
fJ
) ,
(9.50 a)

1 (N; (( + A) N; (f - ( - A))
o (fJ) = - 2 EA (I) + EA (2) '(9.50 b)

Thechange of energy due to this material translation (see Section 3.2) turns out to
be
1
= - A No ICJ = - A IBJ .
2
AD = 0 - 0 (9.51)
(fJ) @ 2
The quantity A [B J may be, therefore, interpreted as the work of the material
(concentrated) force [BJ in the material translation A. The force [BJ acts at the
cross-section x = (and points horizontally in the direction of the stiffer material
(Kienzler, 1993) and not, as mentioned erroneously in Kienzler & Herrmann (1986
a, b), towards the softer material. This force points hence in the direction of larger
E or in the direction of the larger cross-sectional area A, thus implying possible
removal, not acretion, of material in this latter case.

With

;g = _ lim 0 (( + A) - 0 (fJ = dO = + IBJ' (9.52)


~ --+ 0 A dA
[B J may be identified as the energy-release rate ~q (see Section 4.2) due to a
translation of the cross-section at x = (in x-direction. Equations (9.51) and (9.52)
are also valid for arbitrary loading and boundary conditions. Here, No has to be
merely replaced by N( (). For the calculation of [BJ it is necessary to know only
N( () and [ef.
208 9 Bars, Shafts and Beams

Next, we consider a bar (length f) containing at x = (a segment of length 2c with


reduced stiffness (Figure 9.5, state @).

EA (1) = consl. for 0 ~ x < ( - c and ( + c < x ~ e,


E(x)A(x) =
l EA (2) < EA (1) for ( - c < x < (+ c.
(9.53)

@( EA(l) EA(2)

I
EA(1)

~
---J c -l- c j 4 - -
- x
I
I
I
; I
I
.,
,-- I «» -I e
I
I I
®
( EA(l)

:

I EA(2)
I
1
EA(1)

~
- -
I I
x I 1_ _1 I

Ac Ac
Fig. 9.5: Bar with panty reduced stiffness

Again, the bar is subjected to pure tension No = cons!. The change of total energy,
now, due to a self-similar expansion c ~ c (J + -1.) (Figure 9.5, state (13)) is
calculated in the same way as above yielding

I1ll = - k N 2 [C] = 2 Ac [B I . (9.54)

We recall that a physical (concentrated) moment is defined as a limit when the


lever arm a of two equal and opposite forces F approches zero while the product
aF remains finite. In analogy to the physical moment, we introduce the material
(concentrated) moment [HI by

[H] = lim c fB] . (9.55)


c -+ 0

However, c and [BI are both parallel to the x-axis. Thus the quantity [HI is in its
nature a scalar moment or "Fliehmoment" or "virial" as discussed in Section 2.3.

The energy-release rate due to the transformation c ~ c (J + -1.) is


- oll
rJ = - - - = - 2 [BI . (9.56)
o(k)

For shafts in torsion and beams in bending, the jump term [B I is given by
9.5 Energy-Release Rates and Stress-Intensity Factors 209

1 1
{Bj with {C] - - - - - - ,(9.57)
GI (2) GI (1)
p P
1 1
(B] = - i M 2 fC] with fC] = -
EI(2)
- - . (9.58)
EI(l)
2
The energy-release rates due to translation and self-similar expansion can be
calculated directly from (9.52) and (9.56), respectively. In the following, we will
assume that the length c in Figure 9.5 is small in comparison to the length e,
such that the bar, shaft or beam with partly reduced stiffness might be regarded as
a structural element with two symmetric edge cracks.

As mentioned in Section 4.4, energy-release rates play an essential role in fracture


mechanics. The energy-release rate tJ due to crack extension a -? a + !!:.a is
closely related in linear elasticity to stress-intensity factors K, valuable parameters
to assess a structure with cracks (see equation 4.27). Via Rice's I integral, I = II
= Cf} may be interpreted as the crack-driving force, i. e., a material force acting at
the crack tip and pointing in the direction of crack extension.

In Kienzler & Herrmann (1986 b) it was assumed that the energy-release rate W
for crack extension is equal to that for crack widening , i. e., C -? C (1 + A).
Allthough;g corresponds to a virial acting perpendicular to the crack-driving force,
it was postulated that

Cf} = Wid. (9.59)

For dimensional reasons ('fJ is defined in plane elasticity, whereas Wis defined in
a bar or beam theory) a measure of thickness d needs to be introduced.

For rectangular cross-sections (d x h) or piecewise rectangular cross-sections (I,


T-beams) d is the width of the beam. For other cross-sections an effective width
d' was used. Bazant (1990) noted that the hypothesis (9.59) is not exact but
merely a good approximation. He provided the following explanation (see also
Gao & Herrmann, 1992, and Kienzler, 1993).

Consider a crack with initial length a which may subsequently extend by a small
amount !!:.a to length a + !!:.a as shown in Figure (9.6 a) or widen into a fracture
band of width!!:'b ~ d as shown in Figure (9.6 b). The formation of such a crack
in a stressed body may be imagined to completely release the strain energy from
triangular areas 021 and 023, which are limited by "stress diffusion lines" (e. g.,
Knott, 1973). The slope {3 of the triangular areas, which is generally of the order of
unity, is dependent on the geometry and the crack size. If the crack is widened into
a band of width !J"b in the direction of constant stress, the size of the triangular
zones remains the same so that the stress relief zone is changed from area 1231 to
area 45784 (Figure 9.6. b). In comparison, when the crack is extended by !J"a, strain
energy is released from strips 2683 and 2641 (Figure 9.6 a). It is seen that the stress
relief zone 123876541 for crack extension differs from 12387541 for crack
widening only by a triangular area 56725. This triangular area is proportional to !J"a2
210 9 Bars, Shafts and Beams

and may be neglected in comparison with the strips (proportional to Lla), since Lla is
small in comparison to a . Since Llb I 2 = f3 Lla, Bazant (1990) pointed out that
in the limits Lla, Llb ~ 0

(9.60)

a. b.

Llb
•t
~a

I 8
.. ..
I

a I
~I
I
I
I
I
~ '--Lla
I

Fig. 9.6: Energy-release zones at a crack tip for (:1) crack extension and (b) crack widening into a
fracture band.

Thus, the energy-release rate 'fJ, as a work conjugate to the lengthening of the crack
size, is given by
1
'fJ=---
on
d oa
It follows from (9.60) and (9.61) that ;g can be computed from an
lab once f3 is
known. The stress intensity factor K = K J may subsequently be calculated via the
Irwin relation (4.27)
K2
'fJ =- (9.62)
E-
where E - = E for plane stress and E' = EI( J - u 2) for plane strain conditions.

Restricting ourselves to plane stress, the expression for K follows with (9.60) -
(9.62) as

K = 1- ~ E [BI (9.63)
9.6 Examples 211

and, with (9.45), (9.57) and (9.58), the specific expressions are

f3 N A (1)
bars K=N d.A (1) (A (2) - 1) , (9.64)

I
[(1)
shafts K =T f3 T_
__ -p- - 1 (9.65)
d • [(1) ([(2) ) ,
P P

beams K = M (9.66)

For beams and bars, K represents the crack-opening mode I, whereas in torsion, K
is related to the crack-sliding mode Ill.

The correction factor depends on the loading condition, the geometry and the
crack size. Bazant (1990) claimed that f3 can be determined only through optimum
fitting of the exact solutions. Gao & Herrmann (1992) showed, by contrast, that
this correction factor can be obtained through simple matching with standard
asymptotic limiting crack solutions from plane elastostatics.

But, even if one does not adopt these corrections and takes f3 = 1, i. e., one is
concerned solely with elementary beam and bar theory, the approximation is
remarkably accurate. This has been shown by Kienzler & Herrmann (1986 b),
Herrmann & Sosa (1986), Gao & Herrmann (1992) and MUller et al. (1993 a and
b) with several examples.

9.6 Examples

As a first example, consider a barlbeam of rectangular cross-section d x h with


symmetrical edge cracks of length a under pure tensionlbending (Figures 9.7 and
9.8).

The cross-sectional areas are

All) = dh ,

A(2) = d (h - 2a) ,

and the moments of inertia are


212 9 Bars, Shafts and Beams

dh 3
[(l) - -- ,
12
[(2) =d
- (h - 2al·
12

g (a/h) a

3,0
N~f_~-'N
~ a

2,0 /
/'

1,0

a/h
~l ~2 ~3 ~4 ~5

Fig. 9.7: Stress-intensity factor vs. dimensionless crack length for a bar under simple tension with
symmetrical edge cracks (- (eq. 9.67), .... Benthem & Koiter (1973»

a I

6,0

4,0

2,0

~--+---+---+---+--+-- a/h
~l ~2 ~3 ~4 ~5

Fig. 9.8: Stress-intensity factor vs. dimensionless crack length for a beam under pure bending with
symmetrical edge cracks
(- eq. (9.68), .... Benthem & Koiter (1973»

With (9.64) and (9.66) ({3N = 13M = 1), the stress-intensity factors for the bar under
tensile loading are given by
N
K = - g (-)
a (9.67 a)
d/fi. N h
with

l-l-__l_2-~­ - 1 (9.67 b)

h
9.6 Examples 213

and for the beam under pure bending


6M a
K =--gM (-) (9.68 a)
dy'0 h
with

1 -1) . (9.68 b)
- 2 !!-)3
h
The results are compared graphically with those of Benthem & Koiter (1973) in
Figures 9.7 and 9.8. The agreement is quite satisfactory.

As a second example, consider a bar/beam of rectangular cross-section d x h with


a centered crack of length 2a under pure tensionlbending (Figures 9.9 and 9.10).
As in the example above, the cross-sectional areas are

A (1) = dh ,

A (2) = d (h - 2a) .

3,0

2,0

1,0

a/h

Fig. 9.9: Stress-intensity factor vs. dimensionless crack length for a bar under simple tension with
center crack
(- eq. (9.67), .... Benthem & Koiter (1973»

It is, therefore, not possible to distinguish between edge cracks and a center crack
for tension loading. Equation (9.67) is still applicable. The results are compared
graphically with those of Benthem & Koiter (1973) in Figure 9.9..

For bending, the moments of inertia are


3
[(l) dh
- 12 '
214 9 Bars, Shafts and Beams

The stress-intensity factor for the beam under pure bending is with (9.66) given by

(9.69 a)

with

gM = Ij (1 _~ ~J' - 1) . (9.69 b)
n
Again, the results are compared graphically with those of Benthem & Koiter
(1973) (cf. Figure 9.10) and, again, the agreement is satisfactory.

0,8
/,

/,
0,6
;;
;;
0,4 /,
/,
/-
0,2 /-
..-:
...."

°° 0,1 0,2 0,3 0,4 0,5


a/h

Fig. 9.10: Stress intensity factor vs. dimensionless crack length for a beam under pure bending with
center crack
(- eq. (6.69), .... Benlhem & Koiler (1973»

As a third example consider a circular shaft with radius R under pure torsion with
a circumferential edge crack of depth a and a penny-shaped center crack of radius
a (Figures 9.11 and 9.12).

From (9.65) it follows

K = 2T (a ) (9.70 a)
(71' R 5 gT R
with

gT = 1(1 1
-2 .'!-/
- 1 (9.70 b)

for the edge crack and with


9.6 Examples 215

gT " 11 _(~~)' - R
1 (9.70 c)

for the penny-shaped crack. The graphical comparison of (9.70) with the results of
Benthem & Koiter (1973) is depicted in Figures 9.11 and 9.12.

g (aid)
a

200

150
T-.- -lftE-T
a
I
100 I
I
50 /
/

aid
0,1 0,2 0,3 0,4 0,5

Fig. 9.11: Stress-intensity factor Ys. dimensionless crack length for a shaft under pure torsion with
circumferential edge crack
(- eq. (9.70 b), .... Benthem & Koiter (1973»

T-.- -[j ~ ~--T


g (aid)

1,50

1,00
/
/
0,50

o aid
o
Fig. 9.12: Stress-intensity factor Ys. dimensionless crack length for a shaft under pure torsion with
penny-shaped center crack
(- eq. (9.70 c), .... Benthem & Koiter (1973»

It is astonishing, that elementary beam theory may be applied with success even for
circumferentially cracked cylindrical pipes, cf. Gao & Herrmann (1992). As a fourth
example we present here the principal results of that analysis. Consider a pipe
segment depicted in Figure 9.13.

The stress intensity factor K is given by


M -
K = I'TT"R g (9.71 a)
'TT" R2 t
216 9 Bars, Shafts and Beams

g ( o./n) t

4
M(~2R 9 20. 6)M
I
I

3 I
I

o./n
0,1 0,2 0,3 0,4 0,5

Fig. 9.13: Stress-intensity factor vs. nonnalized crack angle a for a circumferentially cracked
cylindrical pipe (radius R, thickness I) under bending (- eq. 9.71, .... Sanders (1982 and
1983»

with the function g (a/TT') defined as

g -TT' (/(1)
- _1) . (9,71 b)
2 /(2)

The bending stiffness of the uncracked pipe is

whereas the bending stiffness of the cracked pipe is


21r - a
E/(1)
E/(2) (cosO + sina /80.
/ TT' - a
a

Figure 9.13 (taken from Gao & Hemnann, 1992) compares the prediction of beam
theory with the result from the complete shell analysis of Sanders (1982 and 1983).
Up to a crack angle of a/TT' = 0.4, corresponding to a nearly half-cracked cylindrical
shell, the results match closely. It is remarkable that the elementary beam theory
can actually reproduce the results generated through a complicated shell analysis.

Likewise it is possible to investigate curved cracked beams in bending and tension


as sketched in Figure 9.14, cf. Muller at aI. (1993 a) and (1993 b).

In closing this Section, the far-reaching and rather unexpected applicability of


theories of strength-of-materials to cracked structural elements of great variety, once
more confirmed the power of these theories which rests mainly on their simplicity
and accuracy as compared to theories of elastic continua.
9.6 Examples 217

F F
.)M

Fig. 9.14: Cracked circular beam subjected to a mount couple M, tensile forces F and three-point
bending P
10

Plates and Shells

10.1 General Remarks

Plate and shell theories attempt to describe the three-dimensional state of


deformation and stress in terms of the defonned and undeformed configurations of
the middle surface. (The middle surface of a plate is plane, that of a shell is
curved). In this sense, plate and shell theories are inherently approximative and their
quality depends on the validity of the a priori assumptions (cf. Bernoulli's normal
hypothesis in beam theory) introduced. The variety of existing plate and shell
theories might be viewed as a measure of their degree of approximation. Since no
"exact" theory is available, various features that characterize a consistent theory
have been introduced to estimate possible errors of the results and to indicate the
range of applicability, e. g., Koiter (1960), Naghdi (1972), Kratzig (1980).

As for bars and beams, conservation laws for plates and shells might be derived in
different ways. Since we have already provided conservation laws in three-
dimensional elasticity (see Chapter 2), the a priori assumptions may be introduced
into those expressions and the integration over the plate or shell thickness can be
performed. However, it is not guaranteed that the invariant integrals in three-
dimensional elasticity are compatible with the plate or shell theory employed. It has
been shown by Sosa et al. (1988) that differences occur, especially when dealing
with the M integral. As pointed out by Li & Shyy (1997), it would be both
technically significant and aesthetically appealing to establish conservation laws
within the original theory, because the partial differential equations which govern
the motion of a particular class of structures may yield their own invariant integrals.
An alternative way would then be, to apply Noether's formalism, the Neutral
Action (NA) method, or any other method mentioned so far, directly to the
governing Lagrangian or to the governing field equations within the proposed
theory. This latter approach will be followed in this Chapter.
220 10 Plates and Shells

10.2 Plate Theories

With reference to an (Xl' X2• x3)-rectangular coordinate system, a plate is considered


as an elastic body bounded by two surfaces x3 == ± h/2 (usually called the faces of
the plate) and by a cylindrical surface perpendicular to the mid-plane x3 == 0 of the
plate. The distance h between the two faces is assumed to be small compared with
the other linear dimensions of the plate and might be, in general, a function of Xl
andx2

The most common theory of plates is referred to as the Poisson-Kirchhoff or


classical theory. Within this theory it is assumed that rotations are small, i. e.,
Ow/Ox]' Ow/Ox2 < 1, where w is the transverse displacement of the mid-surface
(Xl' X 2' 0). It is also assumed that the linear fibers of the plate which are initially
perpendicular to the middle surface remain straight and perpendicular to the
deformed middle surface (sometimes referred to as Kirchhoffs conditions).
Additionally, the shear strains 6/3 and 6 23 are assumed to be negligible (analogously
to the assumptions of the Bernoulli-Euler beam theory in Chapter 9). Finally, it is
assumed that the normal stress a33 is negligibly small in comparison with the
inplane stress components.

Although very useful for many practical problems, this theory leads to some
undesirable consequences such as the occurence of concentrated forces at the
corners of simply-supported plates of polygonal shape and inaccurate results in the
determination of stress concentrations at the edge of holes when the dimensions of
the hole are comparable to the magnitude of the thickness of the plate. These effects
were explained by Kirchhoff (1850), who showed that the assumptions underlying
this theory were responsible for a contraction on the number of boundary conditions
that have to be satisfied at each edge of the plate. For example, along a free edge,
three physical boundary conditions must be satisfied. These are the conditions of
vanishing bending and twisting moments and of vanishing transverse force.
However, in the classical theory the three conditions collapse into two by an
appropriate combination of the shear force with the twisting moment.

Reissner (1944, 1945, 1946) established a theory that accounted for shear
deformations, leading to a higher-order theory in which three boundary conditions
can and must be prescribed along the edge of the plate. As a consequence, due to
the effect of shear deformation, Kirchhoffs conditions do not hold any more.

Mindlin (1951) advanced a dynamic theory, including the effects of transverse shear
and rotatory inertia, and showed that the two-dimensional theory of flexural motions
of isotropic elastic plates could be deduced from the three-dimensional equations of
elasticity.
10.2 Plate Theories 221

Both theories are approximate as compared to the theory of elasticity because of


imposed constraints. Reissner's and Mindlin's theories are distinguished from each
other in that the former takes the effect of linearly weighted averages of the normal
stress distribution in thickness direction into account. The difference between both
theories is insignificant for thin plates, and in the absence of transverse loading they
become identical. Since, in addition, numerical analyses, such as the finite-element
method, usually develop their formulation based on a Mindlin-type plate, the
following considerations rely on the latter theory. We start with the general
elastodynamics of Mindlin plates and introduce later the assumptions of statics and
those of the classical theory.

It is postulated that the displacements have the following form

(10.1 a)

(10.1 b)

(10.1 c)

where uI • u2 and u3 are, respectively, the displacement in the XI' X 2 and x3


directions; t is the time. Here, w is the linearly weighted average (taken over the
plate thickness) of the transverse displacement; 1ftI and 1ft2 are, respectively, the
linearly weighted averages of the rotation about the x2 and XI axes of fibers
normal to the mid-plane before deformation. The variables 1ftI' 1ft2 and w are called
plate displacements with sign conventions depicted in Figure 10.1.

_#:========A- ~X, .~ \112

I~,
Fig. 10.1: Sign conventions of plate stress and displacement components
222 10 Plates and Shells

The bending and twisting moment resultants Ml/' M22 , Ml2' M2/ and the transverse
shear force resultants Q/ and Q2 are related to the plate displacements by

Ml/ = D (1ftl,l + V1ft2.2)' (10.2 a)

M 22 = D (1ft2.2 + v1ftl,l)' (10.2 b)


1 - u
M 12 = M 21 = -2- D (1ftu + 1ft2.1)' (10.2 c)

Q/ = Cs (w,1 + 1ft/), (10.2 d)

Q2 =Cs (w,2 + 1ft2), (10.2 e)

where the flexural stiffness D and the shear stiffness Cs of the plate are given by,
respectivly,
3
D = Eh (10.3)
12 (l - u 2) ,

Cs = k Gh

with E, G, u being Young's modulus, shear modulus and Poisson's ratio,


respectively, while the shear correction factor k takes the value 5/6 (based on energy
considerations in Reissner's theory) or ",2/12 (based on dynamic considerations in
Mindlin's theory). The quantities Ml/' M n , M/2I Q/ and Q2 are functions of (x" x 2' t)
only and are referred to as plate stresses with sign conventions depicted also in
Figure 10.1.

The equations of motion expressed in tenns of plate quantities are obtained by


integrating the corresponding 3-D elasticity equations through the thickness to give

(10.4 a)

o, (10.4 b)

(10.4 c)

with p = p (x/' x2) being the transverse pressure acting in the x3 - direction, p
being the density per unit of volume and I being the moment of inertia per unit
width of the plate. For a inhomogeneous plate, D, Cs, u, p, h and I might be
functions of x/ and Xl'

Finally, the strain energy density W, the kinetic energy T and the potential of
external forces V per unit of area are given by, respectively,

W = 1/ M n 1ft],] + M 22 1ft2,2 + M 12 (1ft2,] + 1ft],2) + (10.5 a)

+ Q] (W,] + 1ft]) + Q2 (W,2 + 1ft2) J.


10.2 Plate Theories 223

T ="21 ! ph W,t2 +
221
+ pI 1fi2,tJ'
pI 1fil,t (10.5 b)

V = - wp. (10.5 c)

Thus, the Lagrangian density L per unit of area of the plate

L =T-(W+V)

is of fIrst order, i. e., it does not depend on higher than fIrst derivatives of the
three dependent variables w, 1ft! and 1ft2'

In order to find solutions for a particular plate problem, the governing fIeld
equations have to be supplemented by a suitable set of initial and boundary
conditions, that are not required in the following considerations. The reader is,
therefore referred to the literature, e. g., Timoshenko & Woinowsky-Krieger (1959)
and Eschenauer et al. (1997).

The classical theory is recovered, if the Kirchhoff condition (normal hypothesis) is


introduced. With

(10.7)

shear strains are neglected. At the same time, an infInite shear stiffness Cs and a
vanishing moment of inertia I are assumed. Thus, the shear forces remain finite
but constitutive relations for the shear forces do not exist. These are obtainable
from the equations of motion.

The equations of the classical theory reduce to

- kinematics

- z w.J, (10.8 a)

- Z w.2 ' (10.8 b)

(10.8 c)

- constitutive relations

(10.9 a)

- D (W,22 + U w,u)' (10.9 b)

- (1 - u) D W.J2 (10.9 c)
224 10 Plates and Shells

- strain energy density

W = -"21 (MlJ w,n + M22 W,22 + 2M12 w,12) (10.10 a)

- kinetic energy

T= i2 phw 2
J
(10.10 b)

- potential of external forces

v = - wp. (10.10 c)

Thus, the Lagrangian density L (per unit of area) of the classical theory

(10.11)

is of second order and depends on tile single dependent variable w. In order to


obtain the associated Euler-Lagrange equation we eliminate the shear forces in
equations of motion with the result

or, in term of the deflection w, by introducing the constitutive relations

(10.13)

Here, the Laplace operator d ( ) = ( }.11 + ( },22 is used. Initial and boundary
conditions have to be adjusted analogously (cf., e. g., Timoshenko &
Woinoswsky-Krieger, 1970).

10.3 Conservation Laws for Elastostatics of Mindlin Plates

As in beam theory, we relegate the lengthy derivation of conservation laws for


inhomogeneous plates, including elastodynamics, to an appendix (Appendix C) and
restrict ourselves to balance laws for statically loaded inhomogeneous plates which
will change to conservation laws in the absence of inhomogeneities, i. e.,
dependence of the Lagrangian explicitely on Xj'

In this Section, we are interested mainly in conservation laws which lead in turn to
path-independent integrals applicable in fracture mechanics. It is not our intention
to strive for completeness. The most efficient way to derive these conservation laws,
then, is to apply the vector operations gradient, curl and divergence to the
Lagrangian L and its moment xL, as has been shown in Sections 2.4 and 6.4. In
10.3 Conservation Laws for Elastostatics of Mindlin Plates 225

static cases, the Lagrangian is just equal to the negative of the sum of the strain
energy density W and the potential of external forces V

L = - (W + V) = L (xk' t/t..,I t/t...,


I,)
W, W')
,I
(10.14)

with

W 1 Q.I (t/t.I
= -
2 {M..IJ ,Ir.. +
'f'j,I
+ W ·)1
,I
(i, j, k = 1, 2) (10.15)

and V given by (10.5 c). For convenience, the equilibrium equations (10.4) are
recalled. Neglecting all time-dependent terms in (10.4) we have

M...
)I,J
-Q.I =0, (10.16 a)

Q.. +p = 0, (10.16 b)
'.'
where the summation convention applies.

In the sequel, the three mathematical operations are applied sequentially.

Gradient

Considering the gradient of the Lagrangian, we find with (10.14), (10.15) and
(10.2)

grad (W + V) = (W + V),k
deW + V)
=M..JI t/t.·k
I,J
+Q.I (t/t.·k
I.
+w·,I k) -p w, k + dX .
k
The explicit derivative of W + V with respect to the independent variable
xk (k = 1, 2) accounts for material inhomogeneities produced, for example, by
stiffness distributions D(xJ and Csfxk ) smoothly or discontinuously variing with X k
(defects), and for physical inhomogeneities, produced by p(xk).

Integration by parts yields

(W + V)k, = (M..JI t/t.·k)·


I..J
- (M...
J',}
- Q.)
I
t/t.,·,k +

(10.18)

Due to (10.16) and (W + V),k = (W + V)J ~jk it is


_d(W+V)
f (W + V) ~·k
J
- M.JI t/t..I, k - Q.}Wk.
. ,J
- j dX
.
k
(10.19)

The left hand side of (10.19) is equal to the Eshelby tensor of plate theory,
226 10 Plates and Shells

B'Jk = (W + V) 8'}k - M l'.. ~'k - Q.J wk' (10.20)


~ "

It corresponds to biJ:' equation (2.53), in three-dimensional elasticity and is the


two-dimensional analogue to the material force B (9.36 a) in the one-dimensional
beam theory.

In the absence of inhomogeneities, the Eshelby tensor (10.20) is divergence-free

Bjk,j = O. (10.21)

H the functions involved are "smooth", integration of (10.19) with respect to the
middle-plane surface S (elements of area dA) and application of the divergence
theorem (1.8) yields with (10.20) the J integral of plate theory

Jk -f B
-
jk
nj ds -_jO(W + V) dA • (10.22)
r s oXk
where r is the boundary enclosing S, s is the arc length along rand n is the unit
outward normal vector.

It can be observed that when the pressure load p is constant and the material is
homogeneous within S, equation (10.22) can be written as a single line integral on
r, and the balance law (10.22) becomes a conservation law.

Curl

This operation is performed on the vector quantity (W + V) x. With the


permutation tensor EAij defined in (2.5) it follows

curl j(W + V) x J= kij j(W


E + V) xji

= Ekij j(W + V),i Xj + (W + V) flji J


= E
ki1. X.J (W + V) ,I.. (10.23)

Of particular interest from a physical point of view regarding plate problems is


the case k = 3. Use of (10.18) and integration by parts yields

- Q.} w.).
,I

We add and substract the term M nj t/ti,1l + Qj t/ti on the right-hand side and fmd

Er IJ ff (W
I
+ V) x.j = Er Ifix' (M
) ,i IJ j J tun
~.
m,l
+ Qnw.) I -
, I ~n
10.3 Conservation Laws for Elastostatics of Mindlin Plates 227

-!M.. )n
Ir. +
'f' n,1
M.. Q. w.)J+
n)
Ir.
'f'"n
+
)
(,'r. +
'f', ,I

+
fM. n)
1ft..I,n + ) J
Q. 1ft.. x· o(W+V)j
1ex. .+
)
I
(10.24)

The tenn in the second square bracket vanishes on substitution of (10.2). It


corresponds to the isotropy condition (2.25) of the three-dimensional theory of
elasticity. If the material behavior is not isotropic, i. e., the constitutive equations
(10.2) have a different form, this tenn has to be treated as an extra source tenn,
cf. (6.69).

Focusing attention on the term in the third square bracket we can write

£3"
I)
(Mn). 1ft..I , n
+ Q.
)
1ft.) = £j"I) (- MIV
. )1ft..
,II
- Q.I )
1ft.)

= £j;j I(-M,,; ~)," + (Mni,,, - Q) ~J

= £1"I) 1- M. 1."1 'U)


1ft.. (10.25)

because of (10.16 a). Upon introduction of (10.25) into (10.24), a rearrangement


leads finally with (10.20) to the desired balance law

(B I V
..x.) M m.•'f'j/,n
_ a (W + V) x· .
£j"
I)
+ Ir.) - £j"
~I ex. ) I
(10.26)

If the material is isotropic and homogeneous, the source term vanishes giving rise
to a divergence-free expression analogously to (2.60 d)

£3"
I)
. x.)
(B I U + M.'U .Ir.)
'f') ,II
= O. (10.27)

Integration of (10.26) and application of the divergence theorem leads


to the L 3 integral of plate theory

Lj -f£3" (B . x
-
I) "')
+ M. 1ft..) n ds -
Ill)"
X. dA. (10.28) -f£j" O(W+V)
I)..J.-)
r s VA.;

Again, if the integrand of the surface integral vanishes, the balance law (10.28)
changes to a conservation law in integral form.

Divergence

This operation is applied to the moment of the sum of strain energy and potential
of external forces per unit of mid-plane surface as

div I(w + V) xJ = I(W + V) xi


= 2 (W + V) + x.1 (W + V).,I . (10.29)
228 10 Plates and Shells

The factor 2 arises from Xu = 8 jj, which is equal to two for plane problems. With
equation (10.15) and (10.18) it is

!f(W
I
+ V) x. 1
, 1,i = M 'J.. tJr...
J,'
+ Q.J (rjI,.J + w.) - 2pw +
,J

+ Xi
f(MJK rjl,k,.)'
I ,J J , ' ,J
~W+V)J.
+ (Q. w.) . + ---'--------"--
ilx
i
Integration by parts leads to
o(W + V)
!f(W + V) x., 1,i~ = (x.' M'
Jk
1/tk,.)'
1 ,J
Q. w.).
+ (x.' J ,I 'J
+x.1
I
ilxi
- 2pw + Qj ~ •

The last term may be transformed with (10.2 d, e) and, again, integration by parts
leads to

= Qj Qj - Q. w.
C J ,J
S

= Qj Qj - (Q. w). - pw. (10.30)


C J ,J
s
Introduction of the last equation and rearrangement yields with (10.20), finally
o(W + V) Q. Q.
(B .. x. + Q. w). = x. - 3pw + _J_J . (10.31)
J" J ,J I ~j Cs
The first term on the right-hand side of (10.31) has the same structure as previous
balance laws (compare, e. g., with (2.67 f) in combination with (2.50), with (9.22)
in combination with (9.19 d) or with (9.38) in combination with (9.36». The
second and third terms are completely different in nature and do not vanish even
for a homogeneous material and constant pressure p. It is also interesting to note
that the last term is dependent on the plate theory employed, indicated by the
presence of the effective shear stiffness Cs'

Integration of (10.31) and application of the divergence theorem leads to the M


integral of plate theory

M = / (Bji Xi + Qj w) nj ds = .f(o(W + V> Xi - 3pw + Qcj Qj ) dA.


r s
ilx i S
(10.32)

The surface integral vanishes only under the quite restrictive conditions that the
material is homogeneous and the plate is subjected to pure bending. As will be
shown below, the third term vanishes in the classical Poisson-Kirchhoff plate
theory.

The integral expressions (10.22), (10.28) and (10.32) have been discussed in terms
of path- and path-domain independent integrals in Sosa (1986), Sosa et al. (1988)
and Sosa & Herrmann (1989). In addition, the physical interpretation of the
integrals as energy-release rates due to cavity translation, rotation and self-similar
lOA Reduction to the Classical Theory 229

expansion were given. Finally, the relationships of Jl , L 3 and M to fracture


parameters, e. g., stress-intensity factors, were established, that make the integrals
valuable tools, from a practical point of view, in fracture mechanics.

10.4 Reduction to the Classical Theory

In order to obtain conservation laws, the same mathematical operations introduced


in Section 10.3 could be performed within the classical Poisson-Kirchhoff theory
governed by equations (10.7) - (10.13). This Section is devoted to show that in the
classical plate theory, invariant integrals can be obtained as a particular case of
equations (10.22), (10.28) and (10.32).

It should be noted, however, that in general such a reduction does not apply in the
presence of a flaw. As it was shown in Sosa (1986), the asymptotic expressions of
the elastic fields in the neighbourhood of a crack for the classical plate theory do
not follow from the fields corresponding to the transverse shear theory. As a
consequence, crack problems need to be solved individually within each plate
theory.

For the following, it is useful to refer the plate field quantities in the integrand of
the contour integral to components nand s nonnal and tangential, respectively, to
the boundary. The relations between the plate stresses in both coordinate systems
then become

(10.32 a)

(10.32 b)

(10.32 c)

where

(10.33 a)

n2 '" cos (n, x) '" sin a (10.33 b)

and a is the angle between the coordinate systems (Xl' X2) and (n, s) as depicted in
Figure 10.2.

Likewise, the following relations are useful

(10.34 a)
230 10 Plates and Shells

t/ls = t/l2 n1 - t/l1 n2 ' (10.34 b)


mv (10.35 a)
an
mv (10.35 b)
~

/-t--.......c.---'-------t----_ Xl

Fig. 10.2: Coordinates along the contour r of integration

Under the condition of neglecting the transverse shear strain components, the three
plate balance laws are considered in sequence.

Jk integral

The analogue of the J integral of the three-dimensional theory of elasticity follows


from (10.22) and is rewritten as

Jk = fB jk nj ds
r

= fl(W
r
-I- V) n k - M nn rft'~k - M ns rfts•k - Q n w,:< Jds
= jO(W -I- V) dA (10.36)
s Oxk
(no summation over nand s).

The strain-energy density W depends merely on W,ik' as indicated in equation


(10.10 a). With the normal hypothesis (10.7) and with (10.35) one can write

.
rftsk-QnWk
,

mv
-.2 +MflS
=MfIn..:J...
mv k
;;l,. - Q" w,k
lOA Reduction to the Classical Theory 231

= M
ow k
- ' - (Q
iJM,lS
+ --) W k - -
iJ
(M W k).
lUI on " d.s' ' d.s' "',
The last tenn vanishes upon integration along a closed curve, since it is a total
differential. The second term is Kirchhoffs so-called ersatz-shear force
_ iJM
Q =Q +~. (10.37)
" " d.s'
In the (XI' xz) coordinate system it is commonly defined as

Q) = Q) + M 12,z • (10.38 a)

Qz = Qz + M Zl,l . (10.38 b)

Equation (10.36) can then be written as

Jk = I'!f(W I + V) n k -
,I, - -Q"
M "" 'I',~k W,k} cis = jiJ(W~+ V) ciA .
r s VA k (10.39)

It may be noted, that (M and QJ and (lfrn Mon and w) are compatible with owl
classical plate theory, which considers only two components of force and
displacement at any boundary of the plate.

L3 integral

The analogue of the L] integral of the three-dimensional theory of elasticity follows


from (10.28) and is rewritten as

L 3 = j£3ij (Bid xj + Mki lfr) n k cis


r

= j£3ij fj j(w + V) n i - M w • lfrll,i - M", lfrs,i - Q ll w,;J+


r
+ (M"" lfrs - M", lfr,) ) cis

iJ(W + V) ciA.
= f£3" X. (10.40)
IJ J &.
s 1

Manipulations along similar lines as above lead to

+ -.!...
d.s'
(M w.)J.
"',1

Focussing attention on the last tenn in the above expression we can write
232 10 Plates and Shells

E31;; XJ'
,
o(M1l'I wJ
d' ' = E3 .. -
IJ
°
d' (x.J M 1l'I w.)
,I
- E31;; _J
'd'
de.
M 1l'I W.
,I

= E ..
3lJ
~
d'
(x. M
J 1l'I
w.) - M
,I 1l'I
~
On
(10.41)

The fIrst tenn will vanish when integrated along a closed path. The last two tenns
in the line integral of (10.40) become

M
nn 'f's
", - M
1l'I
,1,
'f'n
= - M
nn
~
d'
+M
1l'I
~
On
. (10.42)

Therefore, upon introduction of equations (10.41) and (10.42), the tenns M... w,ll
cancel and (10.40) is written as

T
~
= f1e .. X.
/" 3IJ J I
f(W + V) n. - M
I I nn
0l/f; -
On
Qn w.j - M nn
,I
~)
d'
cis
r

(10.43)

M integral

Since in this case the shear stiffness Cs --7 00 the last tenn in (10.32) vanishes and
the analogue to the M integral of the three-dimensional theory of elasticity is
rewritten as

M =
j f(B
I
.. x. +
JI I
Q.J w) n.J ds
r
= fj(W + V) Xi ni - (Mnn 1/tn,i + Mns 1/ts,i + Qn w) Xi + Qn W j ds
r
= f /o(W
+ V) Xi - 3pw dA. j
(10.44)
Sdei
Transfonnations as indicated above and integration by parts where appropriate

:i -
lead to

M = f fi I(w + V) n i -M nn Qn W,;}+ Qn w) ds
r
= flXi o(W + V) - 3pw dA . j (10.45)
s de 1
When the applied pressure p is constant, the surface integrals for Jk and L 3 vanish,
which is not the case for M.

In concluding this Section, it might be mentioned that the conservation laws


presented are useful in the numerical analysis of cracked plates, cf., e. g., Sosa &
Eischen (1986).
10.5 Conservation Laws for Shells 233

10.5 Conservation Laws for Shells

It would be beyond the scope of this text, if a detailed derivation of shell theory,
and the underlying theory of surfaces were to be presented here. Thus the reader is
expected to be familiar with those topics, the aim being a concise discussion of
conservation laws.

The internal geometry of the curved middle surface of a shell imposes the necessity
of using curvlinear coordinates and related to them, the special features of
Riemannian geometry. Covariant and contravariant coordinates have to be
introduced and partial derivatives are to be replaced by covariant derivatives with
the help of Christoffel symbols, cf., e. g., Fliigge (1972, 1973), Niordson (1985),
Eschenauer et a1. (1997). Shell theories of various degrees of approximation have
been derived and are used in practice. Independent of a specific shell theory under
consideration, however, a severe restriction to establish conservation laws in shell
theory becomes apparent if we think in terms of material transformations as
expounded in Chapter 2.

Material translation and rotation were introduced by the comparison of two


neighbouring volume elements which differ from each other by a rigid body
translation (or rotation) in the undeformed state. Because of the curvature of the
shell middle surface, such two elements will not be congruent with each other in
general, giving rise to a distortion, except for shells which enjoy a high degree of
symmetry, e. g., cylindrical shells in axial direction, shells of revolution in
circumferential direction and spherical shells as indicated by Lo (1980) and Kienzler
& Golebiewska-Herrmann (1985). In theories of shallow shells the distortion is an
effect of second order. Thus, for shallow shells, path-independent integrals might be
established for a wider class of mid-surface geometries, cf. Nicholson & Simmonds
(1980) and Li & Shyy (1997).

The aim of the present Section is to give a short account of a shell theory and the
associate Eshelby tensor. We will not give a detailed discussion but wish merely to
acquaint the reader with special features of problems one faces in shell theory.

In the following, it is assumed that the range of greek indices is (1, 2) whereas the
range of latin indices is (1, 2, 3). The displacements of a shell are described
commonly by five independent components ua ' Ifta , w as

(10.46)

The state of deformation consists, therefore, in a combination of the


displacements of a sheet (plane stress) and a plate. The vectors a a and a 3 are
contravariant base vectors of the shell middle surface and the unit vector
perpendicular to the surface, respectively.
234 10 Plates and Shells

The elongation and shear strain of the middle surface are denoted as a..p , the change
of curvature and the twist by f3 aIJ and transverse shear strain by Y3a' The change of
thickness and warping of the cross-sections are neglected in (10.46) but may be
considered in higher-order shell theories (Kienzler, 1982 and Kienzler &
Golebiewska-Herrmann, 1985).

The kinematic relations are given by

aa{3 = "21 (ualf3 + uf3la - 2baf3 w), (10.47 a)

f3a{3 = 1 (1/talf3 + 1/tf3la - b: uplf3 - bI upla + 2 b: bpf3 w), (10.47 b)

Y3a = "21 (1/ta + w,a + b: up)' (10.47 c)

Vertical bars denote covariant differentiation and b'rf3 (or b;) is the curvature
tensor of the middle surface, (not to be confused with the Eshelby tensor bij of
previous chapters, while in this chapter the Eshelby tensor is disignated by BaIJ)'
The tensor of resultant membrane forces, the tensor of moments and the tensor of
resultant shear forces are denoted by Nrf3, M'rf3 = Mf3a and ~ = Q3a = (l',
respectively. In contradistinction to the tensor of moments, the tensor of
membrane forces is not symmetric. In view of (10.47 a), i. e., symmetrical
measures of strain, it is convenient to introduce the symmetrical "pseudo" tensor
of resultant membrane forces (cf. Eschenauer et aI., 1997) as

Na{3 = N af3 + b: Map. (10.48)

With the strain energy density W per unit of area of the shell middle surface, the
constitutive equations are given by

Membrane forces Naf3 = oW, (10.49 a)


o af3a
afJ oW (10.49 b)
M oments M .- - 0 f3 '
f3a

Transverse shear forces Qa = OW. (10.49 c)


oY3a
The external forces pi follow from the potential V as
. OV
(10.50)
P'=-ou.'
I

and the Lagrangian for statics of shells is given as usual by

(10.51)

The equations of equilibrium follow either from the Euler-Lagrange equations


Ek (L) = 0, from considerations of an infinitesimal shell element or from suitable
integrations of the three-dimensional equations with respect to the shell thickness.
10.5 Conservation Laws for Shells 235

They are given by

(10.52 a)

(10.52 b)

= o. (10.52 c)

If the shell middle surface is plane, i. e., bafJ = b; = 0, the above equations of
shell theory uncouple into those of sheet theory (akin to plane elasticity) and plate
theory.

All methods discussed in Chapters I and 2 to establish conservation laws are


obviously applicable also in shell theory. We choose the vector operation gradient
applied to the Lagrangian as the most convenient approach. Thus

grad (W + V) = (W + V),(3

o (W + V) (0 a '?L U I + 0a'?L w )+
o a'?L 0 Uyla ylafJ 0W ,(3

+ 0(W + V) (0 {3 '?L if! /, + 0


{3 '?L U I + 0(3 '?L W a)+
o{3VjL 0 if!yla ya(3 oUyla yla(3 ow 'I'

+ ('0 (W
o Y
+ V) + 0(W + V) )(0 Y3jL
0 r. 0 ",
rft /,
a (3
+ 0 Y3jL W
0w ,a(3
+ 00 Y3jL U /,)+
II a (3
3jL JL3 Y'a ,a a

+ 0' (W + V) U //3 + 0' (W + V) W./3 + 0' (W + V) (10.53)


OU a a ow OX/3

Use of (10.47) - (10.49) leads to

(w + VI
1,(3 = (NaY + b Py Map) U yla(3
I - (NaY + b py Map) b ya W
,(3 +

_ pall /, _p 3 W + 0 (W + V) .
a (3 .(3 OX(3

Addition and subtraction of the terms Nay uylf3a and May if!Y/f3a' integration by
parts and rearrangement yields
236 10 Plates and Shells

+ (NaY U y /(3 )/a + (MaY rjly/f3 )/a + (Qa W,f3 )/a -

- (Na Y/ a - Qa b: + p Y) U y /(3 -

- (Ma Y/a - Q Y) rjI y/(3 +

+ O'(W + V)
O'x f3

Finally, with the equations of equilibrium (10.52), we obtain

/
f(W
I
+ V) fJa -Nay U / -May
f3 y/(3
rjI y/(3/ - Qa W
.(3
J/
/a

= Nay (u / - U / ) + May (rjI / _ tiT / ) + 0' (W + V)


y/a{3 y/{3a y/a(3 'I' y/{3a 0' x(3 •

(10.54)

It shoud be recalled that covariant differentiation on a surface is, in general, not


commutative; rather Gauss' relationship should be used, e. g.,

(10.55)

where

(10.56)

is the Riemann-Christoffel curvature tensor (Naghdi, 1972).

The left-hand side of (10.54) is the Eshelby tensor of shell theory

(l0.57)

Thus, (10.54) is rewritten as

Ba / =R P (Nay U +May rjI) + O'(W + V). (10.58)


(3 fa ya{3 P P 0' x f3

To find the condition for conservation laws, i. e., B; /a = 0, we recall that the
Lagrangian of a shell made of an isotropic material has in general the following
form

All parameters up to the semicolon can contribute to the last term in (10.58). The
condition to be satisfied in order to ensure O'(W + V)/t3x f3 = 0 are
10.5 Conservation Laws for Shells 237

E, U, h = cons!. , (10.59 a)

pj/f3=O, (10.59 b)

b:/f3=O. (10.59 c)

This last condition is satisfied only for shells of constant curvature. In addition,
R p raf3 has to vanish, as seen from (10.58). In shell theory, this condition is
satisfied only for developable surfaces. Thus, conservation laws

(10.60)

will exist for cylindrical shells in the direction of the generators and for shells of
revolution in the circumferential direction, provided the loading pi is constant in
the directions indicated and, of course, the constraint equations (10.59 a) are also
satisfied.

For spherical shells, under certain restrictions, further conservation laws can be
constructed (cf. Lo, 1980 and Kienzler & Golebiewska-Herrmann, 1985).
Appendix A

Conservation Laws for Inhomogeneous Bars under Arbitrary


Axial Loading

The derivation of balance laws for inhomogeneous bars under arbitrary axial
loading has been presented in Chapter 9. The Neutral Action method provides a
means to establish conservation laws in this instance.

To start with, the governing equations of elementary bar theory are recalled from
Chapter 9

force equilibrium N I
= - n, (A.I)

kinematics U
l
= E, (A.2)

Hooke's law N =EAe. (A.3)

After eliminating the axial force N and the axial strain e from equations (A. 1 - A.3)
the governing differential equation becomes

4 = (EA u Y + n = O. (A.4)

As outlined in Chapters 1 and 2, this differential equation has to be multiplied by


the characteristic f It is sufficient to require the characteristic f to be a function of
x, U, and u I only

f = f (x, u, U I) , (A.5)

since higher-order derivatives, say Ull, could be eliminated by (AA). From (1.110)
the characteristic f has to be detennined from

E (f4) = (0 (f4) )11 _(0 (f4) )11 0(f4) +


OU II
au l
au
240 Appendix A

(A.6)

Because of (A.S), equation (A.6) is satisfied only, if the coefficients of the second
(and higher) order derivatives of u vanish independently. It turns out that this is
possible only iff is linear in u and u t, thus

(A.7)

This equation is the same as (1.112) for the homogeneous case. Once f is
determined completely, the conserved current P is given by

P =fEA u l -FEA u + Ifn dx (A.8)

with

pi = fA. = O. (A.9)

At first glance, pi = 0 seems not to be a conservation law because P involves - via


(A.8) - an integral term. But by partial integration and due to (A.6)
if I EA = C = const.) it follows

jfndx fdx =f -C j~ fdxdx (A. 10)

and the integral can always be evaluated explicitly, once the stiffness EA and the
axial load n are given.

H the homogeneous differential equation (n = 0 in (A.4» is considered, the


integral term is not present. This situation of a inhomogeneous bar ((EAy :f:. 0)
will be treated first.

Inhomogeneous bar without distributed loading (n = 0)

Inserting (A.7) into (A.6) and repeated use of the homogeneous (n = 0) equation
(A.4)
U II = _ (EAyt u I
EA (A.ll)
leads, after some manipulations to

(A.12)
Conservation Laws for Inhomogeneous Bars under Arbitrary Axial Loading 241

The tenns in square brackets are neither functions of u nor of u I and, therefore,
must be set equal to zero. The functions t
,f 2 and / 3 follow from integration as

/2 = I~ dx +CO' (Al3 b)

/3 ~ I~ dx 'C" (Al3 c)

Co + Cs constant.

Given an explicit stiffness distribution EA = E(x)A(x), the integrals can readily be


evaluated. In this process, further integration constants are not necessary, for they
are already included in Co' Cj and C4 •

The corresponding conserved current P follows from (A.8) as

P = EA (+ 4C 3 I-L f/t ) dx - 2C2 It - 2 Cj / B +

+ (- 4C3 It J + 2C2
3
H +2_ R+
EA
C

(A. 14)

The constant Co drops out. The quantities B, H, R, N and V have already been
introduced in Chapters 1 and 9 as

material force B = -l EAu /2 (A.I5 a)


2 '
material virial H = -l2 EAuu I
'
(A.I5 b)

second-order material virial R = -l2 EAu 2 ' (AI5 c)

physical force N = EAu I, (A.I5 d)

physical virial V = EAu. (A.I5 e)

The conservation law pi = 0 is valid without any constraint on the stiffness


distribution. As soon as EA = E(x) A(x) is defined, the integrals in (A.l4) can be
242 Appendix A

evaluated and the conservation law appears in explicit form with arbitrarily
specified EA. The special case EA = EA o = cons!. leads to the five independent
conservation laws

Cj "# (): B/ = 0, (A.16 a)

C2 "# 0: (xB -Hy = 0, (A16 b)

C3 "# 0: (x 2B - 2xH + Ry = 0, (A.16 c)

C4 "# 0 : N/ = 0, (A16 d)

Cj "# 0 : (Nx - Vy =0 (A.16 e)

presented already in Chapters 1 and 9.

In the same way, equation (A.14) can be discussed for arbitrary stiffness
distributions EA = E(x)A(x). The most important conservation law, i. e., the
conservation of material force (related to the constant C J) is given by

(E(x)A(x) By = 0 (AI?)

independently of the actual form of E(x)A(x). The conservation law related to the
constant C4 is, independent of the form of E(x)A(x), given as

(AI8)

To evaluate the conservation laws related to the constants C2 , C3 and C5 the


functional form of EA = E(x)A(x) must be given. For example, take

E(x)A(x) = bx + c. (A.19)

Equation (A.14) yields

P = 2EA (~: 2
In (EA) - ~2 In (EA) ~ CJ J
B +

+ 2(- 2 C3 In (EA)
b
+ C2 H
J
/+ 2 ~
EA
R +

+ (~5 In (EA) + C 4JN - ~ R (A.20)

and pi = 0 leads to five independent conservation laws

(EA By = 0, (A.21 a)
Conservation Laws for Inhomogeneous Bars under Arbitrary Axial Loading 243

fe: In (EA) B - H )/ = 0, (A.21 b)

t-EA In 2 (EA) B - -2 In (EA) H


b2 b
+ -
EA
1 R )/ = 0, (A.21 c)

N/ = 0, (A.21 d)
1 1 )/
( b In (EA) N - EA V = O. (A.21 e)

The flrst three laws express the zeroth, first and second order balance of material
force for a bar with the linear stiffness distribuion (A.19), whereas the remainder
are the zeroth and flrst order balance of physical force.

Bars with distributed loading (n :F- 0)

The derivation proceeds in complete analogy to the preceding Subsection. Instead


of (A.ll), the inhomogeneous differential equation (AA) has to be used to
eliminate higher-order derivatives of u. It turns out that the funtions fJ(x) and f
2(X) of (A.7) are the same as in (A.13) butf 3(x) is changed to

(A. 22)

The conserved current for the bar under distributed longitudinal load n(x) turns
out to be

P - 2 [I B + 2EA ( ~)/ H + 2[2/ R +

+ [3 N - [3/ V + / 3 n dx (A.23)

with B, H, R, N und V are defined as before (cf. A.15).

The integral term in (A.23) contains the constants CJ through Cs, the stiffness EA
and the load n. Once the stiffness and the load are specified, the integral can be
evaluated. Therefore P / = 0 is indeed a true conservation law, consisting in fact
of flve independent conserved currents, namely three material and two physical
ones.

As a simple example, formulae for the case n = no = const. and EA = EA o


const. are presented here. The five conservation laws are given as
244 Appendix A

B/ = 0, (A.24 a)

rfsx -H - !... 2
4 EA o
(no x 3 + 3 Nx 2 - 6 Vx J/I) / = 0, (A.24 b)

rfsx
2 - 2Hx + R - !... 2
8 EA o
(no x 4 + 4 Nx 3 -12 Vx 2 )J/
I)
= 0, (A.24 c)

IN + no x J/ = 0, (A.24 d)
Appendix B

B.t Elastodynamics of Inhomogeneous Bernoulli-Euler Beams

The Neutral Action (NA) method is applied to establish conservation laws for
nonuniform beams under arbitrary loading. The governing equations concerning
elastostatics of elementary beam theory, as presented in Chapter 9, have to be
extended to elastodynamcs (see Chapter 6). The time t enters as second independent
variable and in the Lagrangian, the kinetic energy T

T = -
1 m- w '2
2
has to be considered. Here, m
(x) is the inertia term, pA, i. e., mass per unit length
of the beam. Rotatory inertia terms are neglected. The Lagrangian density for
dynamics of inhomogeneous Bernoulli-Euler beam theory is given by

L = l m(x) w -2 - l E(x) l(x) W 1/2 + q(x) w - m(x) WI


2 2
=
I ,W
L (.x , W , W II , W"). (B.1)

and the associated Euler-Lagrange equation is

a = E(L) = mw·· + EI w IV + 2(EIY WIll + (EIYI WII - q - m I.

(B.2)

In the remainder of this Appendix we will not consider moment loading, i. e., we
set m = O.

Given the governing differential equation, one needs to consider

(B.3)

We will assume the characteristic f to be of the fonn

f = f (x, w, w: w·). (B.4)


246 Appendix B

The condition for the existence of conservation laws by the NA method, equation
(B.3), requires

d(fli.) _ (d (f1i.»)1 + (d (fli.) )11 _ (d (fli.) )111


OW eW I eW II eW III

+ (d (fli.»)W _ (d (fli.»)' (d (fli.»)·· + = o. (B.5)


OW W ow' ow"
The characteristic f has to be calculated from (B .5).

Dynamics of beams without transverse loading (q = 0)


Considering q = 0, the solution of (B.5) for f follows after some lengthy
manipulations (Chien, 1992 and Chien et aI., 1993) as

(B.6)

with

- C6 x 2 + C4 X + C.P (B.7 a)

f2 = II- Co ~/ C j
dx 2 + C6 X + C.p (B.7 b)

(B.7 c)

(B.7 d)

where all C;'s are arbitrary constants.

The condition for the existence of conservation laws also imposes the following
constraints on the bending stiffness E(x)/(x) and the inertia term m (x):

(ElY (- 2/11 Co ~/ C j
dx 3 +
B.l Elastodynamics of Inhomogeneous Bernoulli-Euler Beams 247

(B.8 a)

+m -( II -4 Co x + CI dx 2
EI
+ I C2
- C x2
o
EI
- 2C x
1 dx-

(B.8 b)

Having solved for the characteristic f, one can proceed to construct divergence-free
expressions, P: + p/
= 0 from the productj~. The resulting conserved currents are
found to be

Px = - fl (L + M ljIl + Qw I) - [11 M ljI + .!.... til EIw 12 -


2
- f2 (MljI + Qw) + [21 (Mw + EIw 12) +

+ .!....EF w 2) - .!.... f2111 Elw 2 -


f211 (Elww l -
2 2
- f3 Q + f31 M + f311 (Elw - EFw) - f3111 Elw -
l

- f4 (MljI' + Qw') , (B.9 a)

P, F mwlw' + f2 mww' +

(B.9 b)

After substitution of fl, f2, f3 and r into the above equations, the conserved
currents can be written as:

For only Co i:- 0,

Px = (L + M ~ + Qw I) (2 11/~Idx3 + IIi; dx
2
)+

+ MljI (11iI dx
2
+ / ~~ dx)- Qw l I i I dx 2
+

+ Mw
I -x dx - x ljIw - -1
EI 2
X
2 ,12 - -1
'fI w2
2'
(B.IO a)
248 Appendix B

+ mw w 'ff:] 2
dx . (RIO b)

For only C/ ~ 0,

Px = (L +MIf/ +
l
Qw ) (2lJIi] dx 3 + ffi] 2)+
2 dx

+ Mljt (/Ii] dx
2
+ 2/i] dx)- Qw /IiI dx
2
+

+ Mw 11-- E] dx - xljt2 - 1ftw ,


(Rll a)

P,=m w1w' (-2IJ1i] dx


3
-2//il dx
2
) +

+ mw w 'ffil dx
2
. (B.11 b)

For only C2 ':f: 0,

Px = - (L + Mljtl + QwI)Jll-- dx 2 - Mljt


j £1
11-- £1
dx + lljt2, (B.12 a)
2

P -- mw
,
- Iw 'Jl} -
£1
dx 2 . (Rl2 b)

For only C3 ~ 0,

Px = - Qw - M ljt , (Rl3 a)

P, = mww', (Rl3 b)

For only C4 ~ 0,

Px = - x (L + M ljt / + Qw I) - M ljt , (R14 a)

- I •
P, =x m w w . (B,14 b)

For only Cs ~ 0,

(RIS a)

P, = m w1w', (RIS b)

For only C6 ':f: 0,

Px = r (L + M ljt/ + Qw I) - X (Qw - M ljt) + Mw, (B.16 a)

(B,16 b)
B.1 Elastodynamics of Inhomogeneous Bernoulli-Euler Beams 249

For only C 7 # 0,

P =-Q/I~dx2+M/~dx+XWI_W' (B.17 a)
x f EI EI

P mw'l!~ 2 (B.17 b)
I
=
f EI dx .

For only Cs # 0,

(B.18 a)

P =mw'/ll.-dx 2. (B.18 b)
I fEI
For only C9 # 0,

Px = - xQ + M, (B.19 a)

(B.19 b)

For only ClO # 0,

(B.20 a)

P, = mw'. (B.20 b)

For only ell # 0,

Px = - t (M 1ft' + Qw") , (B.21 a)

P, = - t (L - m- w ,2 ). (B.21 b)

For only Cn # 0,

(B.22 a)

(B.22 b)

For only C/3 # 0,

(B.23 a)

P, = tm w' - mw . (B.23 b)

Integration constants are not necessary for all the integrals appearing in the
characteristic f and in the conservation laws.
250 Appendix B

It is noted that the constants C; with i == 7. 8. 9, 10, 12 and 13 do not enter the
constraint equation (B.8). Therefore, conservation laws corresponding to these
C;'s *" 0 are physical balance laws valid for all material properties and each law
holds independently. The law with C/2 *" 0 expresses the balance of energy, i. e., the
time rate of change of the total energy density (Hamiltonian) equals the net rate of
work on a volume element.

The remaining eight divergence-free expressions are material balance laws for the
dynamics of an inhomogeneous Bernoulli-Euler beam. On inspection of equation
(B.8), the two laws which correspond to C j *"
0 and CJI *"
0 cannot exist
independently. These two laws always exist in combination with the remaining six
(i == 0,1,2,4,5,6) which show various balances of material force B = (L + Qw l
+ Mlftl) and wave momentum p = (mw1w'), (cf. Chapter 6).

Given an arbitrary inhomogeneity of the beam, one can construct divergence-free


expressions in material space by solving the material constraint equation (B.8) for
the relations among the C;'s (i == O. 1,2,3.4,5,6, 11). These relations will provide
the necessary combinations of the basic laws listed in expressions (B.lO - B.23) to
obtain conserved quantities. Examples will be given in Section B.2.

For constant bending stiffness and constant mass distribution, i. e., EI, m
= const.,
a connection with the discussion in Section 6.2 can be established, where the
homogeneous bar has been treated. Equations corresponding to (6.11) and (6.12)
*"
follow from Cs 0 and CI2 0 as *"
For Cs *" 0,

Px = b;u = - (~ mw -2 + ~ M~ + QW), wave stress (B.24 a)

wave momentum (B.24 b)

For C/2 *" 0,


Px =bxl (MI/t" +Qw), stress rate of work (B.25 a)

P =b = l
I It 2
mw -2 + l2 EI ~2. Hamiltonian (B.25 b)

The two corresponding conservation laws are, then, rewritten (and coinciding with
(6.12)) as

(E. 26 a)

(E. 26 b)

It may be mentioned that wave stress and Hamiltonian for a beam do not coincide
(by contrast to a bar, where they did coincide up to a minus sign).
B.1 Elastodynamics of Inhomogeneous Bernoulli-Euler Beams 251

Dynamics of beams with transverse loading (q :#- 0)

If one allows for the presence of a transverse distributed loading q(x) acting on the
beam, the condition for the existence of conservation laws by the NA method
requires that:

(B.2? a)

[2 =j)rrCoXEI+Cl
j dx 2 + C X + C3 , (B.2? b)
6

[3 = -IIil lit qdx + JI([4' - [2) qdx J


2
dx
2
+

(B.2? c)

(B.2? d)

where all C;'s are arbitrary constants.

The constraints on the material properties, E(x)l(x) and m(x), are identical to the
dynamic case without loading as given by equations (B.8).

The corresponding conserved currents are

Px = [1 (L + M tj/ + Qw) - tiM f/t + 1 til Elw 12 -

- f2 (Mf/t + Qw) + [21 (Mw+ Elw 12 ) +

+[211 (EI ww I - l (EI) I w 2) -l [2111 Elw 2 -


2 2
- Jt3 q(x) dx - [3 Q + [-'I M + [-'II (Elw I - (EI)Iw) -

- [3111 Elw - f4 (M f/t' + Qw') , (B.28)

P, =fJ mwlw' + f2 mww' + f3 mw ' - f3. mw - f4 (L - mw· 2).


(B.29)

Since both f 3(x) and q(x) are known functions of x, the integral term in (B.28)
can always be evaluated explicitly, such that the above conserved current is a
truly divergence-free expression.
252 Appendix B

Given any q(x), EI(x) and iii (x), conservation laws can be detennined by fIrst
solving the material constraint equations (B.8) for the relation of the constants Cj ,
and then evaluating the precise form of f j,s by equations (B.27). Finally,
conservation laws valid for that system can be obtained from (B.28) and (B.29).

B.2 Reduction to Statics

By obmitting all reference to the inertia term iii and ignoring the time coordinate
t in the derivation of the previous Section, conservation laws can be obtained for a
statical theory of elementary beam bending. Thus, the Lagrangian density reduces to

L = - (W + V) = - l E(x) I(x) W 1/2 + q(x) w (B.30)


2
and the governing differential equation is given by

a =Elw w +2(EI)l w lll+(EI)ll w ll_ q (x) =0. (B.3l)

Assuming the characteristic f to be of the fonn

(B.32)

the condition E(fa) = 0 demands

(B.33)

where the functionst(x) are given in equations (B.7 a, b, c) (C 13 = 0). The material
constraint equation imposed by E(j~) = 0 is identical to (B.8 a) with
Cll = O.

Statics of beams without transverse loading (q = 0)


Having solved for the characteristic f, one can now proceed to construct divergence-
free expressions pi = 0 or P = const., from the product fa. The resulting
conservation law is found to be

p = [1 (W - M 1// - Qw I) - [11 M ljt + iF II EI w 12 - F (M ljt + Qw) +


2
+ f21 (Mw + Elw 12) + f2 II (Elww I - ~ (EI) I w 2) - f [2 III EI w 2 -
B.2 Reduction to Statics 253

(B.34)

After substitution of //, /2 and /3 into the above equation, the conserved currents
are given as:

For only Co "# 0,


2
2
P = - (W - Mf// - Qw
l
) (2IJ1i/ dx
3
+ // ~I dx ) +

+ M1ft (Iii/ dx
2
+ I ~; dx)- Qw I/i/ dx
2
+

+ Mw
I E/
x dx
- x 1ft w - "2I x 2 1ft 2 -"21 2
w . (B.35 a)

For only C/ "# 0,

P = - (W -Mf// -Qw
l
) (21J1-~/ dx
3
+2//i/ dx 2)+

+ M 1ft (jJ~/ dx 2 +2 Ii/ dx) - Qw / / i/ dx 2 +

+ Mw I...!-EI dx - X 1ft2 - 1ftw. (B.35 b)

For only C2 "# 0,

P = (W - M f// - Qw) // i/ dx 2 -

(B.35 c)

For only C3 "# 0,

P=-Qw-M1ft. (B.35 d)

For only C4 "# 0,

P =X (W -Mf// - Qw l ) -Mrft. (B.35 e)

For only C5 "# 0,

P=W-Mf//-Qw/. (B.35 f)

For only C6 "# 0,


254 Appendix B

P = -x 2 (W -Mlj/ - Qw l) -x (Qw -M1/f) +Mw. (B.35 g)

For only ~ '1= 0,

P = -QJI~dx2 +M/~dx +xwl-w. (B.35 h)


j Ef Ef

For only C8 '1= 0,

(B.35 i)

For only C9 '1= 0,

P=-xQ-M. (B.35 j)

For only C10 '1= 0,

P = ~ Q. (B.35 k)

Again, the integration constants appearing in the characteristic f and in the


conservation laws can be absorbed into the constants C3 , C4 , C5 , C6 , C9 and C 1O •

On inspection of the material constraint equation (B.8 a), it is seen that the
constants C 7 • C8 • C9 and C10 do not restrict the class of admissible inhomogeneities.
This implies that conservation laws corresponding to these constants are physical
balance laws valid for any inhomogeneity and each of them holds independently. In
particular, the expression corresponding to C10 '1= 0 relates to the constancy of the
shear force throughout the beam in the absence of a distributed transverse loading.
The expression corresponding to C9 '1= 0 embodies a relationship between the
bending moment and the shear force, and the expressions corresponding to C7 '1= 0
and C8 '1= 0 show higher order balances between the bending moment and the shear
force.

The remaining seven conserved quantities (Co. C/ ' ... , C6 '1= 0) are divergence-free
expressions in material space. Each of them expresses some balance of material
force

(B.36)

of various orders, except C3 '1= O. However, if one inspects equation (B.8 a), one will
observe that no material properties will satisfy the material constraint equation with
only C3 being nonzero, implying that this particular law exists only in combination
with others.

For a given function of the bending stiffness E(x) I(x) , the material constraint
equation (B.8 a) can be solved for the relations among the C/ s (i = 0, 1, ..., 6)
B.2 Reduction to Statics 255

and conservation laws in material space can be established. Two examples regarding
construction of material conservation laws are given below.

For a homogeneous beam, where E(x) I(x) = Elo , the material constraint equation
imposes the following constraints on the C;'s ( i = 0, I, ..., 6),

(B.37 a, b)

(B.37 c)

(B.37 d)

(B.37 e)

and the associated conservation laws in material space are:

For Cs "# 0,

p = (W - M 1// - Qw I) . (B.38 a)

For C) = - 23 c4 '
P x (W - M rj/ - Qw I) + -
3 Qw 1 Mrfr.
= + - (B.38 b)
2 2
4
For C2 = - El o C6 ,
3
P = r
(W - M rj/ - Qw I) +

+ x (Mrfr + 3Qw) - Elo ~ - 3Mw. (B.38 c)

Introducing B, Hand R by (9.36 a - c), respectively, it turns out that the set of
equations (B.38) is identical to (9.40) - (9.42).

As a second example, consider a beam with bending stiffness E(x) I(x) = Elo x4 • In
this case, equation (B.8) requires that

(B.39 a, b)

(B.39 c)

(B.39 d)

(B.39 e)

with the corresponding conservation laws being:

P= r (W - M rj/ - Qw I) - 2 (Mw + x (Qw + M rfr)] -


256 Appendix B

(BAO a)

1 3
p = x (W - M 7f/ - Qw I) + - Qw + - M ljt . (BAO b)
2 2

P = r (W - M 7f/ - Qw I) +X (Qw - M ljt) - Mw. (BAO c)

Again, these three laws show first and second order balances of material force for
the beam considered.

Statics of beams with transverse load (q * 0)


If one allows for the presence of a transverse distributed loading q(x) and
assuming the characteristic to be of the form f =f (x. W, Wi), the condition for the
existence of conservation laws by the NA method, requires that

withfl(x) andf2(x) as given by equation (B.7 a and b) and

cx+c
+
Jl 7
EI
8 dx 2 + C
9
X + C
10'
(B,41)

where all C;'s are arbitrary constants. The constraint on the bending stiffness
E(x) l(x) is identical to the case without loading as given by (B.8 a).

The conserved current valid for this system is found to be

P = F (W + V - M ~ - Qw I) - F 1M ljt + l fIll Elw 12_


2
-lf211IElw2_f2 (Mljt +Qw) +f21(Mw +Elw 12) +
2
+ f211 (Elww I -l (Elf w 2) -

-fi
2
3(X) q(x) dx - F Q + FI M + FII (Elw l - (Elfw) -

- f3111 Elw. (B,42)

Again the term jf3(X) q(x) dx can always be evaluated explicitely. To construct a
divergence-free expression for a beam under arbitrary loading q (x) , we use
B.2 Reduction to Statics 257

equations (B.7 a,b), (B.8 a) and (B.4I) to obtain the characteristic f, and construct
the corresponding conservation law by equation (B.42).

As a simple example, let us consider the case where E(x) l(x) = Elo and q(x) = qo.
For this homogeneous beam under uniform loading, the material constraint equation
has already been evaluated and imposes the constraints on the C/s as given by
(B.37). Equations (B.7 a,b) and (B.4I), then, lead to

f' (x) = - .i C6 xl + C4 X + Cs , (B.43 a)


3
f2 (x) = C6 x - i C4 , (B.43 b)
2

- qo (_ ~ C x.5 + .J..- C x4 - lC x 3) +
EIo 72 6 48 4 6.5

+ _1_ C x 3 + _1_ C x 2 + C9 X + C10 • (B.43 c)


6EIo 7 2EI0 8

With f'(x), f 2(x), and f 3 (x) known, conservation laws can be constructed directly
via equation (B.42), and the results are:

For C4 "* 0,
p = x (W + V - M rj/ - Qw /) + l M If! + i Qw -
2 2

(B.44 a)

p = (W + V - M rj/ - Qw /) (B.44 b)

For C6 "* 0,
P =- .i x 2 (W + V - M rj/ - Q W /) -
3
-x (.i M If! + Qw) + ~ Elo w 12 + Mw -
3 3

Qo ( 1 6 + x.5 .i. Q _ J.- x 4 M - J.- x 3 EI w / + 2.6 x 2 EI w).


- EI + 432 x qo 72 72 18 0 0
o
(B.44 c)

For C7 "* 0,
_ 1 x 4 qo 1 x3 Q _1'l x 2 M EI0 / EI0 w.
P - - ~ __
E
+ +X W - (B.44 d)
258 Appendix B

For Cs :#; 0,

P = - i -1
x 3 Qo Xl Q + xM + EJo w /. (B.44 e)

For C9 :#; 0,

P = -1Xl qo - xQ + M. (8.44 t)

For C/O:#; 0,

P=-xQo-Q. (8.44 g)

The laws corresponding to C5 , C4 , C6 -:/. 0 are conservation laws in material space


expressing the zeroth, first, and second order balance of material force,
respectively; and those corresponding to C7 , Cs • C9 , C/O :#; 0 are laws in physical
space showing various balances of bending moment, shear force and loading.

Given any general loading q(x), and bending stiffness E(x) l(x), corresponding
conservation laws can be constructed as in the previous examples.
Appendix C

C.I Elastodynamics of Mindlin Plates

Conservation laws and path-independent integrals for elastostatics of homogeneous


plates have been derived in Section 10.3 by using vector operations on the strain
energy density. If one is concerned with the elastodynamics of, in general,
inhomogeneous plates, Noether's formalism or the Neutral Action (NA) method are
most appropriate to derive conservation laws. In Appendices A and B, the NA
method was applied in detail. As in Chien et al. (1994), we apply Noether's theorem
involving geometric symmetry in the following.

The equations of motion of Mindlin's plate theory are recalled as

(C.l a)

(C.l b)

a3 = QJ.I + Q2.2 + P - ph w.1f = 0 , (C.I c)

The Lagrangian density in its implicit fonn follows from (10.5) and (10.6) as

L ="21 I ph W,I2 + pI 1fJ12.1 + pI 1fJ22•1 -

- Mil t/tu - M 22 t/t2.2 - M I2 (t/t2.l + t/t1,2) -

- QI (w.J + t/tl) - Q2 (w,2 + t/t2) ) + pw . (C.2)

We have three independent variables (t, Xl' X2), three dependent variables (t/tl' t/t2' w)
and the Lagrangian depends on first derivatives at most. Thus, from the general case
discussed in Chapter I (following equation 1.58) it follows

(C.3)

with m = 3, i. e., = O. f, 2,
260 Appendix C

and Jl == 3, i. e., a == 1, 2, 3.

The Lagrangian L depends explicitely on x I. , as indicated, if the stiffnesses D and


Cs , the inertia terms ph and pI, or the external pressure load p are functions of the
independent variables.

We introduce the notation

Xo == t, (CA a)

x/ == Xl' (CA b)

Xl == Xl' (CA c)

V/ == 1ft/> (CA d)

Vl == 1ftl' (CA e)

(CA f)

and the transformations to starred quantities read

t• ==t+e(o, (C.S a)

Xl
. == X/ + e (/, (C.S b)

X2
. == Xl + e (1' (C.S c)

1ft; == 1ft/ + e cP/, (C.S d)

1ft; == 1ftl + e cPl' (C.S e)

W
. == w + e cP3 . (C.S f)

We restrict ourselves to geometric symmetries, where all transformation functions


(0' ~, (2) cP/, cPl and cP3 depend on t, Xl' Xl' W, 1ft/ and 1ftz only.

The infinitesimal generator of the system governed by the Lagrangian density (C.2)
(cf. equation 1.61 a) is given by

(C.6)

and the first prolongation (cf. equation 1.61 b) written out in full is
C.l Elastodynamics of Mindlin Plates 261

+( de/>] _ d (0
rJr],1 dx
dx] ]

+ ( de/>2
dx]

+ ( de/>2 d(o
dx2
- rJr2,1 dx
2

+ ( de/>] -w -d(o -w -d(]


dt ,I dt ,] dt

+ ( de/>] -w -d(o -w -d(]


,I dx ,1 dx
dx] ] ]

+ ( de/>] -w -d(o -w -d(] (C.?)


,I dx ,1 dx
dx2 2 2

The total differentiation operators are

d 0 0 0 0
- -- +
rJr1,1 ilrJr]
+
rJr2,1 ilrJr +w -
OW
dt Ot 2
,I

-
d
--
0 +
il +
il + w.
il (i = 1,2). (C.8)
dx.I dr.I rJr1,i ilrJr rJr2,i ilrJr ,I
OW
1 2

The requirement for variational symmetries follows from (1.62) as

(C.9)

Plates with end loading

The Lagrangian density for a inhomogeneous Mindlin plate without a transverse


distributed loading p follows from (C.2) simply by neglecting the term pw.

The solution to equation (C.9) with equations (C.2, C.6 - C.8) follows a similar
strategy as outlined in Chapter 2.3 and is found to be (Chien et aI., 1994)
262 Appendix C

(C.1O a)

(C.1O b)

(C.1O c)

(C.1O d)

(C.1O e)

(C.1O f)

Similarly as in beam theory (Appendix B), equation (C.9) also imposes the
following constraints on the inhomogeneity of the material

(C.ll a)

(C.lI c)

(C.lI d)

(C.1I e)

where all A;'s are arbitrary constants.

Having solved for the transformation functions that yield variational symmetries
for this system, the associated conservation laws

(C. 12)

can be constructed directly using Equation (1.64) and the list of correspondence
(C.4) resulting in

(C.l3 a)
C.l Elastodynamics of Mindlin Plates 263

Pi = L f; + :~. (cP1 - 1ft1,t '0 - 1ft1,1 '1 - 1ft1,2 (2) +


1,.

(i = i, 2). (C.B b)

Replacing in (C.l3) the transformation functions '0' 'J' '2'


cP I' cP2 and cP3 by (C.lO),
the conserved currents are determined. In particular, we find the following
conservation laws:

For only A J :F- 0,

P, = - XI (pI 1ft J,' 1ftu + pi 1ft 2,1 1ft2.1 + ph W,. W,I) +

+ ph W W,I , (C.14 a)

(C.14 b)

(C.14 c)

For only A 2 :F- 0,

p. = ph w,.' (C.IS a)

(C.IS b)

(C.IS c)

For only A3 :F- 0,

P, = t (L - pi (1ftli - pi (1ft2i - ph (wi) , (C.16 a)

(C.16 b)

(C.16 c)

For only A4 :F- 0,

P, = L - pI (1ftJi - pi (1ft2i - ph (wi, (C.17 a)


264 Appendix C

PI = Mil Iftl.t + M 12 1ft2; + QI w; , (C.I? b)

P2 = M 2I Iftl,t + M 22 1ft2; + Q2 W" 0


(C.I? c)

For only As 'I; 0,

P, =- (pI Iftl,t Iftu + pI 1ft2,' 1ft2.r + ph w" w.r) , (C.18 a)

P J = L + Mil 1ft/,[ + M 12 1ft2,l + QI W.r ' (C.18 b)

P2 = M 21 Iftu + M 22 1ft2,l + Q2 Wlo (C.18 c)

For only A6 'I; 0,

P t = - (pI Iftl,t IftI,2 + pI 1ft2,' 1ft2,2 + ph W,t w.2) , (C.19 a)

P J = Mil 1ftJ.2 + M I2 1ft2,2 + QI w.2' (C.l9 b)

P 2 = L + M 2I IftI,2 + M 22 1ft2,2 + Q2 w,2 0 (C.19 c)

For only A7 'I; 0,

P, = - X2 (pI 1ft1.1 Iftu + pI 1ft2., 1ft2.1 + ph w,. wrJ + XI (pI Iftl; IftI.2 +

+ pI 1ft2,' 1ft2,2 + ph W,I w 2) - pI 1ft) 1ft2,' + pI 1ft2 IftJ,,' (C.20 a)

PI = x 2 ( L + Mil 1ft/,[ + M 12 1ft2.1 + QI w.J) - XI (Mil IftI.2 + M 12 1ft2.2 +

+ QI w) - Mil 1ft2 + M 12 Iftl , (Co20 b)

P 2 = x 2 (M 2I Iftl./ + M 22 1ft2.J + Q2 w.rJ - x) (L + M 2/ IftI,2 + M 22 1ft2.2 +

+ Q2 w,2) - M2I 1ft2 + M22 1ft/ . (C.20 c)

For only As 'I; 0.

Pt = - XJ ph w; + pI Iftl." (C.21 a)

PI = XI QI - Mil , (C.21 b)

P 2 = XI Q2 - M 12 0 (Co21 c)

For only A 9 'I; 0,

P, = - x 2 ph W" + pI 1ft2, I , (C.22 a)

PI = x2 QJ - M12 , (C.22 b)

P 2 = x2 Q2 - M22 0 (Co22 c)
C.l Elastodynamics of Mindlin Plates 265

Of the nine conservation laws listed above, those corresponding to A2 , As, A9 "# 0
are balance laws in physical space. With A 2 "# 0, a governing equation of the
system, .13 = 0 with p = 0, is expressed in conserved fonn; and As, A9 "# 0 express,
respectively, the combinations of (X J .13 - aJ ) and (X2 .13 - .12) in conserved fonn.
a
Here J , .12 and .13 are the governing equations of this system given by equations
(C.l). As these laws are essentially different representations of the governing
equations, they are valid for any inhomogeneous plate.

The remaining six conservation laws are balance laws in material space (with the
exception of that corresponding to only A4 "# 0). In order to interpret them, it is
useful to define the material momentum tensor, or Eshelby tensor, of elastodynamic
plate theory. The generalization of equation (10.20) to elastodynamics leads, by
analogy to equation (6.21), to

- BJI = L + MJI rfru + M/2 rfr2,J + QJ W,J' (C.23 a)

- B 2J = M2J rfru + Mn rfr2,J + Ql W,l' (C.23 b)

- B/2 = MJI rfrJ,2 + MJl rfr2,1 + QJ w,2' (C.23 c)

- B 22 = L + M2J rfrJ,2 + Mn rfr2,1 + Q2 W,2 ; (C.23 d)

and the wave-momentum or field-momentum density, cf. equation (6.22), as

Btl = pI rfrJ,t rfru + pI rfrl,t rfr2,J + ph W,t w,l , (C.24 a)

(C.24 b)

the rate of work done, cf. equation (6.23), as

(C.25 a)

(C.2S b)

and the total energy density or Hamiltonian of the plate as

- Btl = L - pI (rfrJ") 2 - pI (rfr2,,) 2 - ph (w,,) 2 . (C.26)

The remaining six conservation laws can be rewritten as:

For only A J "# 0,

(C.27 a)

(C.27 b)

(C.27 c)
266 Appendix C

expressing a first (scalar-type) moment balance of material momentum and wave


momentum.

For only A3 i:- 0,

- Pi = t Bit (i = t, I, 2) , (C.28)

expressing, what might be called, the time moment balance of energy and rate of
work done.

For Aj + 4 i:- 0, the conserved currents are

- P'fi 4 0 = B.I) (i = t, 1, 2;j= 0,1, 2), (C29)


I or A j '..

with j = 0 corresponding to only A4 i:- 0, expressing the balance of dissipation of


energy and the flux of rate of work done; and j = 1, j = 2 corresponding to
only A5 i:- 0 or A6 i:- 0, respectively, expressing the balance of material momentum
and wave momentum in the XI - and Xl -directions.

For only A7 i:- 0,

(C30 a)

(C.30 b)

(C30 c)

expressing another first (angular-type) moment balance of material momentum


and wave momentum.

Upon inspection of the material constraint equations (equations CII), the


conservation law corresponding to only A4 i:- 0 is valid for any inhomogeneity. This
conservation law is a balance law in physical space that expresses conservation of
energy. Also, it can be shown that the conservation law corresponding to only A7 i:-
o is valid for a homogeneous plate, and conservation laws corresponding to A5 i:- 0
and A6 i:- 0 are valid for a plate which is homogeneous in the x2- and X I - direction,
respectively. However, the conservation law associated with only A3 i:- 0 is never
valid independently, as no material properties will satisfy the set of material
constraints with only A3 being non-zero. Therefore, the conservation law
corresponding to A3 i:- 0 exists only in combination with others. Some examples
concerning the determination and the construction of conservation laws for
inhomogeneous plates are given below.

For a plate with varying thickness h = h (XI' x 2), the existence of material
conservation laws can be determined using the set of material constraint equations
given by equations (C. 11). With the bending stiffness D (XI' Xl) = Do (hi, the shear
stiffness C s (XI' x 2) = C so h , the moment of inertia 1 (XI' x 2) = 10 (hi, where Do,
Cso and 10 are constants; and with the density p and the Poisson's ratio u of the plate
being constants, the condition for the existence of material conservation laws can be
C.I Elastodynamics of Mindlin Plates 267

readily detennined.

If we are given a plate with varying thickness h (Xl' X2) = ho (Xl + x 2) , where ho is
a constant, equations (C.ll) will impose Al = A3 = A] = 0 and As = - A 6 • This
implies that the only material conservation law is that corresponding to As = - A6 ,
and this law can be constructed using the two basic laws for As :t; 0 and A6 :t; 0
given by equation (C.29). Setting As = 1 and A6 = - 1, the resulting conserved
currents read

(i = t, 1, 2) . (C.3l)

As a second example, we consider a plate with varying thickness h (Xl' X2) = ho


[(Xli + (x2i]. In this case, equations (C.II) require that A/ = A3 = As = A6 = 0,
and this imposes no constraint on A]. Thus the only material conservation law for
this plate is that corresponding to A] :t; 0, and this law is given by equation
(C.30).

For a plate with thickness given by h (Xl' X2) = ho [(xj + (x2i + Xl + x 2],
equations (C.ll) require that A l = A3 = 0, and that As = - A6 = 2A]. Note that
even though the thickness function here is the sum of the thicknesses of the two
previous examples, a conservation law for this plate cannot be obtained by a
superposition of the two previous conservation laws. In this case, a conservation
law is constructed by setting As =1, A6 = -1 and A] = 2. Using equations (C.29)
and (C.30), the conserved currents read

Pt = - (2X2 + 1) B t/ + (2x/ + 1) Bt2 - 2 pI rft/ rft2.t + 2 pI rft2 rftu' (C.32 a)

(C.32 b)

(C.32 c)

As a last example, we consider a plate with thickness h (x/' x 2) = ho {(xj +


(x2i]. For this specific case, equations (C.ll) require that all A/s appearing in this
set of equations are set equal to zero. This implies that, using geometric
symmetries, no material conservation law exists for such a plate.

As illustrated by these examples, to determine the existence of material


conservation laws for any given inhomogeneous plate, one must first use the set
of material constraint equations, equations (C.ll), to obtain the relations between
the A/s appearing in the set. Material conservation laws exist only for the class of
inhomogeneous plates that satisfy equations (C.ll) with at least one Ai :t; 0 . Once
the relations between the A/s are known, material conservation laws can be
constructed using the basic laws listed in equations (C.27) - (C.30).
268 Appendix C

Plates with distributed loading

If one allows for the presence of transverse distributed loading, p = p (x j • x 2), acting
on the plate, one could derive conservation laws in a similar manner as for the case
without loading. The only difference is the inclusion of the term wp corresponding
to the (negative) potential of the external loading in the Lagrangian density. Thus,
the Lagrangian density is given by (C.2).

To obtain conservation laws for this modified system, the condition for variational
symmetries requires that

(0 A3 t + A4 , (C.33 a)

(j Ajxj + A7 X 2 + A5 , (C.33 b)

(2 Aj x2 - A7 X j + A6 , (C.33 c)

4Jj A7 tfr2' (C.33 d)

4J2 - A7 tfrj, (C.33 e)

4J3 - A j w. (C.33 f)

The constraint on the material properties are as given by equations (C. 11 ), and
the additional constraint on the loading is

(C.34)

The corresponding conservation laws for this system with loading corresponding
to Aj, A3 • A4 • A5 • A6• and A7 :f. 0 are as given by equations (C.14), (C.16), (C.17),
(C.18), (C.19) and (C.20), respectively. But the Lagrangian density L appearing in
these equations is now given by equation (C.2) involving the term pw.

The interpretation of these conservation laws is the same as discussed previously


for the case without loading. Also, given any inhomogeneous plate, material
conservation laws, if any, can be constructed similarly to those presented in the
examples above. Since the governing equations for a Mindlin plate with
distributed loading cannot be represented in a conserved form, there exists no
physical conservation law for this system.
C.2 Reduction to Statics 269

C.2 Reduction to Statics

By omitting all reference to the inertia terms, pI and ph, and ignoring the time
coordinate t in the derivations of the previous Section, conservation laws can be
obtained for a statical theory of plate bending. The results describe the flexural
motion of a "statical" Mindlin plate with close resemblances to the earlier Reissner
(1944, 1945, 1947) theory.

Plates with end loading

The Lagrangian density for this special statical system without any transverse
distributed loading is given as

1
L = - - [M" rft, ,I + M 22 rft2,2 + M I2 ( rft21 f rftd +
2
+ Q, (11',1 + rft,) + Q2 (11',2 f rft2)}. (C.35)

The conditions for the existence of conservation laws, i. e., the required
transfonnation functions that yield variational symmetries and the appropriate
material constraints, can be deduced from the results of the dynamics problem.

With the infinitesimal generator of a symmetry group for the statical system given
as

(C.36)

where (; and 4>k are functions of XI' x 2' 11', rft, and rft2 for geometric symmetries, we
require that

(I = Alxl + A 7 x 2 + A" (C.37 a)

(2 = A l x 2 - A 7 X, + An' (C.37 b)

4>, = A 7 rft2 + A8 , (C.37 c)

4>2 = - A 7 rftl + A 9 , (C.37 d)

4>3 = A, 11' - A8 X, - A 9 X2 + A 2 ; (C.37 e)

and that the material propelties satisfy the following constraints


270 Appendix C

Cs.t (A, XI + A7 Xl + A.\) + CS.2 (A, .\'2- A7 XI + Ab ) + 2C s A, = 0,


(C.38 a)

(C.38 b)

(C.38 c)

where all A;'s are arbitrary constants.

The resulting conservation laws are:

For only A, i:- 0,

(C.39 a)

(C.39 b)

expressing a first (scalar-type) moment balance of material momentum.

For only A l i:- 0,

(CAO a)

(CAO b)

which is a physical balance law expressing a governing equation of this system in


conserved fonn.

(i.j~/.2), (CAl)

expressing the balance of material momentum in the xI" and xrdirection.

(CA2 a)

(CA2 b)

expressing another first (angular-type) moment balance of material momentum.

For As i:- 0,

(CA3 a)

(CA3 b)

expressing a physical balance of shear forces and moments evident from the
C.2,Reduction to Statics 271

governing equations of the system.

For A 9 "# 0,

(C.44 a)

(C.44 b)

expressing another physical balance of shear forces and moments evident from
the governing equations of the system.

The form of the B ij in the above equations follows from (C.23) with the Lagrangian
density L = - W as

B ll = W - Mll tflu - MI2 tfl2.1 QI w,I ' (C.4S a)

B]2 .. MIl tfl1,2 M]2 tfl2,2 QI w 2 ' (C.4S b)

B 2I - M2I tfll,l - Mn tfl2.1 Q2 WI' (C.4Sc)

B 22 = W- M2I tfl1,2 M22 tfl2,2 Q2 W,2 ' (C.4S d)

or, in short

(C. 46)

Rewriting the conserved currents for AI' As, AI> and A7 as

(C.47 a)

A .5, A 6 # 0 PI = BI )A) , •, , (j=1,2! (C.47 b)

(C.47 c)

it becomes obvious that the conservation laws Pi'; are identical to those in Chapter
10, cf. equations (10.31), (10.21) and (10.27). If the material is homogeneous,
equation (C.38) does not put any constraint on the constants A" A6 and A7•
Therefore, the path-independent integrals .1k and L J can be established according
to equations (10.22) and (10.28). For a constant shear stiffness C 1" however, the
constraint equation (C.38 a) is not satisfied for A I "# O. Accordingly, the
corresponding M integral, equation (10.32), is not path independent as was
discussed in Chapter 10.

Given any inhomogeneous plate, one could construct path-independent integrals


(material conservation laws), if any, following the procedure illustrated in the
examples in Section C.1.
272 Appendix C

Plates with distributed loading

If one includes the presence of a transverse load p '" P (XI' x 2 ) acting on the plate,
the Lagrangian density of this system will be modified by adding the term pw to L
in (C.35)

For this case, the condition for the existence of conservation laws requires that

(I '" Al XI + A7 x2 + A5 , (CA8 a)

(2 '" A l X2 - A7 Xl + A6 , (CA8 b)

cPi '" A7 1ft2' (CA8 c)

cP2 '" - A7 Iftl' (CA8 d)

cPJ '" Ai W , (CA8 e)

and that the material properties satisfy the constraints as given by equations
(C.38), with the constraint on the loading p (XI' x2) as

The conservation laws for this case, cOlTesponding to AI' A,. A6 and A7 l' 0 are as
given by equations (C.39), (CAl) and (C.42), with the Bij as listed in equation
(CA5), except W is to be replaced by W + V , with V ~ pw. Again, material
conservation laws for arbitrary inhomogeneous plates with loading, if any, can be
constructed in manners similar to those illustrated in Section C.l.

As only geometric symmetries were explored in Appendix C, the results presented


are not exhaustive. If one admits generalized symmetries, additional conservation
laws may be found within the present class of inhomogeneities, which may still be
enlarged. Alternatively, one may employ the Neutral Action method to construct
additional conservation laws for this Lagrangian system.
References

Anderson, T.L. (1995). Fracture Mechanics; Fundamentals and Application, CRC


Press, Boca Raton.
Arnold, V.1. (1989). Mathematical Methods of Classical Mechanics, 2nd ed.
Springer, New York.
Bakker, A. (1984). The three-dimensional I-integral. Ph.D. Thesis, Technical
University Delft.
Barber, 1.R. (1992). Elasticity. Kluwer, Dordrecht.
Bazant, Z.P. (1990). Justification and improvement of Kienzler and Herrmann's
estimate of stress intensity factors of cracked beam, Engng. Fracture Mech.
36, 523-525.
Benthem, J.P. and Koiter, W.T. (1973). Asymptotic approximations to crack
problems. In: Sih, G.c. (ed.), Mi.'c!Iunics ()j' Fracture I. Noordhoff, Leyden,
131-178.
Bessel-Hagen, E. (1921). Ober die Erhaltungssatze der Elektrodynamik. Math. Ann.
84, 258-276.
Biot, M.A. (1956). ThenTIoelasticity and irreversible thermodynamics. 1. Appl. Phys.
27, 240-253.
Bluman, G.W. and Kumei, S. (1989). Symmetries and Dijj'erential EquatioflS. App!.
Math. Sci. No. 81, Springer, New York.
Budiansky, B. and Rice, 1.R. (1973). Conservation laws and energy-release rates. J.
Appl. Mec!I. 40,201-203.
Buggisch, H., Gross, D. and Krliger, K.-H. (1981). Einige Erhaltungssatze der
Kontinuumsmechanik vom J-Integral-Typ. Ing. Arch. 50, 103-111.
Bui, H.D. (1974). Dual path independent integrals in the boundary-value problems
of cracks. Engng. Fracture Mech. 6, 287-296.
Butkov, E. (1968). Mathematical Physics. Addison-Wesley, Reading, Mass.
Chadwick, P. (1975). Application of an energy-momentum tensor in non-linear
e1astostatics. J. Elast. 5, 249-258.
Cherepanov, G.P. (1967). Crack propagation in continuous media. 1. Appl. Math.
Mech. (Engl. transl. of PMM 31 476-488) 31,503-512.
Cherepanov, G.P. (1984). Point defects in solids. In: Bilby, B.A., Muller, KJ. and
Willis, J.R. (eds.), Fundamentals of De/lmnarion and Fracture, IUTAM
Symposium Sheffield, J.D. Eshelby in Memoriam. Camblidge University
Press, Cambridge, U. K. 605-623.
Chien, N. (1992). Conservation Jaws in non-homogeneous and dissipative
mechanical systems. Ph.D. Thesis, Stanford University.
274 References

Chien, N. and Herrmann, G. (1996). Conservation laws for thenno or poroelasticity.


J. App/. Mech. 63, 331-336.
Chien, N., Honein, T. and Herrmann, G. (1993 a). Conservation laws for linear
viscoelasticity. l. angew. Math. Phys. (lAMP) 44, 44-52.
Chien, N., Honein, T. and Herrmann, G. (1993 b). Conservation laws for
nonhomogeneous Bernoulli-Euler beams. lilt. 1. Solids Structures 30, 3321-
3335.
Chien, N., Honein, T. and Herrmann, G. (1994). Conservation laws for non-
homogeneous Mindlin plates. lilt. J. ElIglIg. Sci. 32, 1125-1136.
Chien, N., Honein, T. and Hemnann, G. (1996). Dissipative systems, conservation
laws and symmetries. Int. J. Solids Structures 33, 2959-2968.
Delale, F. and Erdogan, F. (1983). The crack problem for a nonhomogeneous plane.
J. App/. Mech. 50, 609-614.
Dundurs, J. and Mura, T. (1964). Interaction between an edge dislocation and a
circular inclusion. J. Mech. Phys. Sol. 12, 177-189.
Edelen, D.G.B. (1981). Aspects of variational arguments in the theory of elasticity:
fact and folklore. lilt. 1. Solids Structures 17, 729-740.
Ehrlacher, A. (1981). Path independent integral for the calculation of the energy
release rate in elastodynamics. In: Francois, D. (ed.), Advances in Fracture
Research-Fracture 81. Pergamon, Oxford 5,2187-2194.
Eischen, J.W. (1987). Fracture of nonhomogeneous mateJials. Int. J. Fracture 34, 3-
22.
Eischen, J.W. and Hemnann, G. (1987). Energy release rates and related balance
laws in linear elastic defect mechanics. J. App/. Mech. 54, 388-392.
Erdogan, F. (1983). Stress intensity factors. J. Appt. Mech. 50, 992-1002.
Erdogan, F. (1995). Fracture Mechanics of Functionally Graded Materials.
Composites Engng. 5, 753-770.
Ericksen, J.L. (1995). Remarks concerning forces on line defects. Z. angew. Math.
Phys. (lAMP), Special Issue 46, S247-S271.
Eschenauer, H., Olhoff, N. and Schnell, W. (1997). Applied Structural Mechanics.
Splinger, Berlin.
Eshelby, J.D. (1951). The force on an elastic singularity. Phil. Trans. Roy. Soc.
London A 244, 87-112.
Eshelby, J.D. (1956). The continuum theory of lattice defects. In: Seitz, F. and
Turnbull, D. (eds.), Progress ill Solid State Physics. Academic Press, New
York 3, 79-144.
Eshelby, J.D. (1970). Energy relations and the energy-momentum tensor in
continuum mechanics. In: Kanninen, M.F., Adler, W.F., Rosenfield, A.R. and
Jaffee, R.I. (eds.), Inelastic Behavior of So/ids. McGraw-Hill, New York, 77-
115.
Eshelby, J.D. (1975a). The elastic energy-momentum tensor. 1. Etast. 5, 321-335.
Eshelby, J.D. (1975b). The calculation of energy release rates. In: Sih, G.c., van
Elst, H.C. and Broek, D. (eds.), Prospects of Fracture Mechanics.
Noordhoff, Leyden 69-84.
Eshelby, J.D. (1980). The force on a disciination in a liquid crystal. Phi/os. Mag. A
42, 359-367.
Euler, L. (1744). Methodus invellielldi /illeas curvas maximi millimive proprietate
gaudentes sive solutio proIJlematis isoperimetrici latissimo sensu accepti,
Additamentum I: De curvis elasticis. Bousquet, M.M. and Socios, Lausanne.
References 275

Federov, FJ. (1968). The01Y of elastic waves in crystals. Plenum Press, New York.
Fletcher, D.C. (1976). Conservation laws in linear elastodynamics. Arch. Rat. Mech.
Anal. 60, 329-353.
Flilgge, W. (1962). Handbook of Engineering Mechanics. McGraw Hill, New-York.
Flugge, W. (1972). Tensor Analysis and Continuum Mechanics. Springer, Berlin.
Flugge, W. (1973). Stesses in Shells. Springer, New York.
Francfort, G. and Golebiewska-Herrmann, A. (1982). Conservation laws and
material momentum in thennoelasticity. 1. Appl. Mech. 49, 710-714.
Freund, L.B. (1972). Energy flux into the tip of an extending crack in an elastic
solid. J. Elast. 2, 341-349.
Freund, L.B. (1993). Dynamic Fracture Mechanics. Cambridge University Press,
Cambridge, U. K.
Freund, L.B. (1995). On the stability of a biaxially stressed elastic material with a
free surface under variations in surface shape. Z. angew. Math. Phys.
(ZAMP) , Special Issue 46, S 185-S200.
Fung, Y.C. (1965). Foundations of solid mechanics. Prentice-Hall, Englewood
Cliffs, N.J.
Gao, H. (1991). Stress concentration at slightly undulating surfaces. 1. Mech. Phys.
Sol. 39, 443-458.
Gao, H. (1994). Some general properties of stress-driven surface evolution in a
heteroepitaxial thin film structure. .J. Mech. Phys. Solids 42, 741-772.
Gao, H. and Herrmann, G. (1992). On estimates of stress intensity factors for
cracked beams and pipes. Engng. Fracture Mech. 41, 695-706.
Gao, H., Zhang, T.- Y. and Tong, P. (1997). Local and global energy release rates
for an electrically yielded crack in a piezoelectric ceramic. J. Mech. Phys.
Sol. 45, 491-510.
Gdoutos, E.E. (1993). Fracture Mechanics: An Introduction. Kluwer, Dordrecht.
Gelfand, I.M. and Fomin, S.Y. (1963). Calculus of Variations. Prentice-Hall,
Englewood Cliffs, N.J.
Golebiewska-Herrmann, A. (1981). On conservation laws of continuum mechanics.
Int. J. Solids Structures 17, 1-9.
Golebiewska-Herrmann, A. (1982). Characterization of cracks and other defects in
elastic and thennoelastic solids. In: Current Advances in Mechanical Design
and Production, Second Cairo Univ. MDP Conference, Cairo, 97-103.
Golebiewska-Herrmann, A. (1983). On the Lagrangian fonnulation of continuum
mechanics. Physica 118 A, 300-314.
Golebiewska-Herrmann, A. and Herrmann, G. (1981). On energy-release rates for a
plane crack. 1. Appl. Mech. 48, 525-528.
Golebiewska-Herrmann, A. and HeITmann, G. (1983). Influence of boundaries on
energy changes in defonnable solids. In: Proceedings of the Symposium on
Nonlinear Problems in Energy Engineering at Argonne National Laborat01Y.
Argonne, III, 143-148.
Gould, P.L. (1994). Introduction to Linear Elasticity, 2nd ed. Springer, New York.
Griffith, A.A. (1920). The phenomena of rupture and flow in solids. Phil. Trans.
Roy. Soc. London A 221,163-198.
Gross, D. (1996). Bruchmechanik, 2nd. ed. Springer, Berlin.
Gunther, W. (1962). Ober einige Randintegrale del' Elastomechanik. Abh.
Braunschw. Wiss. Ces. 14, 53-72.
276 References

Gurtin, M.E. (1976). On a path-independent integral for elastodynamics. Int. J.


Fracture 12, 643-644.
Gurtin, M.E. (1979). On a path-independent integral for thermoelasticity. Int. J.
Fracture 15, RI69-RI70.
Gurtin, M.E. (1995). The nature of configurational forces. Arch. Rat. Mech. Anal.
131,67-100.
Gurtin, M.E. (1996). Configurational forces and the basic laws for crack
propagation. J. Mech. Phys. Solids 44, 905-927.
Gurtin, M.E. and Podiu-Guidugli, P. (1998). Configurational forces and a
constitutive theory for crack propagation that allows for kinking and curving.
J. Mech. Phys. Solids 46, 1343-1378
Hahn, RG. (1976). Bruchmechanik. Teubner, Stuttgart.
Hahn, H.G. (1985). Elastizitiitstheorie. Teubner, Stuttgart.
Hemnann, G. (1980). Some applications of invariant variational plinciples in
mechanics of solids. In: Nemat-Nasser, S. (ed.), Variational Methods in the
Mechanics of Solids. Pergamon, Oxford 145-150.
Hemnann, G. (1993). Micromechanics: Some basic methods and current trends. In:
Huet, C. (ed.), Micromechanics of Concrete and Cementitious Composites.
Presses Polytechniques et Universitaires Romandes, Lausanne, 1-18.
Hemnann, G. and KienzleI', R. (1999). On the representation of basic laws of
continuum mechanics by 4 x 4 tensors. Mee'h. Res. Comm. 26, 145-150.
Hemnann, G. and Sosa, H. (1986). On bars with cracks. Engng. Fracture Mech. 24,
889-894.
Honein, T., Chien, N. and Hernnann, G. (1991). On conservation laws for
dissipative systems. Phys. Lett. A 155, 223-224.
Honein, T. and Herrmann, G. (1988). The involution correspondence in plane
elastostatics for regions bounded by a circle. .I. Appl. Mech. 55, 566-573.
Honein, T. and Herrmann, G. (1997). Conservation laws in non-homogeneous plane
elastostatics. J. Mech. Phys. Solids. 45, 789-805.
Hubbard, J.H. and West, B.H. (1995). Difji'rential Equations: A Dynamical System
Approach. Springer, New York.
Ilschner, B. and Cherradi, N. (1995). 3rd International Symposium on Structural
and Functional Materials, Proceedings. Presses Poly techniques et
Universitaires Romandes, Lausanne.
Irwin, G.R. (1957). Analysis of stresses and strains near the end of a crack
traversing a plate. J. Appl. Mech. 24, 361-364.
Irwin, G.R. (1958). Fracture. In: Fliigge, S. (ed.), Hal/dhuch del' Physik. Springer,
Berlin VI, 551-590.
Kanninen, M.F. and PopelaI', C.H. (1985). Ad\'(/nced Fracture Mechanics. Clarendon
Press, Oxford.
Khutoryansky, N.M. and Sosa, H. (1995). Dynamic representation and fundamental
solutions for piezoelectricity. ll/t. ./. Solids Structures 32, 3307-3325.
KienzleI', R. (1982). Eine Erweiterung del' klassischen Schalentheorie; del' EinfluB
von Dickenverzerrungen und Querschnittsverwolbungen. II/g. Archiv. 52.
311-322.
KienzleI', R. (1986). On existence and completeness of conservation laws associated
with elementary beam theory. Int. .I. Solids Structures 22, 789-796.
KienzleI', R. (1993). Konzepte del' Bruchmechal/ik. Vieweg, Braunschweig.
References 277

Kienzler, R. and Duan, Z.P. (1987). On the distribution of hoop stresses around
circular holes in elastic sheets. J. Appl. Mech. 54, 110-114.
Kienzler, R. and Golebiewska-Hemnann, A. (1985). Material conservation laws in
higher-order shell theories. Int. J. Solids Structures 21, 1035-1045.
Kienzler, R. and Herrmann, G. (1986 a). On material forces in elementary beam
theory. 1. Appl. Mech. 53, 561-564.
Kienzler, R. and Herrmann, G. (1986 b). An elementary theory of defective beams.
Acta Mech. 62, 37-46.
Kienzler, R. and Herrmann, G. (1997). On the properties of the Eshelby-tensor. Acta
Mech. 125, 73-91.
Kienzler, R. and Kordisch, H. (1990). Calculation of J 1 and J 2 using the Land M
integrals. Int. J. Fracture 43, 213-225.
Kirchhoff, G.R. (1850). Uber das Gleichgewicht und die Bewegung einer elastischen
Scheibe. Crelles J. reine angew. Math. 40, 51-88.
Knott, J.F. (1973). Fundamentals of Fracture Mechanics. Butterworths, London.
Knowles, J.K. and Sternberg, E. (1972). On a class of conservation laws in
linearized and finite elastostatics. Arch. Rat. Mech. Ana\. 44, 187-211.
Koiter, W.T. (1960). A consistent first approximation in the general theory of thin
elastic shells. In: Koiter, W.T. (ed.), Proceedings of the Symposium on the
Theory of Thin Elastic Shells. N0l1h-Holland, Amsterdam 12-33.
Kratzig, W.B. (1980). On the structure of consistent liner shell theories. In: Koiter
W.B. and Mikhailov G.K. (eds.), Proceedings of the 3rd IUTAM Symposium
on Shell Theory. North-Holland, Amsterdam 353-368.
Kroner, E. (1958). Kontinuwnstheorie del' Versetzungen und Eigenspallfllmgell.
Springer, Berlin.
Kroner, E. (1993). Configurational forces and material forces in the theory of
defects in ordered structures. Material Science Forum 123-125,447-454.
Lamb, H. (1932). Hydrodynamics. 6th ed. Dover, New York.
Li, S. and Shyy, W. (1997). On invariant integrals in the Marguerre - von Kanmin
shallow shell. lnt. J. Solids Structures 34, 2927-2944.
Li, X. (1988). Dual conservation laws in elastostatics. Engng. Fracture Mech. 29,
233-241.
Li, X. (1992). Conservation laws and dual conservation laws for crack problems in
homogeneous isotropic conductive solids. Engng. Fracture Mech. 42, 51-57.
Lie, S. (1891). Vorlesungen uber Dljlerentialgleichungen mit bekannten
injinitesimalen Transformationen. Teubner, Leipzig.
Lo, K.K. (1980). Path independent integrals for cylindrical shells and shells of
revolution. Int. J. Solids Structures 16, 701-707.
Maugin, G.A. (1993). Material inhol11ogelleities in elasticity. Chapman & Hall,
London.
Maugin, G.A. (1995). Material forces: concept and applications. Appl. Mech. Rev.
48, 213-245.
McMeeking, R.M. (1990). A J-integral for the analysis of electrically induced
mechanical stress at cracks in elastic dielectrics. Int. 1. Engng. Sci. 28, 605-
613.
Mindlin, R.D. (1951). Influence of rotatory inertia and shear on flexural motions of
isotropic elastic plates. 1. Appl. Mech. 73, 31-38.
Mindlin, R.D. (1967). Micro-structure in linear elasticity. Arch. Rat. Mech. Ana\. 16,
51-78.
278 References

Mindlin, R.D. and Tiersten, H.F. (1962). Effects of couple-stresses in linear


elasticity. Arch. Rat. Mech. Anal. 11,415-448.
Mobius, A.F. (1837). Lehrbuch der Statik, Kapitel 123. Goschen, Leipzig.
Moore, S. and Samdani, G. (1993). New materials set indurance records. Chem.
Eng. 100, 5, 39-43.
Morse, P.M. and Feshbach, H. (1953). Methods of Theoretical Physics. McGraw-
Hill, New York.
MUller, W.H., Herrmann, G. and Gao, H. (1993 a). A note on curved cracked
beams. Int. J. Solids Structures 30, 1527-1532.
MUller, W.H., Hemnann, G. and Gao, H. (1993 b). Elementary strength theory of
cracked beams. Theoret. Appl. Fract. Mech. 18, 163-177.
Mura, T. (1991). Micromechanics of defects in solids, 2nd. ed. Kluwer, Dordrecht.
Naghdi, P.M. (1972). The Theory of Shells and Plates. In: Truesdell, C. (ed.),
Fliigge's Handbuch der Physik. Springer, Berlin IV a/2, 425-640.
Nicholson, J.W. and Simmonds, J.G. (1980). Sanders' energy-release rate integral
for arbitrarily loaded shalow shells and its asymptotic evaluation for circular
cylinders. J. Appl. Mech. 47, 363-369.
Nilsson, F. (1973). A path-independent integral for transient crack problems. Int. J.
Solids Structures 9, 1107-1115.
Nilsson, F. (1990). Dynamic Fracture Theory. In: Klepaczko, J.R. (ed.), Crack
Dynamics in Metallic Materials. Springer, 1-68.
Niordson, F.I. (1985). Shell TheOlY. North-Holland, Amsterdam.
Noether, E. (1918). Invariante Variationsprobleme. Nachr. Ges. Wiss. G6ttingen,
Math.-Phys. Kl. 2, 235-257.
Olver, PJ. (1984). Conservation laws in elasticity II. Linear homogeneous isotropic
elastostatics. Arch. Rat. Mecl!. Anal. 85, 131-160.
Olver, P.J. (1988). Conservation laws in elasticity III. Planar linear anisotropic
elastostatics. Arch. Rat. Mech. Anal. 102,167-181.
Olver, PJ. (1993). Applications of Lie Groups to Differential Equations, 2nd. ed.
Graduate Texts in Mathematics, No 107, Springer, New York.
Pal<, YE. (1990). Crack extension force in a piezoelectric material. J. Appl. Mech.
57,647-653.
Pak, Y.E. (1992). Linear electro-elastic fracture mechanics of piezoelectric
materials. Int. J. Fracture 54, 79-100.
Pak, Y.E. and Herrmann, G. (l986a). Conservation laws and the material
momentum tensor for the elastic dielectric. Int. J. Engng. Sci. 24, 1365-1374.
Pal<, Y.E. and Hemnann, G. (1986b). Crack extension force in a dielectric medium.
Int. J. Engng. Sci. 24, 1375-1388.
Paris, P.C. and Sih, G.C. (1965). Stress analysis of cracks. In: Fracture Toughness
Testing and its Applications. ASTM STP 381. American Society for Testing
and Materials, Philadelphia, 30-81.
Park, S. and Sun, C.-T. (1995). Fracture criteria for piezoelecu'ic ceramics. 1. Am.
Ceram. Soc. 78, 1475-1480.
Peach, M. and Koehler, J.S. (1950). The forces exerted on dislocations and the
stress field produced by them. Phys. Rev. 80, 436-439.
Pohanka, R.C. and Smith, P.L. (1988). Recent advances in piezoelectric ceramics.
In: Levinson, L. M. (ed.), Electronic ceramics. properties. devices and
applications. Marcel Dekker, New York, 45-145.
References 279

Reissner, E. (1944). On the theory of bending of elastic plates. J. Math. Phys. 23,
184-191.
Reissner, E. (1945). The effect of transverse shear deformation on the bending of
elastic plates. J. Appl. Mech. 12, A69-A77.
Reissner, E. (1947). On bending of elastic plates. Quart. Appl. Math. 5, 55-68.
Rice, l.R. (1968). A path independent integral and the approximate analysis of
strain concentration by notches and cracks. 1. Appl. Mech. 35, 379-386.
Riedel, H. (1987) Fracture at High Temperatures. Springer, Berlin.
Rosel, R. (1986). Duality under dependency inversion and Noether theory for
second-order Lagrangians. lilt. 1. Solids Structures 22, 819-832.
Sanders, J.L. (1960). On the Griffith-Irwin fracture theory. J. Appl. Mech. 27, 352-
353.
Sanders, J.L. (1982). Circumferential through-cracks in cylindrical shells under
tension. 1. Appl. Mech. 49, 103-107.
Sanders, J.L. (1983). Circumferential through-crack in a cylindrical shell under
combined bending and tension. J. Appl. Mech. 50, 221.
Schmidt, I. and Gross, D. (1995). A strategy for determining the equilibrium shape
of an inclusion. Arch. Mech. 47, 379-390.
Schmidt, I. and Gross, D. (1997). The equilibrium shape of an elastically
inhomogeneous inclusion. 1. Mech. Phys. Solids 45, 1521-1549
Schweins (1849). Fliehmomente odeI' die Summe (xX + yY) bei Kraften in der
Ebene, und (xX + yY + zZ) bei Kriiften im Raume. Crelles 1. reine angew.
Math. 38, 77-88.
Shield, R.T. (1967). Inverse defonnation results in fmite elasticity. Z. angew. Math.
Phys. (ZAMP) 18, 490-500.
Sih, G.c. (1973). Handbook of Stress Intensity Factors. Lehigh University.
Sosa, H.A. (1986). On the analysis of bars, beams and plates with defects. Ph.D.
Thesis, Stanford University.
Sosa, H.A. (1995). Defect problems in piezoelectric ceramics. In: Erdogan, F. (ed.),
Fracture Mechanics: 25th Volume, ASTM STP 1220. American Society for
Testing and Materials, Philadelphia 61-70.
Sosa, H.A. and Eischen, J.W. (1986). Computation of stress intensity factors for
plate bending via a path-independent integral. Engng. Fracture Mech. 25,
451-462.
Sosa, H.A. and Herrmann, G. (1989). On invariant integrals in the analysis of
cracked plates. Int. J. Fracture 40, 111-126.
Sosa, H. and Khutoryansky, N. (1996). New developments concerning piezoelectric
materials with defects. Int. 1. Solids Structures 33, 3399-3414.
Sosa, H.A., Rafalski, P. and Herrmann, G. (1988). Conservation laws in plate
theories. Ing. Arch. 58, 305-320.
Stark, J.P. (1976). Solid State Dijjilsion. Wiley, New York.
Suo, Z., Kuo, C.-M., Barnett, D.M., and Willis, J.R. (1992). Fracture mechanics of
piezoelectric ceramics. J. Mech. Phys. Sol. 40, 739-765.
Tada, H., Paris, P.c., and Irwin, G.R. (1973). The Stress Analysis of Cracks
Handbook. Hellertown, Pennsylvania.
Tiersten, H.F. (1969). Linear Piezoelectric Plate Vibrations. Plenum, New York.
Timoshenko, S.P. and Goodier, J.N. (1970). Theory of Elasticity, 3rd ed. McGraw-
Hill, New York.
280 References

Timoshenko, S.P. and Woinowsky-Krieger, S. (1970). Theory of Plates and Shells,


2nd ed McGraw-Hill, New York.
Toupin, R.A. (1960). Stress tensors in elastic dielectrics. Arch. Rat. Mech. Anal. 5,
440-452.
Volterra, V. (1907). Sur I'equilibre des corps elastiques multiplement connexes.
AnnIs. Scient. Ec. norm. sup. Paris 24, 401-517.
Weaver, W., Timoshenko, S.P. and Young, D.H. (1990). Vibration Problems in
Engineering, 5th ed. Wiley, New York.
Whitham, G.B. (1974). Linear and Nonlinear Waves. Wiley, New York.
Williams, M.L. (1957). On the stress distribution at the base of a stationary crack.
J. Appl. Meeh. 24, 109-144.
Symbol Index

a crack length 109, 211


d contravariant base vectors of shell middle surface 233
A area, mostly used as differential area elemend dA 8, 16
A cross-sectional area mostly in connection with compressional
stiffness EA 36, 196
A action integral 20, 58
b material loading 200, 204
b, bj , b Burgers vector, its components and its magnitude 100
bij {3 components of Eshelby tensor, material momentum 7, 63
bo:{3' bo: components of curvature tensor of shell middle sUiface (only in
Section 10.5) 234
B material force in one-dimensional problem 38, 200, 204
B body, domain of integration 8, 16
B' deformed configuration of body. B 52
BIJ components of Eshelby tensor of plate theory 226
EO:
{3 components of Eshelby tensor of shell theory 236
C specific heat per unit of volume 180
C ijkl components of tensor of elasticity 55
C compliance 206
Cs shear stiffness for plates and shells 222
C' C ' integral 171
curl ( ) curl operator, E kjj dOj /dx j 77, 148 f., 226
d differential 14
dO/dx j total differential operator 14
D flux ural stiffness for plates and shells 222
Dj components of the electric displacement vector 174
DJ higher-order total derivative 23
(-D )J signed higher-order total derivative 23
div ( ) divergence operator, dO j /dx j 14, 149 f., 227
e dilatation Uj,i 180
components of piezoelectric tensor 175
Young's modulus 36, 56
modified Young's modulus 88
components of electric field vector 174
compressional or axial stiffness 36, 196
bending stiffness 129, 198
282 Symhol Index

E(L), E.,( ) Euler operator in one (Ii = 1) and higher (Ii > l) dimensions 22 f.
g' time average of energy density 158
f,J;. characteristic in one (/I .~ 1) and higher (Ii> 1) dimensions (Neutral
Action method) 44 f.
porosity (only in Section 8.4) 192
dimensionless geometry functions correlated with crack-opening
modes I, 1I and III 107
F energy-absorption rate (only in Section 6.5) 152
F,Fj,F (physical) force vecor, its components and its magnitude 6, 7, 69
.7 time average of energy flux 158
g gravitational acceleration 6
gkij components of piezoelectric tensor 175
G shear modulus 56
GIp torsional stiffness 197 f.
(j, (ji material traction vector acting across a sUlface with normal vector n
and its components 7, 90 f.
G( UJ, k) dispersion relation 156
rJ energy-release rate 108
energy-release rate of strength-of-materials theories 207
grad ( ) gradient operator, d() /dx j 77, 147,225,235
h thickness of beam, plate or shell 211, 220
h material virialloading (only in Sections 9.3 and 9.4) 201, 204
h heat absorbed (only in Section 8.3) 180
H Hamiltonian, total energy density 7, 136
H material scalar momelit (virial) in one-dimensional problem 38, 200,
204
I area moment of inertia 129, 198, 222
Ip polar moment of inertia 197 f.
blj, b I 2' b I 3 three invariants of Eshelby tensor bi} 86 f.
"Ij , crl2' "I! three invariants of stress tensor a jj 54 f.
",12 second invariant of rotation tensor UJ jj 55
j,jj inhomogeneity-force vector (material force density) and its
components 8, 62
I I integral, material force 6, 109
j,lj vector of material forces and its components (components of I
integral) 7, 69 f.
coefficient of diffusion or thermal conductivity 162, 180
permeability coefficient (only in Section 8.4) 193
wave number 156
bulk modulus (only in Section 2.2) 56
total kinetic energy (only in Section 6.5) 152
function (111 = 1) and vector point function (111 > 1) entering in
Bessel-Hagen extension 33, 39
stress-intensity factors correlated with crack-opening modes, 1, II and
III 107
length in strengh-of-materials problem 129, 136, 196
components of (4x4) tensor defined in 142 f.
Lagrangian 19,57, 146
Symbol Index 283

I null Lagrangian 42 f.
L, L n , L vector of material moments, its components (components of L
integral) and L = L 3 in plane problems 69 f.
$£ time average of Lagrangian 158
m number of independent varables 14
m moment load per unit length of beam (only in Chapter 9 and
Appendix B) 198,245
m mass (only in Chapter 1) 6
m- mass per unit of length pA 134
mij components of a (4x4) tensor defined in 141 f.
M material virial (scalar moment), M integral 69 f.
M material content of porous medium (only in Section 804) 193
M bending moment in beam theory 129, 198
M,M.. (physical) moment vector and its components 53 f., 69
Mij bending and twisting moment resultants in plate theory 222
M'r/3 components of tensor of moments in shell theory 234
fl axial load per unit length of bar 196
n, fl j unit outward normal vector and its components 7,16
nJ, nJl, nlJl principal directions of a tensor 54
N axial force in one-dimensional problem 38, 196
N material constant of porous medium (only in Section 804) 192
N"'" components of tensor of membrane forces in shell theory 234
N<>f3 components of pseudo tensor of membrane forces in shell theory
234
OW) tenns of (vanishing) order n and higher 21
p canonical- or wave- momentum density 135, 250
p transverse pressure (load) for plates (only in Chapter 10 and
Appendix C) 222, 259
tluid pressure (only in Section XA) 192
external forces for shells 234
rate of work of the tractions (only in Section 6.5) 152
m-component vector (current), its components and current in a one-
dimensional problem 14 f., 27, 32, 36
pr(n J n-th prolongation 26, 31, 35
q transverse load (force per unit of length) in beam theory 129, 198,
245
vector of body-forces or physical force density and its components
8, 54
Q,Q" characteristic in one (,I = I) and higher (II > I) dimensions
(Noether's fonmlism) 27, 32
transverse shear force in beam theory 129, 198
transverse shear force resultants in plate theory 222
Kirchhoffs ersatz-shear force 231
transverse shear forces in shell theory 234
r radial coordinat in polar coordinate system 10 1
R second-order material scalar moment in one-dimensional problem 41,
200, 204
R,R.. field- or wave-momentum density and its components 139
RPY"B components of Riemann-Christoffel curvature tensor 236
284 Symbol Index

S arc length, mostly used as differential arc-length element ds 16 f.


Sijkl components of the compliance tensor 175
5 surface, domain of integration 8, 16 f.
5, surface of body B with prescribed tractions 58
5u surface of body B with prescribed displacements 58
5 material constant of porous medium (only in Section 8.4) 192
S,5i field intensity and its components 139
t time, as index no summation 133
t torque per unit length of a shaft (only in Chapter 9) 197
! ' !i (physical) traction vector acting across a surface with normal vector
n and its components 7, 54
T kinetic energy (density) 19, 146
T tension (only in Chapter I) 6
T reference temperature (only in Section 8.3) 180
T torque (only in Chapter 9) 197
u, Ui , U displacement vector, its components and longitudinal displacement in
one-dimensional problem 36, 52, 196
displacement gradient 7, 19, 52
dependent variable 14
derivative of dependent variable \'" with respect to independent
variable x;.d\!) dt; 14
partial derivative of "", multi-index notation 23
volume, mostly used as differential volume element dV 8, 16
(physical) virial, scalar moment or Fliehmoment 41, 69, 202
potential of external forces (density) 57, 147
w deflection in beam, plate and shell theory 129, 198,221,233
w infinitesimal generator 26, 31
W strain-energy density 19,56 f., 147
W· creep potential 171
x, Xj, x vector of independent variables, its components and independent
variable in one-dimensional problem 6, 14
a phase velocity 135
a viscosity coefficient (only in Section 7.4) 168
a material constant of a porous medium (only in Section 8.4) 192
a·I quantities not further specified (only in Chapter I) 6
ar coefficient of thennal linear expansion 180
aet{J elongation and shear strain of shell middle surface 234
{3 viscosity coefficient (only in Section 7.4) 168
{3 constant defined by (3 A. + 2)1) a r (only in Section 8.3) 180
{3ik components of the dielectric impermeability tensor 175
{3.,p curvature and twist of shell middle surface 234
Y dimensionless distance (only in Section 4.3) 101
Y angle of twist (only in Chapter 9) 197
Y/, Yl/' Ylll principal values (eigenvalues) of Eshelby tensor bij 87
Yik inverse of dielectric tensor E,k 175 f.
YJ" transverse shear strain of middle slllface 234
r (closed) line or path, domain of integration 16 f.
r specific surface energy 152
Symbol Index 285

;) ( ) variation of ( ) 20
;)jj Kronecker tensor of unity 7, 32
Ll Laplace operator 61, 224
~,~ differential operator, set of differential operators 14
£ infinitesimal group parameter 21, 31
£ strain in one-dimensional problem (only in Chapter 9) 196
£jj components of (linearized) strain tensor 52
£jjk permutation tensor, Levi-Civita tensor 52
E jk components of dielectric tensor 175
TJ viscosity coefficient 165
TJ jj components of incompatibility tensor 53
(f angle of rotation of cross-section (torsion) 197
() increment of temperature above reference temperature T 180
e phase 156
K material constant 60
A Lame constant 55 f.
A' modified Lame constant 60
Ji number of dependent variables 14
Ji Lame constant 55 f.
It couple stress vector 54
V Poisson's ratio 56
(, (; transfonnation function of independent variable in one (m = 1) and
higher (m > 1) dimensions 24, 31
II, II'0t total (potential) energy 6, 82
Ir,It external potential and internal energy 82
p mass density 15, 152,222
p dimensionless distance (only in Section 4.3) 105
a·IJ components of the Cauchy stress tensor, physical momentum 7,
54
principal stresses (eigenvalues) of stress tensor a jj 54
circumferential coordinate in a polar coordinate system 101
transfonnation function of dependent variable in one (;1 = I) and
higher (;1 > 1) dimensions 24, 31
solution of wave elJuation (only Section 6.6) 156
potential (only in Section 7.4) 165
electric potential (only in Section 8.2) 174
displacement in anti plane shear 61
angle of rotation of cross-section (bending) 198,221
angle of transformation (only in Section 3.3) 90
thennodynamic potential 176
frequency (only in Section 6.6) 156
(J, UJ j , UJ vector of rotation, its components and UJ UJ3 in plane problems
52 f., 88
UJ IJ components of rotation tensor 52
n domain of integration 20
d( )/d( )x j partial differential operator 14
d()/d()x f3 ;/J 235
( )'j total derivative, dO/dx; 14
286 Symbol Index

( y, ( )'x total derivative in a one-dimensional problem, d ( )/dx 15


()a~ covariant derivative in shell theory 234
()" derivative with respect to time t 76, 152
dO surface bounding domain Q 32
()' transformed quantity involving transformation of dependent and
independent variables 24, 31
() *() convolution in time 143
{( )J jump of ( ) 111 f.
#J order of multi-index 23
Author Index

A Ericksen, J.L., 5
Anderson, T.L., 107 Eschenauer, H., 223, 233, 234
Arnold, V.I., 1, 176 Eshelby, J.D., Y, 2, 3, 5, 56, 70, 97,
98, 200
B Euler, L., 1, 20, 195
Bakker, A., 100
Barber, J.R., 51, 59,60,75 F
Bazant, Z.P., 209, 210, 211 Federov, F.I., 175
Benthem, J.P., 212, 213, 214, 215 Feshbach, H., 49, 135, 136, 139, 140,
Bernoulli, J., 195 141
Bessel-Hagen, E., 33, 124 Fletcher, D.C., 148
Biot, M.A., 180, 192 Fliigge, W., 143, 233
Bluman, G.W., 172 Fomin, S.Y., 20
Budiansky, B., 3,72,97, 100 FrancfOtt, G., 145
Buggisch, H., 97 Freund, L.B., 94, 97, 107, 116, 146,
Bui, H.D., 72 152
Butkov, E., 161 Fung, Y.c., 207

C G
Cauchy, A.L., 195 Galilei, G., 1
Chadwick, P., 79 Gao, H., 94, 178, 179,209,211,215,
Cherradi, N., 121, 216
Cherepanov, G.P., 3, 105 Gdoutos, E.E., 107, 109
Chien, N., 45, 162, 163, 165, 172, Gelfand, I.M., 20
179, 180,246,259,261 Golebiewska-Herrmann, A., 77, 79,
93, 113, 145,233,234,237
D Goodier, J .N., 51, 58, 60, 92
Delale, F., 172 Gould, P.L., 51
Duan, Z.P., 100 Griffith, A.A., 108, 152
Dundurs, J., 104 Gross, D., 94, 107
Giinther, W., 3, 67, 124,202
E Gurtin, M.E., 4, 143, 145, 186
Edelen, D.G.B., 68
Ehrlacher, A., 153 H
Eischen, J.W., 77, 97, 100, 127, 128, Hahn, H.G., 51
146, 186,232 Hamilton, W.R., 1
Erdogan, F., 121, 127 Herrmann, G., 77,81,90,93,97,100,
288 Author Index

101, 113, 123, 134, 136, 140, N


143, 146, 177, 178, 179, 180, Naghdi, P.M., 219
186,202, 207, 209,211,215, Navier, C.L.M.H., 195
216, 228 Newton, I., 1
Hoff, N.J., 171 Nicholson, lM., 233
Honein, T., 4, 42, 101, 123, 161, 162 Nilsson, F., 146
Hubbart, lH., 89 Niordson, F.L, 233,
Noether, E., 3, 29, 124
I
Ilschner, B., 121 o
Irwin, G.R., 107, 108 Olver, P.l, 17, 18, 23, 26, 29, 33,
43 f., 68, 75, 77, 172
K
Kanninen, M.F., 107 p
Khutoryansky, N.M., 177 Pak, Y.E., 175, 177, 178
Kienzler, R., 70, 81, 90, 93, 97, 100, Paris, P.c., 107
101, 109, 113, 130, 140, 143, Park, S., 179
172,186,202,207,209,211, Peach, M., 102
233, 234, 237 Podiu-Guidugli, P., 4
Kirchhoff, G.R., 220 Pohanka, R.c., 179
Knott, J.F., 209 PopelaI', C.H., 107
Knowles, J.K, 3,67, 124
Kohler, lS., 102 R
Koiter, W.T., 212, 213, 214,215,219 Reissner, E., 220, 269
Kordisch, H., 70, 101, 113 Rice, J.R., 3, 72, 97, 100, 109,
Kriitzig, W.B., 219 Riedel, H., 171, 172
Kroner, E., 5 Rosel, R., 79, 202
Kumei, S., 172
S
L Samdani, G., 121
Lagrange, J.L., 1 Sanders, J.L., 3,216
Lamb, H., 141 Schmidt, I., 94
Lame, G., 195 Schweins, 70
Li, S., 219, 233 Shield, R.T., 79
Li, X., 72 Shyy, W., 219, 233
Lie, S., 3 Sih, G.c., 107, 108
Lo, KK, 233, 237 Simmonds, J.G., 233
Smith, P.L., 179
M Sosa, H.A., 174, 175, 177, 199,211,
Maugin, G.A., 3 219,228,229,232
McMeeking, R.M., 176 Stark, lP., 104
Mindlin, R.D., 54, 220 Sternberg, E., 3, 67, 124
Mobius, A.F., 70 Sun, c.-T., 179
Moore, S., 121 Suo, Z., 176
Morse, P.M., 49, 135, 136, 139, 140,
141 T
MUller, W.H., 211, 216 Tada, H., 108
Mura, T., 81, 104 Tiersten, H.F., 54, 174
289

Timoshenko, S.P., 51, 58, 60, 92,223,


224
Toupin, RA, 177

V
Volterra, V., 100

W
Weaver, W., 143
West, B.H., 89
Whitham, G.B., 160
Williams, M.L., 107, 127
Woinowsky-Krieger, S., 223,224
Subject Index

Boldface indicates a key reference

A three-point, 179, 217


action integral, 20·44, 58,63, 147, Bernoulli-Euler beam theory (classical
172 or elementary theory), 198,
angle 220, 245-252
of twist, 197 Bernoulli's normal hypothesis, 219
of rotation, 198 Bessel-Hagen extension, 13,33 f., 39,
anisotropy, anisotropic, 6, 116, 41,48 f., 51, 62, 67, 124
148 f., 174 f. 179 Betti-Maxwell reciprocal theorem, 75
antisymmetric, 65, 89 binominal expansion, 25
arch 195 boundary
value problem, 4, 74 f., 112 f.
condition, 43, 58-60, 129-131,
B 137, 177,207,220-224
balance bulk modulus, 56
of angular momentum, 9, 54, 73, Burgers vector, 100
141,270
of expanding (scalar) momentum
73, 142 c
of linear momentum 8 f., 42, 54, C· integral, 171 f.
73, 201 calculus
balance laws, IV, 42, 48, 68 f., 72-77, differential, 20
133-156,167,170,173-195, of variation, VI, 20, 42, 48
200-205, 224-229, 239, 250, Cauchy relation, 7, 54, 90 f.
254,265-271 cavity, V, 6, 11,72,96, 100-105,228
bar (theory), VII, 36, 38, 45, 135-138, chain rule, 34
156-158, 195-213,219,239- characteristic, 13, 27-37, 44, 72-78,
250 124, 162-172, 181-191,239,
beam (theory), VII, 9, 13, 34, 116, 245-277
129-131, 156, 195-217,219, characteristic equation, 54 f., 87, 89
224,226, 245-257, 261 f. charge
double cantilever, 128 f. electric, 174
bending, 195-213, 228, 269 Christoffel symbol, 233
292 Subject Index

Clapeyron's theorem, 207 defect, 1-6,53,71,95-100, 110, 146,


compatibility, 121 151, 169-178, 195,225
conductivity deflection, 129 f., 198, 224
coefficient of thennal, 162 degradation
configuration, V, 3,52, 71, 79, 98, electric, 174
116, 198 f., 219 deviatoric part of a tensor, 65, 88
conservation dielectric, VII, 173-176
of angular momentum, 9, 123 differential
of energy, 77, 136, 140 f., 167, form, 14
188,266 total, 231
of linear momentum, 8 f., 123 differentiation, 20, 39, 57
of mass, 141 partial, 14, 23, 30
of waves, 157 of a product (product role), 22, 27,
conservation law, VI, 3 f., 13-48, 51, 30, 35
62-79,95 f., 121-126, 133, total, 30, 75, 168
138-146, 161-173, 181-191, diffusion, 104, 121, 161, 179,209
195-205,219,224-238,239- coefficient of, 162
272 di latation, 180, 193
dual, 72 direction
trivial, 17 f., 42, 188, 191 principal, 55, 81, 86, 89
consistent theory, 219 discontinuity, 206
constitutive relation (equation or law), in stiffness, 205, 225
1,42,48,57-62,76, 79, dislocation, V, 1, 10, 53, 95 f., 100-
147,167, 171, 174-192,204, 106
223-234 dispersion (relation), 157, 159
constraint (condition or equation), 66, displacement, 32, 36, 52-61, 75, 79,
68, 75, 199 f., 221, 237, 241, 84, 99, 107, 109, 117, 122,
246-257, 262-272 128,131,137, 139, 144, 167 f.
kinematic, 198 electric, 174-177
convolution, 133, 143-146 field, 74 f., 82 f., Ill, 128
Cosserat theory, 54, 200 gradient, 19,52-57,67,75 f.,
crack, V, VII, 1,3,9,72, 100, 109- 86-88, 137 f., 177, 185
121, 151-154, 169, 174-179, plate, 221 f.
195, 206-216, 229, 232 transverse, 220 f.
tip, 2 f., 95, 107-116, 127-131, dissipation, 20, 42, 161, 167, 191
145, 152 f., 179,209 f. energy, 3, 167-170, 270
length, 4, 108 f., 179, 209, 211 dissipative system, VII, 48, 161, 172,
size, 211-215 179
creep, 171 divergence
cross-section, 36, 90-93, 132, 134, expression, 14,33,42 f., 74,81,
196-213,234 147-150, 199
current, 27-48, 63-75, 122, 135, 162- -free, 9, 63, 66, 125 f., 133, 166,
188,204-271 174, 226 f., 247-256
curvatur, 198, 234, 237 null, 18
of flux, 15, 140
theorem, 8, 15 f., 28, 32, 43, 69,
D 75, 98, 124 f., 145, 148, 154,
Darcey's law, 193 172, 178, 186,226-228
dead load, 57 duality, 4,51,79, 141,202
Suhject Index 293

Duhamel's integral, 143 equation


dynamics, 76, 156, 245-251, 269 bihannonic, 18
differential, 13-18, 24, 32, 41-49,
59,64,72,89, 105, 123, 161,
E 164, 172,219,239-245,252
eigenvalue, 54, 86, 89, 92 diffusion, 161
problem, 54, 89 energy (-balance) 152-160
elasticity, elastic, 3,19,71-79,95, Euler-Lagrange, 3, 20-23, 27-49,
171 f., 179, 185 f., 209,220 f., 62 f., 134 f., 158 f., 224, 234,
226, 231 f. 245, 263
antiplane, 60 field, 17,58,60,75, 130 f., 174-
plane, 60, 235 177,219,223
theory of, VI, 13,51-60,87,89, harmonic, lR, 61
107, 195, 227, 230 Laplace's, 18, 179
elastodynamics, VI, 79, 133, 143, 146, Navier-Lame, 58-62, 72, 74
151, 221, 224, 245, 259, 265 non-linear wave, 161, 163
elastostatics, VI, 3, 19, 32, 51, 62-n, of compatibility, 53, (il
100, 121, 133 f., 145, 151, of equilibrium, 20, 22, 37, 54-60,
174-177,224,245 123, 149 f., 165-171, 180, 192,
antiplane, 51 200, 204, 225, 234, 266
plane, 51, 59, 101, 121 f 202, of motion, 3, 11, 20, 76 f., 104,
209,211,263 135-148, 222-224
electrostatics, 3 wave, 135
energy equilibrium, 1, 10,53 f., 69,104,106,
complementary, 72 116, 119, 133, 138, 196-203,
flux (or flow) of, 15, 136, 157 f. 239
internal, 82-R4 stable, 10, 106
kinetic, 20, 76, 134, 146, 152, unstable, In, 106
222, 224, 245, 265 Euclidian space, 1,6
potential, 84, 106, 207, 265 expansion
surface, 108, 152 thelmal, 180
theorem, (relation), 29, ] 5R
total, V, 2-10,82-89,92,97-99,
103 f, 108, 116, llR, 137,208 F
energy-absorption rate 152 f failure (criterium), 109, 121, 174
energy density field
potential, 200 f., 265 coupled, 19, 173, 179
strain,13, 19 f., 56-60, 67, elastic, VII, 173-179
76-78, 82-87, 98, 102, 122- electric, VII, 173-179
130, 134, 147, 152, 167-170, intensity, 139
175-191,197-209,222-234, interacting, VII
259 thennal (temperature), VII, 173,
total. See Hamiltonian 179-191
energy-release rate, VI f., 3, 4, 72, 92, first integral, 15, 76 f.
95-100, 103, 108-116, 121, flux, 15,27, 122, 140 f., 149 f., 157,
132,146-155, 173-179, 195, 159,270
205-210,228 force
enthalpy (density), body (volume), 8 f, 54-58, 66-
electric, 176-178 82, 141-154, 161, 165, 174
294 Subject Index

configurational, VI, 2-6 Hilbert's assertation, 29


crack-extension, VI, 2, 108, Ill, hole-dislocation interaction, 100-106
173, 178 fo homogeneity, homogeneous, 1-8, 19,
driving, 2, 6, 209 36,54,61,66-77,81,86,96,
Eshelbian, V 98, 113, 116, 124-133, 138,
external, 53, 99, 234 144, 171, 178,202,226-228,
Galilean, V 240, 250, 255 fo, 263, 270, 275
inertia (d'Alembert), 141, 148 Hooke's law, 55-61, 65, 75,88, 146,
inhomogeneity, 6, 9, 67 fo 196,239
material, V, 2-11, 38, 70, 95,
102-116, 136-141,200-209,
226,241-243,250-258 I
membrane, 234 inclusion, V, 1,3,6,11,81,92,95 f.,
Newtonian, V, 2-6 104
on a defect, V, 1-6, 99 inertia (term), 133, 146, 149, 154,
on a disclination, 5 245 f., 252, 260, 269
on a singularity, V, 2, 97 rotatory, 220, 245
Peach-Kohler, 102 infinitesimal generator, 26, 30 fo, 35,
physical, V, 4-11, 38, 52, 100, 264,273
203, 240-243 intluence surface, 106
(transverse) shear, 129, 198, 220- inhomogeneity, inhomogeneous, 70,
224, 234, 254, 258, 270 121-129,146-154, 177 fo, 195-
thennodynamic, 2, 6 202, 222-226, 239-272
Fourier's law, 193 materiaL 9. 57, 66. 96. 146.225
fracture toughness, 109, 118, 174 physical, 56. 66, 225
frame, 195 initial condition, 137, 144,223 fo
free-body diagram, 4, 9, 82, 95, 114- integral
116 domain-independent, 68, 134, 146,
frequency, 156 228
functionally graded (or gradient) path-independent, 2, 70, 81,95-
material (FGM), 121 113, 121-128, 143-146, 161,
169. 173, 178, 186, 224, 233,
259, 271,
G integrating factor, 76 ro, 138
Gauss' law of electrostatics, 174 interaction
Gauss' relationship, 236 electromechanical, 174
group (theory), VI, 2, 28 ro, 31, 36 invariance, invariant, VI, 5, 13, 28 fo,
infinitesimal, 67 63,
symmetry, 273 172,219,229
condition, 29 ro, 32, 37, 63,
122
H under group action, 13
Hamiltonian, 7, 136-140, 153-159, under reparametrization, 4
250,265 fo invariants. 55, 81, 86-91, 101
Hamilton's principle, 20 inverse deformation. 79
heat involution correspondence, 101
absorbed, 180, 182 Irwin relation, 108, 210
conduction, 161, 180, 193 isotropy, isotropic, 1,6,55,67-73,
specific, 180 145, 149, 220, 227, 236
Subject Index 295

J 232, 271
J integral, 2 f., 70, 72, 81,96 r., 107- mass
118, 121-132, 145, 154, 171, density, 15, 133 f., 141,222,245,
173, 178, 185 f., 209, 226, 250, 266
230, 271 flux, 15
jet (bundle) space, 13, 26, 35 Maxwell tensor of elasticity, 3
jump mechanics
in compliance, 206 classical, V f., 1, 29
in stiffness, 206 configurational, VI, 3
teml, 206, 208 continuum, 4, 51, 79
defect, VI, 2 f., 19
Eshelbian, 3
K fracture, VI f., 2 f., 19, 70, 95,
Kelvin-Voigt model, 165, 168 107, 121, 127, 145, 152,209,
kinematic (relation), 52, 58 f., 61, 224, 229
130, 197,223,234,239 in Material Space (MiMS), V f., 2-
Kirchhoff's 6, 107
condition, 220, 223 in Physical Space (MiPS), V f., 2-
ersatz-shear force 231 6
Kronecker symbol, 32 middle surface (mid-surface, middle-
plane surface, mid-plane), 219 r.,
226 f., 233-235
L Mindlin's theory (Mindlin plate), 221-
L integral, 3, 70-72, 81, 96-100, III 224, 259, 261, 268 f.
115, 124, 142,227,231,271 mode, 95, 108
Lagrangian (density or function), 3, antiplane sliding (III), 108, 211
13,19-48,51,56-77, 121, 134, inplane sliding (II), 108 127
146-161,172,186,200,202, mixed, lO8 f.
219-236, 245, 252, 259-271 opening (I), lO8 f., 127, 178, 211
null, 42-44, 48 f. Mohr's circle, 92
Lame's constants, 55 r., 122 f., 168- moment, 70, 77, 132, 148, 224, 227,
170, 180 234, 266, 270 f.
Laplace transfonn, 146 bending, 129, 195, 198, 220, 222,
Legendre transformation, 176 254, 258
Lie group (theory), VI, 3, 13, 26, 28, material 208
load, loading physical 69, 208
material, 200, 201, 204 of inertia, 129, 198,211,213,
(mechanical), 98 f., 107-109, 222 f., 266
115-119,129 f., 148, 171, scalar (expanding or Flieh-), 38,
174-179, 195 f., 202-213, 69-71, 115,201, 208
237, 239 f., 243, 245 twisting, 195, 220, 222
Newtonian, V vector (angular), 69 f., 115 f.. 274
transverse (pressure), 195, 198, momentum, 4, 116, 135, 140
221-232,246-258,260-272 canonical, 135
material, 7, 33, 78,90, 135,201,
266,270
M physical, 7, 33, 54, 90, 135, 141
M integral, 3, 72, 81, 96-100, 111, wave (field), 136-141, 159 f.,
115, 124-128, 142,219,228, 250, 265 f.
296 Subject Index

multi-index notation, 23, 44 creep, 171


elastic, 57, 207
electric, 174-177
N kinetic, ] 9, 134
Neutral Action (NA) method, 4, 13, of external forces (loads), 57, 69,
42-51, 72, 78, 138, 161-188, 76, 147, 152, ]97-207,222-
201 f., 219, 239, 245 f., 251 227, 234, 268
256, 259, 272 thennodynamic, 176
Newton's law (equation), V, 1, 10, 76, prologation
104 first, 26, 30 f., 260
Noether's higher, 36
method (formalism or approach), second, 35
13, 19,49,78, 138,201,
219,259
second theorem, 29 R
(first) theorem, 3, 28, 33, 47 f.. 51, Reissner's (plate) theory, 220-222, 269
62, 68, 72, 123 relativity, 29
Rice's J integral. See J-integral
Riemann-Christoffel curvature Tensor,
o 236
operator, 26, 171 Riemannian
adjoint, 172 geometry, 233
differential, 26, 51, 77 space, 6
Euler, 23, 42-25, 74 162 rotation, 13, 66-78, 88 f., 92, 95, 99,
Laplace, 224 103, 116, 125, 152, ]54, ]97,
total differential, 14, 22 f, 27, 30, 220 f., 228, 233
261

S
p scaling, 72, 7X
permability coefficient, 193 material, 67, 73, 78
phase, 156-158 physical, 67 f, 73
boundary, V, 2, 4 scattering theory, 18
transition, 206 self-similar expansion, 13, 7] f., 78,
piezoceramic, 174, 178 95,100,103, ]16, 152,208 f.,
piezoelectricity, VII, 19, 173-179 228
pipe shaft, VIl, 195-215
cylindrical, 215 f. shear, 197, 220, 229
plasticity, 172 antiplane, 61
plate theory, VIl, 219-235, 259-272 correction factor, 222
Poisson's ratio, 56, 77, 108, 123-128, modulus, 56, 60 f., 100, 108,
222, 266 125 f., ] 98, 222
Poisson-Kirchhoff plate theory simple, 89
(classical plate theory), 220- shell (theory), VIl, 6, 216, 219, 233-
231 237
porous medium, VIl, 173, 192 f. similarity
potential, 19, 165, 173, 177 material 68
Bouss inesq- Papkovitsch- Neu ber, source tenn, 8, 68, 133, 160, 178 f,
59 201. 204, 207
Subject Index 297

spherical part of a tensor, 65 (variational),VI, 28, 33, 66, 68,


stability 172, 261-272
of equilibrium, 4, 9, 106 contact, 28
of plane boundaries, 92 divergence, 33
of solutions, 18 geometric, point, 28, 37, 39,
statics, 53, 114, 116, 195, 221, 225, ]24 f., 259-272
234, 252, 269 Lie-Backlund, generalized, 28,
stiffness 41 f., 272
bending (fluxural), 198, 205, 2] 6, of differential equations, 172
222, 246-258, 260, 266
compressional (axial), 36, ] 34,
196, 201-209, 240-243 T
shear, 222-232, 260, 266, 271 Taylor series, 21, 25, 35
strain, 58-61,88 f., 107, 122, 128, temperature gradient, 181-188
167 f.,]71, 174-176, 181, tension(-compression), 6, 178 f.,
188, 191, 196,234,240 195 f., 200-217
plane, 59 f., 88, 102, 108 f., tension-torsion analogy, 203
]22, 127, 145, 168,210 tensor
shear, 13], ]68, 197,220,223, (Cauchy) stress, V, 3, 7, 9, 33, 54-
230, 234 57, 64, 66, 79, 91 f., 140, 147-
strength-of-materials, VII, 2, ]9, 114, ]50, 160
139, 195 fo, 216 compliance, 175
stress, 7, 57-61, 76,82-116, 127 f., curvature, 234
133-141,165-188,191,209, dielectric (impermeability), 175 f.
219,250 energy momentum, 3, 135, 139
eigen, 101 Eshelby, V f., 3-9, 33, 63-67, 76,
field, 9, 11, 82 f., 100, 105 fo, 79,81-93, 139 f., 149, 160,
111-116,128,131, 174 173, 177, 186, 225 f., 233-236,
hydrostatic, 89 265
inplane, 220 field, 81
-intensity factor, VI fo, 2, 107- incompatibility, 53
113, 121, 127-132, 145, 195, mass stress, 141
205-216,229 material momentum, V, 3, 33,
internal, 53 135,265
normal, 54, 220 f. of elasticity, 55
plane, 10,59,60, 88 102, 108 f., permutation, Levi-Chivitii, 52, 56,
]25, 127, 145,210,233 148, 226
plate, 221-229 physical momentum, 33
principal, 54 piezoelectric, 175
shear, 54 rotation, 52-55, 88
wave, 136, 250 strain, 52-57, 88, 17]
summation convention, 15, 53, 180, thermodynamics
225 first law of, 15, 108, 136, 167
superposition, 106, 267 second law of, 4
symmetry, symmetric, 9 f., 54, 57, thennoelasticity, VI, 19, 145, 173,
64 f., 81, 86, 89, 106, 130-132, 179-188,192 f.
141, 175,209,233 f. Timoshenko beam (theory), 29, 198 f.
relation (condition), 55, 67, 146 torque, 197
symmetries, torsion, 195, 197-215
298 Subject Index

traction, 4, 54, 86, 98 f., 109-112, traction, 54 58, 90-99


125, 152, 167, 170 velocity, 135 f., 152, 167
material, 2, 7, 92, 112, 116 group, 157
physical, 2, 7, 11, 98, 177 phase, 135, 157, 160
trajectory, 4, 11, 104-106 vibration, 143
transfonnation, 23-44, 63-85, 91, 92, virial, 38, 69-71, 115, 142, 200-209,
101-103,116,122-124,172, 241
208, 232 f., 260-263, 269 second-order, 200-204, 241
generalized, 68 virtual-displacement, or virtual-work
geometric, point, 68 (theorem), 116 f., 202
phase, 18 viscoelasticity, 42, 161, 165-172
scale, similarity, 67 viscosity
symmetry, 122 coefficient of, 165, 168-170
translation, 13, 29, 71 f., 84-86, 93- void, 1, 4, 95
97, 103, 116 f., 123, 125, 147,
152, 154, 178, 205-209, 228
in material space (material), 7, 33, W
66 f., 73, 84-90, 97 f., 116, warping, 234
138, 207, 233 wave
in physical space (physical), 7, 33, action, 159 f.
66 f., 73, 139 dipersive, 156, 158
virtual, 117 f. motion, 134, 156, 160
transport theorem (assumption), 152 f. non-dispersive, 157
twist, 234 number, 156
shot:k, 18, 163
train, 157
U work
uniqueness theorem (unique solution), (external), 82-86, 99,117, 139
18,58 of external forces, 20, 83
of material forces, 207
of sud'ace tractions, 82, 86
V rate of, 5, 136, 152, 167, 170,
value 250, 265 f.
extremal, 86, 91, 93 vil1ual, 117
principal, 81, 86, 89
variation, 20 f., 43-45, 48
variational principle, 3, 42, 72, 177, y
202 yielding
invariant, 3 electrit:al, 179
variational problem, 3, 22 Young's modulus, 36, 56, 108, 127-
trivial, 42 131, 134, 165, 196, 222
vector
contravariant base, 233
couple-stress, 45
displacement, 53, 58, 133, 176
electric-displacement, 174
electric-field, 174
energy-flow, 139 f.
rotation, 52 f.

You might also like