You are on page 1of 358

Static Green’s Functions in

Anisotropic Media

This book provides the basic theory on static Green’s functions in


general anisotropic magnetoelectroelastic media and their detailed
derivations based on the complex variable method, potential method,
and integral transforms. Green’s functions corresponding to the
reduced cases are also presented, including those in anisotropic and
transversely isotropic piezoelectric and piezomagnetic media and
those in purely anisotropic elastic, transversely isotropic elastic,
and isotropic elastic media. Addressed problem domains are three-
dimensional (two-dimensional) infinite, half, and bimaterial spaces
(planes). Although the emphasis is on the Green’s functions related to
the line and point force, those corresponding to the important line and
point dislocation are also provided and discussed. This book provides
a comprehensive derivation and collection of the Green’s functions in
the concerned media, and as such, it should be a good reference book
for researchers and engineers and a textbook and reference book for
both undergraduate and graduate students in engineering and applied
mathematics.

Ernian Pan is a professor of civil engineering at the University of


Akron and a Fellow of both ASME and ASCE. He received his BS
from Lanzhou University, his MS from Peking University, and his
PhD from the University of Colorado at Boulder. His research
interests are in computational mechanics with applications to
anisotropic magnetoelectroelastic solids and nanostructures. As a well-
recognized expert in anisotropic and multilayered Green’s functions,
he has pioneered various benchmark solutions for multiphase and
multilayered composites with functionally graded materials and played
various active roles in, and contributed to, Green’s function research
and education. He has published more than 250 journal articles.

Weiqiu Chen is a professor of engineering mechanics at Zhejiang


University, China. He received his BS and PhD degrees from Zhejiang
University in 1990 and 1996, respectively. He has engaged himself in
the mechanics of smart materials/structures and vibrations/waves in
structures for more than twenty years. He has coauthored more than
three hundred peer-reviewed journal articles, and he has published
books on the elasticity of transversely isotropic elastic and piezoelectric
solids in 2006 and 2001, respectively. He has received numerous awards,
including the National Science Fund for Excellent Young Scholars
from the NSFC in 2007 and the Award of Science and Technology (the
second grade) from the Ministry of Education, China, in 2012.
Static Green’s Functions in
Anisotropic Media

Ernian Pan
University of Akron

Weiqiu Chen
Zhejiang University
32 Avenue of the Americas, New York, NY 10013-2473, USA

Cambridge University Press is part of the University of Cambridge.


It furthers the University’s mission by disseminating knowledge in the pursuit of
education, learning, and research at the highest international levels of excellence.

www.cambridge.org
Information on this title: www.cambridge.org/9781107034808
© Ernian Pan and Weiqiu Chen 2015
This publication is in copyright. Subject to statutory exception
and to the provisions of relevant collective licensing agreements,
no reproduction of any part may take place without the written
permission of Cambridge University Press.
First published 2015
Printed in the United States of America
A catalog record for this publication is available from the British Library.
Library of Congress Cataloging in Publication data
Pan, Ernian.
Static Green’s functions in anisotropic media / Ernian Pan, University of Akron,
Weiqiu Chen, Zhejiang University.
pages cm
Includes bibliographical references and index.
ISBN 978-1-107-03480-8 (hardback)
1. Green’s functions. 2. Differential equations. I. Chen, Weiqiu. II. Title.
QA372.P345 2015
537∙.24460151545–dc23    2014043070
ISBN 978-1-107-03480-8 Hardback
Cambridge University Press has no responsibility for the persistence or accuracy of URLs
for external or third-party Internet Web sites referred to in this publication and does not
guarantee that any content on such Web sites is, or will remain, accurate or appropriate.
In the other class of methods the quantities to be determined
are expressed by definite integrals, the elements of the integrals
representing the effects of singularities distributed over the surface
or through the volume. This class of solutions constitutes an
extension of the methods introduced by Green in the Theory of the
Potential.

—A. E. H. Love, 1944


Contents

Preface page xv
Acknowledgments xvii

1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.0 Introduction 1
1.1 Definition of Green’s Functions 1
1.2 Green’s Theorems and Identities 6
1.3 Green’s Functions of Potential Problems 8
1.3.1 Primary on 2D and 3D Potential Green’s Functions 8
1.3.2 Potential Green’s Functions in Bimaterial Planes 9
1.3.3 Potential Green’s Functions in Bimaterial Spaces 11
1.3.4 Potential Green’s Functions in an Anisotropic Plane or
Space 12
1.3.5 An Inhomogeneous Circle in a Full-Plane 13
1.3.5.1 A Source in the Matrix 14
1.3.5.2 A Source in the Circular Inhomogeneity 16
1.3.6 An Inhomogeneous Sphere in a Full-Space 17
1.3.6.1 A Source in the Sphere 17
1.3.6.2 A Source in the Matrix 21
1.4 Applications of Green’s Theorems and Identities 22
1.4.1 Integral Equations for Potential Problems 22
1.4.2 Boundary Integral Equations for Potential Problems 23
1.5 Summary and Mathematical Keys 24
1.5.1 Summary 24
1.5.2 Mathematical Keys 25
1.6 Appendix A: Equivalence between Infinite Series Summation
and Integral over the Image Line Source 25
1.7 References 27

2 Governing Equations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.0 Introduction 29
2.1 General Anisotropic Magnetoelectroelastic Solids 29

vii
viii Contents

2.1.1 Equilibrium Equations Including Also Those Associated


with the E- and H-Fields 30
2.1.2 Constitutive Relations for the Fully Coupled MEE Solid 30
2.1.3 Gradient Relations (i.e., Elastic Strain-Displacement,
Electric Field-Potential, and Magnetic Field-Potential
Relations) 30
2.2 Special Case: Anisotropic Piezoelectric or Piezomagnetic Solids 32
2.2.1 Piezoelectric Materials 32
2.2.2 Piezomagnetic Materials 33
2.3 Special Case: Anisotropic Elastic Solids 33
2.4 Special Case: Transversely Isotropic MEE Solids 35
2.5 Special Case: Transversely Isotropic Piezoelectric/Piezomagnetic
Solids 37
2.6 Special Case: Transversely Isotropic or Isotropic Elastic Solids 37
2.7 Special Case: Cubic Elastic Solids 39
2.8 Two-Dimensional Governing Equations 39
2.9 Extended Betti’s Reciprocal Theorem 41
2.10 Applications of Betti’s Reciprocal Theorem 41
2.10.1 Relation between Extended Point Forces and Extended
Point Dislocations 41
2.10.2 Relation between Extended Line Forces and Extended
Line Dislocations 45
2.11 Basics of Eshelby Inclusion and Inhomogeneity 46
2.11.1 The Eshelby Inclusion Problem 46
2.11.2 The Eshelby Inhomogeneity Problem 48
2.12 Summary and Mathematical Keys 49
2.12.1 Summary 49
2.12.2 Mathematical Keys 50
2.13 Appendix A: Governing Equations from the Energy Point of
View 50
2.14 Appendix B: Transformation of MEE Material Properties from
One Coordinate System to the Other 51
2.15 Appendix C: Some Important Unit Relations 54
2.16 References 54

3 Green’s Functions in Elastic Isotropic Full and


Bimaterial Planes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
3.0 Introduction 57
3.1 Antiplane vs. Plane-Strain Deformation 57
3.2 Antiplane Solutions of Line Forces and Line Dislocations 58
3.3 Plane Displacements in Terms of Complex Functions 60
3.4 Plane-Strain Solutions of Line Forces and Line Dislocations 62
3.4.1 Plane-Strain Solutions of Line Forces 62
3.4.2 Plane-Strain Solutions of Line Dislocations 63
3.5 Bimaterial Antiplane Solutions of a Line Force and a Line
Dislocation 65
3.5.1 Bimaterial Antiplane Solutions of a Line Force 65
Contents ix

3.5.2 Bimaterial Antiplane Solutions of a Line Dislocation 67


3.6 Bimaterial Plane-Strain Solutions of Line Forces and Line
Dislocations 70
3.7 Line Forces or Line Dislocations Interacting with a Circular
Inhomogeneity 76
3.7.1 Antiplane Solutions 76
3.7.1.1 A Line Force Inside or Outside a Circular
Inhomogeneity 76
3.7.1.2 A Screw Dislocation Inside or Outside a Circular
Inhomogeneity 80
3.7.2 Plane-Strain Solutions 84
3.7.2.1 Line Forces or Edge Dislocations Outside a
Circular Inhomogeneity 84
3.7.2.2 Line Forces or Edge Dislocations Inside a
Circular Inhomogeneity 93
3.8 Applications of Bimaterial Line Force/Dislocation Solutions 101
3.8.1 Image Force of a Line Dislocation 101
3.8.1.1 PK Force on a Screw Dislocation in a Bimaterial
Plane 101
3.8.1.2 PK Force on a Screw Dislocation Interacting with
a Circular Inhomogeneity 102
3.8.1.3 PK Force on an Edge Dislocation in a Bimaterial
Plane 103
3.8.1.4 PK Force on an Edge Dislocation Interacting with
a Circular Inhomogeneity 105
3.8.2 Image Work of Line Forces 107
3.8.2.1 Image Work on an Antiplane Line Force in a
Bimaterial Plane 108
3.8.2.2 Image Work on an Antiplane Line Force
Interacting with a Circular Inhomogeneity 108
3.9 Summary and Mathematical Keys 108
3.9.1 Summary 108
3.9.2 Mathematical Keys 109
3.10 References 110

4 Green’s Functions in Magnetoelectroelastic Full and


Bimaterial Planes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
4.0 Introduction 111
4.1 Generalized Plane-Strain Deformation 111
4.2 Solutions of Line Forces and Line Dislocations in a
2D Full-Plane 113
4.2.1 General Solutions of Line Forces and Line Dislocations 113
4.2.2 Green’s Functions of a Line Force 115
4.2.3 Green’s Functions of a Line Dislocation 117
4.3 Green’s Functions of Line Forces and Line Dislocations in a
Half-Plane 117
4.3.1 Green’s Functions of a Line Force in a Half-Plane 118
x Contents

4.3.2 Green’s Functions of a Line Dislocation in a Half-Plane 118


4.3.3 Green’s Functions of Line Forces and Line Dislocations
in a Half-Plane under General Boundary Conditions 119
4.4 Green’s Functions of Line Forces and Line Dislocations in
Bimaterial Planes 123
4.4.1 General Green’s Functions of Line Forces and Line
Dislocations in Bimaterial Planes 123
4.4.2 Green’s Functions of Line Forces and Line
Dislocations in Bimaterial Planes under Perfect
Interface Conditions 124
4.5 Green’s Functions of Line Forces and Line Dislocations in
Bimaterial Planes under General Interface Conditions 126
4.6 Applications in Semiconductor Industry 129
4.6.1 Basic Formulations of the Eshelby Inclusion and
Quantum Wires 129
4.6.2 Quantum Wires in a Piezoelectric Full Plane 131
4.6.3 Quantum Wires in an MEE Half-Plane 132
4.6.4 Quantum Wires in a Piezoelectric Bimaterial Plane 135
4.7 Summary and Mathematical Keys 136
4.7.1 Summary 136
4.7.2 Mathematical Keys 138
4.8 References 138

5 Green’s Functions in Elastic Isotropic Full and


Bimaterial Spaces. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
5.0 Introduction 140
5.1 Green’s Functions of Point Forces in an Elastic Isotropic
Full-Space 140
5.2 Papkovich Functions and Green’s Representation 143
5.3 Papkovich Functions in an Elastic Isotropic Bimaterial Space
with Perfect Interface 145
5.3.1 A Point Force Normal to the Interface Applied in
Material 1 146
5.3.2 A Point Force Parallel to the Interface Applied in
Material 1 150
5.4 Papkovich Functions in an Elastic Isotropic Bimaterial Space
with Smooth Interface 155
5.4.1 A Point Force Normal to the Interface Applied in
Material 1 155
5.4.2 A Point Force Parallel to the Interface Applied in
Material 1 156
5.5 Papkovich Functions for Both Perfect-Bonded and Smooth
Interfaces of Bimaterial Spaces 156
5.5.1 Papkovich Functions for a Vertical Point Force in
Material 1 156
5.5.2 Papkovich Functions for a Horizontal Point Force in
Material 1 157
Contents xi

5.6 Green’s Displacements and Stresses in Elastic Isotropic


Bimaterial Spaces 158
5.6.1 Green’s Displacements and Stresses in Bimaterial Spaces
by a Vertical Point Force 158
5.6.2 Green’s Displacements and Stresses in Bimaterial Spaces
by a Horizontal Point Force 161
5.7 Brief Discussion on the Corresponding Dislocation Solution 166
5.8 Applications: Uniform Loading over a Circular Area on the
Surface of a Half-Space 166
5.9 Summary and Mathematical Keys 172
5.9.1 Summary 172
5.9.2 Mathematical Keys 172
5.10 Appendix A: Derivatives of Some Common Functions 172
5.11 Appendix B: Displacements and Stresses in a Traction-Free
Half-Space Due to a Point Force Applied on the Surface 173
5.12 Appendix C: Displacements and Stresses in a Half Space
Induced by a Point Force Applied on the Surface with Mixed
Boundary Conditions 174
5.13 References 175

6 Green’s Functions in a Transversely Isotropic


Magnetoelectroelastic Full Space. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
6.0 Introduction 176
6.1 General Solutions in Terms of Potential Functions 176
6.2 Solutions of a Vertical Point Force, a Negative Electric Charge,
or Negative Magnetic Charge 183
6.3 Solutions of a Horizontal Point Force along x-Axis 187
6.4 Various Decoupled Solutions 191
6.4.1 Piezoelectric Green’s Functions 191
6.4.1.1 Solutions of a Vertical Point Force and a Negative
Electric Charge 193
6.4.1.2 Solutions of a Horizontal Point Force
along x-Axis 194
6.4.2 Elastic Green’s Functions 195
6.4.2.1 Solutions of a Vertical Point Force 196
6.4.2.2 Solutions of a Horizontal Point Force
along x-Axis 197
6.5 Technical Applications 198
6.5.1 Eshelby Inclusion Solution in Terms of the Green’s
Functions 198
6.5.2 Elements of the Extended Eshelby Tensor 201
6.5.3 Special Cases 208
6.5.3.1 Special Geometric Cases 208
6.5.3.2 Special Material Coupling Cases 209
6.6 Summary and Mathematical Keys 209
6.6.1 Summary 209
6.6.2 Mathematical Keys 209
xii Contents

6.7 Appendix A: The Extended Green’s Functions and Their


Derivatives 209
6.7.1 The Extended Green’s Displacements 209
6.7.2 Derivatives of the Extended Green’s Displacements 211
6.7.2.1 Derivatives of the Extended Green’s
Displacements Due to the Point Source in
K-direction (K = 3,4,5) 211
6.7.2.2 Derivatives of the Extended Green’s
Displacements Due to the Point Source in
x-direction 211
6.7.2.3 Derivatives of the Extended Green’s
Displacements Due to the Point Source in
y-direction 212
6.7.3 The Scaled Green’s Function Derivatives GJiK (l ) in
Terms of the Unit Vector l 213
6.7.3.1 Due to the Point Source in K-direction
(K = 3,4,5) 214
6.7.3.2 Due to the Point Source in x-direction 214
6.7.3.3 Due to the Point Source in y-direction 215
6.8 Appendix B: Functions Involved in the Eshelby Inclusion
Problem 216
6.8.1 A Spheroidal Inclusion (b = a/c) 216
6.8.2 Three Special Geometric Cases of Inclusion (b = a/c) 217
6.9 References 218

7 Green’s Functions in a Transversely Isotropic


Magnetoelectroelastic Bimaterial Space. . . . . . . . . . . . . . . . . . . . . . . . . . . 220
7.0 Introduction 220
7.1 Problem Description 220
7.2 Green’s Functions in a Bimaterial Space Due to Extended
Point Sources 221
7.2.1 Solutions for a Vertical Point Force, a Negative Electric
Charge, or a Negative Magnetic Charge 221
7.2.2 Solutions for a Horizontal Point Force 225
7.3 Reduced Bimaterial Spaces 230
7.3.1 Green’s Solutions for Piezoelectric Bimaterial Space 230
7.3.1.1 Solutions for a Vertical Point Force and a
Negative Electric Charge 230
7.3.1.2 Solutions for a Horizontal Point Force 232
7.3.2 Green’s Solutions for an Elastic Bimaterial Space 234
7.3.2.1 Solutions for a Vertical Point Force 234
7.3.2.2 Solutions for a Horizontal Point Force 235
7.4 Bimaterial Green’s Functions for Other Interface
Conditions 237
7.4.1 Solutions for a Smoothly Contacting and Perfectly
Conducting Interface 237
Contents xiii

7.4.2 Solutions for a Mechanically Perfect and


Electromagnetically Insulating Interface 238
7.5 Half-Space Green’s Functions 241
7.5.1 Green’s Functions for an MEE Half-Space with Free
Surface 241
7.5.2 Green’s Functions for an MEE Half-Space with Surface
Electrode 242
7.5.3 Surface Green’s Functions 243
7.5.3.1 Extended Boussinesq Solutions for a Vertical
Point Force, Electric Charge, or Magnetic Charge 243
7.5.3.2 Extended Cerruti Solutions for a Horizontal Point
Force 244
7.6 Technical Application: Indentation over an MEE Half-Space 246
7.6.1 Theory of Indentation 246
7.6.2 Indentation over an MEE Half-Space 249
7.7 Summary and Mathematical Keys 257
7.7.1 Summary 257
7.7.2 Mathematical Keys 257
7.8 References 257

8 Green’s Functions in an Anisotropic Magnetoelectroelastic


Full-Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
8.0 Introduction 260
8.1 Basic Equations in 3D MEE Full-Space 260
8.2 Green’s Functions in Terms of Line Integrals 261
8.3 Green’s Functions in Terms of Stroh Eigenvalues 265
8.4 Green’s Functions Using 2D Fourier Transform Method 267
8.5 Green’s Functions in Terms of Radon Transform 274
8.6 Green’s Functions in Terms of Stroh Eigenvalues and
Eigenvectors 275
8.6.1 General Definitions 275
8.6.2 Orthogonal Relations 276
8.6.3 Variation and Integration of Stroh Quantities in the
(m,n)-Plane and the Green’s Functions 276
8.6.4 Derivatives of Extended Green’s Displacements 277
8.7 Technical Applications of Point-Source Solutions 281
8.7.1 Couple Force, Dipoles, and Moments 281
8.7.2 Relations among Dislocation, Faulting, and Force
Moments 282
8.7.3 Equivalent Body Forces of Dislocations 285
8.8 Numerical Examples of Dislocations 287
8.9 Summary and Mathematical Keys 288
8.9.1 Summary 288
8.9.2 Mathematical Keys 290
8.10 Appendix A: Some Basic Mathematical Formulations 291
8.11 References 292
xiv Contents

9 Green’s Functions in an Anisotropic Magnetoelectroelastic


Bimaterial Space. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293
9.0 Introduction 293
9.1 Problem Description 293
9.2 Solutions in Fourier Domain for Forces in Material 1 294
9.3 Solutions in Physical Domain for Forces in Material 1 296
9.4 Solutions in Physical Domain for Forces in Material 1 with
Imperfect Interface Conditions 299
9.4.1 Imperfect Interface Type 1 299
9.4.2 Imperfect Interface Type 2 300
9.5 Special Case: Upper Half-Space under General Surface
Conditions 302
9.6 Bimaterial Space with Extended Point Forces in Material 2 304
9.7 Special Case: Lower Half-Space under General Surface
Conditions 306
9.8 Technical Application: Quantum Dots in Anisotropic
Piezoelectric Semiconductors 307
9.8.1 Analytical Integral over Flat Triangle, the Anisotropic
MEE Full-Space Case 308
9.8.2 Analytical Integral over a Flat Triangle, the Anisotropic
MEE Bimaterial Space Case 311
9.9 Numerical Examples 312
9.9.1 A Pyramidal QD in a Piezoelectric Full-Space 312
9.9.2 QD Inclusion in a Piezoelectric Half-Space 313
9.9.3 Triangular and Hexagonal Dislocation Loops in Elastic
Bimaterial Space 324
9.10 Summary and Mathematical Keys 326
9.10.1 Summary 326
9.10.2 Mathematical Keys 326
9.11 References 327

Index 329
Preface

As one of the most powerful computational methods, boundary integral equation


method (with its discretized version being called boundary element method, or
BEM), has been very successfully applied to various practical engineering problems.
The BEM has also become a regular senior-level graduate and postgraduate course
in various engineering disciplines. Because Green’s functions are the key elements in
the BEM approach, their derivations and behaviors are important to researchers as
well as to students in almost all branches of science and engineering. With advanced
materials/composites of general anisotropy being created and fabricated, and novel
devices of multiphase coupling being designed, new Green’s functions in anisotropic
and multiphase materials are in need.
This book is intended to provide the basic theory on static Green’s functions in
general anisotropic magnetoelectroelastic media and their detailed derivations based
on the complex variable method, potential method, and integral transforms. Green’s
functions corresponding to the reduced (simple) cases are also presented including
those in anisotropic and transversely isotropic piezoelectric and piezomagnetic
media, and those in purely anisotropic elastic, transversely isotropic elastic and
isotropic elastic media. Addressed problem domains are three-dimensional (two-
dimensional) infinite, half, and bimaterial spaces (planes). While the emphasis is on
the Green’s functions related to the line and point forces (the first-order source),
those corresponding to the important (line and point) dislocation source are also
provided and discussed when convenient. It is the authors’ intention that this
book provides a relatively comprehensive derivation and collection of the Green’s
functions in the concerned media, and as such, it should be a good reference book in
the hands of researchers and engineers, and a textbook and reference book for both
undergraduate and graduate students in engineering and applied mathematics.
The book is divided into nine chapters. Chapter 1 is a brief introduction to the
Green’s function method and related theorems. Chapter 2 presents the governing
equations, including the force and charge balance equations, generalized constitutive
relations, and the gradient relations between the extended displacements and
strains. While in Chapter 3 we derive the two-dimensional Green’s functions in
elastic isotropic full and bimaterial planes, the Green’s functions in corresponding
anisotropic magnetoelectroelastic full and bimaterial planes are presented in
Chapter 4. Chapter 5 includes the three-dimensional Green’s functions in elastic

xv
xvi Preface

isotropic full and bimaterial spaces. While Chapter 6 derives the three-dimensional
Green’s functions in a transversely isotropic magnetoelectroelastic full-space, the
three-dimensional Green’s functions in a transversely isotropic magnetoelectroelastic
bimaterial space are derived in Chapter 7. Chapter 8 presents the three-dimensional
Green’s functions in the corresponding anisotropic magnetoelectroelastic full-
space and Chapter 9 those in the corresponding anisotropic magnetoelectroelastic
bimaterial space. Direct and indirect applications of the Green’s functions to various
science and engineering fields are illustrated.
Acknowledgments

Both authors would like to express their sincere thanks to their family members for
their consistent and never-ending loves and supports. This is particularly true to the
first author as he started to put things together since spring 2011 when he took his
sabbatical leave. The first author would also like to thank his former graduate students
and colleagues for their contributions as cited in different chapters. Mr. Amir Molavi
Tabrizi and Mr. Ali Sangghaleh helped draw, respectively, all the figures on Green’s
function solutions in Chapter 5 and Chapter 8. Professor Peter Chung of ARL,
Professor John Albrecht of AFRL/DARPA, and Mr. Roger Green of ODOT shared
their research ideas and collaborated with the first author on various interesting topics.
Many colleagues from Green’s function and BEM communities provided constant
encouragements on writing this Green’s function book. In recent summers, the first
author visited Zhengzhou University and used his evenings and weekends on the
book draft. Constant help and support from colleagues at Zhengzhou University are
also key contributions to this book. The second author would like to thank Professor
Haojiang Ding, his Ph.D. adviser at Zhejiang University for introducing him into the
wonderful field of applied mechanics, especially the mechanics of anisotropic solids
with multifield coupling. He is extremely grateful to the National Natural Science
Foundation of China for its consistent support over the years.
Finally, both authors would like to thank Senior Editor Peter C. Gordon at
Cambridge University Press for his understanding, patience, and encouragement.
Our book draft was initially targeted on February 15, 2013, was then extended to
October 15, 2013, and finally Peter basically sets “no deadline” for this book. Thanks
so much, Peter, for your great patience! Sincere thanks also go to Sara Werden and
Jayashree Prabhu for their help with various publication issues.
The first author would like to dedicate this book to his parents. Although they
never had the chance to enter school, they have provided the first author with the
best possible learning opportunity!

xvii
1 Introduction

1.0 Introduction
Green’s function is named after George Green for his fundamental contributions
to potential theory, reciprocal relations, singular functions, and representative
theorem. Green’s functions can be extremely powerful in solving various differential
equations and are also the essential components in the boundary integral equation
method. As singular solutions to certain differential equations in their most
generalized mathematical form, such solutions can find applications in nearly every
field of science and engineering. They have been and will be continuously utilized
in earth science/geophysics; civil, mechanical, and aerospace engineering; physics
and material science; nanoscience/nanotechnology; biotechnology; information
technology, and so forth. In this chapter, we define the Green’s function and introduce
its basic features, along with derivations of some of the common Green’s functions
in potential problems.

1.1 Definition of Green’s Function


Green was a mathematician and physicist of United Kingdom (Cannell 2001; Challis
and Sheard 2003). He not only developed this powerful tool for solving linear
differential equations, but also contributed to various problems in elasticity. For
instance, he offered a derivation of the governing equations of elasticity without
using any hypothesis on the behavior of the molecular structure of the solids, and
was able to show further that twenty-one elastic constants are required in general to
account for the general anisotropy of elastic property (Timoshenko 1953). He further
explained how symmetry can reduce the independent number of these constants.
Green’s function is also called singular function, which is the fundamental
solution of a (partial or ordinary) differential equation (or system of equations) in
the problem domain (usually of infinite size) where the inhomogeneous term in the
equation is replaced by the Dirac delta function.
As an example, let us consider the following differential equation in a two-
dimensional (2D) or three-dimensional (3D) infinite and homogeneous domain,

Lu(r ) = f (r ) (1.1)

1
2 Introduction

where L is a general linear partial differential operator, f is the given inhomogeneous


term, u is the function to be solved, and r is the plane or space position vector. The
corresponding Green’s function or the fundamental solution G(r;rs) of Eq. (1.1) is
then the solution of

LG(r ; rs ) = δ (r − rs ) (1.2)

where the Dirac delta function δ(r – rs) means that when r = rs (i.e., the field point r
coincides with the source point rs), it has an infinite value; otherwise, it equals zero.
We should also mention that any domain integration of this delta function containing
the source point rs equals 1.
Once one derives the Green’s function solution of Eq. (1.2), the solution of
the corresponding inhomogeneous equation (1.1) can be simply expressed through
the method of superposition. In other words, the solution of Eq. (1.1) can be
expressed as

u(r ) = ∫ G(rs ; r ) f (rs ) dV (rs ) (1.3)


V

Remark 1.1: In an infinite and homogeneous domain, the Green’s function does not
need to satisfy any boundary condition, except for perhaps certain constraints (i.e.,
the regular conditions) at infinity. For half-space, bimaterial space, and so forth, the
Green’s function is required to also satisfy the corresponding surface or interface
conditions.
Remark 1.2: The Green’s function depends on both the source and field points, thus
being also called two-point function.
Remark 1.3: Attention should be paid to the relative location of the source and field
points, except for the Green’s function in a homogeneous infinite domain.
Remark 1.4: Relations (1.1)–(1.3) actually hold for any n-dimensional space.

Example 1: 2D (3D) Poisson’s equations in an infinite domain.


The Poisson’s equation in both 2D and 3D domains with an inhomogeneous
term f(r) can be expressed as

∇2 u(r ) = f (r ) (1.4)

where the Laplacian is defined as

 ∂2 ∂2
 ∂x 2 + ∂ y 2 (in 2D ( x, y ) - plane)

∇2 =  2 2 2 (1.5)
∂ + ∂ + ∂ (in 3D ( x, y, z) - space)
 ∂x 2 ∂y2 ∂z2

The corresponding Green’s function solutions to Eq. (1.4) (i.e., replacing the
inhomogeneous term f(r) by δ(r − rs)) are
1.1 Definition of Green’s Function 3

 −1  1 
 ln   (2D)
G(r − rs ) =  2 π  r − rs  (1.6)
 −1
(3D)
 4 π r − rs

where |r − rs| is the distance between the field and source points (in both 2D and 3D
domains). Then, based on Eq. (1.3), the particular solutions of the Poisson’s equation
(1.4) corresponding to the inhomogeneous “body force” term f(r) can be found as

 1  1 
−
 2π A
∫ f (rs ) ln   d A(rs )
 r − rs 
(2D)
u(r ) =  (1.7)
− 1 f (rs )
 4 π ∫V r − rs dV ( s )
r (3D)

Remark 1.5: The second integral expression in Eq. (1.7) for the 3D case was actually
derived by Green, who introduced and solved the electric potential due to the density
of the electricity (Green 1828).

Example 2: 2D (3D) Poisson formulae in a finite circle (sphere) with described


potential on the boundary of r = a.
For this case, both the 2D and 3D problems can be described as

∇ 2 u = 0 (r < a)

  f (θ) (2D) (1.8)
u r = a =  f (θ, ϕ) (3D)
 

where f(θ) (f(θ,ϕ) in 3D) is the potential given on the circle (sphere) in terms of θ (θ
and ϕ in 3D). For 3D, θ is the zenithal angle measured from the positive z-axis, and ϕ
is the azimuthal angle in the (x,y)-plane measured from the positive x-axis.
To find the potential within the circle (sphere in 3D) produced by the potential
described on the boundary of the circle (sphere in 3D), one can make use of the
Green’s theorem in the following text (Eq. (1.20)). However, one will need the
Green’s function solution within the circle (sphere in 3D), which satisfies the zero-
potential boundary condition (i.e., G = 0) on r = a. In other words, this special Green’s
function is the solution of the following boundary value problem:

∇ 2G = δ(r − rs ) (r < a)
 (1.9)
G r = a = 0

Fortunately, this Green’s function solution within the circle (or sphere) under the
zero-potential boundary condition can be found using the method of image. We take
the 2D case as an example to show the process. As shown in Figure 1.1, the image
point i of the inner source s is selected using the following relation

a2
oi = (1.10a)
os
4 Introduction

It can be shown that (Roach 1982) (Box 1.1)

a if
oi = (1.10b)
sf

for the field point f located arbitrarily on the surface of the circle r = a. In terms of
the position vectors, Eq. (1.10b) can be written as (after also moving all the quantities
to the left-hand side and letting ri = |ri|)

a r − ri
=1 (1.11)
ri r − rs

Box 1.1. Proof of Eq. (1.10b)

Let the angle between of and oi or os as α, then we have

2 2 2
if = of + oi − 2of oi cos α
2 2 2
sf = of + os − 2of os cos α

Making use of Eq. (1.10a) and noticing that of = a , then Eq. (1.10b) holds.

Because the full-plane potential Green’s function is in the form of ln(r) as can be
observed from Eq. (1.6), the Green’s function based on the left-hand side of Eq.
(1.11) (taking the natural logarithm of it) will satisfy the zero-potential boundary
condition on the circle of r = a. In other words, the Green’s function solution of Eq.
(1.9) can be expressed as

−1  a r − ri 
G(r ; rs ) = ln (1.12)
2 π  ri r − rs 

which can be written alternatively as

−1  1  1  1  − 1  a 
G(r ; rs ) = ln + ln + ln (1.13)
2 π  r − rs  2 π  r − ri  2 π  ri 

This is the Green’s function solution of Eq. (1.9), which holds for both the source and
field points at any location within the circle r = a (except for r = rs). It is observed
from Eq. (1.13) that the first term on the right-hand side is the original source Green’s
function, the second term the image source Green’s function, and the third term a
constant added to satisfy the zero-potential boundary condition on the circle r = a.
Then, applying this special Green’s function within the circle to the Green’s
theorem Eq. (1.20), and making use of the boundary condition of the potential
function u on the circle r = a, we have (let φ = u, ψ = G in Eq. (1.20))

u(rs , θ s ) = ∫ f (θ)qr (r ; rs , θ s )dC (r ) (1.14)


C
1.1 Definition of Green’s Function 5

f (r) i(ri)

s(rs)
a θs
o x

Figure 1.1. A circle of radius a, which is under a point source at s (with coordinate rs) within
the circle. The field point at f (with coordinate r) for the special case shown in the figure is
on the circle, and the image point of the inner source s with respect to the circle is at i (with
coordinate ri).

where the source point rs is replaced by its polar coordinates (rs,θs), the integral on
the right-hand side is on the circle r = a, and qr is the normal derivative of the Green’s
function at r, which in our case, equals ∂G/∂r. Expressing the field point r in terms
of the polar coordinate (r,θ) in the Green’s function (1.13), taking its derivative with
respect to r, and substituting the result into Eq. (1.14), one can finally arrive at the
following famous 2D Poisson integral formula

a 2 − rs2 2π f (θ)
u(rs , θ s ) =
2π ∫
0 a 2 + rs2 − 2ars cos(θ − θ s )
dθ (1.15)

This represents the solution of the potential at any inner point (rs,θs) due to a
prescribed potential function f(θ) on the circle r = a, as described by Eq. (1.8).
For the corresponding 3D case, Eq. (1.11) can be equivalently expressed as, for
the field point f at any location on the spherical surface r = a (thinking of Figure 1.1
being a 3D diagram)

1 r 1
− i =0 (1.16)
r − rs a r − ri

Thus the 3D Green’s function of Eq. (1.9) within the sphere r = a (with both the
source and field points within the sphere, except for r = rs) can be constructed as

−1 r 1
G(r ; rs ) = + i (1.17)
4 π r − rs 4 πa r − ri

Similarly, applying the Green’s theorem with one of the functions being this Green’s
function and the other being the real problem as described by Eq. (1.8), we find
the following classic 3D Poisson integral formula (using the spherical coordinates
(rs,θs,ϕs) for the source point rs)

a(a 2 − rs2 ) π 2π f (θ , ϕ )
u(rs , θ s , ϕ s ) =
4π ∫0
sin θ dθ ∫
0 (a + rs − 2ars cos γ )3/ 2
2 2
dϕ (1.18)
6 Introduction

where

cos γ = cosθ cosθ s + sin θ sin θ s cos(ϕ − ϕ s ) (1.19)

Remark 1.6: Notice the following relations when deriving Eq. (1.18)

x = r sin θ cos ϕ
y = r sin θ sin ϕ
z = r cos θ
dS = r 2 sin θ dθ dϕ

1.2 Green’s Theorems and Identities


One of the Green’s theorems is also called the divergence theorem, Gauss’s theorem,
or Green’s second identity. For any two functions that are twice differentiable with
respect to the 3D coordinates (x,y,z), the following identity holds:

∫S (φ∇ψ − ψ∇φ) ⋅ dS = ∫V (φ∇2 ψ −ψ∇2 φ)dV (1.20)

where S is the boundary of the defined volume V, dS is an oriented area element on S,


and ∇ = ix ∂ x + iy ∂ y + iz ∂ z is the gradient operator vector in the Cartesian coordinate
system. Equation (1.20) builds the important relation between the volume integral
of the functions and the boundary (surface) integral of these functions. As such, it
has the following two important applications:
(1) It can be applied to carry out the volume integral on the right hand side if the
surface integral on the left-hand side is computationally more efficient and
convenient than the volume integral. In other words, one does not need to carry
out the complicated volume integral. Similarly, if the involved surface integral
on the left-hand side is difficult to be carried out, one can just calculate the
volume integral on the right-hand side.
(2) One can express one of the terms in terms of the other three if the latter three
can be evaluated efficiently. This is particularly important in the boundary
integral equation method.
If we replace one of them in this equation by the 3D Green’s function −1/(4πr), then
the solution for the other potential can be expressed in terms of its value (potential
and its derivative) on the surface, as will be discussed in Section 1.4.
Remark 1.7: The Green’s theorem (1.20) can be derived through the simple reciprocal
relations or reciprocity that Green derived for electrostatic problems.
Green (1828) also derived the Green’s first identity, which is expressed as

∫S φ∇ ψ ⋅ dS = ∫V (φ∇2 ψ + ∇ φ ⋅∇ ψ)dV (1.21)

It is noted that Eq. (1.20) can be simply derived from Eq. (1.21).
The special case of Eq. (1.20) is when one of the functions, say ψ is constant.
Then Eq. (1.20) is reduced to
1.2 Green’s Theorems and Identities 7

∫ S
∇φ ⋅ dS = ∫ ∇2 φ dV
V
(1.22)

Or in terms of their components in 3D

 ∂φ ∂φ ∂φ   ∂2 φ ∂2 φ ∂2 φ 
∫S  ∂x dydz + ∂y dzdx + ∂z dx dy = ∫V  ∂x 2 + ∂y2 + ∂z2  dV (1.23)

Equations (1.20) to (1.23) are the Green’s relations between the volume integral
V and the surface integral S which bounds the volume V. There are also some
corresponding Green’s relations in 2D that relate a 2D area A to its closed boundary
C, which are discussed in the following text.
The first 2D relation (say in the (x,y)-plane) is similar to the 3D divergence
theorem. We let C be a positively oriented, piecewise smooth, simple, closed curve,
with its outward normal being n. We further let A be a region within the flat (x,y)-
plane enclosed by the curve C. If the vector function u has continuous first-order
partial derivatives in A, then the following relation holds

∫ C u ⋅ ndC = ∫A ∇ ⋅ udA (1.24)

Or in terms of their components

 ∂ux ∂uy 
∫ C (ux dy − u y dx ) = ∫A  ∂x
+
∂y 
dA (1.25)

Equation (1.24) or (1.25) relates the closed line integral to the integral over the area
enclosed by the line. It relates the normal component of a 2D vector function on one
side to its divergence on the other side of the equation.
The second relation between a 3D curve and a 3D surface is also called the
Green’s theorem or the Stokes’s theorem, where the tangential component of the
3D vector function on one side is related to its vector curl on the other side of
the equation. In other words,

∫ C u ⋅ dC = ∫S (∇× u) ⋅ dS (1.26)

where the oriented curve C is defined to be positive when one travels along C in the
positive direction while the enclosed surface S is on the left-hand side of the traveler.
It should be pointed out that the curve C is generally 3D (i.e., it is not required to
be in a given flat plane), and that even the curve C is in a flat plane, the surface S
enclosed can still be in 3D.
It is noted that in terms of their components, Eq. (1.26) can be expressed as

∫ C uk dxk = εkij ∫S ∂i uj dSk (1.27)

where εijk is the permutation tensor (=1 for (ijk) ∈{(123), (231), (312)}; = −1 for (ijk)
∈{(321), (132), (213)}; = 0 if i = j, or j = k, or k = i). Summation convention is implied
over the repeated index.
If the curve C and its enclosed surface A are both in a flat plane, say the (x,y)-
plane, then the Stokes’ theorem (Eq. (1.26) or (1.27)) is reduced to
8 Introduction

 ∂uy ∂ux 
∫ C (ux dx +u y dy ) = ∫A  ∂x

∂y 
dA (1.28)

This compares with the 2D divergence relation Eq. (1.25).


Remark 1.8: In solving various engineering and physics problems, one often ends
up with an integral expression as the problem solution. The Green’s theorems and
identities (Eqs. (1.20)–(1.28)) provide us with the opportunity of converting the
integral into forms that could be integrated easily.
Remark 1.9: If no analytical integration exists for the involved integral, one at
least can convert the integral into one with lower dimensions. For instance, one can
reduce a 3D integral to 2D, or a 2D integral to 1D. The integral with lower dimension
is computationally more efficient. Some of these relations have been frequently
applied in various boundary integral equation formulations.

1.3 Green’s Functions of Potential Problems

1.3.1 Primary on 2D and 3D Potential Green’s Functions


The 2D or 3D potential function u is mathematically only a scalar function with
certain features; however, it can be related to so many engineering and physics
problems, besides the well-known electric and magnetic potentials. This function can
be a quantity associated with elastic torsion, antiplane deformation, and so forth.
It can be also the seepage, aquifer, heat conduction, diffusion, and fluid motion.
As such, a basic understanding of the Green’s functions of potential problems is
important.
The full-plane (or full-space) Green’s function of the potential-type problem
satisfies the following governing equation

k ∇ 2 G = δ( r ) (1.29)

in the whole infinite domain (either 2D or 3D). The physical meaning of the
coefficient k is attached to the physics of the potential field u. For instance, it will be
the thermal conductivity if u represents the temperature change.
The Green’s function G is defined as such that for r ≠ 0, it is the solution of the
following equation

k∇ 2G = 0 (1.30)

and that it also satisfies the following integral relation around the source point r = 0,
due to the property of the delta function δ,

∫V k∇ 2GdV = 1
ε
(1.31)

where Vε is a circle (sphere) of arbitrary radius ε centered at the source point r = 0.


The way to find the Green’s function is to first solve the homogeneous Eq. (1.30) and
1.3 Green’s Functions of Potential Problems 9

then apply the condition (1.31) to determine the unknown coefficients, keeping in
mind that the radius ε is of arbitrary length. By following these steps (Brebbia and
Dominguez, 1996), the Green’s function of Eq. (1.29) at the field point r when the
source is applied at the origin is found to be

 −1 1
 2 πk ln r (2 D)
G(r ) =  (1.32)
 −1 (3D)
 4 πkr

where r = | r − 0 | = | r | is the distance of the field point to the source point at the
origin. This Green’s function solution is essentially the same as that in Eq. (1.6) when
the material property parameter is normalized to k = 1.
Remark 1.10: Adding an arbitrary 2D/3D harmonic function to the right-hand side
of Eq. (1.32) still gives us the Green’s function solution of Eq. (1.29). A special case is
that the harmonic function is a constant, say, equal to b. Then an alternative Green’s
function solution of Eq. (1.29) is

 −1 1
 2 πk ln r + b (2 D)
G(r ) =  (1.33)
 −1 + b (3D)
 4 πkr

In other words, the potential Green’s function could be unique up to an arbitrary


harmonic function.

1.3.2 Potential Green’s Functions in Bimaterial Planes


The method of image presented in this chapter for the potential problem in a
circle (sphere) domain can be utilized to find the potential Green’s functions in
bimaterial planes/spaces. We start with the bimaterial plane case first. We assume
that in the y > 0 half-plane (with material property k(1)), there is a source of unit
magnitude applied at (xs,ys > 0), and the y < 0 half-plane (with material property
k(2)) is free of any source. Along the interface y = 0, we assume the perfect
interface condition, that is, both the potential G and its normal flux qy = k∂G/∂y
are continuous. As such, the governing equations are

k (1) ∇2G(1) = δ ( x − xs )δ ( y − ys ) ( y, ys > 0)


(1.34)
k ( 2 ) ∇ 2 G( 2 ) = 0 ( y < 0)

For the perfect interface continuity conditions on y = 0, we have

G(1) = G( 2) , q(y1) = q(y2) (1.35)

The Green’s function solutions at any field point (x,y) under this perfect interface
condition are
10 Introduction

 −1 1 −(k (1) − k ( 2) ) 1
+
 2πk (1) r 2πk (1) (k (1) + k ( 2) ) ln r
ln ( y > 0)
1 2
G( x, y; xs , ys ) =  (1.36)
 − 1 1
ln ( y < 0)
 π(k (1) + k ( 2) ) r1

where

r1 = ( x − xs )2 + ( y − ys )2
(1.37)
r2 = ( x − xs )2 + ( y + ys )2

are, respectively, the distance between the field point (x,y) and the source (xs,ys) and
the distance between the field point and the image source (xs,−ys).
Remark 1.11: The results for the source in the lower half-plane can be found by
simply switching “(1)” and “(2)” attached to the material coefficient k, and switching
the subscripts between 1 and 2 of the distance r.
Remark 1.12: When the material properties of the two half-planes are identical (i.e.,
k(1) = k(2) = k), Eq. (1.36) is reduced to the Green’s function in the full-plane with
material property k.
Remark 1.13: When k(2) = 0, we have the flux-free boundary condition at y = 0 for the
upper half-plane, and the corresponding half-plane solution with material property
k(1) = k in the y > 0 half-plane is

−1 1
G( x, y; xs , ys ) = ln (1.38)
2 πk r1r2

Remark 1.14: When G = 0 on the boundary y = 0 (i.e., k(2) approaches infinity), the
corresponding half-plane solution in y > 0 is

−1 r
G( x, y; xs , ys ) = ln 2 (1.39)
2 πk r1

Clearly, the preceding Green’s function (1.39) can be used to construct the following
Poisson integral formula

ys ∞ f ( x)
u( xs , ys ) =
πk ∫ −∞ ( x − xs )2 + ys2
dx (1.40)

with f(x) = u(x,0) being prescribed on y = 0. Thus, the potential function u, which
satisfies the 2D Laplace equation and is known on y = 0, can be calculated from Eq.
(1.40) so that its value at any interior point in the upper half-plane can be obtained.
Remark 1.15: It should note here that, the integral in Eq. (1.40) should be over a
closed contour, say with an additional large semicircle of radius R. It can be shown
that, to eliminate the integral over the semicircle, the potential prescribed on the
boundary should satisfy the condition u = O(Rα) with α < 1 as R→∞ (Greenberg
1971).
1.3 Green’s Functions of Potential Problems 11

1.3.3 Potential Green’s Functions in Bimaterial Spaces


The method of image can be equally successfully applied to the bimaterial space
case. In this case, we assume that in the z > 0 half-space (with material property k(1)),
there is a source of unit magnitude at (xs,ys, zs > 0), and the z < 0 half-space (with
material property k(2)) is free of any source. Along the interface z = 0, we consider
the perfect interface condition, that is, both the potential G and its normal flux
qz (= k∂G/∂z) are continuous. As such, the governing equations are

k (1) ∇2G(1) = δ ( x − xs )δ ( y − ys )δ (z − zs ) (z, zs > 0)


(1.41)
k ( 2 ) ∇ 2 G( 2 ) = 0 ( z < 0)

and the interface continuity conditions (on z = 0) are

G(1) = G( 2) , qz(1) = qz( 2) (1.42)

The Green’s function solutions at the field point (x,y,z) under this perfect interface
condition are

 −1 1 −(k (1) − k ( 2) ) 1
 4πk (1) R + 4πk (1) (k (1) + k ( 2) ) R ( z > 0)
G( x, y, z; xs , ys , zs ) =  1 2 (1.43)
−1 1
 (z < 0)
 2 π(k (1) + k ( 2) ) R1

where

R1 = ( x − xs )2 + ( y − ys )2 + (z − zs )2
(1.44)
R2 = ( x − xs )2 + ( y − ys )2 + (z + zs )2

are, respectively, the distance between the field point (x,y,z) and the source (xs,ys, zs),
and the distance between the field point and the image source (xs,ys,−zs).
Remark 1.16: The results for the source in the lower half-space can be found by
simply switching “(1)” and “(2)” attached to the material coefficient k, and switching
the subscripts between 1 and 2 of the distance R.
Remark 1.17: When the material properties of the two half-spaces are identical
(i.e., k(1) = k(2) = k), the solution is then reduced to the full-space Green’s function
with material property k.
Remark 1.18: When k(2) = 0, we have the flux-free boundary condition at z = 0 for the
upper half-space, and the corresponding half-space solution with material property
k(1) = k in the z > 0 half-space is

−1  1 1 
G( x, y, z; xs , ys , zs ) = + (1.45)
4 πk  R1 R2 

Remark 1.19: When G = 0 on the interface z = 0 (i.e., k(2) approaches infinity), then
the corresponding half-space solution is
12 Introduction

−1  1 1 
G( x, y, z; xs , ys , zs ) = − (1.46)

4 πk  R1 R2 

Similarly, when the potential function u is prescribed on z = 0, that is, u(x,y,0) =


f(x,y), we can use the following Poisson integral formula to calculate u at an arbitrary
interior point

zs ∞ ∞ f ( x, y)
u( xs , ys , zs ) =
2π k ∫−∞ ∫−∞
( x − xs )2 + ( y − ys )2 + zs2 
3/ 2
dxdy (1.47)

This formula can be obtained by making use of the Green’s identity (1.20) and
Green’s function (1.46). Just as in the 2D case, certain restrictions on the prescribed
potential function u at infinity should be imposed so that Eq. (1.47) can hold strictly.
This can be shown to be u = O(Rα) with α < 0 as R→∞.

1.3.4 Potential Green’s Functions in an Anisotropic Plane or Space


For the general anisotropic case, the Green’s function (due to a point source at r = 0)
in both 2D and 3D should satisfy

N
∂ 2G
∑ kij ∂x ∂x = δ( r ) (1.48)
i, j = 1 i j

where N = 2 and 3 correspond, respectively, to the 2D and 3D domains, and the


material coefficient matrix [kij] is symmetric (i.e., kij = kji).
The potential Green’s function in the anisotropic 2D (x1,x2)-plane due to a
source of unit magnitude at the origin can be expressed as

−1 1
G= ln (1.49)
2π ∆ r

where
2
∆ = k11k22 − k12 (1.50)

and

r = ( s11 x12 + 2 s12 x1 x2 + s22 x22 )1/ 2


(1.51)
[ sij ] = [kij ]−1

Similarly, the potential Green’s function in the corresponding anisotropic 3D space


due to a source of unit magnitude at the origin is

−1
G= (1.52)
4π ∆ r
1.3 Green’s Functions of Potential Problems 13

where

∆ = det[kij ] (1.53)

and

r = ( s11 x12 + s22 x22 + s33 x32 + 2 s12 x1 x2 + 2 s13 x1 x3 + 2 s23 x2 x3 )1/ 2
(1.54)
[ sij ] = [kij ]−1

It is noticed that for a general anisotropic material, the flux qi is defined as


N
∂G
qi = ∑ kij ∂x j (1.55)
j =1

for both 2D (i,j = 1,2) and 3D (i,j = 1,2,3). This relation is required when solving the
boundary value problem.
Remark 1.20: These 2D and 3D Green’s functions contain the corresponding isotropic
cases as special cases, although the reduced expression from Eq. (1.49) differs to that
in Eq. (1.32) by a constant term.
Remark 1.21: When the source is at an arbitrary point y, one needs only to replace
xi by xi − yi.
Remark 1.22: The accuracy of these anisotropic Green’s functions can be easily
verified by substituting the solution back to Eq. (1.48) when r ≠ 0. The multiplication
factor ∆ is determined by the integral property of the delta function δ(r).
Remark 1.23: More complicated Green’s functions in the anisotropic space (half,
bimaterial, etc.) can be found in Chen and Kuo (2005).

1.3.5 An Inhomogeneous Circle in a Full-Plane


This problem in the (x,y)-plane is better solved using the complex variable method.
In terms of the complex variable z = x + iy, Eq. (1.29) for the source at zs = xs + iys is
expressed as

k∇ 2G = δ (z − zs ) (1.56)

The potential Green’s function G, its flux qi, and the resultant flux Qn along a smooth
curve AB can be expressed by an analytical complex function F(z) as

G = Re[F (z)]
qx − iqy = kF ′(z) (1.57)
B
Qn ≡ ∫ qn ds = k Im[F (z)] B
A
A

Thus, in the full (x,y)-plane, the potential Green’s function will be simply the real
part of the following complex function F
14 Introduction

1
F (z; zs ) = ln(z − zs ) (1.58)
2π k

For the inhomogeneity/matrix bimaterial system, we assume that the material


properties are k(i) and k(m), respectively, for the inhomogeneity and matrix, and we
use superscript “(i)” for the inhomogeneity and “(m)” for the matrix.

1.3.5.1 A Source in the Matrix


We first assume that there is a source of unit magnitude at z = zs within the matrix
outside the circular inhomogeneity (i.e., |zs| > a). In other words,

k ( i ) ∇2G( i ) = 0 ( z < a)
(1.59)
k ( m) ∇2G( m) = δ (z − zs ) ( z , zs > a)

Also, the interface continuity conditions (on the circle |z| = a) are:

G(i ) = G( m) , qr(i ) = qr( m) (1.60)

The Green’s function solution (the real part of F(z)) in terms of the complex function
F(z) can be expressed as (making use of Eq. (1.58) and considering the solution form
similar to Eq. (1.36)) (Smith 1968)

 b ln(z − zs )

( z < a)
F (z) =  1  a2 
 2 πk ( m) ln(z − zs ) + c ln  z − zs  ( z > a) (1.61)

where the second term in the second expression of F(z) in the matrix is associated
with the image of the circle |z| = a (referring to Eq. (1.10)). The two complex
constants b and c are determined by the conditions that Re[F(z)] and kIm[F(z)] are
continuous along the interface between the inhomogeneity and matrix. This gives the
final expression of the Green’s function (the real part of the following expression) in
the inhomogeneity/matrix system, when the source is in the matrix, as

 1
 π(k (i ) + k ( m) ) ln(z − zs ) ( z < a)
F (z) =  (1.62)
1 ( k ( m) − k ( i ) )  a2 

2 π k ( m)
ln( z − zs ) +
2 π k ( m) (k (i ) + k ( m) )
ln  z − zs  ( z > a)


In determining the constants b and c, the following properties of a complex function


F(z) have been used:

Re[F (z)] = Re[F (z )]


(1.63)
Im[F (z)] = − Im[F (z )]

Remark 1.24: When the inhomogeneity and matrix are made of the same material
(i.e., k(i) = k(m) = k), Eq. (1.62) is reduced to the Green’s function in the full-plane
1.3 Green’s Functions of Potential Problems 15

with material property k. In this case, the second term in the second expression of
Eq. (1.62) is identically zero.
Remark 1.25: When k(i) = 0 (qn = 0), we have the flux-free boundary condition at
|z| = a for the matrix, and the corresponding solution with material property k(m) = k
in the matrix domain |z| > a is

1 1  a2 
F (z) = ln(z − zs ) + ln  − zs  (1.64)
2π k 2π k  z 

Remark 1.26: When G = 0 on the boundary of the matrix |z| = a (i.e., k(i) approaches
infinity), then the corresponding solution within the matrix is

1 1  a2 
F (z) = ln(z − zs ) − ln  − zs  (1.65)
2π k 2π k  z 

We point out that, this Green’s function could be inappropriate for certain
applications. A new form can be immediately constructed as

1 1  a2  1 a
F (z) = ln(z − zs ) − ln  − zs  + ln   (1.66)
2π k 2π k  z  π 2 k z

which is obtained by adding to Eq. (1.65) a term [1 / (2π k )]ln(a / z). It is clear that the
real part of this new expression (1.66) identically satisfies the potential equation for
z > a and simultaneously vanishes on z = a.
Remark 1.27: Solution (1.64) is the potential Green’s function in the matrix satisfying
qn = 0 on the circular boundary. On the other hand, solution (1.66) corresponds to the
Green’s function in the matrix satisfying G = 0 on the circular boundary.
Remark 1.28: We assume a potential problem for u outside of |z| = a, on which it
satisfies u = f(θ). Now applying the Green’s theorem (1.20) with one of the functions
being u and the other G, we therefore have the expression similar to Eq. (1.14)

∂G(r ; rs , θ s )
u(rs , θ s ) = ∫ f (θ) dC ( r ) (1.67)
C ∂n

except for the fact that the normal derivative of G, ∂G/∂n, equals −qr. Making use of
Eq. (1.66), letting k = 1 and taking the real part of the expression (also with r = a), we
derive the following 2D Poisson formula outside the circle r = a:

rs2 − a 2 2π f (θ )
u(rs , θ s ) =
2π ∫0 a2 + rs2 − 2ars cos(θ − θ s )
dθ (1.68)
16 Introduction

The preceding Poisson integral formula is the same as the known result in literature
(Greenberg 1971; Chen and Wu 2006). Note that it cannot be derived based on the
Green’s function in Eq. (1.65) due to its inappropriateness (Greenberg 1971).

1.3.5.2 A Source in the Circular Inhomogeneity


We now assume that there is a source of unit magnitude at z = zs within the
inhomogeneity circle (i.e., |zs| < a). Then, the Green’s function solution satisfies

k (i ) ∇2G( i ) = δ (z − zs ) ( z , zs < a)
(1.69)
k ( m ) ∇ 2 G( m ) = 0 ( z > a)

The interface continuity conditions (on the circle |z| = a) are the same as Eq. (1.60).
For this case, the Green’s function solution can be simply obtained by applying
the conformal mapping ζ = a2/z, which maps the inside (outside) region of r = a in
the z-plane onto the outside (inside) region in the ζ-plane. Therefore, from Eq. (1.62)
(switching also k(i) and k(m)), we have

 1
 π(k (i ) + k ( m) ) ln(ζ − ζ s ) ( ζ < a)

F (ζ) =  (1.70)
 1 ln(ζ − ζ s ) + ( k ( i ) − k ( m) )  a2 
 2πk ( i ) 2 π k (k + k )  ζ
( i ) ( i ) ( m )
ln  − ζs 

( ζ > a)

Note that ζ s = a 2 / zs is the image point of the source point zs . When expressed in
terms of z, it reads as

 1 (k (i ) − k ( m) )  a2 a

2 π k (i )
ln( zs − z ) +
2 π k ( i ) (k ( i ) + k ( m ) )  ln( z −
z
) − ln 
z
( z < a)
F (z) =   s s (1.71)
1 (k (i ) − k ( m) ) z

 π(k (i ) + k ( m) )
ln(zs − z) + ln
2 π k ( m ) (k ( i ) + k ( m ) ) a
( z > a)

where the singularity (e.g., the term ln(a2/zzs)) induced by the conformal mapping
within the circle has been removed (Smith 1968). Compared with Smith (1968), some
constant terms have been carefully adjusted to meet the continuity conditions at the
interface z = a.
Remark 1.29: When the inhomogeneity and matrix are made of the same material
(i.e., k(i) = k(m) = k), Eq. (1.71) is reduced to the Green’s function in the full-plane with
material property k. In this case, both the second terms in the two expressions in Eq.
(1.71) are identically zero.
Remark 1.30: When k(m) = 0 and letting k(i) = k, we obtain from the first expression
in Eq. (1.71)

1 1  a2  1 a
F (z) = ln(zs − z) + ln  z − − ln (1.72)
2π k 2π k  zs  2π k zs

However, the second expression of Eq. (1.71), when multiplied by k(m), doesn’t
vanish, and hence the normal flux qn = k ∂G / ∂r calculated from the Green’s function
1.3 Green’s Functions of Potential Problems 17

given by Eq. (1.72) is not zero on the boundary |z| = a. Actually, we have a constant
normal flux qn = 1 / (2π a) over the circular boundary. This situation resembles a
homogeneous circular plate under an internal concentrated force with traction-free
on its surface in Chapter 3 (see Remark 3.30), a problem that is not well posed unless
the concentrated source is in self-balance.
Remark 1.31: When G = 0 on the boundary of the circular inhomogeneity
|z| = a (i.e., k(m) approaches infinity), then the corresponding solution within the
inhomogeneity with material property k is

1 1  a2  1 a
F (z) = ln(zs − z) − ln  z −  + ln (1.73)
2π k 2π k  zs  2π k zs

Remark 1.32: Solution (1.72) is the potential Green’s function within the circle
satisfying qn = 1 / (2π a) on the boundary of the circle. Solution (1.73) is the
corresponding Green’s functions within the circle satisfying G = 0 on the boundary
of the circle. Letting k = 1, and taking the real part of Eq. (1.73), we find that

1 1 a2 1 a
G(r, θ; rs , θ s ) = Re[F (z)] = ln zs − z − ln z − + ln
2π 2π zs 2 π zs
1
= ln r 2 + rs2 − 2rrs cos(θ − θ s ) (1.74)

1
− ln r 2 rs2 / a 2 + a 2 − 2rrs cos(θ − θ s )

which is identical to the one given in Eq. (1.13). The associated Poisson integral
formula has already been given in Eq. (1.15).

1.3.6 An Inhomogeneous Sphere in a Full-Space


The 3D potential problem has been an active research topic in electrostatics. The
method of image was first introduced by Lord Kelvin (Thomson 1845) to solve the
problem of an ideally conducting sphere in front of a static point charge. Neumann
(1883) presented a mathematical formulation of image line charges for a static point
charge inside or outside a dielectric sphere by effectively extending the Kelvin
method. In Neumann’s solution for a dielectric sphere, the image line charge was
introduced (Cai et al. 2007). Due to the importance of the problem, Neumann’s
solution was continuously studied as in Lindell (1992, 1993) and Norris (1995) among
others. In deriving the potential Green’s function solutions for our inhomogeneous
sphere system (we call it sphere/matrix system), the approach in Cai et al. (2007) is
followed.

1.3.6.1 A Source in the Sphere


We first assume that there is a source of unit magnitude at r = rs within the spherical
inhomogeneity (i.e., rs = | rs| < a). Then the Green’s functions must satisfy
18 Introduction

z
(a)
ri R*
f

r
rs θ
x

k(i)
a
k(m)

z
(b)
rs R
f

R*

r
ri θ
x

k(i)
a

k(m)

Figure 1.2. (a) A point source inside the sphere of the sphere/matrix system (rs < a); (b) A
point source inside the matrix of the sphere/matrix system (rs > a), illustrated and viewed in
the (x,z)-plane.

k (i ) ∇2G( i ) = δ(r − rs ) (r, rs < a)


(1.75)
k ( m) ∇2G( m) = 0 ( r > a)

The interface continuity conditions (on the surface of the sphere r = a) are:

G(i ) = G( m) , qr(i ) = qr( m) (1.76)

For the boundary-value problem described by Eqs. (1.75) and (1.76) and as also
illustrated in Figure 1.2a, where there is a unit point source at r = rs inside the sphere
(rs < a) with property k(i), the Green’s function solutions can be expressed as
1.3 Green’s Functions of Potential Problems 19

 ∞
1
 − + ∑ Anr n Pn (cos θ)
 4 πk ( i ) R n = 0
(rs , r < a)
G(r; rs ) =  ∞ (1.77)
 ∑ Bn Pn (cos θ) (r > a)
 n= 0 r n+ 1

where R = |r − rs| is the distance between the field point and source point (Figure 1.2a),
Pn(x) is the Legendre polynomial, and An and Bn are coefficients to be determined
by the interface conditions (1.76). In order to find these coefficients, one can expand
the first term in the first expression of Eq. (1.77) as (Chen and Wu 2006; Cai et al.
2007; see also Appendix A of this chapter)

 1 ∞
 r
n

1
− ∑
 4 πk (i )rs n= 0  rs 
Pn (cos θ) (0 ≤ r ≤ rs )
− = ∞ n (1.78)
4 πk ( i ) R  1  rs 
 4 πk (i) r ∑  r  n
− P (cos θ) (rs ≤ r ≤ a)
 n= 0

Then, making use of these series expansions and reinforcing the interface conditions
(1.76), we find the expansion coefficients in Eq. (1.77) as (for n = 0, 1, 2, …)

1 rsn  1+ β 
An = − +
β 1+ 
4π k a
( i ) 2 n 1
 1 − β + 2n 
(1.79)
rn 1 + β  2β 
Bn = − s (i ) 2 + 
4π k 2  1 − β + 2n 

where β is defined as

k ( i ) − k ( m)
β= (1.80)
k ( i ) + k ( m)

It is noted that β takes values from –1 to 1. While β = 1 (k(m) = 0) reduces the solution
of the sphere/matrix system to the one within the sphere of r = a under boundary
condition qr = 0, the case β = –1 (k(m) = ∞) to the boundary condition G = 0 on the
surface of the sphere.
Thus, the 3D Green’s function solutions for the sphere/matrix system with the
source in the sphere can be finally expressed as:

 ∞
1 1 r n rsn  1+ β 
− ( i )

 4 πk R 4 πk n = 0 a ( i ) ∑ 2 n + 1
β1+
 1 − β + 2 n 
Pn (cos θ) (rs , r < a)
G(r ; rs ) =  ∞
(1.81)
 − 1 rsn 1 + β  2β 
 4 πk n= 0 r
( i ) ∑ n + 1 2  2+
1 − β + 2n 
Pn (cos θ) (r > a)
20 Introduction

Equation (1.81) can be first simplified by introducing the image ri of the source point
rs (Figure 1.2a), which gives (with R* = |r – ri|)

 ∞
1 β ri 1 r n rsn β(1 + β)
 − ( )
− ( i ) *

 4 πk R 4 πk R a 4 πk n = 0 a 2 n + 1 1 − β + 2 n
i ( i ) ∑ Pn (cos θ) (rs , r < a )
G(r ; rs ) =  ∞
 − (1 + β) − 1 rsn 1 + β 2β
 ( i ) ( i ) ∑
4 πk R 4 πk n = 0 r n + 1 2 1 − β + 2 n
Pn (cos θ) (r > a )

 (1.82)
Then, the infinite series terms involved in Eq. (1.82) can be expressed in terms of line
integrals over the image line source (Appendix A). Therefore, the final expression
for the Green’s functions of the sphere/matrix system when the source is within the
sphere is

 1  1 βa β(1 + β) ∞ 1  z 
− ( 1− β ) / 2 
−  + + ∫   dz (rs , r < a)
 4πk  r − rs rs r − ri
( i ) 2a ri r − z  ri 

G(r ; rs ) =  −(1+β)/ 2
 
1  1 + β β(1 + β) rs 1  z  

 − 4 πk (i )  r − r + 2r ∫0 r − z  r  dz (r > a)
  s s s 
 (1.83)
We point out again that ri denotes the image point of rs, and it lies on the radial line
extending from the origin through the source point rs, with its distance from the
origin equal to a2/rs (Figure 1.2a).
Remark 1.33: When the inhomogeneity and the matrix are made of the same material
(i.e., k(i) = k(m) = k and β = 0), Eq. (1.83) is reduced to the Green’s function in the full
space with material property k.
Remark 1.34: When k(m) = 0 (i.e., β = 1 or qr = 0), we have the flux-free boundary
condition at r = a for a homogeneous sphere of radius r = a. In this case, the solution
can be obtained by substituting k(i) = k and β = 1 in the first expression of Eq. (1.83).
It is noted that even for this single sphere case, the solution still contains the image
line integration. This is associated with the ill-posedness of the problem similar to
the 2D potential case discussed after Eq. (1.72) and the 2D antiplane and plane-
strain deformation within a circular plate presented in Chapter 3.
Remark 1.35: When G = 0 on the boundary of the homogeneous sphere r = a
(i.e., k(m) approaches infinity or β = –1), the corresponding solution within the
inhomogeneity is

−1 1 ri
G(r ; rs ) = + (rs ,r < a) (1.84)
4πk r − rs 4πk a r − ri

which is the same as the one given in Eq. (1.17) when k = 1. As discussed previously,
the solution given in Eq. (1.84) can be applied to represent the field within a sphere
due to the given surface potential by applying the Green’s theorem (1.20). In doing
so, one arrives at the Poisson formula for the boundary value problem associated
with the sphere as given in Eq. (1.18).
1.3 Green’s Functions of Potential Problems 21

1.3.6.2 A Source in the Matrix


We now assume that there is a source of unit magnitude at r = rs in the matrix (i.e.,
rs > a), governed by

k ( i ) ∇ 2 G( i ) = 0 ( r < a)
(1.85)
k ( m) ∇ 2G( m) = δ (r − rs ) (r, rs > a)

and subjected to the same interface continuity conditions as given by Eq. (1.76).
Similar to the case where the source is inside the sphere, but as illustrated in
Figure 1.2b, the Green’s function solutions can be assumed as

 ∞
 ∑ Cn r n Pn (cos θ) (r < a)

G(r ; rs ) =  n = 0 ∞ (1.86)
1 Dn
 4πk (m) R n∑
− + P (cos θ) (rs , r > a)
r n+1 n
=0

where Cn and Dn are the coefficients to be determined. By expanding the first term
in the second expression of Eq. (1.86) in terms of the Legendre polynomials and
making use of the interface conditions (1.76), we find that

1 1− β  2β 
Cn = − + 2 + 
4π k ( m ) n
2rs  1 1 − β + 2n 
(1.87)
1 a 2 n+1  1− β 
Dn = β 1− 
4π k ( m ) rsn +1  1 − β + 2 n 

Therefore, the solutions for the sphere/matrix system with the source in the matrix
can be expressed as:

 ∞
1 r n (1 − β)  2β 
 −
 4 πk ( m ) ∑ n + 1 2   2+
1 − β + 2n 
Pn (cos θ) (r < a)
G(r ; rs ) =  n = 0 rs

1 1 a 2 n+ 1  1−β 
− + ∑
 4 πk ( m) R 4 πk ( m) n= 0 rsn+ 1r n+ 1 
β 1−
1 − β + 2n 
Pn (cos θ) (rs , r > a)
 (1.88)

Making use of the image solution for the first infinite series and replacing the second
infinite series by the integral over the image line source, we finally have

 −1  1 − β β(1 − β) ∞ 1  z 
−(1−β)/ 2 

 4 πk ( m ) 
r − rs
+
2rs ∫ 
rs r − z  r  dz (r < a )
G( r; rs ) =   s 
 −(1+ β)/2 
 1  1 βa β(1 − β) ri 1  z 
 − 4 πk ( m)  r − r − r r − r + 2a ∫0 r − z  r  dz (rs , r > a)
  s s i i 
 (1.89)

Remark 1.36: When the sphere and matrix are made of the same material (i.e., k(i) =
k(m) = k, i.e., β = 0), Eq. (1.89) is reduced to the Green’s function in the full space with
material property k.
22 Introduction

Remark 1.37: When k(i) = 0 (i.e., β = –1, or qr = 0), we have the flux-free boundary
condition at r = a. In this case, the solution within the matrix (r > a) with material
property k(m) = k is still expressed by the first expression of Eq. (1.89), but with k(m) =
k and β = –1. Different from the corresponding 2D case, there is no exact closed-form
solution for this 3D problem.
Remark 1.38: When G = 0 on the boundary r = a (i.e., k(i) approaches infinity or
β = 1), the corresponding solution in the matrix is

−1 1 ri
G(r ; rs ) = + (rs , r > a) (1.90)
4πk r − rs 4πk a r − ri

Remark 1.39: Solution (1.90) is the potential Green’s function in the matrix with an
opening sphere under the boundary condition G = 0 on r = a. Similarly, by applying
the Green’s theorem with one of the functions being this Green’s function and the
other being the real problem Eq. (1.8) but for the corresponding exterior problem in
r > a, we find the classic 3D Poisson integral formula (using the spherical coordinates
(rs,θs,ϕs) for the source point rs)

−a(a 2 − rs2 ) π 2π f (θ, ϕ)


u(rs , θ s , ϕ s ) =
4π ∫ 0
sin θ dθ ∫
0 (a 2 + r 2 − 2ar cos γ )3 / 2
s s
dϕ (1.91)

which is the same as in Chen and Wu (2006).

1.4 Applications of Green’s Theorems and Identities

1.4.1 Integral Equations for Potential Problems


We assume the following two different systems, one is the real problem associated
with u (in V and bounded by S, with also suitable boundary conditions on S) and the
other the Green’s function denoted by G. The differential equations of them are

∇ 2 u = f (r ) (r ∈V )
(1.92)
∇2G = δ(r − rs ) (r , rs ∈ E 3 )

Substituting these two functions into Green’s theorem (1.20) gives us

 ∂G(r ; rs ) ∂u(r ) 
u(rs ) = ∫ G(r ; rs ) f (r )dV (r ) + ∫ u(r ) − G(r ; rs ) dS(r ) (1.93)
V S ∂n ∂n 

This is the well-known representative theorem derived by Green (1828) for a given
general potential problem. It is suited for both 2D and 3D spaces.
For the special case where f = 0 and also in view of the 3D Green’s function in
the full space, Eq. (1.93) can be reduced to

1  1 1 
u(rs ) = ∫
4π 
S
u(r )∇ n
r − rs

r − rs
∇n u(r ) ⋅ dS (r )

(1.94)
1.4 Applications of Green’s Theorems and Identities 23

which was also presented by Green (1828), with ∇n being the derivative along the
outward normal direction of the boundary S. This represents the potential in a given
domain in terms of boundary integrals where the corresponding Green’s functions
are the integral kernels.

1.4.2 Boundary Integral Equations for Potential Problems


Let us discuss Eq. (1.93) in more detail under the condition that f = 0, but with given
boundary values on the surface of the problem domain. The real boundary value
problem for both 2D and 3D is given as

∇2 u = 0 (r ∈V )
u=u (r ∈ Su )
qn = q (r ∈ Sq ) (1.95a, b, c, d)
∂u
qn ≡ (S = Su + Sq )
∂n

Then, substituting Eq. (1.95) into Eq. (1.20) gives

u(rs ) = ∫ u (r ) ∂G (r ; rs ) / ∂n − qn (r )G(r ; rs ) dS(r )


Su
(1.96)
+ ∫ u(r )∂G(r ; rs ) / ∂n − q(r )G(r ; rs ) dS(r )
Sq

Equation (1.96) is the general representation of the potential u at any inner point
in terms of its boundary values (u or its normal derivative). The integral kernels are
the associated Green’s function G and its normal derivative ∂G/∂n. It is obvious that
the problem is still unsolvable because we have unknown boundary values on the
right-hand side (qn on Su and u on Sq). However, if one can construct certain special
Green’s functions that can make these unknown terms disappear, one then has the
true representation of the inner potential in terms of its given boundary value (either
u or its derivative). We discuss the following two examples:
Example 3: u is given on the entire boundary S (i.e., Su = S, and Sq = 0). If we can
construct the Green’s function G as such that G = 0 on S, then Eq. (1.96) becomes

u(rs ) = ∫ u (r )∂G(r ; rs ) / ∂n dS(r ) (1.97)


S

For the special circular and spherical domain cases, this expression is reduced to the
celebrated Poisson integral formula for 2D and 3D. Similar integral expression can
be also found for the half-plane and half-space cases where u is given on the surface
and G = 0 on the surface of the half-plane or half-space, as already presented in
Section 1.3.
The opposite case is that qn is given on the entire boundary S (i.e., Sq = S, and
Su = 0). If we can construct the Green’s function G as such that ∂G/∂n = 0 on S, then
Eq. (1.96) becomes

u(rs ) = − ∫ q(r )G(r ; rs )dS(r ) (1.98)


Sq
24 Introduction

Making use of the 2D/3D Green’s functions satisfying this type of boundary
condition for circular and spherical domains and for half-plane and half-space
cases, the potential at any location of the domain induced by the surface flux can be
expressed as a simple integral over the boundary. The process is similar to the one
when deriving the Poisson integral formulae for 2D and 3D domains under surface
potential conditions. It is noticed, however, that the potential inside the domain can
be different by an arbitrary constant.
These examples are for the special cases where only one type of field quantities
(either u or qn) is given on the entire boundary. However, in general, part of the
boundary would be with given u and the remaining part with qn. For this case, one
still faces the challenge of finding the remaining unknown boundary values in
Eq. (1.96) in order for this equation to be applied to represent the inner potential.
To solve this problem, we let the internal point r approach the point on the
boundary (assumed to be smooth). Then, Eq. (1.96) becomes the well-known
boundary integral equation (r ∈S)

1
u(rs ) = ∫ u (r )∂G(r ; rs ) / ∂n − qn (r )G(r ; rs ) dS
2 Su
(1.99)
+ ∫ u(r )∂G(r ; rs ) / ∂n − q(r )G(r ; rs ) dS
Sq

It is noted that if rs is a nonsmooth point on the boundary, then the ½ factor needs
to be replaced by a different value, which depends on the local geometry at rs (i.e.,
Brebbia and Dominguez 1996). Now, Eq. (1.99) can be solved for the unknown
boundary values. After that, Eq. (1.96) can be used to find the internal value of u. The
corresponding internal flux (qi) can be further obtained by taking the derivative of
Eq. (1.96) with respect to the source coordinates rs (xis ). In other words, the following
integral equation can be applied to find the derivative of the field quantity at any
internal point of the domain.

∂u(rs )  ∂ 2G(r ; rs ) ∂G(r ; rs ) 


s
= ∫ u (r ) s
− qn (r )  dS
∂xi Su
 ∂n ∂xi ∂xis 
(1.100)
 ∂ 2G(r ; rs ) ∂G(r ; rs ) 
+ ∫ u(r ) s
− q(r )  dS
Sq
 ∂n ∂xi ∂xis 

Remark 1.40: If the potential or the flux is given on the entire boundary of the
domain (i.e., Su = S or Sq = S), then the derivative of the potential inside will be
completely determined by either the potential or flux given on the boundary. These
special cases resemble the original Green’s representative theorem.

1.5 Summary and Mathematical Keys

1.5.1 Summary
In this chapter, we have defined the Green’s function and presented the Green’s
function solutions for the common infinite and bimaterial plane/space domains.
1.6 Appendix A: Equivalence between Series Summation and Integral 25

These solutions correspond to the potential problems. Green’s identities (theorems)


are also introduced along with the Green’s representative theorem. Making use of
the Green’s representative theorem, the classic Poisson integral formulae are also
presented for both 2D and 3D domains with flat or circular/spherical interface/
surface. It is noted that the very popular boundary integral equation can be also
directly derived using the Green’s representation.

1.5.2 Mathematical Keys


The mathematical key in this chapter is the definition of the Green’s function.
Because the Green’s function is the solution corresponding to a concentrated line/
point source, its expression may not be unique. For instance, the Green’s functions
for the 2D/3D potential problems can be different by any harmonic function. Besides
this key, the Green’s identities and representative theorems are all important keys. It
should be noticed that while there are similarities between the 2D and 3D Green’s
functions, they could be significantly different in certain situations. For instance, while
we have exact closed-form 2D Green’s functions for a circular domain or a matrix
with a circular opening under zero flux, the corresponding 3D Green’s functions
contain a line integral over the image line of the source.

1.6 Appendix A: Equivalence between Infinite Series Summation


and Integral over the Image Line Source
We first list the following two useful expressions.
The first one is on expanding the point-source solution (1/R) in terms of the
infinite series of the Legendre polynomials as (also referring to Figure 1.2a)

1 1
=
4 πkR 4 πk r + rs − 2rrs cos θ
2 2

 1 ∞ rn
 ∑ Pn (cos θ)
 4 πk n = 0 rsn + 1
(r < rs ) (A1)
= ∞
 1 rsn
∑ P (cos θ)
 4 πk n = 0 r n + 1 n
(r > rs )

The second one is associated with the simple integral. It can be easily shown that,
for –1 ≤ β ≤ 1, we have

r 2
∫0 z(1− β )/ 2 + n−1dz = 1 − β + 2n r(1− β )/ 2+ n
(A2a, b)
+∞ 1 2 1
∫r z(1− β )/ 2 + n +1
dz =
1 − β + 2 n r (1− β )/ 2 + n

We now discuss each infinite series term in the Green’s function expressions in
Eqs. (1.82) and (1.88), which cannot be simply expressed by a single image source
contribution.
First, we consider the case where the source is within the sphere.
26 Introduction

For the infinite series in the first expression of Eq. (1.82), that is, for the case
r < a:


1 r n r n β (1 + β )
S(ii ) = −
4π k (i )
∑ a2 n+s 1 1 − β + 2n Pn (cosθ ) (A3)
n=0

which can be expressed as

1 β(1 + β) ∞ 1 2 rn
S(ii ) = −
4 πk
∑ P (cos θ)
n n
n = 0 a 1 − β + 2 n ri
( i ) 2
1 β(1 + β) +∞  ∞ 1 (1−β)/ 2 + n 1 rn 
=−
4 πk ( i ) 2 ∫  ∑ ri ( 1− β )/ 2 + n + 1 n
Pn (cos θ) dz (A4)
ri
 n = 0 a z ri 
(1−β )/ 2
1 β(1 + β) +∞  ∞ ri rn 
=−
4 πk ( i ) 2 a ∫  ∑ (1−β )/ 2 zn + 1
Pn (cos θ) dz
 n=0 z
ri  

or

+∞  −q( ii ) (z) ∞ rn 
S(ii) = ∫
ri

 4π k(i)
∑ zn+1 Pn (cosθ )dz (A5)
n=0 

It is noted that physically, this represents a distributed image charge over ri ≤ z < +∞
with its density per unit length being defined as
(1− β )/ 2
β (1 + β ) ri
q(ii ) (z) = (A6)
2a z(1− β )/ 2

For the infinite series in the second expression of Eq. (1.82), that is, for the case
r > a, we have

β (1 + β ) ∞ rsn 1
S(im) = −
4π k ( i ) ∑ n + 1 1 − β + 2n
Pn (cosθ )
n =0 r
1 β (1 + β ) ∞ rs z−(1+ β )/ 2 zn
=− ∑
4π k (i ) 2rs n =0 ∫0 rs−(1+ β )/ 2 r n +1
Pn (cosθ )dz (A7)

rs −q(im) (z) ∞ zn
=∫
0 ∑
4π k (i ) n = 0 r n + 1
Pn (cosθ )dz

This corresponds to a distributed image charge over 0 ≤ z ≤ ri with its density per
unit length being

β (1 + β ) z−(1+ β )/ 2
q(im) (z) = (A8)
2rs rs−(1+ β )/ 2

We now look at the case when the source is located in the matrix.
For the second infinite series in the first expression of Eq. (1.88), that is, for
r < a, we have
1.7 References 27

1 β (1 − β ) ∞ r n 2
S( mi ) = −
4π k ( m ) 2
∑ n + 1 1 − β + 2n
Pn (cosθ )
n = 0 rs
1 β (1 − β ) +∞ ∞ rs(1− β )/ 2 r n
=−
4π k ( m)

2rs ∫rs n =0 z(1− β )/ 2 zn +1
Pn (cosθ )dz (A9)

+∞ −q( mi ) (z) ∞ rn
=∫
rs 4π k( m)
∑ zn+1 Pn (cosθ )dz
n =0

which represents a distributed image charge over rs< z < +∞ with its density per unit
length being

β (1 − β ) rs(1− β )/ 2
q( mi ) (z) = (A10)
2rs z(1− β )/ 2

For the second infinite series in the second expression of Eq. (1.88), that is, for r > a,
we have

1 β (1 − β ) ∞ a 2 n +1 2
S( mm) = −
4π k ( m) 2
∑ r n+1r n+1 1 − β + 2n Pn (cosθ )
n =0 s
1 β (1 − β ) ri ∞ z−(1+ β )/ 2 zn
=−
4π k (m)

2a ∫0 n =0 ri−(1+ β )/ 2 r n + 1
Pn (cosθ )dz (A11)

ri −q( mm) (z) ∞ zn


=∫
0 ∑
4π k ( m) n= 0 r n+ 1
Pn (cosθ )dz

which corresponds to a distributed image charge over 0 ≤ z ≤ ri with its density per
unit length being defined as

β (1 − β ) z−(1+ β )/ 2
q( mm) (z) = (A12)
2a ri−(1+ β )/ 2

1.7 References
Brebbia, C. A. and Dominguez, J. 1996. Boundary Elements: An Introductory Course. 2nd ed.
Southampton, UK: WIT Press, Computational Mechanics Publications.
Cai, W., Deng, S. Z., and Jacobs, D. 2007. Extending the fast multipole method to charges
inside or outside a dielectric sphere. Journal of Computational Physics 223: 846−64.
Cannell, D. M. 2001. George Green: Mathematician & Physicist, 1793–1841: The Background
to His Life and Work. Philadelphia: Society for Industrial and Applied Mathematics.
Challis, L. and Sheard, F. 2003. The Green of Green functions. Physics Today,
December, 41–6.
Chen, J. T. and Wu, C. S. 2006. Alternative derivations for the Poisson integral formula.
International Journal of Mathematical Education in Science and Technology 37: 165–85.
Chen, T. Y. and Kuo, H. Y. 2005. On linking n-dimensional anisotropic and isotropic Green’s
functions for infinite space, half-space, bimaterial, and multilayer for conduction.
International Journal of Solids and Structures 42: 4099–114.
Green, G. 1828. An Essay on the Application of Mathematical Analysis to the Theories of
Electricity and Magnetism. Printed for the author by T. Wheelhouse, Nottingham, UK.
28 Introduction

Greenberg, M. D. 1971. Application of Green’s Functions in Science and Engineering.


Englewood Cliffs, NJ: Prentice-Hall.
Lindell, I. V. 1992. Electrostatic image theory for the dielectric sphere. Radio Science
27: 1−8.
1993. Image theory for electrostatic and magnetostatic problems involving a material
sphere. American Journal of Physics 61: 39−44.
Neumann, C. 1883. Hydrodynamische Untersuchen nebst einem Anhang uber die Probleme
der Elecktrostatik und der magnetischen Induktion. Leipzig: Teubner, 279−82.
Norris, W. T. 1995. Charge images in a dielectric sphere. IEE Proceedings of Science,
Measurement and Technology 142: 142−50.
Roach, G. F. 1982. Green’s Functions. 2nd ed. New York: Cambridge University Press.
Smith, E. 1968. The interaction between dislocations and inhomogeneities I. International
Journal of Engineering Science 6: 129–43.
Thomson, W. (Lord Kelvin). 1845. Extrait D’une lettre de M. William Thomson a M. Liouville.
Journal de Mathématiques Pures et Appliquées 10: 364.
Timoshenko, S. P. 1953. History of Strength of Materials, with a Brief Account of the History of
Theory of Elasticity and Theory of Structures. New York: Dover Publications.
2 Governing Equations

2.0 Introduction
Material anisotropy and multiphase coupling are two outstanding features among
novel and smart materials and composites. The interesting coupling phenomenon in
piezoelectric materials has been successfully applied to various industries, with most
notable devices based on these materials being sensors and actuators (Yang 2006).
Piezoelectric coupled semiconductors have recently received intensive attentions
due to their potential impact on the optoelectronic properties of semiconductor
devices (Wood and Jena 2008). Piezomagnetic materials came late to the stage but
have also found applications in various devices including those for energy harvesting
(Yuan and Wang 2008; Sharapov 2011). The most striking development in recent
years is the special magnetoelectroelastic (MEE) composites made of piezoelectric
and piezomagnetic phases in the novel field of multiferroic materials (Nan et al.
2008). Due to the product property in MEE composites, a voltage applied to the
piezoelectric phase can generate a magnetic potential in the piezomagnetic phase or
vice versa, the most welcomed coupling between the electric field (E-field, for brevity)
and magnetic field (H-field, for brevity), which is also called the magnetoelectric
(ME) effect (Nan et al. 2008). Although the MEE materials/composites have been
studied in the solid mechanics community for a relatively long time (e.g., Pan 2001),
they only receive considerable attention in recent years due to their close connection
to the multiferroic materials/composites.

2.1 General Anisotropic Magnetoelectroelastic Solids


Consider a linear, anisotropic MEE solid occupying the domain V bounded by
its boundary S. In discussing the Green’s functions, the problem domain and the
corresponding boundary conditions will be clearly described later. We also assume
that the deformation is static, and thus the field equations for such a solid consist of
the following equations (Pan 2001).

29
30 Governing Equations

2.1.1 Equilibrium Equations Including Also Those Associated


with the E- and H-Fields
σ ji, j + fi = 0
Di,i − fe = 0 (2.1)
Bi,i − fh = 0

where σij (= σji), Di, and Bi are, respectively, the stress (N/m2), electric displacement
(C/m2), and magnetic induction (Wb/m2); fi, fe, and fh are, respectively, the body force
(N/m3), and electric (C/m3) and magnetic (Wb/m3) sources. In this and the following
sections, summation from 1 to 3 (1 to 5) over repeated lowercase (uppercase)
subscripts is implied. A subscript comma denotes the partial differentiation. The
three sets of equations in Eq. (2.1) are the equilibrium equations for the elastic,
electric, and magnetic fields, respectively.

2.1.2 Constitutive Relations for the Fully Coupled MEE Solid


σ ij = cijlmγ lm − ekij Ek − qkij Hk
Di = eijk γ jk + ε ij E j + α ij H j (2.2a)
Bi = qijk γ jk + α ji E j + µij H j

where γij is the strain (dimensionless), Ei the electric field (V/m), Hi the magnetic
field (A/m); cijlm, eijk, and qijk are, respectively, the elastic moduli (N/m2), piezoelectric
(C/m2), and piezomagnetic (N/(A×m)) coefficients; εij and μij are, respectively, the
dielectric permittivity (C2/(N×m2)) and magnetic permeability (N×s2/C2) coefficients;
and αij are the ME coefficients (C/(m×A)= Wb/(m×V)).
Remark 2.1: The constitutive relations (2.2a) are often expressed in the following
form in physics and materials communities

γ ij = sijlmσ lm + dkij Ek + qkij H k


Di = dijkσ jk + ε ij E j + α ij H j (2.2b)
Bi = qijkσ jk + α ji E j + µij H j

where sijlm (inverse of cijlm), dijk, and qijk are, respectively, the elastic compliance
coefficients (m2/N), piezoelectric (m/V), and piezomagnetic (m/A) coefficients; εij
and μij are, respectively, the same dielectric permittivity (C2/(N×m2)) and magnetic
permeability (N×s2/C2) coefficients as in Eq. (2.2a); and αij are the same ME
coefficients (C/(m×A)= Wb/(m×V)) as in Eq. (2.2a). It is noticed that though dijk and
qijk are also called piezoelectric and piezomagnetic coefficients (even using the same
symbol qijk), their physical meanings are different from those defined in Eq. (2.2a). In
this book, our discussion is based on the constitutive relations (2.2a).

2.1.3 Gradient Relations (i.e., Elastic Strain-Displacement, Electric


Field-Potential, and Magnetic Field-Potential Relations)

1
γ ij = (ui, j + u j,i )
2 (2.3)
Ei = − φ,i , H i = − ψ ,i
2.1 General Anisotropic Magnetoelectroelastic Solids 31

where ui, φ, and ψ are, respectively, the elastic displacement, electric potential, and
magnetic potential.
Remark 2.2: These governing relations can be derived based on the enthalpy concept
for the MEE materials, as presented in Appendix A.
We now extend the index range from the purely elastic case of 1–3 to 1–5 by
introducing the following definitions for the extended displacement, strain, stress,
traction, and material properties, respectively (Pan 2002),

 ui ( I = i = 1, 2, 3)

uI =  φ ( I = 4) (2.4a)
 ψ ( I = 5)

 γ ij ( I = i = 1, 2, 3)

γ Ij = −E j ( I = 4)  (2.4b)
− H j ( I = 5)

 σij ( J = j = 1, 2, 3)

σiJ =  Di ( J = 4)  (2.4c)
 Bi ( J = 5)

 σij ni ( J = j = 1, 2, 3)

t J = σiJ ni =  Di ni ( J = 4) (2.4d)
 Bi ni ( J = 5)

 cijkl ( J , K = j, k = 1, 2, 3)
 e ( J = j = 1, 2, 3; K = 4)
 lij
 eikl ( J = 4; K = k = 1, 2, 3)
 q ( J = j = 1, 2, 3; K = 5)
ciJKl =  lij (2.4e)
 qikl ( J = 5; K = k = 1, 2, 3)
 − α il ( J = 4, K = 5 or K = 4, J = 5)
−ε ( J , K = 4)
 il
 −µ il ( J , K = 5)

It is noted that we have kept the original symbols instead of introducing new ones
since they can be easily distinguished by the range of their subscript. In terms of this
shorthand notation, the constitutive relations can be unified into a single equation as:

σ iJ = ciJKl γ Kl (2.5)

where the material property coefficients ciJKl can be position-dependent in


domain V.
Similarly, the equilibrium equations in terms of the extended stress can be
recast into

σiJ , i + f J = 0 (2.6)
32 Governing Equations

with the extended force fJ being defined as

 fj ( J = j = 1, 2, 3)

f J =  − fe (J = 4)
(2.7)
 − fh (J = 5)

For the Green’s function solutions, the body force and the electric and magnetic
sources are replaced by the concentrated unit sources, which will be discussed later.
Remark 2.3: The full-size five-dimensional systems of equations can be directly
applied to the various uncoupled cases if we let the coupling material properties eijk
(or dijk), qijk, and αij equal zero. Under this condition, and if the associated body source
and boundary conditions related to E/H-fields are further zero, then the solution will
be purely elastic. If the coupling material properties are zero but the body source
and boundary conditions are related to E/H-fields, then the purely elastic, E-field,
and H-field solutions are all uncoupled.
Remark 2.4: If the coupling material properties qijk and αij are zero, we then solve the
piezoelectric coupled problem.
Remark 2.5: If the coupling material properties eijk (or dijk) and αij are zero, we then
solve the piezomagnetic coupled problem. We present these special cases as follows
for easy reference.
Remark 2.6: The most remarkable feature for the coupled MEE materials or
composites is the intrinsic coupling among the mechanical, electric, and magnetic
fields. In other words, if the three fields are fully coupled, an applied mechanical
force would induce both electric and magnetic fields, and that an applied voltage
would induce a magnetic field or vice versa.

2.2 Special Case: Anisotropic Piezoelectric or Piezomagnetic Solids


For either piezoelectric or piezomagnetic cases, the index range will be 1–4, instead
of 1–5. Thus, the extended notations (2.4), the extended constitutive relations
(2.5), and the extended equilibrium equations (2.6) still hold, but taking the range
from 1–4 only. For these two cases, the extended quantities are defined in the
following text.

2.2.1 Piezoelectric Materials


The piezoelectric materials have the coupling between the mechanical and electric
fields. An applied electric voltage would induce a mechanical deformation, and
a mechanical force would result in an electric field. For this case, the extended
notations (2.4) become:

u ( I = i = 1, 2, 3)
uI =  i (2.8a)
φ ( I = 4)

 γ ij ( I = i = 1, 2, 3)
γ Ij =  (2.8b)
−E j ( I = 4)
2.3╇ Special Case: Anisotropic Elastic Solids 33

σij ( J = j = 1, 2, 3)
σiJ =  (2.8c)
Di ( J = 4)

σij ni ( J = j = 1, 2, 3)
t J = σiJ ni =  (2.8d)
Di ni ( J = 4)

 cijkl ( J , K = j, k = 1, 2, 3)
 e ( J = j = 1, 2, 3; K = 4)
ciJKl =  lij (2.8e)
 eikl ( J = 4; K = k = 1, 2, 3)
 − ε il ( J , K = 4)

The notations (2.8) for piezoelectric materials were first introduced by Barnett and
Lothe (1975), and are sometimes called the Barnett-Lothe notation in literature.
Listed in Table€ 2.1 are the material properties for some of the popular
piezoelectric (semiconductor) materials. It is noted that in this and other tables,
Voigt notation with two indices is used (as explained in Section 2.4), and TI stands
for transversely isotropic material which has a wurtzite crystal structure, while the
cubic material has a zinc-blende crystal structure.

2.2.2╇ Piezomagnetic Materials


The piezomagnetic materials have the coupling between the mechanical and magnetic
fields. An applied magnetic potential would induce a mechanical deformation, and
a mechanical force would result in a magnetic field. The extended notations for this
case can be easily obtained from Eq. (2.8) by replacing the electric quantities with
the magnetic quantities:

u ( I = i = 1, 2, 3)
uI =  i (2.9a)
ψ ( I = 4)

 γ ij ( I = i = 1, 2, 3)
γ Ij =  (2.9b)
− H j ( I = 4)

σij ( J = j = 1, 2, 3)
σiJ =  (2.9c)
 Bi ( J = 4)

σij ni ( J = j = 1, 2, 3)
t J = σiJ ni =  (2.9d)
 Bi ni ( J = 4)

 cijkl ( J , K = j, k = 1, 2, 3)
 q ( J = j = 1, 2, 3; K = 4)
ciJKl =  lij (2.9e)
 qikl ( J = 4; K = k = 1, 2, 3)
 − µ il ( J , K = 4)

2.3╇ Special Case: Anisotropic Elastic Solids


The purely elastic case is well-documented, where the index is taking the range from
1 to 3. Therefore, in Eqs. (2.1)–(2.3), we need only the first relations associated with
34 Governing Equations

Table 2.1. Material Properties for Some Popular Piezoelectric, Including TI Piezoelectric and
Semiconductor Materials (both TI and cubic)

wurtzite c11 c12 c13 c33 c44 e31 e33 e15 ρ ε/ε0

AlN 403 143 104 382 120 −0.66 1.57 −0.39 3.255 9.14
BeO 454 85 77 488 155 −0.02 0.02 — — 7.65//
6.94⊥
GaN 379 140 107 380 98 −0.46 0.77 −0.33 6.095 10.4//
9.5⊥
InN 207 110 95 206 30 −0.60 1.10 −0.44 6.81 15.3
ZnO −0.68 1.31 −0.53 — —
BN 982 134 74 1077 388 0.27 −0.85 — 3.48 5.1//
6.8⊥
SiC 501 111 52 553 163 — 0.20 0.08 3.21 10.03//
9.66⊥
zinc blende e14
GaN 293 159 155 0.72a 6.095 9.8
AlN 304 160 193 1.14a 3.255 9.7
InN 187 125 86 0.83a 6.81 15.3
BN 820 190 480 −0.64 3.45 7.1
SiC 290 235 55 — 3.17 9.72
GaAs 118 54 59 −0.16* 5.318 13.18
InAs 83 45 40 −0.045* 5.667 15.15
AlAs 125 53.4 54.2 −0.23* 3.76 10.06
GaP 140.5 62.0 70.3 −0.10* 4.138 11.11
InP 101.1 56.1 45.6 −0.035* 4.81 12.56
GaSb 88.4 40.3 43.2 −0.013* 5.614 15.69
AlP 133 63 61.5 −0.033** 2.40 9.8
AlSb 87.7 43.4 40.8 −0.084** — —
InSb 68.5 37.4 31.1 −0.159** 5.775 16.8

Notes:
(1) Property units are the same as in Table 2.2 (i.e., cij in 109N/m2, eij in C/m2, εij, in 10−9 C2/Nm2, and ρ in
103kg/m3=g/cm3), with further ε0=8.8542×10−12 C2/Nm2; // parallel to the c-axis and ⊥ perpendicular to
the c-axis.
(2) BN = Boron Nitride, SiC = Silicon Carbide with properties from Levinshtein et al. (2001).
(3) Values of the elastic coefficients are mostly taken from Vurgaftman et al. (2001).
(4) *e14 values are from Adachi (1985, 1994); **e14 values are from Beya-Wakata et al. (2011); ae14 values from
Schulz et al. (2012).
(5) Most static relative dielectric constants and densities are from Singh (2003) and Levinshtein et al.
(2001).
(6) For wurtzite ZnO, eij are taken from Prodhomme et al. (2013); for wurtzite AlN, GaN, and InN, eij are
taken as the average of Prodhomme et al. (2013) and Caro et al. (2013).
(7) Properties of BeO are from Hanada (2009).
(8) For wurtzite structure, cij (GPa) and eij (C/m2) are from Vurgaftman et al. (2001) and Levinshtein et al.
(2001). If the values are different for the same material coefficient from these two references, we then
take their average.
(9) Properties of other important piezoelectric materials are listed:
PZT-4 (lead-zirconate-titanate, Dunn and Taya 1993):
c11=139, c12=77.8, c13=74.3, c33=115, c44=25.6; e15=12.7, e31=−5.2, e33=15.1; ε11=6.4605, ε33=5.61975.
PZT-5J (Kuo and Peng 2013):
c11=82.3, c12=34.1, c13=30.2, c33=59.8, c44=21.3, c66=24.1; e15=14.26, e31=−10.45, e33=16.58;
ε11=14.53, ε33=10.12; μ11=0.0126, μ33=0.0126.
2.4 Special Case: Transversely Isotropic MEE Solids 35

mechanical quantities. Furthermore, the E- and H-field related terms in the first
equation of Eq. (2.2a) should be omitted.

2.4 Special Case: Transversely Isotropic MEE Solids


For this and other special cases followed, the equilibrium equations and gradient
relations are the same and we need only to modify the constitutive relations. We
assume that the transversely isotropic (TI) MEE material has its material symmetric
(or the poling) axis along the z-axis (or x3-axis). We further reduce the indices for
cijkl, eijk, and qijk by following the correspondence between the two and one indices
11⇔1, 22⇔2, 33⇔3, 23⇔4, 13⇔5, 12⇔6 to come to the Voigt notation. Then the
coupled constitutive relations for the TI MEE solid are

σ xx = c11γ xx + c12γ yy + c13γ zz − e31Ez − q31 Hz


σ yy = c12γ xx + c11γ yy + c13γ zz − e31Ez − q31 H z
σ zz = c13γ xx + c13γ yy + c33γ zz − e33 Ez − q33 H z
(2.10a)
σ yz = 2c44γ yz − e15 Ey − q15 H y
σ xz = 2c44γ xz − e15 Ex − q15 H x
σ xy = 2c66γ xy

Dx = 2e15γ xz + ε 11Ex + α 11 H x
Dy = 2e15γ yz + ε 11E y + α 11 H y (2.10b)
Dz = e31 (γ xx + γ yy ) + e33γ zz + ε 33 Ez + α 33 H z

Bx = 2q15γ xz + α 11Ex + µ11 H x


By = 2q15γ yz + α 11Ey + µ11 H y (2.10c)
Bz = q31 (γ xx + γ yy ) + q33γ zz + α 33 Ez + µ33 Hz

where c66=(c11−c12)/2, which always holds for various degenerate materials with
transverse isotropy to be discussed in the following text and will not be repeated
again. It is noticed that for this fully coupled TI MEE solid, there are totally seventeen
independent material coefficients.
Remark 2.7: If the material symmetric axis or poling axis is along the other
coordinate axis, similar constitutive relations can be obtained by simply switching

Table 2.1. Notes (continued)


LiNbO3 (Kuo and Peng 2013):
c11=203, c12=52.9, c13=74.9, c33=243, c44=59.9, c66=74.9, c14=8.99, c56=8.985; e15=3.7, e16=−2.534, e21=−2.538,
e31=0.19, e33=1.31; ε11=0.39, ε33=0.26; μ11=0.05, μ33=0.1.
BaTiO3 (PE) are also given in Table 2.2.
c11=166, c12=77, c13=78, c33=162, c44=43; e15=11.6, e31=−4.4, e33=18.6; ε11=11.2, ε33=12.6; μ11=0.05, μ33=0.1.
Left-hand quartz in a rotated system (Tiersten 1969):
c11=86.74, c12=−8.25, c13=27.15, c14=−3.66, c22=129.77, c23=−7.42, c24=5.70, c33=102.83, c34=9.92, c44=38.61,
c55=68.81, c56=2.53, c66=29.01; e11=0.171, e12=−0.152, e13=−0.0187, e14=0.067, e25=0.108, e26=−0.095,
e35=−0.0761, e36=0.067; ε11=0.03921, ε22=0.03982, ε23=0.00086, ε33=0.04042.
(10) The reader is also referred to Table 1.2.1 in Ding and Chen (2001) for some other piezoelectric
materials.
36 Governing Equations

Table 2.2. Material Properties of MEE Reported in Recent Literature (TI PE/PM with
different volume fraction ratios c= cPE/cPM [pure CoFe2O4 (PM) and pure BaTiO3 (PE) are
also listed as reference])

c 0%(PM) 25%(MEE) 50%(MEE) 75%(MEE) 100%(PE) Ps1 Ps2

c11 286 256 225 196 166 166 mc


c12 173 149 125 101 77 77 mc
c13 170 147 124 101 78 78 mc
c33 269.5 243 216 189 162 162 mc
c44 45.3 44.7 44 43.6 43 43 mc
e31 0 −1.1 −2.2 −3.3 −4.4 −4.4 mc
e33 0 4.65 9.3 14 18.6 18.6 mc
e15 0 2.9 5.8 8.7 11.6 11.6 mc
ε11 0.08 2.86 5.64 8.4 11.2 11.2 mc
ε33 0.093 3.22 6.35 9.47 12.6 12.6 mc
μ11 5.9 4.44 2.97 1.51 0.05 0.05 0.05
μ33 1.57 1.2 0.835 0.46 0.1 0.1 0.1
q31 580 435 290.2 145 0 580 580
q33 700 525 350 175 0 700 700
q15 550 412.5 275 137 0 550 550
α11 0 0 0 0 0 0 0
α33 0 0 0 0 0 0 0
ρ 5.3 5.43 5.55 5.66 5.8 – –

Notes:
(1) Unit: elastic constants, cij, in 109N/m2, piezoelectric constants, eij, in C/m2, piezomagnetic constants,
qij, in N/Am, dielectric constants (permittivity) εij, in 10−9 C2/Nm2, magnetic constants, μij, in 10−4Ns2/
C2, magnetoelectric coefficients αij, in 10−12Ns/VC and ρ in 103kg/m3=g/cm3.
(2) Data in Columns 2–6 are from Xue and Pan (2013).
(3) Data in Columns 7–8 are for the pseudo MEE materials from Pan (2002).
(4) The effective material coefficients for the BaTiO3-CoFe2O4 MEE composite were calculated based
on the micromechanics approach (Kuo and Pan 2011; Kuo and Wang 2012).
(5) Ps1 in Column 7 is for a fully coupled pseudo MEE material with the elastic and piezoelectric
properties being those of the piezoelectric material BaTiO3 and the piezomagnetic coefficients qij
being taken from the magnetostrictive CoFe2O4.
(6) Ps2 in Column 8 is for another pseudo MEE materials with its elastic and piezoelectric properties
being taken from Tiersten (1969) as studied in Pan and Yuan (2000) (where c66 in Pan and Yuan
(2000) should be 29.01×109N/m2, instead of 68.81×109N/m2); the piezomagnetic coefficients qij are
taken from the magnetostrictive CoFe2O4.
(7) mc in Column 8 is for a monoclinic piezoelectric material with the following properties.
c11=86.74, c12=–8.25, c13=27.15, c14=–3.66, c22=129.77, c23=–7.42, c24=5.7, c33=102.83, c34=9.92, c44=38.81,
c55=68.81, c56=2.53, c66=29.01; e11=0.171, e12=–0.152, e13=–0.0187, e14=0.067, e25=0.108, e26=–0.095,
e35=–0.0761, e36=0.067; ε11=0.03921, ε22=0.03982, ε23=0.00086, ε33=0.04042.

the involved indices. Furthermore, to obtain the global material property matrices
for the case in which the local material coordinate system is rotated (with respect
to the global one), we can just apply the coordinate transformation between the
two coordinate systems. These relations are presented in Appendix B.
Listed in Table 2.2 are the material properties of some MEE materials reported
in recent literature.
2.6 Special Case: Transversely Isotropic or Isotropic Elastic Solids 37

2.5 Special Case: Transversely Isotropic


Piezoelectric/Piezomagnetic Solids
For the TI piezoelectric case with the z-axis being the poling direction or axis of
material symmetry, Eq. (2.10) is reduced to

σ xx = c11γ xx + c12γ yy + c13γ zz − e31Ez


σ yy = c12γ xx + c11γ yy + c13γ zz − e31Ez
σ zz = c13γ xx + c13γ yy + c33γ zz − e33 Ez
(2.11a)
σ yz = 2c44γ yz − e15 E y
σ xz = 2c44γ xz − e15 Ex
σ xy = 2c66γ xy

Dx = 2e15γ xz + ε 11Ex
Dy = 2e15γ yz + ε 11E y (2.11b)
Dz = e31 (γ xx + γ yy ) + e33γ zz + ε 33 Ez

Material properties for some of the popular TI piezoelectric materials are listed in
Table 2.1.
Similarly, for the TI piezomagnetic case with the z-axis being the poling direction
or axis of material symmetry, we have

σ xx = c11γ xx + c12 γ yy + c13γ zz − q31 Hz


σ yy = c12 γ xx + c11γ yy + c13γ zz − q31 Hz
σ zz = c13γ xx + c13γ yy + c33γ zz − q33 Hz
(2.12a)
σ yz = 2c44 γ yz − q15 H y
σ xz = 2c44γ xz − q15 H x
σ xy = 2c66γ xy

Bx = 2q15γ xz + µ11 H x
By = 2q15γ yz + µ11 H y (2.12b)
Bz = q31 (γ xx + γ yy ) + q33γ zz + µ 33 H z

For either the piezoelectric or piezomagnetic case, the independent number of


material constants is ten.
Listed in Table 2.3 are the material properties for some of the popular TI
piezomagnetic materials.

2.6 Special Case: Transversely Isotropic or Isotropic Elastic Solids


For the purely elastic but TI case with the z-axis being the axis of material symmetry,
the constitutive relations are
38 Governing Equations

Table 2.3. Material Properties for Some Popular TI Piezomagnetic (or magnetostrictive)
Materials

CoFe2O4 Terfenol-D Inverse NZFO CFO NFO LSMO F-7A1 Alfenol

c11 286 8.541 s11 6.5 6.5 6.5 15 6.62 7.14


c12 173 0.654 s12 −2.4 −2.4 −5 −2.65 −2.14
c13 170.3 3.91
c33 269.5 28.3
c44 45.3 5.55
c66 56.5 18.52
ε11 0.08 0.05
ε33 0.093 0.05 ε33/ε0 10 10 10
μ11 5.9 0.086
μ33 1.57 0.023 μ33/μ0 2 3 3 20 58
q31 580.3 −5.75 q31 556 125 −120
q33 699.7 270.1 q33 −1880 −680 250 −3660 −7100
q15 550 155.56 q11 −918
ρ 5.3 5.37

Notes:
(1) CoFe2O4 and Terfenol-D in Columns 2 and 3 are from Kuo and Peng (2013).
(2) In the inverse relations, the compliances sij are in 10−12m2/N and the piezomagnetic coupling coefficients
qij are in 10−12m2/A.
(3) NZFO (Ni1−xZnxFe2O4) in Column 5 are from Petrov and Srinivasan (2008).
(4) CFO (Cobalt ferrite), NFO (Nickel Ferrite), and LSMO (Lanthanum strontium manganite) in
Columns 6–8 are from Bichurin et al. (2003) where ε0 and μ0 are the permittivity and permeability in
vacuum.
(5) F-7A1 (Ferroxcube 7A1) and Alfenol in Columns 9–10 are from Avellaneda and Harshe (1994).

σ xx = c11γ xx + c12γ yy + c13γ zz


σ yy = c12γ xx + c11γ yy + c13γ zz
σ zz = c13γ xx + c13γ yy + c33γ zz
(2.13)
σ yz = 2c44γ yz
σ xz = 2c44γ xz
σ xy = 2c66γ xy

with only five independent material coefficients. For the corresponding elastic
isotropic case, we have c33=c11, c13=c12, c44=c66, with also c66=(c11−c12)/2. Thus, for
this case, only two elastic constants are required in the constitutive relations. The
relations among these coefficients are

E(1 − ν)
c11 = c33 = λ + 2µ =
(1 + ν)(1 − 2 ν)

c12 = c13 = λ = (2.14)
(1 + ν)(1 − 2 ν)
E
c44 = c66 = µ =
2(1 + ν)

where cij are the elements of the elastic stiffness matrix, E the Young’s modulus, ν
the Poisson’s ratio, λ (= 2 µν / (1 − 2ν )) and μ the Lamé constants. Notice that μ is also
the shear modulus.
2.8 Two-Dimensional Governing Equations 39

2.7 Special Case: Cubic Elastic Solids


For the purely elastic but cubic case, the constitutive relations are

σ xx = c11γ xx + c12 γ yy + c12 γ zz


σ yy = c12 γ xx + c11γ yy + c12 γ zz
σ zz = c12 γ xx + c12γ yy + c11γ zz
(2.15)
σ yz = 2c44γ yz
σ xz = 2c44γ xz
σ xy = 2c44γ xy

where c11, c12 and c44 are the three independent material coefficients. We point out
that most semiconductor materials are cubic, as listed in Table 2.4 of the material
properties for some of the popular elastic materials (with most of them being of
cubic symmetry).

2.8 Two-Dimensional Governing Equations


We assume the general MEE solid is under a generalized plane-strain deformation in
the (x, z)- or (x1, x3)-plane, namely, we have ∂/∂y=0. Therefore, the governing equations
(including both the equilibrium equations and gradient relations) are reduced to

σ xx , x + σ xz,z + fx = 0
σ xy, x + σ yz,z + fy = 0 (2.16a)
σ xz, x + σzz,z + fz = 0

Dx, x + Dz,z − fe = 0
(2.16b)
Bx, x + Bz,z − fh = 0

∂ux ∂u
γ xx = , γ zz = z
∂x ∂z
1  ∂ux ∂uz 
γ xz =  +  (2.16c)
2  ∂z ∂x 
1 ∂uy 1 ∂uy
γ xy = , γ yz =
2 ∂x 2 ∂z

∂φ ∂φ
Ex = − , Ey = 0, Ez = − (2.16d)
∂x ∂z

∂ψ ∂ψ
Hx = − , H y = 0, Hz = − (2.16e)
∂x ∂z

The constitutive relations for the 2D case (generalized plane-strain deformation)


are the same as for the 3D case (since most field quantities are involved, except that
they are independent of the y-coordinate). The generalized plane-strain deformation
40 Governing Equations

Table 2.4. Material Properties for Some Popular Elastic Materials (with cubic symmetry, with
the same units as in Table 2.1)

Material c11 c12 c44 ρ ε/ε0

Ag 124.0 93.4 46.1 10.49


Al 108.2 61.3 28.5 2.79
Au 186.0 157.0 42.0 19.39
Cr 350.0 57.8 101.0 7.19
Cu 168.4 121.4 75.4 8.96
Fe 242.0 146.5 112.0 7.87
Ge 128.9 48.3 67.1 5.32a 16.2a
Ir 580 242 256 22.56
K 4.57 3.74 2.63 0.86
Li 14.8 12.5 10.8 0.53
Mo 460.0 176.0 110.0 10.28
Na 6.03 4.59 5.86 0.97
Nb 246.0 134.0 28.7 8.57
Ni 246.5 147.3 124.7 8.91
Pb 46.6 39.2 14.4 11.34
Pd 227.1 176.0 71.7 12.02
Pt 346.7 250.7 76.5 21.45
Ta 267.0 161.0 82.5 16.69
Th 75.3 48.9 47.8 11.70
Si 165.7 63.9 79.6 2.33a 11.9a
V 228.0 119.0 42.6 5.80
W 521.0 201.0 160.0 19.25
AgBr 56.3 33.0 7.20 6.47
KCl 39.8 6.2 6.25 1.98
LiF 111.2 42.0 62.8 2.64
MgO 286.0 87.0 148.0 3.58
NaCl 48.7 12.4 12.6 2.16
PbS 127.0 29.8 24.8 7.60a 169a
Diamond 1076.0 125.0 576.0 3.52 5.5–10

Notes:
(1) Data are from Hirth and Lothe (1982) under room temperature.
(2) a) Static relative dielectric constants and densities are from Singh (2003).
(3) Extra data of some TI (hexagonal) materials from Bacon et al. (1979) are listed:

Material c11 c12 c13 c33 c44 ρ

Be 292.3 26.7 14.0 336.4 162.5 1.85


Cd 113.8 39.2 40.0 50.6 20.0 8.65
Hf 181.1 77.2 66.1 196.9 55.7 13.20
Mg 59.5 26.1 21.8 61.6 16.4 1.74
Ti 162.4 92.0 69.0 180.7 46.7 4.54
Zn 163.7 36.4 53.0 63.5 38.8 7.13
Zr 143.4 72.8 65.3 164.8 32.0 6.49

(4) Static relative dielectric constants and densities of other materials are from online using Google search.
(5) Based on the material properties for single crystals, one can obtain the corresponding compound
properties using the simple mixture rule. For example, the Silicon-Germanium (Si1−xGex) has a
crystal structure of diamond and its elastic constants can be expressed as c11 = 165.7–36.8x, c12 =
63.9–15.6x, and c44 = 79.6–12.5x (GPa), its density as 2.33+2.99x (g/cm3), and its relative static
dielectric constant as 11.9+4.3x.
(6) The reader is also referred to Table 1.4 in Ding et al. (2006) for some other transversely isotropic
materials.
2.10 Applications of Betti’s Reciprocal Theorem 41

is reduced to the well-known plane-strain deformation when the involved material


possesses certain high symmetry property. In the latter case, the in-plane strain and
antiplane strain deformations are decoupled. It is noted that in this chapter and
also in Chapter 4, we have assumed that the plane-deformation is in the (x,z)-plane,
instead of the (x,y)-plane as in Chapters 1 and 3. The reason for doing so is to have
the same consistent definitions for the Stroh eigenvalues and eigenvectors in the
involved chapters (Chapters 2, 4, 8, and 9).

2.9 Extended Betti’s Reciprocal Theorem


It is well known that Betti’s reciprocal theorem is very important in structural
mechanics analysis. To the community of the boundary element methods, it is the
foundation of the boundary integral equations. Here we further show that this
theorem can also connect sources of different natures, with one of the important
relations being that between the point “force” and point “dislocation”.
Relations among different concentrated sources and their responses can be
studied using Betti’s reciprocal theorem, which states that for two systems (1)
and (2) belonging to the same material space, the following relation holds (Pan
1999a)

σ iJ(1)γ Ji
( 2)
= σ iJ( 2)γ Ji
(1)
(2.17)

This is a direct result of the linear constitutive relations (2.5) along with the symmetric
properties of various material tensors.
From Eq. (2.17), one can easily derive the following integral equation for these
two systems

(1) ( 2)
∫S σiJ uJ ni dS − ∫ σ(iJ1,)i u(J2) dV = ∫ σ(iJ2)u(J1) ni dS − ∫ σ(iJ2,)i u(J1)dV (2.18)
V S V

This integral identity can then be applied to different loading systems made of the
same materials. We now illustrate some of the important results obtained by direct
applications of Eq. (2.18).

2.10 Applications of Betti’s Reciprocal Theorem

2.10.1 Relation between Extended Point Forces and Extended


Point Dislocations
In Eq. (2.18), we let system (1) be the real boundary value problem and (2) the
corresponding extended “point force” Green’s function problem, that is,

σ iJK,i = −δ JK δ ( x pf − x sp ) (2.19)

where the field point is at xf and the extended point force is applied at xs in the
K-direction. It is noted that we have used the superscript K to indicate the direction
of the force and to avoid possible confusion in the notations. Similar superscript
will be also applied to indicate the nature of the source when applying the Betti’s
42 Governing Equations

reciprocal theorem. However, in most expressions in this book, the superscript K will
be moved down as a subscript to represent the direction of the force.
With Eq. (2.19), Eq. (2.18) can be reduced to a well-known integral representation
of the displacement field:

uK ( x sp ) = ∫ [σ iJ ( xqf )uJK ( x sp ; xqf ) − σ iJ


K ( x s ; x f )u ( x f )]n ( x f )dS( x f )
p q J q i q q
S
(2.20)
+ ∫ uJK ( x sp ; xqf ) f J ( xqf )dV ( xqf )
V

where in the Green’s function expressions for the displacements uJK and stresses σ iJK ,
the single superscript K denotes that the Green’s function corresponds to a point
force in the K-direction.
Now, we wish to find the displacement response due to a prescribed dislocation
(displacement discontinuity) across a surface Σ imbedded in V (or the dislocation
Green’s function). Let ni (= ni− = −ni+) be the unit normal to Σ, bJ = uJ+−ui− being the
dislocation on the plane with normal ni. This dislocation along Σ may have any form
provided that the following traction-continuity condition holds:

σ iJ+ ni+ + σ iJ− ni− = 0 (2.21)

We assume that the extended displacement and stress fields satisfy the same
homogeneous boundary condition on the outer boundary S, and apply Eq. (2.20) to
the region bounded internally by Σ and externally by S. We then arrive at (also omit
the volumetric integral term associated with the body force)

K ( x s ; x f )b ( x f )n ( x f )dΣ ( x f )
uK ( x sp ) = ∫ σ iJ (2.22)
p q J q i q q
Σ

It is noted that the kernel function in Eq. (2.22) is the Green’s stress with component
(iJ) at the field point x pf due to a point force at x sp in the K-direction. Alternatively,
the displacement response due to the dislocation density “tensor” bJni can also be
expressed by the kernel displacement function due to a point dislocation, namely,

uK ( xqf ) = ∫ uK
iJ ( x s ; x f )b ( x s )n ( x s )dΣ ( x s )
p q J p i p p (2.23)
Σ

where uK iJ represents the induced displacement in K-direction due to a point

dislocation with its plane normal in the i-direction and its Burgers vector component
in the J-direction.
Comparing Eq. (2.23) with Eq. (2.22), we immediately obtain the following
important equivalence between the stress due to a point force and the displacement
due to a point dislocation

iJ ( x s ; x f ) = σ K ( x f ; x s )
uK (2.24)
p q iJ q p

It is noted that the positions of the source and field points in the point-force Green’s
stresses need to be exchanged in order to obtain the point-dislocation Green’s
displacements. This is the simplest and yet very important relation. Similar results
2.10 Applications of Betti’s Reciprocal Theorem 43

for poroelastic media were derived by Pan (1991). Several important observations
of Eq. (2.24) are listed:
Remark 2.8: In reading Eq. (2.24), one must keep in mind the exact physical meanings
of the two terms on each side; otherwise one may mistake that Eq. (2.24) would
be wrong since the displacement on the left-hand side cannot equal the stress on
the right-hand side. The correct explanation of this equation is that the dislocation-
induced displacement per unit area of the dislocation is equal to the force-induced
stress per magnitude force.
Remark 2.9: In general, once the point-force Green’s functions are solved, the
corresponding point-dislocation Green’s functions can be obtained through the
relation (2.24). In deriving relation (2.24), we have assumed that the system is
linear, but can be fully coupled MEE with general anisotropy and heterogeneity.
In particular, this relation can be used to derive the point-dislocation Green’s
functions in horizontally layered systems, including half-space and bimaterial
domains as special cases. For example, the point-dislocation Green’s functions
in horizontally layered media can be derived in both the Fourier or physical
domains using Eq. (2.24) and the corresponding point-force solutions (Pan 1989,
1990, 1999b).
Remark 2.10: Direct solution of the point-dislocation Green’s functions is also possible
but the procedure may be very complicated. The way to achieve this is to derive
the equivalent body force of the point dislocation, find the related discontinuity of
the physical quantities, and solve for the unknowns, using the method as previously
employed by Pan (1989) for the TI and layered half-space.
Remark 2.11: For the elastic isotropic or TI bimaterial, half-space, or full-space, each
term on the right-hand side of Eq. (2.24) is proportional to various eigenstrains, such
as the misfit lattice strain, the nucleus of strain (or a nucleus of strain multiplied by
the elastic constants), and so forth. With Eq. (2.24), however, it is unnecessary to
add all the related nuclei of strain together and enforce the boundary or interface
condition to solve the coefficients involved. In other words, it is very convenient to
apply Eq. (2.24) to find various dislocation-related solutions once the corresponding
point-force solutions are given.
Remark 2.12: In applying Eq. (2.24), one must keep in mind that on the left-hand
side, xs and xf are the field and source points in the point-dislocation problem,
respectively; while on the right-hand side, xf and xs are the field and source points in
the point-force problem, respectively. Therefore, the Green’s displacements due to a
point dislocation can be obtained from the Green’s stresses due to a point “force” by
exchanging the positions of the field and source points and by assigning the suitable
meanings to the associated indexes.
Remark 2.13: For a homogeneous and infinite domain, expressing the point-force
Green’s stresses by the strains and substituting the result into Eq. (2.24) and then
into Eq. (2.23), one can obtain the extended Volterra relation (Steketee 1958). It
is noted that for this specific case, the point-force Green’s stresses are functions of
the relative vector from the source to field points, that is, xf − xs, and they satisfy the
following relation:
44 Governing Equations

σ iJK ( x pf ; x sp ) = σ iJK ( x pf − x sp ) = −σ iJK ( x sp − x pf ) = −σ iJK ( x sp ; x pf ) (2.25)

We therefore have

iJ ( x s ; x f ) = −σ K ( x s ; x f )
uK (2.26)
p p iJ p p

It should be emphasized that only for the homogeneous infinite domain, can the
Green’s displacements due to a point dislocation be obtained directly from the
Green’s stresses due to a point force without exchanging the field and source
positions! For all other situations, the dislocation-induced Green displacements
should be obtained strictly using Eq. (2.24).
For a homogeneous and infinite 3D solid of purely elastic isotropy, Eq. (2.26) is
reduced to

Eν E
ukij ( x sp ; x pf ) = −σ ijk ( x sp ; x pf ) = − ulk,lδ ij − (uk + ukj ,i ) (2.27)
(1 + ν )(1 − 2ν ) 2(1 + ν ) i, j

where E and ν are the Young’s modulus and Poisson’s ratio, respectively. The
derivatives of the Green’s elastic displacements due to a point force at xs in the k-
direction are taken with respect to the field point xf.
By contrast, if we take the derivatives of the point-force Green displacements
with respect to the source point xs, we then have the following nuclei of strain

Eν E
ε kij ( x sp ; x pf ) = uk δ ij + (uk + ukj ,i ) (2.28)
(1 + ν )(1 − 2ν ) l ,l 2(1 + ν ) i, j

which is opposite to that given by Eq. (2.27).


The concept of nuclei of strain as given by Eq. (2.28) was introduced by
Mindlin (1936) when solving the elastic half-space problem. It is interesting
to note that while the first term corresponds to the center of compression or
dilatation, the other terms are double forces. Therefore, from the physical point
of view, the point-dislocation Green’s functions can be constructed through
superposition of various nuclei of strain (or the derivatives of the point-force
Green’s displacements), with their coefficients being solely related to the elastic
constants. This is a physical explanation for the mathematical and arbitrary
equivalent relation (2.26). It is obvious that if the point-force Green’s functions
can be derived in an exact-closed form (or explicit form), the corresponding point-
dislocation solutions will also have the same features since they are obtained by
the superposition of various nuclei of strain. Detailed analyses can be found in
Maruyama (1964, 1966) for the isotropic elastic case and in Yu et al. (1994) for
the TI elastic case. For materials of either isotropy or transverse isotropy, these
exact closed-form nuclei of strain can also be employed to derive the point-force
Green’s functions in a half-space or a bimaterial space, and to derive the solutions
corresponding to various inclusions.
It is noted that, the Green’s function relations between two different sources
that we derived so far are strictly between those due to a point dislocation and those
2.10 Applications of Betti’s Reciprocal Theorem 45

due to a point force in 3D. For the 2D case, the Green’s functions due to the line
force and those due to the line dislocations (open all the way to the half infinite line)
have the same singularity order. These will be clearly observed from the 2D Green’s
function expressions in Chapter 4, and one should pay particular attention to this
difference, as discussed in the following text.

2.10.2 Relation between Extended Line Forces and Extended


Line Dislocations
For the anisotropic elastic case, Chou and Barnett (1997a, 1997b) derived a relationship
between the deformation field due to a line force and due to a straight dislocation
(or a line dislocation) in an infinite homogeneous anisotropic plane. The Moutier’s
theorem in thermostatics and Stroh formalism for anisotropic plane deformation
were applied to derive the relation. Here, based on the extended notation (including
both electric and magnetic fields), we briefly present the most general and important
relation by simply utilizing the Green’s function solutions of the line force and line
dislocation in Chapter 4.
We let zR = x1f + pR x3f and sR = x1s + pR x3s be, respectively, the Stroh eigenvalue
pR-modified field and source coordinates. From Chapter 4, we observe that the
extended “stress” function in the J-direction at field point xf due to a unit line force
in the K-direction applied at xs in an infinite, anisotropic and uniform plane can be
expressed as (on the right-hand sides of Eqs. (2.29)–(2.31), the superscript for the
force direction K is now moved down as a subscript)

5
1
φJK ( x f ; x s ) = Im ∑ [ BJR ln(zR − sR ) AKR ] (2.29)
π R =1

where A and B are the Stroh eigenmatrices corresponding to the Stroh eigenvalue
pR. We will also observe from Chapter 4 that the extended displacement in the
J-direction at field point xf due to a unit line (or straight) dislocation in the
K-direction at xs in the same infinite plane is

5
1
uJK ( x f ; x s ) = Im ∑ [ AJR ln(zR − sR )BKR ] (2.30)
π R =1

which can be equivalently written as

5
1
uJK ( x f ; x s ) = Im ∑ [ BKR ln(zR − sR ) AJR ] (2.31)
π R =1

Comparing Eq. (2.31) with Eq. (2.29), one immediately observes that the extended
displacement in the J-direction due to a unit line dislocation in the K-direction is the
same as the extended “stress” function in the K-direction due to a unit line force in
the J-direction. In other words, the Green’s displacement due to a unit line dislocation
is just the transpose of the Green’s “stress” function due to a unit line force, without
46 Governing Equations

even exchanging the source and field locations. Symbolically, this extremely simple
yet very important relation can be written as

uJK ( x f ; x s ) line dislocation = φKJ ( x f ; x s ) line force (2.32)

It is clearly observed that this relation not only holds for general fully coupled
MEE materials, but also can be used as a basic relation to derive many interesting
relations between the line-force and line-dislocation solutions in an infinite and
uniform plane (by taking various orders of derivatives of Eq. (2.32)). One should
further notice that while the 3D relation (2.24) between the fundamental solution
due to the point force and that due to the point dislocation holds not only for
general material anisotropy, but also for general material heterogeneity, Eq. (2.32)
of the 2D relation between the line force and line dislocation holds only for a
homogeneous material.

2.11 Basics of Eshelby Inclusion and Inhomogeneity

2.11.1 The Eshelby Inclusion Problem


We now extend the classic Eshelby inclusion problem (Eshelby 1957, 1959, 1961;
Mura 1987) to that in an MEE material. For easy discussion, the extended physical
quantities are simply taken without using “extended.” The Eshelby problem is
associated with a homogeneous matrix M where part of which, say V, is under the
eigenstrain or stress-free transformation strains γ Ij* (Figure 2.1). The procedure
is as follows: One first applies an eigenstrain field γ Ij* to the inclusion in such a
way that both the inclusion and matrix are stress-free. In order to do so, one could
cut out the inclusion and apply the eigenstrain field to the inclusion domain only.
However, due to this applied eigenstrain γ Ij* , the inclusion is deformed. One now
needs to apply certain tractions to its boundary S and also distributed body forces
throughout the whole inclusion domain V to strain the deformed inclusion back
to its original shape. The applied tractions and (equivalent) body forces are σ iJ* ni
and −∂σ iJ* / ∂xi , respectively, with ni being the outward normal of the inclusion
boundary. It is noticed that this special stress field is related to the eigenstrain field
by the following general constitutive relations for the homogeneous matrix M with
material property ciJKl

σ iJ* = ciJKl γ Kl
* (2.33)

It is also noted that after we squeeze the inclusion back into the matrix, there will
be an induced displacement field both in the inclusion and the matrix due to the
applied traction along S and the distributed body force in V. In other words, by
applying the integral expression (2.20), the induced displacement field due to the
traction and body force can be expressed as (Bacon et al. 1979)

∂σ*iJ ( xqf )
uK ( x ps ) = − ∫ uJK ( x ps ; xqf ) dV ( xqf ) + ∫ uJK ( x ps ; xqf )σ*iJ ( xqf )ni ( xqf )dS( xqf ) (2.34)
V ∂xif S
2.11 Basics of Eshelby Inclusion and Inhomogeneity 47

Step 1: M
Cut out V from V
matrix M V

Step 2:
Apply eigenstrain V
field to inclusion V *
γ Kl

M
Step 3:
Apply the body force in ∂σ*iJ (xqf )
the deformed domain and ∂x fi V
the traction over the n
deformed boundary σ *iJ (x fq)ni (x fq)

Figure 2.1. On the basics of Eshelby inclusion.

where uJK ( x sp ; xqf ) represents the Green’s displacement in the J-direction at field
point xf due to the point force in the K-direction applied at xs. Converting the surface
integral to the volume one by integration by parts, we then have

∂uJK ( x sp ; xqf )
uK ( x sp ) = ∫ σ*iJ ( xqf ) dV ( xqf ) (2.35)
V ∂xif

The corresponding strains γKl can be obtained by taking the derivatives of the
displacement field

∂uJK ( x sp ; xqf )
uK ,l ( x sp ) =∫ σ*iJ ( xqf ) dV ( xqf ) (2.36)
V ∂xls ∂xif

If the eigenstrain field is uniform inside the inclusion, Eqs. (2.35) and (2.36) can be
simplified to

∂uJK ( x sp ; xqf )
uK ( x sp ) = ciJLm γ *Lm ∫ dV ( xqf )
V ∂xif (2.37)
= ciJLm γ *Lm ∫ uJK ( x sp ; xqf )ni ( xqf ) dS( xqf )
S

∂uJK ( x sp ; xqf )
uK , p ( x sp ) = ciJLm γ *Lm ∫ dV ( xqf )
V ∂x sp ∂xif
= ciJLm ∫ uJK, p ( x sp ; xqf )ni ( xqf )dS( xqf ) γ *Lm (2.38)
 S 
≡ SKpLm γ *Lm
48 Governing Equations

In the second expressions in Eqs. (2.37) and (2.38), integration by parts is used. The
third expression in Eq. (2.38) defines the extended Eshelby tensor (Eshelby 1961;
Pan 2004a, 2004b), as also discussed in Chapter 4, The final states of the strain and
stress fields in both the inclusion (with superscript “(I)”) and matrix (with superscript
“(M)”) are

(I )
γ Kl = γ Kl − γ Kl
*

σ iJ( I ) = ciJKl (γ Kl − γ Kl
* )

(M )
(2.39)
γ Kl = γ Kl
σ iJ( M ) = ciJKl γ Kl

2.11.2 The Eshelby Inhomogeneity Problem


We assume that σ iJ0 is the applied extended stress field at infinity if the material is
homogeneous, that is, without inhomogeneity. The corresponding extended strain
field is γ Ij0 by the generalized constitutive relations for the infinite domain, that is,
σ iJ0 = ciJLmγ Lm
0 . Therefore, with an inhomogeneity V (of material property c*
iJLm)
in the infinite matrix M (of material property ciJLm), the generalized constitutive
relations in the inhomogeneity V and the matrix can be assumed, based on the classic
Eshelby inhomogeneity relation (Eshelby 1961) as,

p
σ0iJ + σiJ = ciJLm
* ( γ 0Lm + γ Lm − γ Lm ) (V )
(2.40a, b)
σiJ + σiJ = ciJLm ( γ Lm + γ Lm )
0 0 (M − V )

p
where σiJ and γIj are the induced stress and strain fields, and γ Lm is the given
eigenstrain.
Introducing the equivalent eigenstrain field concept γ Ije , we can equally express
Eq. (2.40a, b) as

p
σ 0iJ + σ iJ = ciJLm ( γ 0Lm + γ Lm − γ Lm − γ eLm ) (V )
(2.41a, b)
σ 0iJ + σ iJ = ciJLm ( γ 0Lm + γ Lm ) (M − V )

Let

p
γ **
Ij = γ Ij + γ Ij
e (V ) (2.42)

Equation (2.41) can then be rewritten as

σ0iJ + σiJ = ciJLm ( γ 0Lm + γ Lm − γ **


Lm ) (V )
(2.43a, b)
σiJ + σiJ = ciJLm ( γ Lm + γ Lm )
0 0 (M − V )

Comparing the stress expressions between Eq. (2.43) and those in Eq. (2.39), one
immediately notices the similarity between these expressions so that the solutions
2.12 Summary and Mathematical Keys 49

corresponding to the inhomogeneity problem can be solved. In other words, the


problem described by Eq. (2.43) could be symbolically solved using Eqs. (2.37) and
(2.38) with γ Ij* in Eqs. (2.37) and (2.38) being replaced by γ Ij**. One should, however,
also keep in mind that in deriving Eqs. (2.37) and (2.38), we have assumed that the
eigenstrains γ Ij* are uniform, which is true only for the classic elliptic or ellipsoidal
inhomogeneity discussed in the literature (Eshelby 1961; Mura 1987). In a polygonal
inhomogeneity V in general, the equivalent eigenstrain γ Ije is not uniform (Pan 2004a,
2004b), and therefore the new eigenstrain γ Ij** in V will not be uniform; rather it will
be a function of the position within V. Solutions to these complicated problems can
be found using an iterative approach (Sun et al. 2012).
The eigenstrain problem described in the preceding text for the inhomogeneity
also shares the similar equivalent body force. To see this, we first write the equilibrium
equation for the stresses and the balance for the electric displacements and magnetic
inductions in V (Pan 2004a, 2004b; Jiang and Pan 2004)

σ iJ ,i = 0 (2.44)

Then, substituting Eq. (2.43a) into Eq. (2.44), we have

ciJKl uK ,li = ciJKl γ Kl


**
,i (2.45)

It is clear that the right-hand side of Eq. (2.45) resembles the extended body force as
it would appear on the left-hand side of Eq. (2.45),

f J = −ciJKl γ Kl
**
,i (2.46)

which is the equivalent body force corresponding to the new eigenstrain, similar to
the one in the corresponding inclusion problem (Pan 2004a, 2004b).

2.12 Summary and Mathematical Keys

2.12.1 Summary
In this chapter, we have presented the governing equations associated with
anisotropic and MEE fully coupled materials or composites. The key point is the
coupling among different fields since this phenomenon can be directly applied to
generate a desirable field by using a different field as a source. Besides the general
anisotropic and fully MEE coupling, various field-decoupled cases and also the
materials with high material symmetry (TI, cubic, isotropic) are also presented. The
corresponding 2D governing equations are also listed. Near the end of the chapter,
Betti’s reciprocal theorem is applied to find the relation between the induced field
by the point force and that by the point dislocation, and the relation between the
induced field by the line force and that by the line dislocation. The extended Green’s
representation (2.20) is applied to derive the basic Eshelby relations. It should be
noted that while the Betti’s relation in 3D holds for general heterogeneous materials,
the relation presented in 2D holds only for the homogeneous materials.
50 Governing Equations

2.12.2 Mathematical Keys


The extended Betti’s reciprocal theorem and the extended Green’s representation
should be the mathematical keys in this chapter. The governing equations from the
energy point of view, in the spirit of Green (1828), and the transformation between
different coordinates in Appendices A and B are also important.

2.13 Appendix A: Governing Equations from the Energy Point of View


First, the fully coupled constitutive relations (2.2a) can be derived based on the
energy concept in MEE solids. Similar to the piezoelectric case (Tiersten 1969), we
define the following MEE enthalpy density G (a quadratic function) as

G = U − Ei Di − H i Bi (A1)

where U is the internal energy function of the system with its increment being given
by, according to the first law of thermodynamics

dU = dW + dQ − dK = σ ij dγ ij + Ei dDi + Hi dBi + Tdη (A2)

where W, Q and K are the work done by external stimuli, the heat imported into
the body, and the kinetic energy, respectively, and T and η are the temperature and
entropy, respectively.
Since G is a total differential, it can be observed from Eqs. (A1) and (A2) that

∂G ∂G ∂G ∂G
σij = , Di = − , Bi = − , T= (A3)
∂γ ij ∂Ei ∂H i ∂η

We assume that G is a quadratic function of the strain γij, electric field Ei, and
magnetic field Hi, and omit the effect of thermal field. Then we have

1 1 1
G= cijlmγ ij γ lm − ε ij Ei E j − µij H i H j − eijk Ei γ jk − qijk H i γ jk − α ij Ei H j (A4)
2 2 2

This expression is the first-order approximation of any real physical system with the
full MEE coupling, which assumes no preexisting stresses, electric displacements,
and magnetic inductions before the application of any external stimulus.
Substituting Eq. (A4) into Eq. (A3) gives the constitutive relations as in Eq.
(2.2a).
In order to derive the boundary value problem of the MEE system, we define
the following “potential energy” by extending the purely elastic expression

1
P(ui , φ ,ψ ) = ∫ [
V 2
(σ ij γ ij + Di Ei + Bi Hi ) − fi ui + feφ + fhψ ]dV
(A5)
− ∫ ti ui dS − ∫ Dn φ dS − ∫ Bn ψ dS
St Sd Sb

where V is the problem domain and S is the corresponding boundary. The quantity
with an overbar is the given value on the boundary S = St + Su = Sd + Sφ = Sb + Sψ
2.14 Appendix B: Transformation of MEE Material Properties 51

and St∩Su = Sd∩Sφ = Sb∩Sψ = Ø. It is noted that St and Su are the boundaries where
the traction and displacement are respectively given, Sd and Sφ are the boundaries
where the normal electric displacement and the electric potential are respectively
given, and Sb and Sψ are the boundaries where the normal magnetic induction and
the magnetic potential are respectively given. Also in Eq. (A5), ti , Dn and Bn are,
respectively, the tractions, the negative electric and magnetic charges prescribed on
the appropriate portions of the surface.
We now take the variation of P with respect to ui, φ, and ψ while satisfying the
boundary conditions on Su, Sφ, and Sψ. Then we have

1
δ P(ui , φ, ψ ) = ∫
V 2
( )
δ σij γ ij + Di Ei + Bi Hi dV + ∫ [ − fi δui + fe δφ + fh δψ ] dV
V
   (A6)
− ∫ ti δui dS − ∫ Dn δφ dS − ∫ Bn δψ dS
St Sd Sb

Integrating this expression by parts and further simplifying the result, we arrive at

δ P(ui , φ, ψ ) = ∫ [( −σ ji, j − fi )δui + ( − Di,i + fe )δφ + ( − Bi,i + fh )δψ ]dV


V
   (A7)
+ ∫ (ti − ti )δui dS + ∫ (Dn − Dn )δφ dS + ∫ ( Bn − Bn )δψ dS
St Sd Sb

where the gradient relations between (γij, Ei, Hi) and (ui, φ, ψ), that is, Eq. (2.3),
have been utilized. Since δui, δφ, and δψ are arbitrary variations, δP = 0 implies the
following boundary value problem

σ ji, j + fi = 0 (in V )
Di,i − fe = 0 (in V )
Bi,i − fh = 0 (in V )
σ ji n j − ti = 0 (on St ) (A8)

Dj n j − Dn = 0 (on Sd )
Bj n j − Bn = 0 (on Sb )

The first three sets give the equilibrium laws of the mechanical, electric, and
magnetic fields as specified in Eq. (2.1), while the other three are the natural
boundary conditions.

2.14 Appendix B: Transformation of MEE Material Properties


from One Coordinate System to the Other
We assume that the relation between the two Cartesian coordinate systems is

* =a x
xm (B1)
mi i

where ami are the directional cosines between respective axes of the two coordinates
* and x ). Generally, we have a ≠ a . Equation (B1) is also the transformation
(i.e. xm i mi im
of any vector (or the first-order tensor) between the two coordinate systems.
52 Governing Equations

For a second-order tensor, such as the dielectric permittivity and magnetic


permeability coefficients, we have

ε lm
* =a a ε
li mj ij (B2)

Or, in terms of the matrix expression (all 3×3),

[ε * ] = [a][ε ][a]t (B3)

with the superscript t denoting the transpose of the matrix.


For a third-order tensor, like the piezoelectric coefficients, we have

*
elmn = ali amj ank eijk (B4)

and for a fourth-order tensor, like the elastic stiffness coefficients, we have

*
clmnp = ali amj ank a pq cijkq (B5)

We now introduce the matrix expressions for all the components by using the
convention between two indices and one index (i.e., the Voigt notation). In other
words, their corresponding relation is 11 to 1, 22 to 2, 33 to 3, 23 to 4, 31 to 5, and 12
to 6. As such, the fourth-order elastic tensor cijkl (i,j,k,l vary from 1 to 3) is condensed
to cij (i,j vary from 1 to 6). We further define the following column matrices for the
stresses and strains as

 σ1   σ11   γ 1   γ 11 
σ 2  σ 22   γ 2   γ 22 
[σ]6 ×1  σ 3   σ
≡ σ ≡ σ , [ γ ]6 ×1 ≡  γγ 3  ≡  2γγ33 
33  (B6)
 4   23   4   23 
 σ 5   σ 31   γ 5   2 γ 31 
 σ 6   σ12   γ 6   2 γ 12 

The piezoelectric ekij and piezomagnetic qkij tensors become the matrices of size 3×6
(the last two indices are condensed to one, varying from 1 to 6). In other words,
we have,

 e11 e12 e13 e14 e15 e16 


[e]3×6 ≡ e21 e22 e23 e24 e25 e26  (B7a)
e31 e32 e33 e34 e35 e36 

 q11 q12 q13 q14 q15 q16 


[q]3×6 ≡  q21 q22 q23 q24 q25 q26  (B7b)
 q31 q32 q33 q34 q35 q36 

Thus, the constitutive relation (2.2a) in the index form can be written equivalently
in the matrix form as

[σ ] = [c][γ ] − [e]t [E ] − [q]t [ H ]


[D] = [e][γ ] + [ε ][E] + [α ][ H ] (B8)
[ B] = [q][γ ] + [α ]t [E] + [ µ ][ H ]
2.14 Appendix B: Transformation of MEE Material Properties 53

where [ε], [α], and [μ] are 3×3 matrices, and [E], [H], [D], and [B] are all 3×1 column
matrices.
Similarly, in the coordinate system with *, we have,

[σ * ] = [c* ][γ * ] − [e* ]t [ E* ] − [q* ]t [ H * ]


[D* ] = [e* ][γ * ] + [ε * ][E* ] + [α * ][ H * ] (B9)
[ B* ] = [q* ][γ * ] + [α * ]t [ E* ] + [ µ * ][ H * ]

For a symmetric second-order tensor such as the stress tensor, its transformation
relation can be written in an equivalent matrix form as

[σ * ] = [ K ][σ ] (B10)

where as in Ting (1996),

K1
[K ] =  K 2K 2  (B11)
 3 K 4 

 a11
2 2
a12 2 
a13
 2
[K1 ] = a21 2
a22 a23 
2 (B12)
 2 2 2 
a31 a32 a33 

 a12 a13 a13a11 a11a12 


[K 2 ] = a22 a23 a23a21 a21a22  (B13)
a32 a33 a33a31 a31a32 

a21a31 a22 a32 a23a33 


[K 3 ] =  a31a11 a32 a12 a33a13  (B14)
 a11a21 a12 a22 a13a23 

a22 a33 + a23a32 a23a31 + a21a33 a21a32 + a22 a31 


[K 4 ] =  a32 a13 + a33a12 a33a11 + a31a13 a31a12 + a32 a11  (B15)
 a12 a23 + a13a22 a13a21 + a11a23 a11a22 + a12 a21 

It can be easily shown that the following coordinate transformation relation for the
corresponding strain tensor holds

[γ * ] = ([K ]−1 )t [γ ] (B16)

where

([K ]−1 )t =  2K 1 K2  (B17)


 K 3 K 4 

We also have

[E* ] = [a][E], [ H * ] = [a][ H ]


(B18)
[D* ] = [a][D], [ B* ] = [a][ B]
54 Governing Equations

Then, substituting the original ones for [σ], [γ], [D], [E], [H], and [B] in the constitutive
relations (B9) by the transformed ones with *, we have

[K ][σ ] = [c* ]([K ]−1 )t [γ ] − [e* ]t [a][E] − [q* ]t [a][ H ]


[a][D] = [e* ]([K ]−1 )t [γ ] + [ε * ][a][E] + [α * ][a][ H ] (B19)
[a][ B] = [q* ]([K ]−1 )t [γ ] + [α * ]t [a][E ] + [ µ * ][a][ H ]

Or

[σ ] = [K ]−1 [c* ]([K ]−1 )t [γ ] − [K ]−1[e* ]t [a][E ] − [K ]−1[q* ]t [a][ H ]


[D] = [a]−1[e* ]([K ]−1 )t [γ ] + [a]−1[ε * ][a][E] + [a]−1[α * ][a][ H ] (B20)
[ B] = [a]−1[q* ]([K ]−1 )t [γ ] + [a]−1[α * ]t [a][ E] + [a]−1[ µ * ][a][ H ]

Comparing Eq. (B20) with the original constitutive relations (B8), we have

[c] = [K ]−1[c* ]([K ]−1 )t , [e]t = [K ]−1[e* ]t [a], [q]t = [K ]−1[q* ]t [a]
[e] = [a]−1[e* ]([K ]−1 )t , [ε] = [a]−1[ε* ][a], [α] = [a]−1[α* ][a] (B21)
[q] = [a]−1[q* ]([K ]−1 )t , [α]t = [a]−1 [α* ]t [a], [µ] = [a]−1 [µ* ][a]

Notice that for an orthogonal transformation, we have [a]t=[a]−1. Therefore, different


transformation relations for the same tensor are equivalent to each other.

2.15 Appendix C: Some Important Unit Relations

J
1V = 1
C
N V
1 =1 (C1)
C m
N/C A ⋅ m (C/s) ⋅ m m
= = =
N / ( A ⋅ m) C C s

The first equation means that 1J of work must be done to move 1C charge through
a potential difference of 1V.

A⋅s J W ⋅ s C C2 C2 s2 ⋅ C 2 s4 ⋅ A 2 s
F= = 2 = 2 = = = = 2 = 2 = (C2)
V V V V J N ⋅ m m ⋅ kg m ⋅ kg Ω

In both equations, V = volt, J = joule, C = coulomb, N = newton, m = meter, A =


ampere, s = second, F = farad, W = watt, kg = kilogram, Ω = ohm. Eq. (C2) is from
online Wikipedia.

2.16 References
Adachi, S. 1985. GaAs, AlAs, and AlxGa1-xAs: Material parameters for use in research and
device applications. Journal of Applied Physics 58: R1–R29.
1994. GaAs and Related Materials: Bulk Semiconducting and Superlattice Properties.
Singapore: World Scientific Publishing Co. Ltd.
2.16 References 55

Avellaneda, M. and Harshe, G. 1994. Magnetoelectric effect in piezoelectric/magnetostrictive


multilayer (2–2) composites. Journal of Intelligent Material Systems and Structures 5:
501–13.
Bacon, D. J., Barnett, D. M., and Scattergood, R. O. 1979. Anisotropic continuum theory of
lattice defects. Progress in Materials Science 23: 51–262.
Barnett, D. M. and Lothe, J. 1975. Dislocations and line charges in anisotropic piezoelectric
insulators. Physica Status Solidi B 67: 105–11.
Beya-Wakata, A., Prodhomme, P. Y., and Bester, G. 2011. First- and second-order
piezoelectricity in III-V semiconductors. Physical Review B 84: 195207.
Bichurin, M. I., Petrov,V. M., and Srinivasan, G. 2003.Theory of low-frequency magnetoelectric
coupling in magnetostrictive-piezoelectric bilayers. Physical Review B 68: 054402.
Caro, M. A., Schulz, S., and O’Reilly, E. P. 2013. Theory of local electric polarization and its
relation to internal strain: Impact on polarization potential and electronic properties of
group-III nitrides. Physical Review B 88: 214103.
Chou, Y. T. and Barnett, D. M. 1997a. Reciprocal relationship between the stress fields of a
line force and a dislocation in anisotropic media. Philosophical Magazine A 76: 221–9.
1997b. On a reciprocal relationship between the deformation fields of a line force and a
straight dislocation. Philosophical Magazine Letters 76: 139–41.
Ding, H. J. and Chen, W. Q. 2001. Three Dimensional Problems of Piezoelasticity. New York:
Nova Science Publishers.
Ding, H. J., Chen, W. Q., and Zhang, L. C. 2006. Elasticity of Transversely Isotropic Materials.
Dordrecht, The Netherlands: Springer.
Dunn, M. L. and Taya, M. 1993. An analysis of piezoelectric composite materials containing
ellipsoidal inhomogeneities. Proceedings of the Royal Society of London A 443:
265–87.
Eshelby, J. D. 1957. The determination of the elastic field of an ellipsoidal inclusion, and
related problems. Proceedings of the Royal Society of London A 241 (1226): 376–96.
1959. The elastic field outside an ellipsoidal inclusion. Proceedings of the Royal Society of
London A 252 (1271): 561–9.
1961. Elastic inclusions and inhomogeneities. In: Progress in Solid Mechanics (Eds.
Sneddon, I. N. and Hill, R.), Vol. 2, pp. 89–140. Amsterdam: North-Holland.
Green, G. 1828. An Essay on the Application of Mathematical Analysis to the Theories of
Electricity and Magnetism. Printed for the author by T. Wheelhouse, Nottingham, UK.
Hanada, T. 2009. Basic properties of ZnO, GaN and related materials. In: Oxide and Nitride
Semiconductors Processing, Properties, and Applications (Eds. Yao, T. and Hong, S. K.),
Vol. 12, pp. 1–19. Berlin: Springer.
Hirth, J. P. and Lothe, J. 1982. Theory of Dislocations. 2nd ed. New York: Wiley.
Jiang, X. and Pan, E. 2004. Exact solution for 2D polygonal inclusion problem in anisotropic
magnetoelectroelastic full-, half-, and bimaterial-planes. International Journal of Solids
and Structures 41: 4361–82.
Kuo, H. Y. and Pan, E. 2011. Effective magnetoelectric effect in multicoated circular fibrous
multiferroic composites. Journal of Applied Physics 109: 104901.
Kuo, H. Y. and Peng, C. Y. 2013. Magnetoelectricity in coated fibrous composites of
piezoelectric and piezomagnetic phases. International Journal of Engineering Science
62: 70–83.
Kuo, H. Y. and Wang, Y. L. 2012. Optimization of magnetoelectricity in multiferroic fibrous
composites. Mechanics of Materials 50: 88–99.
Levinshtein, M. E., Rumyantsev, S. L., and Shur, M. S. 2001. Properties of Advanced
Semiconductor Materials: GaN, AlN, InN, BN, SiC, SiGe. New York: John Wiley
& Sons.
Maruyama, T. 1964. Statical elastic dislocations in an infinite and semi-infinite medium.
Bulletin of the Earthquake Research Institute 42: 289–368.
1966. On two-dimensional elastic dislocations in an infinite and semi-infinite medium.
Bulletin of the Earthquake Research Institute 44: 811–71.
Mindlin, R. D. 1936. Force at a point in the interior of a semi-infinite solid. Physics 7:
195–202.
56 Governing Equations

Mura, T. 1987. Micromechanics of Defects in Solids. 2nd ed. Dordrecht, The Netherlands:
Kluwer Academic Publishers.
Nan, C. W., Bichurin, M. I., Dong, S. X., Viehland, D., and Srinivasan, G. 2008. Multiferroic
magnetoelectric composites: Historical perspective, status, and future directions. Journal
of Applied Physics 103: 031101.
Pan, E. 1989. Static response of a transversely isotropic and layered half-space to general
dislocation sources. Physics of the Earth and Planetary Interiors 58: 103–17.
1990. Thermoelastic deformation of a transversely isotropic and layered half-space by
surface loads and internal sources. Physics of the Earth and Planetary Interiors 60:
254–64.
1991. Dislocation in an infinite poroelastic medium. Acta Mechanica 87: 105–15.
1999a. A BEM analysis of fracture mechanics in 2-D anisotropic piezoelectric solids.
Engineering Analysis with Boundary Elements 23: 67–76.
1999b. Green’s functions in layered poroelastic half-spaces. International Journal for
Numerical and Analytical Methods in Geomechanics 23: 1631–53.
2001. Exact solution for simply supported and multilayered magneto-electro-elastic plates.
Journal of Applied Mechanics 68: 608–18.
2002. Three-dimensional Green’s functions in anisotropic magneto-electro-elastic
bimaterials. Zeitschrift für angewandte Mathematik und Physik 53: 815–38.
2004a. Eshelby problem of polygonal inclusions in anisotropic piezoelectric full- and half-
planes. Journal of the Mechanics and Physics of Solids 52: 567–89.
2004b. Eshelby problem of polygonal inclusions in anisotropic piezoelectric bimaterials.
Proceedings of the Royal Society of London A 460: 537–60.
Pan, E. and Yuan, F. G. 2000. Three-dimensional Green’s Functions in Anisotropic
Piezoelectric Bimaterials. International Journal of Engineering Science 38: 1939–60.
Petrov, V. M. and Srinivasan, G. 2008. Enhancement of magnetoelectric coupling in
functionally graded ferroelectric and ferromagnetic bilayers. Physical Review B 78:
184421.
Prodhomme, P. Y., Beya-Wakata, A., and Bester, G. 2013. Nonlinear piezoelectricity in
wurtzite semiconductors. Physical Review B 88: 121304 (R).
Schulz, S., Caro, M. A., O’Reilly, E. P., and Marquardt, O. 2012. Piezoelectric properties of
zinc blende quantum dots. Physica Status Solidi B 249: 521–5.
Sharapov, V. 2011. Piezoceramic Sensors. Berlin and Heidelberg: Springer-Verlag.
Singh, J. 2003. Electronic and Optoelectronic Properties of Semiconductor Structures. New
York: Cambridge University Press.
Steketee, J. A. 1958. On Volterra’s dislocation in a semi-infinite elastic medium. Canadian
Journal of Physics 36: 192–205.
Sun, L. G., Xu, K. Y., and Pan, E. 2012. Irregular inhomogeneities in an anisotropic
piezoelectric plane. Journal of Applied Mechanics 79: 021014.
Tiersten, H. F. 1969. Linear Piezoelectric Plate Vibrations. New York: Plenum.
Ting, T. C. T. 1996. Anisotropic Elasticity. New York: Oxford University Press.
Vurgaftman, I., Meyer, J. R., and Ram-Mohan, L. R. 2001. Band parameters for III-V
compound semiconductors and their alloys. Journal of Applied Physics 89: 5815–75.
Wood, C. and Jena, D. 2008. Polarization Effects in Semiconductors: From Ab Initio Theory
to Device Applications. New York: Springer Science+Business Media.
Xue, C. X. and Pan, E. 2013. On the longitudinal wave along a functionally graded magneto-
electro-elastic rod. International Journal of Engineering Science 62: 48–55.
Yang, J. S. 2006. Analysis of Piezoelectric Devices. Singapore: World Scientific.
Yu, H. Y., Sanday, S. C., and Chang, C. I. 1994. Elastic inclusions and inhomogeneities in
transversely isotropic solids. Proceedings of the Royal Society of London A 444:
239–52.
Yuan, F. G. and Wang, L. 2008. Vibration energy harvesting by magnetostrictive material.
Smart Materials and Structures 17: 045009.
3 Green’s Functions in Elastic Isotropic
Full and Bimaterial Planes

3.0 Introduction
Fundamental solutions in 2D isotropic elastic domains are generally solved based
on the complex variable method (or complex function method). Even though the
2D antiplane problem is usually described by the potential theory, the 2D potential
functions are either the real or imaginary part of an analytical function. In this
chapter, in terms of convenience, both the potential theory and complex variable
method will be applied to derive the Green’s functions in full, half, and bimaterial
planes. Presented are also the fundamental solutions for the circular inhomogeneity
in an infinite 2D matrix. Both the line-force and line-dislocation solutions will be
provided and the features of these two types of solutions will be discussed. When
solving the in-plane deformation problem, Muskhelishvili’s formulation based on
the complex variable method for plane elasticity (Muskhelishvili 1953) will be briefly
presented for easy reference.

3.1 Antiplane vs. Plane-Strain Deformation


We assume that the 2D linear isotropic elastic problem is within the (x,y)-plane.
Then, for the antiplane deformation, we have the following governing equations

σ xz, x + σ yz, y = − fz ( x ) (3.1)

σ xz = 2µγ xz , σ yz = 2µγ yz (3.2)

1 1
γ xz = 2 uz, x , γ yz = 2 uz, y (3.3)

Equation (3.1) is the stress (σxz, σyz) equilibrium equation in the z-direction (balance
the given line force fz), Eq. (3.2) is the constitutive relation between the stress and
strain (γxz, γyz), and Eq. (3.3) is the strain and displacement (uz) relation. In Eq. (3.2),
μ is the shear modulus.

57
58 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

From Eqs. (3.1) to (3.3), one can easily find that, in terms of the out-of-plane
displacement uz ≡ u, the antiplane deformation is governed by the 2D Poisson’s
equation as

fz ( x )
∇ 2u = − (3.4)
µ

with ∇ 2 = ∂ 2 / ∂x 2 + ∂ 2 / ∂y2 being the 2D Laplacian.


For a plane-strain problem of isotropic elasticity, the equilibrium equation,
constitutive relation, and strain-displacement relations can be expressed as
(Love 1944)

σ αβ ,α = − fβ ( x ) (3.5)

2 µν
σ αβ = δαβ γ ττ + 2 µγ αβ (3.6)
1 − 2ν
1
γ αβ = 2 (uα,β + uβ,α ) (3.7)

where Greek indexes are varying from 1 to 2, and the repeated ones take the
summation over 1 and 2. In terms of the in-plane elastic displacements u1 ≡ ux and
u2 ≡ uy, we have the following 2D plane-strain elasticity equations:

1 − fβ ( x )
uα ,αβ + uβ ,ττ = (3.8)
1 − 2ν µ

where fα (α = 1,2) are the line forces, and ν is the Poisson’s ratio.

3.2 Antiplane Solutions of Line Forces and Line Dislocations


Let us first assume that there is a concentrated line force with magnitude f applied
at x = 0, that is, fz(x) = fδ(x). Then Eq. (3.4) becomes

f δ ( x)
∇ 2u = − (3.9)
µ

The Green’s function solution of Eq. (3.9) is, according to Eq. (1.32),

f  1
u( x ) = ln   (3.10)
2πµ  r 

where r = | x | = x 2 + y2 . The corresponding Green’s stresses are

− fx − fy
σzx = , σzy = (3.11)
2 πr 2 2 πr 2

Remark 3.1: In isotropic elasticity, the two antiplane Green’s stresses are
independent of the shear modulus μ!
3.2 Antiplane Solutions of Line Forces and Line Dislocations 59

For the antiplane line dislocation (or screw dislocation) case, the antiplane
governing equation (3.4) will be free of the “body” source. In other words, the equation
governing the antiplane displacement u satisfies the 2D Laplace’s equation

∇2u = 0 (3.12)

However, at the location where the antiplane line dislocation, or the screw
dislocation, is applied, the antiplane displacement to be solved needs to satisfy
certain discontinuity conditions, besides being a solution of Eq. (3.12).
We assume that there is a uniformly distributed line (screw) dislocation with
Burgers vector b in the z-direction, acting along the entire negative x-axis. Then
the jump condition across the line dislocation is (in terms of the polar coordinates
r and θ)

u (r , π ) − u (r , − π ) = b (3.13)

We define the complex function ln(z) with z = x+iy as

ln r (θ = 0)
ln z =  (3.14)
ln r ± iπ (θ = ± π)

Then it can be shown that the following displacement is harmonic and also satisfies
the dislocation jump condition (3.13)

−b
u( x ) = Re[i ln z] (3.15)

Equation (3.15) can be simply written as


u( x ) = (3.16)

Taking the derivative of Eq. (3.15), we have

−b  i  −by
∂ x u( x ) = Re   =
2π  z  2 πr 2 (3.17)
−b  −1  bx
∂ yu( x ) = Re   =
2π  z 2 πr 2

Therefore, the stresses are

−µby µbx
σzx = , σzy = (3.18)
2 πr 2 2 πr 2

Comparing Eqs. (3.16) and (3.18) of the screw-dislocation solution with Eqs.
(3.10) and (3.11) of the out-of-plane line-force solution, we have the following
remarks.
Remark 3.2: While the displacement of the line force is the real part of ln(z), that of
the screw dislocation is the imaginary part of ln(z).
60 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

Remark 3.3: The displacement of the line force is inversely proportional to the shear
modulus μ, while that of the line dislocation is independent of the modulus.
Remark 3.4: The stresses of the line force are independent of the modulus, while
those of the screw dislocation are proportional to the modulus.
Remark 3.5: The stresses of both line force and line dislocation have the same order
of singularity.
Remark 3.6: These similarities between the solutions of the line force and line
dislocation are the representation of Betti’s reciprocal theorem discussed in Chapter 2.
In particular, if one has the stress solution for the line force, the corresponding line
dislocation solution can be found as

f µb f µb
σbzx = σzy , σbzy = − σzx (3.19)
f f

Or, if the screw dislocation solution is given, then the corresponding line-force
solution can be found as

f f f f
σzx = − σbzy , σzy = σbzx (3.20)
µb µb

Remark 3.7: In Eqs. (3.19) and (3.20), the superscripts “b” and “f” denote the Green’s
line-source solutions corresponding to the screw dislocation with magnitude b and
to the antiplane line force with magnitude f, respectively.
Remark 3.8: Should b and f be both of unit magnitude, the stresses of the line force
and line dislocation are connected by the shear modulus μ only.

3.3 Plane Displacements in Terms of Complex Functions


In this section, we briefly introduce the complex function method that Muskhelishvili
(1953) first used in solving various 2D elastic problems. It can be shown that for the
force-free case (i.e., the right-hand side of Eq. (3.8) is zero), the plane displacements
and stresses can be expressed in terms of complex functions in the z = x+iy complex
plane as

2µ(ux + iuy ) = κφ(z) − zφ ′(z) − ψ (z)


σ xx + σ yy = 2[φ ′(z) + φ ′(z)] (3.21)
σ yy − σ xx + 2 iσ xy = 2[z φ ′′(z) + ψ ′(z)]

where κ = (3 − 4ν ) (For plane stress elasticity, we have κ=(3 −ν ) / (1 + ν )). Then,


it can be shown that the resultant force of the traction exerted on an arc AB
(from the positive side of the normal n on the arc) is (notice that nx = dy/ds, and
ny = −dx/ds)
3.3 Plane Displacements in Terms of Complex Functions 61

B B
∫A (tx + it y ) ds = ∫A [(σ xx nx + σ yx ny ) + i(σ xy nx + σ yy ny )] ds (3.22)
= − i[φ(z) + zφ′(z) + ψ (z)]B
A

A bar over the whole function denotes complex conjugate. However, one should pay
attention to the overbar that applies only to the function, but not its variable. That is,
the overbar only extends over the function as defined

F (z) = F (z ) (3.23)

As an example, if F(z) is a rational function defined as

a0 zn + a1zn −1 +  + an
F (z) = (3.24)
b0 zm + b1zm −1 +  + bm

Then, we have

a0 zn + a1zn−1 +  + an
F (z) = (3.25)
b0 zm + b1zm−1 +  + bm

It can be shown that the conjugate of F(z) can be expressed as

F (z) = F (z ) (3.26)

An important application: It is easily observed that, if F(z) is holomorphic (=analytic)


in a given region S, then F (z) is holomorphic in the region S, obtained from S by
reflection with respect to the real axis.
+ −
Example: If function F(z) is defined in S (y > 0), we define a new region S (y < 0)
+
which is separated from S by the real axis x = t. Then, the function F (z) will be
− +
defined in the region S . Further, if the boundary value F (t) exists, then the boundary

value F (t ) also exists and that

F − (t ) = F + (t ) (3.27)

− +
Notice that if z → t from S , then, z → t from S . Similarly,


F + (t ) = F (t ) (3.28)

These simple relations are very useful when solving the problems in a bimaterial
domain (either an inhomogeneity in an infinite matrix or an infinite plane made of
two half-planes).
62 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

3.4 Plane-Strain Solutions of Line Forces and Line Dislocations

3.4.1 Plane-Strain Solutions of Line Forces


From Eq. (3.21) which was obtained by Muskhelishvili (1953), we notice that in
order to solve a plane-elasticity problem, one needs only to find the two complex
functions φ and ψ. For a line force (fx+ify) applied at the origin z = (x+iy) = 0, these
two functions can be assumed to be of the logarithmic type having a singularity at
the source point z = 0

φ (z) = A ln(z),ψ (z) = B ln(z) (3.29)

The complex constants A and B are determined by the following two conditions: (1)
The displacement should be single valued; and (2) The stress field should balance
the applied line force. The second condition requires that for any closed path of
integration C with z = 0 inside, we should have

∫ C [(σ xx nx + σ yx ny ) + i(σ xy nx + σ yy ny )] ds − ( fx + ify ) = 0 (3.30)

Using Eq. (3.21), the elastic displacements are found as

zA
2µ(ux + iuy ) = κ A ln(z) − − Bln(z) (3.31)
z

Because

ln(z) = ln(re iθ ) = ln(r ) + iθ (3.32)

the θ-term in Eq. (3.31) should be zero to meet the single-valuedness condition of
the displacement. This leads to

κA+ B = 0 (3.33)

Applying Eq. (3.30) to the unit circle centered at z = 0 with the stresses in terms of
the two complex functions given by Eq. (3.29), we find

zA

−i 
z =1
[ A ln(z) +
z
+ Bln(z)]ds = −( fx + if y ) (3.34)

This gives us

2 π( A − B) = −( fx + if y ) (3.35)

Solving Eqs. (3.33) and (3.35), we have the complex coefficients A and B as

1 κ
A= − ( fx + if y ), B= ( fx − if y ) (3.36)
2 π(κ + 1) 2 π(κ + 1)
3.4 Plane-Strain Solutions of Line Forces and Line Dislocations 63

Therefore, the two complex functions are

f x + if y κ ( f x − if y )
φ(z) = − ln(z), ψ (z) = ln(z) (3.37)
2 π(κ + 1) 2 π(κ + 1)

The displacements and stresses are found using Eq. (3.21)

2µ(ux + iuy ) = κ A ln(z) − Az / z − B ln(z )


σ xx + σ yy = 2( A /z + A / z ) (3.38)
σ yy − σ xx + 2 iσ xy = 2(− Az /z2 + B / z)

Or, making use of coefficients A and B, the displacement and stress fields are

(3 − 4 ν)( f x + if y ) f x − if y
z
ux + iu y = − [ln(z) + ln(z)] +
16 πµ(1 − ν) 16 πµ(1 − ν) z
−1
σ xx + σ yy = [( fx + if y ) / z + ( fx − if y ) / z ] (3.39)
4 π(1 − ν)
1
σ yy − σ xx + 2 iσ xy = [( fx + if y )z / z + (3 − 4 ν)( fx − if y )]
4π(1 − ν)z

3.4.2 Plane-Strain Solutions of Line Dislocations


We assume that a line dislocation (edge dislocation) with components (bx, by) is
applied at z = 0. The two complex functions are also assumed to be of logarithmic-
type as in Eq. (3.29), as for the line-force case. For the dislocation case, however,
the resultant force is zero while the displacement field should satisfy the following
discontinuity condition

(ux + iuy )θ = π − (ux + iuy )θ = − π = bx + iby (3.40)

Therefore, the force equilibrium condition (3.30) gives us

( A − B) = 0 (3.41)

while condition (3.40) gives

iµ(bx + iby )
κA + B = (3.42)
π

We can thus solve for the complex constants A and B

iµ iµ
A= − (bx + iby ), B= (bx − iby ) (3.43)
π(κ + 1) π(κ + 1)
64 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

Thus, in terms of the new coefficients A and B in Eq. (3.43) for the line-dislocation
case, the displacement and stress fields can be also represented by Eq. (3.38) as for
the line-force case.
Substituting the coefficients A and B in Eq. (3.43) into Eq. (3.38), one can find
the displacements and stresses due to a line dislocation as

i(bx + iby ) i(bx − iby ) z


ux + i u y = [ −κ ln(z) + ln(z)] −
2 π(κ + 1) 2 π(κ + 1) z
2 iµ
σ xx + σ yy = [ −(bx + iby ) / z + (bx − iby ) / z ] (3.44)
π(κ + 1)
2 iµ
σ yy − σ xx + 2 iσ xy = [(bx + iby )z / z + (bx − iby )]
π(κ + 1)z

Comparing Eq. (3.44) of the line-dislocation solution with Eq. (3.39) of the line-
force solution, we have the following remarks.
Remark 3.9: These two sets of expressions are very similar to each other and
the singularities in displacements are the same and those in the stresses are also
the same.
Remark 3.10: For both line-force and line-dislocation solutions, the Green’s
displacements are even functions of the variables and the stresses are odd functions
of the variables.
Remark 3.11: These Green’s functions are for the line sources at z = 0; for the general
case where the source is at z = zs, one only needs to replace z by z – zs (and also z
by z − zs ).
Remark 3.12: While the Green’s displacements and stresses for a concentrated
line source at z = zs may be obtained according to the preceding simple coordinate
transformation, that for the two complex potentials is a little more complicated
(England 1971). In fact, we have

φ (z) = φs (z − zs ), ψ (z) = ψ s (z − zs ) − zsφs′ (z − zs ) (3.45)

where φs and ψ s are the complex potentials when the origin of the coordinates is set
at z = zs.
We point out that the line-force and line-dislocation induced fields can be
expressed in real form, instead of the complex form in Eqs. (3.39) and (3.44). For
example, letting by = 0, the displacements and stresses due to the edge dislocation bx
can be reduced to the following simple real form (r2 = x2+y2):

bx  −1 y xy 
ux = tan + 
2π  x 21 ν
( − )r 2
 (3.46)
bx  1 − 2ν x 2 − y2 
uy = −  ln r 2 + 
2π  4(1 −ν ) 4(1 −ν )r 2 
3.5 Bimaterial Antiplane Solutions of a Line Force and a Line Dislocation 65

µ bx y(3 x 2 + y2 )
σ xx = −
2π (1 −ν )r 4
µ b y( x 2 − y2 )
σ yy = x (3.47)
2π (1 −ν )r 4
µ b x( x 2 − y2 )
σ xy = x
2π (1 −ν )r 4

Or, in terms of the polar coordinates (x = r cosθ, y = r sin θ), they are


bx cos θ sin θ 
ux = θ + 

 2(1 −ν ) 
(3.48)
b  1 − 2ν cos2 θ − sin 2 θ 
uy = − x  ln r 2 + 
2π  4(1 −ν ) 4(1 −ν ) 

µbx sin θ
σ xx = − (3 cos2 θ + sin 2 θ)
2π(1 − ν)r
µbx sin θ
σ yy = (cos2 θ − sin 2 θ) (3.49)
2π(1 − ν)r
µbx cos θ
σ xy = (cos2 θ − sin 2 θ)
2 π(1 − ν)r

Remark 3.13: Similar real expressions of the displacements and stresses under the
edge dislocation by can be also obtained.

3.5 Bimaterial Antiplane Solutions of a Line Force and


a Line Dislocation

3.5.1 Bimaterial Antiplane Solutions of a Line Force


We let the upper half-plane (y > 0) be Material 1 with shear modulus μ1 and the lower
half-plane (y < 0) be Material 2 with shear modulus μ2. The antiplane displacement
u (≡uz) in Materials 1 and 2 are denoted by u(1) and u(2), respectively. We further
assume that an antiplane line force of unit magnitude is applied at (x,y) = (0,d>0) in
Material 1. Therefore, the antiplane governing equations in terms of the antiplane
displacements in Materials 1 and 2 are

δ ( x)δ ( y − d)
∇ 2 u(1) = −
µ1 (3.50)
∇ 2 u( 2) =0

The Green’s function solution u(1) in Material 1 can be expressed as

1 1 c 1
u(1) ( x ) = ln   + 1 ln   (3.51)
2πµ1  r1  2πµ1  r2 
66 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

where

r1 = x 2 + ( y − d)2 , r2 = x 2 + ( y + d)2 (3.52)

The corresponding stresses are

x c x y − d c1 ( y + d)
σ(zx1) = − − 1 , σ(zy1) = − − (3.53)
2 πr12 2 πr22 2 πr12 2 πr22

In Material 2, the displacement and stresses can be expressed as

c2 1
u(2) ( x ) = ln   (3.54)
2πµ2  r1 

c2 x c2 ( y − d)
σ(zx2) = − , σ(zy2) = − (3.55)
2 πr12 2 πr12

The two constants c1 and c2 are determined by the continuity conditions of the
displacement u and stress σzy on the interface y = 0, which result in the following
two relations:

c2 µ 1
1 + c1 = , 1 − c1 = c2 (3.56)
µ2

Solving the two constants in Eq. (3.56), we have

1− Γ 2Γ
c1 = , c2 = (3.57)
1+ Γ 1+ Γ

where Γ = µ 2 / µ1 is the shear modulus ratio of the bimaterial. Thus, the two-point
Green’s functions of the displacement and stress fields in both the upper and lower
half-planes are found (with the source (0,d) being replaced by the general location
(xs,ys) with ys > 0) in the following.
In Material 1, that is, the upper (source) half-plane (y > 0):

1 1 1− Γ 1
u(1) ( x, y; xs , ys ) = ln   + ln  
2πµ1  r1  2π (1 + Γ )µ1  r2 
1 x − xs 1− Γ x − xs
σ z(1x) ( x, y; xs , ys ) = − − (3.58)
2π r1 2 2π (1 + Γ ) r22
(1) 1 y − ys 1− Γ y + ys
σ zy ( x, y; xs , ys ) = − −
2π r1 2 2π (1 + Γ ) r22

In Material 2, that is, the lower half-plane (y < 0):


3.5 Bimaterial Antiplane Solutions of a Line Force and a Line Dislocation 67

1 1
u(2) ( x, y; xs , ys ) = ln  
π (1 + Γ )µ1  r1 
( 2) Γ x − xs
σ zx ( x, y; xs , ys ) = − (3.59)
π (1 + Γ ) r12
(2) Γ y − ys
σ zy ( x, y; xs , ys ) = −
π (1 + Γ ) r12

where

r1 = ( x − xs )2 + ( y − ys )2 , r2 = ( x − xs )2 + ( y + ys )2 (3.60)

Remark 3.14: The solutions to the case in which a line force is located in Material
2, that is, in the lower half-plane, can be found from the preceding expressions by
simply exchanging the superscripts and subscripts between 1 and 2.
Remark 3.15: These solutions reduce to the infinite plane solution if μ1 = μ2 (i.e.,
Г = 1).
Remark 3.16: For the upper half-plane under the traction-free surface condition,
the solution can be reduced from the preceding bimaterial solution by letting μ2 = 0.
Therefore, we have

1 1 1 1
u(1) ( x, y; xs , ys ) = ln   + ln  
2πµ1  r1  2πµ1  r2 
(1) x − xs x − xs
σ zx ( x, y; xs , ys ) = − − (3.61)
2π r12 2π r22
(1) y − ys y + ys
σ zy ( x, y; xs , ys ) = − −
2π r12 2π r22

Remark 3.17: For the upper half-plane under the rigid displacement condition (u =
0) on y = 0, the solution can also be reduced from the preceding bimaterial solution
by letting μ2 = ∞. Therefore, we have

1  1 1  1
u(1) ( x, y; xs , ys ) = ln − ln
2 πµ1  r1  2 πµ1  r2 
x − xs x − xs
σ(zx1) ( x, y; xs , ys ) = − + (3.62)
2 πr12 2 πr22
y − ys y + ys
σ(zy1) ( x, y; xs , ys ) = − +
2 πr12 2 πr22

3.5.2 Bimaterial Antiplane Solutions of a Line Dislocation


We assume that there is a uniformly distributed line dislocation (i.e., screw
dislocation) with Burgers vector b in the z-direction located at zs (= xs + iys with ys > 0)
68 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

in the upper half-plane (y > 0). We introduce the local polar coordinates (r,θ)
originated at zs, where r is the distance between z and zs, and θ is the angle from
the horizontal axis with the origin at zs. Then the jump condition across the line
dislocation source at zs is given by Eq. (3.13). We also recall the definition of
the complex function ln(z−zs) with z = x+iy. Then in Material 1, the antiplane
displacement can be expressed as

−b c
u(1) (z; zs ) = Re[i ln(z − zs )] + 1 Re[i ln(z − zs )] (3.63)
2π 2π

with c1 being a constant to be determined. From Eq. (3.63), we have

b( y − ys ) c1 ( y + ys )
∂ x u(1) ( x, y; xs , ys ) = − +
2π r12 2π r22
(3.64)
b( x − xs ) c1 ( x − xs )
∂ y u(1) ( x, y; xs , ys ) = −
2π r12 2π r22

with r1 and r2 being the distances defined in Eq. (3.60). The stresses are

(1) µ1b( y − ys ) µ1c1 ( y + ys )


σ zx ( x, y; xs , ys ) = − +
2π r12 2π r22
(3.65)
(1) µ b( x − xs ) µ1c1 ( x − xs )
σ zy ( x, y; xs , ys ) = 1 −
2π r12 2π r22

In Material 2 (the lower half-plane), the displacement and stresses can be


expressed as

c2
u( 2) (z; zs ) = Re[i ln(z − zs )] (3.66)

(2) µ2 c2 ( y − ys )
σ zx ( x, y; xs , ys ) =
2π r12
(3.67)
(2) µ c (x − x )
σ zy ( x, y; xs , ys ) = − 2 2 2 s
2π r1

where c2 is another constant. These two constants can be determined by the


continuity conditions of the displacement u and traction σzy on the interface y = 0,
which require

−b − c1 = c2 ; b − c1 = −c2 µ2 / µ1 (3.68)

These help us solve the two constants as

b(1 − Γ ) 2b
c1 = ; c2 = − (3.69)
1+ Γ 1+ Γ
3.5 Bimaterial Antiplane Solutions of a Line Force and a Line Dislocation 69

Therefore, the screw-dislocation solutions in Materials 1 and 2 can be finally found.


In Material 1, the displacement is

b b(1 − Γ )
u(1) (z; zs ) = − Re[i ln(z − zs )] + Re[i ln(z − zs )] (3.70)
2π 2 π(1 + Γ )

and the stresses are

(1) µ1b y − ys µ1b(1 − Γ ) y + ys


σ zx ( x, y; xs , ys ) = − +
2π r12 2π (1 + Γ ) r22
(3.71)
(1) µ b x − xs µ1b(1 − Γ ) x − xs
σ zy ( x, y; xs , ys ) = 1 −
2π r12 2π (1 + Γ ) r22

In Material 2, the displacement is

b
u( 2) (z; zs ) = − Re[ i ln(z − zs )] (3.72)
π(1 + Γ )

and the stresses are

(2) µ2 b y − ys
σ zx ( x, y; xs , ys ) = −
π (1 + Γ ) r12
(3.73)
(2) µ 2 b x − xs
σ zy ( x, y; xs , ys ) =
π (1 + Γ ) r12

Remark 3.18: Solutions for the screw dislocation in the lower half-plane can be found
from the solutions presented by simple exchange of superscripts and subscripts
between 1 and 2. One should remember also that ys becomes negative.
Remark 3.19: These solutions reduce to the infinite plane solution if μ1 = μ2.
Remark 3.20: For the upper half-plane (Material 1 with y > 0) under the traction-
free surface condition, the solution can be reduced from the preceding expressions
by letting μ2 = 0. Therefore, we have

b b
u(z; zs ) = − Re[ i ln(z − zs )] + Re[ i ln(z − zs )]
2π 2π
µb y − ys µb y + ys
σzx ( x, y; xs , ys ) = − + (3.74)
2 π r12 2 π r22
µb x − x s µb x − x s
σ zy ( x, y; xs , ys ) = −
2 π r12 2 π r22

Remark 3.21: For the upper half-plane under rigid displacement condition on the
surface, the solution can be reduced from the preceding expressions by letting
μ2 = ∞. Therefore, we have
70 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

b b
u(z; zs ) = − Re[i ln(z − zs )] − Re[i ln(z − zs )]
2π 2π
µb y − ys µb y + ys
σzx ( x, y; xs , ys ) = − − (3.75)
2 π r12 2 π r22
µb x − x s µb x − x s
σ zy ( x, y; xs , ys ) = +
2 π r12 2 π r22

3.6 Bimaterial Plane-Strain Solutions of Line Forces


and Line Dislocations
For the bimaterial plane-strain problem with a line force or a line dislocation (edge
dislocation) applied at zs = xs + iys with ys > 0 in Material 1 (y > 0), we assume the
complex functions in Eq. (3.21) as follows.
In Material 1:

φ1 (z) = φ1 p (z) + φ1c (z)


(3.76)
ψ 1 (z) = ψ 1 p (z) + ψ 1c (z)

where the subscript p denotes the given “particular” solution in the infinite plane with
material property of Material 1, with the source location being at zs. The subscript
c denotes the complementary part which is used to satisfy the interface condition
at y = 0. It is assumed that φ1c (z) and ψ 1c (z) are analytic for y > 0, while φ1p (z) and
ψ 1p (z) are obviously analytic for y < 0. Thus, according to Section 3.3, φ1c (z) and
ψ 1c (z) will be analytic for y < 0, while φ1p (z) and ψ 1p (z) be analytic for y > 0.
In Material 2, the two complex functions are assumed to be φ2(z) and ψ2(z), both
analytic for y < 0. Accordingly, φ2 (z) and ψ 2 (z) will be analytic for y > 0.
The remaining four complex functions (φ1c, ψ1c, φ2 and ψ2) are determined by
the interface conditions at z = x (y = 0), which require that the displacements and
resultant forces from both sides of the half-plane materials should be continuous.
In other words, when z = x on the interface, the following relations should hold (for
the resultant force, the directions should be opposite along the interface from both
material domains):

Γ {κ 1φ1 ( x) − [ xφ1′ ( x) +ψ 1 ( x)]}+ = {κ 2φ2 ( x) − [ xφ2′ ( x) +ψ 2 ( x)]}−


(3.77)
[φ1 ( x) + xφ1′ ( x) + ψ 1 ( x)]+ = [φ2 ( x) + xφ2′ ( x) + ψ 2 ( x)]−

where κ i = 3 − 4ν i , the superscripts “+” and “–” indicate the limiting value of the
functions when approaching y = 0 from y > 0 and y < 0, respectively, that is, from the
upper and lower half-planes, respectively.
Substituting the complex functions Eq. (3.76) in Material 1 to Eq. (3.77),
we have

Γκ 1[φ1 p ( x)]+ + Γκ 1[φ1c ( x)]+ − Γ [ xφ1′ p ( x) + ψ 1 p ( x)]+ − Γ [ xφ1′c ( x) + ψ 1c ( x)]+


= κ 2 [φ2 ( x)]− − [ xφ2′ ( x) + ψ 2 ( x)]−
(3.78)
[φ1 p ( x)]+ + [φ1c ( x)]+ + [ xφ1′ p ( x) +ψ 1 p ( x)]+ + [ xφ1′c ( x) +ψ 1c ( x)]+
= [φ2 ( x)]− + [ xφ2′ ( x) + ψ 2 (x)]−
3.6 Bimaterial Plane-Strain Solutions of Line Forces and Line Dislocations 71

which can be reordered as

Γκ 1[φ1c ( x)]+ − Γ [ xφ1′ p ( x) + ψ 1 p ( x)]+ + [ xφ2′ ( x) + ψ 2 ( x)]−


= κ 2 [φ2 ( x)]− − Γκ 1[φ1 p ( x)]+ + Γ [ xφ1′c ( x) + ψ 1c ( x)]+
(3.79)
[φ1c ( x)]+ + [ xφ1′ p ( x) + ψ 1 p ( x)]+ − [ xφ2′ ( x) + ψ 2 ( x)]−
= [φ2 ( x)]− − [φ1 p ( x)]+ − [ xφ1′c ( x) + ψ 1c ( x)]+

The preceding equations can be further written as

{
lim Γκ 1φ1c (z) − Γ [zφ1′ p (z) +ψ 1 p (z)] + [zφ2′ (z) + ψ 2 (z)]
y→ 0 +
}
{
= lim κ 2φ2 (z) − Γκ 1φ1 p (z) + Γ [zφ1′c (z) + ψ 1c (z)]
y→0 −
}
(3.80)
{
lim φ1c (z) + [zφ1′ p (z) + ψ 1 p (z)] − [zφ2′ (z) + ψ 2 (z)]
y→ 0 +
}
{
= lim φ2 (z) − φ1 p ( x) − [zφ1′c (z) + ψ 1c (z)]
y→0 −
}
In these two equations, the left-hand side is analytic in Material 1 (y > 0) and the right-
hand side is analytic in Material 2 (y < 0). Therefore, from the analytic continuation
theory, we have, in Material 1 (y > 0)

Γκ 1φ1c (z) − Γ [zφ1′ p (z) + ψ 1 p (z)] + [zφ2′ (z) + ψ 2 (z)] = 0


(3.81)
φ1c (z) + [zφ1′ p (z) + ψ 1 p (z)] − [zφ2′ (z) + ψ 2 (z)] = 0

and in Material 2 (y < 0)

κ 2φ2 (z) − Γκ 1φ1 p (z) + Γ [zφ1′c (z) + ψ 1c (z)] = 0


(3.82)
φ2 (z) − φ1 p (z) − [zφ1′c (z) + ψ 1c (z)] = 0

Thus, from Eq. (3.81), we have

Γ −1 α −β
φ1c (z) = [zφ1′ p (z) + ψ 1 p (z)] = [zφ1′ p (z) + ψ 1 p (z)]
Γκ 1 + 1 1+ β
(3.83)
Γ(κ 1 + 1) 1+ α
zφ 2′ (z) + ψ 2 (z) = [zφ1′ p (z) + ψ 1 p (z)] = [zφ1′ p (z) + ψ 1p (z)]
Γκ 1 + 1 1+ β

and from Eq. (3.82), we have

Γ(κ 1 + 1) 1+ α
φ2 (z) = φ1 p (z) = φ1 p (z)
κ2 + Γ 1−β
(3.84)
Γκ 1 − κ 2 α+β
zφ1′ c (z) + ψ 1c (z) = φ1 p (z) = φ1 p (z)
κ2 + Γ 1−β
72 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

In the preceding equations, we have introduced the Dundurs parameters (Dundurs


1968) for two-phase plane elasticity, α and β , as follows

Γ (κ 1 + 1) − (κ 2 + 1) Γ (κ 1 − 1) − (κ 2 − 1)
α= , β= (3.85)
κ
Γ ( 1 1) ( 2 1)
+ + κ + Γ (κ 1 + 1) + (κ 2 + 1)

Therefore, the four unknown complex functions are finally solved from Eqs. (3.83)
and (3.84) as

α −β
φ1c (z) = [zφ1′ p (z) + ψ 1 p (z)]
1+ β
α+β
ψ 1c (z) = φ1 p (z) − zφ1′ c (z)
1−β
α+β α −β
= φ1 p (z) − [zφ1′ p (z) + z2 φ1′′p (z) + zψ 1′ p (z)] (3.86)
1−β 1+ β
1+ α
φ 2 (z) = φ1 p (z)
1−β
2(1 + α)β 1+ α
ψ 2 (z) = − zφ1′ p (z) + ψ 1 p (z)
(1 + β)(1 − β) 1+ β

It is worth mentioning that the preceding relations are universal in the sense that
they are valid for any kind of concentrated source (e.g., concentrated force, edge
dislocation, concentrated moment, or quantum wire with certain eigenstrain). It is
easy to show that these results are the same as those derived by Suo (1989) using a
slightly different formula.
Notice that the particular solutions of the complex functions for the line force
and line dislocation are located at z = zs, instead of z = 0. Then, in view of Eq. (3.45),
the complex potentials of the particular solutions should read as

zs
φ1 p (z) = A ln(z − zs ), ψ 1 p (z) = B ln(z − zs ) − A (3.87)
z − zs

where the complex coefficients are given in the following.


For the line force (fx, fy) at z = zs

f x + if y κ 1 ( f x − if y )
A= − , B= (3.88a)
2 π(1 + κ 1 ) 2 π(1 + κ 1 )

For the line dislocation (bx,by) at z = zs

iµ 1 (bx + iby ) iµ 1 (bx − iby )


A= − , B= (3.88b)
π(1 + κ 1 ) π(1 + κ 1 )
3.6 Bimaterial Plane-Strain Solutions of Line Forces and Line Dislocations 73

In terms of the particular solutions (3.87), the four complex functions in Eq. (3.86)
can be finally expressed as

α − β  z − zs 
φ1c (z) = A + B ln(z − zs )
1 + β  z − zs 
(α + β) A α −β A+ B z − zs 
ψ 1c (z) = ln(z − zs ) − z −A 
1−β 1 + β  z − zs (z − zs )2 
(3.89)
(1 + α) A
φ2 (z) = ln(z − zs )
1− β
2(1 + α)β A z 1+ α  z 
ψ 2 (z) = − +  B ln(z − zs ) − A s 
1−β 2 z − zs 1 + β  z − zs 

Remark 3.22: If μ1 = μ2 and ν1 = ν2, then α = β = 0 and the complementary complex


functions in Material 1 are zero while those in Material 2 are reduced to

φ2 (z) = A ln(z − zs )
zs (3.90)
ψ 2 (z) = B ln(z − zs ) − A
z − zs

These are identical to the particular solutions (3.87). Thus, when substituting these
complex functions into Eq. (3.21), the displacement and stress fields obtained will be
the same as those in the infinite plane with the line source located at z = zs.
Remark 3.23: If μ2 = 0 (α = −1), we will have the traction-free half-plane solution in
Material 1 (y > 0), as

 z − zs 
φ1c (z) = −  A + B ln(z − zs )
 z − zs 
(3.91)
A+ B z − zs 
ψ 1c (z) = − A ln(z − zs ) + z  −A 
 z − zs (z − zs )2 

Remark 3.24: If μ2 = ∞ (α = 1, β = (κ − 1) / (κ + 1)), we will have the solution in the


half-plane of Material 1 (y > 0) bonded by a rigid surface, as

1  z − zs 
φ1c (z) = A + B ln(z − zs )
κ  z − zs 
(3.92)
1 A+ B z − zs 
ψ 1c (z) = κ A ln(z − zs ) − z  −A 
κ  z − zs (z − zs )2 

We now derive the displacement and stress fields in the bimaterial plane due to
either a line force or line dislocation using the complex functions in Eq. (3.89) along
with the complex functions of the particular solution at z = zs. In order to do so, we
first write Eq. (3.89) as
74 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

z − zs
φ1c (z) = c1 A + c1 B ln(z − zs )
z − zs
z z(z − zs )
ψ 1c (z) = c2 A ln(z − zs ) − c1 ( A + B) + c1 A
z − zs (z − zs )2 (3.93)
φ2 (z) = c3 A ln(z − zs )
z z
ψ 2 (z) = c5 A + c4 B ln(z − zs ) − c4 A s
z − zs z − zs

where the five real constants ci (i = 1–5) are defined as

α −β µ − µ1 α + β κ 1µ 2 − κ 2 µ 1
c1 = = 2 , c2 = = ,
1 + β κ 1µ 2 + µ 1 1−β κ 2 µ1 + µ 2
1+ α µ (κ + 1) 1+ α µ (κ + 1)
c3 = = 1 + c2 = 2 1 , c4 = = 1 + c1 = 2 1 , (3.94)
1−β κ 2 µ1 + µ 2 1+ β κ 1µ 2 + µ 1
2(1 + α)β µ (κ − 1) + µ 2 (1 − κ 1 )
c5 = − = c4 − c3 = c1 − c2 = µ 2 (κ 1 + 1) 1 2
1 − β2 (κ 1µ 2 + µ 1 )(κ 2 µ 1 + µ 2 )

First, in Material 1 where the source is located, the total solution is the summation
of the particular solution and the complementary part. For the particular solution,
we have,

[2µ1 (ux + iuy )]1 p = κ 1 A ln(z − zs ) − A(z − zs ) / (z − zs ) − B ln(z − zs )


[σ xx + σ yy ]1 p / 2 = A / (z − zs ) + A / (z − zs ) (3.95)
[σ yy − σ xx + 2iσ xy ]1 p / 2 = − A(z − zs ) / (z − zs )2 + B / (z − zs )

Using Eqs. (3.21) and (3.93), the complementary part of the solution in Material 1
is found to be

 z − zs 
[2µ1 (ux + iuy )]1c = κ 1 c1 A + c1 B ln(z − zs )
 z − zs  (3.96a)
z−z (z − z )(z − zs )
− c1 ( A + B) − c2 A ln(z − zs ) + c1 A
z − zs (z − zs )2

A+ B z − zs A+ B z − zs 
[σ xx + σ yy ]1c / 2 = c1  −A 2
+ −A 
 z − zs (z − zs ) z − zs (z − zs )2 
 B z − zs  zs
[σ yy − σ xx + 2iσ xy ]1c / 2 = c1z  − 2
+ 2A 3
+ c1 B (3.96b)
 ( z − zs ) ( z − z )
s  ( z − zs )2
(c2 − c1 ) A (z − zs ) z(z − zs )
+ + c1 A + 2c1 A s
z − zs (z − zs ) 2 (z − zs )3

The total displacement and stress fields in Material 1 (in the upper source half-plane
y > 0) are the summation of those in Eqs. (3.95) and (3.96).
In Material 2 (the lower half-plane without the source y < 0), the displacement
and stress fields are found by using Eqs. (3.21) and Eq. (3.93). The results are
3.6 Bimaterial Plane-Strain Solutions of Line Forces and Line Dislocations 75

2µ 2 (ux + iuy ) = κ 2 c3 A ln(z − zs ) − c4 B ln(z − zs )


z−z z − zs (3.97a)
− c3 A − c4 A
z − zs z − zs

 A A 
(σ xx + σ yy ) / 2 = c3  +
 z − zs z − zs 
(3.97b)
z−z z − zs c A + c4 B
(σ yy − σ xx + 2 iσ xy ) / 2 = c3 A − c4 A + 5
(z − zs )2 (z − zs )2 z − zs

Remark 3.25: If μ1 = μ2 = μ and ν1 = ν2 = ν, then c1 = c2 = 0. Thus, the complementary


parts of the displacement and stress fields in Material 1 given by Eq. (3.96) are zero,
and the total solution is reduced to the particular solution given by Eq. (3.95), which
gives the Green’s displacements and stresses for an infinite homogeneous plane
subjected to a concentrated source at z = zs.
Remark 3.26: Also, if μ1 = μ2 = μ and ν1 = ν2 = ν, then c3 = c4 = 1 and c5 = 0. The
displacement and stress fields in Material 2 given by Eq. (3.97) are reduced to the
following expressions

z − zs
2µ(ux + iuy ) = κ A ln(z − zs ) − A − B ln(z − zs ) (3.98a)
z − zs

 A A 
(σ xx + σ yy ) / 2 =  +
 z − zs z − zs 
(3.98b)
z − zs B
(σ yy − σ xx + 2 iσ xy ) / 2 = − A 2
+
(z − zs ) z − zs

which coincide with Eq. (3.95), that is, the solution of the displacement and stress
fields in an infinite plane with the source at z = zs.
Remark 3.27: If μ2 = 0 (μ1 = μ, ν1 = ν and κ1 = κ) we will have the traction-free half-
plane solution in Material 1 (y > 0). For this case, the constants c1 = c2 = –1, and the
complementary solution (3.96) is reduced to

 z − zs 
[2µ(ux + iuy )]1c = −κ  A + B ln(z − zs )
 z − zs  (3.99a)
z−z (z − z )(z − zs )
+ ( A + B) −A + A ln(z − zs )
z − zs (z − zs )2

A+ B z − zs A+ B z − zs 
[σ xx + σ yy ]1c / 2 = −  −A 2
+ −A 
 z − zs (z − zs ) z − zs (z − zs )2 
 B z − zs  Bzs
[σ yy − σ xx + 2 iσ xy ]1c / 2 = z  2
− 2A − (3.99b)
 (z − zs ) (z − zs )  (z − zs )2
3

(z − zs ) z(z − zs )
−A 2
− 2A s
(z − zs ) (z − zs )3
76 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

Remark 3.28: If μ2 = ∞ (μ1 = μ, ν1 = ν and κ1 = κ), we will have the solution in the
half-plane Material 1 (y > 0) bonded by a rigid surface. For this case, the constants
c1 = 1/κ, c2 = κ, and the complementary solution (3.96) is reduced to

z − zs
[2µ(ux + iuy )]1c = A + B ln(z − zs )
z − zs
(3.100a)
A (z − z )(zs − zs ) B z − z
+ − − κ A ln(z − zs )
κ (z − zs )2 κ z − zs

1 A+ B z − zs A+ B z − zs 
[σ xx + σ yy ]1c / 2 =  −A + −A 
κ  z − zs (z − zs ) 2 z − zs (z − zs )2 
z B z − zs  B zs
   [σ yy − σ xx + 2 iσ xy ]1c / 2 = − + 2A + (3.100b)
κ  (z − zs )2 (z − zs )3  κ (z − zs )2
1 (κ 2 − 1) A A (z − zs ) 2 A z(zs − zs )
+ + +
κ z − zs κ (z − zs )2 κ (z − zs )3

3.7 Line Forces or Line Dislocations Interacting with a


Circular Inhomogeneity

3.7.1 Antiplane Solutions


3.7.1.1 A Line Force Inside or Outside a Circular Inhomogeneity
We first assume that a line force of unit magnitude is applied at z = zs (|zs| < a)
within the circle of radius a. The circular inhomogeneity has a shear modulus μI
(≡μ1) and is perfectly connected to the matrix material with shear modulus μM (≡μ2).
The antiplane displacement u (≡uz) in the circular inhomogeneity (Material 1) and
the matrix (Material 2) are denoted by u(1) and u(2), respectively. Therefore, the
antiplane governing equations in terms of the antiplane displacements in Materials
1 and 2 are

∇2 u(1) = −δ(z − zs ) / µ 1 ( z < a & zs < a)


(3.101)
∇ 2 u( 2 ) =0 ( z > a)
where z = x + i y = r ei θ and zs = xs + i ys = rs ei θ s . Besides, the following continuity
conditions should be satisfied on the circle |z| = a:

(1) ( 2)
u(1) = u( 2) , σ zr = σ zr (3.102)

(i )
where σ zr = µi ∂u(i ) / ∂r (i = 1 or 2) is the shear stress. The problem is very similar to
that treated in Section 1.3.5.2, and therefore the two-phase Green’s function solution
may be obtained as

u(i ) (z) = Re[F (i ) (z)] (3.103)

with
3.7 Line Forces or Line Dislocations Interacting with a Circular Inhomogeneity 77

1 (µ 1 − µ 2 )  a2 a
F (1) (z) = − ln(z − zs ) −  ln(z − ) − ln  ( z < a)
2 πµ 1 2 πµ 1 (µ 1 + µ 2 )  zs zs  (3.104)
1 (µ 1 − µ 2 ) z
F ( 2) (z) = − ln(z − zs ) − ln ( z > a)
π(µ 1 + µ 2 ) 2πµ 2 (µ 1 + µ 2 ) a

The displacements in real form are

1 K   a
u(1) = − ln r1 +  ln r2 − ln    ( r < a)
2 πµ1 2 πµ1   rs  
(3.105)
1− K K r
u( 2) =− ln r1 + ln (r > a)
2 πµ1 2 πµ 2 a

where K = (Γ − 1) / (Γ + 1) is the Dundurs parameter for two-phase antiplane


elasticity (Dundurs 1968), and

r1 = z − zs = r 2 + rs2 − 2rrs cos(θ − θ s ),


a4 a2 (3.106)
r2 = z − a 2 / zs = r 2 + 2
− 2r cos(θ − θ s )
rs rs

are, respectively, the distance between the field point z and the source point zs
and that between the field point z and the image source point a 2 / zs (Figure 3.1).
Note that, when z is on the circle |z|=a, we have r2 = ar1 / rs (also see Eq. (1.11)). The
corresponding stresses are

(1) 1 r − rs cos(θ − θ s ) K r − (a 2 / rs )cos(θ − θ s )


σ zr =− +
2π r12 2π r22
1 rs sin(θ − θ s ) K (a 2 / rs )sin(θ − θ s )
σ z(θ1) = − +
2π r12 2π r22
(3.107)
( 2) 1 − K r − rs cos(θ − θ s ) K 1
σ zr =− + ,
2π r12 2π r
1 + K rs sin(θ − θ s )
σ z(θ2) =−
2π r12

It is easy to show that the continuity conditions at the interface as specified in Eq.
(3.102) are satisfied.
Remark 3.29: If μ1 = μ2 = μ, then K = 0. Thus, the expressions in Eqs. (3.105)
and (3.107) correspond to the Green’s displacement and stresses for an infinite
homogeneous plane subjected to a concentrated line force at z = zs, as already
discussed in Section 3.2.
Remark 3.30: If μ2 = 0, then K = –1. Unlike the half-plane case, the stress σ zr thus
obtained doesn’t vanish at r = a, as may be seen from Eq. (3.107). This observation
is similar to that in Section 1.3.5.2 for the ill-posed potential problem. Here it is
physically obvious that the concentrated line force inside the circular plate needs to
be balanced and thus the surface of circular plate cannot be traction free.
78 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

θ1, θ2

r2
z a2 / z–s

a r1
zs
x

Figure 3.1. Geometry of a circular inhomogeneity with a source inside.

Remark 3.31: If μ2 = ∞, then K = 1. We thus have the Green’s function solution in


r < a due to a concentrated line force applied in a circular disc with a rigidly fixed
boundary:

1  ar 
u=− ln  1 
2πµ  r2 rs 
1 r − rs cos(θ − θ s ) 1 r − (a 2 / rs )cos(θ − θ s )
σ zr =− + (3.108)
2π r12 2π r22
1 rs sin(θ − θ s ) 1 (a 2 / rs )sin(θ − θ s )
σ zθ = − +
2π r12 2π r22

where μ = μ1 and the superscript (1) has been omitted for simplicity.
When the line force of unit magnitude is applied at z = zs (|zs| > a) within the
matrix (Figure 3.2), the antiplane governing equations in terms of the antiplane
displacements in Materials 1 and 2 are

∇2 u(1) = 0 ( z < a)
(3.109)
∇2 u( 2) = −δ(z − zs ) / µ 2 ( z > a, zs > a)

along with the interface continuity conditions as described by Eq. (3.102). According
to Section 1.3.5.1, we have Eq. (3.103) again, but with F (i ) (z) being given by

1
F (1) (z) = − ln(z − zs ) ( z < a)
π(µ1 + µ 2 )
(3.110)
1 (µ 2 − µ 1 )  a2 
F ( 2) (z) = − ln(z − zs ) − ln  − zs  ( z > a)
2 πµ 2 πµ µ µ
2 2( 1 + 2)  z 

The displacements in real form are found to be


3.7 Line Forces or Line Dislocations Interacting with a Circular Inhomogeneity 79

y
θ2, θ1
z
r1

zs
r 2
a

a2 / z–s x

Figure 3.2. Geometry of the circular inhomogeneity in a matrix with a source in the matrix.

(1 − K )
u(1) = − ln r1 (r < a)
2 πµ 1
(3.111)
1 K   r 
u( 2) = − ln r1 −  ln r2 − ln    ( r > a)
2 πµ 2 2 πµ 2   rs  

with r1 and r2 still being given by Eq. (3.106), as also shown in Figure 3.2. We note
that the relation r2 = ar1 / rs still holds when the field point is on the circle |z| = a.
Thus, the displacement continuity at the interface is obvious from Eq. (3.111). The
stresses are calculated as

1 − K r − rs cos(θ − θ s ) 1 − K rs sin(θ − θ s )
σ(zr1) = − σ(z1θ) = −
2π r12 2π r12
1 r − rs cos(θ − θ s ) K  r − (a 2 / rs )sin(θ − θ s ) 1 
σ(z2r) = − +  −  (3.112)
2π r12 2π  r22 r
1 rs sin(θ − θ s ) K (a 2 / rs )sin(θ − θ s )
σ(z2θ) = − −
2π r12 2π r22

The shear stress σ zr is also continuous across the interface.

Remark 3.32: If μ1 = μ2 = μ, then K = 0. We again obtain the Green’s function solution


for an infinite homogeneous plane subjected to a concentrated line force at z = zs.

Remark 3.33: If μ1 = 0, then K = 1. We have the Green’s function solution in the


matrix material r > a due to a concentrated line force in an infinite plane (μ2 = μ)
with a circular hole, the boundary of which (r = a) is traction-free:
80 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

1   r 
u=−  ln r1 + ln r2 − ln   
2πµ   rs  
1 r − rs cos(θ − θ s ) 1  r − (a 2 / rs )sin(θ − θ s ) 1 
σ zr =− +  −  (3.113)
2π r12 2π  r22 r
1 rs sin(θ − θ s ) 1 (a / rs )sin(θ − θ s )
2
σ zθ =− −
2π r12 2π r22

where the superscript (2) has been omitted. As noticed in Section 1.3.5.1, such a
Green’s displacement is unbounded at infinity, and may need to be amended for
certain applications.
Remark 3.34: If μ1 = ∞, then K = –1. We have the Green’s function solution in the
matrix material r > a due to a concentrated line force in an infinite plane (μ2 = μ)
with a rigid circular inhomogeneity of radius a inside:

1   r 
u=−  ln r1 − ln r2 + ln   
2πµ   rs  
1 r − rs cos(θ − θ s ) 1  r − (a 2 / rs )sin(θ − θ s ) 1 
σ zr =− −
 −  (3.114)
2π r12  2π r22 r
1 rs sin(θ − θ s ) 1 (a 2 / rs )sin(θ − θ s )
σ zθ =− +
2π r12 2π r22

3.7.1.2 A Screw Dislocation Inside or Outside a Circular Inhomogeneity


We now assume that there is a screw dislocation of Burgers vector b in the z-direction
applied at z = zs (|zs| < a) within the circle. Then both the antiplane displacements in
Materials 1 and 2 satisfy the Laplace’s equation (3.12), and subjected to a displacement
discontinuity at z = zs as shown in Eq. (3.13). Similar to the line-force case, we can also
assume that the displacements are in form of Eq. (3.103), but with F (i ) (z) being as

bi (µ − µ 2 )bi   a 2  a
F (1) (z) = − ln(z − zs ) − 1  ln  − z + ln  ( z < a)
2π 2π(µ 1 + µ 2 )   zs  zs 
(3.115)
µ 1b i (µ − µ 2 )bi z
F ( 2) (z) = − ln(z − zs ) + 1 ln ( z > a)
πµ 2 (µ1 + µ 2 ) 2 π (µ 1 + µ 2 ) a

The displacements in real form are found to be

b
u(1) = [θ1 − K (π − θ 2 + θ s )] ( r < a)
2π (3.116)
b
u( 2) = [(1 − K )θ1 + K θ] (r > a )

where we have denoted z − zs = r1 ei θ1 and z − a 2 / zs = r2 ei θ 2 (for simplicity, here we


assume 0 ≤ arg(z) ≤ 2π , also see Figure 3.1). Thus, when z is on the interface (r = a),
we have
3.7 Line Forces or Line Dislocations Interacting with a Circular Inhomogeneity 81

θ1 + θ 2 = θ + θ s + π (3.117)

It is readily seen that the displacement continuity condition is met by the preceding
solution (3.116). The corresponding stresses are

(1) µ1b  rs sin(θ − θ s ) (a 2 / rs )sin(θ − θ s ) 


σ zr =−  + K 
2π  r12 r22 
µ1b  r − rs cos(θ − θ s ) r − (a 2 / rs )cos(θ − θ s ) 
σ z(θ1) =  + K 
2π  r12 r22  (3.118)
( 2) µ2 b rs sin(θ − θ s )
σ zr =− (1 − K ) ,
2π r12
µ2 b  r 2 − rrs cos(θ − θ s ) 
σ z(θ2) = (1 − K ) + K 
2π  r12 

where use has been made of the following derivatives:

∂θ1 r sin(θ − θ s ) ∂θ1 r 2 − rrs cos(θ − θ s )


=− s , =
∂r r12 ∂θ r12
(3.119)
∂θ2 (a2 / rs )sin(θ − θ s ) ∂θ 2 r 2 − r(a 2 / rs )cos(θ − θ s )
=− , =
∂r r22 ∂θ r22

Clearly, the shear stress σ zr is also continuous across the interface.


Remark 3.35: If μ1 = μ2 = μ, then K = 0. We obtain the Green’s function solution for
an infinite homogeneous plane subjected to a concentrated line screw dislocation at
z = zs, as obtained in Section 3.2.
Remark 3.36: If μ2 = 0, then K = –1. We get the Green’s function solution in r < a due
to a screw dislocation in a circular disc of Material 1 (μ1 = μ), the boundary of which
(r = a) is traction-free:

b
u= (θ1 − θ 2 + π + θ s )

µ b  r sin(θ − θ s ) (a 2 / rs )sin(θ − θ s ) 
σ zr =− 1  s −  (3.120)
2π  r12 r22 
µ1b  r − rs cos(θ − θ s ) r − (a 2 / rs )cos(θ − θ s ) 
σ zθ =  − 
2π  r12 r22 

where the superscript (1) has been omitted.


Remark 3.37: If μ2 = ∞, then K = 1. We get the Green’s function solution in r < a due
to a screw dislocation in a circular disc of Material 1 (μ1 = μ), the boundary of which
(r = a) is fixed:
82 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

b
u= (θ1 + θ 2 − π − θ s )

µ b  r sin(θ − θ s ) (a 2 / rs )sin(θ − θ s ) 
σ zr =− 1  s +  (3.121)
2π  r12 r22 
µ1b  r − rs cos(θ − θ s ) r − (a 2 / rs )cos(θ − θ s ) 
σ zθ =  + 
2π  r12 r22 

If the screw dislocation is located at z = zs (|zs| > a) within the matrix, we can assume,
instead of Eq. (3.115),

µ 2 bi
F (1) (z) = − ln(z − zs ) ( z < a)
π(µ1 + µ 2 )
(3.122)
bi (µ − µ 2 )bi  a 2 
F ( 2) (z) = − ln(z − zs ) − 1 ln − zs  ( z > a)
2π 2 π(µ1 + µ 2 )  z 

The displacements in real form are found to be

b
u(1) = (1 + K )θ1 ( r < a)
2π (3.123)
b
u( 2) = [θ1 − K (θ2 − π − θ s − θ)] (r > a )

The corresponding stresses are

(1) µ1 b r sin(θ − θ s )
σ zr =− (1 + K ) s
2π r12
µb r 2 − rrs cos(θ − θ s )
σ z(θ1) = 1 (1 + K )
2π r12
(3.124)
( 2) µ2 b  rs sin(θ − θ s ) (a 2 / rs )sin(θ − θ s ) 
σ zr =−  −K 
2π  r12 r22 
µ 2 b  r − rs cos(θ − θ s ) r − (a 2 / rs )cos(θ − θ s ) 1
σ z(θ2) =  −K +K 
2π  r12 2
r2 r

It is easy to check that the continuity conditions at the interface r = a are satisfied by
the preceding solutions.

Remark 3.38: If μ1 = μ2 = μ, then K = 0. The preceding results become the Green’s


function solution in an infinite homogeneous plane subjected to a concentrated
screw dislocation at z = zs.

Remark 3.39: If μ1 = 0, then K = 1. We obtain the Green’s function solution in r > a


due to a concentrated screw dislocation in an infinite plane of Material 2 (μ2 = μ)
with a circular hole, the boundary of which (r = a) is traction-free:
3.7 Line Forces or Line Dislocations Interacting with a Circular Inhomogeneity 83

b
u= ( 2θ1 − π )

µ b  r sin(θ − θ s ) (a 2 / rs )sin(θ − θ s ) 
σ zr =− 2  s −  (3.125)
2π  r12 r22 
µ2 b  r − rs cos(θ − θ s ) r − (a 2 / rs )cos(θ − θ s ) 1 
σ zθ =  − + 
2π  r12 r22 r

Remark 3.40: If μ1 = ∞, then K = –1. We get the Green’s function solution in r > a due
to a concentrated line screw dislocation in an infinite plane of Material 2 (μ2 = μ)
with a rigid circular inhomogeneity of radius a inside:

b
u= (θ1 + θ 2 − θ s − θ − π )

µ b  r sin(θ − θ s ) (a 2 / rs )sin(θ − θ s ) 
σ zr =− 2  s +  (3.126)
2π  r12 r22 
µ2 b  r − rs cos(θ − θ s ) r − (a 2 / rs )cos(θ − θ s ) 1 
σ zθ =  + − 
2π  r12 r22 r

Remark 3.41: Because we have already derived the bimaterial plane solution, an
alternative way to find the solution for a circle of radius a within a matrix is to apply
the powerful conformal mapping method. It can be easily shown that the following
functions map the circle-in-matrix in the z-plane to the bimaterial ζ (ξ+iη)-plane:

a−z i −ζ
ζ= i, z=a (3.127)
z+ a i+ζ

We take as an example the case of a line force acting at z = zs (|zs| < a) inside the
circular inhomogeneity. With the mapping functions in Eq. (3.127), we map the
source point on the z-plane onto ζ s = ξ s + iηs = i(a − zs ) / (zs + a) on the upper half-
plane of the ζ-plane. Then, according to Section 3.5.1, we may take in Material 1 (the
upper half-plane η > 0):

1 µ1 − µ2
F (1) (ζ ; ζ s ) = − ln(ζ − ζ s ) − ln(ζ − ζ s ) (3.128a)
2πµ1 2πµ1 ( µ1 + µ2 )

and in Material 2, the lower half-plane

1
F ( 2) (ζ ; ζ s ) = − ln(ζ − ζ s ) (3.128b)
π ( µ1 + µ 2 )

The antiplane displacements are the real parts of the preceding complex functions
for the two half-planes, respectively. It can be readily shown that Eq. (3.128) will lead
to Eq. (3.104) when expressed in terms of z by removing certain rigid-body motions
as well as the singularity (z = −a) caused by the transformation (3.127).
84 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

3.7.2 Plane-Strain Solutions


This section presents the plane-strain counterparts of Green’s solutions for the
same two-phase (circle/matrix) material system considered in the last subsection.
For plane-strain problems, we will adopt the complex variable formalism. When
expressed in polar coordinates, we have

2µ(ur + iuθ ) = e − iθ [κφ(z) − zφ ′(z) − ψ (z)]


σ rr + σθθ = 2[φ ′(z) + φ ′(z)] (3.129)
z
σ rr + iσ rθ = φ ′(z) + φ ′(z) − [z φ ′′(z) + ψ ′(z)]
z

The resultant force of the traction exerted on an arc is still given by Eq. (3.22).
From the theory of complex variables, we note that, if f1 (z) is analytic in the
interior of the circle r < a and can be expanded as


f1 (z) = ∑ Cn zn (3.130)
n =0

then


a2 n
f1 (a 2 / z) = ∑ Cn zn
(3.131)
n=0

is analytic in the exterior of the circle r > a, including the point at infinity. Conversely,
if f2 (z) is analytic in the exterior of the circle r > a, including the point at infinity and
can be expanded as


D
f2 (z) = ∑ znn (3.132)
n=0

then


D
f2 (a 2 / z) = ∑ a2nn zn (3.133)
n =0

is analytic within the circle r < a. The transformation f (a2 /z) = f (a2 /z ) of f (z) is
referred to as the hat transformation, following Honein and Herrmann (1990).

3.7.2.1 Line Forces or Edge Dislocations Outside a Circular Inhomogeneity


First, let us consider the case when a line force (f = fx+ify) or edge dislocation
(b = bx+iby) is applied at z = zs (|zs| > a) outside the circular inhomogeneity of radius
a. Unless stated otherwise, the notations adopted before will be followed.
3.7 Line Forces or Line Dislocations Interacting with a Circular Inhomogeneity 85

The complex functions in Eq. (3.129) are assumed to be

φ1 (z) = φ1 (z), ψ 1 (z) = ψ 1 (z) ( z < a)


(3.134)
φ 2 (z) = φ 2 p (z) + φ 2 c (z), ψ 2 (z) = ψ 2 p (z) + ψ 2 c (z) ( z > a)

where φ2 p (z) and ψ 2 p (z) are the “particular solution” given as

zs
φ2 p (z) = C ln(z − zs ), ψ 2 p (z) = D ln(z − zs ) − C (3.135)
z − zs

with

fx + if y κ 2 ( fx − if y )
C=− , D= (3.136a)
2 π(1 + κ 2 ) 2 π(1 + κ 2 )

for the line force (Fx,Fy) at z = zs, and

iµ 2 (bx + iby ) iµ 2 (bx − iby )


C=− , D= (3.136b)
π(1 + κ 2 ) π(1 + κ 2 )

for the line dislocation (bx,by) at z = zs. This particular solution is obtained from Eq.
(3.87) simply by switching the subscripts 1 and 2, and by replacing A and B by C and
D. Note that both φ2 p (z) and ψ 2 p (z) are analytic for r < a.
We now assume that φ1 (z) and ψ 1 (z) are analytic for r < a, while φ2c (z) and
ψ 2c (z) are analytic for r > a. From the hat transformation properties, we know that
φ2 p (a 2 / z),ψ 2 p (a 2 / z), φ1 (a 2 / z) and ψ 1 (a 2 / z) will be analytic for r > a, and φ2 c (a 2 / z)
and ψ 2 c (a 2 / z) will be analytic for r < a.
Similar to the bimaterial plane case (Section 3.6), the four complex functions
(φ1, ψ1, φ2c and ψ2c) should satisfy and be determined from the following interface
conditions at r = a:

Γ{κ 1φ1 (t ) − [t φ1′ (t ) + ψ 1 (t )]}+ = {κ 2 φ 2 (t ) − [t φ 2′ (t ) + ψ 2 (t )]}−


(3.137)
[φ1 (t ) + t φ1′ (t ) + ψ 1 (t )]+ = [φ 2 (t ) + t φ 2′ (t ) + ψ 2 (t )]−

where t = a eiτ is a point on the circle r = a, and Γ = µ2 / µ1 is the modulus ratio.


The superscripts “+” and “–” indicate the limiting value of the functions when
approaching the interface (r = a) from r < a and r > a, respectively, that is, from the
circular inhomogeneity and the matrix, respectively.
Substituting Eq. (1.134) into the preceding continuity conditions yields

Γκ 1[φ1 (t )]+ − Γ [tφ1′ (t ) + ψ 1 (t )]+


= κ 2 [φ2 p (t )]− + κ 2 [φ2 c (t )]− − [tφ2′ p (t ) + ψ 2 p (t )]− − [tφ2′ c (t ) + ψ 2 c (t )]−
   (3.138)
[φ1 (t )]+ + [tφ1′ (t ) + ψ 1 (t )]+
= [φ2 p (t)]− + [φ2c (t)]− + [tφ2′ p (t) + ψ 2 p (t)]− + [tφ2′ c (t) + ψ 2 c (t)]−
86 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

which can be rearranged as follows

Γκ 1[φ1 (t )]+ − κ 2 [φ2 p (t )]− + [tφ2′ c (t ) + ψ 2c (t )]−


= κ 2 [φ2c (t )]− − [tφ2′ p (t ) + ψ 2 p (t )]− + Γ [tφ1′ (t ) + ψ 1 (t )]+
(3.139)
[φ1 (t )]+ − [φ2 p (t )]− − [tφ2′ c (t ) + ψ 2 c (t )]−
= [φ2 c (t )]− + [tφ2′ p (t ) + ψ 2 p (t )]− − [tφ1′ (t ) + ψ 1 (t )]+

Because on r = a, we have tt = a 2 , the preceding equation can be rewritten as

{
lim Γκ 1φ1 (z) − κ 2φ2 p (z) + [zφ2′ c (a 2 / z) +ψ 2c (a 2 / z)]
z→ t
}
z <a

{
= lim κ 2φ2 c (z) − [zφ2′ p (a 2 / z) + ψ 2 p (a 2 / z)] + Γ [zφ1′ (a 2 / z) + ψ 1 (a 2 / z)]
z→t
}
z >a (3.140)
{
lim φ1 (z) − φ2 p (z) − [zφ2′ c (a2 / z) + ψ 2 c (a2 / z)]
z→ t
}
z <a

{
= lim φ2c (z) + [zφ2′ p (a 2 / z) + ψ 2 p (a 2 / z)] − [zφ1′ (a 2 / z) +ψ 1 (a 2 / z)]
z→ t
}
z >a

Unlike the bimaterial plane case (i.e., Eq. (3.80)), the right-hand sides contain
the variable z, which will cause undesired singularities at infinity. To remove such
singularities, we rewrite Eq. (3.140) as

{
lim Γκ 1φ1 (z) − κ 2 φ 2 p (z) + [zφ 2′ c (a 2 / z) + ψ 2 c (a 2 / z)] + zφ 2′ p (0) − Γ zφ1′ (0)
z→ t
}
z <a

{ }
= lim κ 2 φ 2 c (z) − [zφ 2′ p (a 2 / z) − zφ 2′ p (0) + ψ 2 p (a 2 / z)] + Γ[zφ1′ (a 2 / z) − zφ1′ (0) + ψ 1 (a 2 / z)]
z→ t
z >a

{
lim φ1 (z) − φ 2 p (z) − [zφ 2′ c (a 2 / z) + ψ 2 c (a 2 / z)] − zφ 2′ p (0) + zφ1′ (0)
z→ t
}
z <a

{
= lim φ2 c (z) + [zφ 2′ p (a 2 / z) − zφ 2′ p (0) + ψ 2 p (a 2 / z)] − [zφ1′ (a 2 / z) − zφ1′ (0) + ψ 1 (a 2 / z)]
z→ t
}
z >a
(3.141)

Then in each of these two equations, the left-hand side is analytic in Material 1
(r < a) and the right-hand side is analytic in Material 2 (r > a). Therefore, from the
analytic continuation theory, we have, in Material 1 (r < a)

Γκ 1φ1 (z) − κ 2φ2 p (z) + [zφ2′ c (a 2 / z) + ψ 2c (a 2 / z)] + zφ2′ p (0) − Γ zφ1′ (0) = 0
(3.142)
φ1 (z) − φ2 p (z) − [zφ2′ c (a 2 / z) + ψ 2c (a 2 / z)] − zφ2′ p (0) + zφ1′ (0) = 0
3.7 Line Forces or Line Dislocations Interacting with a Circular Inhomogeneity 87

and in Material 2 (r > a)

κ 2φ2c (z) − [zφ2′ p (a 2 / z) − zφ2′ p (0) +ψ 2 p (a 2 / z)]


+ Γ [zφ1′ (a 2 / z) − zφ1′ (0) + ψ 1 (a 2 / z)] = 0
(3.143)
φ2c (z) + [zφ2′ p (a 2 / z) − zφ2′ p (0) + ψ 2 p (a 2 / z)]
− [zφ1′ (a 2 / z) − zφ1′ (0) + ψ 1 (a 2 / z)] = 0

From Eq. (3.142), we can obtain

1 + κ2 Γ −1 1+α* α* − β*
φ1 (z) = φ2 p (z) + zφ1′ (0) = φ 2 p ( z) − zφ1′ (0)
1 + Γκ 1 1 + Γκ 1 1− β* 1− β*
κ − Γκ 1 Γ + Γκ 1
zφ2′ c (a 2 / z) + ψ 2 c (a 2 / z) = 2 φ2 p (z) − zφ2′ p (0) + zφ1′ (0) (3.144)
1 + Γκ 1 1 + Γκ 1
α* + β* 1 −α*
= φ2 p (z) − zφ2′ p (0) + zφ1′ (0)
1− β * 1− β*

and from Eq. (3.143), we have

1− Γ
φ2 c (z) = zφ2′ p (a 2 / z) − zφ2′ p (0) + ψ 2 p (a 2 / z) 
κ2 + Γ  
α* − β*
=  zφ2′ p (a 2 / z) − zφ2′ p (0) + ψ 2 p (a 2 / z) 
1+ β*  
1 + κ2 (3.145)
zφ1′ (a 2 / z) + ψ 1 (a 2 / z) − zφ1′ (0) = zφ2′ p (a 2 / z) − zφ2′ p (0) + ψ 2 p (a 2 / z) 
κ2 + Γ  
1+α*
= zφ2′ p (a 2 / z) − zφ2′ p (0) +ψ 2 p (a 2 / z) 
1+ β*  

where α * and β * also are Dundurs parameters but with the roles played by the two
phases being switched:

(κ 2 + 1) − Γ (κ 1 + 1) (κ − 1) − Γ (κ 1 − 1)
α* = , β* = 2 (3.146)
(κ 2 + 1) + Γ (κ 1 + 1) (κ 2 + 1) + Γ (κ 1 + 1)

From the first expression in Eq. (3.144), we can derive

φ1′ (0) = M1φ2′ p (0) − M2φ2′ p (0), φ1′ (0) = M1φ2′ p (0) − M 2φ2′ p (0) (3.147)

where

(1 + α * )(1 − β * ) (1 + α * )(α * − β * )
M1 = , M 2 = (3.148)
(1 − β * )2 − (α * − β * )2 (1 − β * )2 − (α * − β * )2
88 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

In view of Eq. (3.147), the four unknown complex functions are finally solved from
Eqs. (3.144) and (3.145) as

1+α* α* − β*
φ1 (z) = φ2 p (z) − [ M1φ2′ p (0) − M2φ2′ p (0)]z
1− β * 1− β*
1+α* 2(1 + α * )β *
ψ 1 (z) = ψ 2 p (z) − f (z)
1− β * (1 + β * )(1 − β * )
α* − β*
φ2c (z) = f (a 2 / z)
1+ β*
α* + β* α * − β * a4
ψ 2c (z) = φ2 p (a 2 / z) + f ′(a 2 / z)
1− β * 1 + β * z3
 1−α*  a2 1−α* a2
− 1 − (M1 − M 2 ) φ2′ p (0) + M 2 [φ2′ p (0) − φ2′ p (0)]
 1− β 1− β
* *
 z z (3.149)

where we have introduced the following auxiliary function

a2
f (z) = [φ2′ p (z) − φ2′ p (0)] + ψ 2 p (z) (3.150)
z

and f (a 2 / z) and f ′(a 2 / z) denote the hat transforms of f (z) and f ′(z), respectively.
It can be easily shown that if φ2 p (z) were chosen such that φ2′ p (0) is real, then Eq.
(3.149) reduces to the same expressions as presented in Honein and Herrmann
(1990).
In terms of the particular solutions (3.135), the four complex functions in Eq.
(3.149) can be finally expressed as

1 + α* α * − β*  C C
φ1 (z) = C ln(z − z ) − M2 − M1  z
1 − β* 
s
1 − β* zs zs 
1 + α* zs  2(1 + α* )β*  1 1  a2
ψ 1 (z) =  D ln(z − zs ) − C z − z  − 1 − β*2 C  z − z + z  z
1 + β* s s s
α −β  
* * 1 1 z 
φ 2c (z) = C  2 +  z + D ln(a 2 / z − zs ) − C 2 s 
1 + β*
  a / z − z s z s  a / z − s
z
α * + β* (3.151)
ψ 2c (z) = C ln(a 2 / z − zs )
1 − β*
α * − β* a 4  zs − z  1 1  z2 D 
+ C − C  + + 
1 + β* z3  (a 2 / z − zs )2  a 2 / z − zs zs  a 2 a 2 / z − zs 
 1 − α*  C a 2 1 − α*  C C  a2
+ 1 − (M1 − M2 ) + M2  − 
 1−β *
 zs z 1 − β
*  zs zs  z

Remark 3.42: If μ1 = μ2 and ν1 = ν2, then α * = β * = 0 and M1 = 1, M 2 = 0. The


complementary complex functions in the matrix (Material 2) are zero while those in
the inhomogeneity (Material 1) are reduced to

φ1 (z) = C ln(z − zs )
(3.152)
ψ 1 (z) = D ln(z − zs ) − Czs / (z − zs )
3.7 Line Forces or Line Dislocations Interacting with a Circular Inhomogeneity 89

These are identical to the particular solutions (3.135). Consequently, when substituting
these complex functions into Eq. (3.21), the displacement and stress fields obtained will
be the same as those in the infinite plane with the concentrated line source at z = zs.
Remark 3.43: If μ1 = 0 (α * = −1, M1 = M2 = 0), we have an infinite plane of Material
2 with a circular hole subjected to a traction-free boundary condition at r = a. The
complementary solution in the matrix plane is

  1 1 z 
φ 2 c (z) = − C  2 +  z + D ln(a 2 / z − zs ) − C 2 s 
  a / z − zs zs  a / z − zs 
C a2
ψ 2 c (z) = −C ln(a 2 / z − zs ) +
   zs z (3.153)
a 
4 z −z  1 1  z2 D 
− 3 C 2 s 2
− C  +  + 2 
z  (a / z − zs ) 2
 a / z − zs zs  a 2 a / z − zs 

Remark 3.44: If μ1 = ∞ (α = 1, β = (κ − 1) / (κ + 1)), we will have an infinite plane


of Material 2 with a rigid circular inhomogeneity of radius a. The complementary
solution in the matrix plane is

1  1 1 z 
φ 2 c (z) = C  2 +  z + D ln(a 2 / z − zs ) − C 2 s 
κ   a / z − zs zs  a / z − zs 
ψ 2 c (z) = κC ln(a 2 / z − zs )
   1 a4  zs − z  1 1  z2 D  (3.154)
+ C − C  +  + 2 
κ z  (a / z − zs )
3 2 2  2
a / z − zs zs  a 2 a / − s
z z
 (κ + 1)2  C a 2 κ + 1  C C  a 2
+ 1 − + −
 4  zs z 2  zs zs  z

The displacement and stress fields then can be obtained from the complex functions
in Eqs. (3.151) and (3.135). Similar to the bimaterial plane case, we first write Eq.
(3.151) as

 C C
φ1 (z) = c3*C ln(z − zs ) − c6*  M2 − M1  z
 zs zs 
 z   1 1  a2
ψ 1 (z) = c4*  D ln(z − zs ) − C s  + c5*C  + 
 z − zs   z − zs zs  z
  1 1 zs 
   φ2 c (z) = c1 C  a 2 / z − z + z  z + D ln(a / z − zs ) − C a 2 / z − z 
* 2
(3.155)
 s s s
C a2  C C  a2
ψ 2c (z) = c2*C ln(a 2 / z − zs ) + c7* + c8*  − 
zs z  zs zs  z
a4  zs − z  1 1  z2 D 
+ c1* C 2 2
−C 2 +  2+ 2 
z3  (a / z − zs )  a / z − zs zs  a a / z − zs 
90 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

where

α* − β* µ1 − µ 2 * = α + β = κ 2 µ 1 − κ 1µ 2 ,
* *
c1* = = , c2
1 + β* κ 2 µ1 + µ 2 1 − β* κ 1µ 2 + µ 1
1+ α * µ
* = 1 2 (κ + 1 ) 1 + α * µ (κ + 1)
c3* = = 1 + c 2 , c * =
4 = 1 + c1* = 1 2 ,
1− β * κ 1µ 2 + µ 1 1+ β * κ 2 µ1 + µ 2
2(1 + α* )β* µ (κ − 1) + µ1 (1 − κ 2 )
c5* =− = c4* − c3* = c1* − c2* = µ 1 (κ 2 + 1) 2 1 (3.156)
1−β * 2 (κ 2 µ1 + µ 2 )(κ 1µ 2 + µ 1 )
α* − β* µ1 − µ 2 1 − α * µ1 (1 + κ 1 )(µ1 − µ 2 )
c6* = = , c8* = M2 =
1−β * κ 1µ 2 + µ 1 1−β * (κ 1µ 2 + µ 1 )(2µ 1 + µ 2 κ 1 − µ 2 )
1 − α* µ 1µ 2 (1 + κ 1 )(1 + κ 2 )
c7* = 1− ( M1 − M 2 ) = 1 −
1 − β* (κ 1µ 2 + µ 1 )(2µ 1 + µ 2 κ 1 − µ 2 )

In the matrix (Material 2) where the source is located, the total solution is the
summation of the particular solution and the complementary parts. For the particular
solution, we have,

[ 2µ 2 eiθ (ur + iuθ )]2 p = κ 2C ln(z − zs ) − C (z − zs ) / (z − zs ) − D ln(z − zs )


[σ rr + σθθ ]2 p / 2 = C / (z − zs ) + C / (z − zs )
   (3.157)
[σrr + iσrθ ]2 p = C / (z − zs ) + C / (z − zs ) + Cz / (z − zs )2
− (z / z)[D / (z − zs ) + Czs / (z − zs )2 ]

Using Eqs. (3.129) and (3.155), the complementary part in the matrix is found to be

  1 1 z 
[2µ 2 eiθ (ur + iuθ )]2c = κ 2 c1* C  2 +  z + D ln(a 2 / z − zs ) − C 2 s 
  a / z − zs zs  a / z − zs 
  1 1 1 a2  zs  1 a2 
− c1*z C  2 +  +C 2  1 −  − D 2 
  a / z − zs zs  a / z − zs z 2 
2
(a / z − zs ) z z
C a2  C C  a2
− c2*C ln(a 2 / z − zs ) − c7* − c8*  − 
zs z  zs zs  z
a4  zs − z  1 1  z2 D 
− c1* C 2 −C  2 +  2 + 2 
z3  (a / z − zs )
2  a / z − zs zs  a a / z − zs 
 (3.158a)

  1 1 1 a2  zs  1 a2 
[σ rr + σθθ ]2 c / 2 = c1* C  2 +  +C 2  1 −  − D 
  a / z − zs zs  (a / z − zs )2 z  z a 2 / z − zs z2 
  1 1 1 a2  zs  1 a2 
+ c1* C  2 +  +C 2  1 −  − D 2 
  a / z − zs zs 
2
(a / z − zs ) z z a / z − zs z 2 
  1 1 1 a2  zs  1 a2 
[σrr + iσrθ ]2c = c1* C  2 +  +C 2  1 −  − D 2 
  a / z − zs zs 
2
(a / z − zs ) z z a / z − zs z2 
  1 1 1 a2  zs  1 a2 
+ c1* C  2 +  +C 2  1 −  − D 
  a / z − zs zs  (a / z − zs )2 z  z a 2 / z − zs z 2 
3.7 Line Forces or Line Dislocations Interacting with a Circular Inhomogeneity 91

 1 a2 2 a4  zs  1 a2  zs 
− c1* C 2 2
+C 2  1 −  − C 2 2  1 − 
 (a / z − zs ) z (a / z − zs ) z 2
3 z (a / z − zs ) z z
1 a 2 zs 1 a4 2 a2 
+C 2 −D 2 + D
(a / z − zs )2 z 2 (a / z − zs )2z3 a 2 / z − zs z 2 
1 a2 C a2  C C  a2
+ c2*C 2 + c7* + c8*  − 
a / z − zs zz zs zz  zs zs  zz
a4  zs − z  1 1  z2 D 
+ 3c1* 3 C 2 2
− C  2
+  2
+ 2 
zz  (a / z − zs )  a / z − zs zs  a a / z − zs 
a4  2(z − z ) a 2 1  1 1 z
− c1* 2 C 2 s −C 2 − 2C  2 +
zz  (a / z − zs )3 z 2 (a / z − zs )2  a / z − zs zs  a2
1 D a2 
−C + 2
(a2 / z − zs ) 2 (a / z − zs ) z 2 
2
 (3.158b)

The total displacement and stress fields in the matrix (Material 2, r > a) are the
summation of those in Eqs. (3.157) and (3.158).
In the inhomogeneity (Material 1, r < a), the displacement and stress fields
are found by using Eqs. (3.129) and the first two expressions in Eq. (3.155). The
results are

 C C
2µ 1eiθ (ur + iuθ ) = c3*κ 1C ln(z − zs ) − c6*κ 1  M2 − M1  z
 zs zs 
z  C C
− c3*C + c6*  M 2 − M1  z (3.159a)
z − zs  zs zs
 z   1 1  a2
− c4*  D ln(z − zs ) − C s  − c5*C  + 
 z − zs   z − zs zs  z

 C C  * C C
(σrr + σθθ ) / 2 = c3*  +  − c6 (M2 − M1 )  + 
 z − zs z − zs   zs zs 
 C C  * C C
σrr + iσ rθ = c3*  +  − c6 (M2 − M1 )  + 
 z − zs z − zs  zs zs 
z  1 zs z (3.159b)
+ c3*C − c4*  D +C
(z − zs ) 2
 z − zs (z − zs )  z
2

 1 a2  1 1  a2 
+ c5*C  +  +  
 (z − zs ) z  z − zs zs  zz 
2

Remark 3.45: If μ1 = μ2 = μ and ν1 = ν2 = ν, then c1* = c2* = 0. Thus, the complementary


parts of the displacement and stress fields in Material 2 given by Eq. (3.158) are
zero, and the total solution is reduced to the particular solution given by Eq. (3.157),
which gives the Green’s displacements and stresses for an infinite homogeneous
plane subjected to a concentrated source at z = zs.
Remark 3.46: Also, if μ1 = μ2 = μ and ν1 = ν2 = ν, then c3* = c4* = 1 and c5* = c6* = 0 . The
displacement and stress fields in Material 1 given by Eq. (3.159) are reduced to the
following expressions
92 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

z − zs
2µ eiθ (ur + iuθ ) = κC ln(z − zs ) − C − D ln(z − zs ) (3.160a)
z − zs

C C
(σrr + σθθ ) / 2 = +
z − zs z − zs
(3.160b)
C C z  1 zs z
σrr + iσrθ = + +C − D + C
z − zs z − zs (z − zs )2  z − zs (z − zs )2  z

which coincide with Eq. (3.157), that is, the solution of the displacement and stress
fields in an infinite plane with the source at z = zs.
Remark 3.47: If μ1 = 0, we have an infinite plane of Material 2 with a circular hole
subjected to a traction-free boundary condition at r = a. For this case, c1* = c2* = −1,
c7* = 1, and c8* = 0. The complementary solution (3.158) is then reduced to

  1 1 z 
[2µ 2 eiθ (ur + iuθ )]2 c = −κ 2 C  2 +  z + D ln(a 2 / z − zs ) − C 2 s 
  a / z − zs zs  a / z − zs 
  1 1 1 a2  zs  1 a2 
+ z C  2 +  +C 2  1 −  − D 
  a / z − zs zs  (a / z − zs )2 z  z a 2 / z − zs z 2 
C a2 a4  z −z  1 1  z2 D 
+ C ln(a 2 / z − zs ) − + 3 C 2 s 2
− C  2
+  2
+ 2 
zs z z  (a / z − zs )  a / z − zs zs  a a / z − zs 

 (3.161a)

  1 1 1 a2  zs  1 a2 
[σrr + σθθ ]2c / 2 = − C  2 +  +C 2  1 −  − D 2 
  a / z − zs zs 
2
(a / z − zs ) z z a / z − zs z2 
  1 1 1 a2  zs  1 a2 
− C  2 +  +C 2  1 −  − D 
  a / z − zs zs  (a / z − zs )2 z  z a 2 / z − zs z 2 
  1 1 1 a2  zs  1 a2 
[σ rr + iσ rθ ]2 c = − C  2 +  +C 2  1 −  − D 2 
  a / z − zs zs 
2
(a / z − zs ) z z a / z − zs z2 
  1 1 1 a2  zs  1 a2 
− C  2 +  +C 2  1 −  − D 
  a / z − zs zs  (a / z − zs )2 z  z a 2 / z − zs z 2 
 1 a2 2 a4  zs  1 a2  zs 
+ C 2 +C 2  1 −  − C 2  1 − 
2
 (a / z − zs ) z
3
(a / z − zs ) z 2 z (a / z − zs ) z 
2 z
1 2
a zs 1 a 4 2 a 
2
+C 2 −D 2 +D 2 
(a / z − zs )2 z 2 (a / z − zs )2 z 3 a / z − zs z 2 
1 a2 C a2
−C +
a 2 / z − zs zz zs zz
3a 4  z −z  1 1  z2 D 
− 3 C 2 s 2
−  C 2
+  2
+ 2 
zz  (a / z − zs )  
a / z − zs zs a a / z − zs 
a4  2(zs − z ) a 2 1  1 1 z
+ C 2 −C 2 − 2C  2 +
2
zz  (a / z − zs )3 z 2 (a / z − zs )2  a / z − zs zs  a 2
1 D a2 
−C + 
(a 2 / z − zs )2 (a 2 / z − zs )2 z 2  (3.161b)

3.7 Line Forces or Line Dislocations Interacting with a Circular Inhomogeneity 93

Remark 3.48: If μ1 = ∞, we have an infinite plane of Material 2 with a rigid


circular inhomogeneity of radius a inside. For this case, c1* = 1 / κ 2 , c2* = κ 2 , c7* = 1,
and c8* = 0. The complementary solution (3.158) is reduced to

  1 1 z 
[2µ 2 eiθ (ur + iuθ )]2 c = C  2 +  z + D ln(a 2 / z − zs ) − C 2 s 
  a / z − zs z s  a / z − zs
z   1 1 1 a2  zs  1 a2 
− C  2 +  +C 2  1 −  − D 2 
κ 2   a / z − zs zs  2
(a / z − zs ) z z a / z − zs z 2 
C a2
− κ 2C ln(a 2 / z − zs ) −
zs z
1 a4  zs − z  1 1  z2 D 
− C 2 −C  2 +  2 + 2 
κ 2 z  (a / z − zs )
3 2  a / z − zs zs  a a / z − zs 
 (3.162a)

1   1 1 1 a2  zs  1 a2 
[σ rr + σθθ ]2 c / 2 = C +  +C  1 −  − D 2
κ 2   a / z − zs zs 
2 2 2
(a / z − zs ) z z a / z − zs z2 
1   1 1 1 a2  zs  1 a2 
+ C +  +C  1 −  − D 
κ 2   a 2 / z − zs zs  (a 2 / z − zs )2 z  z a 2 / z − zs z 2 
1   1 1 1 a2  z  1 a2 
[σ rr + iσ rθ ]2 c = C +  +C  1− s  −D 
κ 2   a 2 / z − zs zs  (a 2 / z − zs ) 2 z  z a 2 / z − zs z 2 
1   1 1 1 a2  zs  1 a2 
+ C +  +C  1 −  − D 2 
κ 2   a / z − zs zs 
2 2 2
(a / z − zs ) z z a / z − zs z 2 
1  1 a2 2 a4  zs  1 a2  zs 
− C 2 +C  1 −  − C 2  1 − 
κ 2  (a / z − zs ) z
2 2 3
(a / z − zs ) z 2 z (a / z − zs ) z 
2 z
1 a 2 zs 1 a4 2 a2 
+C −D +D 
(a 2 / z − zs )2 z 2 (a 2 / z − zs )2 z 3 a 2 / z − zs z 2 
1 a2 C a2
+ κ 2C +
a 2 / z − zs zz zs zz
3 a4  zs − z  1 1  z2 D 
+ C − C  +  + 
κ 2 zz  (a / z − zs )
3 2 2  a / z − zs zs  a
2 2 2
a / z − zs 
1 a4  2( zs − z ) a 2 1  1 1 z
− C −C − 2C  + 
κ 2 zz  (a / z − zs ) z
2 2 3 2 2
(a / z − zs ) 2  a / z − zs zs  a 2
2

1 D a2 
−C +
(a 2 / z − zs )2 (a 2 / z − zs )2 z 2 
 (3.162b)

3.7.2.2 Line Forces or Edge Dislocations Inside a Circular Inhomogeneity


Now, let us consider the case when a concentrated line force (f = fx+ify) or edge
dislocation (b = bx+iby) is applied at z = zs (|zs| < a) within the circular inhomogeneity
of radius a.
94 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

The complex functions in Eq. (3.129) are assumed to be

φ1 (z) = φ1 p (z) + φ1c (z), ψ 1 (z) = ψ 1 p (z) + ψ 1c (z) ( z < a)


(3.163)
φ2 (z) = φ2 p (z) + φ 2c (z), ψ 2 (z) = ψ 2 p (z) + ψ 2c (z) ( z > a)

where φ1p (z) and ψ 1p (z) are the “particular solution” given in Eqs. (3.87) and (3.88),
while φ2 p (z) and ψ 2 p (z) now read as

φ2 p (z) = C2 ln(z / a), ψ 2 p (z) = D2 ln(z / a) (3.164)

with C2 and D2 to be determined. The singularity behavior in the matrix at infinity


described by Eq. (3.164) is caused by the nonvanishing resultants of tractions along
the circular boundary (r = a), or the discontinuity in the displacement along the
boundary (England 1971).
We assume that the complementary functions φ1c (z) and ψ 1c (z) are analytic for
r < a, and φ2c (z) andψ 2c (z) are analytic for r > a. From the theory of complex variables
(the properties of the hat transformation), we know that φ1c (a 2 / z), ψ 1c (a 2 / z) will
be analytic for r > a, while φ2 c (a 2 / z) and ψ 2 c (a 2 / z) will be analytic for r < a. We
also note that φ1p (z) and ψ 1p (z) are not analytic at infinity, which, to make use of the
analytic continuation theory, may be rewritten as:

φ1 p (z) = A ln(z / a) + φ10 (z); ψ 1 p (z) = B ln(z / a) + ψ 10 (z) (3.165)

where φ10 (z) and ψ 10 (z) become analytic for r > a. Accordingly, φ10 (a 2 / z) and
ψ 10 (a 2 / z) are analytic for r < a.
The four complex functions (φ1c, ψ1c, φ2c, and ψ2c) as well as the two constants C2
and D2 are to be determined from the interface conditions (r = a) as specified in Eq.
(3.137), which, in view of Eqs. (3.163)–(3.165), can be written as

Γκ 1[ A ln(t / a)]+ + Γκ 1[φ10 (t )]+ + Γκ 1[φ1c (t )]+


− Γ [t A / t + B ln( t / a)]+ − Γ [tφ10 ′ (t ) + ψ 10 (t )]+ − Γ [tφ1′c (t ) + ψ 1c (t )]+
= κ 2 [C2 ln(t / a)] + κ 2 [φ2c (t )]− − [tC2 / t + D2 ln( t / a)]− − [tφ2′ c (t ) + ψ 2c (t )]−

[ A ln(t / a)]+ + [φ10 (t )]+ + [φ1c (t )]+ + [t A / t + B ln( t / a)]+


′ (t ) + ψ 10 (t )]+ + [tφ1′c (t ) + ψ 1c (t )]+
+ [tφ10
= [C2 ln(t / a)]− + [φ2 c (t )]− + [tC2 / t + D2 ln( t / a)]− + [tφ2′ c (t ) + ψ 2c (t )]−
 (3.166)

Because ln(z/a) is neither analytic for r < a nor for r > a, we need to eliminate it from
Eq. (3.166) in order to apply the analytic continuation theory later. Noticing tt = a 2 ,
we therefore obtain

Γκ 1 A + Γ B = κ 2C 2 + D2
(3.167)
A − B = C 2 − D2

from which, we get


3.7 Line Forces or Line Dislocations Interacting with a Circular Inhomogeneity 95

Γκ 1 + 1 Γ −1
C2 = A+ B≡C
κ2 + 1 κ2 + 1
(3.168)
Γκ 1 − κ 2 Γ + κ2
D2 = A+ B≡D
κ2 + 1 κ2 + 1

where C and D are given in Eq. (3.136). Then, Eq. (3.166) can be rearranged as
follows

Γκ 1[φ1c (t )]+ − Γ [t A / t ]+ + [tC / t ]− − Γ [tφ10 ′ (t ) + ψ 10 (t )]+ + [tφ2′ c (t ) + ψ 2c (t )]−


= κ 2 [φ2 c (t)]− − Γκ 1[φ10 (t)]+ + Γ [tφ1′c (t) +ψ 1c (t)]+
(3.169)
[φ1c (t )]+ + [t A / t ]+ − [tC / t ]− + [tφ10
′ (t ) + ψ 10 (t )]+ − [tφ2′ c (t ) + ψ 2 c (t )]−
= [φ2c (t)] − [φ10 (t)] − [tφ1′c (t) + ψ 1c (t)]+
− + 

where we have replaced C2 by C. The preceding equation can be further


rewritten as

{ ′ (a 2 / z) + ψ 10 (a 2 / z)]
lim Γκ 1φ1c (z) − Γ Az2 / a 2 + Cz2 / a 2 − Γ [zφ10
z→ t
z <a

+[zφ2′ c (a2 / z) + ψ 2 c (a2 / z)] − Γ zφ1′c (0) }


{
= lim κ 2φ2 c (z) − Γκ 1φ10 (z) + Γ [zφ1′c (a 2 / z) − zφ1′c (0) + ψ 1c (a 2 / z)]
z→ t
}
z >a
(3.170)
{ ′ (a 2 / z) + ψ 10 (a 2 / z)]
lim φ1c (z) + Az2 / a 2 − Cz2 / a 2 + [zφ10
z→ t
z <a

−[zφ2′ c (a 2 / z) + ψ 2c (a 2 / z)] + zφ1′c (0) }


{
= lim φ2 c (z) − φ10 (z) − [zφ1′c (a 2 / z) − zφ1′c (0) + ψ 1c (a 2 / z)]
z→ t
}
z >a

where the terms containing φ1′c (0) are added to eliminate the singularity at infinity.
The left-hand sides of Eq. (3.170) are analytic in Material 1 (r < a) and the right-hand
sides are analytic in Material 2 (r > a). By the analytic continuation theory, we have,
in Material 1 (r < a)

Γκ 1φ1c (z) − Γ Az2 / a 2 + Cz2 / a 2 − Γ [zφ10 ′ (a 2 / z) + ψ 10 (a 2 / z)]


+ [zφ2′ c (a 2 / z) + ψ 2c (a 2 / z)] − Γ zφ1′c (0) = 0
(3.171)
φ1c (z) + Az2 / a 2 − Cz2 / a 2 + [zφ10 ′ (a 2 / z) + ψ 10 (a2 / z)]
− [zφ2′ c (a / z) + ψ 2c (a / z)] + zφ1′c (0) = 0
2 2

and in Material 2 (r > a)

κ 2φ2c (z) − Γκ 1φ10 (z) + Γ [zφ1′c (a 2 / z) − zφ1′c (0) +ψ 1c (a 2 / z)] = 0


(3.172)
φ2 c (z) − φ10 (z) − [zφ1′c (a 2 / z) − zφ1′c (0) + ψ 1c (a 2 / z)] = 0

We can obtain from Eq. (3.171)


96 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

φ1c (z) = c1[zφ10 ′ (a 2 / z) + ψ 10 (a 2 / z) + zφ1′c (0) + Az2 / a 2 ]


zφ2′ c (a 2 / z) + ψ 2 c (a 2 / z) (3.173)
′ (a 2 / z) + ψ 10 (a 2 / z) + zφ1′c (0) + Az2 / a 2 ] − Cz2 / a 2
= c4 [zφ10

and from Eq. (3.172)

φ2 c (z) = c3φ10 (z)


(3.174)
zφ1′c (a 2 / z) − zφ1′c (0) + ψ 1c (a 2 / z) = c2φ10 (z)

where ci are material coefficients defined in Eq. (3.94), all related to the two Dundurs
parameters α and β. It can be seen that Eqs. (3.173) and (3.174) are very similar to
those for a bimaterial plane as presented in Eqs. (3.83) and (3.84).
From the first expression in Eq. (3.173), we can derive

c1 c
φ1′c (0) = 2
(M3 + c1M3 ), φ1′c (0) = 1 2 (c1M3 + M3 ) (3.175)
1 − c1 1 − c1

where M3 = (d[ψ 10 (a 2 / z)] / d z)z=0 = −( Bzs + Azs ) / a 2 . Then, we can obtain the four
complex functions from Eqs. (3.173) and (3.174) as

z2 c12
φ1c (z) = c1 g (a 2 / z) + c1 A + (c1M3 + M3 )z
a 2 1 − c12
a4 a2
ψ 1c (z) = c2φ10 (a 2 / z) + c1 3
g ′(a 2 / z) + c1M3 − 2c1 A (3.176)
z z
φ2 c (z) = c3φ10 (z)
c1 a2 a2
ψ 2 c (z) = c5 g(z) + c3ψ 10 (z) + (M3 + c1 M3 ) + (c4 A − C ) 2
1 − c1 z z

where we have introduced the following auxiliary function

a2
g(z) = ′ (z) + ψ 10 (z)
φ10 (3.177)
z

It is readily seen that the expressions in Eq. (3.176) are identical to those derived by
Honein and Herrmann (1990).
In terms of the particular solutions (3.87), the four complex functions in Eq.
(3.176) can be finally expressed as

z − zs  a 2 − zzs  c2 z
φ1c (z) = c1 A + c1B ln   − 1 2 [c1( Bzs + Azs ) + ( Bzs + Azs )] 2
a2 / z − zs  a  1 − c1 a
 a 2 − zzs   a2 zzs2 (z − zs )  zs2
ψ 1c (z) = c2 A ln  − c A  + 1 −  + c B
a 
1 2 1
  a − zzs (a 2 − zzs )2  a 2 − zzs (3.178)
φ 2 c (z) = c3 A ln(z − zs ) − c3 A ln(z / a)
z a2  a(z − zs )  z a2
ψ 2 c (z) = c5 A 2 s + c4 B ln   − c4 A s + (c4 A − C ) 2
z (z − zs )  z  z − zs z
c1 1
− [( Bzs + Azs ) + c1( Bzs + Azs )]
1 − c1 z

3.7 Line Forces or Line Dislocations Interacting with a Circular Inhomogeneity 97

Remark 3.49: If μ1 = μ2 and ν1 = ν2, then c1 = c2 = c5 = 0, c3 = c4 = 1, and A = C , B = D.


The complementary complex functions in the inhomogeneity (Material 1) are zero
while those in the matrix (Material 2) are reduced to

φ2 c (z) = A ln(z − zs ) − A ln(z / a)


(3.179)
ψ 2 c (z) = B ln(z − zs ) − B ln(z / a) − Azs / (z − zs )

The summation of Eqs. (3.179) and (3.164) gives the particular solutions (3.135).
Consequently, when substituting these complex functions into Eq. (3.21), the
displacement and stress fields obtained will be the same as those in the infinite plane
with the concentrated source at z = zs.
Remark 3.50: If μ2 = 0 (c1 = c2 = −1), we have a circular plate of Material 1 with
traction-free conditions at r = a. The complementary solution in the plate is

z − zs  (a 2 − zzs )  1 z
φ1c (z) = − A − B ln   − (Bzs + Azs ) 2
a2 / z − zs  a  2 a
(3.180)
 a 2 − zzs   a2 zz 2 (z − zs )  zs2
ψ 1c (z) = − A ln   + A 2 + 1 − 2s  − B
 a   a − zzs (a − zzs )2  a 2 − zzs

Note that when c1 = −1 the first expression in Eq. (3.173) gives φ1′c (0) + φ1′c (0) = − M3,
indicating that M3 = −( Bzs + Azs ) / a 2 should be real. This means that only the edge
dislocation case (B = A ) is allowed, for which the circular boundary could be kept
traction-free. In this case, instead of the first expression in Eq. (3.176), we will have

z2 M 3 z
φ1c (z) = − g (a 2 / z) − A + (3.181)
a2 2

Here we have actually mathematically proved that one cannot apply a concentrated
(unbalanced) line force to the interior of a circular plate while keeping its surface
traction free.
Remark 3.51: If μ2 =∞ (c1 = 1 / κ 1 , c2 = κ 1 ), we will have a circular plate of Material 1
with fixed boundary condition at r = a. The complementary solution in the plate is

A z − zs B  (a 2 − zzs )  [(Bzs + Azs ) + κ 1 (Bzs + Azs )] z


φ1c (z) = + ln −
κ 1 a 2 / z − zs κ 1  a  κ 1 (κ 12 − 1) a2
(3.182)
 a 2 − zzs  A  a 2 zz 2 (z − zs )  B zs2
ψ 1c (z) = κ 1 A ln  −  2 + 1 − 2s +
 a  κ 1  a − zzs (a − zzs )2  κ 1 a 2 − zzs

The displacement and stress fields can be obtained from the complex functions in
Eqs. (3.87), (3.164), and (3.178). Note that, the total solution in each phase is the
summation of the particular solution and the complementary one. The particular
solution in the inhomogeneity in terms of the rectangular coordinates is given in Eq.
(3.95), while in terms of the polar coordinates, it becomes
98 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

[2µ1eiθ (ur + iuθ )]1 p = κ 1 A ln(z − zs ) − A(z − zs ) / (z − zs ) − B ln(z − zs )


[σ rr + σθθ ]1 p / 2 = A / (z − zs ) + A / (z − zs )
(3.183)
[σ rr + iσ rθ ]1 p = A / (z − zs ) + A / (z − zs ) + Az / (z − zs )2
− (z / z)[ B / (z − zs ) + Azs / (z − zs )2 ]

For the particular solution in the matrix (ref. to Eq. (3.164)), we have,

[ 2µ 2 eiθ (ur + iuθ )]2 p = κ 2C ln(z / a) − Cz / z − D ln(z / a)


[σrr + σθθ ]2 p / 2 = C / z + C / z (3.184)
[σrr + iσrθ ]2 p = C / z + 2C / z − D / z

Using Eqs. (3.129) and (3.178), the complementary part in the inhomogeneity is
found to be

[ 2µ 1eiθ (ur + iuθ )]1c


 z(z − zs )  a 2 − zzs  c12 z 
= κ 1 c1 A 2 + c1 B ln   − 2
[c1 ( Bzs + Azs ) + ( Bzs + Azs )] 2 
 a − zzs  a  1 − c1 a 
 z a (z − zs ) 
2 z (z − zs )
− c1 A  2 + 2 2
z + c1 B s 2
 a − zz s (a − zz )
s  a − zzs
c 2
z
+ 1 2 [c1 ( Bzs + Azs ) + (Bzs + Azs )] 2
1 − c1 a
 a 2 − zzs   a2 zz2 (z − zs ) 
− c2 A ln  + c1 A  2 + 1 − 2s
 a 
  a − zzs (a − zzs )2 
 (3.185a)

 z a 2 (z − zs )  z
[σ rr + σθθ ]1c / 2 = c1 A  2 + 2 2
− c1 B 2 s
 a − zz s ( a − zzs )  a − zzs
2  z
c 1 a 2 (z − zs ) 
− 1 2 [c1 (Bzs + Azs ) + (Bzs + Azs )] 2 + c1 A  2 + 2 2
1 − c1 a  a − zzs (a − zzs ) 
z c2 1
− c1 B 2 s − 1 2 [c1 ( Bzs + Azs ) + ( Bzs + Azs )] 2
a − zzs 1 − c1 a
[σ rr + iσ rθ ]1c
 z a 2 (z − zs )  zs c12 1
= c1 A  2 + 2 2  − c1 B 2
− 2
[c1 (Bzs + Azs ) + (Bzs + Azs )] 2
 a − zzs (a − zzs )  a − zzs 1 − c1 a
 z 2
a (z − zs ) 
2 z c 1
+ c1 A  2 + 2 2
− c1 B 2 s − 1 2 [c1 ( Bzs + Azs ) + ( Bzs + Azs )] 2
 a − zz s ( a − zz s ) a − zz s 1 − c 1 a
 1 zzs a 2 (z − zs )zs  zzs2
− 2c1 A  2 + 2 2
+ 2 3  z + c1 B (a2 − zz )2
 a − zzs (a − zzs ) (a − zzs )  s

 z  a2 zs zs2 (2z − zs ) 2zzs3 (z − zs )  zs3  z


+ c2 A 2 s + c1 A  2 2
− 2 2
− 2 3  − c1 B 2 2 
 a − zzs  (a − zzs ) (a − zzs ) (a − zzs )  (a − zzs )  z

(3.185b)
3.7 Line Forces or Line Dislocations Interacting with a Circular Inhomogeneity 99

The total displacement and stress fields in the inhomogeneity (Material 1, r < a) are
the summation of those in Eqs. (3.183) and (3.185).
In the matrix (Material 2, r > a), the displacement and stress fields are found by
using Eqs. (3.129) and (3.178). The results are

[2µ 2 eiθ (ur + iuθ )]2 c


 a(z − zs )  zzs z a2  a(z − zs ) 
= c3κ 2 A ln   − c3 A − c5 A 2 s − c4 B ln  
 z  z (z − zs ) z (z − zs )  z
z a2 c 1
 + c4 A s − (c4 A − C ) 2 + 1 [( Bzs + Azs ) + c1 ( Bzs + Azs )] (3.186a)
z − zs z 1 − c1 z

zs zs
[σ rr + σθθ ]2 c / 2 = c3 A + c3 A
z(z − zs ) z (z − zs )
zs zs z (2z − zs )
[σ rr + iσ rθ ]2 c = c3 A + c3 A + c3 A s
z(z − zs ) z (z − zs ) z (z − zs )2
2
z a (3z − 2zs ) zs zzs (3.186b)
+ c5 A s 2 2
− c4 B − c4 A
zz (z − zs ) z(z − zs ) z(z − zs )2
a 2 c 1
+ 2(c4 A − C ) 2 − 1 [( Bzs + Azs ) + c1 ( Bzs + Azs )]
zz 1 − c1 zz

The total displacement and stress fields in the matrix (Material 2, r > a) are the
summation of those in Eqs. (3.184) and (3.186).
Remark 3.52: If μ1 = μ2 = μ and ν1 = ν2 = ν, then c1 = c2 = 0 . Thus, the complementary
parts of the displacement and stress fields in the inhomogeneity (Material 1) given
by Eq. (3.185) are zero, and the total solution is reduced to the particular solution
given by Eq. (3.183), which gives the Green’s displacements and stresses for an
infinite homogeneous plane subjected to a concentrated source at z = zs.
Remark 3.53: Also, if μ1 = μ2 = μ and ν1 = ν2 = ν, then c3 = c4 = 1, c5 = 0 and
C = A, D = B. The displacement and stress fields in the matrix (Material 2) given by
Eq. (3.186) are reduced to the following expressions

[2µeiθ (ur + iuθ )]2c


 a(z − zs )  zz − zzs  a(z − zs )  (3.187a)
= κ A ln   −A s − B ln  
 z  z (z − zs )  z

zs zs
[σ rr + σθθ ]2 c / 2 = A +A
z(z − zs ) z (z − zs )
zs zs z (2z − zs ) zs zs z
[σrr + iσrθ ]2c =A +A +A s −B −A
z(z − zs ) z (z − zs ) z (z − zs )2 z(z − zs ) z(z − zs )2

 (3.187b)
Adding Eq. (3.187) to Eq. (3.184) gives the solution of the displacement and stress
fields in an infinite plane with the source at z = zs, which is given in Eq. (3.183).
100 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

Remark 3.54: If μ2 = 0, we have a circular plate of Material 1 subjected to a traction-


free boundary condition at r = a. For this case, c1 = c2 = −1. The complementary
solution (3.185) is then reduced to

[ 2µ1eiθ (ur + iuθ )]1c


z(z − zs )  a 2 − zzs  κ 1 − 1 z
= −κ 1 A 2
− κ 1 B ln   − ( Bzs + Azs ) 2
a − zzs  a  2 a
 a 2 z(z − zs ) a 2 − zz zz2 (z − zs )  (3.188a)
+ A 2 − 2 − 1 + 2s 
 (a − zzs )
2 a − zzs (a − zzs )2 
z (z − zs )  a 2 − zzs 
− B s2 + A ln  
a − zzs  a

 z a 2 (z − zs )   z a 2 (z − zs ) 
[σ rr + σθθ ]1c / 2 = − A  2 + 2 2  − A  2
+ 2 2
 a − zzs (a − zzs )   a − zzs (a − zzs ) 
1 z z
− (Bzs + Azs ) 2 + B 2 s +B 2 s
a a − zzs a − zzs
[σ rr + iσ rθ ]1c
 z a 2 (z − zs )   z a 2 (z − zs ) 
= −A  2 + 2 2  − A 2 + 2 2
 a − zzs (a − zzs )   a − zzs (a − zzs ) 
1 z z
− ( Bzs + Azs ) 2 + B 2 s +B 2 s
a a − zzs a − zzs
 1 zz a (z − zs )zs 
2 zz2
+ 2A  2 + 2 s 2+ 2 3  z−B 2 s 2
 a − zzs (a − zzs ) (a − zzs )  (a − zzs )
 z  a 2 zs z2 (2z − zs ) 2zzs3 (z − zs )  z3  z
− A 2 s + A 2 2
− s2 2
− 2 3  −B 2 s 2 
 a − zzs  (a − zzs ) (a − zzs ) (a − zzs )  (a − zzs )  z
 (3.188b)

The complete solution is the summation of Eqs. (3.188) and (3.183). We point out
that the solution is only meaningful when the interior of the circular plate is under
the line or edge dislocation (referring to Remark 3.50).
Remark 3.55: If μ2 = ∞, we have a circular plate of Material 1 subjected to a rigid
boundary condition at r = a. For this case, c1 = 1 / κ 1 , c2 = κ 1 . The complementary
solution (3.185) is reduced to

[2µ1eiθ (ur + iuθ )]1c


z(z − zs )  a2 − zzs  ( Bzs + Azs ) z
=A + B ln  −
a2 − zzs a  κ1 a2
A z a 2 (z − zs )  B zs (z − zs ) (3.189a)
−  + 2  z+
κ 1  a − zzs (a − zzs ) 
2 2 κ 1 a 2 − zzs
 a 2 − zzs  A  a 2 zzs2 (z − zs ) 
− κ 1 A ln  +  + 1 −
 a 
 κ 1  a 2 − zzs (a 2 − zzs )2 
3.8 Applications of Bimaterial Line Force/Dislocation Solutions 101

A z a 2 (z − zs )  A  z a 2 (z − zs ) 
[σ rr + σθθ ]1c / 2 = + + +
κ 1  a 2 − zzs (a 2 − zzs )2  κ 1  a 2 − zzs (a 2 − zzs )2 
B zs B zs (Bzs + Azs ) + ( Bzs + Azs ) 1
− − −
κ 1 a 2 − zzs κ 1 a 2 − zzs κ 1 (κ 1 − 1) a2
A z a 2 (z − zs )  A  z a 2 (z − zs ) 
[σrr + iσrθ ]1c = + 2 + +
κ 1  a − zzs (a − zzs )2  κ 1  a 2 − zzs (a2 − zzs )2 
 2  

B zs B zs ( Bzs + Azs ) + ( Bzs + Azs ) 1


− − −
κ 1 a − zzs κ 1 a − zzs
2 2 κ 1 (κ 1 − 1) a2
2A  1 zz a 2 (z − zs )zs  B zzs2
−  + 2 s 2+ 2  z+
κ 1  a − zzs (a − zzs )
2 (a − zzs ) 3 κ 1 (a − zzs )2
2

 z A  a 2 zs zs2 (2z − zs ) 2zzs3 (z − zs )  B zs3  z


+ κ 1 A 2 s +  − −  − 
 a − zzs κ 1  (a − zzs )
2 2 2
(a − zzs ) 2 (a − zzs )  κ 1 (a − zzs )  z
2 3 2 2

 (3.189b)

The complete solution is the summation of Eqs. (3.189) and (3.183).

3.8 Applications of Bimaterial Line Force/Dislocation Solutions

3.8.1 Image Force of a Line Dislocation


In this section, we will first evaluate the image force or the Peach-Koehler force
(abbreviated as PK force hereafter) acting on the dislocation (with Burgers vector
bi ) defined as (Peach and Koehler 1950; Dundurs 1968)

Fi = −ε ijk bl t jσ kl (3.190)

where ε ijk is the permutation tensor (or Levi-Civita symbol), ti is the unit tangential
vector on the point of interest on the dislocation, and σ ij is the total stress at that
point. For a straight dislocation parallel to the z-axis and oriented in the positive z-
axis, we can compute from the preceding equation

Fx = bxσ xy + byσ yy + bzσ yz


Fy = −bxσ xx − byσ xy − bzσ xz (3.191)
Fz = 0

This indicates that for the straight dislocation parallel to the z-axis, the PK force is
confined in the (x, y)-plane though the dislocation has all three components (one
screw and two edges). Note that the PK force alternatively can be computed from
the change of interaction energy (Eshelby 1951; Dundurs 1968).

3.8.1.1 PK Force on a Screw Dislocation in a Bimaterial Plane


For a screw dislocation located at the source point (xs, ys), we have bx = by = 0,
bz = b ≠ 0. Thus, Eq. (3.191) becomes

Fx = bσ yz , Fy = −bσ xz ( x = xs , y = ys ) (3.192)
102 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

Note that the limit of σ ij should be taken by approaching the source point (xs, ys)
along its normal that is perpendicular to the particular component of the force being
computed (Dundurs 1968), in which the singular part corresponding to the full plane
should also be removed. Then, according to Eq. (3.71), we can obtain

µ1 b 2 K
Fx = 0, Fy = (3.193)
4π ys

where K is the antiplane Dundurs parameter defined just below Eq. (3.105). Equation
(3.193) shows that the image force is always normal to the interface between
Material 1 and Material 2. Note that a positive (negative) Fy indicates that Material
2 repels (attracts) the dislocation. Thus, for Γ > 1, we have Fy > 0 and the dislocation
is repelled from the interface, while for Γ < 1, we have Fy < 0 and the dislocation is
pulled to the interface. In other words, the dislocation always tends to move away
from (approach) the material with the higher (lower) shear modulus.

3.8.1.2 PK Force on a Screw Dislocation Interacting with a Circular


Inhomogeneity
Without loss of generality, we assume that the screw dislocation is always located on
the x-axis. We first consider the interior dislocation case, that is, when the dislocation
is inside the circular inhomogeneity (rs < a). Then, according to Eq. (3.118), we
obtain

µ 1bK rs / a
σ(xz1) = 0, σ(yz
1)
=− ( x = xs , y = ys ) (3.194)
2 πa 1 − (rs / a)2

Thus,

µ1 b 2 K rs / a
Fx = − , Fy = 0 (3.195)
2π a 1 − (rs / a)2

This shows that the image force is always along the radial direction. In addition, the
sign of Fx still depends on the Dundurs parameter K only. For a stiff inhomogeneity,
we have Γ < 1, K < 0, and Fx > 0, indicating that the dislocation tends to move toward
the interface. For a soft inhomogeneity, Γ > 1, K > 0, Fx < 0, and the dislocation tends
to move toward the center of the inhomogeneity.
When the dislocation is at the point (rs,0) outside the inhomogeneity (rs>a), we
obtain from Eq. (3.124) that

( 2) ( 2) µ 2 bK  1 1
σ xz = 0, σ yz =−  −  (3.196)
2π  rs − (a / rs ) rs 
2

Then one obtains

µ 2 b2 K 1
Fx = − , Fy = 0 (3.197)
2π a (rs / a)[(rs / a)2 − 1]
3.8 Applications of Bimaterial Line Force/Dislocation Solutions 103

The image force is again along the radial direction, and its sign is determined by
the Dundurs parameter K. For a stiff inhomogeneity, we have Γ < 1, K < 0, and
Fx > 0. This indicates that the inhomogeneity repels the dislocation. While for a
soft inhomogeneity, Γ > 1, K > 0, Fx < 0, and the dislocation tends to move toward
the inhomogeneity. It should be noted that rs represents the x-coordinate of the
dislocation, and can be negative in Cartesian coordinates. In the preceding, we have
assumed tacitly that rs > 0. For rs < 0, the conclusion keeps unchanged.
Remark 3.56: The formula of the PK force on a screw dislocation near a straight
interface, Eq. (3.193), can be obtained either from Eq. (3.195) or (3.197) by letting
a → ∞, and ys = rs .
Remark 3.57: The PK force on a screw dislocation near a free or fixed surface can be
readily obtained by letting Γ = 0 or Γ → ∞. For example, when the screw dislocation
is within a circular inhomogeneity (rs < a), and if the outside matrix is absent, we
have Γ = 0 and K = −1. Then from Eq. (3.195), one obtains

µ 1b 2 rs / a
Fx = , Fy = 0 (3.198)
2π 1 (rs / a)2
a −

This formula shows that Fx is larger than zero, and the dislocation tends to approach
the free surface of the inhomogeneity.

3.8.1.3 PK Force on an Edge Dislocation in a Bimaterial Plane


For an edge dislocation located at the source point (xs, ys > 0) in Material 1, we have
bz = 0, and one or both of bx and by are nonzero. Thus, Eq. (3.191) becomes

Fx = bx σ xy + by σ yy
Fy = −bx σ xx − by σ xy
( x = xs , y = ys ) (3.199)

Because the expressions for stresses for an edge dislocation with both components
given in Eqs. (3.95) (for the particular part) and (3.96) (for the complementary part)
are rather lengthy, we here consider, for simplicity, two separated cases, that is, when
the Burgers vector is either normal or parallel to the interface.
If the Burgers vector of the edge dislocation is normal to the interface, we have
bx = 0, and Eq. (3.199) further reduces to

Fx = by σ yy , Fy = −by σ xy ( x = xs , y = ys ) (3.200)

Meanwhile, Eq. (3.96b) gives

(c1 + c2 )µ1by 1
[σ yy ( xs , ys )]1c = 0, [σ xy ( xs , ys )]1c = − (3.201)
π (1 + κ 1 ) 2 ys

wherein the particular part in Eq. (3.95), which is singular at (xs, ys) and doesn’t
contain the parameter ys, has been omitted (Dundurs 1968). Thus, we obtain the PK
force as
104 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

(c1 + c2 )µ1by2
Fx = 0, Fy = (3.202)
2π (1 + κ 1 ) ys

Equation (3.202) indicates that, when the Burgers vector of the dislocation is
perpendicular to the interface, the PK force is also perpendicular to that interface.
If the Burgers vector of the edge dislocation is parallel to the interface, we have
by = 0, and Eq. (3.199) becomes

Fx = bx σ xy , Fy = −bx σ xx ( x = xs , y = ys ) (3.203)

We can obtain from Eq. (3.96b)

(c1 + c2 )µ1bx 1
[σ xx ( xs , ys )]1c = 0, [σ xy ( xs , ys )]1c = (3.204)
π (1 + κ 1 ) 2 ys

The PK force is calculated as

(c1 + c2 )µ1bx2
Fx = , Fy = 0 (3.205)
2π (1 + κ 1 ) ys

Thus, when the Burgers vector of the dislocation is parallel to the interface, the PK
force is also parallel to that interface.
Remark 3.58: The expression for Fy in Eq. (3.202) is identical to that for Fx in (3.205)
except for the directions of Burgers vector and the image force. The sign of Fy or Fx
thus solely depends on the sign of c1 + c2 (with fixed ys > 0). According to Eq. (3.94),
one has

2(α + β 2 )
c1 + c2 = (3.206)
1− β 2

Because 1 − β 2 > 0 (see Figure 3.3 for the possible values of the two Dundurs
parameters α and β), the behavior of the edge dislocation depends only on the
value of α + β 2 (Dundurs 1968). Consequently, when the Burgers vector of the
dislocation in Material 1 (ys > 0) is perpendicular to the interface, there exist three
possibilities on the dislocation mobility: (1) the dislocation is attracted toward the
interface if α + β 2 < 0; (2) the dislocation is repelled away from the interface if
α + β 2 > 0; and (3) the dislocation is in equilibrium if α + β 2 = 0. Similar conclusions
can be drawn when the vertical edge dislocation is located in Material 2 (ys < 0).
When the Burgers vector of the dislocation is parallel to the interface, we have the
following three possibilities on the dislocation mobility: (1) the dislocation tends
to move in the negative x-direction if α + β 2 < 0; (2) the dislocation tends to move
in the positive x-direction if α + β 2 > 0; and (3) the dislocation is in equilibrium if
α + β 2 = 0.
3.8 Applications of Bimaterial Line Force/Dislocation Solutions 105

(1,0.5)

κ1=3, κ2=1 Γ=∞


α

Γ=0 κ1=1, κ2=3

(–1,–0.5)
Figure 3.3. The (α, β)-plane. A point in the shaded region represents the possible values that
α, β may take (Dundurs 1968).

3.8.1.4 PK Force on an Edge Dislocation Interacting with a Circular


Inhomogeneity
We first consider the case when the edge dislocation is at the point (xs,ys) = (rs,0)
outside the circular inhomogeneity (rs > a) with its Burgers vector along the x-
direction. The PK force is then calculated according to Eq. (3.203). From Eq. (3.158),
we obtain

µ2 bx a 2  (c1* + c2* ) 1
[σ xy (rs , 0)]2 c = [σ rθ (rs , 0)]2 c =  + (3c1* − c2* ) 3 
π (1 + κ 2 )  rs (rs2 − a 2 ) rs  (3.207)
[σ xx (rs , 0)]2c = [σ rr (rs , 0)]2c =0

Thus,

µ 2 bx2 a 2  (c1* + c2* ) 1


Fx =  + (3c1* − c2* ) 3  , Fy = 0 (3.208)
π(1 + κ 2 )  rs (rs − a )
2 2 rs 

This shows that the image force is always along the radial direction.
When the Burgers vector of the edge dislocation at (rs,0) is along the y-direction,
the PK force should be determined from Eq. (3.200). In this case, we obtain from
Eq. (3.158) that

[σ yy (rs , 0)]2c = [σ θθ (rs , 0)]2c = 0


µ2 by a 2  (c1* + c2* ) 1 (3.209)
[σ xy (rs , 0)]2c = [σ rθ (rs , 0)]2c = −  + (c1* + c7* ) 3 
π (1 + κ 2 )  rs (rs − a )
2 2 rs 

Then one obtains

µ2 by2 a 2  (c1* + c2* ) 1


Fx = 0, Fy = + (c1* + c7* ) 3  (3.210)
π (1 + κ 2 )  rs (rs2 − a 2 ) rs 

The image force is found to be parallel to the Burgers vector direction.


Remark 3.59: Both PK force expressions, Fx in Eq. (3.208) and Fy in Eq. (3.210), are
complex. However, we can discuss two particular cases that will enable us to get
some insight into the behavior of the dislocation in an analytical manner.
106 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

The first case is when the dislocation is very close to the inhomogeneity, for
which we have from Eqs. (3.208) and (3.210) respectively

µ2 bx2 a 2 (c1* + c2* ) 2 µ2 bx2 α * + ( β * )2 a2


Fx = = (by = 0)
π (1 + κ 2 ) rs (rs − a ) π (1 + κ 2 ) 1 − (β ) rs (rs − a 2 )
2 2 * 2 2
   (3.211)
µ2 by2 a 2 (c1* + c2* ) 2 µ2 by2 α * + ( β * )2 a2
Fy = = (bx = 0)
π (1 + κ 2 ) rs (rs2 − a 2 ) π (1 + κ 2 ) 1 − ( β * )2 rs (rs2 − a 2 )

We point out: (1) Equation (3.206) is also valid when the quantities in it are all
replaced with those with a star. (2) The two expressions in Eq. (3.211) are essentially
identical to Eqs. (3.205) and (3.202) by noting that rs = a + d , where d  a represents
the distance to the interface. Thus, the discussion presented in Remark 3.58 also
applies here.
The second case is when the dislocation is far away from the inhomogeneity. For
this case, we have

4 µ2 bx2 a 2 α * − β * 1
Fx = (by = 0)
π (1 + κ 2 ) 1 + β * rs3
(3.212a,b)
4 µ2 by2 a 2 β* 1
Fy = (bx = 0)
π (1 + κ 2 ) 1 + α − 2 β rs3
* *

Because 1 + β * > 0 and 1 + α * − 2 β * > 0 (see Figure 3.3, and notice that α * = −α ,
β * = − β ), the behavior of this distant edge dislocation is determined by the signs
of α * − β * in (3.212a) and β * in (3.212b). It is easily observed that if α * − β * > 0
(α * − β * < 0), the dislocation on the x-axis with Burgers vector b = (bx , 0) will
be repelled away from (attracted to) the inhomogeneity. Thus, there will be an
intermediate point where the PK force on the dislocation becomes zero (i.e., α* = β*).
Discussion on the mobility of the edge dislocation in the y-direction can be carried
out similarly.
Now let’s consider the case when the edge dislocation is at the point (xs,ys) =
(rs,0) inside the circular inhomogeneity (rs < a) with its Burgers vector along the
x-direction. The PK force is determined by Eq. (3.203). From Eq. (3.185), we obtain

µ1bx (c1 + c2 )rs


[σ xy (rs , 0)]1c = [σ rθ (rs , 0)]1c = −
π (1 + κ 1 ) (a2 − rs2 ) (3.213)
[σ xx (rs , 0)]1c = [σ rr (rs , 0)]1c = 0

Accordingly,

µ1bx2 (c1 + c2 )rs


Fx = − , Fy = 0 (3.214)
π (1 + κ 1 ) (a 2 − rs2 )

The PK force is also perpendicular to the interface, and its behavior depends on the
sign of c1 + c2 or α + β 2 (see Eq. (3.206)).
When the Burgers vector of the inside edge dislocation is along the y-direction,
the PK force is determined by Eq. (3.200). We can obtain from Eq. (3.185)
3.8 Applications of Bimaterial Line Force/Dislocation Solutions 107

[σ yy (rs , 0)]1c = [σ θθ (rs , 0)]1c = 0


µ1by  (c1 + c2 )rs 4c12 rs  (3.215)
[σ xy (rs , 0)]1c = [σ rθ (rs , 0)]1c =  2 2 + 
π (1 + κ 1 )  (a − rs ) 1 − c1 a 2 

Thus, one obtains

µ1by2  (c1 + c2 )rs 4c12 rs 


Fx = 0, Fy = −  2 2 +  (3.216)
π (1 + κ 1 )  (a − rs ) 1 − c1 a 2 

Remark 3.60: When the dislocation is very near to the interface, Eqs. (3.214) and
(3.216) reduce to

µ1bx2 (c1 + c2 )rs


Fx = − (by = 0)
π(1 + κ 1 ) (a 2 − rs2 )
(3.217)
µ1by2 (c1 + c2 )rs
Fy = − (bx = 0)
π(1 + κ 1 ) (a 2 − rs2 )

These again agree essentially with Eqs. (3.205) and (3.202) by noting that rs = a − d ,
where d  a represents the distance to the interface. By contrast, when the dislocation
is in the vicinity of the center of the inhomogeneity, we have

µ1bx2 (c1 + c2 )rs 2µ1bx2 rs α + β2


Fx = − = − (by = 0)
π(1 + κ 1 ) (a 2 − rs2 ) π(1 + κ 1 ) (a 2 − rs2 ) 1 − β 2
µ 1by 
2 4c1  rs
2
Fy = −  c1 + c2 + (3.218)
π(1 + κ 1 )  1 − c1  a 2
µ 1by2 rs (1 + α)[α(1 − 2β) + 3β 2 ]
=− (bx = 0)
π(1 + κ 1 ) a 2 (1 − β 2 )(1 − α + 2β)

Hence, the behavior of the dislocation near the center of the inhomogeneity will be
determined by the signs of α + β 2 and α (1 − 2 β ) + 3β 2 . For example, if α + β 2 > 0,
then Fx is always directed toward the center, indicating that the center is a stable
equilibrium position for the dislocation with Burgers vector b = (bx , 0).

3.8.2 Image Work of Line Forces


The analytical results of interactions between dislocations and inhomogeneities
presented in the last subsection are crucial to the understanding of some
micromechanisms in materials and also to the development of materials with desired
mechanical properties (strength, ductility, etc.). The analysis can be extended to the
interactions between applied forces and inhomogeneities as well. We define the
image work as

E = fi uic (3.219)
108 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

where fi are the components of force applied at the source point, and uic are
the displacement components at the source point excluding the singular part
corresponding to the full plane, that is, only the image part or the complementary part
of the solution would be considered. This subsection will give a simple illustration
for the antiplane case only, and the readers can readily deal with more complicated
cases by following the same idea. In the antiplane case, Eq. (3.219) reads

E = fuc (3.220)

where f is the magnitude of the out-of-plane line force, and uc is the complementary
antiplane displacement induced at the source point.

3.8.2.1 Image Work on an Antiplane Line Force in a Bimaterial Plane


According to Eq. (3.58), where the force is of unit magnitude, we can obtain

Kf 2
E= ln(2 ys ) (3.221)
2πµ1

It is shown that, no matter what kind of the material combination is, there will be a
fixed equilibrium point, that is, ys = 1 / 2 , where the image work is zero.

3.8.2.2 Image Work on an Antiplane Line Force Interacting with a Circular


Inhomogeneity
We first consider that the source is within the inhomogeneity at the point (xs,ys) =
(rs,0) with rs < a. From Eq. (3.105), where the force is also of unit magnitude, one
obtains

Kf 2  a 2 − rs2 
E= ln   (3.222)
2πµ1  a 

It can be shown that, when the source point is very close to the interface, Eq. (3.222)
degenerates into Eq. (3.221) identically.
When the source is in the matrix, that is, outside the inhomogeneity, at the point
(xs,ys) = (rs,0) with rs > a, we obtain from Eq. (3.111) that

Kf 2  rs2 − a 2 
E=− ln   (3.223)
2πµ2  rs 

When the source point is very close to the interface, Eq. (3.223) again agrees
essentially with Eq. (3.221).

3.9 Summary and Mathematical Keys

3.9.1 Summary
This chapter presents several typical Green’s function solutions due to a concentrated
line force or dislocation acting in an elastic isotropic plane. The plane can be fully
homogeneous, or composed of two half-planes or a circular inhomogeneity embedded
3.9 Summary and Mathematical Keys 109

in an infinite homogeneous plane. For the inhomogeneous cases, perfect bonding


between the two phases is assumed. Both antiplane and in-plane 2D problems are
investigated. Various degenerate situations are also discussed, including the half-
plane under a free or fixed boundary condition, and an infinite plane with a circular
opening hole or a rigid circular plate. Green’s displacements and stresses have also
been derived explicitly, which may be used as an important reference.
The Green’s function solutions presented in this chapter are very useful.
For example, the Green’s solutions of force can be used as kernel functions in
the method of boundary integral equation (BIE) or boundary element method
(BEM). The Green’s solutions of dislocation can be used to construct the BIE for
dealing with cracks of general configurations. In particular, the Green’s solutions
for inhomogeneous material systems are important for understanding various
mechanisms (such as plastic deformation, strengthening, and hardening) occurring
in materials. These are discussed in Section 3.8 by presenting the image force (or
work) on the dislocation (or force) induced by the interaction between dislocations
(forces) and inhomogeneities.

3.9.2 Mathematical Keys


This chapter derives the Green’s function solutions for concentrated line force/
dislocation under antiplane and in-plane 2D deformation. For the homogeneous plane
problem, the trial-and-error method has been employed based on the mathematical
knowledge (i.e., the potential theory) in Chapter 1. For the inhomogeneous plane
problem, the method of image, which has also been discussed in Chapter 1, has been
adopted to construct the Green’s function solutions in the antiplane deformation
case. The complex variable method along with the analytical continuation has
been used to analyze the in-plane deformation, which is very powerful especially
for problems with complex configurations. The idea of this method is to construct
analytical functions in a domain outside the definition domain of known analytical
functions through the continuity conditions at an interface (or boundary conditions
at a boundary). To do this, the singular part of the known functions, except that
containing removable singularities, should be carefully pretreated. This has been well
demonstrated in Section 3.7.2.2 for the problem of a line force/dislocation applied
inside the circular inhomogeneity.
The image force on the dislocation has been computed according to the PK
formula in Eq. (3.190). This force can also be computed from the interaction energy
Eint as

∂Eint ∂Eint
Fx = − , Fy = − (3.224)
∂xs ∂ys

in which, when external forces are absent,

1
2 ∫S
Eint = bi n j σ ij d S (3.225)

where S signifies the dislocation cut, and ni is a unit normal on S+ and pointing

toward S . The calculation based on Eq. (3.224) seems more involved than that using
110 Green’s Functions in Elastic Isotropic Full and Bimaterial Planes

Eq. (3.190). However, Eq. (3.224) tells us that the image force physically equals the
negative gradient of the interaction energy with respect to the coordinate of the
dislocation.

3.10 References
Dundurs, J. 1968. Elastic interaction of dislocation with inhomogeneities. In: Mathematical
Theory of Dislocation (Ed. Mura, T.), pp. 70–115. New York: American Society of
Mechanical Engineers.
England, A. H. 1971. Complex Variable Methods in Elasticity. London: Wiley-Interscience.
Eshelby, J. D. 1951. The force on an elastic singularity. Philosophical Transactions of the Royal
Society, A 244 (877): 87–112.
Honein, T. and Herrmann, G. 1990. On bonded inclusions with circular or straight boundaries
in plane elastostatics. Journal of Applied Mechanics 57 (4): 850–6.
Love, A. E. H. 1944. A Treatise on the Mathematical Theory of Elasticity. 4th ed. New York:
Dover Publications.
Muskhelishvili, N. I. 1953. Some Basic Problems of the Mathematical Theory of Elasticity.
Groningen, Holland: P. Noordhoff Ltd.
Peach, M. D. and Koehler, J. S. 1950. The force exerted on dislocations and the stress field
produced by them. Physical Review 80: 436–9.
Suo, Z. 1989. Singularities interacting with interfaces and cracks. International Journal of
Solids and Structures 25 (10): 1133–42.
4 Green’s Functions in
Magnetoelectroelastic Full and
Bimaterial Planes

4.0╇ Introduction
Most advanced materials are elastically anisotropic and many of them are even
smart in the sense that stimulus in one field would induce response in the other
field. For instance, in piezoelectric materials, an applied mechanical force would
induce an electric field, and an applied electric voltage would induce a mechanical
deformation. In this chapter, by applying the extended Stroh formalism (Ting 1996),
we solve the extended line-force and line-dislocation problems in infinite, half, and
bimaterial planes. While our solutions for the half-plane case include many different
surface boundary condition cases, our bimaterial solution contains various imperfect
interface cases. All these solutions are expressed in a very elegant simple form.
Briefly given is also the relation of the solutions between the line force and line
dislocation, a result that was presented in Chapter€ 2 when we studied the Betti’s
reciprocal theorem. As applications of the solutions, various quantum-wire problems
from current semiconductor industry are analyzed.

4.1╇ Generalized Plane-Strain Deformation


Our 2D problem will be in the (x1, x3)-plane, and it is under the assumption that
all quantities are independent of the x2-variable (i.e., ∂()/∂x2=0). Therefore, the
Green’s functions presented are rigorously for the generalized 2D plane-strain case.
Furthermore, the reason that we select the (x1, x3)-plane for the 2D problem is that
our solutions for the half-plane and bimaterial plane cases would correspond to
those in 3D, with the surface and interface being all at x3 = 0 in both 3D and 2D, as
already mentioned in Section 2.8.
It is noted that the line-dislocation Green’s function cannot be derived from the
line-force Green’s function by taking the derivatives, as stated in Chapter€2 based
on the Betti’s reciprocal theorem. Here for the 2D case, the line-dislocation Green’s
function and the line-force Green’s function have the same order of singularity at
the source point. The reason is that the line dislocation is obtained from the point
dislocation by integrating over the dislocation surface, instead of line integral as in
the line-force case. We define the following extended displacement and two stress
vectors,€as

111
112 Green’s Functions in Magnetoelectroelastic Full and Bimaterial Planes

u = (u1 , u2 , u3 , φ ,ψ )t
t1 = (σ 11 , σ 12 , σ 13 , D1 , B1 )t (4.1)
t 3 = (σ 31 , σ 32 , σ 33 , D3 , B3 )t

and assume that the extended displacement vector is proportional to a complex


function€as

u = a f (z) (4.2)

where a = (a1, a2, a3, a4, a5)t is a complex vector, and f(z) is a complex function of
variable z = x1+px3 with p to be determined in the following text. Note that the
superscript t denotes matrix transpose. The extended displacement vector, which is
real, can be taken as either the real part or the imaginary part of the right-hand side
of Eq. (4.2). If we take the real part, we have u = Re[af(z)]; if we take the imaginary
part, we then have u = Im[af(z)]. Thus, the operation on the extended displacement
vector u can be carried out just over af(z) on the right-hand side of Eq. (4.2). After
that, one can simply use 2Re[*] or 2Im[*] over the results [*] to find the physical
quantities. We point out that the factor 2 is due to the fact that the associated
eigenvalues and eigenvectors in Eq. (4.2) always appear as complex conjugate pairs
and therefore we can use only half of these eigensolutions (Ting 1996). This will be
seen clearly later in this chapter.
Substituting Eq. (4.2) into Eq. (2.2a) and then Eq. (2.1) in Chapter€ 2 for the
2D case in the (x1, x3)-plane, one arrives at the following eigenrelation for the
eigenvalue€p and the corresponding eigenvector€a:

[Q + p( R + R t ) + p2T ] a = 0 (4.3)

with

QIK = c1IK 1 , RIK = c1IK 3 , TIK = c3 IK 3 (4.4)

where ciJKl are the elastic, electric, and magnetic coefficients defined in Eq. (2.4e) in
Chapter€2.
The two vectors t1 and t3 in Eq. (4.1) can be expressed in terms of the derivatives
of the complex function f(z)€as

t1 = (Q + pR)af ′(z)
(4.5)
t 3 = ( Rt + pT)af ′(z)

We now introduce the extended “stress” function vector€φ,€as

t1 = − φ,3 , t 3 = φ,1 (4.6)

and also define a new vector€b€as

1
b = ( Rt + pT )a = − (Q + p R)a (4.7)
p
4.2 Solutions of Line Forces and Line Dislocations in a 2D Full-Plane 113

Denoting by pm, am, and bm (m = 1–10) the Stroh eigenvalues and the associated
eigenvectors of Eqs. (4.3) and (4.7), we then order them in such a way so that

Im pJ > 0 , pJ + 5 = pJ , a J + 5 = a J , bJ + 5 = bJ ( J = 1 − 5) (4.8)
A = [a1 , a2 , a3 , a4 , a5 ] B = [b1 , b2 , b3 , b4 , b5 ]

where an overbar denotes the complex conjugate. We have also assumed that pm are
distinct and the eigenvectors am, and bm satisfy the normalization relation (Barnett
and Lothe 1975; Ting 1996)

bm
t a + at b = δ
n m n mn (4.9)

with δmn being the 10×10 Kronecker delta (i.e., the 10×10 identity matrix). Then, in
terms of the real values, the general solutions of the extended displacement and
stress function vectors can be written as

u = 2 Im[ A < f (z* ) > q]


(4.10)
φ = 2 Im[ B < f (z* ) > q]

where the matrices A and B are eigenvector matrices defined in Eq. (4.8), q is the
unknown complex vector to be determined for the given problem, and <f(z*)> is the
diagonal matrix, related to the eigenvalues and the coordinates, defined as

< f (z* ) >= diag[ f (z1 ), f (z2 ), f (z3 ), f (z4 ), f (z5 )] (4.11)

where zJ = x1+pJx3.
Remark 4.1: The corresponding formulations for anisotropic elastity were first
developed by Stroh (see Barnett and Lothe 1975), and are frequently referred to as
Stroh formalism in literature (Ting 1996).
Remark 4.2: The case of repeated eigenvalues pJ can be avoided by using slightly
perturbed material coefficients with negligible errors (Pan and Amadei 1996;
Pan 1997). In so doing, the simple structure of the solutions (4.10) can always be
employed.

4.2 Solutions of Line Forces and Line Dislocations in a 2D Full-Plane

4.2.1 General Solutions of Line Forces and Line Dislocations


We assume that there is a uniformly distributed (extended) line force f (per unit
length) along the x2-axis, and a uniformly distributed line dislocation with (extended)
Burgers vector b along the x2-axis, then the jump conditions across these sources are
(see Figure 4.1 and Eqs. (4.18) and (4.19) in the following text)

∆ u ≡ u( r , π ) − u( r , − π ) = b
(4.12)
∆ φ ≡ φ(r , π ) − φ(r , − π ) = f
114 Green’s Functions in Magnetoelectroelastic Full and Bimaterial Planes

x3

x2
x1
o
f3

Figure€ 4.1a.╇ A general line force (f3) applied along the x2-axis, with the associated
branch€cut.

x3
x2

x1
o
b1

Figure€4.1b.╇ A line dislocation (b1) applied along the x2-axis, with the associated branch€cut.

The Green’s function solutions due to these line sources can be assumed€as

1
u= Im[ A < ln(z* ) > q ∞ ]
π (4.13)
1
φ = Im[ B < ln(z* ) > q ∞ ]
π

where q∞ is an unknown complex vector to be determined, and the complex function


ln(z) takes the value in its single-valued branch defined€as

−π < Im[ln z] < π (4.14)

Making use of the force and dislocation discontinuity conditions in Eq. (4.12), we
find€that

2 Re[ Aq ∞ ] = b, 2 Re[ Bq ∞ ] = f (4.15)

Or

 A A  q ∞   b 
 B B   q ∞  =  f  (4.16)

Therefore, the solutions for q∞€are

 q∞   Bt At   b 
q ∞  =  B t At   f 
(4.17)

which is derived from Eq. (4.16) by invoking the normalization relation (4.9). It
is noted that the line-force and line-dislocation solutions have the same order of
singularity, which is different from their corresponding 3D counterparts. This is
explained physically in Figure€4.1, and mathematically in Box€4.1.
4.2 Solutions of Line Forces and Line Dislocations in a 2D Full-Plane 115

Box 4.1. Mathematical relations between the line source and point
source
The line - force solution

= ∫ [The point - force solution ]dx2 ;
−∞
The line - dislocation solution
0 ∞
=∫ dx1 ∫ [The point - dislocation solution ]dx2 .
−∞ −∞

Actually, the singularities of a line force and line dislocation are described as
(Figure 4.1)

f ( x1 , x3 ) = f δ ( x1 ) δ ( x3 )
Δt 3 ( x1 , x3 ) = − f δ ( x1 ) (4.18)
Δ φ = − ∫ Δ t 3 dS = f

u( x1 , x3 ) = b δ ( x3 ) H( − x1 )
b δ ( x3 ) ( x1 <0) (4.19)
u( x1 , x3 ) = 
0 ( x1 >0)

where H(x) is the Heaviside function.


Remark 4.3: The line dislocation and 2D displacement discontinuity are physically
different. Thus, in order to find the induced fields by a 2D displacement discontinuity,
one needs to apply the Betti’s reciprocity (2.22) to a 2D plane with bj being the
displacement discontinuities (and the surface integral being the line integral along
the displacement discontinuity). The solutions due to Δu in Eq. (4.19) are strictly
those due to the line dislocation, not due to the 2D displacement discontinuity!

4.2.2 Green’s Functions of a Line Force


We now consider only the line force in Eq. (4.17), and let the line-force vector
f t= (f1, f2, f3, −fe, −fh) equal to (1,0,0,0,0), (0,1,0,0,0), (0,0,1,0,0), (0,0,0,1,0), and (0,0,0,0,1).
Then the line-force Green’s functions can be written in terms of their components
(J is the component of the displacement, stress function, and traction vectors at
x(x1,x3) and K is the direction of the force at y(y1,y3))

5
1
uJK ( x; y) = Im ∑ [ AJR ln(zR − sR ) AKR ]
π R=1
5
(4.20a, b)
1
φK
J ( x; y) = Im ∑ [ BJR ln(zR − sR ) AKR ]
π R=1

5
1 p n − n3
t JK ( x; y) = − Im ∑ [ BJR R 1 AKR ] (4.21)
π R=1 zR − sR
116 Green’s Functions in Magnetoelectroelastic Full and Bimaterial Planes

In these equations, “Im” again stands for the imaginary part of the complex value;
AIJ and BIJ are two eigenmatrices related to the MEE material property; n1 and n3
(functions of x) are the unit outward normal components along the x1- and x3-directions;
pR (R = 1–5) are the Stroh eigenvalues, and zR = x1+pRx3 and sR = y1+pRy3 are related to
the field x(x1,x3) and source y(y1,y3) points, respectively. These displacements, stress
functions, and tractions are required in the conventional boundary integral equation
formulation to solve the general boundary value problems in MEE solids. In order
to find the elastic strain, and electric and magnetic fields, one only needs to take
the derivative of the Green’s displacement (4.20a) with respect to the field point x.
The derivatives of the Green’s functions with respect to the source point y can be
also found, which may be useful in other applications (i.e., higher-order boundary
integral equations for fracture analysis, as in Pan (1999)). The corresponding stresses,
electric displacements, and magnetic inductions can be obtained through the general
MEE constitutive relations (2.2) in Chapter 2.
The traction Green’s functions (4.21) with the arbitrary pair of outward normal
(n1,n3) can be found using the definition of the extended tractions (4.5)

t = t1n1 + t 3 n3
(4.22)
= [(Q + pR)n1 + ( Rt + pT )n3 ]af ′(z)

which can be also written as

t = t1n1 + t 3 n3 = −[ B ( pn1 − n3 )] f ′(z) (4.23)

In Eqs. (4.22) and (4.23),

1
f (z) = Im[< ln z* > q ∞ ] (4.24)
π

It is noted that the second expression in Eq. (4.22) is just from Eq. (4.5), and that in
deriving (4.23), use has been made of Eqs. (4.7) and (4.8) for the Stroh vector b and
matrix B. Making use of Eq. (4.17) with b = 0 and letting the line force be in different
directions (K = 1 to 5), we then find the extended traction Green’s functions from
Eq. (4.23) as given by Eq. (4.21).
An alternative way to find t JK is to directly take the derivatives of the stress
function in Eq. (4.20b) (with respect to the field point x) and make use of

t JK ≡ σ αKJ nα = σ 1KJ n1 + σ 3KJ n3


∂φ JK ∂φ JK (4.25)
=− n1 + n3
∂x3 ∂x1

In Eq. (4.25), σ αKJ represents the Green’s stress σαJ due to a line force in the K-
direction. This equation gives all the stress components once the stress function φJK is
found. While the stresses can be found by taking the derivative of the stress function
vector, the strains can be found by taking the derivative of the displacement Green’s
functions uJK . The results of the derivatives with respect to the field point x are
4.3 Green’s Functions of Line Forces and Line Dislocations in a Half-Plane 117

1 5
 1 
uJK,1 ( x; y) = Im ∑  AJR AKR 
π R=1  zR − sR  (4.26)
1 5
 pR 
uJK, 3 ( x; y) = Im ∑  AJR AKR 
π R=1  zR − sR 

4.2.3 Green’s Functions of a Line Dislocation


The Green’s functions (including the displacement, stress function, and traction)
at x due to a generalized line dislocation (Burgers vector, electric and magnetic
potential discontinuities) bt = (Δu1, Δu2, Δu3, Δφ, Δψ) of unit magnitude at y (i.e., with
bt = (1, 0, 0, 0, 0), (0, 1, 0, 0, 0), (0, 0, 1, 0, 0), (0, 0, 0, 1, 0), (0, 0, 0, 0, 1), respectively)
can be found as

5
1
uJK ( x; y) = Im ∑ [ AJR ln(zR − sR )BKR ]
π R=1 (4.27a, b)
5
1
φJK ( x; y) = Im ∑ [ BJR ln(zR − sR )BKR ]
π R=1

5
1  p n − n3 
t JK ( x; y) = − Im ∑  BJR R 1 BKR  (4.28)
π R=1  zR − sR 

The derivatives of the displacements are

1 5
 1 
uJK,1 ( x; y) = Im ∑  AJR BKR 
π R=1  zR − sR 
(4.29)
1 5
 pR 
u ( x; y) = Im ∑  AJR
K
J ,3 BKR 
π R=1  zR − sR 

Comparing Eqs. (4.27)–(4.28) with Eqs. (4.20)–(4.21), we notice that the line-force
and line-dislocation Green’s functions are very similar to each other and that they
have the same order of singularity. In order to find the line-dislocation solution,
one needs only to replace the second matrix A in Eqs. (4.20), (4.21), and (4.26) by
matrix B.

4.3 Green’s Functions of Line Forces and Line Dislocations


in a Half-Plane
The half-plane can be the upper (x3 > 0) or the lower (x3 < 0) half-plane. The solutions
are in a unified form and they are the same as presented in the following text. We
assume that the extended traction vector is zero on the surface of the half-plane
(x3 = 0).
118 Green’s Functions in Magnetoelectroelastic Full and Bimaterial Planes

4.3.1 Green’s Functions of a Line Force in a Half-Plane


The half-plane Green’s functions for the displacements, stress functions, and tractions
(the J-th component) with outward normal n1 and n3 (at x) due to a line force at y
with component K can be expressed as (similar to those in Pan (1999) and Jiang and
Pan (2004))

5  5 
uJK ( x; y) =
1  v ]
Im ∑  AJR ln(zR − sR ) AKR + ∑[ AJR ln(zR − sv )QRK 
π 
R=1  v= 1 
5  5  (4.30a, b)
1  v ]
φK
J ( x; y ) = Im ∑  BJR ln(zR − sR ) AKR + ∑[ BJR ln(zR − sv )QRK 
π 
R =1  v= 1 

5 
1  p n − n3 5
p n − n3 v 
t JK ( x; y) = − Im ∑  BJR R 1 AKR + ∑[ BJR R 1 QRK ] (4.30c)
π 
R=1  zR − sR v= 1 zR − sv 

where

5 5
QRN
v
= − ∑ ∑ [ BRS
−1
B SP ( I v ) PP ANP ] (4.31)
S =1 P =1

with

I 1 = diag[1, 0, 0, 0, 0], I 2 = diag[0, 1, 0, 0, 0]


(4.32)
I 3 = diag[0, 0, 1, 0, 0], I 4 = diag[0, 0, 0, 1, 0], I 5 = diag[0, 0, 0, 0, 1]

The derivatives of the displacements are

5 
1  1 5
 1  
uJK,1 ( x; y) = Im ∑  AJR AKR + ∑  AJR QRK
v

π 
R=1  zR − sR v= 1  zR − sv   (4.33a, b)
5 
1  pR 5
 pR  
uJK,3 ( x; y) = Im ∑  AJR AKR + ∑  AJR QRK
v

π 
R= 1  zR − sR v=1  zR − sv  

Other extended stress components can be found by taking derivatives of the stress
function in Eq. (4.30b) with respect to the field coordinate x.

4.3.2 Green’s Functions of a Line Dislocation in a Half-Plane


Similarly, the half-plane Green’s functions for the displacements (and derivatives),
stress functions, and tractions (the J-th component) with outward normal n1 and n3
(at x) due to the line dislocation at y with component K can be expressed as

5  5 
1  v ]
uJK ( x; y) = Im ∑  AJR ln(zR − sR )BKR + ∑[ AJR ln(zR − sv )QRK 
π 
R=1  v=1 
5  5  (4.34a, b)
1  v ]
φK
J ( x; y ) = Im ∑  BJR ln(zR − sR )BKR + ∑[ BJR ln(zR − sv )QRK 
π 
R=1  v= 1 
4.3 Green’s Functions of Line Forces and Line Dislocations in a Half-Plane 119

5 
1  p n − n3 5
p n − n3 v 
t JK ( x; y) = − Im ∑  BJR R 1 BKR + ∑[BJR R 1 QRK ] (4.34c)
π 
R=1  zR − sR v= 1 zR − sv 

5 
1  1 5
 1  
uJK,1 ( x; y) = Im ∑  AJR BKR + ∑  AJR QRK
v

π 
R=1  zR − sR v=1  zR − sv   (4.35a, b)
5 
1  pR 5
 pR  
u ( x; y) = Im ∑  AJR
K
J ,3 BKR + ∑  AJR QRK
v

π 
R=1  zR − sR v=1  zR − sv  

where

5 5
QRN
v
= ∑ ∑ [ BRS
−1
B SP ( I v ) PP BNP ] (4.36)
S=1 P =1

As for the line-force case, other extended stress components can be found by
taking derivatives of the stress function in Eq. (4.34b) with respect to the field
coordinate x.

4.3.3 Green’s Functions of Line Forces and Line Dislocations


in a Half-Plane under General Boundary Conditions
Similar Green’s function expressions can be obtained for the general boundary
conditions on the surface of the anisotropic MEE half-plane. These solutions
can be found by following Pan (2002) as presented in the following text. The
corresponding elastic half-plane case was discussed in Ting and Wang (1992) and
Pan (2002).
We first write the general boundary conditions on the surface of the half-plane
by a simple vector equation that is similar to the purely elastic counterpart (Ting and
Wang 1992)

Iu u + I t t3 = 0 (4.37)

where Iu and It are 5×5 diagonal matrices whose five diagonal elements are either
one or zero, and satisfy the conditions

Iu + I t = I , Iu I t = 0 (4.38)

with I being the unit matrix.


Equation (4.37) under the constraints (4.38) includes a total of twenty-five
different boundary condition sets, among which the following twelve sets are
of particular interests. They are listed as follows in terms of the components of
the extended displacements u (u1,u2,u3,u4,u5) and traction t3 (t1,t2,t3,t4,t5, equally
for the components of the extended stress function vector φ in place of the
components of t3)
120 Green’s Functions in Magnetoelectroelastic Full and Bimaterial Planes

t1 = 0 , t2 = 0 , t3 = 0 , t4 = 0 , t5 = 0
t1 = 0 , t2 = 0 , t3 = 0 , u4 = 0 , u5 = 0
t1 = 0 , t2 = 0 , t3 = 0 , t4 = 0 , u5 = 0
t1 = 0 , t2 = 0 , t3 = 0 , u4 = 0 , t5 = 0
u1 = 0 , u2 = 0 , u3 = 0 , t4 = 0 , t5 = 0
u1 = 0 , u2 = 0 , u3 = 0 , u4 = 0 , u5 = 0
(4.39)
u1 = 0 , u2 = 0 , u3 = 0 , t4 = 0 , u5 = 0
u1 = 0 , u2 = 0 , u3 = 0 , u4 = 0 , t5 = 0
t1 = 0 , t2 = 0 , u3 = 0 , t4 = 0 , t5 = 0
t1 = 0 , t2 = 0 , u3 = 0 , u4 = 0 , u5 = 0
t1 = 0 , t2 = 0 , u3 = 0 , t4 = 0 , u5 = 0
t1 = 0 , t2 = 0 , u3 = 0 , u4 = 0 , t5 = 0

For these boundary conditions, the diagonal matrices Iu and It take the following
diagonal elements:

Iu = diag[0, 0, 0, 0, 0] , I t = diag[1, 1, 1, 1, 1]
Iu = diag[0, 0, 0, 1, 1] , I t = diag[1, 1, 1, 0, 0]
Iu = diag[0, 0, 0, 0, 1] , I t = diag[1, 1, 1, 1, 0]
Iu = diag[0, 0, 0, 1, 0] , I t = diag[1, 1, 1, 0, 1]
Iu = diag[1, 1, 1, 0, 0] , I t = diag[0, 0, 0, 1, 1]
Iu = diag[1, 1, 1, 1, 1] , I t = diag[0, 0, 0, 0, 0]
(4.40)
Iu = diag[1, 1, 1, 0, 1] , I t = diag[0, 0, 0, 1, 0]
Iu = diag[1, 1, 1, 1, 0] , I t = diag[0, 0, 0, 0, 1]
Iu = diag[0, 0, 1, 0, 0] , I t = diag[1, 1, 0, 1, 1]
Iu = diag[0, 0, 1, 1, 1] , I t = diag[1, 1, 0, 0, 0]
Iu = diag[0, 0, 1, 0, 1] , I t = diag[1, 1, 0, 1, 0]
Iu = diag[0, 0, 1, 1, 0] , I t = diag[1, 1, 0, 0, 1]

It is apparent that the first three values in each row of Eq. (4.39) correspond to the
mechanical boundary conditions: The first four equations are for traction-free, the
fifth to eighth for rigid, and the last four for slippery. The fourth and fifth values
in each row of Eq. (4.39) are for the electric and magnetic boundary conditions,
respectively. For example, the first row in Eq. (4.39) represents the electrically and
magnetically insulating, and the second the electrically and magnetically conducting,
and so forth. To show a clear connection between the names of the different boundary
conditions and the physical quantities involved, these sets of boundary conditions
are further listed in Table 4.1.
To solve the half-plane problem under the general boundary conditions (4.37),
we introduce the following new complex matrix K of 5×5 as

K = Iu A + I t B (4.41)

which is a linear combination of the Stroh eigenmatrices A and B that are properly
coupled with the boundary conditions. For instance, for the first (traction-free,
4.3 Green’s Functions of Line Forces and Line Dislocations in a Half-Plane 121

Table 4.1. Physical Meanings of the Twelve MEE Boundary Conditions (E = electric, M =
magnetic, cond. = conducting, ins. = insulating)

Name of B.C.s Mechanical Electric Magnetic

Traction-free, E & M insulating σ31=σ32=σ33=0 D3=0 B3=0


Traction-free, E & M conducting σ31=σ32=σ33=0 φ=0 ψ=0
Traction-free, E ins. & M cond. σ31=σ32=σ33=0 D3=0 ψ=0
Traction-free, E cond. & M ins. σ31=σ32=σ33=0 φ=0 B3=0
Rigid, E & M insulating u1= u2=u3=0 D3=0 B3=0
Rigid, E & M conducting u1=u2=u3=0 φ=0 ψ=0
Rigid, E ins. & M cond. u1=u2=u3=0 D3=0 ψ=0
Rigid, E cond. & M ins. u1=u2=u3=0 φ=0 B3=0
Slippery, E & M insulating σ31=σ32=0, u3=0 D3=0 B3=0
Slippery, E & M conducting σ31=σ32=0, u3=0 φ=0 ψ=0
Slippery, E ins. & M cond. σ31=σ32=0, u3=0 D3=0 ψ=0
Slippery, E cond. & M ins. σ31=σ32=0, u3=0 φ=0 B3=0

electrically and magnetically insulating), the sixth (rigid, electrically and magnetically
conducting), and the eleventh (slippery, electrically insulating and magnetically
conducting) boundary conditions in Eq. (4.39), the matrix K has, respectively, the
following expression:

 B11 B12 B13 B14 B15 


 B21 B22 B23 B24 B25 
K fii =  B31 B32 B33 B34 B35  (4.42)
 B41 B42 B43 B44 B45 
 B51 B52 B53 B54 B55 

 A11 A12 A13 A14 A15 


 A21 A22 A23 A24 A25 
K rcc =  A31 A32 A33 A34 A35  (4.43)
 A41 A42 A43 A44 A45 
 A51 A52 A53 A54 A55 

 B11 B12 B13 B14 B15 


 B21 B22 B23 B24 B25 
K sic =  A31 A32 A33 A34 A35  (4.44)
 B41 B42 B43 B44 B45 
 A51 A52 A53 A54 A55 

where the three subscripts to K are introduced to indicate the corresponding


boundary conditions of the mechanical, electric, and magnetic quantities. Therefore,
Eq. (4.42) stands for the traction-free (the first index f), electrically insulating (the
second index i), and magnetically insulting (the third index i); Eq. (4.43) stands for the
rigid (the first index r), electrically conducting (the second index c), and magnetically
conducting (the third index c); and Eq. (4.44) stands for the slippery (the first index
s), electrically insulating (the second index i), and magnetically conducting (the third
index c).
122 Green’s Functions in Magnetoelectroelastic Full and Bimaterial Planes

For a half-plane with the general boundary conditions as described by Eq. (4.37),
we can assume the solution as

5
1 1
u= Im[ A < ln(z* − s* ) > q ∞ ] + Im ∑ [ A < ln(z* − sv ) > qv ]
π π v= 1 (4.45)
5
1 1
φ = Im[ B < ln(z* − s* ) > q ∞ ] + Im ∑ [ B < ln(z* − sv ) > qv ]
π π v=1

On the surface x3 = 0, we have

5
1 1
u= Im[ A < ln( x1 − s* ) > q ∞ ] + Im ∑ [ A ln( x1 − sv )qv ]
π π v=1 (4.46)
5
1 1
φ = Im[ B < ln( x1 − s* ) > q ∞ ] + Im ∑ [ B ln( x1 − sv )qv ]
π π v=1

Replacing the first terms by their equivalent forms, that is, the negative values of
their complex conjugates, and also making use of the following relation,

5
< ln( x1 − s* ) >= ∑ [ln( x1 − sv ) I v ] (4.47)
v =1

the boundary condition on the surface can be expressed as

 5 
Im  ∑ ln( x1 − sv )(− KI v q ∞ + Kqv ) = 0 (4.48)
 v=1 

Therefore, the unknown vector qv can be found as, for the general boundary condition
(4.37) on the surface of the half-plane x3 = 0,

qv = K −1 KI v q ∞ (4.49)

or

qv = (Iu A + I t B)−1 (Iu A + I t B) I v q ∞ (4.50)

For the line-force case, we let the line-force vector f t = (f1, f2, f3, −fe, −fh) in Eq.
(4.17) equal to (1,0,0,0,0), (0,1,0,0,0), (0,0,1,0,0), (0,0,0,1,0), and (0,0,0,0,1), while
b = 0. Then we obtain the extended Green’s function solution (in matrix form
with two indices) of the extended line force. Similarly, letting the line-dislocation
vector bt in Eq. (4.17) equal to (1,0,0,0,0), (0,1,0,0,0), (0,0,1,0,0), (0,0,0,1,0), and
(0,0,0,0,1), with f = 0 gives us the extended Green’s functions of the extended
line dislocations.
4.4 Green’s Functions of Line Forces and Line Dislocations in Bimaterial Planes 123

With the solved Green’s displacement and stress functions in Eq. (4.45), one
can take their derivatives (with respect to the field coordinate x) to obtain the
corresponding Green’s strain and stress components.

4.4 Green’s Functions of Line Forces and Line Dislocations


in Bimaterial Planes

4.4.1 General Green’s Functions of Line Forces and Line Dislocations


in Bimaterial Planes
Due to the relative locations of the source and field points, there are four sets of
Green’s functions for the bimaterial case. We assume that Materials 1 and 2 occupy
the half-planes x3 > 0 and x3 < 0, respectively. Let us again assume that a line force
f t = (f1, f2, f3, −fe, −fh) or line dislocation bt = (Δu1, Δu2, Δu3, Δφ, Δψ) is applied at the
source point (y1, y3) in one of the half-planes. To derive the Green’s functions, it
is sufficient to find the displacement vector u and the traction vector t due to the
line force or line dislocation, which are presented in the following text for different
combinations of the source and field points. We point out that the solution presented
is valid when the source is in either the upper or lower half-plane.
We assume that the source point (y1, y3) is in the half-plane of Material λ (λ = 1
or 2). Then if the field point (x1, x3) is in the source plane (i.e., the half-plane of
Material λ), the displacement, stress function, and traction vectors can be expressed
as (Ting 1992; Jiang and Pan 2004):

5
1 1
u( λ ) = Im  A( λ ) 〈 ln(z*(λ ) − s*(λ ) )〉q ∞,λ  + Im ∑  A( λ ) 〈 ln(z*(λ ) − sJ( λ ) )〉q(Jλ ) 
π   π  
J =1
5
1 1
Im  B( λ ) 〈 ln(z*λ − s*λ )〉q ∞, λ  + Im ∑  B( λ ) 〈 ln(z*(λ ) − sJ( λ ) )〉q(Jλ ) 
( ) ( )
φ( λ ) =
π   π   (4.51)
J =1

1  p( λ ) n1 − n3  1 5  p λ n1 − n3 (λ ) 
( )
t ( λ) = −
π
Im  B( λ ) 〈 *( λ ) ( λ ) 〉q ∞,λ  −
z* − s* π
Im ∑  B( λ ) 〈 *( λ )
z* − sJ(λ )
〉q J 
  J =1 
 

If the field point (x1, x3) is in the other half-plane of Material μ (μ≠λ) (λ, μ = 1 or 2),
then the displacement, stress function, and traction vectors can be expressed as:

5
1
u(µ ) = Im ∑  A(µ ) 〈 ln(z*(µ ) − s(Jλ ) )〉q(Jµ ) 
π  
J =1
5
1
Im ∑  B (µ ) 〈 ln(z*µ − sJλ )〉qJµ 
( ) ( ) ( )
φ(µ ) = (4.52)
π  
J =1
5  (µ ) 
1 p n1 − n3
t (µ ) = − Im ∑  B(µ ) 〈 *(µ ) ( λ ) 〉q(Jµ ) 
π J =1 
 z* − s J 

In Eqs. (4.51) and (4.52), the superscripts (λ) and (μ) denote the quantities associated
with Materials 1 and 2; p(Jλ ) , A(λ), and B(λ) (λ = 1 and 2 for the two half-planes) are the
124 Green’s Functions in Magnetoelectroelastic Full and Bimaterial Planes

Stroh eigenvalues and the corresponding eigenmatrices as given before. Also in Eqs.
(4.51) and (4.52), we have defined:

〈 ln(z*λ − s*λ )〉 = diag[ln(z1λ − s1λ ), ln(z2λ − s2λ ), ln(z3λ − s3λ ),


( ) ( ) ( ) ( ) ( ) ( ) ( ) ( )

ln(z4λ − s4λ ), ln(z5λ − s5λ )]


( ) ( ) ( ) ( )

(λ) (λ) (λ) (λ) (λ) (λ) (λ) (4.53)


〈 ln(z*λ − sJ )〉 = diag[ln(z1 − sJ ), ln(z2 − sJ ), ln(z3 − sJ ),
( )

ln(z4(λ ) − sJ(λ ) ), ln(z5(λ ) − sJ(λ ) )]

where z(Jλ ) and s(Jλ ) (λ = 1, 2) are the complex variables associated with the field and
source points, respectively. They are defined as:

z(Jλ ) = x1 + p(Jλ ) x3 , s(Jλ ) = y1 + p(Jλ ) y3 (4.54)

We further observe that the first term in Eq. (4.51) corresponds to the full-plane
Green’s functions in Material λ with (see Eq. (4.17))

q ∞,λ = ( A( λ ) )t f (4.55)

for the line-force case, and

q ∞,λ = ( B ( λ ) )t b (4.56)

for the line-dislocation case.


The second term in Eq. (4.51) and the term in Eq. (4.52) are the complementary
parts of the Green’s function solutions. The complex vectors q(Jλ ) (λ = 1,2; J = 1–5) in
Eq. (4.51) and q(Jµ ) (μ = 1,2; J = 1–5) in Eq. (4.52) are determined using the continuity
conditions along the interface of the two half-planes, which will be discussed in the
following text.

4.4.2 Green’s Functions of Line Forces and Line Dislocations in Bimaterial


Planes under Perfect Interface Conditions
We first assume the perfect interface conditions along the interface x3 = 0 of the two
half-planes. That is, along the interface, we have

u(λ ) = u(µ ) , φ(λ ) = φ(µ ) (4.57)

In terms of the solutions (4.51) and (4.52), we obtain as in Ting (1996),

5
1 1
Im  A( λ ) 〈 ln( x1 − s*(λ ) )〉q ∞, λ  + Im ∑  A( λ ) ln( x1 − sJ( λ ) )q(Jλ ) 
π   π  
J =1
5
1
= Im ∑  A(µ) ln( x1 − s(Jλ ) )q(Jµ) 
π  
J =1
5 (4.58)
1 1
Im  B( λ ) 〈 ln( x1 − s*(λ ) )〉q ∞,λ  + Im ∑  B( λ ) ln( x1 − sJ( λ ) )q(Jλ ) 
π   π  
J =1
5
1
= Im ∑  B( µ) ln( x1 − s(Jλ ) )q(Jµ) 
π  
J =1
4.4 Green’s Functions of Line Forces and Line Dislocations in Bimaterial Planes 125

Notice that

1 1
Im  A(λ ) 〈 ln( x1 − s*(λ ) )〉q ∞,λ  = − Im  A(λ ) 〈 ln( x1 − s*(λ ) )〉q ∞,λ 
π   π  
(4.59)
1 1
Im  B(λ ) 〈 ln( x1 − s*(λ ) )〉q ∞,λ  = − Im  B(λ ) 〈 ln( x1 − s*(λ ) )〉q ∞,λ 
π   π  

and that

5
〈 ln( x1 − s*(λ ) )〉 = ∑ [ln( x1 − sv(λ ) ) I v ] (4.60)
v =1

we then have (taking also the conjugate of the right-hand side and multiplying by
“–1” to be equal)

5 5
1 1
− Im ∑  A(λ ) ln( x1 − sJ(λ ) ) I J q ∞,λ  + Im ∑  A(λ ) ln( x1 − sJ(λ ) )q(Jλ ) 
π   π  
J =1 J =1
5
1
= − Im ∑  A(µ ) ln( x1 − sJ( λ ) )qJ(µ ) 
π  
J =1
5 5 (4.61)
1 1
− Im ∑  B( λ ) ln( x1 − sJ ) I J q ∞, λ  + Im ∑  B( λ ) ln( x1 − sJ )qJ 
(λ) (λ) (λ)
π   π  
J =1 J =1
5
1
= − Im ∑  B(µ ) ln( x1 − sJ( λ ) )q J(µ ) 
π  
J =1

In Eqs. (4.60) and (4.61), IJ (J = 1–5) are the diagonal matrices defined in Eq. (4.32).
Therefore, we finally have the following two matrix equations for the unknown
vectors q(λ,μ) (and also for J = 1–5)

A(λ ) q(Jλ ) + A( µ ) q J( µ ) = A(λ ) I J q ∞,λ


(4.62)
B(λ ) q(Jλ ) + B( µ ) q J( µ ) = B(λ ) I J q ∞,λ

The two unknown vectors can be solved as (λ, μ = 1 or 2, but μ ≠ λ):

q(Jλ ) = ( A(λ ) )−1 ( M ( λ ) + M ( µ ) )−1 ( M ( µ ) − M ( λ ) ) A(λ ) I J q ∞,λ (4.63)

for the unknowns in Eq. (4.51), and

q(Jµ ) = ( A( µ ) )−1 ( M (λ ) + M ( µ ) )−1 ( M (λ ) + M (λ ) ) A( λ ) I J q ∞,λ (4.64)

for the unknowns in Eq. (4.52).


In Eqs. (4.63) and (4.64), the matrix M (α ) is the impedance tensor defined as:

M (α ) = − iB(α ) ( A(α ) )−1 (α = 1 , 2) (4.65)

For the line-force case, we let the line-force vector f t = (f1, f2, f3, −fe, −fh) in Eq. (4.55)
equal to (1,0,0,0,0), (0,1,0,0,0), (0,0,1,0,0), (0,0,0,1,0), and (0,0,0,0,1), while b = 0 in
126 Green’s Functions in Magnetoelectroelastic Full and Bimaterial Planes

Eq. (4.56). Then we obtain the extended Green’s function solutions in Eqs. (4.51)
and (4.52) (in matrix form with two indices) due to the extended line force. Similarly,
letting the line-dislocation vector bt in Eq. (4.56) equal to (1,0,0,0,0), (0,1,0,0,0),
(0,0,1,0,0), (0,0,0,1,0), and (0,0,0,0,1), while f = 0 in Eq. (4.55) gives us the extended
Green’s functions due to the extended line dislocations. The strain and other stress
components can be obtained by taking the derivatives of these Green’s functions
with respect to the field point x.

4.5 Green’s Functions of Line Forces and Line Dislocations in


Bimaterial Planes under General Interface Conditions
We only take two of imperfect interface models as examples to show the process
of deriving the solution. Similar Green’s function expressions can be derived for
the bimaterial plane with general (or imperfect) interface conditions. Detailed
discussion can be found in Pan (2003a, 2003b).
The first imperfect interface model is a simple one where we assume that the
imperfect interface conditions along the interface x3 = 0 between the two half-
planes are

u(Jλ ) = u (Jµ) , φ(Jλ ) = φ(Jµ) ( J = 1 − 4)


(4.66)
u5( λ ) = u5(µ ) =0

In other words, along the interface, the elastic displacements, stress tractions, electric
potential, and the normal electric displacement are continuous, while the magnetic
potential is zero on both sides of the interface.
Following the same process for the perfect interface case, we still find that, for
row I = 1–4 of each equation set, we have (i.e., consider only the first four equations
in each)

A(λ ) q(Jλ ) + A( µ ) qJ( µ ) = A(λ ) I J q ∞,λ


(4.67)
B(λ ) q(Jλ ) + B( µ ) q J( µ ) = B(λ ) I J q ∞,λ

The fifth rows I = 5 of the two sets of equations are from the second sets of equations
in Eq. (4.66) where u5 = 0 along the interface. We have (only for the fifth row of
each set)

A(λ ) q(Jλ ) = A(λ ) I J q ∞,λ


(4.68)
A( µ ) qJ( µ ) = 0

These two equations can be equivalently written as (by simple addition and
subtraction of them)

A(λ ) q(Jλ ) + A( µ ) qJ( µ ) = A(λ ) I J q ∞,λ


(4.69)
A(λ ) q(Jλ ) − A( µ ) qJ( µ ) = A(λ ) I J q ∞,λ
4.5 Green’s Functions under General Interface Conditions 127

Therefore, the two unknown vectors q(λ,μ) (and also for J = 1–5) can be solved from
the following two sets of matrix equations, which are symbolically similar to Eq.
(4.62) for the perfect interface case, but with the newly defined B matrices in the two
materials (matrix A are still the same as in the perfect interface case)

A( λ ) q (Jλ ) + A(µ) q J(µ) = A( λ ) I J q ∞, λ


(µ) (λ) (4.70)
 ( λ)q( λ) + B
B  
q J(µ) = B I J q ∞, λ
J

where

(λ)
 B11 B12 B13 B14 B15 
 B21 B22 B23 B24 B25 
 (λ )
B =  B31 B32 B33 B34 B35 
 B41 B42 B43 B44 B45 
 A51 A52 A53 A54 A55 
(µ ) (4.71)
 B11 B12 B13 B14 B15 
 B21 B22 B23 B24 B25 
 (µ)
B =  B31 B32 B33 B34 B35 
 B41 B42 B43 B44 B45 
 − A51 − A52 − A53 − A54 − A55 

The solutions to Eq. (4.70) have the same form as Eqs. (4.63) and (4.64), but with the
newly defined B matrices in Eq. (4.71). Thus, the bimaterial Green’s functions for
this simple imperfect interface conditions are completely solved.
The second imperfect interface model is relatively complicated where the
imperfect interface conditions along the interface x3 = 0 are as follows: The two half-
planes are in smooth contact so that the normal displacement is continuous and the
interface shear stresses are zero, while the electric and magnetic potentials are also
zero. Namely

u3( λ ) = u3(µ) , φ(3λ ) = φ(3µ)


φ(αλ ) = φ(αµ ) = 0 (α = 1, 2 ) (4.72)
(λ) (µ )
uJ = uJ =0 ( J = 4, 5)
Then, the following equations hold for the first two rows of each vector relations as
follows

B(λ ) q(Jλ ) = B(λ ) I J q ∞,λ


(4.73)
B( µ ) qJ( µ ) = 0

For the third rows in both equations in the following text, we have

A(λ ) q(Jλ ) + A( µ ) q J( µ ) = A(λ ) I J q ∞,λ


(4.74)
B(λ ) q(Jλ ) + B( µ ) q J( µ ) = B(λ ) I J q ∞,λ
128 Green’s Functions in Magnetoelectroelastic Full and Bimaterial Planes

Finally, for the fourth and fifth rows in both equations as follows, we have

A(λ ) q(Jλ ) = A(λ ) I J q ∞,λ


(4.75)
A( µ ) qJ( µ ) = 0

It is observed that, for the first two rows, the relations (4.73) can be expressed
equivalently as

B(λ ) q(Jλ ) + B( µ ) qJ( µ ) = B(λ ) I J q ∞,λ


(4.76)
B(λ )q(Jλ ) − B( µ )q J( µ ) = B(λ ) I J q ∞,λ

Similarly for the fourth and fifth rows in Eq. (4.75), we have

A(λ ) q(Jλ ) + A( µ ) qJ( µ ) = A(λ ) I J q ∞,λ


(4.77)
A(λ ) q(Jλ ) − A( µ ) qJ( µ ) = A(λ ) I J q ∞,λ

Therefore, using the modified A and B matrices, the interface conditions for
determining the unknown vectors q can be written, similar to the perfect interface
case Eq. (4.62), as
(µ ) (λ )
 (λ )q (λ ) + A
A  
q J(µ ) = A I J q ∞, λ
J
(µ ) (λ) (4.78)
 (λ)q(λ) + B
B  
q J(µ ) = B I J q ∞, λ
J

where the newly modified A and B matrices are defined as


(λ)
 B11 B12 B13 B14 B15 
 B21 B22 B23 B24 B25 
 (λ )
A =  A31 A32 A33 A34 A35 
 A41 A42 A43 A44 A45 
 A51 A52 A53 A54 A55 
(µ) (4.79)
 B11 B12 B13 B14 B15 
 B21 B22 B23 B24 B25 
 (µ)
A =  A31 A32 A33 A34 A35 
 A41 A42 A43 A44 A45 
 A51 A52 A53 A54 A55 

(λ )
 B11 B12 B13 B14 B15 
 B21 B22 B23 B24 B25 
 (λ)
B =  B31 B32 B33 B34 B35 
 A41 A42 A43 A44 A45 
 A51 A52 A53 A54 A55  (4.80)
(µ )
 − B11 − B12 − B13 − B14 − B15 
(µ )  − B21 − B22 − B23 − B24 − B25 

B =  B31 B32 B33 B34 B35 
 − A41 − A42 − A43 − A44 − A45 
 − A51 − A52 − A53 − A54 − A55 

Thus, the solutions to Eq. (4.78) have the same forms as Eqs. (4.63) and (4.64), but
with the modified A and B matrices defined by Eqs. (4.79) and (4.80). Hence, the
bimaterial Green’s functions for this complicated imperfect interface conditions are
completely solved.
4.6 Applications in Semiconductor Industry 129

x3

Interface Material 1 x1
o
Material 2

V
γ*Ij S

Figure 4.2. A general inclusion of polygonal shape in an anisotropic MEE bimaterial plane
where the inclusion is located in the lower half-plane (x3 < 0) under an extended eigenstrain
γ *Ij ( γ *ij , − E *j , − H *j ). Reproduced from Pan (2004b): Proceedings of the Royal Society of London
A 460: 537–60. Used with permission © Royal Society Publishing.

4.6 Applications in Semiconductor Industry

4.6.1 Basic Formulations of the Eshelby Inclusion and Quantum Wires


Let us assume that there is an extended general eigenstrain γ Ij* (γ ij* , −E*j , − H *j ) within
the domain V bounded by its surface S (see Figure 4.2 for illustration), which is
further located in one of the half-plane of the bimaterial system.
From Chapter 2 and assuming that the eigenstrain is uniform within the domain
V, then the induced extended displacements can be found as

uK ( y) = ciJLm γ *Lm ∫ uJK ( x; y)ni ( x )dS( x ) (4.81)


S

where ni(x) is the outward normal to the boundary S and uJK ( x; y) is the extended
displacement in J-direction at x due to an extended concentrated line-force in K-
direction applied at y.
To find the elastic strain, electric, and magnetic fields, we take the derivatives of
Eq. (4.81) with respect to the field point y (i.e., the source point of the line-source
Green’s functions), which yields

1
γ kp ( y) = 2 γ *Lm ciJLm ∫ [uJk, yp ( x; y) + u pJ, yk ( x; y)]ni ( x)dS( x) (k, p = 1, 2, 3) (4.82a)
S

(4.82b)
E p ( y) = − γ *Lm ciJLm ∫ uJ5, yp ( x; y)ni ( x )dS( x ) ( p = 1, 2, 3)
S

(4.82c)
H p ( y) = − γ *Lm ciJLm ∫ uJ5, yp ( x; y)ni ( x)dS( x) ( p = 1, 2, 3)
S

We remark that Eqs. (4.82a–c) together can be expressed alternatively using the
extended Eshebly tensor SIjLm (Eshelby 1957, 1961; Mura 1987; Dunn and Taya 1993;
Dunn and Wienecke 1997), as

γ Ij = SIjLmγ Lm
* (4.83)
130 Green’s Functions in Magnetoelectroelastic Full and Bimaterial Planes

where the elements of the extended Eshebly tensor S are readily obtained by comparing
Eq. (4.83) with Eq. (4.82). Furthermore, the stresses, electric displacements, and magnetic
inductions can be obtained using the constitutive relations (2.2) in Chapter 2.
Remark 4.4: Equations (4.81) and (4.82) are very useful, because for a uniform
eigenstrain within a homogeneous MEE solid, the elastic, electric, and magnetic
fields can be obtained by performing an integral over the boundary of the inclusion,
using the available Green’s function (see, e.g., Pan 1999, 2002) as the integrand. Thus,
if the Green’s functions are in relatively simple form and the inclusion domain is
also simple, instead of numerically carrying out the boundary integrals in Eqs. (4.81)
and (4.82), one could obtain the analytical expressions of the Eshelby inclusion-
induced field. We will show some 2D examples in this chapter as applications to the
semiconductor quantum wires case.
Semiconductor quantum wires (QWRs) are a special case of Eshelby inclusion
where due to the lattice mismatch between the inclusion material (QWR) and the
substrate matrix, a mismatch strain field or eigenstrain field will be generated within
the QWR. As a good approximation, the material property of the QWR could be
assumed to be the same as the matrix. Thus the QWR problem becomes a standard
Eshelby inclusion problem.
For 2D Green’s functions of the extended displacement presented in this chapter,
one can observe that the Green’s displacements contain only the logarithmic function,
which can be integrated out for certain integral paths. We now consider only the case
in which the QWR is of a polygonal shape of N sides. In other words, the boundary
of the QWR is made of N straight-line segments. While the application is for the
semiconductor QWRs, it can be for any Eshebly problem that can be described by
the MEE coupled or various uncoupled material systems.
Let us define a line segment in the (x1, x3)-plane starting from point 1 (a1,c1) and
ending at point 2 (a2,c2), in terms of the parameter t (0 ≤ t ≤1), as

x1 = a1 + (a2 − a1 )t
(4.84)
x3 = c1 + (c2 − c1 )t

Then, the outward normal component ni(x) (i = 1 and 3) along the line segment is
constant, given by

n1 = (c2 − c1 ) / l, n3 = −(a2 − a1 ) / l (4.85)

where l = (a2 − a1 )2 + (c2 − c1 )2 is the length of the line segment. It is obvious that
the differential length is dS = ldt.
As an illustration, we now assume that the polygonal QWR is in an infinite
plane. Thus the induced extended displacement field can be found using Eq. (4.81)
with the Green’s functions being those in Eq. (4.20a), that is,

5
1
uK ( y) = ciJLm γ *Lm Im ∑ [ AJR ∫ ln(zR − sR )ni ( x)dS( x)AKR ] (4.86)
π R=1
S

where we recall that zR = x1+pRx3 and sR = y1+pRy3 are related to the field x(x1,x3) and
source y(y1,y3) points, respectively, of the line-force Green’s functions.
4.6 Applications in Semiconductor Industry 131

Because along each segment of the polygon, the outer normal ni is constant, we
thus only need to carry out the following integral

1
hR ( y) ≡ ∫ ln(zR − sR )dt (4.87)
0

or,

1
hR ( y) = ∫ ln{[(a2 − a1 ) + pR (c2 − c1 )]t + [(a1 + pR c1 ) − sR ]}dt (4.88)
0

Integration of this expression gives

(a1 + pR c1 ) − sR  a + pR c2 − sR 
   hR ( y) = ln 2 + ln[a2 + pR c2 − sR ] − 1 (4.89)
(a2 − a1 ) + pR (c2 − c1 )  a1 + pR c1 − sR 

Therefore, the induced elastic displacements, electric and magnetic potentials, due
to the contribution of a straight-line segment along the boundary of the polygonal
QWR, can be obtained in the following exact closed-form

5
l
uK ( y) = ni ciJLmγ Lm
* Im ∑ [ AJR hR ( y) AKR ] (4.90)
π R =1

By adding the contributions from all line segments of the boundary, the solution for
an inclusion with a general polygonal shape in a full-plane is then obtained in an
exact closed-form!
The exact closed-form solution for the strain, electric, and magnetic fields
can be obtained by simply taking the derivative of Eq. (4.90) with respect to
the coordinate y = (y1,y3). After that, the constitutive relations can be utilized to
find the extended stress field. Because the Green’s functions for the extended
displacements are all in terms of the simple logarithmic functions in both half and
bimaterial planes, one can find the exact closed-form solutions for a polygonal
QWR in more complicated problem domains. Furthermore, based on the principle
of superposition, the corresponding multiple QWR problem can be also solved
in exact closed-form by adding all the contributions of different QWRs together.
In the numerical examples in the following text, we let x = x1 and z = x3 for easy
presentation.

4.6.2 Quantum Wires in a Piezoelectric Full Plane


We now apply the exact closed-form solutions, that is, Eq. (4.90) and its derivatives
with respect to the field point y, to a square QWR in a piezoelectric GaAs (001)
full-plane. For this example, the dimension of the Green’s matrix (uJK ) is 4×4 instead
of 5×5. The magnetic field and related coefficients are set to be zero. The QWR has
a dimension of 20nm×20nm. The misfit-strain is hydrostatic, that is, γ xx
* = γ * = 0.07.
zz
The material properties for GaAs are listed in Pan (2004a), which can also be found
in Chapter 2.
Shown in Figures 4.3a and 4.3b are, respectively, the contours of the strain
component γxx and hydrostatic strain γxx+γzz in the square QWR within the GaAs
132 Green’s Functions in Magnetoelectroelastic Full and Bimaterial Planes

(a) (b)
10 10

8 8

6 0.062 6 0.098
0.060 0.097
4 0.058 4 0.097
0.056 0.097
2 0.054 2 0.096
0.052 0.095
0.050 0.095
0 0
z (nm)

0.094

z (nm)
0.048
0.046 0.094
–2 0.044 –2 0.094
0.093
0.042 0.093
–4 0.040 –4 0.092
0.038 0.091
–6 0.036 –6 0.091
0.034 0.090
–8 –8

–10 –10
–10 –8 –6 –4 –2 0 2 4 6 8 10 –10 –8 –6 –4 –2 0 2 4 6 8 10
x (nm) x (nm)

Figure 4.3. Contours of strain component γxx (a) and hydrostatic strain γxx+γzz (b) in a square
QWR of 20nm×20nm within the GaAs (001) full-plane (where x≡x1,z≡x3). Reproduced with
permission from Pan (2004a): Journal of the Mechanics and Physics of Solids 52: 567–89.
© 2004 Elsevier.

(001) full-plane (Pan 2004a). It is observed from Figure 4.3a that while the two
equal maximums of γxx are reached in the middle of the left and right sides of the
square QWR with a value (= 0.062) that is slightly less than the misfit-strain, the two
equal minimums are reached in the middle of the top and bottom sides of the square
with a value (= 0.036) slightly over half of the misfit-strain. The hydrostatic strain
(Figure 4.3b), however, has a very gentle variation in the square QWR, with the
maximum difference less than 10 percent within the QWR. Notice further that these
normal strains are finite at the four corners.

4.6.3 Quantum Wires in an MEE Half-Plane


Figure 4.4 shows the geometry of a typical T-shaped QWR in an MEE half-
plane substrate, obtained from experimental observations (Goldoni et al. 1997;
Rossi et al. 1997, 1999; Grundmann et al. 1998; Itoh et al. 2003). The T-shaped
QWR is formed by intersecting two quantum-well structures where the carrier is
confined in one direction. The surface of the half-plane is assumed to be extended
traction-free (i.e., traction-free, electric and magnetic insulating as in the first
row in Table 4.1). The eigenstrain field within the T-shaped QWR is assumed to
be γ xx
* = γ * = 1, and the half-plane is assumed to be a fully coupled MEE solid,
zz
with its material properties being given in Jiang and Pan (2004), which can also
be found in Chapter 2. It is noted that these properties are chosen from the
composite materials consisting of piezoelectric BaTiO3 fiber reinforcement and
magnetostrictive CoFe2O4 matrix.
Figures 4.5, 4.6, and 4.7, show, respectively, the hydrostatic strain, total electric
field, and total magnetic field in the MEE half-plane. It is observed clearly that
while the elastic field inside the QWR is about a magnitude of one-order larger
than that outside (Figure 4.5), the electric and magnetic fields inside are only about
twice large as those outside (Figures 4.6 and 4.7). This is due to the fact that only
4.6 Applications in Semiconductor Industry 133

x
0 S R

–20

P Q
z (nm)

–40

–60

–80
–80 –40 0 40 80
x (nm)
Figure 4.4. A typical T-shaped QWR within an MEE half-plane substrate (z < 0) (where x≡x1
and z≡x3). Reproduced with permission from Jiang and Pan (2004): International Journal of
Solids and Structures 41: 4361–82. © 2004 Elsevier.

(inside) (outside)

2.2 0.335
–5 0.3 0.2 2.185
0.32
5
18

0.3
0.

2.17
0.1 0.29
85 2.155
0.2

0.275
–10
2.14 0.26
2.1

2.125 0.245
0.23
2.1
z (nm)

–15 2.11
2.1

2.1 0.215
2.1 2.1
0.2
2.08 0.185
–20
5

2.065 0.17
0.1

0.15
15 2.05
0. 0.14
2.035
–25 0.125
2.02
0.11
2 0.1
2

–30 1.99 0.08

–20 –15 –10 0 –5 5 10 15 20


x (nm)
Figure 4.5. Hydrostatic strain γxx+γzz distribution inside and outside the T-shaped QWR in the
half-plane MEE substrate (z < 0). Reproduced with permission from Jiang and Pan (2004):
International Journal of Solids and Structures 41: 4361–82. © 2004 Elsevier.

elastic eigenstrain is given inside the QWR. It is pointed out that while the extended
displacement and traction vectors are required to be continuous across the interface
between the QWR and its matrix, there is no requirement on the continuity of the
corresponding extended strain field when passing the interface.
134 Green’s Functions in Magnetoelectroelastic Full and Bimaterial Planes

(inside) (outside)

200 105
190 100
–5

50

40
180 95
170 90
160 85

50
–10 40 150 80
140 75
130 70

150
120 65

150
z (nm)

–15
100 110 60
25

50
40
50 100 55
40 50 90 50
–20 40 80 45
70 40
60 35
25
50 30
–25 40 25
30 20
20 15
10 10
–30
–20 –15 –10 –5 0 5 10 15 20
x (nm)

Figure 4.6. Total electric field Ex2 + Ez2 (×107V/m) distribution inside and outside the T-shaped
QWR in the half-plane MEE substrate (z < 0). Reproduced with permission from Jiang and
Pan (2004): International Journal of Solids and Structures 41: 4361–82. © 2004 Elsevier.

(inside) (outside)

4.75 2.9
–5 4.5 2.75
4.25 2.6
4 2.45
3.75
1

2.3
–10 3.5 2.15
3.25
2
4 3
1.85
4

2.75
4

–15
z (nm)

4 1.7
1 2.5
1.55
2.25
1 1.4
2
–20 1.25
1.75
1.5 1.1
1.25 1
1 0.8
–25
0.75 0.65
1 1
0.5 0.5
0.25 0.35
–30 0 0.2

–20 –15 –10 –5 0 5 10 15 20


x (nm)

Figure 4.7. Total magnetic field H x2 + Hz2 (×107A/m) distribution inside and outside the
T-shaped QWR in the half-plane MEE substrate (z < 0). Reproduced with permission from Jiang
and Pan (2004): International Journal of Solids and Structures 41: 4361–82. © 2004 Elsevier.
4.6 Applications in Semiconductor Industry 135

4.6.4 Quantum Wires in a Piezoelectric Bimaterial Plane


We now apply our bimaterial Green’s function solution to an inclusion in a
piezoelectric bimaterial plane. The bimaterials are made of two typical piezoelectric
materials: one is a left-hand quartz in a rotated coordinate system (Tiersten 1969),
and the other one is the poled lead-zirconate-titanate (PZT-4) ceramic (Dunn
and Taya 1993). Their elastic constants, piezoelectric coefficients, and dielectric
constants are given in Pan (2004b) and can also be found in Chapter 2.
We remark that while the quartz is a weakly coupled piezoelectric material,
the ceramic is a strongly coupled one, with their electromechanical coupling
factors, defined as g = emax / (ε max cmax ) , being equal to 0.07 and 0.5, respectively.
In the definition of the coupling factor g, emax denotes the maximum among
all piezoelectric coefficients eij, εmax denotes the maximum among all dielectric
permittivity coefficients εij, and cmax denotes the maximum among all elastic
coefficients cij.
Two bimaterial cases are considered. For case 1, named quartz/ceramic,
Material 1 (i.e., the upper half-plane with z > 0) is quartz and Material 2 (i.e.,
the lower half-plane with z < 0) is ceramic. For case 2, named ceramic/quartz,
Material 1 is ceramic and Material 2 is quartz. For both cases, there is a square
QWR of 20nm×20nm in Material 2, and it is symmetrically located with respect
to the z-axis with its upper side having a distance of 5nm to the interface. The
eigenstrain in the inclusion is assumed to be hydrostatic with γ xx * = γ * = 0.07, a
zz
typical magnitude in the InAs/GaAs quantum structure (Bimberg et al. 1999).
Plotted in Figure 4.8a is the contour of the electric field component Ex (V/m)
within the square QWR in Material 2 for the bimaterial quartz/ceramic case. The
corresponding result is shown in Figure 4.8b for the bimaterial ceramic/quartz case.
We notice that the induced Ex is completely different for the two bimaterial systems.
Their contour shapes and magnitudes are clearly different from each other. In
particular, the maximum magnitude in quartz/ceramic (Figure 4.8a) is roughly twice
of that in ceramic/quartz (Figure 4.8b). Also, we observe that the distribution of the
electric field component Ex for the ceramic/quartz case (Figure 4.8b) is asymmetric,
due to the fact that the quartz has been rotated to become a low-symmetry monoclinic
material. It should be noted that while in Figure 4.8a there is a field concentration
at each corner of the square QWR, there is only one concentration near the center
in Figure 4.8b.
Figures 4.9a and 4.9b show the contours of the electric field component Ez
(V/m) within the square QWR in Material 2 of quartz/ceramic (a) and of ceramic/
quartz (b). Again their magnitudes and shapes are completely different, with the
maximum magnitude in quartz/ceramic being about five times larger than that in
ceramic/quartz. Different from the Ex-field in Figure 4.8, here we observe that while
there is a field concentration near the middle of each side in Figure 4.9a, each corner
of the square QWR shows a concentration (Figure 4.9b).
136 Green’s Functions in Magnetoelectroelastic Full and Bimaterial Planes

(a)
–6

–8
1.8E+008
1.6E+008
–10 1.4E+008
1.2E+008
1.0E+008
–12 8.0E+007
6.0E+007
4.0E+007
–14
z (nm)

2.0E+007
0.0E+000
-2.0E+007
–16 -4.0E+007
-6.0E+007
-8.0E+007
–18 -1.0E+008
-1.2E+008
-1.4E+008
–20 -1.6E+008
-1.8E+008
-2.0E+008
–22

–24

–10 –8 –6 –4 –2 0 2 4 6 8 10
x (nm)

(b)
–6

–8
6.0E+007
–10 5.0E+007
4.0E+007
–12 3.0E+007
2.0E+007
1.0E+007
–14
z (nm)

0.0E+000
-1.0E+007
–16 -2.0E+007
-3.0E+007
–18 -4.0E+007
-5.0E+007
–20 -6.0E+007
-7.0E+007
-8.0E+007
–22

–24

–10 –8 –6 –4 –2 0 2 4 6 8 10
x (nm)

Figure 4.8. Contours of electric field components Ex(V/m) within the square QWR of
20nm × 20nm in Material 2 of the quartz/ceramic bimaterial (a) and the ceramic/quartz
bimaterial (b). Reproduced from Pan (2004b): Proceedings of the Royal Society of London A
460: 537–60. Used with permission © Royal Society Publishing.

4.7 Summary and Mathematical Keys

4.7.1 Summary
Because the problems discussed depend only on two spatial variables, the complex
variable function method is most suited to derive the Green’s functions. Furthermore,
the elegant Stroh formalism is applied to solve the eigenvalue problem of the
homogeneous governing equations and to express the corresponding eigenvalues
and eigenvectors.
The 2D Green’s function solutions corresponding to the extended line force and
line dislocation are very similar to each other and further possess the same order
of singularity. This feature is in contrast to the corresponding 3D solutions in which
the Green’s function solution to the point dislocation has a higher-order singularity
than that to the point force. This is explained physically in this chapter along with the
illustrative diagram in Figure 4.1 and the mathematical relations in Box 4.1.
4.7 Summary and Mathematical Keys 137

(a)
–6

–8
-3.8E+007
-4.0E+007
–10 -4.2E+007
-4.4E+007
-4.6E+007
–12 -4.8E+007
-5.0E+007
-5.2E+007
–14 -5.4E+007
z (nm)

-5.6E+007
-5.8E+007
-6.0E+007
–16 -6.2E+007
-6.4E+007
-6.6E+007
–18 -6.8E+007
-7.0E+007
-7.2E+007
-7.4E+007
–20 -7.6E+007
-7.8E+007
-8.0E+007
–22 -8.2E+007

–24

–10 –8 –6 –4 –2 0 2 4 6 8 10
x (nm)

(b)
–6

–8
3.5E+008
–10 3.0E+008
2.5E+008
2.0E+008
–12
1.5E+008
1.0E+008
–14 5.0E+007
z (nm)

0.0E+000
–16 -5.0E+007
-1.0E+008
-1.5E+008
–18 -2.0E+008
-2.5E+008
–20 -3.0E+008
-3.5E+008
-4.0E+008
–22

–24

–10 –8 –6 –4 –2 0 2 4 6 8 10
x (nm)

Figure 4.9. Contours of electric field components Ez(V/m) within the square QWR of
20nm × 20nm in Material 2 of the quartz/ceramic bimaterial (a) and the ceramic/quartz
bimaterial (b). Reproduced from Pan (2004b): Proceedings of the Royal Society of London A
460: 537–60. Used with permission © Royal Society Publishing.

The Green’s functions of a half-plane with general (but specially homogeneous)


boundary conditions and those of a bimaterial plane with general imperfect (but
specially homogeneous) interfacial conditions can be equally found and the results
share the same simple compact form as those corresponding to the conventional
boundary and interfacial conditions.
Application examples are given for the QWR semiconductors, and are limited
only to a single polygonal inclusion. While applications to multiple QWRs or array
of QWRs can be found as in Han et al. (2006), a more complicated inclusion shape
can be approximated by a suitable number of straight-line segments (as in Jiang
and Pan 2004), or by the conformal mapping method (as in Ru 1999, 2000). More
recently, the extended Stroh formalism of complex variables was applied by Zou
and Pan (2012) to solve exactly the inclusion problems in the bimaterial MEE plane
where the inclusion can be of arbitrary shape, described by a Laurent polynomial, a
polygon, or the one bounded by a Jordan curve.
The inclusion problem is not limited to the QWR semiconductors. There are
many problems in materials science and engineering that require the solution of
138 Green’s Functions in Magnetoelectroelastic Full and Bimaterial Planes

an inclusion problem. Even for the polygonal inclusion case, there are a couple of
advanced research topics. One is to solve the corresponding inhomogeneity problem
that is important in finding the effective material property of the composite where
the cross-section of fibers is not in a circular or elliptical shape but rather in a
polygonal shape. The first step to do so is to predict accurately the inhomogeneity-
induced field in the full-plane (Sun et al. 2012a). Furthermore, in this chapter, we
have assumed that the inclusion is under a uniform eigenstrain field. In a QWR or an
inclusion, the eigenstrain field is usually nonuniform. A couple of interesting works
appeared recently that would be good references for future effort (Sun et al. 2012b;
Chen et al. 2014).

4.7.2 Mathematical Keys


The expressions of the Green’s functions due to a line force and a line dislocation
are very similar to each other. The induced displacements are all proportional to
the logarithmic function and the induced strains/stresses are an inverse function of
the distance between the source and field points. As such, many physical features
observed from the line-dislocation solution should also exist in the solutions due
to the line force. One should keep in mind that, in contrast to the corresponding
3D case in which the solutions due to the point force and point dislocation have
different orders of singularities, here for the 2D case, their orders of singularity are the
same. Furthermore, because the line-source induced displacements are logarithmic
functions, they are singular both at the source point and at infinity.
Due to the analytical expressions of the 2D line-source Green’s functions, they
can be directly applied to many interesting problems by making use of the Green’s
representation presented in Chapter 2. While illustrated already in this chapter,
more complicated and interesting Eshelby problems can be also solved based on the
Green’s functions and the Green’s representation.

4.8 References
Barnett, D. M. and Lothe, J. 1975. Dislocations and line charges in anisotropic piezoelectric
insulators. Physica Status Solidi B 67: 105–11.
Bimberg, D., Grundmann, M., and Ledentsov, N. N. 1999. Quantum Dot Heterostructures.
Chichester, UK: John Wiley and Sons.
Chen, Q. D., Xu, K. Y., and Pan, E. 2014. Inclusion of arbitrary polygon with graded eigenstrain
in an anisotropic piezoelectric half plane. International Journal of Solids and Structures
51:53–62.
Dunn, M. L. and Taya, M. 1993. An analysis of piezoelectric composite materials containing
ellipsoidal inhomogeneities. Proceedings of the Royal Society of London A443: 265–87.
Dunn, M. L. and Wienecke, H. A. 1997. Inclusions and inhomogeneities in transversely
isotropic piezoelectric solids. International Journal of Solids and Structures 34:
3571–82.
Eshelby, J. D. 1957. The determination of the elastic field of an ellipsoidal inclusion, and
related problems. Proceedings of the Royal Society of London A 241: 376–96.
1961. Elastic inclusions and inhomogeneities. In: Progress in Solid Mechanics (Eds.
Sneddon, I. N. and Hill, R.), Vol. 2, pp. 89–140. Amsterdam: North-Holland.
Goldoni, G., Rossi, F., and Molinari, E. 1997. Quantum interference in nanometric devices:
Ballistic transport across array of T-shaped quantum wires. Applied Physics Letters 71:
1519–21.
4.8 References 139

Grundmann, M., Stier, O., and Bimberg, D. 1998. Electronic states in strained cleaved-edge-
overgrowth quantum wires and quantum dots. Physical Review B 58: 10557–61.
Han, F., Pan, E., and Albrecht, J. D. 2006. Strain and piezoelectric fields in embedded quantum
wire arrays. Superlattices and Microstructures 40: 125–36.
Itoh, H., Hayamizu, Y., Yoshita, M., Akiyama, H., Pfeiffer, L. N., West, K. W., Szymanska, M.
H., and Littlewood, P. B. 2003. Polarization-dependent photoluminescence-excitation
spectra of one-dimensional exciton and continuum states in T-shaped quantum wires.
Applied Physics Letters 83: 2043–5.
Jiang, X. and Pan, E. 2004. Exact solution for 2D polygonal inclusion problem in anisotropic
magnetoelectroelastic full-, half-, and bimaterial-planes. International Journal of Solids
and Structures 41: 4361–82.
Mura, T. 1987. Micromechanics of Defects in Solids. 2nd rev. ed. Dordrecht, The Netherlands:
Kluwer Academic Publishers.
Pan, E. 1997. A general boundary element analysis of 2-D linear elastic fracture mechanics.
International Journal of Fracture 88: 41–59.
1999. A BEM analysis of fracture mechanics in 2-D anisotropic piezoelectric solids.
Engineering Analysis with Boundary Elements 23: 67–76.
2002. Mindlin’s problem for an anisotropic piezoelectric half space with general boundary
conditions. Proceedings of the Royal Society of London A458: 181–208.
2003a. Three-dimensional Green’s functions in anisotropic elastic bimaterials with
imperfect interfaces. Journal of Applied Mechanics 70: 180–90.
2003b. Some new three-dimensional Green’s functions in anisotropic piezoelectric
bimaterials. Electronic Journal of Boundary Elements 1: 236–69 (also with 2D bimaterial
Green’s functions under general interfacial conditions).
2004a. Eshelby problem of polygonal inclusion in anisotropic piezoelectric full- and half-
planes. Journal of the Mechanics and Physics of Solids 52: 567–89.
2004b. Eshelby problem of polygonal inclusions in anisotropic piezoelectric bimaterials.
Proceedings of the Royal Society of London A460: 537–60.
Pan, E. and Amadei, B. 1996. Fracture mechanics analysis of 2-D cracked anisotropic media
with a new formulation of the boundary element method. International Journal of
Fracture 77: 161–74.
Rossi, F., Goldoni, G., and Molinari, E. 1997. Shape-independent scaling of excitonic
confinement in realistic quantum wires. Physical Review Letters 78: 3527–30.
Rossi, F., Goldoni, G., Mauritz, O., and Molinari, E. 1999. Theory of excitonic confinement in
semiconductor quantum wires. Journal of Physics: Condensed Matter 11: 5969–88.
Ru, C. Q. 1999. Analytical solution for Eshelby’s problem of an inclusion of arbitrary shape
in a plane or half-plane. Journal of Applied Mechanics 66: 315–22.
2000. Eshelby’s problem for two-dimensional piezoelectric inclusions of arbitrary shape.
Proceedings of the Royal Society of London A456: 1051–68.
Sun, L. G., Xu, K. Y., and Pan, E. 2012a. Irregular inhomogeneities in an anisotropic
piezoelectric plane. Journal of Applied Mechanics 79: 021014.
2012b. Inclusion of arbitrary polygon with graded eigenstrain in an anisotropic piezoelectric
full plane. International Journal of Solids and Structures 49: 1773–85.
Tiersten, H. F. 1969. Linear Piezoelectric Plate Vibrations. New York: Plenum.
Ting, T. C. T. 1992. Image singularities of Green’s functions for anisotropic elastic half-spaces
and bimaterials. The Quarterly Journal of Mechanics and Applied Mathematics 45:
119–39.
1996. Anisotropic Elasticity. Oxford: Oxford University Press.
Ting, T. C. T., and Wang, M. Z. 1992. Generalized Stroh formalism for anisotropic elasticity
for general boundary conditions. Acta Mechanica Sinica 8: 193–207.
Zou, W. N. and Pan, E. 2012. Eshelby’s problem in an anisotropic multiferroic bimaterial
plane. International Journal of Solids and Structures 49: 1685–700.
5 Green’s Functions in Elastic Isotropic
Full and Bimaterial Spaces

5.0 Introduction
This chapter is on an old topic. While the Green’s function solutions in an elastic
isotropic full-space or bimaterial space are well known, the methodologies employed
to derive the solutions are worthy to be discussed. Furthermore, most Green’s
function solutions in elastic isotropic bimaterial spaces are for the potential functions
only and at most, the Green’s displacement solutions only. In this chapter, we present
a systematic study on the methods employed in deriving the Green’s functions and
list the complete Green’s function solutions (potential functions, displacements, and
stresses) in the bimaterial space. Both perfectly bonded and smooth interfaces are
considered. Various reduced cases are also discussed and plots are shown for the
surface circular loading case over a half-space.
We also point out that the methodology utilized in deriving the full-space and
bimaterial space Green’s functions in this chapter is different from that employed in
Chapters 6 and 7 when deriving the corresponding Green’s functions in transversely
isotropic solids. More specifically, in this chapter, we employ the Galerkin potential
functions to derive the full-space Green’s functions and the Papkovich functions
to derive the bimaterial Green’s functions. It is worthy to mention that the original
Green’s potential functions in bimaterial potential fields and the original Green’s
theorem are both employed to derive the Green’s functions in an isotropic elastic
bimaterial space (Rongved 1955).

5.1 Green’s Functions of Point Forces in an Elastic Isotropic Full-Space


The governing equations in a linearly elastic isotropic solid are given as:
The equilibrium equations
σ ij ,i = − f j ( x ) (5.1)

where σij is the elastic stress tensor, and fj the body force vector.
The displacement-strain relations

1
γ ij = 2 (ui, j + u j,i ) (5.2)

140
5.1 Green’s Functions of Point Forces in an Elastic Isotropic Full-Space 141

where γij is the elastic strain tensor, and uj the displacement vector.
The constitutive relations

σ ij = cijlmγ lm (5.3)

where cijkl is the fourth-order stiffness tensor, which, for the isotropic case, is
reduced to

cijkl = λδ ij δ kl + µ(δ ik δ jl + δ il δ jk ) (5.4)

where

2 µν
λ= (5.5)
(1 − 2ν )

and μ are two Lamé constants, while ν is the Poisson’s ratio.


Thus the constitutive relations (5.3) are reduced to

2µν
σij = δ ij γ kk + 2µγ ij
1 − 2ν (5.6)
2µν
= δ ij uk ,k + µ(ui, j + u j,i )
1 − 2ν

There are a couple of methods to find the full-space point-force Green’s functions, or
the Kelvin’s solutions (Love 1927). The solution process in the following text is based
on the Galerkin potential function method as in Brebbia and Dominguez (2001).
First, in terms of the elastic displacements, the equilibrium equations (5.1)
become

1 − f j ( x)
um,mj + u j ,ll = (5.7)
1 − 2ν µ

For a point-force of unit magnitude applied at y = 0 in the k-direction, the body


force in Eq. (5.7) becomes f j ( x ) = δ jk ( x ) (≡δjkδ(x)). We now let ulk ( x ) stand for the
displacement in the l-direction at x induced by this unit force in the k-direction at
the origin y = 0. Then

1 k k
−δ jk ( x )
um , mj + u j ,ll = (5.8)
1 − 2ν µ

Representing the Green’s elastic displacements in terms of the Galerkin’s vector gk


(Brebbia and Dominguez 2001) due to the point force in the k-direction, we have

1
ukj = g kj ,mm − gk (5.9)
2(1 −ν ) m,mj
142 Green’s Functions in Elastic Isotropic Full and Bimaterial Spaces

Then, Eq. (5.8) becomes

−δ jk ( x )
g kj ,llmm = (5.10)
µ

We further introduce an intermediate vector function Fk due to the point-force in


the k-direction as

Fjk = g kj ,ll (5.11)

Then, Eq. (5.10) becomes a standard Poisson’s equation as

−δ jk ( x)
Fjk,mm = (5.12)
µ

with one of its particular solutions being

δ jk
Fjk = (5.13)
4π r µ

where r = |x| is the distance between the field point x and the source point y = 0.
Substituting solution (5.13) back to Eq. (5.11) results in the differential equations
that the Galerkin vector should satisfy

δ jk
g kj ,ll = (5.14)
4π r µ

which has the following particular solution

r
g kj = δ jk (5.15)
8πµ

Finally, substituting Eq. (5.15) back to Eq. (5.9), we arrive at the Green’s function
solutions for the elastic displacement in the j-direction at the field point x due to the
unit point force in the k-direction applied at y = 0

1  1 
ukj ( x ) = r,llδ jk − r, jk  (5.16)
8πµ  2(1 − ν ) 

where the indices jk following the subscript prime “,” indicate the derivatives of r
with respect to xj and xk. For the point force applied at the source point y (y1,y2,y3)
with unit magnitude in the k-direction, the elastic displacement in the j-direction at
the field point x (x1,x2,x3) is

1  x j − y j xk − yk 
ukj ( x; y) = (3 − 4ν )δ jk +  (5.17)
16πµ (1 −ν )r  r r 
5.2 Papkovich Functions and Green’s Representation 143

where r, the distance between the field (x) and source (y) points, is now
expressed as

r = x − y = ( x1 − y1 )2 + ( x2 − y2 )2 + ( x3 − y3 )2 (5.18)

Remark 5.1: The point-force-induced elastic displacement is symmetric with respect


to the direction of the point force and the direction of the displacement. In other
words, k and j are exchangeable.
Remark 5.2: This Green’s function is also symmetric with respect to the source
coordinate and the field coordinate. In other words, the coordinates y and x are
exchangeable. Mathematically, the following relations hold

ukj ( x; y) = ukj ( x; y) = ukj ( y; x ) (5.19)

The preceding relations can also be obtained from the simple argument of the
reciprocal theorem.
Remark 5.3: The strain due to the point force can be obtained by taking the derivatives
of the Green’s displacement with respect to the field point x

1  1 
γ lm
k =
r, jjmδ lk + r, jjl δ mk − r,lkm  (5.20)
16πµ  (1 − ν ) 

Or the strain components can be expressed as:

 δ lm ( xk − yk ) − (1 − 2ν )[δ mk ( xl − yl ) + δ kl ( xm − ym )] 
1  r 
γ lm
k =  3( x − y )( x − y )( x − y )  (5.21)
16πµ (1 −ν )r 2 − k k l l m m 
 r3 

Remark 5.4: Using the constitutive relations (5.6), the Green’s stress components
can be expressed as:

1 (1 − 2ν )[δ lm ( xk − yk ) − δ lk ( xm − ym ) − δ mk ( xl − yl )] / r 
σ lm
k =
 (5.22)
8π (1 −ν )r  −3( xk − yk )( xl − yl )( xm − ym ) / r 3
2

5.2 Papkovich Functions and Green’s Representation


To derive the Green’s displacements and stresses in the corresponding elastic
bimaterial space, we first introduce the Papkovich functions and related Green’s
representations. As is well known (e.g., Wang 2002), the displacement u at position x
induced by the body force f(y) can be expressed in terms of the Papkovich functions
P and q as

1
u=P− ∇ ( x ⋅ P + q) (5.23a)
4(1 −ν )
144 Green’s Functions in Elastic Isotropic Full and Bimaterial Spaces

where ∇ is the gradient operator vector in the Cartesian coordinate system (referring
to the definition after Eq. (1.20) in Chapter 1). In terms of its components, Eq. (5.23a)
can be written as

1
ui = Pi − ( x j Pj + q),i
4(1 −ν )
(5.23b)
3 − 4ν 1
= Pi − ( x j Pj ,i + q,i )
4(1 −ν ) 4(1 −ν )

where P and q are the Papkovich functions satisfying the following vector and scalar
forms of the Poisson’s equation

µ∇ 2 P = − f
(5.24)
µ∇ 2 q = x ⋅ f

Now let us recall the Green’s theorem between volumetric and surface integrals for
two arbitrary functions, as in Eq. (1.20) in Chapter 1,

∫S (φ∇ψ − ψ∇ φ) ⋅ dS = ∫V (φ∇2 ψ −ψ∇2 φ)dV (5.25)

and let these two functions (φ and ψ) correspond to u and G, which satisfy the
following two equations:

∇ 2 u = g( x ) (5.26)
∇ 2 G = δ( x − y )

We then have

u( y) = ∫ u( x )∇G( x; y) ⋅ d S + ∫ G( x; y) g( x )d V ( x ) (5.27)
S V

In the preceding, we have assumed that G is the potential Green’s function in the
upper half-space where the source is located at the point (y1,y2,y3 > 0) with fixed
surface at x3 = 0 so that the other surface integral in Eq. (5.25) vanishes. Such a
potential Green’s function can be easily found as (see, Eq. (1.46) in Chapter 1)

−1  1 1 
G= − (5.28)
4 π  r1 r2 

where

r1 = ( x1 − y1 )2 + ( x2 − y2 )2 + ( x3 − y3 )2 (5.29)

and

r2 = ( x1 − y1 )2 + ( x2 − y2 )2 + ( x3 + y3 )2 (5.30)
5.3 Papkovich Functions with Perfect Interface 145

Then the representative relation (i.e., expressing any function u in terms of its surface
value if the body source is zero) gives

1 ∂G( x; y) 1
u( y) = −
4π ∫S
u( x )
∂n( x )
dS ( x ) −
4 π ∫V
G( x; y) g( x )dV ( x ) (5.31)

We note that the potential Green’s function G has the following symmetry
property

G( x; y) = G( y; x ) (5.32)

It is also clear from Eq. (5.31) that the first term (i.e., the surface integral) is
harmonic in the problem domain, and the second term (i.e., the volumetric integral)
is generally not. Furthermore, symbolically, we can express u as the sum of the surface
and volume terms

u = us + uv (5.33)

Equations (5.31) and (5.33) will be frequently used for deriving the Papkovich
functions.

5.3 Papkovich Functions in an Elastic Isotropic Bimaterial Space


with Perfect Interface
The Green’s functions of point forces in an elastic isotropic bimaterial space can be
derived using different approaches (Rongved 1955; Chen and Tang 1997; Guzina and
Pak 1999; Wang 2002). In this chapter, we follow the approach by Rongved (1955).
The most attractive feature of using Rongved’s approach is that in deriving the
bimaterial Papkovich functions, we utilize the bimaterial potential Green’s functions
combined with the integral equation method.
We assume that the shear modulus and Poisson’s ratio in Materials 1 and 2 of
the bimaterial elastic isotropic system are (μ1,ν1) and (μ2,ν2), respectively. We assume
further that there is a point force vector in Material 1, that is, x3 > 0, or z > 0 and
(y1,y2,y3) = (0,0,c > 0), of the bimaterial system with Material 2 in the lower half-
space x3 < 0.
We first present the detailed derivation of obtaining the Papkovich functions in
the bimaterial space where the two half-spaces are well bonded along the interface
x3 = 0 (z = 0). In other words, we have the following continuity conditions along the
interface (assuming its normal unit vector being iz, i.e., along the positive x3-axis)

u(1) = u( 2)
(5.34)
iz ⋅ σ (1) = iz ⋅ σ ( 2)

where the stress tensor can be expressed in terms of the displacement vector as
(Rongved, 1955)

σ = λ∇⋅uI + µ (∇ u + u ∇ ) (5.35)
146 Green’s Functions in Elastic Isotropic Full and Bimaterial Spaces

In terms of the Papkovich functions, the interface conditions (5.34) can be


expressed as

1 1
P (1) − ∇( x ⋅ P (1) + q(1) ) = P ( 2) − ∇( x ⋅ P ( 2) + q( 2) )
4(1 − ν1 ) 4(1 − ν2 )
µ1
iz ⋅ [(1 − 2 ν1 )(∇ P (1) + P (1) ∇) + 2 ν1 ∇ ⋅ P (1) I − (∇∇ P (1) ) ⋅ x − ∇∇q(1) ] (5.36)
2(1 − ν1 )
µ2
= iz ⋅ [(1 − 2 ν2 )(∇ P ( 2 ) + P ( 2 ) ∇) + 2 ν2 ∇ ⋅P ( 2 ) I − (∇∇ P ( 2 ) ) ⋅ x − ∇∇q( 2) ]
2(1 − ν2 )

Furthermore, in terms of the components, Eq. (5.36) can be written as

1 1
Pi(1) − ∂ i ( x j Pj(1) + q(1) ) = Pi( 2 ) − ∂ i ( x j Pj( 2 ) + q( 2 ) )
4(1 − ν1 ) 4(1 − ν2 )
µ1
δ 3 j [(1 − 2 ν1 )(∂ i Pj(1) + ∂ j Pi(1) ) + 2 ν1∂ m Pm(1) δ ij − xm ∂ i ∂ j Pm(1) − ∂ i ∂ j q(1) ]
2(1 − ν1 )
µ2
= δ 3 j [(1 − 2 ν2 )(∂ i Pj( 2 ) + ∂ j Pi( 2 ) ) + 2 ν2 ∂ m Pm( 2 ) δ ij − xm ∂ i ∂ j Pm( 2 ) − ∂ i ∂ j q( 2 ) ]
2(1 − ν2 )
 (5.37a, b)

where ∂ i = ∂ / ∂xi . With these preparations, we are ready to derive the Papkovich
functions. These are discussed separately when the point force is in the z-direction
and in the x-direction.

5.3.1 A Point Force Normal to the Interface Applied in Material 1


For this case, we can assume, due to the symmetric feature of the solution, that in the
two half-spaces (z > 0 and z < 0; we use x = x1, y = x2, and z = x3 thereafter),

Px(1) = Py(1) = 0 (z ≥ 0 )
(5.38)
Px(2) = Py(2) = 0, P (2) ≡ Pz(2) (z ≤ 0 )

Then the interface conditions (5.37a, b) at z = 0 are reduced to

1 − ν1 1 − ν1
∂ x q(1) = ∂ x q( 2) , ∂ y q(1) = ∂ y q( 2)
1 − ν2 1 − ν2
(5.39a)
1 − ν1
κ 1 Pz(1) − ∂z q(1) = κ 2 P (2) − ∂ zq(2) 
1 − ν2 

µ 2 1 − ν1
(1 − 2 ν1 )∂ x Pz(1) − ∂ x ∂ z q(1) = [(1 − 2 ν2 )∂ x P (2) − ∂ x ∂z q(2) ]
µ 1 1 − ν2
µ 1 − ν1
(1 − 2 ν1 )∂ y Pz(1) − ∂ y ∂ z q(1) = 2 [(1 − 2 ν2 )∂ y P ( 2) − ∂ y ∂ z q( 2) ] (5.39b)
µ 1 1 − ν2
µ 1 − ν1
2(1 − ν1 )∂ z Pz(1) − ∂ z∂ z q(1) = 2 [2(1 − ν2 )∂ z P ( 2) − ∂ z ∂ z q(2) ]
µ 1 1 − ν2
5.3 Papkovich Functions with Perfect Interface 147

where κi = 3 − 4vi (i = 1, 2). Because the first and second expressions in Eqs. (5.39a)
and (5.39b) involve the same functions, the solutions of Eqs. (5.39) may be converted
to the solutions of the following equations

1 − ν1 (2)
q(1) = q
1 − ν2
1 − ν1
κ 1 Pz(1) − ∂ z q(1) = [κ 2 P (2) − ∂ zq(2) ]
1 − ν2
    (5.40)
1 µ 1 − ν1
(1 − 2 ν1 )Pz( ) − ∂zq(1) = 2 [(1 − 2 ν2 )P (2) − ∂z q(2) ]
µ 1 1 − ν2
µ 1 − ν1
2(1 − ν1 )∂z Pz(1) − ∂z ∂z q(1) = 2 [ 2(1 − ν2 )∂ z P ( 2) − ∂ z ∂ zq( 2) ]
µ 1 1 − ν2

On the source side (Material 1), we can express the unknown Papkovich functions as
the sum of the surface and volume integral-related terms, as in Eq. (5.33)

Pz ≡ P = Ps + Pv
(5.41)
q = qs + qv

Because the Green’s function G is symmetric on its variables x and y, therefore, for
the volume integral-related part, we can write (by switching the source and field
points)

1
4 π ∫V
uv ( x ) = − G( y; x ) g( y) dV ( y) (5.42)

Applying Eq. (5.42) to Pv and qv in Material 1 (noticing that no superscript “(1)” is


needed for Material 1), with g being specified by the right-hand side terms in Eq.
(5.24), we find

1
4 πµ 1 ∫V
Pv ( x ) = G( y; x ) fz ( y)dV ( y)
(5.43)
−1
4 πµ 1 ∫V
qv ( x ) = G( y; x ) y3 fz ( y)dV ( y)

Using

fz ( y) = fz δ( y1 )δ( y2 )δ( y3 − c) (5.44)

we finally obtain

fz
Pv ( x ) = (1 / R1 − 1 / R2 )
4πµ 1
(5.45)
−cfz
qv ( x ) = (1 / R1 − 1 / R2 )
4πµ 1
148 Green’s Functions in Elastic Isotropic Full and Bimaterial Spaces

where

R1 = x12 + x22 + ( x3 − c)2


(5.46)
R2 = x12 + x22 + ( x3 + c)2

Notice that along the interface x3 = 0 (z = 0),

∂  1 1  ∂  1  ∂2  1 1 
1 / R1 = 1 / R2 ,  −  = −2   ,  − =0 (5.47)
∂z  R1 R2  ∂z  R2  ∂z  R1 R2 
2

then in terms of the surface and volume terms on the source side, the interface
conditions on z = 0 can be rewritten as (Pv = qv = 0 on the interface) (quantities in
the lower half-space are still denoted by superscript “(2)”)

1 − ν1 ( 2)
qs = q
1 − ν2
cf ∂  1  1 − ν1
κ 1 Ps − ∂z qs − z = [κ 2 P (2) − ∂ z q(2) ]
2πµ 1 ∂z  R2  1 − ν2
cf ∂  1  µ 2 1 − ν1
(1 − 2 ν1 )Ps − ∂ z qs − z = [(1 − 2 ν2 )P (2) − ∂z q(2) ]
2 πµ1 ∂z  R2  µ1 1 − ν2
(1 − ν1 ) fz ∂  1  µ 1 − ν1
2(1 − ν1 )∂z Ps −   − ∂z∂ z qs = 2 [2(1 − ν2 )∂ z P ( 2) − ∂ z∂ zq( 2) ]
πµ 1 ∂z  R2  µ 1 1 − ν2
 (5.48 a, b, c, d)

These four equations can be used to find the four functions (Ps and qs in Material 1
and P(2) and q(2) in Material 2). This is done as shown in the following text.
It is noted that in Eq. (5.48), its left-hand sides are harmonic in Material 1 (z ≥
0) and its right-hand sides are harmonic in Material 2 (z ≤ 0). Therefore, the solutions
to Eq. (5.48) should be such that the left-hand sides are the images with respect to
z = 0 of the right-hand sides, following directly from the uniqueness theorem of the
Dirichlet problem (i.e., Wang 2002). In other words, the relation between the left-
and right-hand sides for any preceding relation can be symbolically written as

L( x, y, z) = R( x, y, −z) (5.49)

Equation (5.49) indicates that when one takes the derivative with respect to z once
or integrate with respect z once on both sides of Eq. (5.49), there will be a sign
change in the expression. For instance, taking the derivative of Eq. (5.48a) on both
sides with respect to z, and substituting the result into Eq. (5.48b), we have

1 − ν1 ( 2) 1 − ν1
q = qs , ∂z q( 2) = −∂z qs
1 − ν2 1 − ν2
(5.50a, b, c)
1 − ν1 ( 2) κ 1 2 1 cfz ∂  1 
P = Ps − ∂ z qs −
1 − ν2 κ2 3 − 4 ν2 κ 2 2 πµ 1 ∂z  R2 
5.3 Papkovich Functions with Perfect Interface 149

Equations (5.50a, c) are the solution for the potential functions q(2) and P(2) in
Material 2 in terms of qs and Ps in Material 1. We now substitute them back into Eqs.
(5.48c, d), integrate Eq. (5.48d) with respect to z (causing a sign change on the right-
hand side), and carry out some algebraic calculations (addition, subtraction, etc.) to
find the following two expressions for functions Ps and ∂zqs

(µ1 − µ 2 )cfz z + c (1 − ν1 ) fz  1 
Ps = + (5.51)
2 πµ1 (µ1 + µ 2 κ 1 ) R2
3 π(µ1 + µ 2 κ 1 )  R2 

(1 − ν1 )cfz ∂  1 
∂z qs = −
π(µ1 + µ 2 κ 1 ) ∂z  R2 
(1 − ν1 )[µ1κ 2 (1 − 2 ν1 ) − µ 2 κ 1 (1 − 2 ν2 )] fz  1 
+  R 
π(µ1κ 2 + µ 2 )(µ1 + µ 2 κ 1 ) 2 (5.52)

Equation (5.52) can be integrated on both sides with respect to z to obtain

−(1 − ν1 )cfz 1 (1 − ν1 )[µ 1κ 2 (1 − 2 ν1 ) − µ 2 κ 1 (1 − 2 ν2 )] fz


qs = + ln R2* (5.53)
π(µ 1 + µ 2 κ 1 ) R2 π(µ 1κ 2 + µ 2 )(µ 1 + µ 2 κ 1 )

where R2* = R2 + z + c. Therefore, the final Papkovich functions in Material 1 can be


simply obtained by adding the surface- and volume-integral parts together; namely,
we have

Pz = Ps + Pv
F  1 (µ − µ 2 )  3 − 4 ν1 2c(z + c)   (5.54)
= z  + 1 + 
4πµ 1  R1 µ 1 + µ 2 κ 1  R2 R23  

fz  c (µ1 − µ 2 )κ 1 c 
q=− +
4πµ1  R1 (µ1 + µ 2 κ 1 ) R2 
(5.55)
f 4µ1 (1 − ν1 )[µ1κ 2 (1 − 2 ν1 ) − µ 2 κ 1 (1 − 2 ν2 )]
+ z ln R2*
4πµ1 (µ1κ 2 + µ 2 )(µ1 + µ 2 κ 1 )

For the Papkovich functions in Material 2 (i.e., the lower half-space free of any
source), we can make use of Eqs. (5.50a, c). However, to make use of Ps and qs, we
need to change z by −z, and R2 by R1, which gives us (from Eqs. (5.51) and (5.52))

(µ1 − µ 2 )cfz − z + c (1 − ν1 ) fz 1
Ps = + (5.56)
2 πµ1 (µ1 + µ 2 κ 1 ) R13 π (µ1 + µ 2 κ 1 ) R1

(1 − ν1 )cfz ∂  1 
∂z qs =
π(µ1 + µ 2 κ 1 ) ∂z  R1 
(5.57)
(1 − ν1 )[µ1κ 2 (1 − 2 ν1 ) − µ 2 κ 1 (1 − 2 ν2 )] fz  1 
+  R 
    π(µ 1κ 2 + µ 2 )(µ 1 + µ 2 κ 1 ) 1
150 Green’s Functions in Elastic Isotropic Full and Bimaterial Spaces

These finally give us the Papkovich potential functions in Material 2 (with q(2) being
obtained from Eqs. (5.50a) and (5.53)), which will be utilized to find the corresponding
displacement and stress fields.

5.3.2 A Point Force Parallel to the Interface Applied in Material 1


For this case with a point force in the x (x1)-direction applied in Material 1 at
(0,0,c > 0), we assume the potential functions in the two half-spaces (x3 > 0 and x3 < 0,
or z > 0 and z < 0) as

Py(1) = 0 (z ≥ 0 )
(5.58)
Py( 2) =0 (z ≤ 0 )

Then the interface conditions (5.37a, b) at z = 0 are reduced to

1 − ν1
κ 1 Px(1) − x∂ x Px(1) − ∂ x q(1) = [κ 2 Px(2) − x∂ x Px(2) − ∂ x q(2) ]
1 − ν2
1 − ν1
   x Px(1) + q(1) = ( x Px( 2) + q( 2) )  (5.59a, b, c)
1 − ν2
1 − ν1
κ 1 Pz(1) − x∂z Px(1) − ∂z q(1) = [κ 2 Pz( 2) − x∂z Px( 2) − ∂zq( 2) ]
1 − ν2

µ1 1 1 1
[(1 − 2 ν1 )(∂ x Pz( ) + ∂ z Px( ) ) − x∂ x ∂ z Px( ) − ∂ x ∂ z q(1) ]
(1 − ν1 )
µ2
= [(1 − 2 ν2 )(∂ x Pz( 2) + ∂z Px( 2) ) − x∂ x ∂z Px( 2) − ∂ x ∂z q( 2) ]
(1 − ν2 )
µ1 µ2
[(1 − 2 ν1 ) Pz(1) − x∂ z Px(1) − ∂ z q(1) ] = [(1 − 2 ν2 ) Pz(2) − x∂ z Px( 2) − ∂z q( 2) ]
(1 − ν1 ) (1 − ν2 )
µ1
[2(1 − ν1 )∂z Pz(1) + 2 ν1∂ x Px(1) − x∂z ∂z Px(1) − ∂z ∂z q(1) ]
(1 − ν1 )
µ2
= [2(1 − ν2 )∂z Pz( 2) + 2 ν2 ∂ x Px(2) − x∂z ∂z Px(2) − ∂z ∂z q( 2) ] (5.60a, b, c)
(1 − ν2 )

On the source side (Material 1), we now express the unknown functions as a sum of
the surface and volume integral-related terms (Pz is harmonic because the force is
in the x-direction)

Px ≡ P = Ps + Pv
(5.61)
q = qs + qv

Making use of the Green’s function integral expression (5.43), we have

1
4 πµ1 ∫V
Pv ( x ) = G( y; x ) fx ( y)dV ( y)
(5.62a, b)
−1
4 πµ1 ∫V
qv ( x) = G( y; x) y1 fx ( y)dV ( y)
5.3 Papkovich Functions with Perfect Interface 151

Substituting

fx ( y) = fx δ( y1 )δ( y2 )δ( y3 − c) (5.63)

into Eqs. (5.62a, b) gives

fx
Pv ( x ) = (1 / R1 − 1 / R2 )
4πµ 1 (5.64a, b)
qv ( x ) = 0

In other words, q in the source side, that is, Material 1, is also harmonic, and Px is the
only function that has the nonharmonic term.
Taking the derivative with respect to x of Eq. (5.60b) and adding the result to
Eq. (5.60a), we have, on the interface z = 0,

Px(1) = Px( 2) (5.65)

Because Pv = 0 on the interface z = 0 (R1 = R2 on the interface), Eq. (5.65) can be


rewritten as

Ps = Px( 2) (5.66)

Similarly, taking the derivative of Eq. (5.60b) with respect to x and subtracting the
result from Eq. (5.60a), we have on the interface z = 0,

µ2
∂z Px(1) = ∂z Px( 2) (5.67)
µ1

Replacing Px(1) in Material 1 by Ps+Pv, we have (∂zR1 = – ∂zR2 on z = 0, and replacing


R1 by R2)

fx ∂ 1 µ
∂z Ps − = 2 ∂z Px( 2) (5.68)
2 πµ1 ∂z R2 µ1

Replacing the right-hand side of Eq. (5.68) by Eq. (5.66), we have (noticing again
L(z) = R(−z))

fx ∂ 1 µ
∂z Ps − = − 2 ∂z Ps (5.69)
2 πµ1 ∂z R2 µ1

Therefore, we find the solution for Ps from the preceding equation as

fx 1 µ
Ps − = − 2 Ps (5.70)
2πµ 1 R2 µ1

which gives the solution of Ps as


−1
 µ  fx 1
Ps =  1 + 2  (5.71)
 µ1  2πµ1 R2
152 Green’s Functions in Elastic Isotropic Full and Bimaterial Spaces

By adding Pv in Eq. (5.64a), we finally have Px in Material 1 as

µ 2 −1 f x 1 f
Px = (1 + ) + x (1 / R1 − 1 / R2 )
µ1 2 πµ 1 R2 4 πµ 1
(5.72)
f  1 µ − µ2 1 
= x  + 1
4πµ 1  R1 µ 1 + µ 2 R2 

Making use of Eq. (5.66), we also have Px( 2) in Material 2 as (also replacing R2
by R1),

fx 1
Px( 2) = (5.73)
2 π(µ 1 + µ 2 ) R1

Now, from Eq. (5.60b), we have, on the interface z = 0,

x fx 1 1 − ν1  x fx 1 
+ q(1) = ( 2)
 2 π(µ + µ ) R + q  (5.74)
2 π(µ 1 + µ 2 ) R2 1 − ν2  1 2 1 

where we have made use of Eqs. (5.72) and (5.73), and also noticed the fact that
R2 = R1 on z = 0.
Moving the R1-term to the left-hand side and remembering that R2 = R1, we have
the following relation between q(1) and q(2)

( ν1 − ν2 ) fx x 1 − ν1 ( 2)
+ q(1) = q (5.75)
2 π(1 − ν2 )(µ1 + µ 2 ) R2 1 − ν2

Replacing x/R2 by a new harmonic function, which has the same value as x/R2 on the
interface z = 0, Eq. (5.75) can be rewritten as

( ν1 − ν2 ) f x x(R2 + c ) 1 − ν1 ( 2)
+ q(1) = q (5.76)
2 π(1 − ν2 )(µ1 + µ 2 ) R2 R2* 1 − ν2

Substituting Eqs. (5.72), (5.73) and (5.76) to Eqs. (5.60a, b, c), we find, from Eq.
(5.60a)

(1 − ν2 )κ 1 (1) 2(1 − ν2 )  µ  (1 − ν2 ) cf x  1
Pz − ∂ z q(1) −  1 − 2  ∂x
(1 − ν1 )κ 2 (1 − ν1 )κ 2  µ 1  (1 − ν1 ) 2 πκ 2 (µ 1 + µ 2 )  R2 
( ν1 − ν2 ) f x (5.77)
+ ∂ x (ln R2* ) = Pz( 2 )
2 π (1 − ν1 )κ 2 (µ1 + µ 2 )

where we have made use of the following relation:

(R2 + c) 1 z
= − (5.78)
R2 R2* R2 R2 R2*

and the following relation on the interface z = 0,


5.3 Papkovich Functions with Perfect Interface 153

x c
∂z = ∂x (5.79)
R2 R2

From Eq. (5.60b), we have

µ 1 (1 − ν2 )(1 − 2 ν1 ) (1) (µ 2 + µ 1 )(1 − ν2 )


Pz − ∂ z q(1)
µ 2 (1 − ν1 )(1 − 2 ν2 ) µ 2 (1 − 2 ν2 )(1 − ν1 )
(5.80)
( ν1 − ν2 ) fx
+ ∂ x (ln R2* ) = Pz( 2)
2 π (µ 1 + µ 2 )(1 − ν1 )(1 − 2 ν2 )

From Eq. (5.60c), we have

µ1 1  µ   µ  fx  c 
∂ z Pz(1) +  1 − 1  ∂ z ∂ z q(1) +  1 − 1  ∂ x ∂z  
µ2 2(1 − ν1 )  µ2   µ 2  4 π(1 − ν1 )(µ 1 + µ 2 )  R2 
(5.81)
 µ  (1 − 2 ν1 ) fx
+ 1− 1  ∂ x ∂ z (ln R2* ) = ∂ z Pz( 2 )
 µ 2  4 π(1 − ν1 )(µ 1 + µ 2 )

In deriving Eq. (5.81), we have made use of the following relations on the interface
z = 0:

x 1 c
∂z ∂z = ∂x + ∂ x ∂z
R2 R2 R2
−x
∂ x ∂z ln R2* = ∂z ∂ x ln R2* = 3 (5.82a, b, c)
R2
z −2
∂z ∂z =
R2 R2* R23

Integrating both sides of Eq. (5.81) with respect to z, and recalling that L(z) = R(−z),
we obtain

µ 1 (1) 1  µ   µ  fx  c 
Pz + 1 − 1  ∂ z q(1) +  1 − 1  ∂x
µ2 2(1 − ν1 )  µ2   µ 2  4π(1 − ν1 )(µ 1 + µ 2 )  R2 
   (5.83)
 µ  (1 − 2 ν1 ) fx
+ 1 − 1  ∂ x (ln R2* ) = −Pz( 2 )
 µ 2  4 π(1 − ν1 )(µ1 + µ 2 )

Inserting Eq. (5.80) into Eq. (5.77), and Eq. (5.80) into Eq. (5.83) (in order to get rid
of Pz( 2)), we then obtain the following two equations

(1 − ν2 )κ 1 (1) µ 1 (1 − ν2 )(1 − 2 ν1 ) (1) 2(1 − ν2 ) (µ 2 + µ 1 )(1 − ν2 )


Pz − Pz − ∂z q(1) + ∂ z q(1)
(1 − ν1 )κ 2 µ 2 (1 − ν1 )(1 − 2 ν2 ) (1 − ν1 )κ 2 µ 2 (1 − 2 ν2 )(1 − ν1 )
 µ  (1 − ν2 ) cfx  1 (ν1 − ν2 ) fx
− 1− 2  ∂x   + ∂ x (ln R2* )
 µ 1  (1 − ν1 ) 2 πκ 2 (µ 1 + µ 2 )  R2  2π (1 − ν1 )κ 2 (µ 1 + µ 2 )
(ν1 − ν2 ) fx
− ∂ x (ln R2* ) = 0
2π (µ 1 + µ 2 )(1 − ν1 )(1 − 2 ν2 )
 (5.84)
154 Green’s Functions in Elastic Isotropic Full and Bimaterial Spaces

µ 1 (1) µ 1 (1 − ν2 )(1 − 2 ν1 ) (1) 1  µ  (µ 2 + µ 1 )(1 − ν2 )


Pz + Pz +  1 − 1  ∂ z q(1) − ∂ z q ( 1)
µ2 µ 2 (1 − ν1 )(1 − 2 ν2 ) 2(1 − ν1 )  µ2  µ 2 (1 − 2 ν2 )(1 − ν1 )
 µ  fx  c   µ  (1 − 2 ν1 ) f x
+ 1− 1  ∂x + 1− 1  ∂ x (ln R2* )
 µ 2  4π (1 − ν1 )(µ1 + µ 2 )  R2   µ 2  4 π (1 − ν1 )(µ 1 + µ 2 )
( ν1 − ν2 ) fx
+ ∂ x (ln R2* ) = 0
2 π (µ 1 + µ 2 )(1 − ν1 )(1 − 2 ν2 )
 (5.85)

Equations (5.84) and (5.85) can be rewritten as

[κ 1 (1 − 2 ν2 )µ 2 / µ 1 − κ 2 (1 − 2 ν1 )]Pz(1) + (µ 2 / µ + κ 2 )∂ z q(1)
µ  µ  (1 − 2 ν2 )cfx  1  µ 2 (ν1 − ν2 ) fx (5.86)
− 2 1 − 2  ∂x − ∂ x (ln R2* ) = 0
µ1  µ 1  2π (µ 1 + µ 2 )  R2  µ 1 π (µ 1 + µ 2 )

2[(1 − ν2 )(1 − 2 ν1 ) + (1 − 2 ν2 )(1 − ν1 )]Pz(1) − (κ 2 + µ 2 / µ 1 )∂ z q(1)


 µ  (1 − 2 ν2 ) fx  c 
− 1− 2  ∂x
 µ1  2 π(µ 1 + µ 2 )  R2  (5.87)
µ [2( ν1 − ν2 ) − (1 − 2 ν2 )(1 − 2 ν1 )(µ1 / µ 2 − 1)] fx
+ 2 ∂ x (ln R2* ) = 0
µ1 2 π (µ 1 + µ 2 )

Solving Pz(1) and q(1) in Material 1 from these two equations, we finally have

(µ 1 − µ 2 ) f x  cx 1 − 2 ν1 x 
Pz(1) = − +
2 π (µ 1 + κ 1µ 2 )  µ 1 R2 µ 1 + µ 2 R2 R2* 
 3

fx  (1 − 2 ν1 )(µ 1 − µ 2 )cx Ax 
q(1) = + *
2 π (µ 1 + µ 2 )(µ 1 + κ 1µ 2 )  R2 R2* R2  (5.88a,b)

where

[µ 2 κ 1 (1 − 2 ν2 ) − µ 1κ 2 (1 − 2 ν1 )](1 − 2 ν1 )(µ 1 − µ 2 ) − 2µ 2 ( ν1 − ν2 )(µ 1 + κ 1µ 2 )


A=
µ 2 + κ 2 µ1
 (5.89)

The other two functions Pz( 2) and q(2) in Material 2 can be solved from Eqs. (5.76)
and (5.80) as

(1 − 2 ν2 )(µ 1 − µ 2 ) fx x
Pz( 2) =
2 π (µ 1 + µ 2 )(µ 2 + κ 2 µ 1 ) R1 R1*
(1 − ν2 ) fx
q( 2) = (5.90a,b)
2 π (1 − ν1 )(µ 1 + µ 2 )(µ 1 + κ 1µ 2 )
  ( ν − ν2 )(µ 1 + κ 1µ 2 )  cx
×  (1 − 2 ν1 )(µ 1 − µ 2 ) + 1  R R*
  1 − ν2  1 1
 
( ν − ν2 )(µ 1 + κ 1µ 2 ) x  
+ A + 1  R* 
 1 − ν2  1 
5.4 Papkovich Functions with Smooth Interface 155

where R1* = R1 − (z − c). Thus when a concentrated horizontal force Fx is applied


at (0,0,c > 0) in Material 1, we have derived the Papkovich functions, which will be
utilized to find the corresponding displacement and stress fields.

5.4 Papkovich Functions in an Elastic Isotropic Bimaterial Space


with Smooth Interface
For the smooth interface case, we have, along the interface x3≡z = 0,

uz(1) = uz( 2) , σ(zz1) = σ(zz2)


(5.91)
σ(xz1) = σ(xz2) = 0, σ(yz
1) ( 2)
= σ yz = 0

By following the same process but using the smooth interface condition (5.91), the
Papkovich functions for the two material half-spaces can be found (Dundurs and
Hetenyi 1965).

5.4.1 A Point Force Normal to the Interface Applied in Material 1


When a vertical point force Fz is applied in Material 1 at (0,0,c > 0), the nonzero
Papkovich functions in the two half-spaces with smooth interface are:
In Material 1 (z > 0):

 1 Ds
fz κ 2c(z + c)  
Pz( 1) =  − + (1 − D)  1 + 
4 πµ 1
 R1 R2  R2 R23  
(5.92)
f 1 c [κ 1 − D(κ 1 − 1)]c 
q( 1)
= z  (1 − D)(κ 12 − 1) ln R2* − − 
4πµ 1  2 R1 R2 

In Material 2 (z < 0):


fz 1 2Ds (κ 2 + 1) c(z − c) 
Pz( 2) =  D(κ 2 + 1) − 
4 πµ 2
 R1 (κ 1 + 1) R13 
(5.93)
f  D(1 − κ 22 ) c D(1 − κ 22 ) 
q( 2) = z − + ln R1* 
4πµ 2  (κ 1 + 1) R1 2 

In Eqs. (5.92) and (5.93), while parameter Ds will be defined later, D is defined as
follows:

(κ 1 + 1)Γ 1
D= ≡ (1 + α) (with Γ ≡ µ 2 / µ1 ) (5.94)
(κ 1 + 1)Γ + κ 2 + 1 2

where α is one of the Dundurs’ parameters defined in Eq. (3.85).


156 Green’s Functions in Elastic Isotropic Full and Bimaterial Spaces

5.4.2 A Point Force Parallel to the Interface Applied in Material 1


When a horizontal point force in the x-direction fx is applied in Material 1 at
(0,0,c > 0), the nonzero Papkovich functions in the two half-spaces with smooth
interface are:
In Material 1 (z > 0):

fx  cx (κ 1 − 1)2 x 
q(1) = (1 − D) (κ 1 − 1) −
4 πµ 1  R2 R2* 2 R2* 
(5.95)
f  1 1 f  x 2cx 
P x
(1)
= x  +  , Pz(1) = x (1 − D) (κ 1 − 1) − 3
4πµ 1  R1 R2  4 πµ 1  R2 R2 R2 
*

In Material 2 (z < 0):

fx D(κ 22 − 1)  2cx x 
q( 2) = − + (κ 1 − 1) * 
8πµ 2 (κ 1 + 1)  R1 R1* R1 
(5.96)
f D(κ 2 + 1)  x 2cx 
P( 2)
x = 0, P z
(2)
= x (κ 1 − 1) − 3
4 πµ 2 (κ 1 + 1)  R1 R1 R1 
*

5.5 Papkovich Functions for Both Perfect-Bonded and Smooth


Interfaces of Bimaterial Spaces
By comparing the Papkovich functions in Section 5.4 for the smooth interface case
with those in Section 5.3 for the perfect-bonded interface case, we can uniformly
express the Papkovich potential functions for both perfect-bonded and smooth
interface cases. This is presented in the following text.

5.5.1 Papkovich Functions for a Vertical Point Force in Material 1


In Material 1 (z > 0):

fz  1 Ds κ 2c(z + c)  
Pz(1) =  − + k1  1 + 
4πµ 1  R1 R2  R2 R23  
(5.97)
f  c c 
q(1) = z  k3 ln R2* − − k2
4 πµ 1  R1 R2 

In Material 2 (z < 0):

k4 fz c(z − c)
Pz( 2 ) = − Ds m1 fz
πR1 πR13
(5.98)
k cf k f
q (2)
= − 5 z + 6 z ln R1*
πR1 π

In Eqs. (5.97) and (5.98), the involved material coefficients are defined as follows.
5.5 Papkovich Functions for Both Perfect-Bonded 157

For the perfect-bonded interface, the parameters Ds and ki are defined by

(1 − Γ )
Ds = 0, k1 = , k2 = k1κ 1
1 + κ 1Γ
(1 + κ 1 )[κ 2 (κ 1 − 1) − κ 1 (κ 2 − 1)Γ ]
k3 = (5.99)
2(κ 2 + Γ )(1 + κ 1Γ )
(1 + κ 2 ) (1 + κ 2 ) (1 + κ 2 )
k4 = , k5 = , k6 = k3
4µ 1 (κ 2 + Γ ) 4µ1 (1 + κ 1Γ ) 4µ1 (1 + κ 1 )

For the smooth interface case, they are

(κ 2 + 1) 1
Ds = D, m1 = , k1 = 1 − D, k2 = κ 1 − D(κ 1 − 1), k3 = (1 − D)(κ 12 − 1)
2µ 2 (κ 1 + 1) 2
 k4 = D(κ 2 + 1) D ( 1 − κ 2)
2 D(1 − κ 2)
2
, k5 = , k6 =
4µ 2 4µ 2 (κ 1 + 1) 8µ 2
(5.100)

5.5.2 Papkovich Functions for a Horizontal Point Force in Material 1


In Material 1 (z > 0):

fx  cx x 
q(1) =  k8 + k9 * 
2π  R2 R2 *
R2 
(5.101)
f  1 1 f  x cx 
P ( 1)
= x  + 2 µ 1 k7 , Pz(1) = x  k8 − k10 3 
R2 
x
4 πµ 1  R1 2 π  R2 R2* R2 

In Material 2 (z < 0):

fx  cx x 
q( 2) =  k11 + k12 * 
2π  R1 R1
*
R1 
(5.102)
( 2) f k F  x cx 
Px = x 13 , Pz( 2) = x  k14 + Ds m2 3 
2π R1 2π  R1 R1
*
R1 

In Eqs. (5.101) and (5.102), the involved material constants are defined as follows.
For the perfect-bonded interface, we have

1− Γ k (κ − 1)
Ds = 0; k7 = , k8 = 7 1
2µ 1 (1 + Γ ) 1 + κ 1Γ
B 1− Γ
k9 = , k10 =
µ 1 (1 + Γ )(1 + κ 1Γ ) µ 1 (1 + κ 1Γ )
k7 (κ 2 + 1)(κ 1 − 1) (κ 2 − κ 1 )
k11 = + (5.103)
(κ 1 + 1)(1 + κ 1Γ ) µ 1 (κ 1 + 1)(1 + Γ )
(κ + 1)k9 (κ 2 − κ 1 ) 1 k (κ − 1)
k12 = 2 + , k13 = , k14 = 7 2
(κ 1 + 1) µ1 (κ 1 + 1)(1 + Γ ) µ1 (1 + Γ ) Γ + κ2
[κ (κ − 1)Γ − κ 2 (κ 1 − 1)](κ 1 − 1)(1 − Γ ) − 2(κ 2 − κ 1 )( 1 + κ 1Γ )Γ
B= 1 2
4(Γ + κ 2 )
158 Green’s Functions in Elastic Isotropic Full and Bimaterial Spaces

For the smooth interface case, we have

−(κ 2 + 1)
Ds = 1, m2 = ,
µ 2 (κ 1 + 1)
1 (1 − D)(κ 1 − 1) (D − 1)(κ 1 − 1)2 (1 − D)
k7 = , k8 = , k9 = , k10 =
2µ 1 2µ 1 4µ 1 µ1
D(κ 2 − 1)
2 D(κ 2 − 1)(κ 1 − 1)
2
D(κ 2 + 1)(κ 1 − 1)
k11 = − , k12 = , k13 = 0, k14 =
2µ 2 (κ 1 + 1) 4µ 2 (κ 1 + 1) 2µ 2 (κ 1 + 1)
 (5.104)

5.6 Green’s Displacements and Stresses in Elastic Isotropic


Bimaterial Spaces

5.6.1 Green’s Displacements and Stresses in Bimaterial Spaces


by a Vertical Point Force
To derive the point-force induced displacement fields (for both perfect-bonded
and smooth interfaces), we just make use of Eq. (5.23b) and of the derivatives in
Appendix A. The Green’s displacements due to a concentrated vertical force fz
applied in Material 1 at (0,0,c) are listed as follows.
In Material 1 (z > 0):

fz x  z − c k1κ 1z − k2 c − Ds z 6k1cz(z + c) k 
    uxz =  + + − 3 *
4 π (κ 1 + 1)µ 1  R1 3
R23
R25
R2 R2  (5.105a)

uyz ( x, y, z; 0, 0, c) = uxz ( y, x, z; 0, 0, c) (5.105b)

fz  κ 1 (z − c)2 k1κ 12 − k3 − Ds κ 1 6k1cz(z + c)2


uzz = + + +
4 πµ 1 (κ 1 + 1)  R1 R13 R2 R25
   (5.105c)
k κ (z + 2c)(z + c) − 2k1cz − (k2 c + Ds z)(z + c) 
+ 1 1 
R23 

In Material 2 (z < 0):

fz x  k4 z − k5 c 3Ds m1cz(z − c) k 
uxz = − − 6* (5.106a)
π (κ 2 + 1)  R13 R15 R1 R1 

uyz ( x, y, z; 0, 0, c) = uxz ( y, x, z; 0, 0, c) (5.106b)

fz  k4 κ 2 k6 cz(z − c)2
uzz =  + − 3Ds m1
π (κ 2 + 1)  R1 R1 R15
   (5.106c)
k z(z − c) k5 c(z − c) zc κ c(z − c) 
+ 4 3 − + Ds m1 3 − Ds m1 2 3 
R1 R13
R1 R1 
5.6 Green’s Displacements and Stresses in Elastic Isotropic Bimaterial Spaces 159

Taking the derivatives of the obtained Green’s displacements with respect to the
field point (x,y,z) and making use of the constitutive relations (5.6), we can also find
the Green’s stresses induced by the concentrated vertical force fz applied at (0,0,c >
0). These are listed as follows:
In Material 1 (z > 0):

2 µ 1 ν1 z fz  z − c k1κ 1z − k2 c − Ds z 6k1cz(z + c) k 
σ zxx = e1 +  + + − 3 *
1 − 2 ν1 2 π(κ 1 + 1)  R1 3
R23
R25
R2 R2 
fz x 2  3(z − c) 3(k1κ 1 z − k2 c − Ds z)
+  − R5 −
2π (κ 1 + 1)  1 R25
30k1cz(z + c)  1 1 
− + k3  3 * + 2 * 2  
R27
 R2 R2 R2 R2  
 (5.107a)

z ( x, y, z; 0, 0, c) = σ z ( y, x, z; 0, 0, c)
σ yy (5.107b)
xx

2µ 1 ν1 z
σ zzz = e1
1 − 2 ν1
fz  (κ 1 − 2)(z − c) 3(z − c)3 30k1 cz(z + c)3
+ − − −
2π(κ 1 + 1)  R13 R15 R27
−k κ (κ − 2)(z + c) + k1 (κ 1 − 2)c + Ds (1 − κ 1 )(z + c) − (k2 c + Ds z) + k3 (z + c)
+ 1 1 1
R23
6k c(z + c)(4z + c) − 3k1κ 1 (z + 2c)(z + c)2 3(k2 c + Ds z)(z + c)2 
+ 1 + 
R25 R25 
 (5.107c)

fz x  (κ 1 − 1) − k1κ 1 (κ 1 − 1) / 2 + k3 + Ds (κ 1 − 1) / 2
σ zxz = − +
2 π (κ 1 + 1)  2 R13 R23
3(z − c)2 30k1cz(z + c)2
− −
R15 R27
3k c(3z + c) −3k1κ 1 (z + c)2 + 3(k2 c + Ds z)(z + c) 
+ 1 5 + 
R2 R25  (5.107d)

z ( x, y, z, 0, 0, c) = σ z ( y, x, z, 0, 0, c)
σ yz (5.107e)
xz

fz xy  3(z − c) 3(k1κ 1z − k2 c − Ds z)
σ zxy = − −
2 π (κ 1 + 1)  R15 R25
     (5.107f)
30k1cz(z + c)  1 1 
− + k3  3 * + 2 * 2  
R27
 R2 R2 R2 R2  
160 Green’s Functions in Elastic Isotropic Full and Bimaterial Spaces

where the dilatation e ≡ ux, x + uy, y + uz,z in Material 1 (z > 0) is given by

fz (κ 1 − 1)  (z − c) 6k1c(z + c)2 − k1κ 1 (z + c) + 2k1c + Ds (z + c) 


e1z = − R3 − +  (5.108)
4π (κ 1 + 1)µ 1  1 R25 R23 

In Material 2 (z < 0):

2 µ 2 ν2 z 2µ 2 fz  k4 z − k5 c 3Ds m1cz(z − c) k 
σ zxx = e2 +  − − 6 *
1 − 2 ν2 π(κ 2 + 1)  R1 3
R15
R1 R1 
2µ 2 fz x 2  3(k4 z − k5 c) 15Ds m1cz(z − c)  1 1 
+ − + + k6  3 * + 2 *2  
π(κ 2 + 1)  R15
R17
 R1 R1 R1 R1  
 (5.109a)

σ zyy( x, y, z; 0, 0, c) = σ zxx ( y, x, z; 0, 0, c) (5.109b)

2µ 2 ν2 z
σ zzz = e2
1 − 2 ν2
2µ 2 fz  3(k5 c − k4 z)(z − c)2 3Ds m1 (κ 2 − 1)c(z − c)2 9Ds m1cz(z − c)
+  + −
π (κ 2 + 1)  R15 R15 R15
[k4 (κ 2 − 1) + k6 ](z − c) k4 z − k5 c Ds m1 (κ 2 − 1)c 15Ds m1 cz(z − c)3 
− + − + 
R13 R13 R13 R17 
 (5.109c)

µ 2 fz x  k4 (κ 2 − 1) + 2k6 6(k4 z − k5 c)(z − c)


σ zxz = − −
π (κ 2 + 1)  R13 R15
  
3Ds m1 (κ 2 − 1)c(z − c) − 3Ds m1cz 30 Ds m1cz(z − c)2  (5.109d)
+ + 
R15 R17 

z ( x, y, z; 0, 0, c) = σ z ( y, x, z; 0, 0, c)
σ yz (5.109e)
xz

2µ 2 fz xy  3(k4 z − k5 c) 15Ds m1cz(z − c)  1 1 


σzxy =
− + + k6  3 * + 2 *2  
π (κ 2 + 1)  R15
R17
 R1 R1 R1 R1  

(5.109f)

where the dilatation in Material 2 (z < 0) is given by

fz (κ 2 − 1)  k4 (z − c) Ds m1 c 3Ds m1 c(z − c)2 


e2z = − − +
π (κ 2 + 1)  R13 R13 R15 

(5.110)
5.6 Green’s Displacements and Stresses in Elastic Isotropic Bimaterial Spaces 161

5.6.2 Green’s Displacements and Stresses in Bimaterial Spaces by a Horizontal


Point Force
Making use of Eq. (5.23b), we find the bimaterial Green’s displacements due to a
horizontal point force fx applied in Material 1 at (0,0,c > 0) as follows.
In Material 1 (z > 0):

fx  κ1 x2 k7 κ 1 − k8 (k7 + k8 )x 2 + k10 cz
uxx =  + + +
2 π (κ 1 + 1)  2µ 1 R1 2µ 1 R13 R2 R23
(5.111a)
3k czx 2 (k8 − k9 ) (k8 − k9 ) x 2 
− 10 5 + −
R2 R2* R2 R2*2 

f x xy  1 k +k 3k cz (k − k ) 
uyx =  + 7 3 8 − 105 − 8 *29  (5.111b)
2π (κ 1 + 1)  2µ 1 R13 R2 R2 R2 R2 

fx x  (z − c) (k7 + k8 )(z + c) − k10 κ 1c 3k10 cz( z + c) k9 + k8 κ 1 


uzx =  + − + 
2 π (κ 1 + 1)  2µ 1 R13 R23 R25 R2 R2* 
 (5.111c)

In Material 2 (z < 0):


fx  k11c + k14z  1  k c + k14z  x 2 zc czx 2
uxx =   k13κ 2 +  +  k13 − 11  − Ds m2 + 3Ds m2
2 π (κ 2 + 1)  z−c R1  z − c  R13 R13 R15
 k c + k14z  1  k c + k14z  x 2 
−  k12 + 11  +  k12 + 11  
 z − c  R1 
*
z − c  R1R1*2 
 (5.112a)

fx xy  k14 z + k11c  1 3Ds m2 cz  k z + k11c  1 


uyx =   k13 −  3 + +  k12 + 14 
2 π (κ 2 + 1)  z−c R1 R15  z − c  R1 R1*2 
 (5.112b)

fx x  k13 (z − c) − (k14 z + k11c ) + Ds m2 κ 2 c 3Ds m2 zc( z − c) k14 κ 2 − k12 


uzx = + +
2π (κ 2 + 1)  
R13 R15 R1 R1* 
 (5.112c)

Taking the derivatives of the Green’s displacements and making use of the constitutive
relations (5.6), we can also derive the following bimaterial Green’s stresses.
In Material 1 (z > 0):

2µ 1 ν1 x
σ xxx = e1
1 − 2 ν1
µ 1 fx x  (κ 1 − 2) 3x 2 − k7 (κ 1 − 2) + 3k8 3(k7 + k8 ) x 2 + 9k10 cz
+ − − + −
π (κ 1 + 1)  2µ 1 R13 2µ 1 R15 R23 R25
15k10 czx 2 3(k8 − k9 )  1 2 
+ − + (k8 − k9 ) x 2  3 *2 + 2 *3  
R27
R2 R2
*2
 R2 R2 R2 R2  
 (5.113a)
162 Green’s Functions in Elastic Isotropic Full and Bimaterial Spaces

2µ 1 ν1 x µ 1 fx x  1 (k + k ) 3k cz (k − k ) 
σ xyy = e1 +  + 7 3 8 − 105 − 8 *29 
 1 − 2 ν1 π (κ 1 + 1)  2µ 1 R1
3
R2 R2 R2 R2 
µ 1 fx xy 2  3 3(k7 + k8 ) 15k10 cz  1 2 
+ − − + + (k8 − k9 )  3 *2 + 2 *3  
π (κ 1 + 1)  2µ 1 R1
5
R25
R27
 R2 R2 R2 R2  
(5.113b)

2µ 1 ν1 x
σ zz
x
= e1
1 − 2 ν1
µ 1 fx x  1 3(z − c)2 k7 − k9 − k8 (κ 1 − 1) 15k10 cz( z + c) 2
+  − + + (5.113c)
π (κ 1 + 1)  2µ 1 R1
3
2µ 1 R15 R23 R27
3k10 κ 1c(z + c) − 3k10 c(2 z + c) − 3(k7 + k8 )(z + c)2 
+ 
R25 

µ 1 fx  (κ 1 − 1)(z − c) [2k8 − k7 (κ 1 − 1)](z + c) − k10 (κ 1 − 1)c 6k10 cz(z + c)


σ xxz = − + −
2 π (κ 1 + 1)  2µ 1 R13 R23 R25
6 x 2 (z − c) 3k10 (κ 1 − 1)c − 6(k7 + k8 )(z + c) 2 30k10 czx 2 (z + c) 
− + x + 
2µ 1 R15 R25 R27 
µ 1 fx  2k9 + k8 ( f1 − 1)  1 1 
+  − [2k9 + k8 ( f1 − 1)]x 2  3 * + 2 *2  
2 π (κ 1 + 1)  R2 R2*  R2 R2 R2 R2  
 (5.113d)

µ 1 fx xy  3(z − c) 3k10 (κ 1 − 1)c − 6(k7 + k8 )(z + c) 30k10 cz(z + c)


σ xyz = − + +
2 π (κ 1 + 1)  µ 1 R15 R25 R27
 1 1 
−[2k9 + k8 (κ 1 − 1)]  3 * + 2 *2  
 R2 R2 R2 R2  
 (5.113e)

µ 1 fx y  −(κ 1 − 1) 2k8 − k7 (κ 1 − 1) 6k10 cz 2(k8 − k9 ) 


σ xxy =  + − − 
2 π (κ 1 + 1)  2µ 1 R13 R23 R25 R2 R2*2 
µ f x2 y  3 6(k7 + k8 ) 30k10 cz  1 2 
+ 1 x − − + + 2(k8 − k9 )  3 *2 + 2 *3  
2π (κ 1 + 1)  µ 1 R1 5
R25
R27
 R2 R2 R2 R2  
 (5.113f)

where the dilatation in Material 1 (z > 0) is given by

fx x(κ 1 − 1)  1 (k + k ) 3k c(z + c) 
   e1x = −  + 7 3 8 − 10 5  (5.114)
2π (κ 1 + 1)  2µ 1 R1
3
R2 R2 
5.6 Green’s Displacements and Stresses in Elastic Isotropic Bimaterial Spaces 163

In Material 2 (z < 0):

2µ 2 ν 2 x
σ xxx = e2
1 − 2 ν2
µ 2 fx x   3(k11c + k14 z)  1 cz czx 2
+  −  k13 (κ 2 − 2) +  R3 + 9 Ds m 2 − 15 Ds m2
π (κ 2 + 1)   z−c  1 R15 R17
 k c + k14 z  x 2
 k c + k14 z  1
− 3  k13 − 11  + 3  k12 + 11 
 z − c  R1 5  z − c  R1R1*2
 k c + k14 z   1 2  
− x 2  k12 + 11   + 2 *3  
 z − c   R1 R13 * 2
R1 R1  
 (5.115a)

2µ 2 ν2 x
σ xyy = e2
1 − 2 ν2
µ 2 fx x  k14 z + k11c  1 3Ds m2cz  k z + k11c  1 
+  k13 −  + +  k12 + 14 
π (κ 2 + 1)  z − c  R13 R15  
z − c  R1R1*2 
µ 2 fx xy 2   k14z + k11c  1 15Ds m2cz  k z + k11c   1 2 
−  3  k13 −  5 + +  k12 + 14   3 *2 + 2 *3  
π (κ 2 + 1)   z−c R1 R17  z−c  R1 R1 R1 R1  
 (5.115b)

2 µ 2 ν2 x
σ zz
x
= e2
1 − 2 ν2
µ 2 fx x  k13 − k12 + k14 (k2 − 1) 3Ds m2 (κ 2 − 1)c( z − c) + 3Ds m2 zc
+  −
π (κ 2 + 1)  R13 R15
3[k13 (z − c) − (k14 z + k11c)]( z − c) 15Ds m2 zc( z − c)2  (5.115c)
− − 
R15 R17 

µ 2 fx
σ xxz =
2 π (κ 2 + 1)
 (k + k )c 1  2(k11c + k14z)  (z − c) Ds m2 (κ 2 − 1)c
×  − 11 14 2 − k13 (κ 2 − 1) +  R3 +
 ( − z + c ) R1  z−c  1 R13
6Ds m2zc( z − c) (k11 + k14 )c 1  k c + k14 z  1 
+ + +  k14κ 2 − 2k12 − 11  
R15
(− z + c) R1 
2 *
z − c  R1R1* 
µ 2 fx x 2
+
2 π (κ 2 + 1)
 (k + k )c 1 6[k13 (z − c) − (k14 z + k11c)] 3Ds m2 (κ 2 − 1)c 30 Ds m2zc( z − c)
×  11 14 2 − − −
 ( −z + c) R1 R15 R15 R17
3

(k + k )c 1  k c + k14z   1 1  
− 11 14 2 −  k14 m2 − 2k12 − 11   + 2 *2  
(−z + c) R1R1 
*2
z − c   R1 R1 R1 R1  
3 *

 (5.115d)
164 Green’s Functions in Elastic Isotropic Full and Bimaterial Spaces

µ 2 fx xy
σ xyz =
2 π (κ 2 + 1)
 (k + k14 )c  1 1  6[k13 (z − c) − (k14z + k11c)]
×  11 2 
− *2 

 ( − z + c)  R1 R1R1  R15
3

3Ds m2 (κ 2 − 1)c 30Ds m2zc( z − c)  k z + k11c   1 1  


− − −  k14κ 2 − 2k12 − 14   3 * + 2 *2  
R15
R17  z−c  R1 R1 R1 R1  
 (5.115e)

µ 2 fx y
σ xxy =
π (κ 2 + 1)
  k (κ + 1) k11c + k14 z  1  k11c + k14 z  1 3Ds m2 cz 
×  −  13 2 +  R 3 +  k12 + z − c  R R*2 + 
  2 z − c  1 1 1 R15 
µ f x2 y
− 2 x
π (κ 2 + 1)
  k z + k11c  1 15Ds m2 cz  k c + k14 z   1 2 
×  3  k13 − 14  + +  k12 + 11   + 2 *3  
  z − c  R1 R1  z − c   R1 R1 R1 R1  
5 7 3 * 2

 (5.115f)

where the dilatation in Material 2 (z < 0) is given by

fx x(κ 2 − 1)  k14 − k13 3Ds m2 c( z − c) 


e2x = − (5.116)
2 π (κ 2 + 1)  R13 R15 

Remark 5.5: It can be shown that the perfect-bonded or smooth conditions of


displacements and tractions on the interface from both sides of the half-spaces are
satisfied (letting z approach the interface from both half-spaces).
Remark 5.6: When the source point (0,0,c > 0) approaches the interface (from
Material 1), our bimaterial Green’s displacements and stresses are then reduced
to the special interfacial Green’s functions for both perfect-bonded and smooth
interface cases.
Remark 5.7: The bimaterial Green’s function solutions for the perfect-bonded
interface case are reduced to the Kelvin’s solutions, that is, Eqs. (5.17) and (5.22), in a
homogeneous full-space if we let μ1 = μ2 = μ,ν1 = ν2 = ν. For the smooth interface case,
they are reduced to the solutions for an infinite homogenous space with a smooth
horizontal plane at z = 0.
For the perfect-bonded interface, the parameters in Eqs. (5.99) and (5.103) are,
respectively, reduced to

1
Ds = 0, k1 = 0, k2 = 0, k3 = 0, k4 = k5 = , k6 = 0 (5.117)

1
Ds = 0, k7 = 0 , k8 = 0, k9 = 0, k10 = 0, k11 = 0, k12 = 0, k13 = , k14 = 0 (5.118)

5.6 Green’s Displacements and Stresses in Elastic Isotropic Bimaterial Spaces 165

For the smooth interface, the parameters in Eqs. (5.100) and (5.104) are,
respectively, reduced to

1 1 1
Ds = D = , m1 = , k1 = , k2 = 2(1 − ν), k3 = 2(1 − ν)(1 − 2 ν)
2 2µ 2
    (5.119)
1− ν (1 − 2 ν) (1 − ν)(1 − 2 ν)
k4 = , k5 = − , k6 = −
2µ 4µ 2µ

1 1 1 1 − 2ν
Ds = D = , m2 = − , k7 = , k8 = ,
2 µ 2µ 4µ
    (5.120)
(1 − 2 ν)2
k9 = −k12 , k10 = k7 , k11 = −k8 , k12 = , k13 = 0, k14 = k8

Remark 5.8: For the reduced traction-free half-space case of Material 1 only (σiz =
0 at z = 0; i = x,y,z), we have μ2 = 0; μ1 = μ, ν1 = ν. Therefore, for both the perfect-
bonded and smooth interface cases, the parameters in Eqs. (5.99) and (5.103) and
those in Eqs. (5.100) and (5.104) are reduced to the same expressions as follows:

Ds = 0, D = 0, k1 = 1, k2 = 3 − 4 ν, k3 = 4(1 − ν)(1 − 2 ν) (5.121)

1 1 − 2ν (1 − 2 ν)2 1
Ds = 0, D = 0, k7 = , k8 = , k9 = − , k10 = (5.122)
2µ µ µ µ

Therefore, the solutions of the traction-free half-space for both perfect-bonded


and smooth interface cases are exactly the same. It can be further shown that the
displacement and stress fields in Material 1 for the traction-free half-space case
are reduced to those in Mindlin (1936, 1953), which can be further reduced to the
Boussinesq (1885) or Cerruti (1888) solutions when a vertical or horizontal point
force is applied on the surface of the half-space. For easy reference, the displacement
and stress fields in the traction-free half-space due to the vertical and horizontal
point forces applied on the surface are listed in Appendix B.
Remark 5.9: For the reduced rigid half-space in Material 1 only, we have (μ2 = ∞;
μ1 = μ, ν1 = ν). Thus, for the perfect-bonded interface (reducing to the half-space
with a rigid surface, i.e. ui = 0 at z = 0 (i = x, y, z); Lorentz 1896; Wang 2002), the
parameters in Eqs. (5.99) and (5.103) are reduced, respectively, to

1
Ds = 0, k1 = − , k2 = −1, k3 = 0 (5.123)
(3 − 4 ν)

1 1
Ds = 0, k7 = − , k8 = 0, k9 = 0, k10 = − (5.124)
2µ µ (3 − 4ν)

For the corresponding smooth interface (reducing to the mixed boundary condition
with uz = 0 and σxz = σyz = 0 at z = 0), the parameters in Eqs. (5.100) and (5.104) are,
respectively, reduced to

Ds = D = 1, k1 = 0, k2 = 1, k3 = 0 (5.125)
166 Green’s Functions in Elastic Isotropic Full and Bimaterial Spaces

1
Ds = D = 1, k7 = , k8 = 0, k9 = 0, k10 = 0 (5.126)

It is interesting to point out that for the special case in which the vertical or horizontal
force is applied on the surface of the half-space, the displacement and stress fields are
all zero for the half-space with a rigid surface, i.e. ui = 0 (i = x, y, z) at z = 0; however,
they are not zero for the half-space under the mixed boundary condition (uz = 0 and
σxz = σyz = 0 at z = 0). These nonzero components are listed in Appendix C.
Remark 5.10: When reducing to the special solutions, the following two relations are
found to be very useful (similar relations exist for R1 and R1*)

z+ c 1 1  1 1  1 1
= − * , (z + c)  3 * + 2 *2  = 3 − (5.127)
R2 R2 R2 R2
*
 R2 R2 R2 R2  R2 R2 R2*2

5.7. Brief Discussion on the Corresponding Dislocation Solution


Using the relation between the point-force and point-dislocation solution presented in
Chapter 2, one can directly obtain the point-dislocation-induced displacements (from
the point-force-induced stresses). Taking the derivatives of the point-dislocation-
induced displacement and making use of the constitutive relations will give the
induced stress field. For a finite dislocation surface, we only need to integrate these
expressions over the surface to find the finite dislocation-induced displacement and
stress fields. These statements between the point-force and point-dislocation solutions
can be applied to infinite, half-space, and bimaterial cases, except that one needs to be
cautious on the relation between the source and field points when deriving the point-
dislocation solution from the corresponding point-force solution.

5.8. Applications: Uniform Loading over a Circular Area on the


Surface of a Half-Space
Surface loading over a circular area on the surface of a half-space has important
applications in various engineering branches. In this section as the application, we
first present the integral expressions of the surface displacements and stresses due
to uniform normal (vertical) or shear (horizontal) loading over a circle of radius a
on the surface of an isotropic half-space. Outside the circle, it is traction free. Then,
based on these integral expressions, we present some numerical results.
Under a uniform load of density q in vertical direction over a circle r = a on
the surface, the Boussinesq solutions (Boussinesq 1885; see Appendix B) can be
integrated to find the induced displacements and stresses. For example, the induced
vertical displacement uz and normal stress σxx on the surface can be analytically
expressed as

µuzz 1 − ν 2 π a 1
qa
= ∫ ∫
2 πa 0 0 [r + ρ − 2rρ cos(θ − β)]1/ 2
2 2
ρ dρdβ (5.128)
5.8. Applications: Uniform Loading over a Circular Area 167

z
   σ xx = (1 − 2 ν)
2 π a r 2 + ρ2 − 2rρ cos(θ − β) − 2(r sin θ − ρ sin β)2

2 π ∫0 ∫0
ρdρdβ (5.129)
q [r 2 + ρ2 − 2rρ cos(θ − β)]2

where

r = x 2 + y2 , θ = arctan( y / x) (5.130)

Similarly, under the same uniform vertical load q over the circle of r = a, the induced
stress component σzz in the half-space can be analytically expressed as

σ zzz 3z3 2 π a 1
q
=−
2π ∫0 ∫0 [r + ρ − 2rρ cos(θ − β) + z2 ]5/ 2
2 2
ρdρdβ (5.131)

Under a uniform shear (horizontal) load q in the x-direction over the surface circle
r = a, we can also find the induced displacements and stresses by integrating the
Cerruti solutions (Cerruti 1888). Integral expressions of the induced fields by the
uniform horizontal loading over a surface circle on the surface can be expressed as

µuxx 1 − ν 2 π a 1
qa
=
2 πa ∫0 ∫0 [r + ρ − 2rρ cos(θ − β)]1/ 2
2 2
ρdρdβ
(5.132)
ν 2π a (r cos θ − ρ cos β)2
2 πa ∫0 ∫0 [r 2 + ρ2 − 2rρ cos(θ − β)]3/ 2
+ ρdρdβ

µuyx ν 2 π a (r cos θ − ρ cos β)(r sin θ − ρ sin β)


2πa ∫0 ∫0 [r 2 + ρ2 − 2rρ cos(θ − β)]3/2
= ρdρdβ (5.133)
qa

µuzx 1 − 2 ν 2 π a r cos θ − ρ cos β


qa
=
4 πa ∫0 ∫0 r + ρ2 − 2rρ cos(θ − β)
2
ρdρdβ (5.134)

σ xxx 1 − 2 ν 2 π a (r cos θ − ρ cos β) 3(r cos θ − ρ cos β)(r sin θ − ρ sin β)2
q
=
2π ∫0 ∫0 [r + ρ − 2rρ cos(θ − β)]
2 2 3 / 2

[r 2 + ρ2 − 2rρ cos(θ − β)]5/ 2
ρdρdβ

3 2π a (r cos θ − ρ cos β)3



2π ∫0 ∫0 [r + ρ2 − 2rρ cos(θ − β)]5 / 2
2
ρdρdβ
 (5.135)

σ xyy
3(1 − 2 ν) 2 π a r cos θ − ρ cos β (r cos θ − ρ cos β)3
q 2π
= ∫0 ∫0 [r + ρ − 2rρ cos(θ − β)]
2 2 3 / 2
− 2
[r + ρ2 − 2rρ cos(θ − β)]5/ 2
ρdρdβ

3 2 π a (r cos θ − ρ cos β)(r sin θ − ρ sin β)2


2π ∫0 ∫0
− ρ dρdβ
[r 2 + ρ2 − 2rρ cos(θ − β)]5 / 2
 (5.136)

Remark 5.11: The double integrals in Eqs. (5.128), (5.129), (5.131)–(5.136) can be
reduced to a single integral from 0 to 2π, although the involved integrands are
168 Green’s Functions in Elastic Isotropic Full and Bimaterial Spaces

2 µuzz /qa
0.65
0.60
0.55
0.50
0.45
y/a

0 0.40
0.35
0.30
0.25
0.20
0.15
–2 0.10

–4
–4 –2 0 2 4
x/a

Figure 5.1. Contours of normalized vertical displacement μuz/(qa) on the surface of the half-
space due to a uniform vertical load q over the circle r = a based on Eq. (5.128) with ν = 0.3.

σ zxx /q
0.05
1 –0.00
–0.05
–0.10
–0.15
–0.20
–0.25
–0.30
y/a

0 –0.35
–0.40
–0.45
–0.50
–0.55
–0.60
–1 –0.65
–0.70
–0.75

–2
–2 –1 0 1 2
x/a

Figure 5.2. Contours of the normalized stress component σxx/q on the surface of the half-
space due to a uniform vertical load q over the circle r = a based on Eq. (5.129) with ν = 0.3.

complicated. Furthermore, these double integral expressions can be reduced to a


single integral from 0 to +∞. Actually, for a layered isotropic elastic half-space under
uniform circular loading on its surface, the induced displacement and stress fields
can all be expressed in terms of simple line integration from 0 to +∞ (i.e., Pan et al.
2007). These results have many applications in engineering and physics where the
structures are made of layered elastic materials.
Numerical results of the induced fields by these uniform surface loadings
(vertical or horizontal) over the circle of r = a are presented in Figures 5.1 to 5.8
based on the integral expressions (5.128) to (5.136). We point out that the results
are dimensionless and the Poisson’s ratio is fixed at ν = 0.3. Figure 5.1 shows the
contours of the normalized vertical displacement μuz/(qa) on the surface of the
half-space due to a uniform vertical load q over the circle r = a based on Eq. (5.128).
It can be seen clearly that due to the symmetry of the problem, the contours are
axisymmetric around the center of the surface circle. Figure 5.2 shows the contours
5.8. Applications: Uniform Loading over a Circular Area 169

3
σ zzz /q
–0.05
–0.15
–0.25
–0.35
z/a

2 –0.45
–0.55
–0.65
–0.75
–0.85
–0.95
1

0
–2 –1 0 1 2
x/a

Figure 5.3. Contours of the normalized stress component σzz /q on the vertical plane y = 0 due
to a uniform vertical load q over the circle r = a based on Eq. (5.131).

µu xx /qa
2 0.80
0.75
0.70
0.65
0.60
0.55
y/a

0 0.50
0.45
0.40
0.35
0.30
0.25
0.20
–2 0.15
0.10

–4
–4 –2 0 2 4
x/a

Figure 5.4. Contours of normalized horizontal displacement μux/(qa) on the surface of the
half-space due to a uniform horizontal load q in the x-direction over the circle r = a based on
Eq. (5.132) with ν = 0.3.

of the normalized stress component σxx/q on the surface of the half-space due to the
same uniform vertical load q based on Eq. (5.129). It is observed that this normal
stress component is symmetric about both the x- and y-axes. These two figures show
further that the magnitude of the displacement and stress inside the circle is much
larger than that outside and that their magnitude decays with increasing distance
to the center of the circle. Figure 5.3 plots the contours of the normalized stress
component σzz /q on the vertical plane y = 0 due to the uniform vertical load q over
the circle r = a based on Eq. (5.131) where the surface traction condition is clearly
satisfied. The stress contour is in the bulb shape as is well known in various civil
engineering applications (e.g., Craig 1992).
Figure 5.4 shows the contours of normalized horizontal displacement μux/(qa)
on the surface of the half-space due to a uniform horizontal load q in the x-direction
over the circle r = a based on Eq. (5.132). Compared with Figure 5.1, its contours are
170 Green’s Functions in Elastic Isotropic Full and Bimaterial Spaces

µu xy /qa
0.04
0.03
0.02
y/a

0 0.01
0.00
–0.01
–0.02
–0.03
–0.04
–2

–4
–4 –2 0 2 4
x/a

Figure 5.5. Contours of normalized horizontal displacement μuy/(qa) on the surface of the
half-space due to a uniform horizontal load q in the x-direction over the circle r = a based on
Eq. (5.133) with ν = 0.3.

2
µu xz /qa
0.09
0.07
0.05
0.03
y/a

0 0.01
–0.01
–0.03
–0.05
–0.07
–0.09
–2

–4
–4 –2 0 2 4
x/a

Figure 5.6. Contours of normalized vertical displacement μuz/(qa) on the surface of the half-
space due to a uniform horizontal load q in the x-direction over the circle r = a based on Eq.
(5.134) with ν = 0.3.

no longer in circle but in elliptic because the loading is no longer symmetric. Under
the same horizontal loading, the contours of the normalized horizontal displacement
μuy/(qa) based on Eq. (5.133) are shown in Figure 5.5. It is interesting that these
contours are antisymmetric with respect to the x- and y-axes, but symmetric with
respect to the axis which makes 45 degrees to either the x- or y-axis. It should be
also noticed that the magnitude of uy is much smaller than ux and its maximum
and minimum values are not within the circle of the loading but on the circle of
r = a. Contours of normalized vertical displacement μuz/(qa) on the surface of the
half-space due to the same uniform horizontal load are presented in Figure 5.6 based
on Eq. (5.134). It is noted that its magnitude is about twice of uy. Furthermore, its
5.8. Applications: Uniform Loading over a Circular Area 171

1
σ xxx /q
2.5
2.0
1.5
1.0
y/a

0 0.5
0.0
–0.5
–1.0
–1.5
–2.0
–2.5
–1

–2
–4 –2 0 2 4
x/a

Figure 5.7. Contours of normalized stress component σxx /q on the surface of the half-space
due to a uniform horizontal load q in the x-direction over the circle r = a based on Eq. (5.135)
with ν = 0.3.

1
σ xyy /q
0.7
0.6
0.5
0.4
y/a

0 0.3
0.2
0.1
0.0
–0.1
–0.2
–0.3
–1 –0.4
–0.5
–0.6
–0.7

–2
–2 –1 0 1 2
x/a

Figure 5.8. Contours of normalized stress component σyy/q on the surface of the half-space
due to a uniform horizontal load q in the x-direction over the circle r = a based on Eq. (5.136)
with ν = 0.3.

contour shapes are symmetric with respect to the x-axis and antisymmetric with
respect to the y-axis. Similarly, its maximum and minimum are reached on the circle
of r = a. Figures 5.7 and 5.8 show the contours of normalized stress components σxx /q
and σyy /q on the surface of the half-space due to a uniform horizontal load q in the
x-direction over the circle r = a based on Eqs. (5.135) and (5.136). It is observed
that both stress components have similar distributions with their maximum and
minimum on the circle r = a. Both contours are symmetric with respect to the x-axis
and antisymmetric with respect to the y-axis. The magnitude of σxx is about four
times larger than that of σyy.
172 Green’s Functions in Elastic Isotropic Full and Bimaterial Spaces

5.9 Summary and Mathematical Keys

5.9.1 Summary
We have presented the point-force Green’s functions in elastic isotropic full
and bimaterial spaces. The solutions include all the stresses as well as all the
displacements. Our solutions can be reduced to the full-space Green’s functions,
or the Kelvin’s solutions, the Mindlin solutions for a traction-free half-space, and
the Lorentz solutions for a half-space under the fixed surface conditions. It should
be noticed that when the point force is located on the interface, we then have the
interfacial Green’s functions in the bimaterial space. For the half-space case and
when the point force is located on its surface, we have the Boussinesq solutions and
the Cerruti solutions. The complete Green’s stresses in the bimaterial space and both
the interfacial displacements and stresses are believed to be presented only in this
chapter. As numerical examples, displacement and stress fields due to vertical and
horizontal uniform loadings over a circular domain on the surface of a half-space
are plotted.

5.9.2 Mathematical Keys


In order to solve the full-space Green’s functions, the Galerkin’s vector is employed.
To solve the corresponding bimaterial problem, we follow the approach by
Rongved (1955). The most attractive feature of using Rongved’s approach is that in
deriving the bimaterial Green’s functions using the Papkovich functions, we utilize
the bimaterial potential Green’s functions combined with the integral equation
method.
In solving the displacements and stresses from the Papkovich functions, the first
and second derivatives of various harmonic functions are needed.
In Chapters 6 and 7, a totally different method will be presented to derive the
extended Green’s displacements and stresses in transversely isotropic MEE full, half,
and bimaterial spaces.

5.10 Appendix A: Derivatives of Some Common Functions


We first define the following functions

R = x12 + x22 + x32


(A1)
Rα* = R + sα x3 , s1 = −1, s2 = 1

We then have the following useful partial derivatives (α,β = 1,2; i = 1,2,3)

∂ ln Rα* xβ ∂ ln Rα* sα
= , = (A2)
∂xβ RRα* ∂ x3 R

∂  1  nxi
 n  = − n + 2 (A3)
∂xi R R
5.11 Appendix B: Displacements and Stresses in a Traction-Free Half-Space 173

∂  1  nxβ ∂  1  nsα
 *  =− *(
,  *  =− (A4)
n
∂xβ  Rα  RRα n + 1) ∂x3  Rα 
n RRα* n

∂  1   1 1  ∂  1  s
= − xβ  3 * + 2 * 2  , = − α3 (A5)
∂xβ  RRα*   R Rα R Rα  ∂x3  RRα*  R

∂  1   1 2  ∂  1   1 1 
 * 2  = − xβ  3 * 2 + 2 * 3  ,  * 2  = − sα  3 * + 2 * 2  (A6)
∂xβ  RRα   R Rα R Rα  ∂x3  RRα   R Rα R Rα 

5.11 Appendix B: Displacements and Stresses in a Traction-Free


Half-Space Due to a Point Force Applied on the Surface
For this case, the traction-free boundary conditions on the surface z = 0 is σiz = 0
(i = x, y, z), and the solutions of the displacements and stresses are

fz x  z (1 − 2 ν) 
uxz =  − (B1)
4 πµ  R 3 R(R + z) 

fz  2(1 − ν) z2 
uzz = + 3 (B2)
4 πµ  R R 

fx  1 x 2 1 − 2 ν  x2  
uxx =  + 3+  1−  (B3)
4 πµ  R R R+z  R(R + z)  

fx xy  1 (1 − 2 ν) 
uyx = − (B4)
4πµ  R 3 R(R + z)2 

fx x  z (1 − 2 ν) 
uzx =  + (B5)
4 πµ  R 3 R(R + z) 

fz  (1 − 2 ν)z (1 − 2 ν) 3 x 2 z  1 1 
σzxx =  − − 5 + (1 − 2 ν)x 2  3 + 2 
2π  R 3 R(R + z) R  R (R + z) R (R + z)  
2
(B6)

3 fz z3
σzzz = − (B7)
2 πR 5

3 fz xz2
σzxz = − (B8)
2πR 5

fz xy  3z  1 1 
σzxy =  − 5 + (1 − 2 ν)  3 + 2  (B9)
2π  R  R (R + z) R (R + z)  
2
174 Green’s Functions in Elastic Isotropic Full and Bimaterial Spaces

f x x  (1 − 2 ν) 3x 2 3(1 − 2 ν)  1 2 
σ xxx =  − 5 − + (1 − 2 ν)x 2  3 + 2  (B10)
2π  R 3 R R(R + z) 2  R (R + z) 2 R (R + z)  
3

f x x  (1 − 2 ν) (1 − 2 ν) 3 y2  1 2 
σ xyy =  − − 5 + (1 − 2 ν) y2  3 + 2  (B11)
2π  R 3 R(R + z) 2 R  R (R + z) 2 R (R + z)  
3

3 fx xz2
σzz
x =− (B12)
2 πR 5

3 fx x 2 z
σ xxz = − (B13)
2 πR 5

3 fx xyz
σ xyz = − (B14)
2 πR5

fx y  (1 − 2 ν) 3x2  1 2 
   σ xxy = − − 5 + (1 − 2 ν) x 2  3 + 2  (B15)
2 π  R(R + z) 2 R  R (R + z) 2 R (R + z)  
3

It is noted that Eqs. (B1), (B2), and (B6)–(B9) correspond to the Boussinesq solution
(the expression for u yz can be obztained from u xz by simply switching x and y), while
Eqs. (B3)–(B5) and (B10)–(B15) to the Cerruti solution.

5.12 Appendix C: Displacements and Stresses in a Half Space


Induced by a Point Force Applied on the Surface with
Mixed Boundary Conditions
For this case, the boundary condition on the surface z = 0 is uz = 0 and σxz = σyz = 0,
and the solutions of the displacements and stresses are

uxz = 0, uzz = 0 (C1)

fx  3 − 4ν x2 
uxx = + 3 (C2)
8 πµ(1 − ν)  R R 

fx xy
uyx = (C3)
8 πµ(1 − ν)R 3

f x xz
uzx = (C4)
8 πµ(1 − ν)R 3

σzxx = 0, σzzz = 0, σzxz = 0, σzxy = 0 (C5)

− fx x  1 − 2 ν 3x 2 
σ xxx = + 5
4 π(1 − ν)  R 3 R 
(C6)
5.13 References 175

fx x  1 − 2 ν 3 y 2 
σ xyy = − 5
4 π(1 − ν)  R 3 R  (C7)

fx x  1 − 2 ν 3z2 
σzz
x = − 5
4 π(1 − ν)  R 3 R  (C8)

fx z  1 − 2 ν 3x 2 
σ xxz = − + 5
4 π(1 − ν)  R 3 R  (C9)

3 f x xyz
σ xyz = −
4 π(1 − ν)R 5 (C10)

fx y  1 − 2 ν 3 x 2 
σ xxy = − + 5 (C11)
4 π(1 − ν)  R 3 R 

5.13 References
Boussinesq, J. 1885. Application des potentiels a l’etude de l’equilibre et du mouvement des
solides elastiques. Paris: Gauthier-Villars.
Brebbia, C. A. and Dominguez, J. 2001. Boundary Elements: An Introductory Course. 2nd ed.
Southampton, UK: Computational Mechanics Publications.
Cerruti, V. 1888. Sulla deformazione di un corpo elastico isotropo per alsune speziali
condizioni ai limiti. Rendiconti Reale Accademia dei Lincei Roma 4 (4): 785–92.
Chen, M. C. and Tang, R. J. 1997. An explicit tensor expression for the fundamental solutions
of a bimaterial space problem. Applied Mathematics and Mechanics 18: 331–40.
Craig, R. F. 1992. Soil Mechanics. 5th ed. New York: Chapman and Hall.
Dundurs, J. and Hetenyi, M. 1965. Transmission of force between two semi-infinite solids.
Journal of Applied Mechanics 32: 671–4.
Guzina, B. B. and Pak, R. Y. S. 1999. Static fundamental solutions for a bi-material full-space.
International Journal of Solids and Structures 36: 493–516.
Lorentz, H. A. 1896. Ein allgemeiner Staz, die Bewegung einer reibenden Flüssigkeit
betreffend, nebst einigen Anwendungen desselben. Zittingsverlag Akad. v. Wet.
Amerstadm 5: 168–75.
Love, A. E. H. 1927. A Treatise on the Mathematical Theory of Elasticity. 4th ed. Cambridge:
Cambridge University Press.
Mindlin, R. D. 1936. Force at a point in the interior of a semi-infinite solid. Phys. 7: 195–202.
1953. Force at a point in the interior of a semi-infinite solid. In: Proceedings of the First
Conference on Solid Mechanics, pp. 56–9.
Pan, E., Bevis, M., Han, F., Zhou, H., and Zhu, R. 2007. Surface deformation due to loading
of a layered elastic half-space: A rapid numerical kernel based on a circular loading
element. Geophysical Journal International 88: 90–100.
Rongved, L. 1955. Force interior to one of two joined semi-infinite solids. In: Proceedings of
the Second Conference on Solid Mechanics (ed. by Bogdanov, J. L.), pp. 1–13.
Wang, M. Z. 2002. Advanced Elasticity. Beijing: Beijing University Press (in Chinese).
6 Green’s Functions in a Transversely
Isotropic Magnetoelectroelastic
Full Space

6.0 Introduction
In addition to the hexagonal crystals of class 6mm, most commercial piezoceramics
(e.g., BaTiO3) also exhibit the characteristic of transverse isotropy after poling, with
their axisymmetric axis (the isotropic plane) coincident with (perpendicular to)
the polarization direction. A detailed account of three-dimensional treatments of
transversely isotropic piezoelectric bodies can be found in Ding and Chen (2001). Some
piezomagnetic materials behave similarly, with the isotropic plane perpendicular to
the magnetization direction (e.g., CoFe2O4). In this chapter, we shall devote ourselves
directly to transversely isotropic magnetoelectroelastic (MEE) materials, which may
be composite in nature, obtained, for example, by blending a piezoelectric phase with
a piezomagnetic phase. We consider the response of a full MEE space, subjected to
external point mechanical, electric, or magnetic stimulus (see Figure 6.1). The analysis
starts from establishing the general solution of the governing equations using the
operator theory; a trial-and-error method is then employed to derive the exact
fundamental solutions (i.e., the Green’s functions). As an application, the derived
Green’s functions are also employed to solve the corresponding Eshelby inclusion
problem. Note that the comprehensive analytical results presented in this chapter can
be reduced, by letting appropriate coupling material constants vanish, to those for the
piezoelectric, piezomagnetic, and elastic bodies with transverse isotropy.

6.1 General Solutions in Terms of Potential Functions


In the Cartesian coordinates (x,y,z) the constitutive relations of transversely isotropic
MEE materials can be expressed in terms of the elastic displacements and electric
and magnetic potentials as (from Eq. (2.10))

σ xx = c11ux, x + c12 uy, y + c13uz,z + e31φ,z + q31ψ ,z


σ yy = c12 ux, x + c11uy, y + c13uz,z + e31φ,z + q31ψ ,z
σ zz = c13ux, x + c13uy, y + c33uz,z + e33φ,z + q33ψ ,z
(6.1a)
σ yz = c44 (uy,z + uz, y ) + e15φ, y + q15ψ , y
σ xz = c44 (ux,z + uz, x ) + e15φ, x + q15ψ , x
σ xy = c66 (ux, y + uy, x )

176
6.1 General Solutions in Terms of Potential Functions 177

o x

S (0,0,h)
y

Figure 6.1. An infinite MEE space subjected to concentrated sources at the point S(0,0,h).
The dashed vector represents the mechanical force (fm), applied along an arbitrary direction,
while the solid dot denotes either the electric charge (fe) or the magnetic charge (fh).

Dx = e15 (ux,z + uz, x ) − ε 11φ, x − α 11ψ , x


Dy = e15 (uy,z + uz, y ) − ε 11φ, y − α 11ψ , y (6.1b)
Dz = e31 (ux, x + uy, y ) + e33uz,z − ε 33φ,z − α 33ψ ,z

Bx = q15 (ux,z + uz, x ) − α 11φ, x − µ11ψ , x


By = q15 (uy,z + uz, y ) − α 11φ, y − µ11ψ , y (6.1c)
Bz = q31 (ux, x + uy, y ) + q33uz,z − α 33φ,z − µ33ψ ,z

where c66 = (c11 − c12 ) / 2. Here we have assumed that the isotropic plane of the
material is parallel to the (x,y)-plane of the coordinates.
Substituting the constitutive relations (6.1) into the equilibrium equations with
free body sources (i.e., Eq. (2.1)) leads to

c11ux, xx + c66 ux, yy + c44 ux,zz + (c12 + c66 )uy, yx + (c13 + c44 )uz,zx
+ (e15 + e31 )φ, xz + (q15 + q31 )ψ , xz = 0
c66 uy, xx + c11uy, yy + c44 uy,zz + (c12 + c66 )ux, xy + (c13 + c44 )uz,zy
+ (e15 + e31 )φ, yz + (q15 + q31 )ψ , yz = 0
c44 ∆uz + c33uz,zz + (c13 + c44 )(ux, xz + uy, yz )
(6.2)
+ e15 ∆φ + e33φ,zz + q15 ∆ψ + q33ψ ,zz = 0
e15 ∆uz + e33uz,zz + (e15 + e31 )(ux, xz + uy, yz )
− ε 11∆φ − α 11∆ψ − ε 33φ,zz − α 33ψ ,zz = 0
q15 ∆uz + q33uz,zz + (q15 + q31 )(ux, xz + uy, yz )
− α 11∆φ − µ11∆ψ − α 33φ,zz − µ33ψ ,zz = 0

where ∆ = ∂ 2 / ∂x 2 + ∂ 2 / ∂y2 is the two-dimensional Laplacian. Now following Ding


et al. (1996a), we introduce two potential functions, Φ and Ψ, such that
178 Green’s Functions in a Transversely Isotropic Magnetoelectroelastic Full Space

ux = −Φ, x +Ψ , y , uy = −Φ, y −Ψ , x (6.3)

Then, we obtain from Eq. (6.2)

 ∂2 
 c66 ∆ + c44 2 Ψ = 0 (6.4)
 ∂z 

 Φ  0 
  0 
D uφz  = 0 (6.5)
 ψ  0 
   

where D is a differential operator matrix defined as

 ∂2 ∂ ∂ ∂ 
 c11 ∆ + c44 ∂z2 −(c13 + c44 )
∂z
−(e15 + e31 )
∂z
−(q15 + q31 )
∂z 
 ∂2 ∂2 ∂2
 −(c13 + c44 )∆ ∂ c44 ∆ + c33 2 e15 ∆ + e33 2 q15 ∆ + q33 2 
 ∂z ∂z ∂z ∂z 
D= ∂  ∂2  ∂2 ∂ 2  (6.6)
 (e15 + e31 )∆ −  e15 ∆ + e33 2  ε11 ∆ + ε 33 2 α 11 ∆ + α 33 2 
 ∂z  ∂z  ∂z ∂z 
 ∂  ∂2  ∂2 ∂2 
 (q15 + q31 )∆ ∂z −  q15 ∆ + q33 2 
 ∂z 
α 11 ∆ + α 33 2
∂z
µ 11 ∆ + µ 33 2 
∂z 

Its determinant D can be computed to be

∂8 ∂6 4 2
D = n0 + n ∆ + n ∆ 2 ∂ + n ∆3 ∂ + n ∆4 (6.7)
1 2 3 4
∂z8 ∂z6 ∂ z4 ∂ z2

where
n0 = c44 Π  (6.8a)

n1 = c11Π + [c44
2 − (c + c )2 ]Π
13 44 11 + [c44 ε 11 + (e15 + e31 ) ]Π 22 + [c44 µ11 + (q15 + q31 ) ]Π 33
2 2

+ 2[c44 e15 − (c13 + c44 )(e15 + e31 )]Π 12 + 2[c44 q15 − (c13 + c44 )(q15 + q31 )]Π 13
+ 2[c44α 11 + (e15 + e31 )(q15 + q31 )]Π 23 ,
 (6.8b)
n2 = c11 (c44 Π 11 + e15 Π 12 + q15 Π 13 ) + c44 (c33 Γ 11 + e33 Γ 12 + q33 Γ 13 )
2 − (c + c )2 ]Ω + (e + e )2 Ω + (q + q )2 Ω
+ [c11c33 + c44 13 44 11 15 31 22 15 31 33
+ [c11e33 + c44 e15 − 2(c13 + c44 )(e15 + e31 )]Ω12
+ [c11q33 + c44 q15 − 2(c13 + c44 )(q15 + q31 )]Ω13 + 2(e15 + e31 )(q15 + q31 )Ω 23 , (6.8c)

n3 = c44 Γ + [c11c33 − (c13 + c44 )2 ]Γ 11 + [c11ε 33 + (e15 + e31 )2 ]Γ 22 + [c11µ 33 + (q15 + q31 )2 ]Γ 33
+ 2[c11e33 − (c13 + c44 )(e15 + e31 )]Γ 12 + 2[c11q33 − (c13 + c44 )(q15 + q31 )]Γ 13
+ 2[c11α 33 + (e15 + e31 )(q15 + q31 )]Γ 23 ,
 (6.8d)
 (6.8e)
n4 = c11Γ
6.1 General Solutions in Terms of Potential Functions 179

with Πij and Γ ij being the cofactors of the matrices Π and Γ defined by

 c33 e33 q33   c44 e15 q15 


Π =  −e33 ε 33 α 33  , Γ =  − e15 ε 11 α 11 
 −q33 α 33 µ 33   −q15 α 11 µ 11 

and

Π = det( Π ), Γ = det( Γ )

Ω11 = ε 11 µ33 + ε 33 µ11 − 2α 11α 33 , Ω 22 = c44 µ33 + c33 µ11 + 2q15 q33 ,
Ω 33 = c44ε 33 + c33ε 11 + 2e15 e33 , Ω12 = e33 µ11 + e15 µ33 − α 11q33 − α 33 q15 ,
Ω 13 = q33ε 11 + q15ε 33 − α 11e33 − α 33 e15 , Ω 23 = −(c44α 33 + c33α 11 + e15 q33 + e33 q15 ).

According to the operator theory, the general solution to Eq. (6.5) can be written as

Φ = Ai1F , uz = Ai 2 F , φ = Ai 3 F , ψ = Ai 4 F (i = 1, 2, 3, or 4) (6.9)

where Aij are the cofactors of the matrix D, and F satisfies

DF =0 (6.10)

It is noted that, the index i in Eq. (6.9) can be taken rather arbitrarily without
affecting the final results (Ding et al. 1996a). In the following, we will assume i = 4 in
Eq. (6.9) for the sake of consistence.
Equations (6.4) and (6.10) can be rewritten as

 ∂2 
 ∆ + 2 Ψ = 0 (6.11)
 ∂z0 

 ∂2  ∂2  ∂2  ∂2 
 ∆ + 2   ∆ + 2   ∆ + 2   ∆ + 2  F = 0 (6.12)
 ∂z1  ∂z2  ∂z3  ∂z4 

where zi = si z (j = 0−4), s0 = c66 / c44 and si (i = 1−4) are the four roots (with positive
real part) of the following algebraic eigen-equation:

n0 s 8 − n1 s 6 + n2 s 4 − n3 s 2 + n4 = 0 (6.13)

According to the generalized Almansi’s theorem (Ding et al. 1996a), the function F
can be expressed in terms of four harmonic functions (more precisely, quasiharmonic
functions)

 F1 + F2 + F3 + F4 (for distinct si )
 F1 + F2 + F3 + zF4 (for s1 ≠ s2 ≠ s3 = s4 )
F= 2
(6.14)
F1 + F2 + zF3 + z F4 (for s1 ≠ s2 = s3 = s4 )
 F1 + zF2 + z F3 + z3 F4
2 (for s1 = s2 = s3 = s4 )
180 Green’s Functions in a Transversely Isotropic Magnetoelectroelastic Full Space

where

 ∂2 
 ∆ + ∂z2  Fi = 0 (i = 1, 2, 3, 4) (6.15)
 i 

In the following, we will confine ourselves to the simplest case, that is, the eigenroots
of Eq. (6.13) are distinct from each other. In such a case, substitution of the first
expression in Eq. (6.14) into Eq. (6.9) leads to

∂Ψ 4 ∂ 6 Fi ∂Ψ 4 ∂ 6 Fi
ux = − ∑ (κ i 1 / si ) 5
, u y = − − ∑ (κ i 1 / si )
∂y i = 1 ∂x∂zi ∂x i = 1 ∂y∂zi5
4 4 4 (6.16)
∂ 6 Fi ∂ 6 Fi ∂6 F
uz = ∑ κ i 2 6
, φ = ∑ κ i3 6
, ψ = ∑ κ i 4 6i
i =1 ∂zi i =1 ∂zi i= 1 ∂zi

where

κ ij = a j si6 − bj si4 + d j si2 − g j ( j = 1, 2, 3, 4)

a1 = −(c13 + c44 )Π13 + (e15 + e31 )Π 23 + (q15 + q31 )Π 33

b1 = −(c13 + c44 )Ω13 + (e15 + e31 )Ω 23 + (q15 + q31 )Ω 33

d1 = −(c13 + c44 )Γ 13 + (e15 + e31 )Γ 23 + (q15 + q31 )Γ 33

g1 = 0

a2 = −c44 Π13

b2 = − c11Π 13 − c44 Ω13 − (c13 + c44 )[α 33 (e15 + e31 ) − ε 33 (q15 + q31 )]
− (e15 + e31 )[q33 (e15 + e31 ) − e33 (q15 + q31 )]

d2 = − c11Ω13 − c44 Γ 13 − (c13 + c44 )[α 11 (e15 + e31 ) − ε11 (q15 + q31 )]
− (e15 + e31 )[q15 (e15 + e31 ) − e15 (q15 + q31 )]

g2 = − c11Γ 13

a3 = c44 Π 23

b3 = c11Π 23 + c44 Ω 23 + (c13 + c44 )[α 33 (c13 + c44 ) + e33 (q15 + q31 )]
+ (e15 + e31 )[q33 (c13 + c44 ) − c33 (q15 + q31 )]

d3 = c11Ω 23 + c44 Γ 23 + (c13 + c44 )[α 11 (c13 + c44 ) + e15 (q15 + q31 )]
+ (e15 + e31 )[q15 (c13 + c44 ) − c44 (q15 + q31 )]

g3 = c11Γ 23
6.1 General Solutions in Terms of Potential Functions 181

a4 = c44 Π 33

b4 = c11Π 33 + c44 Ω 33 − (c13 + c44 )[ε 33 (c13 + c44 ) + e33 (e15 + e31 )]
− (e15 + e31 )[e33 (c13 + c44 ) − c33 (e15 + e31 )]

d4 = c11Ω 33 + c44 Γ 33 − (c13 + c44 )[ε11 (c13 + c44 ) + e15 (e15 + e31 )]
− (e15 + e31 )[e15 (c13 + c44 ) − c44 (e15 + e31 )]

g4 = c11Γ 33

Equation (6.16) can be further simplified by letting

∂ 5 Fi
(κ i 1 / si ) = Ψi (i = 1, 2, 3, 4) (6.17)
∂zi5

Utilizing the preceding equation, and writing Ψ as Ψ 0, yields

∂ Ψ 0 4 ∂Ψ i ∂Ψ 0 4 ∂Ψ i
ux = −∑ , uy = − −∑
∂y i = 1 ∂x ∂x i = 1 ∂y
4 (6.18)
∂Ψ i
wk = ∑ βik (k = 1, 2, 3)
i=1 ∂zi

where use has been made of the compact notations w1 = uz, w2 = φ , and w3 = ψ , and

βi 1 = κ i 2 si / κ i 1 , βi 2 = κ i 3 si / κ i 1 , βi 3 = κ i 4 si / κ i 1 (i = 1 − 4)

The functions Ψ i are harmonic, that is,

 ∂2 
 ∆ + ∂z2  Ψ i = 0 (i = 0, 1, ..., 4) (6.19)
 i 

In cylindrical coordinates (r,θ , z), the axial component of the mechanical displacement
uz , the electric potential φ , as well as the magnetic potential ψ still take the same
form as given in Eq. (6.18), while the radial and circumferential components of the
mechanical displacement ur and uθ will be

1 ∂Ψ 0 4 ∂Ψ i ∂Ψ 0 4 1 ∂Ψ i
ur = −∑ , uθ = − −∑ (6.20)
r ∂θ i =1 ∂r ∂r i =1 r ∂θ

For convenience, we introduce the following complex quantities

U = ux + i uy = ei θ (ur + i uθ ), σ1 = σ xx + σ yy = σ rr + σθθ (6.21a)

σ 2 = σ xx − σ yy + 2 i σ xy = e2 i θ (σrr − σθθ + 2 i σrθ ) (6.21b)


182 Green’s Functions in a Transversely Isotropic Magnetoelectroelastic Full Space

τ z1 = σ xz + i σ yz = ei θ (σ rz + i σθz ), τ z2 = Dx + i Dy = ei θ (Dr + i Dθ ) (6.21c)

τ z3 = Bx + i By = ei θ ( Br + i Bθ ), σ z1 = σ zz , σ z2 = Dz, σ z3 = Bz (6.21d)

Then, in view of Eqs. (6.1) and (6.18), we obtain

 4   4 
U = − Λ  ∑ Ψi + i Ψ0  , σ 2 = −2c66 Λ 2  ∑ Ψi + i Ψ0  (6.22a)
 i =1   i =1 

4
∂ 2 Ψi
σzk = ∑ γ ik (k = 1, 2, 3, 4) (6.22b)
i=1 ∂zi2

 4 ∂Ψ i ∂Ψ 0 
τ zk = Λ  ∑ ϖik − i s0 νk (k = 1, 2, 3) (6.22c)
 i =1 ∂zi ∂z0 

where σ z4 = σ1, Λ = ∂ / ∂x + i ∂ / ∂y = ei θ [∂ / ∂r + i(1 / r )∂ / ∂θ] , and

γ i 1 = c13 + c33 si βi 1 + e33 si βi 2 + q33 si βi 3 (6.23a)

γ i 2 = e31 + e33 si βi 1 − ε 33 si βi 2 − α 33 si βi 3 (6.23b)

γ i 3 = q31 + q33 si βi 1 − α 33 si βi 2 − µ 33 si βi 3 (6.23c)

γ i 4 = 2[(c11 − c66 ) + c13 si βi 1 + e31 si βi 2 + q31 si βi 3 ] (6.23d)

ϖi 1 = −c44 si + c44β i 1 + e15βi 2 + q15βi 3 (6.23e)

ϖi 2 = −e15 si + e15β i1 − ε11β i 2 − α 11βi 3 (6.23f)

ϖi 3 = −q15 si + q15β i1 − α 11β i 2 − µ 11β i 3 (6.23g)

ν1 = c44 , ν 2 = e15, ν 3 = q15 . (6.23h)

We also note the following identities

γ ij si = ϖ (i = 1, 2, 3, 4, j = 1, 2, 3) (6.24)

which can be easily verified by direct substitution. It is noted here that verification of
Eq. (6.24) can be used to check the correctness of the lengthy expressions presented
in the preceding text.
As mentioned earlier, the general solution in Eq. (6.18) is valid only for the
particular case in which si are distinct from each other, and was presented in Hou
6.2 Solutions of a Vertical Point Force 183

et al. (2003). When Eq. (6.13) has equal roots, the form of general solution changes
accordingly (Hou et al. 2005). It is noted that a general solution in a slightly different
form was derived a little earlier by Wang and Shen (2002). Chen et al. (2004) derived
a concise general solution for magneto-electro-thermoelasticity when the thermal
effect is also involved.
For convenience of later derivation, we first define Ri and Ri* , and list some
important relations on the derivatives associated with them

Ri = ( x − X )2 + ( y − Y )2 + si2 (z − Z )2 , Ri* = Ri + si z − Z (6.25)

∂ ln Ri* x − X ∂ ln Ri* s
= , = sgn(z − Z ) i (6.26)
∂x Ri Ri* ∂z Ri

∂  1  n( x − X ) ∂  1  nsi2 (z − Z )
=− , = −
∂x  Rin  Rin+ 2  
∂z  Rin  Rin+ 2
(6.27)
∂  1  n( x − X ) ∂  1  nsi
=− , = − sgn(z − Z )
∂x  Ri*n  Ri Ri*( n+ 1) ∂z  Ri*n  Ri Ri*n

∂  1   n m 
  = −( x − X )  n + 2 *m + 
∂x  Rin Ri*m  Ri Ri
*(
Rin +1 Ri m + 1)

(6.28)
∂  1   n n−m 
 n *m  = − sgn(z − Z )si  + 2 *( m−1) − n +1 *m 
∂z  Ri Ri   Rin Ri Ri Ri 

where sgn(·) is the sign function. Derivatives with respect to y can be obtained from
those with respect to x by simply replacing x − X with y − Y.

6.2 Solutions of a Vertical Point Force, a Negative Electric Charge,


or Negative Magnetic Charge
Without loss of generality, we assume that at the point (x,y,z) = (0,0,h), there is an
axisymmetric point source (i.e., either a vertical point force fz(x,y,z) = fzδ(x)δ(y)δ(z−h),
a negative electric charge −fe(x,y,z) = feδ(x)δ(y)δ(z-h), or a negative magnetic
charge −fh(x,y,z) = fhδ(x)δ(y)δ(z−h)). Because these sources are either along the axis
of symmetry (vertical point force) or scalar (the electric or magnetic charge), we can
assume the following potential functions (Ding and Jiang 2004b)

Ψ0 = 0, Ψi = Ai sgn(z − h) ln Ri* (i = 1, 2, 3, 4) (6.29)

where Ai (i = 1–4) are four unknown coefficients to be determined later, and

Ri = x 2 + y2 + si2 (z − h)2 , Ri* = Ri + si z − h (i = 0, 1,..., 4) (6.30)

which corresponds to X = Y = 0 and Z = h in Eq. (6.25).


184 Green’s Functions in a Transversely Isotropic Magnetoelectroelastic Full Space

We point out that the unknown functions have to possess the following features:
(1) They have to satisfy the scaled 3D Laplace’s equations or the quasiharmonic
equations, that is, Eq. (6.19) in the full space everywhere, except at the source
point where the functions are singular;
(2) The symmetric or antisymmetric features of the functions on both sides of the
source level z = h. This can be confirmed directly physically or by looking at
the existing solutions corresponding to the purely elastic cases (isotropic and/or
transversely isotropic).
In earlier works, the sign function and absolute value of z−h appearing in Eq. (6.29)
are not contained in Green’s functions for infinite transversely isotropic elastic or
piezoelectric media (Elliot 1948; Hu 1956; Pan and Chou 1976; Fabrikant 1989; Dunn and
Wienecke 1996). Hanson (1998) pointed out that, when the point source is taken into
consideration, the potential functions, like Ψ i in Eq. (6.22) need not be harmonic, that
is, satisfy Eq. (6.19), but rather satisfy a Poisson equation with a properly determined
inhomogeneous term, which may cause certain mathematical difficulty. He showed
that, for a point force tangential to the isotropic plane, the difficulty could be readily
overcome by introducing the absolute value operation as in Eq. (6.29), while for a
normal (or axisymmetric) point force, there is no need to do so (Hanson 1998, 1999).
Fabrikant (2003) argued that, due to the singular behavior of the problem, Hanson’s
derivation is not sufficiently rigorous, and that the Green’s functions by him (Fabrikant
1989) are essentially the same as those obtained by Hanson (1998, 1999).
In Eq. (6.29), we have adopted the sign function and the absolute value of z−h
for the axisymmetric sources to avoid the singularity of the function ln[Ri + si (z − h)]
at the location different from the source point (Ding et al. 1996b). It is clear that the
argument of this logarithm function ln[Ri + si (z − h)] adopted by most researchers
becomes zero on the z-axis when z < h. Furthermore, to make sure that the potential
functions are harmonic, we divide the whole space into two half-spaces that are
bonded perfectly at the source level z = h , so that the point sources are just located on
the surface of either half-space. With such a treatment, the mathematical complexity
due to the presence of a point source in the interior of a body, as noticed by Hanson
(1998), can be avoided completely, and the potential functions will satisfy the scaled
Laplace equation, that is, Eq. (6.19) within each half-space.
By taking the derivatives of the potential functions, we have

Ai x
Ψ i, x = sgn(z − h)
Ri Ri*
Ay
Ψ i, y = sgn(z − h) i * (6.31)
Ri Ri
As
Ψ i ,z = i i
Ri

Ai si x Ai si y
Ψi, xz = − , Ψi, yz = −
Ri3 Ri3
(6.32)
 1 1 
Ψi, xy = − sgn(z − h) Ai xy  3 * + 2 *2 
 Ri Ri Ri Ri 
6.2 Solutions of a Vertical Point Force 185

 A  1 1 
Ψ i, xx = sgn(z − h)  i * − Ai x 2  3 * + 2 *2  
 Ri Ri  Ri Ri Ri Ri  
 A  1 1 
Ψ i, yy = sgn(z − h)  i * − Ai y2  3 * + 2 *2   (6.33)
 Ri Ri  Ri Ri Ri Ri  
Ai si3 (z − h)
Ψ i,zz = −
Ri3

The preceding derivatives are not necessarily valid for z = h , which is the surface of
both half-spaces; those on z = h may be seen as a limit by taking z → h from either
side of the surface. By contrast, if we consider the infinite medium as a whole, then
the Dirac delta function will emerge inevitably in the derivatives of Ψ i with respect
to z due to the presence of the sign function (Hanson 1999). In this sense, the sign
function introduced in the preceding equations works only for z − h ≠ 0. From Eq.
(6.33), we can easily verify that Ψ i (i = 0–4) satisfy Eq. (6.19) within each half-space.
Then the extended displacements and stresses are

4
( x + i y)
U = − sgn(z − h)∑ Ai (6.34)
i =1 Ri Ri*

4
β ik
wk = ∑ Ai (k = 1, 2, 3) (6.35)
i =1 Ri

4  1 1 
σ 2 = 2c66 sgn(z − h)∑ Ai ( x + i y)2  3 * + 2 *2  (6.36)
i =1  Ri Ri Ri Ri 

4
γ ik si (z − h)
σ zk = −∑ Ai (k = 1, 2, 3, 4) (6.37)
i =1 Ri3

4
ϖ ik ( x + i y)
τ zk = −∑ Ai (k = 1, 2, 3) (6.38)
i =1 Ri3

As we can see, the use of sign function and the absolute value operation can indicate
explicitly the necessary symmetry or antisymmetry properties (with respect to the
source level) of the physical variables. For transversely isotropic elastic materials,
Fabrikant (2003) showed that the conventional form of Green’s functions, which
employs the function ln[Ri + si (z − h)], also presents such properties as they should
be, but through a rather cumbersome mathematical manipulation. Furthermore,
the elastic, electric, and magnetic fields all exhibit a singular behavior at the source
point (0,0,h), as expected. The four coefficients Ai (i = 1–4) are determined from the
following continuity conditions at the interface (z=h) between the two half-spaces
(or at the source level).
(1) The extended displacements uI (I=1–5) are continuous:
186 Green’s Functions in a Transversely Isotropic Magnetoelectroelastic Full Space

While for I = 3–5, these are automatically satisfied because they are even
functions of z; for I = 1 and 2 (they are both odd functions of z – h and should vanish
identically at z = h ), we have
4
∑ Ai = 0 (6.39)
i =1

In view of this, the use of sign function seems very natural and also important because
Eqs. (6.34)–(6.38) become valid even at z = h .
(2) For the extended traction, the shear components σxz and σyz should be continuous
across the source level. These again are automatically satisfied because these
two stress components are even functions of z – h, as clearly indicated by Eq.
(6.38).
(3) The normal components of the extended traction are also continuous except at
the source point. The integration of the jumps in the normal components (σzz, Dz,
Bz) across the source level has to balance the magnitudes of the corresponding
applied point sources (vertical point force f1 = fz, negative electric charge
f2 = − fe, or negative magnetic charge f3 = − fh ). Considering the equilibrium
of an MEE layer bounded by planes z = h ± ε (Hou et al. 2005), we have the
following three integral expressions
∞ ∞
∫−∞ ∫−∞ [σzk ( x, y, h + ε) − σzk ( x, y, h − ε)]d x d y + fk = 0 (k = 1, 2, 3) (6.40)

where ε can be arbitrary. The preceding equations can also be formally obtained by
integrating the third force equilibrium, and the electric charge and magnetic charge
balance equations on both sides of the interface. The integrals involved in Eq. (6.40)
can be carried out easily:

∞ ∞
∫−∞ ∫−∞ [σzk ( x, y, h + ε) − σzk ( x, y, h − ε)]d x d y
4
∞ 1
= − ∑ Ai γ ik si × 2ε × 2π ∫ rdr (6.41)
i =1
0 (r 2 + si2 ε 2 )3/ 2
4
= −4π ∑ Ai γ ik
i =1

where r 2 = x 2 + y2. Therefore, we have the following three extra equations for
determining the coefficients Ai:

4
4π ∑ Ai γ ik = fk (k = 1, 2, 3) (6.42)
i =1

Equation (6.39) and the three equations in Eq. (6.42) form four linear algebraic
equations for the four unknown coefficients Ai, which can be easily solved to give

−1 −1
 A1   1 1 1 1  0  1 1 1 1   0 
 A2  1 γ 11 γ 21 γ 31 γ 41   f1  1 γ 11 γ 21 γ 31 γ 41   fz 
A  = γ γ γ γ  f  = γ γ 22 γ 32 γ 42   − f  (6.43)
 A3  4π γ 12 γ 22 γ 32 γ 42   f2  4π γ 12 γ 23 γ 33 γ 43  − f e 
 4  13 23 33 43   3  13  h
6.3 Solutions of a Horizontal Point Force along x-Axis 187

For the Green’s function problem, we can just assume (fz, −fe, −fh) respectively equal
(1,0,0), (0,1,0), and (0,0,1), and therefore, the corresponding Green’s functions due
to the vertical point force, point electric charge, and point magnetic charge can
be finally found. The extended Green’s displacements are summarized in Eq. (A2)
in Appendix A in terms of the relative coordinates between the source and field
points.
Now we can make a further comment on the form of the potential functions Ψ i .
For a normal (axisymmetric) point source, we can add an arbitrary plane harmonic
function of r to the potential functions in Eq. (6.29), that is, we may take

Ψ0 = 0, Ψi = Ai sgn(z − h) ln Ri* + h(r ) (i = 1, 2, 3, 4) (6.44)

where h(r ) satisfies the planar Laplace equation ∆h(r ) = 0. It can be seen that, in
consideration of Eq. (6.39), the extended displacements calculated from Eq. (6.44)
are the same as those in Eqs. (6.34) and (6.35), and hence the extended tractions given
in Eqs. (6.36)–(6.38) remain the same as well. Such an uncertainty in the potential
functions has been noted by Hanson (1998, 1999). Regardless of the singular behavior
on the axisymmetric axis r = 0, we thus can take either Ψi = Ai ln [Ri + si (z − h)] or
Ψi = Ai ln[Ri − si (z − h)] (Fabrikant 2003). In Hanson’s analysis (Hanson 1998,
1999), he adopted the form of ln {[Ri + si (z − h)] / r} , which also eliminates the
singular behavior on the negative z-axis. However, each term in the expressions for
some displacement and stress components still contains such singularities because
Ri + si (z − h) appears in the denominator. Although they cancel out as a whole in
view of Eq. (6.39), they may cause significant numerical difficulty in the calculation.

6.3 Solutions of a Horizontal Point Force along x-Axis


We assume that there is a point force in the positive x-direction fx, applied at (x,y,z) =
(0,0,h), that is, fx(x,y,z) = fxδ(x)δ(y)δ(z−h). For this case, the four potential functions
can be assumed as

B0 y Bx
Ψ0 = *
, Ψ i = i* ( i = 1, 2, 3, 4) (6.45)
R0 Rj

with Bi (i = 0–4) being the five coefficients to be determined.


The derivatives of these functions are found as

B0 xy B B y2
Ψ 0, x = − * 2
, Ψ 0, y = 0* − 0 *2
R0 R0 R0 R0 R0
B0 s0 y
Ψ 0,z = − sgn(z − h)
R0 R0*
(6.46)
B B x2 B xy
Ψ i, x = *i − i *2 , Ψ i, y = − i *2
Ri Ri Ri Ri Ri
Bsx
Ψ i,z = − sgn(z − h) i i *
Ri Ri
188 Green’s Functions in a Transversely Isotropic Magnetoelectroelastic Full Space

B0 x  1 2 
Ψ 0, xy = − + B0 xy2  3 *2 + 2 *3 

*
R0 R0 2
 R0 R0 R0 R0 
 1 1 
Ψ 0, xz = sgn(z − h)B0 s0 xy  3 * + 2 *2 
 R0 R0 R0 R0 
B s  1 1 
Ψ 0, yz = − sgn(z − h) 0 0* + sgn(z − h)B0 s0 y2  3 * + 2 *2 
R0 R0 R R
 0 0 R0 R0  (6.47)
By  1 2 
Ψ i, xy = − i *2 + Bi x 2 y  3 *2 + 2 *3 
Ri Ri  Ri Ri Ri Ri 
Bi si  1 1 
Ψ i, xz = − sgn(z − h) *
+ sgn(z − h)Bi si x 2  3 * + 2 *2 
Ri Ri R R
 i i Ri Ri 
 1 1 
Ψ i, yz = sgn(z − h)Bi si xy  3 * + 2 *2 
R R
 i i Ri Ri 

B0 y  1 2 
Ψ 0, xx = − + B0 x 2 y  3 *2 + 2 *3 

*
R0 R0 2
 R0 R0 R0 R0 
3B0 y  1 2 
Ψ 0, yy = − + B0 y3  3 *2 + 2 *3 

*
R0 R0 2
 R0 R0 R0 R0 
3Bi x  1 2 
Ψ i, xx = − * 2
+ Bi x 3  3 *2 + 2 *3  (6.48)
Ri Ri  Ri Ri Ri Ri 
Bi x  1 2 
Ψ i, yy = − * 2
+ Bi xy2  3 *2 + 2 *3 
Ri Ri  Ri Ri Ri Ri 
Bi si2 x
Ψ i,zz =
Ri3

Just as for the normal (axisymmetric) point source, these derivatives are not
necessarily all valid at z = h .
Then, the corresponding field quantities are

 1 i y( x + i y)  4  1 x( x + i y) 
* 2  ∑ i  R*
U = B0  * + − B − (6.49)
R
 0 R R
0 0  i =1  i Ri Ri*2 

4
βik x
wk = − sgn(z − h)∑ Bi (k = 1, 2, 3) (6.50)
i =1 Ri Ri*

 2( x + i y)  1 2 
σ 2 = −2c66 B0  *2
+ i y( x + i y) 2 +
 R 3 R0*2 R 2 R0*3  

 R0 R0  0 0  
(6.51)
4  2( x + i y)  1 2 
+ 2c66 ∑ Bi  * 2
− x( x + i y)2  3 *2 + 2 *3  
i =1  Ri Ri  Ri Ri Ri Ri  
6.3 Solutions of a Horizontal Point Force along x-Axis 189

4
γ ik x
σ zk = ∑ Bi (k = 1, 2, 3, 4) (6.52)
i =1 Ri3

 1  1 1 
τ zk = − sgn(z − h)s0ν k B0  *
+ i y( x + i y)  3 * + 2 *2  
 
 R0 R0  R0 R0 R0 R0  
(6.53)
4  1  1 1 
− sgn(z − h)∑ϖ ik Bi  *
− x( x + i y)  3 * + 2 *2   (k = 1, 2, 3)
i =1  Ri Ri  Ri Ri Ri Ri  

Similarly, the five coefficients Bi (i = 0–4) can be determined by the following


continuity conditions across the source level z = h.
(1) The extended displacements uI (I = 1–5) should be continuous:
While for I = 1–2, these are automatically satisfied because they are even
functions of z; for I = 3–5 (they are odd functions of z – h and should vanish
identically at z = h ), we have

4
∑ βik Bi = 0 (k = 1, 2, 3) (6.54)
i =1

(2) For the extended traction, we should have the normal components (σzz, Dz and
Bz) continuous across the source level at z = h . These again are automatically
satisfied because these components are even functions of z – h.
(3) The shear components σxz and σyz should be continuous across the source level
except at the source point, which gives

4
B0 s0ν 1 − ∑ϖ i 1 Bi = 0 (6.55)
i =1

In view of Eq. (6.54),ν 1 = c44, and the expression for ϖ i1 in Eq. (6.23), we can rewrite
the preceding equation in a very simple form:

4
∑ si Bi = 0 (6.56)
i= 0

In addition, considering the force equilibrium in the x-direction of the MEE layer
bounded by planes z = h ± ε (Hou et al. 2005), we have

∞ ∞
∫−∞ ∫−∞ [σ xz ( x, y, h + ε ) − σ xz ( x, y, h − ε )]d x d y + fx = 0 (6.57)

The expression for σxz can be easily extracted from Eq. (6.53) (for k = 1) as the
real part:
190 Green’s Functions in a Transversely Isotropic Magnetoelectroelastic Full Space

 1  1 1 
σ xz = Re[τ z1 ] = − sgn(z − h)s0ν 1 B0  *
− y 2  3 * + 2 *2  
 R R0 R R0 
 R0 R0  0 0  
(6.58)
4  1  1 1 
− sgn(z − h)∑ϖ i 1 Bi  *
− x 2  3 * + 2 *2  
i =1  Ri Ri  Ri Ri Ri Ri  

Because the 0-term (the first term) and the i-terms (the second term) share the similar
function structure, we can look at the 0-term in detail. In other words, we have

[σ xz ( x, y, h + ε ) − σ xz ( x, y, h − ε )]0
 1  1 1   (6.59)
= −2 s0 c44 B0  − y2  3 + 2 
 R0 (R0 + s0ε )  R0 (R0 + s0ε ) R0 (R0 + s0ε )2  

The involved integrals can be carried out in terms of polar coordinates, with the result
being (the coefficient −2 s0 c44 B0 is omitted in the following text for simplicity)

∞ 2π 1  1 1  
∫0 ∫0  − y2  3 + 2   r d r dθ
 R0 (R0 + s0ε )  R0 (R0 + s0ε ) R0 (R0 + s0ε )   2

∞ 2π   1  1 1  
=∫ ∫  − r 2 sin 2 θ  3 + 2   r d r dθ
0 0
 R0 (R0 + s0 ε )  R0 (R0 + s0 ε ) R0 (R0 + s0 ε )2  
∞ 2  1 1  
=π∫  − r2  3 + 2  r d r
0 R (R + s ε )
 0 0 0  R0 (R0 + s0ε ) R0 (R0 + s0ε )  2

∞  2  2 R0 s0ε  
=π∫  − r2  3 +  r d r
0 R (R + s ε )  R0 (R0 + s0ε )2 R03 (R0 + s0ε )2  
 0 0 0
 (6.60)

The last expression can be further changed to


∞ 2  2 R0 s0ε  
π∫  − r2  3 + 3 r d r
0 R (R + s ε )  R0 (R0 + s0ε ) 2 R0 (R0 + s0ε )  
2
 0 0 0
∞  2 R 2 (R0 + s0 ε ) − 2 R0 r 2 − r 2 s0 ε 
=π∫  0 rdr
0 
 R03 (R0 + s0ε )2  (6.61)
∞  2 R0 (R0 + s0ε ) − r 2  ∞  2 R (R + s ε ) − r 2 
=π∫   r d r = π s0ε ∫  0 3 0 0 2  r d r
0  R (R + s ε )
3 2
 0  R (R + s ε ) 
 0 0 0  0 0 0
∞ 1
= π s0ε ∫ r d r = −π s0ε / R0 |rr =∞ =0 = π
0 R3
0

Using the similar result for the second term (i-terms), we therefore arrive at the
following equation

4
2π c44 B0 s0 + 2π ∑ϖ i 1 Bi − fx = 0 (6.62)
i =1
6.4 Various Decoupled Solutions 191

Again, this equation can be simplified as

4
2π c44 s0 B0 − 2π c44 ∑ si Bi = f x (6.63)
i =1

Equations (6.54), (6.56), and (6.63) form a total of five linear algebraic equations
that can be used to determine the five unknown coefficients Bi (i = 0–4) as

fx
B0 = (6.64)
4π c44 s0

−1
 B1   s1 s2 s3 s4  1
 B2  fx  β11 β 21 β 31 β 41  0 
B  = −  β12 β 22 β 32 β 42  0  (6.65)
3
B  4π c44 0 
 4  β13 β 23 β 33 β 43   

For the Green’s function solutions, we just let fx=1. In order to obtain the Green’s
function solutions due to a horizontal point load along the y-direction, one can
simply replace x with y and y with −x respectively on both sides of the expressions in
Eq. (6.45) and then follow the same procedure as outlined in the preceding text. The
consequence of such actions is that the final expressions of the extended Green’s
displacements and stresses due to the y-direction force can be obtained from
those due to the x-direction force by simply switching x and y on both sides of the
expressions of the Green’s displacements and stresses. These statements apply to
the decoupled cases discussed in the following text. For easy future reference, the
extended Green’s displacements are summarized in Eqs. (A3) and (A4) in Appendix
A in terms of the relative coordinates between the source and field points.

6.4 Various Decoupled Solutions


We present two decoupled Green’s function solutions here: piezoelectric Green’s
functions and purely elastic Green’s functions. The corresponding piezomagnetic
Green’s functions can be obtained from the piezoelectric ones by simply replacing
the piezoelectric/electric coefficients and electric point source by the corresponding
piezomagnetic/magnetic ones.
It is noted here that, the two-dimensional Green’s functions for an MEE plane
can be found in Ding and Jiang (2004a), as also presented in Chapter 4. We also
point out that the three-dimensional Green’s functions due to point heat flux were
obtained by Hou et al. (2009).

6.4.1 Piezoelectric Green’s Functions


The Green’s function solutions in the corresponding piezoelectric full space can
be directly obtained from those in the fully coupled MEE space through a proper
degenerate analysis. For example, the eigen-equation (6.13) is reduced to

n1 s6 − n2 s 4 + n3 s 2 − n4 = 0 (6.66)
192 Green’s Functions in a Transversely Isotropic Magnetoelectroelastic Full Space

which determines three characteristic roots si (i = 1, 2, 3) with positive real part for
the piezoelectric material. In the preceding equation, we have

n1 = c44 (c33 ε 33 + e33


2 ), n4 = c11 (c44 ε 11 + e15
2 ) (6.67a)

n2 = c44 (c33ε 11 + e33 e15 ) + [c11c33 + c44


2 − (c + c )2 ]ε + (e + e )2 c
13 44 33 15 31 33
(6.67b)
+ [c11e33 + c44 e15 − 2(c13 + c44 )(e15 + e31 )]e33

n3 = c44 (c44ε 11 + e15


2 ) + [c c − (c + c )2 ]ε + [c ε + (e + e )2 ]c
11 33 13 44 11 11 33 15 31 44
(6.67c)
+ 2[c11e33 − (c13 + c44 )(e15 + e31 )]e15

These can be obtained by simply setting µ 33 = 0 and µ11 = 1 along with qij = α ij = 0
in Eq. (6.8).
The extended displacements are given by

∂ Ψ 0 3 ∂Ψ i ∂Ψ 0 3 ∂Ψ i
ux = −∑ , uy = − −∑
∂y i = 1 ∂x ∂x i = 1 ∂y
3 (6.68)
∂Ψ i
wk = ∑ βik (k = 1, 2)
i =1 ∂zi

where Ψ i (i = 0, 1, 2, 3) satisfy Eq. (6.19), and

βi 1 = κ i 2 si / κ i 1 , βi 2 = κ i 3 si / κ i 1 (i = 1, 2, 3) (6.69)

κ ij = a j si6 − bj si4 + d j si2 − g j (i, j = 1, 2, 3) (6.70)

which is consistent with the fully coupled MEE case. In Eq. (6.70),

a1 = 0, g1 = 0 (6.71a)

b1 = −ε 33 (c13 + c44 ) − e33 (e15 + e31 ) (6.71b)

d1 = −ε 11 (c13 + c44 ) − e15 (e15 + e31 ) (6.71c)

a2 = 0, b2 = − c44 ε 33 (6.71d)

d2 = −c11ε 33 − c44ε 11 − (e15 + e31 )2 , g2 = −c11ε 11 (6.71e)

a3 = 0, b3 = −c44 e33 (6.71f)

d3 = −c11e33 + c44 e31 + c13 (e15 + e31 ), g3 = −c11e15 (6.71g)


6.4 Various Decoupled Solutions 193

The extended tractions then can be obtained from Eq. (6.68) by differentiation.
Using the complex notation introduced in the last section, we can write all physical
variables as

 3  3
∂Ψ i
U = − Λ  ∑ Ψi + i Ψ0  , wk = ∑ βik (k = 1, 2) (6.72)
 i =1  i =1 ∂zi

 3 
σ 2 = −2c66 Λ 2  ∑ Ψi + i Ψ0  (6.73a)
 i =1 

3
∂ 2 Ψi
σzk = ∑ γ ik (k = 1, 2, 4) (6.73b)
i=1 ∂zi2

 3 ∂Ψ i ∂Ψ 0 
τ zk = Λ  ∑ ϖik − i s0 νk (k = 1, 2) (6.73c)
 i =1 ∂zi ∂z0 

where Λ is the complex gradient operator defined just before Eq. (6.23), and

γ i 1 = c13 + c33 si βi 1 + e33 si βi 2 (6.74a)

γ i 2 = e31 + e33 si βi1 − ε 33 si βi 2 (6.74b)

γ i 4 = 2[(c11 − c66 ) + c13 si βi1 + e31 si βi 2 ] (6.74c)

ϖi 1 = − c44 si + c44βi 1 + e15βi 2 (6.74d)

ϖi 2 = −e15 si + e15βi 1 − ε11βi 2 (6.74e)

ν1 = c44 , ν2 = e15 (6.74f)

These expressions can be obtained by simply setting qij = α ij = 0 in Eq. (6.23). The
relation (6.24) still holds for i = 1, 2, 3 and j = 1, 2.

6.4.1.1 Solutions of a Vertical Point Force and a Negative Electric Charge


The potential functions are assumed in the form of Eq. (6.29). The corresponding
extended displacements and stresses are derived as

3
( x + i y)
U = − sgn(z − h)∑ Ai (6.75)
i =1 Ri Ri*

3
β ik
wk = ∑ Ai (k = 1, 2) (6.76)
i =1 Ri
194 Green’s Functions in a Transversely Isotropic Magnetoelectroelastic Full Space

3  1 1 
σ 2 = 2c66 sgn(z − h)∑ Ai ( x + i y)2  3 * + 2 *2  (6.77)
i =1  Ri Ri Ri Ri 

3
γ ik si (z − h)
σ zk = −∑ Ai (k = 1, 2, 4) (6.78)
i =1 Ri3

3
ϖ ik ( x + i y)
τ zk = −∑ Ai (k = 1, 2) (6.79)
i =1 Ri3

The three coefficients Ai (i=1–3) are determined by

−1 −1
 A1  1  1 1 1   0  1  1 1 1   0 
 A2  = γ 11 γ 21 γ 31   f1  = γ 11 γ 21 γ 31   fz  (6.80)
 A3  4π γ 21 γ 22 γ 32   f2  4π γ 21 γ 22 γ 32  − fe 

For the Green’s function problem, we can assume (fz, −fe) respectively equals (1,0)
and (0,1), and therefore, the corresponding Green’s functions due to the vertical
point force and point electric charge can be finally found.

6.4.1.2 Solutions of a Horizontal Point Force along x-Axis


The potential functions are assumed in the form of Eq. (6.45), from which we
obtain

 1 i y( x + i y)  3  1 x( x + i y) 
* 2  ∑ i  R*
U = B0  * + − B − (6.81)
R
 0 R R
0 0  i =1  i Ri Ri*2 

3
βik x
wk = − sgn(z − h)∑ Bi (k = 1, 2) (6.82)
i =1 Ri Ri*

 2( x + i y)  1 2 
σ 2 = −2c66 B0  + i y( x + i y)2  3 *2 + 2 *3  
 R R0
 R0 R0
* 2
 0 R0 R0  
(6.83)
3  2( x + i y)  1 2 
+ 2c66 ∑ Bi  * 2
− x( x + i y)2  3 *2 + 2 *3  
i =1  Ri Ri  Ri Ri Ri Ri  

3
γ ik x
σ zk = ∑ Bi (k = 1, 2, 4) (6.84)
i =1 Ri3

 1  1 1 
τ zk = − sgn(z − h)s0ν k B0  *
+ i y( x + i y)  3 * + 2 *2  
 R0 R0  R0 R0 R0 R0  
(6.85)
3  1  1 1 
− sgn(z − h)∑ϖ ik Bi  *
− x( x + i y)  3 * + 2 *2   (k = 1, 2)
i =1  Ri Ri  Ri Ri Ri Ri  
6.4 Various Decoupled Solutions 195

Similarly, the four coefficients Bi (i = 0–3) can be determined from the continuity
conditions across the source level z = h as well as the force balance of the layer
bounded by the planes z = h ± ε . The results are

fx
B0 = (6.86)
4π c44 s0

−1
 B1  fx  s1 s2 s3   1
 B2  = −  β11 β 21 β 31  0  (6.87)
 B3  4π c44  β12 β 22 β 32  0 

For the Green’s function solutions, we let fx = 1. Just as for the fully coupled MEE
case, the extended Green’s displacements and stresses due to the horizontal point
load along the y-direction can be obtained simply by switching x and y on both sides
of the displacement and stress expressions due to the x-direction force.

6.4.2 Elastic Green’s Functions


The Green’s function solutions corresponding to the purely elastic full space case
can be obtained from the piezoelectric case through a proper degenerate analysis.
The eigen-equation is further reduced to

n2 s 4 − n3 s 2 + n4 = 0 (6.88)

where

n2 = c44 c33 , 2 − 2c c ,
n3 = c11c33 − c13 n4 = c11c44 . (6.89)
13 44

which can be obtained from Eq. (6.67) by setting ε 33 = 0 and ε 11 = 1 along with eij = 0.
The elastic displacements are given by

∂ Ψ 0 2 ∂Ψ i ∂Ψ 0 2 ∂Ψ i 2
∂Ψ i
ux = −∑ , uy = − −∑ , uz = ∑ βi 1 (6.90)
∂y i = 1 ∂x ∂x i = 1 ∂y i=1 ∂zi

where Ψ i (i = 0, 1, 2) satisfy Eq. (6.19), and βi1 are given by Eqs. (6.69) and (6.70) for
i, j = 1, 2. Furthermore,

a1 = 0, b1 = 0, g1 = 0, d1 = c13 + c44
(6.91)
a2 = 0, b2 = 0, d2 = c44 , g2 = c11 .

Using the compact complex notation, we can write the displacements and
stresses as

 2  2
∂Ψ i
U = − Λ  ∑ Ψi + i Ψ0  , w1 = ∑ βi 1 (6.92)
 i =1  i =1 ∂zi
196 Green’s Functions in a Transversely Isotropic Magnetoelectroelastic Full Space

 2 
σ 2 = −2c66 Λ 2  ∑ Ψ i + i Ψ 0  (6.93)
 i =1 

2
∂ 2 Ψi
σzk = ∑ γ ik (k = 1, 4) (6.94)
i=1 ∂zi2

 2 ∂Ψ i ∂Ψ0 
τ z1 = Λ  ∑ ϖi1 − i s0 c44 (6.95)
 i =1 ∂zi ∂z0 

where

γ i1 = c13 + c33 si βi1 , γ i 4 = 2[(c11 − c66 ) + c13 si βi 1 ], ϖi 1 = − c44 si + c44βi 1 + e15βi 2

These expressions can be obtained by simply setting eij = 0 in Eq. (6.74). The relation
(6.24) still holds for i = 1, 2 and j = 1.

6.4.2.1 Solutions of a Vertical Point Force


The displacements and stresses are

2
( x + i y)
U = − sgn(z − h)∑ Ai (6.96)
i =1 Ri Ri*

2
βi1
uz = ∑ Ai (6.97)
i =1 Ri

2  1 1 
σ 2 = 2c66 sgn(z − h)∑ Ai ( x + i y)2  3 * + 2 *2  (6.98)
i =1  Ri Ri Ri Ri 

2
γ ik si (z − h)
σ zk = −∑ Ai (k = 1, 4) (6.99)
i =1 Ri3

2
ϖ i 1 ( x + i y)
τ z1 = −∑ Ai (6.100)
i =1 Ri3

The two coefficients Ai (i = 1,2) are determined from the following equation:

{AA } = 41π γ1 {f0 }


−1
1 1 
(6.101)
2 11 γ 21  z

For the Green’s function problem, we assume fz=1, and therefore, the corresponding
Green’s functions due to the vertical point force can be finally found.
6.4 Various Decoupled Solutions 197

6.4.2.2 Solutions of a Horizontal Point Force along x-Axis


In accordance with the potential functions assumed in Eq. (6.45), we have

 1 i y( x + i y)  2  1 x( x + i y) 
* 2  ∑ i  R*
U = B0  * + − B − (6.102)
R
 0 R R
0 0  i =1  i Ri Ri*2 

2
β i1 x
uz = − sgn(z − h)∑ Bi (6.103)
i =1 Ri Ri*

 2( x + i y)  1 2 
σ 2 = −2c66 B0  + i y( x + i y)2  3 *2 + 2 *3  

 R0 R0
* 2
 R0 R0 R0 R0  
(6.104)
2  2( x + i y)  1 2 
+ 2c66 ∑ Bi  * 2
− x( x + i y)2  3 *2 + 2 *3  
i =1  Ri Ri  Ri Ri Ri Ri  

2
γ ik x
σ zk = ∑ Bi (k = 1, 4) (6.105)
i =1 Ri3

 1  1 1 
τ z1 = − sgn(z − h)s0ν 1 B0  *
+ i y( x + i y)  3 * + 2 *2  
 
 R0 R0  R0 R0 R0 R0  
(6.106)
2  1  1 1 
− sgn(z − h)∑ϖ i 1 Bi  *
− x( x + i y) +
 R 3 Ri* R 2 Ri*2  

i =1  Ri Ri  i i  

Similarly, the three coefficients Bi (i = 0–2) can be determined from the continuity
conditions across the source level as well as the force balance of the layer bounded
by the planes z = h ± ε :

fx
B0 = (6.107)
4π c44 s0

{ } {01}
−1
B1 fx  s1 s2 
B2 = − 4π c  β11 β 21  (6.108)
44

For the Green’s function solutions, we let fx=1. Again, the Green’s displacements
and stresses due to the horizontal point load along the y-direction can be obtained
simply by switching x and y on both sides of the displacement and stress expressions
due to the x-direction force.
For transversely isotropic elastic materials, Ding et al. (1997) derived a unified
form of Green’s functions that are valid either for transversely isotropic materials
or the degenerated isotropic materials. Such unified-form Green’s functions
are beneficial to numerical calculations because for isotropic materials we have
s1 = s2 = 1 which results in the indeterminate expression with zero divided by
198 Green’s Functions in a Transversely Isotropic Magnetoelectroelastic Full Space

zero (Ding et al. 1997). Fabrikant (2003) also presented the unified-form Green’s
functions in a complex notation, and he further showed that such unified form is
actually not unique. We have not tried to obtain these unified Green’s functions for
MEE materials here; in practice, we can avoid the numerical difficulty by slightly
changing the material constants so that the eigenroots si become different from
each other, an approach that can be very efficient because one can directly use the
solutions presented in this chapter (Pan 1997).

6.5 Technical Applications

6.5.1 Eshelby Inclusion Solution in Terms of the Green’s Functions


As one of the most important and immediate applications of the Green’s function
solutions provided in the preceding text, we analyze the Eshelby problem of an
inclusion in an infinite transversely isotropic MEE space (see Figure 6.2). While the
derivation of the Eshelby solution in terms of the Green’s functions is presented in
Chapter 2, here we offer a different way to derive the same expression.
For an extended general eigenstrain ε Ij* at x = (x, y, z) within the domain V, the
induced extended displacement at any point X = (X, Y, Z) can be found by virtue of
the Eshelby theory. In other words, the response is an integral, over the whole space
E, of an equivalent body force in the square bracket in the following text, multiplied
by the point-source Green’s function (i.e., Mura 1987; Pan 2004)

uK ( X ) = − ∫ uJK ( x; X )[ciJLm ε*Lm ( x )],i dV ( x ) (6.109)


E

where uJK ( x; X ) is the J component of the extended Green’s displacements (elastic


displacements, electric and magnetic potentials) at x due to an extended point source
(point force, point electric and magnetic charges) in the K-direction applied at X,
as have been derived earlier in this chapter and also summarized in Appendix A
in terms of the relative coordinates between the source and field points. It should
be noted that due to the discontinuity of the extended eigenstrain field across the
interface between the inclusion and matrix, the equivalent body force in Eq. (6.109)
is generally not zero (even for the uniform eigenstrain case).
Integrating by parts and noticing that the extended eigenstrain is nonzero only
in part V of the whole space E, Eq. (6.109) can be written alternatively as

uK ( X ) = ∫ uJK, x ( x; X )ciJLm ε*Lm ( x )dV ( x ) (6.110)


V i

If we further assume that the extended eigenstrain is uniform within the domain V,
then Eq. (6.110) can be reduced to

uK ( X ) = ciJLm ε*Lm ∫ uJK, x ( x; X )dV ( x ) (6.111)


V i
6.5 Technical Applications 199

∂V
V
ε*Ij

E-V

Figure 6.2. An arbitrary inclusion V with uniform extended eigenstrain ε Ij


* embedded in an
infinite MEE space E. The boundary of the inclusion is ∂V.

This expression is the same as the first expression in Eq. (2.37). Because in the
infinite space, the solution depends only on the relative position of the field and
source points, Eq. (6.111) can be rewritten as

(6.112)
uK ( X ) = ciJLm ε*Lm ∫ uJK, x ( x − X )dV ( x )
V i

Furthermore, due to the fact that the derivatives of Green’s displacements in a three-
dimensional full space are proportional to 1/R2, where R = x − X , we can introduce
the following unit vector l = (l1, l2, l3)t

l = (x − X ) / R (6.113)

to express the derivatives of the Green’s functions as (Dunn and Wienecke 1997)

uJK, x ( x − X ) ≡ GJiK (l ) / R 2 (6.114)


i

This changes Eq. (6.112) to

(6.115)
uK ( X ) = ε*Lm ∫ ciJLmGJiK (l )dΩ∫ dR

where Ω is the surface of the unit sphere (see Figure 6.3), and

dV = R2 dRdΩ ≡ R2 dR sin θ dθ dϕ (6.116)


200 Green’s Functions in a Transversely Isotropic Magnetoelectroelastic Full Space

X3

X2
ϕ

X1

Figure 6.3. A unit sphere with the spherical angles (θ,ϕ ).

is the differential volume element of the inclusion.


Integrating R in Eq. (6.115) gives
(6.117)
uK ( X ) = ε*Lm ∫ ciJLmGJiK (l )R(l )dΩ

where R(l) defines the boundary of the (convex) inclusion. In what follows, we
discuss the ellipsoidal inclusion case (with semiaxes a1, a2, and a3 being all different
from each other) with X in the inclusion. For this case, R(l) is the positive root of the
following equation (Eshelby 1957)

( X 1 + Rl1 )2 ( X 2 + Rl2 )2 ( X 3 + Rl3 )2


+ + =1 (6.118)
a12 a22 a32

which can be solved as

R(l ) = − f / g + ( f 2 / g 2 + e / g)1/ 2 (6.119)

where,

f = X 1l1 / a12 + X 2 l2 / a22 + X 3 l3 / a32


g = l12 / a12 + l22 / a22 + l32 / a32 (6.120)
e = 1 − (X 12 / a12 + X 22 / a22 + X 32 / a32 )

Because the derivatives of the Green’s displacements are odd functions of the
orientation l, and (f2/g2+e/g)1/2 is even in l, their product integration over Ω (the unit
sphere surface) is zero. Thus, Eq. (6.117) can be changed to

α s ciJLmGJiK (l ) (6.121)
uK ( X ) = − X s ε*Lm ∫ dΩ
Ω g
6.5 Technical Applications 201

where α i = li / ai2 (i = 1, 2, 3) . Taking the derivative of Eq. (6.121) with respect to the
field point X (i.e., the source point of the point-force/point-charges Green’s function),
we have

α p ciJLmGJiK (l ) (6.122)
uK , X p ( X ) = − ε*Lm ∫ dΩ
Ω g

from which, we can immediately derive

ε Kp = SKpLmε Lm
* (6.123)

where

 1 α p ciJLmGk (l ) + α k ciJLmG p (l )
− ∫ Ji Ji
dΩ (K = k = 1, 2, 3)
 2 Ω g
   SKpLm = (6.124)
 α p ciJLmGJiK (l )
 −∫

dΩ (K = 4, 5)
 g

is the extended Eshelby tensor. Because the integrand is independent of X, the


induced extended strains inside the ellipsoidal inclusion are uniform when a uniform
extended eigenstrain field is applied, and are further linearly proportional to the
eigenstrain field, with the proportional coefficients being the extended Eshelby
tensor. Therefore, the key is to calculate this Eshelby tensor.
As shown by Eq. (6.124), for the general ellipsoidal inclusion case, the extended
Eshelby tensor is expressed as the integral of the derivatives of the point-source
Green’s function over the surface of the unit sphere. Such an integral can be
carried out using, say, the Gauss quadrature method as in Li (2000). However, for
a spheroidal inclusion, with axes a1=a2≡a and the axis a3≡c being normal to the
plane of isotropy, the integrals in Eq. (6.124) can be carried out analytically so that
we have exact-closed form expressions for the extended Eshelby tensor. This is
discussed in detail in the following text.

6.5.2 Elements of the Extended Eshelby Tensor


Looking at Eq. (6.124), one only needs to deal with the following typical integral (of
the newly introduced tensor PKpJi)

α pGJiK (l ) (6.125)
PKpJi = ∫ dΩ
Ω g

In terms of this tensor, the Eshelby tensor can be expressed as

 −( PkpJi + PpkJi )ciJLm / 2 (K = k = 1, 2, 3)


SKpLm =  (6.126)
 − PKpJi ciJLm (K = 4, 5)
202 Green’s Functions in a Transversely Isotropic Magnetoelectroelastic Full Space

Now we express the unit vector l in terms of the spherical angles (θ,ϕ ) as

l = (sin θ cos ϕ , sin θ sin ϕ , cos θ ) (6.127)

Then g and αi can be expressed as

g = (c 2 sin 2 θ + a 2 cos2 θ ) / (a 2 c 2 )
(6.128)
α = (sin θ cos ϕ / a 2 , sin θ sin ϕ / a 2 , cos θ / c 2 )

With these changes, Eq. (6.125) can be rewritten as

π sin θ dθ 2π
PKpJi = ∫ ∫0 α pGJiK (l )dϕ (6.129)
0 g(θ )

where GJiK (l ) are the derivatives of the extended Green’s displacements with their
components being given in Appendix A. In terms of its p-components, Eq. (6.129)
can be expressed as

( PK 1Ji , PK 2 Ji , PK 3 Ji )
π sin θ dθ 2π  sin θ cos ϕ sin θ sin ϕ cos θ  K (6.130)
=∫
g(θ ) ∫0 
 , , 2  GJi (l )dϕ
0 a2 a2 c 

We now calculate the individual components involved in Eq. (6.130) for each of
these terms.
K with K = 3,4,5, J = 3,4,5, and m = 1,2,3
(1) For GJm
From Eq. (A21), we have

( PK 1Jm , PK 2 Jm , PK 3 Jm )
q
π sin θ dθ 2π  sin θ cos ϕ sin θ sin ϕ cos θ  4 AiK β i, J − 2 si lm
= −∫
0 g(θ )
×∫ 
0  a2
,
a2
,  ∑
c 2  i =1 ri3

(6.131)

where ri = sin 2 θ + si2 cos2 θ , q=0 for m = 1 and 2, and q = 2 when m = 3.


For m = 1,2,3, it can be separately written as

π π (1, 0, 0)sin 3 θ dθ 4 AiK β i, J − 2


( PK 1J 1 , PK 2 J 1 , PK 3 J 1 ) = −
a2 ∫0 g(θ )
∑ ri3
i =1
4 AiK β i, J − 2
π π (0, 1, 0)sin 3 θ dθ
( PK 1J 2 , PK 2 J 2 , PK 3 J 2 ) = −
a2 ∫0 g(θ )
∑ ri3
(6.132)
i =1
2π π (0, 0, 1)sin θ cos 2 θ dθ 4 AiK β i, J − 2 si2
( PK 1 J 3 , PK 2 J 3 , PK 3 J 3 ) = −
c2 ∫0 g(θ ) ri3

i =1

which can be expressed as


6.5 Technical Applications 203

4
( PK 1J 1 , PK 2 J 1 , PK 3 J 1 ) = −π (1, 0, 0)∑ AiK β i, J − 2 H1 (b, si )
i =1
4
(PK 1J 2 , PK 2 J 2 , PK 3 J 2 ) = −π (0, 1, 0)∑ AiK β i, J − 2 H1 (b, si ) (6.133)
i =1
4
( PK 1J 3 , PK 2 J 3 , PK 3 J 3 ) = −2π (0, 0, 1)∑ AiK β i, J − 2 si2 H 2 (b, si )
i =1

where

1 π sin 3 θdθ sin 3 θdθ


π
H1 (b, si ) =
a2 ∫0 g(θ)ri3
≡∫
0 (cos2 θ + b2 sin 2 θ)r 3
,
i
(6.134)
1 π sin θ cos2 θdθ π sin θ cos2 θdθ
H 2 (b, si ) = 2 ∫ ≡ b 2
∫0 (cos2 θ + b2 sin2 θ)ri3
c 0 g(θ)ri3

The two functions H1 and H2 can be integrated by introducing t = cos θ and letting
b = a/c. The detailed results are given in Appendix B of this chapter.
K with K = 3,4,5
(2) For G11
For this component, we have (from Eq. (A18)1)

( PK 111 , PK 211 , PK 311 )


π sin θ dθ 2π  sin θ cos ϕ sin θ sin ϕ cos θ  4 AiK
= −∫
0 g(θ )
sgn(l3 )∫ 
0  a2
,
a2
,  ∑
c 2  i =1 ri (ri + si l3 )

π sin θ dθ 2π  sin θ cos ϕ sin θ sin ϕ cos θ 


+∫ sgn(l3 )∫  , , 2  (6.135)
0 g(θ ) 0  a2 a2 c 
4  1 1 
× ∑ AiK l12  3 + 2 2  dϕ
i =1  ri (ri + si l3 ) ri (ri + si l3 ) 

It can be reduced to

4 AiK
2π (0, 0, 1) π sin θ cos θ dθ
( PK 111 , PK 211 , PK 311 ) = − ∫0 g(θ ) sgn(cos θ )∑
c2 i =1 ri (ri + si cos θ )
π (0, 0, 1) π sin 3 θ cos θ dθ 4  1 1 
+ ∫ sgn(cos θ ) ∑ AiK  3 + 2 
g(θ )  ri (ri + si cos θ ) ri (ri + si cos θ ) 
c 2 0 2
i =1
 (6.136)

which can be further expressed as

4
(PK 111 , PK 211 , PK 311 ) = π (0, 0, 1)∑ AiK [ −2 H 3 (b, si ) + H 4 (b, si ) + H 5 (b, si )]
i =1 (6.137)
204 Green’s Functions in a Transversely Isotropic Magnetoelectroelastic Full Space

where

1 π sin θ cos θ sgn(cos θ )dθ π sin θ cos θ sgn(cosθ )dθ


H 3 (b, si ) =
c2 ∫0 g(θ )ri (ri + si cos θ )
≡ b2 ∫
0 (cos2 θ + b2 sin 2 θ )r (r + s cos θ )
i i i
1 π sin 3 θ cos θ sgn(cos θ )dθ π sin 3 θ cosθ sgn(cosθ )dθ
H 4 (b, si ) =
c2 ∫0 g(θ )ri3 (ri + si cos θ )
≡ b 2
∫0 (cos2 θ + b2 sin2 θ )ri3 (ri + si cosθ )
1 π sin 3 θ cos θ sgn(cosθ )dθ π sin 3 θ cosθ sgn(cosθ )dθ
H 5 (b, si ) =
c2 ∫0 g(θ )ri (ri + si cosθ )
2 2
≡ b2 ∫
0 (cos θ + b2 sin 2 θ )r 2 (r + s cos θ )2
2
i i i

(6.138)

Again, these integrals are discussed in Appendix B of this chapter.


K with K = 3,4,5
(3) For G12
From Eq. (A18)2, we have

(PK 112 , PK 212 , PK 312 ) = (0, 0, 0) (6.139)

K with K = 3,4,5
(4) For G13
From Eq. (A18)3, we have

4
( PK 113 , PK 213 , PK 313 ) = π (1, 0, 0)∑ AiK si H1 (b, si ) (6.140)
i =1

K with K = 3,4,5
(5) For G21
From Eq. (A20)1, we have

( PK 121 , PK 221 , PK 321 ) = ( PK 112 , PK 212 , PK 312 ) (6.141)

K with K = 3,4,5
(6) For G22
From Eq. (A20)2, we have

( PK 122 , PK 222 , PK 322 ) = (PK 111 , PK 211 , PK 311 ) (6.142)

K with K = 3,4,5
(7) For G23
From Eq. (A20)3, we have

4
( PK 123 , PK 223 , PK 323 ) = π (0, 1, 0)∑ AiK si H1 (b, si ) (6.143)
i =1

Notice that, in the preceding expressions (1) to (7) for different K (K = 3,4,5), the
corresponding coefficients AiK are different because they correspond to different
source Green’s functions as indicted by the superscript K to them.
6.5 Technical Applications 205

1
(8) For G11
From Eq. (A22)1, we have

(P1111 , P1211 , P1311 )


π sin θ dθ 2π  sin θ cos ϕ sin θ sin ϕ cos θ 
=∫
0 g(θ ) ∫0 
, , 2  dϕ
 a2 a2 c 
 − B0 l1  2
l2 l1 2l22 l1  (6.144)
× 2
+ B 0  3 2
+ 2 3 
 r0 (r0 + s0 l3 )  r0 (r0 + s0 l3 ) r0 (r0 + s0 l3 ) 
4  3l1 l13 2l13  
+ ∑ Bi  2
− 3 2
− 2 3 
i =1  ri (ri + si l3 ) ri (ri + si l3 ) ri (ri + si l3 )  

which can be written as

( P1111 , P1211 , P1311 )


π sin θ dθ 2π  sin θ cos ϕ sin θ sin ϕ cos θ 
=∫
g(θ ) ∫0 
 , , 2  dϕ
0 a2 a2 c 
 − B sin θ cos ϕ  sin θ cos ϕ sin ϕ 2 sin 3 θ cos ϕ sin 2 ϕ 
3 2
(6.145)
× 0 2
+ B 0  3 + 
 r0 (r0 + s0 l3 )  r0 (r0 + s0 l3 )2 r02 (r0 + s0 l3 )3 
4  3 sin θ cos ϕ sin 3 θ cos3 ϕ 2 sin 3 θ cos3 ϕ  
+ ∑ Bi  − 3 − 
i =1  ri (ri + si l3 )
2 ri (ri + si l3 )2 ri2 (ri + si l3 )3  

Integrating it over ϕ, we have

π (1, 0, 0) π dθ
( P1111 , P1211 , P1311 ) =
a2 ∫0 g(θ )
 − B0 sin 3 θ B  1 2 
× 2
+ 0 sin 5 θ  3 2
+ 2 3  (6.146)
 r0 (r0 + s0 l3 ) 4  r0 (r0 + s0 l3 ) r0 (r0 + s0 l3 ) 
4  3 sin 3 θ 3 5  1 2   
+ ∑ Bi  2
− sin θ +
 r 3 (ri + si l3 )2 r 2 (ri + si l3 )3   

i =1  ri (ri + si l3 ) 4  i i   

or,

π
( P1111 , P1211 , P1311 ) = (1, 0, 0){B0 [−4 H 6 (b, s0 ) + H 7 (b, s0 ) + 2 H8 (b, s0 )]
4
4 (6.147)
+ ∑ 3Bi [4 H6 (b, si ) − H 7 (b, si ) − 2 H8 (b, si )]}
i =1

where
206 Green’s Functions in a Transversely Isotropic Magnetoelectroelastic Full Space

1 π sin 3 θ dθ π sin 3 θ dθ
H6 (b, si ) =
a2 ∫0 g(θ )ri (ri + si cosθ )2 ∫0 (cos2 θ + b2 sin 2 θ )ri (ri + si cos θ )2

1 π sin 5 θ dθ π sin 5 θ dθ
H 7 (b, si ) =
a2 ∫0 g (θ )ri3 (ri + si cosθ ) 2
≡∫
0 (cos θ + b sin 2 θ )ri3 (ri + si cosθ )2
2 2

1 π sin 5 θ dθ π sin 5 θ dθ
H8 (b, si ) =
a2 ∫0 g(θ )ri2 (ri + si cos θ )3
≡ ∫0 (cos2 θ + b2 sin 2 θ )r 2 (r + s cos θ )3
i i i
 (6.148)

These integrals are discussed in Appendix B of this chapter.


1
(9) For G12
From Eq. (A22)2, we have

(P1112 , P1212 , P1312 )


π sin θ dθ 2π  sin θ cos ϕ sin θ sin ϕ cos θ 
=∫
0 g(θ ) ∫0 
, , 2  dϕ
 a2 a2 c 
  l23 2l23  3B0 l2 (6.149)
×  B0  3 2
+ 2 −
3  r (r + s l )2
r ( r
  0 0 0 3 + s l ) r (
0 0 r + s l
0 3 )  0 0 0 3
4  l2 2
l2 l1 2l2 l12  
+ ∑ Bi  2
− 3 2
− 2 3 
i =1  ri (ri + si l3 ) ri (ri + si l3 ) ri (ri + si l3 )  

which can be expressed as

π
( P1112 , P1212 , P1312 ) = (0, 1, 0){3B0 [ −4 H 6 (b, s0 ) + H 7 (b, s0 ) + 2 H 8 (b, s0 )]
4
4
+ ∑ Bi [ H6 (b, si ) − H 7 (b, si ) − 2 H8 (b, si )]}
i =1 (6.150)

1
(10) For G13
From Eq. (A22)3, we have

( P1113 , P1213 , P1313 ) = π (0, 0, 1){B0 s0 [ −2 H 3 (b, si ) + H 4 (b, s0 ) + H 5 (b, s0 )]


4 (6.151)
+ ∑ Bi si [ 2 H 3 (b, si ) − H 4 (b, si ) − H 5 (b, si )] }
i =1

1
(11) For G21
From Eq. (A23)1, we have

4
π
( P1121 , P1221 , P1321 ) = (0, 1, 0)∑ Bi [4 H 6 (b, si ) − H 7 (b, si ) − 2 H 8 (b, si )] (6.152)
4 i =0
6.5 Technical Applications 207

1
(12) For G22
From Eq. (A23)2, we have

4
π
   ( P1122 , P1222 , P1322 ) = (1, 0, 0)∑ Bi [4 H6 (b, si ) − H 7 (b, si ) − 2 H8 (b, si )] (6.153)
4 i =0

1
(13) For G23
From Eq. (A23)3, we have

( P1123 , P1223 , P1323 ) = (0, 0, 0) (6.154)

1 with J = 3,4,5
(14) For GJ1
From Eq. (A24)1, we have

4
( P11J 1 , P12 J 1 , P13 J 1 ) = π (0, 0, 1)∑ Bi β i, J − 2 [ −2 H 3 (b, si ) + H 4 (b, si ) + H 5 (b, si )] (6.155)
i =1

(15) For GJ1 2 with J = 3,4,5


From Eq. (A24)2, we have

( P11J 2 , P12 J 2 , P13 J 2 ) = (0, 0, 0) (6.156)

(16) For GJ1 3 with J = 3,4,5


From Eq. (A24)3, we have

4
( P11J 3 , P12 J 3 , P13 J 3 ) = π (1, 0, 0)∑ Bi β i, J − 2 si H1 (b, si ) (6.157)
i =1

(17) For G12j with J = 1,2,3

From Eq. (A25), we have

( P211 j , P221 j , P231 j ) = ( P112 j , P122 j , P132 j ) (6.158)

2
(18) For G21
From Eq. (A26)1, we have

π
( P2121 , P2221 , P2321 ) = (1, 0, 0){B0 [ −4 H 6 (b, s0 ) + H 7 (b, s0 ) + 2 H 8 (b, s0 )]
4
   4 (6.159)
+ ∑ Bi [4 H 6 (b, si ) − H 7 (b, si ) − 2 H8 (b, si )]}
i =1
208 Green’s Functions in a Transversely Isotropic Magnetoelectroelastic Full Space

2
(19) For G22
From Eq. (A26)2, we have

π
( P1112 , P1212 , P1312 ) = (0, 1, 0){B0 [ −4 H 6 (b, s0 ) + H 7 (b, s0 ) + 2 H 8 (b, s0 )]
4
   4 (6.160)
+ ∑ 3Bi [ 4 H 6 (b, si ) − H 7 (b, si ) − 2 H 8 (b, si )]}
i =1

2
(20) For G23
From Eq. (A26)3, we have

(P2123 , P2223 , P2323 ) = (P1113 , P1213 , P1313 ) (6.161)

2 with J = 3,4,5
(21) For GJ1
From Eq. (A27)1, we have

( P21 J 1 , P22 J 1 , P23 J 1 ) = ( P11 J 2 , P12 J 2 , P13 J 2 ) (6.162)

(22) For GJ22 with J = 3,4,5


From Eq. (A27)2, we have

4
( P21J 2 , P22 J 2 , P23 J 2 ) = π (0, 0, 1)∑ Bi β i, J − 2 [ −2 H 3 (b, si ) + H 4 (b, si ) + H 5 (b, si )]
i =1
 (6.163)

(23) For GJ23 with J = 3,4,5


From Eq. (A27)3, we have

4
( P21 J 3 , P22 J 3 , P23 J 3 ) = π (0, 1, 0)∑ Bi β i, J − 2 si H1 (b, si ) (6.164)
i =1

6.5.3 Special Cases

6.5.3.1 Special Geometric Cases


For the special geometric cases, the expressions of the Eshelby tensor in terms of
the functions Hi (i = 1–8) will be the same but the Hi functions will be simplified.
The three common cases are: (1) the spherical case (a=c, or the semi-axis ratio b=1);
(2) infinite circular cylinder (a remains fixed but c approaches infinity; in other
words, b=0); and (3) disklike inclusion (a remains fixed but c approaches zero; in
other words, b approaches infinity). The corresponding Hi functions are given in
Appendix B.
6.7 Appendix A: The Extended Green’s Functions and Their Derivatives 209

6.5.3.2 Special Material Coupling Cases


The Eshelby tensor for the corresponding piezoelectric case can be obtained directly
from the Eshelby tensor for the fully coupled MEE case by assigning the range of
the indices J and K from 1 to 4 only and by setting µ33 = 0 and µ11 = 1 along with
qij = α ij = 0. All the involved coefficients from the Green’s functions correspond to
the reduced piezoelectric Green’s functions given in Section 6.4.1. Similarly, for the
purely elastic case, we take the range of the indices J and K from 1 to 3 only, and
further let ε 33 = 0 and ε 11 = 1 along with eij = 0 in the corresponding piezoelectric
Eshelby tensor. Again, the involved Green’s function coefficients should be those
from Section 6.4.2. Finally, for the piezomagnetic case, one can directly use the
Eshelby tensor corresponding to the piezoelectric case by replacing the piezoelectric/
electric quantities by the corresponding piezomagnetic/magnetic ones.

6.6 Summary and Mathematical Keys

6.6.1 Summary
The potential function method is applied to derive the Green’s function solutions
in a transversely isotropic MEE full space. The sources are the point forces in the
coordinate directions and the point electric/magnetic charges. The Green’s functions
also contain those corresponding to the decoupled material full spaces, namely, the
transversely isotropic piezoelectric or piezomagnetic full space and the transversely
isotropic elastic full space. Different from the previous Green’s functions for the
decoupled material systems (Pan and Chou 1976; Karapetian et al. 2000), here our
Green’s functions are singular only at the source point! It was achieved by introducing
the sign function and the absolute value of the relative z-coordinate, based on the
intuitive prediction of the Green’s function behavior under a given point source. As
a direct and important application of the derived Green’s functions, we have also
presented the corresponding Eshelby inclusion solution, that is, the extended Eshelby
tensor, when the inclusion is in a spheroidal shape. The Eshelby tensor for the reduced
geometric and material decoupled cases is also discussed.

6.6.2 Mathematical Keys


We introduced the special potential functions containing the sign functions and the
absolute value of the relative z-coordinate so that the derived Green’s function is
singular only at the source point. Based on the analytical Green’s function solution, we
have also derived the extended Eshelby tensor in analytical form, which should be very
useful to various 3D problems associated with Eshelby inclusion and inhomogeneity.

6.7 Appendix A: The Extended Green’s Functions and Their Derivatives

6.7.1 The Extended Green’s Displacements


We first summarize the expressions of uJK ( x; X ) for the J-th extended Green’s
displacements (elastic displacements, electric and magnetic potentials) at x due to
an extended unit point source (point force, point electric, and magnetic charges) in
210 Green’s Functions in a Transversely Isotropic Magnetoelectroelastic Full Space

the K-direction applied at X. We provide the results in terms of the relative distance
Ri (i = 0–4) between the source (x) and field (X)

2
Ri = ( x − X )2 + ( y − Y )2 + si2 z − Z (A1)

The extended displacements given in Eqs. (6.34) and (6.35) due to a vertical point
force (K = 3), a negative point electric charge (K = 4), and a negative point magnetic
charge (K = 5) can be expressed as

4
AiK
u1K ( x − X ) = − sgn(z − Z )( x − X )∑
i = 1 Ri (Ri + si z − Z )
4
AiK
u2K ( x − X ) = − sgn(z − Z )( y − Y )∑ (A2)
i = 1 Ri (Ri + si z − Z )
4
AiK βi, J − 2
uJK ( x − X ) = ∑ ( J = 3, 4, 5)
i =1 Ri

where for different K (K = 3,4,5), the coefficients Ai (i=1–4) are different. In other
words, Ai are the functions of the point-source direction K, as explained after Eq.
(6.43).
The extended displacements in Eqs. (6.49) and (6.50) due to a point force in the
horizontal x-direction (K = 1) can be expressed as

4  
− Bi Bi ( x − X )2
u11 ( x − X ) = ∑  + 2 
i = 1  Ri + si z − Z Ri (Ri + si z − Z ) 
B0 ( y − Y )2 B0
− +
R0 (R0 + s0 z − Z )2 R0 + s0 z − Z
4 (A3)
Bi
u21 ( x − X ) = ( x − X )( y − Y )∑ 2
R
i=0 i ( R +
i si z − Z )
4 Bi βi, J − 2
u1J ( x − X ) = − sgn(z − Z )( x − X )∑ ( J = 3, 4, 5)
i = 1 Ri (Ri + si z − Z )

The coefficients Bi (i = 0–4) are determined by the approach described in


Section 6.3.
Similarly, the extended displacements due to a point force in the horizontal y-
direction (K = 2) can be expressed as

4
Bi
u12 ( x − X ) = ( x − X )( y − Y )∑ 2
≡ u21 ( x − X )
R
i=0 i ( Ri si z − Z )
+
4  
− Bi Bi ( y − Y )2
u22 ( x − X ) = ∑  + 2 
i = 1  Ri + si z − Z Ri (Ri + si z − Z ) 
(A4)
B0 ( x − X )2 B0
− 2
+
R0 (R0 + s0 z − Z ) R0 + s0 z − Z
4
Bi βi, J − 2
uJ2 ( x − X ) = − sgn(z − Z )( y − Y )∑ ( J = 3, 4, 5)
i = 1 Ri (Ri + si z − Z )
6.7 Appendix A: The Extended Green’s Functions and Their Derivatives 211

Again, the coefficients Bi (i = 0–4) are determined by the approach described in


Section 6.3.

6.7.2 Derivatives of the Extended Green’s Displacements


6.7.2.1 Derivatives of the Extended Green’s Displacements Due to
the Point Source in the K-direction (K = 3,4,5)
We now take the derivatives of Eq. (A2) with respect to the i-th variable of (x – X),
with the second subscript being the derivative component.

u1K,1 ( x − X ) = − sgn(z − Z )
4  1  1 1  
× ∑ AiK  − ( x − X )2  3 + 2 2 
i =1  Ri (Ri + si z − Z )  Ri (Ri + si z − Z ) Ri (Ri + si z − Z )  
4  1 1 
u1K,2 ( x − X ) = sgn(z − Z )( x − X )( y − Y )∑ AiK  3 + 2 2 
i =1  Ri (Ri + si z − Z ) Ri (Ri + si z − Z ) 
4 AiK si
u1K, 3 ( x − X ) = ( x − X )∑
Ri3 i =1
 (A5)

u2K,1 ( x − X ) = u1K, 2 ( x − X )
u2K, 2 ( x − X ) = − sgn(z − Z )
4  1  1 1  
× ∑ AiK  − ( y − Y )2  3 + 2 2 
i =1  Ri (Ri + si z − Z )  Ri (Ri + si z − Z ) Ri (Ri + si z − Z )  
4 AK s
u2K, 3 ( x − X ) = ( y − Y )∑ i 3
i

i =1 R i
 (A6)

4
AiK βi, J − 2
uJK,1 ( x − X ) = −( x − X )∑
i =1 Ri3
4
AiK βi, J − 2
uJK,2 ( x − X ) = −( y − Y )∑ ( J = 3, 4, 5) (A7)
i =1 Ri3
4
AiK βi, J − 2 si2
uJK,3 ( x − X ) = −(z − Z )∑
i =1 Ri3

6.7.2.2 Derivatives of the Extended Green’s Displacements Due to the


Point Source in x-direction
Similarly, by taking the derivatives of Eq. (A3) with respect to the i-th variable of
(x – X), we have
212 Green’s Functions in a Transversely Isotropic Magnetoelectroelastic Full Space

 1  1 2  
u11,1( x − X ) = B0 ( x − X )  − 2
+ ( y − Y )2  3 2
+ 2 3 
 R0 (R0 + s0 z − Z )  R0 (R0 + s0 z − Z ) R0 (R0 + s0 z − Z )  

  
4 
3 1 2
+ ( x − X )∑ Bi  2
− ( x − X )2  3 2
+ 2 3 
i=1 R (
 i i iR + s z − Z ) R (
 i i i R + s z − Z ) R i ( Ri + si z − Z )  
 (A8)

 3  1 2  
u11, 2 ( x − X ) = B0 ( y − Y )  − 2
+ ( y − Y )2  3 2
+ 2 3 
 R0 (R0 + s0 z − Z )  R0 (R0 + s0 z − Z ) R0 0 + 0 −
( R s z Z )  
4  1  1 2  
+ ( y − Y )∑ Bi  2
− ( x − X )2  3 2
+ 2 3 
i=1 R ( R
 i i i+ s z − Z ) R ( R
 i i i+ s z − Z ) Ri ( Ri + si z − Z )  
 (A9)

 1  1 1  
u11, 3 ( x − X ) = sgn(z − Z )B0 s0  − + ( y − Y )2  3 + 2 2 
R ( R
 0 0 0 + s z − Z ) R ( R
 0 0 0 + s z − Z ) R0 ( R0 + s0 z − Z )  
4  1  1 1  
+ ∑ sgn(z − Z )Bi si  − ( x − X )2  3 + 2 2 
i =1  Ri (Ri + si z − Z )  Ri (Ri + si z − Z ) Ri (Ri + si z − Z )  
 (A10)

4    
 Bi ( y − Y ) 1 2
u21,1( x − X ) = ∑  2
− Bi ( y − Y )( x − X )2  3 2
+ 2 3 
 Ri (Ri + si z − Z )
i=0   Ri (Ri + si z − Z ) Ri (Ri + si z − Z )  
4    
 Bi ( x − X ) 1 2
u21,2 ( x − X ) = ∑  2
− Bi ( x − X )( y − Y )2  3 2
+ 2 3 
 Ri (Ri + si z − Z )
i=0   Ri (Ri + si z − Z ) Ri (Ri + si z − Z )  
4  
1 1
u21, 3 ( x − X ) = − sgn(z − Z )( x − X )( y − Y )∑ Bi si  3 + 2 2
i=0 R (
 i i iR + s z − Z ) R i ( Ri + si z − Z ) 
 (A11)

4  −1  1 1  
u1J ,1 ( x − X ) = sgn(z − Z )∑ Biβi, J − 2  + ( x − X )2  3 + 2 2 
i=1  Ri (Ri + si z − Z )  Ri (Ri + si z − Z ) Ri (Ri + si z − Z )  
4    
 1 1
u1J , 2 ( x − X ) = sgn(z − Z )( x − X )( y − Y )∑  Biβi, J − 2  3 + 2 2 

i=1   Ri (Ri + si z − Z ) Ri (Ri + si z − Z )  
4
Biβi, J − 2 si
u1J , 3 ( x − X ) = ( x − X )∑
i=1 Ri3
( J = 3, 4, 5)
 (A12)

6.7.2.3 Derivatives of the Extended Green’s Displacements


Due to the Point Source in y-direction
Finally, by taking the derivatives of Eq. (A4) with respect to the i-th variable of
(x – X), we have
6.7 Appendix A: The Extended Green’s Functions and Their Derivatives 213

u12, j ( x − X ) = u21, j ( x − X ) ( j = 1, 2, 3)  (A13)

4  1  1 2  
u22,1 ( x − X ) = ( x − X )∑ Bi  2
− ( y − Y )2  3 2
+ 2 3 
i =1 R ( R
 i i i + s z − Z ) R ( R
 i i i+ s z − Z ) R i ( Ri + si z − Z )  
 −3  1 2  
+ ( x − X ) B0  2
+ ( x − X )2  3 2
+ 2 3 
 R0 ( R0 + s0 z − Z )  R0 (R0 + s0 z − Z ) R0 (R0 + s0 z − Z )  

 (A14)

4  3  1 2  
u22, 2 ( x − X ) = ( y − Y )∑ Bi  2
− ( y − Y )2  3 2
+ 2 3 
i=1  Ri (Ri + si z − Z )  Ri (Ri + si z − Z ) Ri (Ri + si z − Z )  
 −1  1 2  
+ ( y − Y )B0  2
+ ( x − X )2  3 2
+ 2 3 
 R0 (R0 + s0 z − Z )  R0 (R0 + s0 z − Z ) R0 (R0 + s0 z − Z )  

 (A15)

4  1  1 1  
u22, 3 ( x − X ) = sgn(z − Z )∑ Bi si  − ( y − Y )2  3 + 2 2 
i=1  Ri (Ri + si z − Z )  Ri (Ri + si z − Z ) Ri ( Ri + si z − Z )  
 −1  1 1  
+ sgn(z − Z )B0 s0  + ( x − X )2  3 + 2 2 
 R0 ( R0 + s0 z − Z )  R0 (R0 + s0 z − Z ) R0 (R0 + s0 z − Z )  
 (A16)

uJ2,1( x − X ) = u1J , 2 ( x − X )
uJ2, 2 ( x − X ) = sgn(z − Z )

  
4 
−1 1 1
× ∑ Biβi, J − 2  + ( y − Y )2  3 + 2 2 
i =1 R ( R
 i i i+ s z − Z ) R ( R
 i i i + s z − Z ) Ri ( Ri + si z − Z )  
4
Biβi, J − 2 si
uJ2, 3 ( x − X ) = ( y − Y )∑ ( J = 3, 4, 5)
i =1 Ri3
 (A17)

6.7.3 The Scaled Green’s Function Derivatives GJiK (l ) in Terms of the


Unit Vector l
In the calculation of the Eshelby tensor, ref. Eq. (6.130), only the scaled Green’s
function derivatives are needed. Based on the Green’s function derivatives uJK,i given
in Eqs. (A5)–(A17), we find
214 Green’s Functions in a Transversely Isotropic Magnetoelectroelastic Full Space

6.7.3.1 Due to the Point Source in K-direction (K = 3,4,5)

4  1  1 1  
3 ∑ Ai  − l12  3
K = −sgn(l )
G11 K + 2 2 
i =1  ri (ri + si l3 )  ri (ri + si l3 ) ri (ri + si l3 )  
4  l1l2 l1l2 
3 ∑ Ai  3
K = sgn(l )
G12 K + 2 (A18)
r ( r + s l ) r ( r + s l ) 2
i =1 i i i 3 i i i 3 
4 AK s l
∑ ir 3 1
K = i
G13
i =1 i

where

Ri
ri = = l12 + l22 + si2 l32 = sin 2 θ + si2 cos2 θ (A19)
R

K = GK
G21 12
4  1  1 1  
3 ∑ Ai  − l22  3
K = −sgn(l ) K + 2
G22 2   (A20)
i =1  ri (ri + si l3 )  ri (ri + si l3 ) ri (ri + si l3 )  
4 AK s l
∑ ir 3
K = i 2
G23
i =1 i

4
AiK βi, J − 2 l1
GJK1 = − ∑
i =1 ri3
4
AiK βi, J − 2 l2
GJK2 = − ∑ ( J = 3, 4, 5) (A21)
i =1 ri3
4 AiK βi, J − 2 si2 l3
GJK3 = − ∑
i =1 ri3

6.7.3.2 Due to the Point Source in x-direction

1 =B
 l22 l1 2l22 l1  B0 l1
G11 0 3 2
+ 2 3 −
 r0 (r0 + s0 l3 ) r0 (r0 + s0 l3 )  r0 (r0 + s0 l3 )2
4  3l1 l13 2l13 
+ ∑ Bi  2
− 3 2
− 2 3 
i =1  ri (ri + si l3 ) ri (ri + si l3 ) ri (ri + si l3 ) 

1 =B
 l23 2l23  3B0 l2
G12 0 3 2
+ 2 −
 r0 (r0 + s0 l3 ) r0 (r0 + s0 l3 )  r0 (r0 + s0 l3 )2
3
(A22)
4  l2 l2 l12 2l2 l12 
+ ∑ Bi  − 3 − 
i =1  ri (ri + si l3 )
2 ri (ri + si l3 )2 ri2 (ri + si l3 )3 

1 = sgn(l )B s −
 1 l22 l22 
G13 3 0 0  + 3
+ 2 2 
 r0 (r0 + s0 l3 ) r0 (r0 + s0 l3 ) r0 (r0 + s0 l3 ) 
4  1 l12 l12 
+ sgn(l3 )∑ Bi si  − 3 − 2 2 
i =1  ri (ri + si l3 ) ri (ri + si l3 ) ri (ri + si l3 ) 
6.7 Appendix A: The Extended Green’s Functions and Their Derivatives 215

4  l2 l2 l12 2l2 l12 


1 =
G21 ∑ i r ( r + s l ) 2 r 3 ( r + s l ) 2 r 2 ( r + s l )3 
B  − −
i =0 i i i 3 i i i 3 i i i 3 
4  l l l 2 2 l l 2 
∑ Bi  r (r + s1 l )2 − r 3 (r + s2 l )2 − r 2 (r + s2 l )3 
1 = 1 1
G22 (A23)
i =0 i i i 3 i i i 3 i i i 3 
4  l l l l 
3 ∑ Bi si  3
1 = −sgn(l ) 12 12
G23 + 2 
i =0 ir (ri + si l 3 ) ri (ri + si l3 )2 

4  
−1 l12 l12
GJ11 = sgn(l3 )∑ Bi βi, J − 2  + 3 + 2 2
i =1  ri (ri + si l3 ) ri (ri + si l3 ) ri (ri + si l3 ) 
4  
ll l1l2
GJ1 2 = sgn(l3 )∑ Bi βi, J − 2  3 1 2 + 2 2
( J = 3, 4, 5)
i =1  ri (ri + si l3 ) ri (ri + si l3 ) 
4
Bi βi, J − 2 si l1
GJ1 3 = ∑
i =1 ri3
 (A24)

6.7.3.3 Due to the Point Source in y-direction

G12j = G21 j ( j = 1, 2, 3)  (A25)

4  l1 l1l22 2l1l22 
2 =
G21 ∑ r ( r + s l ) 2 r 3 ( r + s l ) 2 r 2 (r + s l ) 3 
Bi  − −
i =1 i i i 3 i i i 3 i i i 3 
 −3l1 l13 2l13 
+ B0  2
+ 3 2
+ 2 3 
 r0 (r0 + s0 l3 ) r0 (r0 + s0 l3 ) r0 (r0 + s0 l3 ) 
4  3l2 l23 2l23 
2 =
G22 ∑ r (r + s l ) 2 r 3 (r + s l ) 2 r 2 ( r + s l ) 3 
Bi  − −
i =1 i i i 3 i i i 3 i i i 3  (A26)
 − l2 l12 l2 2l12 l2 
+ B0  2
+ 3 2
+ 2 3 
 r0 (r0 + s0 l3 ) r0 (r0 + s0 l3 ) r0 (r0 + s0 l3 ) 
4  1 l22 l22 
2
G23 = sgn(l3 )∑ Bi si  − 3 − 2 2 
i =1  ri (ri + si l3 ) ri (ri + si l3 ) ri (ri + si l3 ) 
 −1 l12 l12 
+ sgn(l3 )B0 s0  + 3 + 2 2 
 r0 (r0 + s0 l3 ) r0 (r0 + s0 l3 ) r0 (r0 + s0 l3 ) 

GJ21 = GJ1 2
4  
−1 l22 l22
GJ22 = sgn(l3 )∑ Bi βi, J − 2  + 3 + 2 2  ( J = 3, 4, 5)
i =1  ri (ri + si l3 ) ri (ri + si l3 ) ri (ri + si l3 ) 
4
Bi βi, J − 2 si l2
GJ23 = ∑
i= 1 ri3
 (A27)
216 Green’s Functions in a Transversely Isotropic Magnetoelectroelastic Full Space

6.8 Appendix B: Functions Involved in the Eshelby Inclusion Problem

6.8.1 A Spheroidal Inclusion (b = a/c)

1 (1 − t 2 )
H1 (b, si ) = 2 ∫ dt (B1)
0 [1 + (b2 − 1)t 2 ][1 + ( s 2 − 1)t 2 ]3 / 2
i

1 t2
H 2 (b, si ) = 2b2 ∫ dt (B2)
0 [1 + (b2 − 1)t 2 ][1 + ( si2 − 1)t 2 ]3/ 2

1 td t
H 3 (b, si ) = 2b2 ∫
0
[1 + (b2 − 1)t 2 ] 1 + ( si2 − 1)t 2 [ 1 + ( si2 − 1)t 2 + si t ]
1 t(1 − t 2 )dt
H 4 (b, si ) = 2b2 ∫ (B3)
0
[1 + (b2 − 1)t 2 ][1 + ( si2 − 1)t 2 ]3/ 2 [ 1 + ( si2 − 1)t 2 + si t ]
1 t(1 − t 2 )dt
H 5 (b, si ) = 2b2 ∫
0
[1 + (b2 − 1)t 2 ][1 + (si2 − 1)t 2 ][ 1 + (si2 − 1)t 2 + si t ]2

1 (1 − t 2 )dt
H6 (b, si ) = 2 ∫
0
[1 + (b2 − 1)t 2 ] 1 + ( si2 − 1)t 2 [ 1 + ( si2 − 1)t 2 + si t ]2
1 (1 − t 2 )2 dt
H 7 (b, si ) = 2 ∫ (B4)
0
[1 + (b2 − 1)t 2 ][1 + ( si2 − 1)t 2 ]3/ 2 [ 1 + ( si2 − 1)t 2 + si t]2
1 (1 − t 2 )2 dt
H8 (b, si ) = 2 ∫
0
[1 + (b2 − 1)t 2 ][1 + (si2 − 1)t 2 ][ 1 + ( si2 − 1)t 2 + si t ]3

From the expressions for the tensor PKpJi given in Eqs. (6.131)–(6.164), it is observed
that the elements of this tensor are actually related to the following four functions
only: H1, H2, F1 ≡ −2H3+H4+H5, and F2 ≡ −4H6+H7+2H8. Thus, we only need to find
the analytical integrations for these four functions, which are given as follows.

 s b2 tan −1 b2 / si2 − 1 
H1 (b, si ) = 2  2 i 2 +  (B5)
 si − b (b2 − si2 )3 / 2 
 

 1 tan −1 b2 / si2 − 1 
H 2 (b, si ) = −2b2  +  (B6)
 si ( si2 − b2 ) (b2 − si2 )3/ 2 
 

 1 si tan −1 b2 / si2 − 1 
F1 (b, si ) = 2b2  2 +  (B7)
 si − b2 (b2 − si2 )3/ 2 
 
6.8 Appendix B: Functions Involved in the Eshelby Inclusion Problem 217

 s b2 tan −1 b2 / si2 − 1 
F2 (b, si ) = −2  2 i 2 +  (B8)
 si − b (b2 − si2 )3/ 2 
 

6.8.2 Three Special Geometric Cases of Inclusion (b = a/c)


(1) b = 1 (spherical case: a = c)

 s tan −1 1 / si2 − 1 
H1 ( si ) = 2  2 i +  (B9)
 si − 1 (1 − si2 )3/ 2 
 
 1 tan −1 1 / si − 1 
2
H 2 ( si ) = −2  + 
 si ( si2 − 1) (1 − si2 )3/ 2 
  (B10)

 1 si tan −1 1 / si2 − 1 
F1 (si ) = 2  2 + 
 si − 1 (1 − si2 )3/ 2 
  (B11)

 s tan −1 1 / si2 − 1 
F2 (si ) = −2  2 i +  (B12)
 si − 1 (1 − si2 )3/ 2 
 

(2) b = 0 (infinite circular cylinder, with a being fixed but c approaches


infinity):
H1 ( si ) = 2 / si (B13)

H 2 ( si ) = 0 (B14)

F1 (si ) = 0 (B15)

F2 (si ) = −2 / si (B16)

(3) b = ∞ (disklike inclusion, with a being fixed but c approaches zero):


H1 ( si ) = 0 (B17)

H 2 (si ) = 2 / si (B18)

F1 (si ) = 0 (B19)

F2 ( si ) = 0 (B20)
218 Green’s Functions in a Transversely Isotropic Magnetoelectroelastic Full Space

6.9 References
Chen, W. Q., Lee, K.Y., and Ding, H. J. 2004. General solution for transversely isotropic
magneto-electro-thermo-elasticity and the potential theory method. International
Journal of Engineering Science 42: 1361–79.
Ding, H. J. and Chen, W. Q. 2001. Three Dimensional Problems of Piezoelasticity. New York:
Nova Science Publishers.
Ding, H. J. and Jiang, A. M. 2004a. A boundary integral formulation and solution for
2D problems in magneto-electro-elastic media. Computers & Structures 82 (20–1):
1599–607.
2004b. Fundamental solutions for transversely isotropic magneto-electro-elastic media
and boundary integral formulation. Science in China E, 46 (6): 607–19.
Ding, H. J., Chen, B., and Liang, J. 1996a. General solutions for coupled equations for
piezoelectric media. International Journal of Solids and Structures 33: 2283–98.
Ding, H. J., Liang, J., and Chen, B. 1996b. Fundamental solutions for transversely isotropic
piezoelectric media. Science in China A 39 (7): 766–75.
1997. The united point force solution for both isotropic and transversely isotropic.
Communications in Numerical Methods in Engineering 13 (2): 95–102.
Dunn, M. L. and Wienecke, H. A. 1996a. Green’s functions for transversely isotropic
piezoelectric solids. International Journal of Solids and Structures 33: 4571–81.
1997. Inclusions and inhomogeneities in transversely isotropic piezoelectric solids.
International Journal of Solids and Structures 27: 3571–82.
Elliot, H. A. 1948. Three-dimensional stress distributions in hexagonal aeolotropic crystals.
Proceedings of Cambridge Philosophical Society 44: 522–33.
Eshelby, J. D. 1957. The determination of the elastic field of an ellipsoidal inclusion and
related problems. Proceedings of the Royal Society of London A 241: 376–96.
Fabrikant, V. I. 1989. Applications of Potential Theory in Mechanics: A Selection of New
Results. Dordrecht: Kluwer.
2003. A new form of the Green function for a transversely isotropic body. Acta Mechanica
167 (1–2): 101–11.
Hanson, M. T. 1998. Some observations on the potential functions for transverse isotropy
in the presence of body forces. International Journal of Solids and Structures 35:
3793–813.
1999. Elastic fields for point and partial line loading in transversely isotropic linear
elasticity. Journal of Elasticity 55: 143–62.
Hou, P. F., Chen, H. R., and He, S. 2009. Three-dimensional fundamental solution for
transversely isotropic electro-magneto-thermo-elastic materials. Journal of Thermal
Stresses 32 (9): 887–904.
Hou, P. F., Ding, H. J., and Chen, J. Y. 2005. Green’s functions for transversely isotropic
magnetoelectroelastic media. International Journal of Engineering Science 43: 826–58.
Hou, P. F., Leung, A. Y. T., and Ding, H. J. 2003. The elliptical Hertzian contact of transversely
isotropic magnetoelectroelastic bodies. International Journal of Solids and Structures 40
(11): 2833–50.
Hu, H. C. 1956. On the equilibrium and vibration of a transversely isotropic elastic body.
Scientia Sinica 5: 1–18. [First published in Chinese in Acta Physica Sinica 11 (1955):
219–38.]
Karapetian, E., Sevostianov, I., and Kachanov, M. 2000. Point force and point electric charge
in infinite and semi-infinite transversely isotropic piezoelectric solid. Philosophical
Magazine B 80: 331–59.
Li, J. Y. 2000. Magnetoelectroelastic multi-inclusion and inhomogeneity problems and their
applications in composite materials. International Journal of Engineering Science 38:
1993–2011.
Mura, T. 1987. Micromechanics of Defects in Solids. 2nd ed. Dordrecht: Martinus Nijhoff
Publishers.
Pan, E. 1997. A general boundary element analysis of 2-D linear elastic fracture mechanics.
International Journal of Fracture 88: 41–59.
6.9 References 219

2004. Eshelby problem of polygonal inclusions in anisotropic piezoelectric full- and half-
planes. Journal of the Mechanics and Physics of Solids 52: 567–89.
Pan, Y. C. and Chou, T. W. 1976. Point force solution for an infinite transversely isotropic
solid. Journal of Applied Mechanics 43: 608–12.
Wang, X. and Shen, Y. P. 2002. The general solution of three-dimensional problems in
magnetoelectroelastic media. International Journal of Engineering Science 40: 1069–80.
7 Green’s Functions in a Transversely
Isotropic Magnetoelectroelastic
Bimaterial Space

7.0 Introduction
We will extend here the analysis in Chapter 6 to the bimaterial case, in which the
effect of interface should be considered. Just as the infinite full-space case, exact
Green’s solutions can be obtained when the bimaterial is subjected to a concentrated
force, an electric charge or a magnetic charge at an arbitrary point, provided that
the interface is parallel to the material isotropic plane of both half-spaces. The
derivations are also based on the concise general solution in Eq. (6.18), and are quite
similar to those in Chapter 6. The difference is that, the complete solution can be
divided into two parts: one corresponds to the full space as already presented in
Chapter 6, and the other is complementary to account for the interface effect. The
complementary part can be obtained by generalizing the method of images (Lur’e
1964); the generalization includes the introduction of several measures of distance
between two points, which fully accounts for the material anisotropy as well as
multifield coupling, as embodied by the characteristic roots of Eq. (6.13) for both
phases in the bimaterial system. Different interface conditions are also discussed,
including the half-space as a special case.
The approach described was first proposed by Ding et al. (1997) for deriving
the Green’s functions in a piezoelectric bimaterial space. It was also adopted by
Ding et al. (2005) to obtain the exact Green’s function solutions for an MEE
bimaterial space. The presentation in this chapter follows the work of Ding et al.
(2005) closely, but with a more compact notation. The two-dimensional case
was treated in parallel by Jiang and Ding (2005) for an MEE half-plane and
by Jiang et al. (2006) for a general two-phase MEE plane. Pan and Han (2005)
derived the Green’s functions for a layered system using the propagator matrix
method. Further incorporation of the thermal effect was explored by Hou et al.
(2008; 2009a; 2009b) for various three-dimensional configurations (with surface
or interface effect), and by Xiong and Ni (2012) for the Green’s function in the
corresponding half-plane.

7.1 Problem Description


We consider a bimaterial system where each half-space is made of different transversely
isotropic MEE materials but with their polarization directions both along the
220
7.2 Green’s Functions in a Bimaterial Space Due to Extended Point Sources 221

z
Material 1

S(0,0,h)

y
o
Interface
x

Material 2

Figure 7.1. An MEE bimaterial system consisting of two half-spaces. Concentrated sources
are applied at the point S(0,0,h>0) in Material 1. The reader is referred to Figure 6.1 for
further details.

z-direction. The Cartesian coordinate system (x,y,z) is attached to the bimaterial


system such that the (x,y)-plane coincides with the interface, and the z-axis is normal
to it, see Figure 7.1. We further assume, without loss of generality, that in the bimaterial
domain there is a source (not necessary of unit magnitude at this stage) applied on the
z-axis at (0,0,h > 0) in the upper half-space (z > 0). This point source can be a vertical
or horizontal point force, a negative electric charge or negative magnetic charge. The
upper half-space with the source is further named as Material 1, and the lower half-
space of source-free is named Material 2. Material properties and related parameters,
and solutions in Materials 1 and 2 are distinguished from each other by attaching
superscripts “(1)” and “(2)” to these quantities. However, for simplicity, the superscript
“(1)” will be omitted for Material 1 when there is no confusion. The interface (z = 0) is
assumed to be perfect. In other words, the extended displacement and traction vectors
are all continuous across the interface. Our goal is, therefore, to find the solutions of the
Green’s functions in both half-spaces. We point out that solutions to the corresponding
smooth interface case are also derived for future reference.

7.2 Green’s Functions in a Bimaterial Space Due to Extended


Point Sources

7.2.1 Solutions for a Vertical Point Force, a Negative Electric Charge,


or a Negative Magnetic Charge
We assume that at point (x,y,z) = (0,0,h > 0) in the upper half-space, there is an
axisymmetric point source along the z-axis, that is, either a vertical point force
fz(x,y,z) = fzδ(x)δ(y)δ(z−h), or a negative electric charge −fe(x,y,z) = feδ(x)δ(y)δ(z−h),
or a negative magnetic charge −fh(x,y,z) = fhδ(x)δ(y)δ(z−h). Because these sources
are either along the axis of symmetry (vertical point force) or scalar (the electric or
magnetic charge), we assume the potential functions of the complementary part of
222 Green’s Functions in a TI MEE Bimaterial Space

the solution, which should be added to the infinite-space Green’s functions derived
in Chapter€6, in the upper half-space (Material€1)€as

4
Ψ c0 = 0, Ψ ci = ∑ Aij ln Rij* (i = 1, 2, 3, 4) (7.1)
j =1

where the superscript (c) is attached to signify the complementary part of the
solution,€and (i,j = 0–4)

Rij* = Rij + zij , Rij = x 2 + y2 + zij2 , zij = si z + s j h (7.2)

In Eq. (7.2), s0 = c66 / c44 and si (i = 1–4) are the four eigenroots (with positive
real part) of Eq. (6.13). It is easily seen that when i = j, we have Rij (−z) = Ri (z) , as
defined in Eq. (6.30). Thus, this corresponds to the image source due to the presence
of the interface located at z = 0. The unknown constants Aij (i, j = 1, 2, 3, 4) are to be
determined from the interface conditions at z€=€0.
The first and second derivatives of the potential functions in Eq. (7.1) can be
found€to€be

4 4 4
x y 1
Ψ ci, x = ∑ Aij *
, Ψ c =
i, y ∑ Aij *
, Ψ c =
i ,z ∑ Aij si (7.3a)
j =1 Rij Rij j =1 Rij Rij j= 1 Rij

4 4
x y
Ψ ci, xz = − ∑ Aij si 3
, Ψ c
i , yz = − ∑ Aij si 3
j =1 Rij j =1 Rij
(7.3b)
4  1 1 
Ψ ci, xy = − ∑ Aij xy  3 * + 2 * 2 
j =1  Rij Rij Rij Rij 

4  1  1 1 
Ψ ic, xx = ∑ Aij  − x 2 + 
 Rij Rij*  Rij3 Rij* Rij2 Rij* 2  
j =1  
4  1  1 1  
Ψ ic, yy = ∑ Aij  − y 2 +  (7.3c)
 Rij Rij*  Rij3 Rij* Rij2 Rij* 2  
j =1  
4 z
Ψ ic,zz = − ∑ Aij si2 3
ij

j =1 Rij

Thus, according to Eqs. (6.18) and (6.22), the complementary part of the extended
displacement solution can be obtained€as

4 4
( x + i y)
U c = uxc + i uyc = − ∑ ∑ Aij
i =1 j =1 Rij Rij*
4 4
(7.4)
1
wkc = ∑ ∑ Aij βik (k = 1, 2, 3)
i =1 j =1 Rij
7.2 Green’s Functions in a Bimaterial Space Due to Extended Point Sources 223

where βik are parameters related to the material coefficients and they are defined
in Chapter 6 (in the expressions following Eq. (6.18)). Note also that, following
Chapter 6, wkc (k = 1,2,3) in the second expression of Eq. (7.4), correspond,
respectively, to the vertical displacement component uzc , the electric potential φc, and
the magnetic potential ψ c, see the definition following Eq. (6.18).
The complementary part of the stresses is

4 4  1 1 
σ c2 = 2c66 ∑ ∑ Aij ( x + i y)2  3 * + 2 * 2  (7.5a)
i =1 j =1  Rij Rij Rij Rij 

4 4 zij
σ czk = − ∑ ∑ Aij γ ik (k = 1, 2, 3, 4) (7.5b)
i =1 j =1 Rij3

4 4
( x + i y)
τ czk = − ∑ ∑ Aij ϖik (k = 1, 2, 3) (7.5c)
i =1 j =1 Rij3

The formulations (7.4) and (7.5) are very similar to those in Eqs. (6.34)–(6.38), with
notations following directly from Eq. (6.21) and σ z4 = σ1. However, it should be
noted that the extended complementary displacements and stresses are regular in
the entire upper half-space. The complete expressions for the Green’s displacements
and stresses in the upper half-space are the sum of the complementary regular part in
Eqs. (7.4) and (7.5), and the singular part in Eqs. (6.34)–(6.38). The complete solutions
will be labeled by the superscript “(1),” for example, U (1), wk(1), σ zk
(1)
, and so forth.
For the lower half-space (Material 2) where it is free from any point source, we
have only the potential functions corresponding to the complementary part of the
solution. In other words, the potential functions in the lower half-space (Material 2)
can be assumed as

4
Ψ (02) = 0, Ψ (i 2) = ∑ Aij( 2) ln rij* (i = 1, 2, 3, 4) (7.6)
j =1

where, as mentioned earlier, we have used the superscript “(2)’” to distinguish the
lower half-space from the upper half-space, and

rij = x 2 + y2 + (zij( 2) )2 , rij* = rij − zij( 2) , zij( 2) = si( 2)z − s j h (7.7)

The corresponding derivatives are

4 4 4
x y 1
Ψ(i,2x) = ∑ Aij( 2) *
( )
, Ψi,2y = ∑ Aij(2) rij rij* , Ψ(i,2z) = − ∑ Aij( 2) si( 2) (7.8a)
j =1 rij rij j =1 j =1 rij

4
x 4
y 4  1 1 
Ψ(i,2xz
)
= ∑ Aij( 2) si( 2) , Ψ(i,2yz
)
= ∑ Aij(2) si(2) , Ψ ( 2)
i, xy = − ∑ Aij( 2) xy  3 * + 2 *2 
j =1 rij3 j =1
3
rij j =1  rij rij rij rij 
 (7.8b)
224 Green’s Functions in a TI MEE Bimaterial Space

4  1  1 1 
Ψ i(,2xx) = ∑ Aij( 2)  * − x 2  3 * + 2 *2  
 rij rij  rij rij rij rij  
j =1  
4  1  1 1  
Ψ i(,2yy) = ∑ Aij( 2)  * − y2  3 * + 2 *2   (7.8c)
 rij rij  rij rij rij rij  
j =1  
( 2)
4 zij
Ψ i(,2zz) = ∑ Aij( 2) ( si( 2) )2 3
j =1 rij

Thus, the corresponding extended displacements and stresses are

4 4
( x + i y)
U (2) = − ∑ ∑ Aij(2)
i =1 j =1 rij rij*
4 4
(7.9)
1
wk(2) = −∑ ∑ Aij(2)β(ik2) (k = 1, 2, 3)
i =1 j =1 rij

4 4  1 1 
σ (22) = 2c66
( 2)
∑ ∑ Aij(2) ( x + i y)2  r 3r* + r 2 r*2  (7.10a)
i= 1 j = 1  ij ij ij ij 

4 4 (2)
z
= ∑∑ ( 2 ) ( 2 ) ij
σ (zk2) Aij γ ik 3 (k = 1, 2, 3, 4) (7.10b)
i =1 j =1 rij

4 4
( x + i y)
τ (zk2) = ∑ ∑ Aij(2) ϖ(ik2) (k = 1, 2, 3) (7.10c)
i =1 j =1 rij3

Let us now consider the perfect bonding case, that is, the two half-spaces are perfectly
bonded to each other at the interface. Then, the continuity conditions of the extended
displacements and tractions at z = 0 can be expressed as:

U (1) = U (2) , wk(1) = wk(2) , σ(zk


1)
= σ(zk2) , τ(z11) = τ(z21) (k = 1, 2, 3) (7.11)

from which the following thirty-two algebraic equations can be derived:

4 4
Ai − ∑ Aji = − ∑ A(ji2) (7.12a)
j =1 j =1

4 4
β ik Ai + ∑ β jk Aji = − ∑ β jk
( 2) ( 2)
Aji (7.12b)
j =1 j =1

4 4
ϖ i 1 Ai + ∑ ϖ j 1 Aji = − ∑ ϖ (j 12) A(ji2) (7.12c)
j =1 j =1
7.2 Green’s Functions in a Bimaterial Space Due to Extended Point Sources 225

4 4
γ ik Ai − ∑ γ jk Aji = − ∑ γ (jk2) A(ji2) (7.12d)
j =1 j =1

where i = 1,2,3,4, k = 1,2,3, and use has been made of the following identities:

Ri = R ji = rji , Ri* = R*ji = rji* , zji = −z(ji2) = si h (z = 0 ) (7.13)

Then, the thirty-two unknowns Aij and Aij( 2 ) can be uniquely determined from
Eq. (7.12), where Ai are given by Eq. (6.43). For the Green’s function solutions, we
should let (fz, −fe, −fh) in Eq. (6.43) equal (1,0,0), (0,1,0) and (0,0,1), respectively.

7.2.2 Solutions for a Horizontal Point Force


When the point force is along the x-direction (fx), the complementary solution in the
upper half-space (z>0) can be assumed to be

4
y x
Ψ c0 = B00 *
R00
, Ψ ci = ∑ Bij R* (i = 1, 2, 3, 4) (7.14)
j =1 ij

with Rij* defined in Eq. (7.2). The following derivatives can be obtained

xy  1 y2  y
Ψc0, x = − B00 * 2
, Ψ0c , y = B00  * − * 2
, Ψc0,z = −B00 s0 *
R00 R00  R00 R00 R00  R00 R00

(7.15a)

 1 1 
Ψ 0c, xz = B00 s0 xy  2 * 2 + 3 * 
R R
 00 00 R 00 R00 
 1  1 1 
Ψ 0c, yz = B00 s0  − *
+ y2  2 * 2 + 3 *
   (7.15b)
 R00 R00  R00 R00 R00 R00  
 1  1 2 
Ψ 0c, xy = B00 x  − + y2  3 * 2

+ 2 * 3 
R R
 00 00
* 2
 00 R00
R R00 R00  

 1  1 2 
Ψ 0c, xx = B00 y  − + x2  3 * 2 + 2 * 3 

*
 R00 R00
2
 R00 R00 R00 R00  
 3  1 2 
Ψ 0c, yy = B00 y  − + y2  3 * 2 + 2 * 3  

(7.15c)
 R00 R00
* 2
 R00 R00 R00 R00  
y
Ψ 0c,zz = B00 s02 3
R00

4  1 x2  4
xy 4
x
Ψci, x = ∑ Bij  * − , Ψci, y = − ∑ Bij , Ψ ci,z = − ∑ Bij si
j =1  Rij Rij Rij* 2  j =1 Rij Rij* 2 j =1 Rij Rij*
 (7.16a)
226 Green’s Functions in a TI MEE Bimaterial Space

4  1  1 1 
Ψ ic, xz = ∑ Bij si  − + x 2 + 
 Rij Rij*  Rij2 Rij* 2 Rij3 Rij*  
j =1  
4  1 1 
Ψ ic, yz = ∑ Bij si xy  2 * 2 + 3 *  (7.16b)
 Rij Rij 
j =1  Rij Rij
4  1  1 2 
Ψ ic,xy = ∑ Bij y  − + x 2 + 
 Rij Rij* 2  Rij3 Rij* 2 Rij2 Rij* 3  
j =1  

4  3  1 2 
Ψ ic, xx = ∑ Bij x  − + x 2 + 
 Rij Rij* 2  Rij3 Rij* 2 Rij2 Rij* 3  
j =1  
4  1  1 2 
Ψ ic, yy = ∑ Bij x  − + y2  3 * 2 + 2 * 3   (7.16c)
 Rij Rij* 2  Rij Rij Rij Rij  
j =1  
4
x
Ψ ic,zz = ∑ Bij si2 3
j =1 R ij

Then, according to Eqs. (6.18) and (6.22), the corresponding field quantities are

4 4  1 x( x + i y)   1 i y( x + i y) 
U c = − ∑ ∑ Bij  * − * 2  + B00  * + * 2 
i=1 j =1  Rij Rij Rij   R00 R00 R00 
4 4
(7.17)
x
wkc = − ∑ ∑ Bij βik (k = 1, 2, 3)
i =1 j =1 Rij Rij*

4 4  2( x + i y)  1 2 
σ c2 = −2c66 ∑ ∑ Bij  − + x ( x + i y) 2
 3 *2
+ 
 Rij Rij* 2  Rij Rij Rij2 Rij* 3  
i =1 j =1  (7.18a)
 2( x + i y)  1 2 
− 2c66 B00  * 2
+ i y( x + i y)2  3 * 2 + 2 * 3  
 R00 R00  R00 R00 R00 R00  

4 4
x
σ czk = ∑ ∑ Bij γ ik (k = 1, 2, 3, 4) (7.18b)
i =1 j =1 Rij3

4 4  1  1 1 
τ czk = ∑ ∑ Bij ϖik  − + x ( x + i y )  + 
 Rij Rij* 2 *2
 Rij Rij Rij3 Rij*  
i=1 j =1 
(7.18c)
 1  1 1  
− B00 s0 νk  *
+ i y( x + i y )  2 R* 2
+ 3 R*   (k = 1, 2, 3)
 R00 R00  R00 00 R00 00   

For the same horizontal force fx, the complementary part of the solutions in the
lower half-space (z < 0, Material 2) can be obtained by assuming
7.2 Green’s Functions in a Bimaterial Space Due to Extended Point Sources 227

4
y x
Ψ (02) = B00
( 2)
*
r00
, Ψ (i 2) = ∑ Bij(2) r* (i = 1, 2, 3, 4) (7.19)
j =1 ij

where rij* are defined in Eq. (7.7). It can be shown that

xy ( 2)  1 y2  y
Ψ(02, x) = − B00
( 2)
*2
, Ψ(02, y) = B00 −
 r * r r * 2  , Ψ(02,z) = B00
( 2) ( 2)
s0 * (7.20a)
r00 r00 00 00 00 r00 r00

 1 1 
Ψ 0(,2xz
)
= − B00( 2) ( 2)
s0 xy  2 * 2 + 3 * 
 r r00
 00 r00 r00 
 1  1 1 
Ψ 0(,2yz
) ( 2) ( 2)
= B00 s0  − y2  2 * 2 + 3 *  
 r r00
(7.20b)
*
 r00 r00  00 r00 r00  
 1  1 2 
Ψ 0(,2xy
)
= − B00( 2)
x * 2
− y2  3 * 2 + 2 * 3  

 r00 r00  r00 r00 r00 r00  

 1  1 2 
Ψ 0(,2xx
) ( 2)
= B00 y − * 2
+ x2  3 * 2
 r r00
+ 2 r* 3 
 r00 r00  00 r00 00  
 3  1 2 
Ψ 0(,2yy
) ( 2)
= B00 y − * 2
+ y2  3 * 2 + 2 * 3   (7.20c)
 r00 r00  r00 r00 r00 r00  

( 2) ( 2) 2  1 1 
Ψ 0(,2zz
)
= B00 ( s0 ) y  2 * 2 + 3 *
 r r00 
 00 r00 r00 

4  1 x2  4
xy 4
x
Ψ(i,2x) = ∑ Bij( 2)  * − *2  , Ψ(i,2y) = − ∑ Bij( 2) , Ψ(i,2z) = ∑ Bij( 2) si( 2)
j =1  rij rij rij  j =1 rij rij*2 j =1 rij rij*
 (7.21a)

4  1  1 1 
Ψ i(,2xz) = ∑ Bij( 2) si( 2)  * − x 2  2 *2 + 3 *  
 rij rij  rij rij rij rij  
j =1  
4  1 1 
Ψ i(,2yz) = − ∑ Bij( 2) si( 2) xy  2 *2 + 3 *  (7.21b)
 r r r 
ij rij 
j =1  ij ij
4  1  1 2 
Ψ i(,2xy) = ∑ Bij(2) y  − *2 + x 2  3 *2 + 2 *3  
 rij rij  rij rij rij rij  
j =1  

4  3  1 2 
Ψ i(,2xx) = ∑ Bij( 2) x  − *2 + x 2  3 *2 + 2 *3  
 rij rij  rij rij rij rij  
j =1  
4  1  1 2 
Ψ i(,2yy) = ∑ Bij( 2) x  − *2 + y2  3 *2 + 2 *3   (7.21c)
 rij rij  rij rij rij rij  
j =1  
4
( 2) ( 2) 2 x
Ψ i,zz = ∑ Bij ( si ) 3
( 2)

j =1 rij
228 Green’s Functions in a TI MEE Bimaterial Space

Then, the corresponding field quantities are

4 4  1 x( x + i y)  (2)  1 i y( x + i y) 
U (2) = −∑ ∑ Bij(2)  * − *2 
+ B00  r * + r r* 2  (7.22a)
i = 1 j =1 r
 ij r r
ij ij   00 00 00 

4 4
x
wk( 2) = ∑ ∑ Bij( 2)β(ik2) (k = 1, 2, 3) (7.22b)
i =1 j =1 rij rij*

4 4  2( x + i y)  1 2 
σ 2(2) = 2c66
( 2)
∑ ∑ Bij(2)  r r*2 − x(x + i y)2  r 3r*2 + r 2 r*3  
i =1 j =1  ij ij  ij ij ij ij  
(7.23a)

( 2 ) ( 2 ) 2( x + i y)  1 2 
− 2c66 B00  + i y( x + i y)2 +
*2 
 r 3 r00
*2 2 r* 3 
 r00 r00  00 r00 00  

4 4
x
       σ(zk2) = ∑ ∑ Bij( 2) γ (ik2) (k = 1, 2, 3, 4) (7.23b)
i =1 j =1 rij3

4 4  1  1 1 
τ(zk2) = ∑ ∑ Bij(2) ϖ(ik2)  * − x( x + i y)  2 *2 + 3 *  
i=1 j =1  rij rij  rij rij rij rij  

   (7.23c)
(2) (2) (2)
 1  1 1 
+ B00 s0 νk  *
+ i y( x + i y)  2 * 2 + 3 *   (k = 1, 2, 3)
 r00 r00  r00 r00 r00 r00  

The continuity conditions at the perfect interface z = 0 are given in Eq. (7.11), from
which, one can derive the following thirty-four algebraic equations:

( 2)
B0 + B00 = B00 (7.24a)

4 4
Bi + ∑ Bji = ∑ B(ji2) (7.24b)
j =1 j =1

4 4
βik Bi − ∑ β jk Bji = ∑ β(jk2) B(ji2) (7.24c)
j =1 j =1

s0ν 1 B0 − s0ν 1 B00 = s0( 2)ν 1( 2) B00


( 2)
(7.24d)

4 4
ϖ i 1 Bi − ∑ ϖ j 1 Bji = ∑ ϖ (j 12) B(ji2) (7.24e)
j =1 j =1

4 4
γ ik Bi + ∑ γ jk Bji = ∑ γ (jk2) B(ji2) (7.24f)
j =1 j =1
7.2 Green’s Functions in a Bimaterial Space Due to Extended Point Sources 229

where i = 1,2,3,4; k = 1,2,3, and the constants B0 and Bi (i = 1, 2, 3, 4) are given


in Eqs. (6.64) and (6.65). We can solve immediately from Eq. (7.24a) and
(7.24d) that

s0ν 1 − s0( 2)ν 1( 2)


B00 = B0
s0ν 1 + s0( 2)ν 1( 2)
(7.25)
( 2) 2 s0ν 1
B00 = B0
s0ν 1 + s0( 2)ν 1( 2)

where the involved parameters in Eq. (7.25) bear the same definitions as in Chapter 6.
It is seen that if the two half-spaces are made of the same MEE material, we have
( 2)
B00 = 0 and B00 = B0, which is just the case for a homogeneous infinite space. The
remaining thirty-two unknown constants (Bij , Bij( 2)) can be determined from the
remaining thirty-two equations in Eqs. (7.24a−f). For the Green’s function solutions,
we should let fx = 1 in Eqs. (6.64) and (6.65).
Remark 7.1: Just like the infinite space case, if the horizontal point force applied
at (0,0,h > 0) is along the y-direction, that is, we have fy, instead of fx, then by just
replacing x with y, and y with −x in the potential function expressions Eqs. (7.14) and
(7.19), one finds the solutions to the horizontal point force in the y-direction. This
can be easily understood from a 90°-rotation of the coordinates along the z-axis,
along with the isotropy symmetry of the material in the (x, y)-plane. Consequently,
the final expressions of the extended Green’s displacements and stresses due to the
y-direction force can be obtained from those due to the x-direction force by simply
switching x and y on both sides of the expressions of the Green’s displacements and
stresses. These statements also apply to the decoupled cases to be discussed in this
chapter.
Remark 7.2: We here emphasize that all the formulations presented in this chapter
are only valid for the case when the eigenroots si are distinct from each other.
For equal eigenroots, the reader is referred to Ding et al. (1997) for piezoelectric
materials. Or, one could directly perturb the material properties slightly to make
the eigenroots distinguish from each other so that the present formulations can be
directly used (Pan 1997).
Remark 7.3: If the two half-spaces are identical, the material coefficients in Material
1 are the same as those in Material 2. Then, the involved coefficients have the
following simple expressions

Aij = 0 (for i, j = 1, 2, 3, 4)
(2)
Aij =0 (for i, j = 1, 2, 3, 4 but i ≠ j ) (7.26)
(2)
Aii = − Ai (for i = 1, 2, 3, 4)

This leads to the solutions of the MEE field in the infinite space, which was considered
in Chapter 6. It can be verified that this solution satisfies Eq. (7.12) exactly. We also
point out that a similar conclusion can be made for the horizontal force case.
230 Green’s Functions in a TI MEE Bimaterial Space

Remark 7.4: If the source is located in the lower half-space (i.e., h < 0 in Material 2),
then one only needs to switch all the material indices between “1” and “2.” This is
different from the 3D anisotropic bimaterial case (see Chapter 9), where one cannot
simply switch the material domains when the source location is switched.

Remark 7.5: Solutions for different interface conditions can be also obtained easily.
This will be elaborated in Section 7.4.

Remark 7.6: The solutions for the corresponding half-space case can be derived as
the limiting case of the present bimaterial solution; and also for those corresponding
to different boundary conditions on the surface of the half-space. This will be shown
later in Section 7.5.

Remarks 7.5 and 7.6 are similar to those for the anisotropic half-space with
general homogeneous conditions and the anisotropic bimaterial space with general
homogeneous interface conditions.

7.3 Reduced Bimaterial Spaces


The Green’s solutions for an MEE bimaterial space derived in the last section can
be readily reduced to those for materials with less field couplings. In this section,
we consider two reducible cases, which are typical and have also been considered
in Chapter 6. One is the piezoelectric material case, in which the piezomagnetic
effect is absent, and the other is the elastic material case, in which only the elastic
field is taken into consideration. As a prerequisite, in both cases we require that
the axis of material symmetry for transverse isotropy or the polarization is along
the z-axis, or the material has a higher order of symmetry (including isotropy).

7.3.1 Green’s Solutions for Piezoelectric Bimaterial Space


In this case, the general solution is expressed in terms of four potential functions
as given by Eqs. (6.72) and (6.73). The Green’s function solutions for an infinite
transversely isotropic piezoelectric space have been derived in Section 6.4.1.1 when
a vertical point force or a negative electric charge is applied at (0,0,h > 0), and in
Section 6.4.1.2 when a horizontal point force along the x-direction is applied at
(0,0,h > 0). Thus one only needs to consider the corresponding complementary parts
as presented in the following text.

7.3.1.1 Solutions for a Vertical Point Force and a Negative Electric Charge
For a piezoelectric bimaterial space with a perfect interface at z = 0, the
complementary part of the Green’s function solutions in the upper (source) half-
space (Material 1) is

3 3 3 3
( x + i y) 1
U c = − ∑ ∑ Aij , wkc = ∑ ∑ Aij βik (k = 1, 2) (7.27)
i=1 j =1 Rij Rij* i=1 j =1 Rij
7.3 Reduced Bimaterial Spaces 231

3 3  1 1 
σ 2c = 2c66 ∑ ∑ Aij ( x + i y)2  3 * + 2 * 2  (7.28a)
 R R Rij Rij

i =1 j =1  ij ij 

3 3 zij
σczk = − ∑ ∑ Aij γ ik (k = 1, 2, 4) (7.28b)
i =1 j =1 Rij3

3 3
( x + i y)
τ czk = − ∑ ∑ Aij ϖik (k = 1, 2) (7.28c)
i =1 j =1 Rij3

where the material parameters β ik , γ ik and ϖ ik are defined in Eqs. (6.69) and (6.74).
Equations (7.27) and (7.28) are obtained by assuming the potential functions
Ψ i (i = 0, 1, 2, 3) in the same form as Eq. (7.1), except that the summation over j is
from 1 to 3, instead of 1 to 4.
The corresponding Green’s function solutions in the lower half-space (Material
2) are

3 3 3 3
( x + i y) 1
U (2) = − ∑ ∑ Aij(2) , wk
(2)
= − ∑ ∑ Aij(2)β(ik2) rij (k = 1, 2) (7.29)
i=1 j =1 rij rij* i =1 j =1

3 3  1 1 
σ 2( 2) = 2c66
( 2)
∑ ∑ Aij(2) ( x + i y)2  r 3r* + r 2 r*2  (7.30a)
i =1 j = 1  ij ij ij ij 

3 3 (2)
z
= ∑∑ ( 2 ) ( 2 ) ij
σ(zk2) Aij γ ik 3 (k = 1, 2, 4) (7.30b)
i =1 j =1 rij

3 3
( x + i y)
τ(zk2) = ∑ ∑ Aij( 2) ϖ(ik2) (k = 1, 2) (7.30c)
i =1 j =1 rij3

Then, from the continuity conditions in Eq. (7.11), we have

3 3
Ai − ∑ Aji = − ∑ A(ji2) (7.31a)
j =1 j =1

3 3
β ik Ai + ∑ β jk Aji = − ∑ β jk
( 2) ( 2)
Aji (7.31b)
j =1 j =1

3 3
ϖ i 1 Ai + ∑ ϖ j 1 Aji = − ∑ ϖ (j 12) A(ji2) (7.31c)
j =1 j =1

3 3
γ ik Ai − ∑ γ jk Aji = − ∑ γ (jk2) A(ji2) (7.31d)
j =1 j =1
232 Green’s Functions in a TI MEE Bimaterial Space

where now i = 1,2,3; k = 1,2, and the constants Ai (i = 1, 2, 3) are given in Eq. (6.80).
The preceding eighteen algebraic equations then completely determine the eighteen
unknowns Aij and Aij( 2) (i, j = 1, 2, 3) .

7.3.1.2 Solutions for a Horizontal Point Force


The complementary part of the Green’s function solutions in the upper half-space
(Material 1) can be similarly obtained as

3 3  1 x( x + i y)   1 i y( x + i y) 
U c = − ∑ ∑ Bij  * − *2 
+ B00  * + * 2 
i =1 j =1 R
 ij R R
ij ij  R
 00 R00 R00 
3 3
(7.32)
x
wkc = − ∑ ∑ Bij βik (k = 1, 2)
i =1 j =1 Rij Rij*

3 3  2( x + i y)  1 2 
σ 2c = −2c66 ∑ ∑ Bij  − + x ( x + i y) 2 + 
 Rij Rij* 2  Rij3 Rij* 2 Rij2 Rij* 3  
i =1 j =1   (7.33a)
 2( x + i y)  1 2  
− 2c66 B00  * 2
+ i y( x + i y)2  3 * 2 + 2 * 3  
R
 00 00R R R
 00 00 R00 R00   

3 3
x
   σ czk = ∑ ∑ Bij γ ik (k = 1, 2, 4) (7.33b)
i =1 j =1 Rij3

3 3  1  1 1 
τ czk = ∑ ∑ Bij ϖik  − + x ( x + i y)  + 
 Rij Rij* 2 *2
 Rij Rij Rij3 Rij*  
   i=1 j =1  (7.33c)
 1  1 1 
− B00 s0 νk  *
+ i y( x + i y)  2 * 2 + 3 *   (k = 1, 2)
 R00 R00  R00 R00 R00 R00  

The corresponding Green’s function solutions in the lower half-space (Material


2) are

3 3  1 x( x + i y)  (2)  1 i y( x + i y) 
U (2) = −∑ ∑ Bij(2)  * − * 2  + B00  * + *2 
(7.34a)
i = 1 j =1  rij rij rij   r00 r00 r00 

3 3
x
wk( 2) = ∑ ∑ Bij( 2)β(ik2) (k = 1, 2) (7.34b)
i =1 j =1 rij rij*

3 3  2( x + i y)  1 2 
σ 2(2) = 2c66
( 2)
∑ ∑ Bij(2)  r r*2 − x(x + i y)2  r 3r*2 + r 2 r*3  
i =1 j =1  ij ij  ij ij ij ij  
(7.35a)

( 2 ) ( 2 ) 2( x + i y)  1 2 
− 2c66 B00  + i y( x + i y)2  3 * 2 + 2 * 3  
* 2  r r00 r00 r00  
 r00 r00  00
7.3 Reduced Bimaterial Spaces 233

3 3
x
σ (zk2) = ∑ ∑ Bij(2) γ (ik2) (k = 1, 2, 4) (7.35b)
i =1 j =1 rij3

3 3  1  1 1 
τ(zk2) = ∑ ∑ Bij( 2) ϖ(ik2)  * − x( x + i y)  2 *2 + 3 *  
i=1 j =1  rij rij  rij rij rij rij  
 (7.35c)
(2) (2) (2)  1  1 1 
+ B00 s0 νk  *
+ i y( x + i y)  2 * 2 + 3 *   (k = 1, 2)
 r00 r00  r00 r00 r00 r00  

Then, from the continuity conditions in Eq. (7.11), we have

( 2)
B0 + B00 = B00 (7.36a)

3 3
Bi + ∑ Bji = ∑ B(ji2) (7.36b)
j =1 j =1

3 3
β ik Bi − ∑ β jk Bji = ∑ β(jk2) B(ji2) (7.36c)
j =1 j =1

s0ν 1 B0 − s0ν 1 B00 = s0( 2)ν 1( 2) B00


( 2)
(7.36d)

3 3
ϖ i 1 Bi − ∑ ϖ j 1 Bji = ∑ ϖ (j 12) B(ji2) (7.36e)
j =1 j =1

3 3
γ ik Bi + ∑ γ jk Bji = ∑ γ (jk2) B(ji2) (7.36f)
j =1 j =1

where i = 1,2,3; k = 1,2, and the constants Bi (i = 0, 1, 2, 3) are given in Eqs. (6.86)
( 2)
and (6.87). We point out that B00 and B00 are still given by Eq. (7.25), while the
remaining eighteen unknown constants Bij , Bij( 2) (i, j = 1, 2, 3) are determined from
the eighteen equations in Eqs. (7.36b), (7.36c), (7.36e), and (7.36f). The interesting
( 2)
feature that B00 and B00 are the same as those for MEE materials lies in the fact that
the associated deformation is purely elastic. Such a purely elastic and actually shear
deformation (within the plane of isotropy) has been observed in many situations for
materials with transverse isotropy (Pan and Heyliger 2002; Chen et al. 2005; Ding
et al. 2006).
Remark 7.7: For poled ceramics, the polarization direction can be either along
the positive z-axis or the negative z-axis. The two differ only by the sign of the
piezoelectric coefficients, according to the tensor transformation rule.
Remark 7.8: As mentioned in Section 6.4, the corresponding piezomagnetic Green’s
functions can be obtained from the piezoelectric ones by simply replacing the
234 Green’s Functions in a TI MEE Bimaterial Space

piezoelectric/electric coefficients and electric point source by the piezomagnetic/


magnetic ones.

7.3.2 Green’s Solutions for an Elastic Bimaterial Space


The corresponding general solution is expressed in terms of three potential functions
as given by Eqs. (6.92) and (6.93). The Green’s function solutions for an infinite
transversely isotropic elastic space have been derived in Section 6.4.2.1 when a
vertical point force is applied at (0,0,h > 0), and in Section 6.4.2.2 when a horizontal
point force along the x-direction is applied at (0,0,h > 0).

7.3.2.1 Solutions for a Vertical Point Force


The complementary part of the Green’s function solutions in the upper half-space
(Material 1) is

2 2 2 2
( x + i y) 1
U c = − ∑ ∑ Aij , w1c = ∑ ∑ Aij βi 1 (7.37)
i=1 j =1 Rij Rij* i =1 j =1 Rij

2 2  1 1 
σ 2c = 2c66 ∑ ∑ Aij ( x + i y)2  3 * + 2 * 2  (7.38a)
 
i =1 j =1  Rij Rij Rij Rij 

2 2 zij
σczk = − ∑ ∑ Aij γ ik (k = 1, 4) (7.38b)
i =1 j =1 Rij3

2 2
( x + i y)
τ zc1 = −∑ ∑ Aijϖ i 1 (7.38c)
i =1 j =1 Rij3

where the material parameters β i1, γ i1 , γ i4 and ϖ i1 are defined in Eq. (6.94). Equations
(7.37) and (7.38) are simply obtained by taking the potential functions Ψ i (i = 0, 1, 2)
in the same form as Eq. (7.1), with summation over j now taking from 1 to 2 only.
The corresponding Green’s function solutions in the lower half-space (Material
2) are

2 2
( x + i y)
U ( 2) = −∑ ∑ Aij( 2)
i =1 j =1 rij rij*
2 2
(7.39)
1
w1( 2) = −∑ ∑ Aij( 2) β i(12)
i =1 j =1 rij

2 2  1 1 
σ 2(2) = 2c66
( 2)
∑ ∑ Aij(2) ( x + i y)2  r 3r* + r 2 r*2  (7.40a)
i =1 j =1  ij ij ij ij 
7.3 Reduced Bimaterial Spaces 235

2 2 (2)
z
= ∑∑ ( 2 ) ( 2 ) ij
σ (zk2) Aij γ ik 3 (k = 1, 4) (7.40b)
i =1 j =1 rij

2 2
( x + i y)
τ z(12) = ∑ ∑ Aij( 2)ϖ i(12) (7.40c)
i =1 j =1 rij3

Then, from the continuity conditions in Eq. (7.11), we have

2 2
Ai − ∑ Aji = − ∑ A(ji2) (7.41a)
j =1 j =1

2 2
β i 1 Ai + ∑ β j 1 Aji = − ∑ β j(12) A(ji2) (7.41b)
j =1 j =1

2 2
ϖ i 1 Ai + ∑ ϖ j 1 Aji = − ∑ ϖ (j 12) A(ji2) (7.41c)
j =1 j =1

2 2
γ i 1 Ai − ∑ γ j 1 Aji = − ∑ γ (j 12) A(ji2) (7.41d)
j =1 j =1

where i = 1,2, and the constants Ai (i = 1, 2) are given in Eq. (6.101). The preceding
eight algebraic equations completely determine the eight unknowns Aij and
Aij( 2) (i, j = 1, 2).

7.3.2.2 Solutions for a Horizontal Point Force


In this case, the complementary part of the Green’s function solutions in the upper
half-space (Material 1) is

2 2  1 x( x + i y)   1 i y( x + i y) 
U c = −∑ ∑ Bij  * − * 2  + B00  * + * 2 
i =1 j =1  Rij Rij Rij   R00 R00 R00 
(7.42)
2 2
x
w1c = −∑ ∑ Bij β i 1
i =1 j =1 Rij Rij*

2 2  2( x + i y)  1 2 
σ c2 = −2c66 ∑ ∑ Bij  − + x ( x + i y)2
 3 *2
+ 
 Rij Rij* 2  Rij Rij Rij2 Rij* 3  
i =1 j =1  (7.43a)
 2( x + i y)  1 2  
− 2c66 B00  * 2
+ i y( x + i y)2  3 * 2 + 2 * 3  
R
 00 00R R R
 00 00 R00 R00   
236 Green’s Functions in a TI MEE Bimaterial Space

2 2
x
σczk = ∑ ∑ Bij γ ik (k = 1, 4) (7.43b)
i=1 j =1 Rij3

2 2  1  1 1 
τ zc1 = ∑ ∑ Bijϖ i 1  − + x ( x + i y )  + 
 Rij Rij*  Rij2 Rij* 2 Rij3 Rij*  
i =1 j =1   (7.43c)
 1  1 1 
− B00 s0ν k  *
+ i y( x + i y ) 
 R 2 R00 * 2
+ 3 R*  
 R00 R00  00 R00 00  

The corresponding Green’s function solutions in the lower half-space (Material


2) are

2 2  1 x( x + i y)  (2)  1 i y( x + i y) 
U (2) = −∑ ∑ Bij(2)  * − *2 
+ B00  r * + r r* 2  (7.44a)
i = 1 j =1 r
 ij r r
ij ij   00 00 00 

2 2
x
w1( 2) = ∑ ∑ Bij( 2) β i(12) (7.44b)
i =1 j =1 rij rij*

2 2  2( x + i y)  1 2 
σ 2(2) = 2c66
( 2)
∑ ∑ Bij(2)  r r*2 − x(x + i y)2  r 3r*2 + r 2 r*3  
i =1 j =1  ij ij  ij ij ij ij  
(7.45a)
( 2 ) ( 2 )  2( x + i y)  1 2 
− 2c66 B00  + i y( x + i y)2  3 * 2 + 2 * 3  
* 2  r00 r00  
 r00 r00  r00 r00

2 2
x
σ(zk2) = ∑ ∑ Bij(2) γ (ik2) (k = 1, 4) (7.45b)
i =1 j =1 rij3

2 2  1  1 1 
τ z(12) = ∑ ∑ Bij( 2)ϖ i(12)  * − x( x + i y)  2 *2 + 3 *  
 rij rij  rij rij  
i =1 j =1  rij rij  (7.45c)
( 2) ( 2) ( 2)
 1  1 1 
+ B00 s0 ν k  + i y( x + i y)  2 * 2 + 3 *  
r r *  r r r
 00 00  00 00 00 r00  

Then, from the continuity conditions in Eq. (7.11), we have

( 2)
B0 + B00 = B00 (7.46a)

2 2
Bi + ∑ Bji = ∑ B(ji2) (7.46b)
j =1 j =1
7.4 Bimaterial Green’s Functions for Other Interface Conditions 237

2 2
β i 1 Bi − ∑ β j 1 Bji = ∑ β(j21) B(ji2) (7.46c)
j =1 j =1

s0ν 1 B0 − s0ν 1 B00 = s0( 2)ν 1( 2) B00


( 2)
(7.46d)

2 2
ϖ i 1 Bi − ∑ ϖ j 1 Bji = ∑ ϖ (j 12) B(ji2) (7.46e)
j =1 j =1

2 2
γ i 1 Bi + ∑ γ j 1 Bji = ∑ γ (j 12) B(ji2) (7.46f)
j =1 j =1

where i = 1,2, and the constants Bi (i = 0, 1, 2) are given in Eqs. (6.107) and (6.108).
( 2)
As expected, Eqs. (7.46a) and (7.46d) can be solved for B00 and B00 which are the
same as in Eq. (7.25). The remaining eight unknown constants Bij , Bij( 2) (i, j = 1, 2) are
determined from the eight equations in Eqs. (7.46b), (7.46c), (7.46e), and (7.46f).

7.4 Bimaterial Green’s Functions for Other Interface Conditions


Section 7.2 presents the Green’s function solutions for an MEE bimaterial space
when the two half-spaces are bonded perfectly to each other. Here we explore
several situations when the interface is not perfect.

7.4.1 Solutions for a Smoothly Contacting and Perfectly Conducting


Interface
We first consider the case in which the two half-spaces contact mechanically
smoothly with each other, but the continuity conditions associated with the electric
and magnetic fields are satisfied. In this case, we have, instead of Eq. (7.11)

wk(1) = wk( 2) , σ(zk


1)
= σ(zk2) , τ(z11) = 0, τ(z21) = 0 (k = 1, 2, 3) (7.47)

Then, for a vertical point force, a negative electric charge or a negative magnetic cha­
rge, the thirty-two algebraic equations to determine the thirty-two unknowns become

4 4
β ik Ai + ∑ β jk Aji = − ∑ β jk
( 2) ( 2)
Aji (7.48a)
j =1 j =1

4
ϖ i 1 Ai + ∑ ϖ j 1 Aji = 0 (7.48b)
j =1

4
∑ϖ (j12) A(ji2) = 0 (7.48c)
j =1
238 Green’s Functions in a TI MEE Bimaterial Space

4 4
γ ik Ai − ∑ γ jk Aji = − ∑ γ (jk2) A(ji2) (7.48d)
j =1 j =1

where i = 1,2,3,4; k = 1,2,3.


For a horizontal point force, the algebraic equations for determining the thirty-
four unknowns become

4 4
βik Bi − ∑ β jk Bji = ∑ β(jk2) B(ji2) (7.49a)
j =1 j =1

B0 − B00 = 0 (7.49b)

( 2)
B00 =0 (7.49c)

4
ϖ i 1 Bi − ∑ ϖ j 1 Bji = 0 (7.49d)
j =1

4
∑ϖ (j12) B(ji2) = 0 (7.49e)
j =1

4 4
γ ik Bi + ∑ γ jk Bji = ∑ γ (jk2) B(ji2) (7.49f)
j =1 j =1

where i = 1,2,3,4; k = 1,2,3.


( 2)
Remark 7.9: As we have noticed in the end of Section 7.3.1, B00 and B00 are associated
with the purely elastic shear deformation within the plane of isotropy. Thus, for a
horizontal force applied in the upper half-space (Material 1) and when the two half-
spaces can slip freely with respect to each other, such a shear deformation will not
(2)
be induced in the lower half-space (Material 2), that is, we have B00 = 0 as given by
Eq. (7.49c).
Remark 7.10: For the smoothly contacting interface considered here, the expressions
for all field variables are the same as those for a perfect interface, except that the
unknown constants assumed in the potential functions are determined from the
algebraic equations presented in the preceding text.

7.4.2 Solutions for a Mechanically Perfect and Electromagnetically


Insulating Interface
In this case, the interface conditions become

U (1) = U ( 2) , w1(1) = w1( 2) , σ z(11) = σ z( 21) , τ z(11) = τ z( 21) , σ z(12) = σ z(22) = 0, σ z(13) = σ z(23) = 0 (7.50)
7.4 Bimaterial Green’s Functions for Other Interface Conditions 239

where, in addition, we have assumed that the interface is free from electric charge
and current (magnetic flux).
Then, for a vertical point force, a negative electric charge or a negative magnetic
charge, we have

4 4
Ai − ∑ Aji = − ∑ A(ji2) (7.51a)
j =1 j =1

4 4
β i 1 Ai + ∑ β j 1 Aji = − ∑ β j(12) A(ji2) (7.51b)
j =1 j =1

4 4
ϖ i 1 Ai + ∑ ϖ j 1 Aji = − ∑ ϖ (j 12) A(ji2) (7.51c)
j =1 j =1

4 4
γ i 1 Ai − ∑ γ j 1 Aji = − ∑ γ (j 12) A(ji2) (7.51d)
j =1 j =1

4
γ ik Ai − ∑ γ jk Aji = 0 (7.51e)
j =1

4
∑ γ (jk2) A(ji2) = 0 (7.51f)
j =1

where i = 1,2,3,4; k = 2,3. There are totally thirty-two equations in Eq. (7.51) for
determining the thirty-two unknowns, Aij and Aij( 2) (i, j = 1, 2, 3, 4).
For a horizontal point force, the thirty-four algebraic equations are

( 2)
B0 + B00 = B00 (7.52a)

4 4
Bi + ∑ Bji = ∑ B(ji2) (7.52b)
j =1 j =1

4 4
β i 1 Bi − ∑ β j 1 Bji = ∑ β(j21) B(ji2) (7.52c)
j =1 j =1

s0ν 1 B0 − s0ν 1 B00 = s0( 2)ν 1( 2) B00


( 2)
(7.52d)

4 4
ϖ i 1 Bi − ∑ ϖ j 1 Bji = ∑ ϖ (j 12) B(ji2) (7.52e)
j =1 j =1
240 Green’s Functions in a TI MEE Bimaterial Space

4 4
γ i 1 Bi + ∑ γ j 1 Bji = ∑ γ (j 12) B(ji2) (7.52f)
j =1 j =1

4
γ ik Bi + ∑ γ jk Bji = 0 (7.52g)
j =1

4
∑ γ (jk2) B(ji2) = 0 (7.52h)
j =1

where i = 1,2,3,4; k = 2,3.


Remark 7.11: We can also consider the case when the interface is electrically
conducting and magnetically insulating (without magnetic flux) or vice versa. For
the former, we have the following interface conditions:

U (1) = U ( 2) , wk(1) = wk( 2) , σ(z11) = σ(z21) , τ(z11) = τ(z21) , σ(z13) = σ(z32) = 0 (7.53)

where k = 1,2. For the latter, we have the following interface conditions:

U (1) = U (2) , wk(1) = wk(2) , σ (z11) = σ (z21), τ(z11) = τ(z21) , σ (z12) = σ (z22) = 0 (7.54)

where k = 1,3. We point out that, as listed and discussed in Chapter 4, there are
many other types of possible interfaces obtained from different combinations of
mechanical (rigid, smooth, fixed, …), electric (conducting, insulating, electroded, …)
and magnetic (conducting, insulating, …) fields. Remarked in the following text is
such an example.
Remark 7.12: If the two half-spaces are separated from each other completely, then
the problem is reduced to a half-space one with different surface conditions, as
will be discussed in the next section. If the two half-spaces are only mechanically
separated, they can be still coupled through the electric and magnetic quantities at
the interface. In fact, the interface conditions read as

σ (z11) = σ (z21) = 0, τ(z11) = τ(z21) = 0, wk(1) = wk( 2) , σ (zk1) = σ (zk2) (7.55)

where k = 2,3. Thus, for a vertical point force, a negative electric charge or a negative
magnetic charge, we have

4
β i 1 Ai + ∑ β j 1 Aji = 0 (7.56a)
j =1

4
∑ β j(12) A(ji2) = 0 (7.56b)
j =1
7.5 Half-Space Green’s Functions 241

4
ϖ i 1 Ai + ∑ ϖ j 1 Aji = 0 (7.56c)
j =1

4
∑ϖ (j12) A(ji2) = 0 (7.56d)
j =1

4 4
β ik Ai + ∑ β jk Aji = − ∑ β jk
( 2) ( 2)
Aji (7.56e)
j =1 j =1

4 4
γ ik Ai − ∑ γ jk Aji = − ∑ γ (jk2) A(ji2) (7.56f)
j =1 j =1

where k = 2,3. Then, it is clearly seen that the constants in the upper half-space and
those in the lower half-space are coupled through Eqs. (7.56e) and (7.56f). Thus,
a source in the upper half-space will generally induce the magnetic, electric, and
elastic fields in the lower half-space, although they are not mechanically contacting
with each other.

7.5 Half-Space Green’s Functions


The Green’s function solutions for an MEE half-space (z ≥ 0) subjected to point
sources at (0,0,h ≥ 0) can be obtained in a way similar to the bimaterial case. They
are also composed of two parts, that is, the part corresponding to the infinite space,
and the complementary part due to the half-space boundary. Obviously, they depend
on the surface conditions at z = 0, and two typical ones (i.e., free and electrode) will
be exemplified in the following subsections.
We first consider the situation for which the Green’s solutions are known as
the extended Mindlin solutions for free boundary case and the extended Lorentz
solution for fixed boundary case (Ding et al. 2005).

7.5.1 Green’s Functions for an MEE Half-Space with Free Surface


The boundary conditions at z = 0 for an MEE half-space (occupying Material 1) free
from surface tractions, electric charge, and magnetic charge are

1)
σ(zk = 0, τ(z11) = 0 (7.57)

where k = 1,2,3. Thus, for a vertical point force, a negative electric charge or a negative
magnetic charge applied at (0,0,h > 0), we can derive from Eqs. (6.34)–(6.38), (7.4)
and (7.5) that

4
γ ik Ai − ∑ γ jk Aji = 0 (7.58a)
j =1
242 Green’s Functions in a TI MEE Bimaterial Space

4
ϖ i 1 Ai + ∑ ϖ j 1 Aji = 0 (7.58b)
j =1

where i = 1,2,3,4; k = 1,2,3. There are sixteen algebraic equations, sufficient to uniquely
determine the sixteen unknown constants Aij (i, j = 1, 2, 3, 4).
For a horizontal point force, the following seventeen algebraic equations are
obtained

B00 = B0 (7.59a)

4
ϖ i 1 Bi − ∑ ϖ j 1 Bji = 0 (7.59b)
j =1

4
γ ik Bi + ∑ γ jk Bji = 0 (7.59c)
j =1

where i = 1,2,3,4; k = 1,2,3.

7.5.2 Green’s Functions for an MEE Half-Space with Surface Electrode


When the surface of the MEE half-space is free from surface tractions and magnetic
charge, but with an electrode, the boundary conditions at z = 0 are

σ(z11) = σ(z13) = 0, τ(z11) = 0, w2(1) = 0 (7.60)

Then, for a vertical point force, a negative electric charge or a negative magnetic
charge applied at (0,0,h > 0), we can derive from Eqs. (6.34)–(6.38), (7.4) and
(7.5) that

4
γ ik Ai − ∑ γ jk Aji = 0 (7.61a)
j =1

4
ϖ i 1 Ai + ∑ ϖ j 1 Aji = 0 (7.61b)
j =1

4
β i 2 Ai + ∑ β j 2 Aji = 0 (7.61c)
j =1

where i = 1,2,3,4; k = 1,3. There are sixteen algebraic equations in Eq. (7.61) that are
sufficient to determine the sixteen unknown constants Aij (i, j = 1, 2, 3, 4).
For a horizontal point force, the following seventeen algebraic equations are
obtained

B00 = B0 (7.62a)
7.5 Half-Space Green’s Functions 243

4
β i 2 Bi − ∑ β j 2 Bji = 0 (7.62b)
j =1

4
ϖ i 1 Bi − ∑ ϖ j 1 Bji = 0 (7.62c)
j =1

4
γ ik Bi + ∑ γ jk Bji = 0 (7.62d)
j =1

where i = 1,2,3,4, k = 1,3. Again, the purely elastic shear deformation associated with
B0 and B00 is decoupled from other coupled elastic, electric, and magnetic fields.

7.5.3 Surface Green’s Functions


In the preceding derivation, we have confined ourselves to the interior source case
(i.e., with h > 0). If the sources act exactly on the surface of the half-space, we will
have h = 0, and these situations are frequently encountered in many areas, say
indentation studies. The Green’s functions for a half-space with h = 0 are usually
called surface Green’s functions, or more specifically, the extended Boussinesq
solutions for a vertical point force, electric charge, or magnetic charge, and extended
Cerruti solutions for a horizontal point force.

7.5.3.1 Extended Boussinesq Solutions for a Vertical Point Force,


Electric Charge, or Magnetic Charge
When h = 0, noticing that z ≥ 0, we have Rij = Ri , Rij* = Ri* = Ri + zi , and the
complementary part in Eqs. (7.1) has the same form as the infinite space solution
given by Eq. (6.29). Thus, the surface Green’s solution for a vertical point force,
electric charge, or magnetic charge can be assumed as

Ψ0 = 0, Ψi = Ai ln Ri* (i = 1, 2, 3, 4) (7.63)

with the extended displacements and stresses being

4
( x + i y)
U = −∑ Ai (7.64a)
i =1 Ri Ri*

4
1
wk = ∑ Ai βik (k = 1, 2, 3) (7.64b)
i=1 Ri

4  1 1 
σ 2 = 2c66 ∑ Ai ( x + i y)2  3 * + 2 *2  (7.65a)
i =1  Ri Ri Ri Ri 
244 Green’s Functions in a TI MEE Bimaterial Space

4
zi
   σzk = − ∑ Ai γ ik (k = 1, 2, 3, 4) (7.65b)
i=1 Ri3

4
( x + i y)
τ zk = − ∑ Ai ϖik (k = 1, 2, 3) (7.65c)
i =1 Ri3

Because the surface is free from tractions everywhere except at the source point, we
obtain from Eqs. (7.65b) and (7.65c)

4
∑ Aiϖ i1 = 0 (7.66)
i =1

where the identities in Eq. (6.24) have been applied. The equilibrium of the MEE
layer bounded by the surface z = 0 and the plane z = ε demands

∞ ∞
∫−∞ ∫−∞ σzk ( x, y, ε)d x d y + fk = 0 (k = 1, 2, 3) (7.67)

where f1 = fz, f2 = − fe, and f3 = − fh . Then, by making use of Eq. (6.41), we obtain

4
2 π ∑ Ai γ ik = fk (k = 1, 2, 3) (7.68)
i =1

This differs from the infinite space counterpart, Eq. (6.42), by a factor of 2. Thus, the
unknown constants Ai (i = 1, 2, 3, 4) are solved from Eqs. (7.66) and (7.68) as

−1 −1
 A1  ϖ 11 ϖ 21 ϖ 31 ϖ 41  0 ϖ 11 ϖ 21 ϖ 31 ϖ 41   0 
 A2  1  γ 11 γ 21 γ 31 γ 41   f1  1  γ 11 γ 21 γ 31 γ 41   fz 
A  = f  = − f 
γ γ γ γ  γ γ γ γ 
 A3  2π  γ 12 γ 22 γ 32 γ 42   f2  2π  γ 12 γ 22 γ 32 γ 42  − f e 
 4  13 23 33 43   3  13 23 33 43   h
 (7.69)

7.5.3.2 Extended Cerruti Solutions for a Horizontal Point Force


Similarly, the complementary part in Eqs. (7.14) has the same form as the infinite-
space solution given by Eq. (6.45). Thus, we can simply take

B0 y Bi x
Ψ0 = , Ψi = ( i = 1, 2, 3, 4) (7.70)
R0* R*j

The corresponding extended displacements and stresses are

 1 i y( x + i y)  4  1 x( x + i y) 
* 2  ∑ i  R*
U = B0  * + − B − (7.71a)
R
 0 R R
0 0  i =1  i Ri Ri*2 

4
x
wk = − ∑ Bi βik (k = 1, 2, 3) (7.71b)
i =1 Ri Ri*
7.5 Half-Space Green’s Functions 245

 2( x + i y)  1 2 
σ 2 = −2c66 B0  + i y( x + i y)2  3 *2 + 2 *3  
 R R0
 R0 R0
* 2
 0 R0 R0  
(7.72a)
4  2( x + i y)  1 2 
+ 2c66 ∑ Bi  * 2
− x( x + i y)2  3 *2 + 2 *3  
i =1  Ri Ri  Ri Ri Ri Ri  

4
x
σzk = ∑ Bi γ ik (k = 1, 2, 3, 4) (7.72b)
i =1 Ri3

 1  1 1 
τ zk = − s0 νk B0  + i y( x + i y)  + 
 R0 R0
*
 R03 R0* R02 R0*2  
   (7.72c)
4  1  1 1 
− ∑ ϖik Bi  *
− x( x + i y)  3 * + 2 *2   (k = 1, 2, 3)
i =1  Ri Ri  Ri Ri Ri Ri  

Vanishing of the extended normal stresses σ zk (k = 1, 2, 3) and shear stress


component τ yz = Im[τ z1 ] on the surface requires

4
∑ Bi γ ik = 0 (k = 1, 2, 3) (7.73)
i =1

4
s0ν 1 B0 − ∑ϖ i 1 Bi = 0 (7.74)
i =1

The second equation also makes the shear stress component τ xz = Re[τ z1 ] vanish
everywhere except at the source point. The equilibrium of the MEE layer bounded
by the surface z = 0 and the plane z = ε is now expressed as

∞ ∞
∫−∞ ∫−∞ τ xz ( x, y, ε )d x d y + fx = 0 (7.75)

which gives

4
fx
s0ν 1 B0 + ∑ϖ i 1 Bi = (7.76)
i =1 π

Then, we can solve from Eqs. (7.73)–(7.75) the involved coefficients to find

fx fx
B0 = = (7.77)
2πν 1 s0 2π c44 s0

−1
 B1  ϖ 11 ϖ 21 ϖ 31 ϖ 41  1
 B2  f x  γ 11 γ 21 γ 31 γ 41  0 
 B3  = 0  (7.78)
γ γ γ γ 
 B  2π  γ 12 γ 22 γ 32 γ 42  0 
 4  13 23 33 43   
246 Green’s Functions in a TI MEE Bimaterial Space

It is interesting to note that, the inverse matrix for the horizontal force in Eq. (7.78)
is the same as that for the vertical force (or electric/magnetic charge) in Eq. (7.69).
Remark 7.13: In this subsection, we have obtained the surface Green’s functions
in a very straightforward way. Similarly, one can also derive the interface Green’s
functions when the source point is located on the interface of the bimaterial space.
This, however, is left to the interesting reader.

7.6 Technical Application: Indentation over an MEE Half-Space


The instrumented indentation technique, in which the load and indentation depth
are continuously measured and recorded, has become one of the most useful and
convenient methods to characterize mechanical properties of advanced as well as
conventional materials (Oliver and Pharr 1992; Bischoff 2004). They can be applied
at macroscale, microscale, and nanoscale, and hence are very attractive. Contact
mechanics, primarily established based upon the masterpiece of Hertz (1882), plays
a significant role in relating the indentation responses to various properties (e.g.,
elastic modulus, hardness, fracture toughness, yield strength, residual stress, etc.) of
the indented specimen. There are a few excellent monographs on this topic (Gladwell
1980; Johnson 1985; Galin and Gladwell 2008).

7.6.1 Theory of Indentation


We first give here a brief review on the theory and analysis of the indentation technique
proposed by Oliver and Pharr (1992) for isotropic elastic materials so that the reader
will have a basic understanding of the theoretical works to be presented in subsequent
subsections. The indentation theory is established completely based on Hertz contact
equations. In practice, however, some additional complexities need to be considered
in the analysis of the recorded data for a correct and precise interpretation of the
indentation response (Oliver and Pharr 1992; Fischer-Cripps 2007).
The typical indenter used in the instrumented indentation techniques is the
three-sided Berkovich indenter with the face half-angle being θ = 65.27°. It can be
theoretically (and approximately) treated as a conical indenter of half-apical angle β
regarding the equivalence of the projected contact area:

A = 3 3hc2 tan 2 θ = π hc2 tan 2 β (7.79)

where hc is the vertical distance between the indenter tip and the contact edge, see
Figure 7.2(a). The preceding equation yields β = 70.296°, which in turn gives

A(hc ) = 24.5hc2 (7.80)

A(hc ) is known as the contact area function.


Thus, we will pay our attention only to the indentation of an isotropic elastic
half-space by a conical indenter (with half-apical angle β = 70.296°) in the remaining
part of this subsection. From the analysis of Sneddon (1948), we have the following
Hertz contact equations:
7.6 Technical Application: Indentation over an MEE Half-Space 247

(a) Undeformed surface

Unloaded
ha a

hr
hmax
hc Loaded

(b) P

Pmax

Loading

Unloading S

h
hmax
hr

hc

Figure 7.2. (a) Schematic of geometry of indentation at loaded and unloaded stages for a
conical indenter. (b) Typical load-displacement (P-h) curves for loading and unloading. The
unloading deformation is assumed to be purely elastic, while the loading process involves
inelastic effect, leading to a residual depth hr at the center of contact after the load is
completely removed.

π * 2  π r
P= E a cot β, uz =  −  a cot β (r ≤ a) (7.81)
2  2 a

where P is the applied load, a is the contact radius, uz is the surface sediment (within
the contact area). E* is the reduced modulus of the indentation system, defined by

1 1 −ν 2 1 −ν i2
= + (7.82)
E* E Ei

with E,ν and Ei ,ν i being Young’s modulus and Poisson’s ratio of the half-space
(specimen) and the indenter, respectively.
Taking h = uz r =0, that is, the depth of penetration, we get from Eq. (7.81)

2E* 2 d P 4E*
P= h tan β , = h tan β (7.83)
π dh π
248 Green’s Functions in a TI MEE Bimaterial Space

Combining the second equation in the preceding text with Eq. (7.79) leads to

dP 1 π
E* = (7.84)
dh 2 A

As has been shown by Pharr et al. (1992), Eq. (7.84) actually holds for any
axisymmetric indenter with a smooth generatrix.
Because the loading process may involve inelastic deformation (e.g., plastic
deformation due to the sharpness of the indenter) and other complexities, Oliver
and Pharr (1992) suggested the unloading curve only be used for data interpretation.
Applying the Hertz contact equations to the unloading state, and omitting the effect
of perturbed surface (with residual impression), we find

π  π 1
ha =  − 1 a cot β, he = a cot β, Pmax = She (7.85)
2  2 2

where ha is the surface deflection at the perimeter and he = hmax − hr is the recoverable
penetration depth, and S = d P / d h is the indentation stiffness evaluated at P = Pmax .
We then arrive at

Pmax
hc = hmax − ha = hmax − ε (7.86)
S

where ε = 2 − 4 / π ≈ 0.72 . Having obtained hc, we can calculate the contact area from
Eq. (7.80), and then use Eq. (7.84), which can be written as

π S
E* = (7.87)
2 A

to obtain the reduced Young’s modulus. Meanwhile, Pharr et al. (1992) proposed to
use the following formula to calculate the hardness of the indented material:

Pmax
H= (7.88)
A

which has been shown to work quite well.


Remark 7.14: To accurately calculate the value of contact stiffness S, the unloaded
data are used to fit the following equation

P = B(h − hr )m (7.89)

where theoretically, for an ideal conical indenter, m = 1.5 , but it ranges actually from
1.1 to 1.8, depending on the specimen material (Fischer-Cripps 2007).
Remark 7.15: Several important issues need to be considered for the correct
interpretation of the P−h curve (Fig. 7.2b). These include the incorporation of load
frame compliance, the correction of the contact area function to account for the nonideal
sharpness at the tip of the indenter, as well as the modification of Eq. (7.87) into
7.6 Technical Application: Indentation over an MEE Half-Space 249

1 π S
E* = (7.90)
β 2 A

by comparing with the numerical simulations, here β = 1.034 is a correction factor.


Issues related to the specimen material may be also critical. The reader is referred to
Fischer-Cripps (2007) for more detailed discussion.
Remark 7.16: Depending on the specimen material and the loading range, spherical
indenter may be preferred for which only the loading curve will be utilized
(Ramamurty et al. 1999).
Remark 7.17: From the preceding analysis, we can see that the Hertz contact equations
play a central role in interpreting the load-displacement curve. Thus, it becomes
emergent to develop appropriate indentation theories for advanced materials in
which material anisotropy, multifield coupling, viscoelasticity, and so forth may be
encountered. A great deal of works have already been done in these directions. For
examples, the effect of material anisotropy was discussed by Vlassak and Nix (1994)
and Swadener and Pharr (2001); the electromechanical coupling was considered by
Suresh et al. (Giannakopoulos and Suresh 1999; Giannakopoulos 2000), Chen et al.
(Chen and Ding 1999; Chen et al. 1999; Chen 2000), Ding et al. (1999, 2000), and Wang
et al. (2009); and the magnetomechanical coupling was considered by Giannakopoulos
and Parmaklis (2007). In particular, Suresh et al. (Ramamurty et al. 1999; Saigal et al.
1999; Sridhar et al. 1999) conducted a series of experiments to probe the properties
of piezoelectric materials using the instrumented indentation technique. By contrast,
Kalinin et al. (Kalinin and Bonnell 2002; Kalinin et al. 2004) explored the imagining
mechanism of piezoresponse force microscopy using the generalized Hertz contact
equations for piezoelectric materials.

7.6.2 Indentation over an MEE Half-Space


In this subsection, we consider the response of an MEE half-space indented by an
axisymmetric rigid indenter (Chen et al. 2010). It is worthy to mention the related
studies by Hou et al. (2003) for elliptical Hertz contact, Rogowski and Kalinski
(2012) for a truncated conical punch, and Rogowski (2012) for a concave indenter.
The indenter is assumed to be electrically and magnetically conducting and its
shape can be spherical, conical, or flat-ended (see Figure 7.3). Here we will present
some particular results, and a complete analysis can be found in the paper of Chen
et al. (2010).
The problem can be described as a mixed boundary-value problem of an MEE
half-space, with the following boundary conditions prescribed on its surface z = 0:

0 ≤ r ≤ a : wk (r, 0) = gk (r ) (k = 1, 2, 3) (7.91)

  r > a : σzk (r, θ, 0) = 0 (k = 1, 2, 3) (7.92)

r ≥ 0 : τ z1 (r, 0) = 0 (7.93)
250 Green’s Functions in a TI MEE Bimaterial Space

Undeformed surface
β
R

h h
h

2a 2a 2a
Figure 7.3. Indentation of an MEE half-space by an axisymmetric rigid indenter. (a) Spherical
(b) Conical (c) Circular and flat-ended.

where for simplicity, we have used the cylindrical coordinates (r, θ , z), a is the radius
of contact, and

h − r 2 / (2 R) (for sphere)

g1 (r ) = h − r cot β (for cone) (7.94)
h (for flat)

g2 (r ) = φ0 (7.95)

g3 (r) =ψ 0 (7.96)

where h is the indentation depth, R is the radius of the spherical indenter, β is the
half-apical angle of the conical indenter, and φ0 and ψ0 are the electric and magnetic
potentials prescribed on the conducting indenter, respectively. We note here that the
expression for g1(r) for the spherical indenter is not exact but approximate regarding
the fact that the radius of the contact is much less than the radius of the indenter.
Also, different electric and/or magnetic conditions (i.e., electrically insulating and
magnetically conducting, electrically conducting and magnetically insulating, and
electrically and magnetically insulating) can be imposed on the indenter, and the
reader is referred to Chen et al. (2010) for the detailed discussion.
To solve the problem, we here adopt the surface Green’s functions presented in
Section 7.5.3. The problem is axisymmetric, and the extended Boussinesq solutions
(7.64) and (7.65) can be rewritten in the cylindrical coordinates as

4 4
r 1
ur = −∑ Ai *
, wk (r , z) = ∑ β ik Ai (7.97)
i =1 Ri Ri i =1 Ri

4
zi
σzk (r, z) = −∑ γ ik Ai (k = 1, 2, 3, 4)
i =1 Ri3
4
r
τ zk (r, z) = − ∑ γ ik si Ai (k = 1, 2, 3) (7.98)
i =1 Ri3
4
r2  1 1 
σ 2 (r, z) = 2c66 ∑ Ai +
i =1 Ri2 Ri*  Ri Ri* 

where the identities in Eq. (6.24) have also been used.The constants Ai are determined
from Eq. (7.69), or they can be rewritten as (Chen et al. 2010)
7.6 Technical Application: Indentation over an MEE Half-Space 251

3
Ai = ∑ I ij f j (7.99)
j =1

where f1 = fz , f2 = − fe, f3 = − fh , and

 I1 j  −1
γ 11 s1 γ 21 s2 γ 31 s3 γ 41 s4   0 
 I 2 j  1  γ 11 γ 21 γ 31 γ 41  δ 1 j 
I  = δ 2 j  (7.100)
3j 2π  γ 12 γ 22 γ 32 γ 42 
I   γ 13 γ 23 γ 33 γ 43  δ 
 4j   3j 

with δ ij being the Kronecker delta.


The indentation-induced extended displacements can be expressed, using the
method of superposition, in terms of integrals over the circular contact area S:
0 ≤ r ≤ a of the pressure, electric charge, and magnetic flux, which are yet unknown.
This immediately gives us

3 p j (ρ, θ0 )
2π a
wk (r, θ, 0) = ∑ ξkj ∫0 ∫0 R0
ρ d ρ d θ0 (k = 1, 2, 3) (7.101)
j =1

where R0 = r 2 + ρ 2 − 2r ρ cos(θ − θ0 ) is the distance between points (r, θ , 0) and


( ρ , θ 0 , 0) on the surface, and

pk (r, θ) = −σ zk (r, θ, 0) (k = 1, 2, 3). (7.102)

4
The coefficients in Eq. (7.101) are defined as ξkj = ∑ β ik I ij , and we have ξ kj = ξ jk ,
i =1
which can be verified through the reciprocity theorem established in Appendix B of
Chen et al. (2010).
From Eq. (7.101), we obtain

2π a pj (ρ, θ0 )
∫0 ∫0 R0
ρ d ρ d θ0 = q j ( j = 1, 2, 3) (7.103)

where

1 3
qj = ∑ ηkj wk (r,θ , 0)
η k =1
(7.104)

with η = ξkj being the determinant, and ηkj = η jk the cofactors of the matrix [ξkj ].
The three integral equations in Eq. (7.103) bear the same mathematical structure,
and they can be solved by making use of the potential theory method in Fabrikant
(1989). First, they can be transformed into, according to Fabrikant (1989),

r dx a ρd ρ  x2 
    4 ∫ ∫x (ρ2 − x 2 )1/ 2  rρ  pj (ρ, θ) = qj (r, θ)
L ( j = 1, 2, 3) (7.105)
0 (r 2 − x 2 )1/ 2
252 Green’s Functions in a TI MEE Bimaterial Space

where

1 2π
L ( k ) f (θ ) = ∫0 λ (k,θ − θ 0 ) f (θ 0 )d θ 0 (7.106)

is the Poisson operator, with

1 − k2
λ (k , θ ) = (7.107)
1 + k 2 − 2k cos θ

By inverting two Abel operators and one Poisson operator in Eq. (7.105), we can
obtain its solution as

1  χ j (a, r, θ ) a d t d 
π 2  a 2 − r 2 ∫r t 2 − r 2 d t
p j (r, θ , 0) =  − χ j (t, r, θ ) (7.108)

with

1 t x d x d   xr  
χ j (t , r , θ ) =
t ∫0  xL  2  q j ( x, θ ).
t 2 − x2 d x   t  
(7.109)

The right-hand side of Eq. (7.108) will be singular at r = a if χ j (a, a,θ ) is not zero. For
the three indenters shown in Figure 7.3, we can obtain

 η1 j t 2
q0 j − η R (for sphere)

 πt η1 j
χ j (t, r, θ) = q0 j − cot β (for cone) (7.110)
 2 η
q0 j (for flat)

1
where q0 j = (η11h + η21φ0 + η31ψ 0 ). For a smooth indenter (e.g., conical or spherical
η
indenter), it is usually required, from physical consideration, that no stress singularity
occur at the contact edge r = a. Thus, we have

π a η11
q01 − cot β = 0 (7.111)
2 η

for the conical indenter, and

η11 a 2
q01 − =0 (7.112)
η R
7.6 Technical Application: Indentation over an MEE Half-Space 253

for the spherical indenter. Equation (7.111) or (7.112) provides the relation for
determining the contact radius for the conical or spherical indenter. For the circular
flat-ended cylindrical indenter, the situation is, however, quite different: The contact
radius is predetermined and equal to the radius of the indenter, and stress singularity
inevitably occurs at the contact edge.
Substituting Eq. (7.110) into Eq. (7.108), we obtain the following exact
expressions for the pressure, electric charge and magnetic flux, all θ -independent,
within the contact area:

 2 η1 j η1 j (h − a 2 / R) + η2 j φ0 + η3 j ψ 0 1
 2 a2 − r 2 + (for sphere)
 π η R π 2η
a − r2
2
 η1 j cot β  a  η1 j (h − πa cot β / 2) + η2 j φ0 + η3 j ψ 0 1
p j (r ) =  cosh −1   + (for cone)
 2 πη  r π2η a2 − r2
 η1 j h + η2 j φ0 + η3 j ψ 0 1
 (for flat)
 π 2η
a 2 − r2

 (7.113)

By integrating Eq. (7.113) over the contact area, we get

a
Pj = 2 π ∫ p j (r )r d r
0
 2a  2  η1 j η21   η1 j η31  
  η1 j h +  η2 j −  φ0 +  η3 j −  ψ0  (for sphere)
 πη  3  3η11   3η11  
 a   η1 j η21   η1 j η31   (7.114)
=   η1 j h +  2 η2 j −  φ0 +  2 η3 j −  ψ0  (for cone)
 πη   η11   η11  
 2a(η1 j h + η2 j φ0 + η3 j ψ 0 )
 (for flat)
 πη

Note that the resultant force P, total electric charge Q and total magnetic flux M
applied on the indenter are given by P = P1, Q = − P2 and M = − P3 . The relations in
Eq. (7.114) are of particular importance in indentation techniques.
When the indenter is electrically insulating or magnetically insulating, the
indentation analysis is quite similar. Tables 7.1 through 7.4 summarize various
important analytical results of the indentation responses for indenters with different
electric and magnetic properties.
Remark 7.18: In view of Eqs. (7.111) and (7.112), we can obtain from Eq.
(7.113) that

η11 cot β a


p1 (r) = cosh −1   (7.115)
2πη r
254

Table 7.1. Analytical Results for Indentation Responses of an MEE Half-space by Electrically and Magnetically Conducting Circular Indenters

Quantity Cone Sphere Flat-ended punch

Contact radius (a) 2 ηP 3πηRP constant


3
η11 cot β 4 η11
Penetration depth (h) πa η φ + η31ψ 0 a2 η21φ0 + η31ψ 0 πηP η21φ0 + η31ψ 0
cot β − 21 0 − −
2 η11 R η11 2aη11 η11

Resultant force (P) a 4a 2a


( η11h + η21φ0 + η31ψ 0 ) ( η11h + η21φ0 + η31ψ 0 ) ( η11h + η21φ0 + η31ψ 0 )
πη 3πη πη
Contact pressure (p1) η11 cot β  a 2 η11 η11h + η21φ0 + η31ψ 0 1
cosh −1   a2 − r 2
2 πη  r π 2 ηR π2η a2 − r2
Total electric charge (Q) a   η12 η21  2a  2  η12 η21  2a( η12 h + η22 φ0 + η32 ψ 0 )

−  η12 h +  2 η22 − φ0 −  η12 h +  η22 − φ0
πη   η11  πη  3  3η11  πη
 η η    η η  
+  2 η32 − 12 31  ψ 0  +  η32 − 12 31  ψ 0 
 η11    3η11  
Contact surface electric η12 cot β  a 1 2 η12 1 η12 h + η22φ0 + η32 ψ 0 1
charge (p2) cosh −1   + a2 − r 2 +
2 πη  r a2 − r 2 π 2 ηR a − r2
2 π2η a2 − r2
η (h − πa cot β / 2) + η22 φ0 + η32 ψ 0 η (h − a / R) + η22φ0 + η32 ψ 0
2
× 12 × 12
π2 η π2η
Total magnetic charge (M) a   η13 η21  2a  2  η13 η21  2a( η13h + η23φ0 + η33 ψ 0 )

−  η13h +  2 η23 − φ0 −  η13h +  η23 − φ0
πη   η11  πη  3  3η11  πη
 η η    η η  
+  2 η33 − 13 31  ψ 0  +  η33 − 13 31  ψ 0 
 η11    3η11  
Contact surface magnetic η13 cot β  a 1 2 η13 1 η13h + η23φ0 + η33ψ 0 1
charge (p3) cosh −1   + a2 − r 2 +
2 πη  r a2 − r 2 π 2 ηR a2 − r 2 π2η a2 − r2
η (h − πa cot β / 2) + η23φ0 + η33 ψ 0 η (h − a 2 / R) + η23φ0 + η33ψ 0
× 13 × 13
π2η π2η
7.6 Technical Application: Indentation over an MEE Half-Space 255

Table 7.2. Analytical Results for Indentation Responses of an MEE Half-space by Electrically Conducting
and Magnetically Insulating Circular Indenters

Quantity Cone Sphere Flat-ended punch

Contact radius 2 η33 P 3πη33RP constant


(a) 3
ξ 22 cot β 4ξ 22

Penetration depth πa ξ φ a2 ξ12φ0 πη33 P ξ 21φ0


(h) cot β + 12 0 + +
2 ξ 22 R ξ 22 2aξ 22 ξ 22
Resultant force (P) ξ a 2 cot β 4ξ 22 a3 2a
22 (ξ 22 h − ξ12φ0 )
2 η33 3πη33 R πη33
Contact pressure ξ 22 cot β  a 2ξ 22 ξ 22 h − ξ12 φ0 1
(p1) cosh −1   a2 − r 2
2πη33  r π 2 η33R π 2 η33 a2 − r 2
Total electric ξ 21a 2 cot β 2a(ξ11 − ξ12
2 / ξ )φ 4ξ 21a3 2a(ξ11 − ξ12
2 / ξ )φ 2a
− 22 0

22 0 (ξ 21h − ξ11φ0 )
charge (Q) πη33
2 η33 πη33 3πη33 R πη33
Contact surface ξ 21 cot β  a 2ξ 21 −ξ 21h + ξ11φ0 1
electric charge − cosh −1   − a2 − r 2
2πη33  r π 2 η33R π 2 η33 a2 − r 2
(p2) (ξ11 − ξ12
2 / ξ )φ 1
(ξ11 − ξ12 / ξ 22 )φ0
2 1 22 0
+ +
π 2 η33 a2 − r 2 π 2 η33 a2 − r2
Contact surface 1 1   r2   1
− [ η13 (h − r cot β) + η23φ0 ] − (η13h + η23φ0 )
magnetic −  η13  h − + η23φ0 
η33 η33   2R  
η33
potential (ψ)

Table 7.3. Analytical Results for Indentation Responses of an MEE Half-space by Electrically Insulating
and Magnetically Conducting Circular Indenters

Quantity Cone Sphere Flat-ended punch

Contact radius 2 η22 P 3πη22 RP constant


(a) 3
ξ 33 cot β 4ξ 33
Penetration depth πa ξ ψ a2 ξ13ψ 0 πη22 P ξ 31ψ 0
(h) cot β + 13 0 + +
2 ξ 33 R ξ 33 2aξ 33 ξ 33
Resultant force (P) ξ a 2 cot β 4ξ 33a3 2a
33 (ξ 33h − ξ13ψ 0 )
2 η22 3πη22 R πη22
Contact pressure ξ 33 cot β  a 2ξ 33 ξ 33h − ξ13ψ 0 1
(p1) cosh −1   a2 − r 2
2πη22  r π 2 η22 R π 2 η22 a2 − r 2
Total magnetic ξ 31a 2 cot β 2a(ξ11 − ξ13
2 / ξ )ψ 4ξ 31a3 2a(ξ11 − ξ13
2 / ξ )ψ 2a

33 0

33 0 (ξ 31h − ξ11ψ 0 )
charge (M) πη22
2 η22 πη22 3πη22 R πη22
Contact surface ξ 31 cot β 2ξ 31 −ξ 31h + ξ11ψ 0 1
−1  a  − a2 − r 2
magnetic charge − 2πη cosh  r  π η22 R
2
π 2 η22
22 a2 − r 2
(p3) (ξ11 − ξ13
2 / ξ )ψ
(ξ11 − ξ13
2 / ξ )ψ
33 0 1 33 0 1
+ +
π 2 η22 a2 − r 2 π 2 η22 a2 − r 2
Contact surface 1 1   r2   1
− [η12 (h − r cot β) + η32 ψ 0 ] − (η12 h + η32 ψ 0 )
electric −  η12  h − + η32 ψ 0 
η22 η22   2R  
η22
potential (φ)
256 Green’s Functions in a TI MEE Bimaterial Space

Table 7.4. Analytical Results for Indentation Responses of an MEE Half-space by Electrically
and Magnetically Insulating Circular Indenters

Quantity Cone Sphere Flat-ended punch

Contact radius (a) 2ξ11P 3πξ11RP constant


3
cot β 4
Penetration depth (h) πa a2 πξ11P
cot β
2 R 2a
Resultant force (P) a 2 cot β 4a 3 2ah
2ξ11 3πξ11R πξ11
Contact pressure (p1) cot β  a 2 h 1
cosh −1   a2 − r 2
2πξ11  r π 2 ξ11R π 2ξ11 a 2 − r 2
Contact surface electric ξ 21 ξ 21  r2  ξ 21h
potential (φ) (h − r cot β) h−
ξ11 ξ11  2R  ξ11
Contact surface magnetic ξ 31 ξ 31  r2  ξ 31h
potential (ψ) (h − r cot β)  h−
ξ11 ξ11  2R  ξ11

for the conical indenter, and

2η11
p1 (r ) = a2 − r 2 (7.116)
π 2η R

for the spherical indenter. It can be seen that both pressures vanish at the contact
edge r = a, that is, there is no stress singularity there. This is in opposition to the case
of flat-ended cylindrical indenter, for which there is an inverse square singularity
at r = a. We also note that, for the conical indenter, there is a logarithmic stress
singularity at the center of the contact area (Chen et al. 1999), which is clearly
induced by the geometric singularity of the apex of the indenter.
Remark 7.19: It is interesting to note that, for the given electric and magnetic properties
of the indenter, the application of electric or magnetic potential to the indenter will
not change the contact radius or the projected contact area. However, the magnitude
of the imposed electric or magnetic potential will affect the penetration depth of the
indenter. A few more interesting features regarding the indentation responses can
also be observed from Tables 7.1 through 7.4.
Remark 7.20: Having obtained the pressure, electric charge and magnetic flux
distributions as shown in Eq. (7.113), we can further derive the expressions for the
MEE field at any point within the half-space using the method of superposition.
Actually, these can all be expressed in terms of elementary functions and the results
can be found in Appendix D of Chen et al. (2010).
Remark 7.21: Various decoupled and degenerated cases, including indentation on
piezoelectric, piezomagnetic, or elastic materials can be readily analyzed in a similar
way. The reader is also referred to Appendix E of Chen et al. (2010) for a detailed
discussion.
7.8 References 257

7.7 Summary and Mathematical Keys

7.7.1 Summary
This chapter extends the potential function method to the bimaterial space case. In
contrast to the infinite space case considered in Chapter 6, the complete Green’s
function solutions consist of two parts, one corresponding to the infinite space that
has already been obtained in Chapter 6, and the other being complementary. The
latter is obtained using the method of images, which can rightly account for the
effect of interface. Green’s function solutions for a half-space are then obtained as
a special case. All the derivations are actually based on a trial-and-error method (or
inverse method), with only the involved unknown constants determined from the
equilibrium conditions and the continuity or boundary conditions. The 3D MEE fields
are obtained in terms of elementary functions, which are very convenient for use.
The surface Green’s function solution is particularly presented due to its
practical importance. Then, it is used, by the method of superposition, to derive the
indentation responses of three common circular indenters. The analytical results
have been summarized in Tables 7.1 through 7.4 for different cases considering the
electric and magnetic properties of the indenter. All the results are in exact closed-
form, which certainly can be used in the indentation technique for MEE materials.
Various decoupled and degenerated cases can be considered similarly, with some
results given in this chapter, but a few more left for the interested reader as exercises.
For the piezoelectric material case and the purely elastic material case, the reader is
also referred to Ding and Chen (2001) and Ding et al. (2006), respectively.

7.7.2 Mathematical Keys


The method of superposition is applied so that the solution to the bimaterial
space can be expressed by the summation of the infinite-space solution and a
complementary part (or image part) induced to satisfy the interface continuity
condition. The correct expression for the complementary part is found by the trial-
and-error method keeping in mind that the potential functions in this part are all
regular. In application to the indentation problem, the potential theory method by
Fabrikant (1989) is a key.

7.8 References
Bischoff, J. E. 2004. Static indentation of anisotropic biomaterials using axially asymmetric
indenters – a computational study. Journal of Biomechanical Engineering 126 (4):
498–505.
Chen, W. Q. 2000. On piezoelastic contact problem for a smooth punch. International Journal
of Solids and Structures 37 (16): 2331–40.
Chen, W. Q. and Ding, H. J. 1999. Indentation of a transversely isotropic piezoelectric half-
space by a rigid sphere. Acta Mechanica Solida Sinica 12 (2): 114–20.
Chen, W. Q., Lee, K. Y., and Ding, H. J. 2005. On free vibration of non-homogeneous
transversely isotropic magneto-electro-elastic plates. Journal of Sound and Vibration
279 (1–2): 237–51.
Chen, W. Q., Shioya, T., and Ding, H. J. 1999. The elasto-electric field for a rigid conical punch
on a transversely isotropic piezoelectric half-space. Journal of Applied Mechanics 66
(3): 764–71.
258 Green’s Functions in a TI MEE Bimaterial Space

Chen, W. Q., Pan, E., Wang, H. M., and Zhang, C. Z. 2010. Theory of indentation on multiferroic
composite materials. Journal of the Mechanics and Physics of Solids 58 (1): 1524–51.
Ding, H. J. and Chen, W. Q. 2001. Three Dimensional Problems of Piezoelasticity. New York:
Nova Science Publishers.
Ding, H. J., Chen, B., and Liang, J. 1997. On the Green’s functions for two-phase transversely
isotropic piezoelectric media. International Journal of Solids and Structures 34 (28):
3041–57.
Ding, H. J., Chen, W. Q., and Zhang, L. C. 2006. Elasticity of Transversely Isotropic Materials.
Dordrecht: Springer.
Ding, H. J., Hou, P. F., and Guo, F. L. 1999. Elastic and electric fields for elliptical contact for
transversely isotropic piezoelectric bodies. Journal of Applied Mechanics 66 (2): 560–2.
2000. The elastic and electric fields for three-dimensional contact for transversely isotropic
piezoelectric materials. International Journal of Solids and Structures 37 (23): 3201–29.
Ding, H. J., Jiang, A. M., Hou, P. F., and Chen, W. Q. 2005. Green’s functions for two-phase
transversely isotropic magneto-electro-elastic media. Engineering Analysis with
Boundary Elements 29 (6): 551–61.
Fabrikant, V. I. 1989. Applications of Potential Theory in Mechanics: A Selection of New
Results. Dordrecht: Kluwer.
Fischer-Cripps, A. C. 2007. Introduction to Contact Mechanics. 2nd ed. Berlin: Springer.
Galin, L. A. and Gladwell, G. M. L. 2008. Contact Mechanics. Berlin: Springer.
Giannakopoulos, A. E. 2000. Strength analysis of spherical indentation of piezoelectric
materials. Journal of Applied Mechanics 67: 409–16.
Giannakopoulos, A. E. and Parmaklis, A. Z. 2007. The contact problem of a circular rigid
punch on piezomagnetic materials. International Journal of Solids and Structures 44:
4593–612.
Giannakopoulos, A. E. and Suresh, S. 1999. Theory of indentation of piezoelectric materials.
Acta Materialia 47: 2153–64.
Gladwell, G. M. L. 1980. Contact Problems in the Classical Theory of Elasticity. Alphen aan
den Rijn: Sijthoff and Noordhoff.
Hertz, H. 1882. Uber die Beruhrung fester elastischer Korper. Journal für die reine und
angewandte Mathematik 92: 156–71.
Hou, P. F., Leung, A. Y. T., and Ding, H. J. 2003. The elliptical Hertzian contact of transversely
isotropic magnetoelectroelastic bodies. International Journal of Solids and Structures 40
(11): 2833–50.
2008. A point heat source on the surface of a semi-infinite transversely isotropic electro-
magneto-thermo-elastic material. International Journal of Engineering Science 46 (3):
273–85.
Hou, P. F., Teng, G. H., and Chen, H. R. 2009a. Three-dimensional Green’s function for a
point heat source in two-phase transversely isotropic magneto-electro-thermo-elastic
material. Mechanics of Materials 41 (3): 329–38.
Hou, P. F., Yi, T., and Leung, A. Y. T. 2009b. Green’s functions for semi-infinite transversely
isotropic electro-magneto-thermo-elastic material. International Journal of Applied
Electromagnetics and Mechanics 29 (2): 83–100.
Jiang, A. M. and Ding, H. J. 2005. Green’s functions and boundary element method for a
magneto-electro-elastic half-plane. Structural Engineering and Mechanics 20 (2):
259–64.
Jiang, A. M., Ding, H. J., and Wu, G. Q. 2006. Green’s functions for a two-phase infinite
magneto-electro-elastic plane. Multidiscipline Modeling in Materials and Structures 2
(1): 67–82.
Johnson, K. L. 1985. Contact Mechanics. Cambridge: Cambridge University Press.
Kalinin, S. V. and Bonnell, D. A. 2002. Imaging mechanism of piezoresponse force microscopy
of ferroelectric surfaces. Physical Reviews B 65: no. 125408.
Kalinin, S.V., Karapetian, E., and Kachanov, M. 2004. Nanoelectromechanics of piezoresponse
force microscopy. Physical Reviews B 70: no. 184101.
Lur’e, A. I. 1964. Three-Dimensional Problems of the Theory of Elasticity. New York:
Interscience Publishers.
7.8 References 259

Oliver, W. C. and Pharr, G. M. 1992. Improved technique for determining hardness and
elastic modulus using load and displacement sensing indentation experiments. Journal
of Materials Research 7 (6): 1564–83.
Pan, E. 1997. A general boundary element analysis of 2-D linear elastic fracture mechanics.
International Journal of Fracture 88: 41–59.
Pan, E. and Han, F. 2005. Green’s functions for transversely isotropic piezoelectric functionally
graded multilayered half spaces. International Journal of Solids and Structures 42:
3207–33.
Pan, E. and Heyliger, P. R. 2002. Free vibrations of simply supported and multilayered
magneto-electro-elastic plates. Journal of Sound and Vibration 252: 429–42.
Pharr, G. M., Oliver, W. C., and Brotzen, F. R. 1992. On the generality of the relationship
among contact stiffness, contact area, and elastic modulus during indentation. Journal
of Materials Research 7 (3): 613–7.
Ramamurty, U., Sridhar, S., Giannakopoulos, A. E., and Suresh, S. 1999. An experimental
study of spherical indentation on piezoelectric materials. Acta Materialia 47 (8):
2417–30.
Rogowski, B. 2012. A concave indenter on a piezoelectromagneto-elastic substrate or a layer
elastically supported. Journal of Strain Analysis for Engineering Design 47 (6): 362–78.
Rogowski, B. and Kalinski, W. 2012. Indentation of piezoelectromagneto-elastic half-space by
a truncated conical punch. International Journal of Engineering Science 60 (1): 77–93.
Saigal, A., Giannakopoulos, A. E., Pettermann, H. E., and Suresh, S. 1999. Electrical response
during indentation of a 1–3 piezoelectric ceramic-polymer composite. Journal of Applied
Physics 86 (1): 603–6.
Sneddon, I. N. 1948. Boussinesq’s problem for a rigid cone. Proceedings of Cambridge
Philosophical Society 44: 492–507.
Sridhar, S., Giannakopoulos, A. E., Suresh, S., and Ramamurty, U. 1999. Electrical
response during indentation of piezoelectric materials: A new method for material
characterization. Journal of Applied Physics 85 (1): 380–7.
Swadener, J. G. and Pharr, G. M. 2001. Indentation of elastically anisotropic half-spaces by
cones and parabolae of revolution. Philosophical Magazine A 81 (2): 447–66.
Vlassak, J. T. and Nix, W. D. 1994. Measuring the elastic properties of anisotropic materials
by means of indentation experiments. Journal of the Mechanics and Physics of Solids
42 (8): 1223–45.
Wang, J. H., Chen, C. Q., and Lu, T. J. 2009. Indentation responses of piezoelectric films.
Journal of the Mechanics and Physics of Solids 56: 3331–51.
Xiong, S. M. and Ni, G. Z. 2009. 2D Green’s functions for semi-infinite transversely isotropic
electro-magneto-thermo-elastic composite. Journal of Magnetism and Magnetic
Materials 321 (12): 1867–74.
8 Green’s Functions in an Anisotropic
Magnetoelectroelastic Full-Space

8.0 Introduction
Since Lord Kelvin’s solution on the point-force Green’s functions in an isotropic
elastic full-space, seeking the solutions to the point-force Green’s functions in a solid
full-space made of various elastic anisotropic materials, piezoelectric anisotropic
materials, and even anisotropic MEE materials has been always an interesting
topic. The full-space Green’s functions have been developed employing different
mathematical approaches, and this chapter presents some of the common ones
utilized to derive the Green’s function solution in an anisotropic MEE full-space.
The advantages and disadvantages among these approaches and their corresponding
different expressions of the solutions are also discussed for future users.

8.1 Basic Equations in 3D MEE Full-Space


We first briefly review the basic equations for the point-force Green’s function
problem in an anisotropic MEE 3D full-space. These are expressed as

σ iJ = ciJPqγ Pq (8.1)

σ iJ ,i = − f J ( x ) (8.2)

where σiJ, γPq, ciJPq, and fJ are, respectively, the extended stresses, strains, material
coefficients, and forces, as defined in Eqs. (2.4) and (2.7) in Chapter 2.
Replacing fJ by the extended point forces of unit magnitude δJMδ(x) in M-direction
(for the source at the origin without loss of generality), substituting Eq. (8.1) into Eq.
(8.2), and making use of the extended displacement and strain relations (2.3), we
obtain the following differential equations for the extended displacement field

ciJPquPM,qi + δ JM δ ( x) = 0 (8.3)

where uPM is the Green’s extended displacement in the P-direction at the field point x
due to the point force in the M-direction applied at the source y = 0. It should be noted
that P = 1–3 indicates the elastic displacement, P = 4 the electric potential, and P = 5
the magnetic potential. Similar definition holds for the direction of the point force.
260
8.2 Green’s Functions in Terms of Line Integrals 261

8.2 Green’s Functions in Terms of Line Integrals


We first define the following 3D Fourier transform, using the extended displacement
Green’s function as an example,

+∞
u PM ( K ) = ∫ ∫ ∫ uPM ( x )eix⋅ K dx1dx2 dx3 (8.4)
−∞

Applying the 3D Fourier transform (8.4) to Eq. (8.3), we find that in the transformed
domain, Eq. (8.3) becomes

ciJPq Ki Kq u PM (Κ ) = δ JM (8.5)

Solving for the transformed displacements and taking the inverse Fourier
transform, we find that the Green’s displacements in the physical domain can be
expressed as

1 +∞
uPM ( x) ≡
( 2 π )3 ∫ ∫ ∫ −∞
u PM ( x)e − i x⋅ K d K1 d K2 d K3

1 +∞ (8.6)
3 ∫ ∫ ∫ −∞
= (KK )−PM 1
e − i x ⋅ K d K1 d K 2 d K 3
( 2 π)

where we define the (KK) matrix as

(KK ) JP = Ki ciJPq Kq (8.7)

with its inverse being expressed as

APM ( K )
(KK )−PM
1 = (8.8)
D( K )

In Eq. (8.8), D(K) and APM(K) are, respectively, the determinant and adjoint matrix
of (KK)PM. We point out that the adjoint matrix A is different from the Stroh matrix
A used later in this chapter.
Introducing the following variable transform

k = Κ /K, K = Κ
(8.9)
d K 1d K 2 d K 3 = K 2 d Kd S

with dS being the surface element on the unit sphere S in the K-space, the Green’s
displacement (8.6) can be rewritten as

1 +∞
uPM ( x ) =
(2 π)3 ∫0 dK ∫ (kk )−PM
S
1 e − iK x ⋅ k d S ( k ) (8.10)

where we have made use of the following relations

(KK ) JP = K 2 (kk ) JP , (kk ) JP = ki ciJPq kq (8.11)


262 Green’s Functions in an Anisotropic Magnetoelectroelastic Full-Space

Because (kk)JP is an even function of k and the inner integral over k is on the whole
unit sphere surface, Eq. (8.10) can be further written as

1 +∞
uPM ( x) =
2(2 π)3 ∫−∞ dK ∫S (kk)PM
−1 e − iK x⋅k dS(k ) (8.12)

Using the following δ-function relation (Gel’fand et al. 1966)

1 +∞ ± ikx
2 π ∫−∞
δ ( x) = e dk (8.13)

we can reduce Eq. (8.12) to

1
uPM ( x) =
8π 2 ∫S δ( x ⋅ k )(kk)−PM
1 dS ( k ) (8.14)

Before further reducing the area integral in Eq. (8.14) to a line integral, we first find
the derivatives of the Green’s displacements (with respect to the field point), which
are required for finding the extended Green’s strain and stress fields.
Taking the derivative of Eq. (8.14) with respect to the field point x, we obtain

1 dδ( x ⋅ k )
uPM,i ( x ) =
8π 2 ∫S d( x ⋅ k )
−1 dS(k )
ki (kk )PM (8.15)

Denoting the angle between k and x by θ (0 to π), the angle ϕ (0 to 2π) on the L-
plane with its normal being x, and e = x/r, with r being the magnitude of x (Figure 8.1
in Box 8.1), we then have

e ⋅ k = cos θ, d(e ⋅ k ) = − sin θ dθ, dS(k ) = sin θ dθ dϕ (8.16)


Making use of Eq. (8.16), we can change Eqs. (8.14) and (8.15) to

1 1 2π
uPM ( x) =
8π 2 ∫−1δ ( x ⋅ k )d(e ⋅ k )∫0 (kk )−PM
1 dϕ (8.17)

1 1 dδ ( x ⋅ k ) 2π
uPM,i ( x) = ∫
8π r −1 d( x ⋅ k )
2
d( x ⋅ k )∫ ki (kk )−PM
0
1 dϕ (8.18)

As for Eq. (8.18), we integrate it by parts to find


1 ∂[δ ( x ⋅ k ) ∫0 ki (kk ) PM dϕ ]
−1
1
8π 2 r ∫−1
uPM,i ( x) = d( x ⋅ k )
∂( x ⋅ k )

(8.19)
1 1 ∂∫ ki (kk )−PM
1 dϕ

8π 2 r ∫−1
0
− δ ( x ⋅ k) d( x ⋅ k )
∂( x ⋅ k )
8.2 Green’s Functions in Terms of Line Integrals 263

Box 8.1. Basic relations between various unit vectors

It is noted that the L-plane is normal to the vector x (and the unit vector e).
The unit vector m is along the cross line of the L-plane and the (x1, x2)-plane so
that m is perpendicular to both iz and e, and hence also to the vector x′, which
is the projection of x onto the (x1, x2)-plane. As a result, the angle between m
and ix is equal to that between x′ and iy, and equals π/2-Φ. The angle between
the projection vector of k onto the L-plane is denoted by ϕ. Thus we have

e = sinΘ cosΦ ix + sinΘ sin Φi y + cos Θ iz  (F8.1)

m = sinΦix − cos Φi y + 0iz  (F8.2)

n is determined from m and e as

n = e × m = cosΘ cosΦi x + cos Θ sin Φi y − sin Θiz  (F8.3)

If vector k is located within the L-plane, we then have

kx   sin Φ  cosΘ cosΦ 


k = k y  = cosϕ m + sin ϕ n =  −cosΦ  cosϕ +  cosΘ sinΦ  sin ϕ (F8.4)
 kz   0   −sinΘ 

x3

Θ x

L
θ k
e
o x2
m
n x'
ϕ
Φ

x1
Figure 8.1. The L-plane is normal to x, and ϕ is an angle within the L-plane, say,
measured from x1. The other orthonormal system attached to the fixed coordinate
system (o;x1,x2,x3) is m × n = e (e = x/r).
264 Green’s Functions in an Anisotropic Magnetoelectroelastic Full-Space

The first integral is zero due to the delta function feature. The integral from −1 to
1 in the second integral can be carried out. Finally, when evaluating at k·x = 0, the
Green’s displacements and their derivatives can be expressed as

1 2π
uPM ( x ) =
8π r
2 ∫0
(kk )−PM
1 dϕ (8.20)

1 2π ∂ki (kk )−PM


1

8π 2 r 2 ∫0
uPM,i ( x ) = − dϕ (8.21)
∂ (e ⋅ k )

There are two immediate issues associated with Eqs. (8.20) and (8.21) when
implementing them for numerical calculation, as remarked in the following text.
Remark 8.1: The first issue is to deal with the derivatives in Eq. (8.21). To do so, we
first let m and n be any two orthogonal unit vectors on the L-plane that is normal to
x so that m × n = e forms an orthogonal unit vector base (as we carry out the integral
for ϕ). Then the arbitrary integral variable unit vector k can be expressed in terms
of θ and ϕ as

k = cosθ e + sinθ cos ϕ m + sin θ sin ϕ n (8.22)

In Eqs. (8.20) and (8.21), the integral is for ϕ, which is on the L-plane normal to
x. In other words, the value of θ in Eq. (8.22) should be fixed at θ = π/2 after the
operations. For instance, we should have

k = cos ϕ m + sin ϕ n, ∂ki / ∂θ θ =π / 2 = −ei ,


(8.23)
∂(e ⋅ k ) / ∂θ θ =π / 2 = −1, ∂ki / ∂(e ⋅ k ) θ =π / 2 = ei .

where ei is the i-th component of the unit vector e in the fixed coordinate system
(xi). The last expression comes from the second and third ones by the chain rule of
partial derivatives.
Furthermore, making use of

(kk )−PM
1 (kk )
MN = δ PN (8.24)

we have

∂(kk )−PM
1 ∂(kk )QN
= −(kk )−PQ
1 (kk )−NM
1 (8.25)
∂(e ⋅ k ) ∂(e ⋅ k )

Thus Eqs. (8.20) and (8.21) can be finally reduced to

1 π
4π 2 r ∫0
uPM ( x ) = (kk )−PM
1 dϕ (8.26)

1
{ }
π
2 2 ∫0
uPM, i ( x ) = − ei (kk )PM
−1
− ki (kk )PQ
−1
[(ek )QN + (ke)QN ](kk )NM
−1

4π r
   (ek ) ≡ ei ciQNq kq (8.27)
QN
8.3 Green’s Functions in Terms of Stroh Eigenvalues 265

where the integral interval in Eqs. (8.26) and (8.27) is reduced to [0,π] due to the
fact that the involved integrands are even functions of their integral variable ϕ. We
further point out that various high-order derivatives of the Green’s displacements
can be also expressed in terms of the simple line integral. For further details, the
reader is referred to Han (2009).
Remark 8.2: The second issue is on the selection of the unit vector k. We let the unit
vector e along x be (also see Eq. (F8.1) in Box 8.1)

e = sinΘcosΦ ix + sinΘ sin Φ i y + cos Θ iz (8.28)

Then the unit vector k = (kxix+kyiy+kziz), which should be normal to x, can be


expressed as

kx   sin Φ  cosΘ cosΦ 


k =  k y  =  −cosΦ  cosϕ +  cosΘ sinΦ  sin ϕ ≡ cosϕ m + sin ϕ n (8.29)
 kz   0   −sinΘ 

It is noted that Eq. (8.29) further defines the unit vectors m and n as required in
Eqs. (8.22) and (8.23). Then, the matrix (kk)JP, its determinant, and adjoint matrix
can be evaluated for different values of the integral variable ϕ. This finishes the
derivation of the Green’s displacements and their derivatives in terms of a line
integral.

8.3 Green’s Functions in Terms of Stroh Eigenvalues


First, in terms of the unit vectors m and n and the integral variable ϕ, the Green’s
displacements (8.20) can be expressed as

1 π / 2 APM (cos ϕ m + sin ϕ n)


4π 2 r ∫−π / 2 D(cos ϕ m + sin ϕ n)
uPM ( x ) = dϕ (8.30)

In terms of the new variable p defined as follows

ϕ = tan −1 p, or p = tan ϕ (8.31)

we have

1 +∞ APM [(m + pn)]


uPM ( x ) =
4π r
2 ∫−∞ D[(m + pn)]
dp (8.32)

In deriving Eq. (8.32), we have made use of the fact that both A and D are even
functions of their variable p with det(A) being 2-order lower than D.
Therefore, for the given unit vectors m and n (or the field point as connected by
Eqs. (8.28) and (8.29)), Eq. (8.32) can be evaluated using the Cauchy’s residue theory.
We assume that the first five (distinct) roots are those with positive imaginary part
(Im(pL)>0; L = 1,2,3,4,5) and the remaining five are their corresponding conjugates.
Then, D can be expressed as
266 Green’s Functions in an Anisotropic Magnetoelectroelastic Full-Space

10 5
D[(m + pn)] = ∑ aL pL = a10 ∏ ( p − pL )( p − pL ) (8.33)
L= 0 L= 1

It should be pointed out that in the 3D case, the Stroh eigenvalues depend on the
orientation (m,n) as well as on the material properties (Ting 1996), a feature differing
from that in the 2D case studied in Chapter 4.
In terms of the Stroh eigenvalues only, we have therefore, the Green’s
displacements as

Im 5 APM (m + pK n)
uPM ( x ) = − ∑
2π r K =1 5
(8.34)
a10 ( pK − pK ) ∏ ( pK − pL )( pK − pL )
L =1
L≠ K

For the derivatives of the extended Green’s displacements, we follow the same
procedure as for the Green’s displacements. In other words, we first write Eq.
(8.27) as

 ei APM (cos ϕ m + sin ϕ n) 


− 
 D(cos ϕ m + sin ϕ n) 
1 π /2  APQ (cos ϕ m + sin ϕ n)(ek)QN ANM (cos ϕ m + sin ϕ n) 
uPM,i ( x ) =
4π r
2 2 ∫− π / 2  + ki
D2 (cos ϕ m + sin ϕ n) dϕ
 
 APQ (cosϕ m + sin ϕ n)(ke)QN ANM (cos ϕ m + sin ϕ n) 
 + ki D2 (cos ϕ m + sin ϕ n) 
 
 (8.35)

where e (or its components ei) again is the unit vector of the field point x as in Eq.
(8.28), and k (or its components ki) is a vector function as in Eq. (8.29).
Then, we change the integral variable from ϕ to p as in Eq. (8.31), which
gives us

 ei APM (m + pn) 
− 
 D ( m + pn ) 
1 ∞  APQ ( m + p n)(ek )QN ANM ( m + pn) 
4π 2 r 2 ∫−∞ 
uPM,i ( x) = +ki dp (8.36)
D2 (m + pn) 
 APQ (m + pn)(ke)QN ANM (m + pn) 
+
 i k 
 D2 (m + pn) 

where k = m+pn.
The integrals in Eq. (8.36) can be evaluated using the Cauchy’s residue theory,
and can be expressed as

1
uPM,i ( x) = [u(1) ( x) + u(PMi
2)
( x) + u(PMi
3)
( x)] (8.37)
2π r 2 PMi
8.4 Green’s Functions Using 2D Fourier Transform Method 267

where

5
APM (m + pK n)
1)
u(PMi ( x ) = ei Im ∑ 5
(8.38)
K =1
a10 ( pK − pK ) ∏ ( pK − pL )( pK − pL )
L=1
L≠ K

 
 
5  
1 ∂  ki APQ (m + pn)(ek )QN ANM (m + pn) 
(2)
uPMi ( x ) = − Im 2 ∑ (8.39)
a10 J = 1 ∂pJ  5 


∏ ( p − pL )2 ( p − pL )2 

L= 1
 L≠ J 

 
 
5  
1 ∂  ki APQ (m + pn)(ke)QN ANM (m + pn) 
uPMi ( x ) = − Im 2 ∑
( 3)
(8.40)
a10 J = 1 ∂pJ  5 


∏ ( p − pL )2 ( p − pL )2 

L= 1
 L≠ J 

In Eqs. (8.39) and (8.40), the partial derivative ∂/∂pJ means that we first take the
derivative with respect to p and then evaluate it at p = pJ.
Therefore, Eq. (8.34) gives the extended Green’s displacements in terms of
the Stroh eigenvalues, while Eqs. (8.37)–(8.40) are the derivatives of the extended
Green’s displacements also in terms of the Stroh eigenvalues.

8.4 Green’s Functions Using 2D Fourier Transform Method


The 2D Fourier transform method is particularly powerful for deriving the Green’s
functions in half, bimaterial, and multilayered spaces. The solutions in these systems
are usually composed of a singular full-space part and a regular image part. In
problem analysis associated with these systems, therefore, the corresponding full-
space solutions based on the 2D Fourier transform method are more convenient in
many cases. This is the motivation to present the Green’s functions in the full-space
using the 2D Fourier transform method. The corresponding elastic, piezoelectric, and
MEE solutions can be found in Pan and Yuan (2000), Pan and Tonon (2000) and Pan
(2002). The Green’s functions in the MEE full-space presented in the following text
is a compact summary of those in Pan (2002).
We assume that, in a general anisotropic MEE full-space, there is an extended
point force f = (f1, f2, f3, −fe, −fh)t that is applied at the source point y(0,0,d) with the
field point being denoted by x(x1, x2, x3 ≡ z).
In order to solve the problem, we artificially divide the problem domain into two
regions, z > d and z < d, so that these two domain are free of any body source. Across
the source level at z = d, however, the following continuity conditions must hold:
268 Green’s Functions in an Anisotropic Magnetoelectroelastic Full-Space

u J ( x1 , x2 , d + 0) − u J ( x1 , x2 , d − 0) = 0
(8.41a,b)
t J ( x1 , x2 , d + 0) − t J ( x1 , x2 , d − 0) = −δ ( x1 )δ ( x2 ) f J

where tJ is the extended traction at z = constant as defined in Chapter 4. While Eq.


(8.41a) is obviously the continuity condition of the extended displacements, Eq.
(8.41b) can be obtained by integrating the equilibrium Eq. (8.2) over z from z = d–ε
to z = d+ε, and then let ε approach zero. In other words, we have

d +ε d +ε
∫d −ε σ iJ ,i dz + ∫d −ε f J ( x1 , x2 , z)dz = 0 (8.42)

Because the extended body force is a concentrated one at (0,0,d), we then have
(when ε → 0)

d +ε
∫d −ε σ zJ ,zdz + fJ δ ( x1 )δ ( x2 ) = 0 (8.43)

or

[σ zJ ( x1 , x2 , d + ε ) − σ zJ ( x1 , x2 , d − ε )] ε →0 + f J δ ( x1 )δ ( x2 ) = 0 (8.44)

which is the same as Eq. (8.41b) when ε approaches zero.


We define the following 2D Fourier transforms (over the two horizontal
coordinates x1 and x2) of the two-point extended displacement

u I (k1 , k2 , z; y ) = ∫∫ uI ( x1 , x2 , z; y ) eikα xα dx1dx2 (8.45)

where summation from 1 to 2 is assumed over the repeated Greek letter (here it is
for α, unless otherwise stated). We point out that the pair (k1,k2) denotes specially
for the Fourier transform of the in-plane horizontal variables (x1,x2) or (x,y) only (in
this chapter and in Chapter 9). It should be noted that this 2D vector is completely
different from the 3D vector K introduced in Eq. (8.4) for the 3D Fourier transform.
Applying the 2D Fourier transform (8.45) to Eq. (8.3) with zero body source, we
find the following equation in the transformed domain,

cα IKβ kα kβ u K + i(cα IK 3 + c3 IK α )kα u K ,3 − c3 IK 3u K ,33 = 0 (8.46)

where again α, β take the summation from 1 to 2. We now introduce the polar
coordinates (η,θ) that are related to the Fourier variables (k1,k2) as

 k1   m1   cosθ  0 
 k2  = η m , m =  m2  =  sin θ  , n = 0  (8.47)
 0   0   0   1

where η = k12 + k22 , m and n are two orthogonal unit vectors that will enter into the
Stroh formalism.
The general solution of Eq. (8.46) can be expressed as

u (k1 , k2 , z ) = ae− ipηz (8.48)


8.4 Green’s Functions Using 2D Fourier Transform Method 269

with p and a being the eigenvalues and eigenvectors of the following extended Stroh
eigenequation

[Q + p( R + R t ) + p2T ] a = 0 (8.49)

where the superscript t denotes the matrix transpose, and

QIK = c jIKs m j ms ≡ (mm), RIK = c jIKs m j ns ≡ (mn), TIK = c jIKs n j ns ≡ (nn) (8.50)

We also introduce

1
b = ( Rt + pT )a = − (Q + p R)a (8.51)
p

Then the quadratic Stroh eigenrelation (8.49) can be changed into the following
linear Stroh eigenrelation

 N1 N 2  a  a 
N t 
= p  (8.52)
 3 N1  b b

where

N 1 = −T −1 Rt , N 2 = T −1 , N 3 = RT −1 Rt − Q (8.53)

Equation (8.52) is the extended MEE Stroh eigenrelation in the oblique plane spanned
by m and n defined in Eq. (8.47). We point out that the eigenvalues of Eq. (8.52) are
either complex or purely imaginary as discussed in Chapter 4 for the corresponding
2D case. Once the eigenproblem (8.52) is solved, the extended displacements in the
Fourier transformed domain are then obtained from Eq. (8.48).
In order to find the extended stresses in the Fourier transformed domain,
we start with the physical-domain relation. In the physical domain, the extended
traction vector t on the x3 = constant plane and the extended in-plane stress vector s
are related to the extended displacements as

t ≡ (σ13 , σ 23 , σ 33 , D3 , B3 )t
= (c31Kl uK ,l , c32 Kl uK ,l , c33 Kl uK ,l , c34 Kl uK ,l , c35 Kl uK ,l )t (8.54)

s ≡ (σ 11 , σ 12 , σ 22 , D1 , D2 , B1 , B2 )t
= (c11Kl uK ,l , c12 Kl uK ,l , c22 Kl uK ,l , c14 Kl uK ,l , c24 Kl uK ,l , c15Kl uK ,l , c25Kl uK ,l )t
 (8.55)

Taking the Fourier transform, we then find that the transformed extended traction
and in-plane stress vectors can be expressed as

t = − iη be − ipηz (8.56)

s = − iη ce − ipηz (8.57)
270 Green’s Functions in an Anisotropic Magnetoelectroelastic Full-Space

with

c = Ea (8.58)

where the matrix E is defined by

 c111α mα + pc1113 c112α mα + pc1123 c113α mα + pc1133 c114α mα + pc1143 c115α mα + pc1153 
 c121α mα + pc1213 c122α mα + pc1223 c123α mα + pc1233 c124α mα + pc1243 c125α mα + pc1253 
c221α mα + pc2213 c222α mα + pc2223 c223α mα + pc2233 c224α mα + pc2243 c225α mα + pc2253 
E =  c141α mα + pc1413 c142α mα + pc1423 c143α mα + pc1433 c144α mα + pc1443 c145α mα + pc1453 
c241α mα + pc2413 c242α mα + pc2423 c243α mα + pc2433 c244α mα + pc2443 c245α mα + pc2453 
 c151α mα + pc1513 c152α mα + pc1523 c153α mα + pc1533 c154α mα + pc1543 c155α mα + pc1553 
c251α mα + pc2513 c252α mα + pc2523 c253α mα + pc2533 c254α mα + pc2543 c255α mα + pc2553 
 (8.59)

with α taking the summation from 1 to 2.


Because the eigenvalues cannot be real, we can let pm, am, and bm (m = 1, 2, …,
10) be the eigenvalues and the associated eigenvectors, and order them as

Im pJ > 0 , pJ + 5 = pJ , aJ + 5 = aJ , bJ + 5 = bJ ( J = 1 , 2 , 3 , 4, 5)
A = [a1 , a2 , a3 , a4 , a5 ], B = [b1 , b2 , b3 , b4 , b5 ], C = [c1 , c 2 , c 3 , c 4 , c 5 , c6 , c 7 ]
 (8.60)

where Im stands for the imaginary part and the overbar denotes the complex
conjugate. In our analysis in the following text, we assume that pJ are distinct and
that the eigenvectors aJ, and bJ satisfy the following normalization relation (their
other relations are presented later by Eqs. (8.71) and (8.101))

bIt a J + aIt bJ = δ IJ (8.61)

or equivalently in terms of the following matrix form

B t A + At B = I (8.62)

In Eq. (8.61), δIJ is the Kronecker delta of 5 × 5.


Remark 8.3: We point out that the Stroh eigenvalues and eigenmatrices presented
here are similar to those discussed in Chapter 4, except that while in Chapter 4
those values depend only on the material properties, they also depend on the
Fourier transformation variables k1 and k2 in this chapter. In other words, the Stroh
eigenvalues and eigenmatrices are presented in the Fourier-transformed domain.
The general solutions of Eq. (8.46) in the transformed domain can be obtained
by superposing the ten eigensolutions of Eq. (8.52), that is

u (k1 , k2 , z) = iη−1 A < e − ip* ηz > q + iη−1 A < e − ip*ηz > q ′


t(k1 , k2 , z) = B < e − ip* ηz > q + B < e − ip* ηz > q ′ (8.63)
s(k1 , k2 , z) = C < e − ip* ηz > q + C < e − ip* ηz > q ′

where q and q′ are two arbitrary complex vectors to be determined and


8.4 Green’s Functions Using 2D Fourier Transform Method 271

〈 e − ip* ηz 〉 = diag[e− ip1ηz , e − ip2 ηz , e − ip3 ηz , e − ip4 ηz , e − ip5 ηz ] (8.64)

We emphasize again that, besides their obvious dependence on material properties,


the vectors q and q′, and the Stroh eigenvalues pJ and matrices A, B, C, are also
functions of the unit vector m. It is noted that due to the special arrangement on
the eigenvalues pJ in Eq. (8.60), the general solutions associated with the first terms
on the right-hand side of Eq. (8.63) are finite in the half-space z > 0, and the second
terms are finite in the half-space z < 0.
Therefore, for the source at z = d, the solutions in the upper (z > d) and lower
(z < d) half-spaces can be expressed, respectively, as

u (k1 , k2 , z) = iη−1 A < e − ip* η(z − d ) > q


t(k1 , k2 , z) = B < e − ip* η(z − d ) > q (z − d > 0 ) (8.65)
s(k1 , k2 , z) = C < e− ip* η(z− d ) > q

u (k1 , k2 , z) = iη−1 A < e − ip* η(z − d ) > q ′


t(k1 , k2 , z) = B < e − ip* η(z − d ) > q ′ (z − d < 0 ) (8.66)
s(k1 , k2 , z) = C < e− ip* η(z− d ) > q ′

The two arbitrary complex vectors q and q′ can be determined by the continuity
conditions of the extended displacement and traction as described by Eq. (8.41),
which in the Fourier transformed domain, can be expressed as

u (k1 , k2 , d + 0) − u (k1 , k2 , d − 0) = 0
(8.67a, b)
t(k1 , k2 , d + 0) − t(k1 , k2 , d − 0) = − f

Therefore, at the source level z = d, these two vectors q and q′ should satisfy

Aq − Aq ′ = 0
(8.68)
Bq − Bq ′ = − f

where the common nonzero factor iη−1 in the first relation has been removed.
Premultiplying the first relation by B t and the second one by At , we then have

B t Aq − B t Aq ′ = 0
(8.69)
At Bq − At Bq ′ = − At f

Adding these two relations, we have

q = − At f (8.70)

where we have made use of the following two relations

B t A + At B = I
(8.71)
B t A + At B = 0
272 Green’s Functions in an Anisotropic Magnetoelectroelastic Full-Space

Similarly, Eq. (8.68) can be also changed to (by premultiplying Bt and At,
respectively)

B t Aq − B t Aq ′ = 0
(8.72)
At Bq − At Bq ′ = − At f

Adding these two equations gives

q ′ = At f (8.73)

where we have made use of the following two obvious relations, the conjugates of
Eq. (8.71)

B t A + At B = I
(8.74)
B t A + At B = 0

Therefore, we have solved the two arbitrary complex vectors q and q′. It is also
interesting to note that these two complex vectors are not independent, one being
the negative conjugate of the other. Finally, the solution in the Fourier transformed
domain for a point “force” in a general anisotropic MEE space can be expressed as

u (k1 , k2 , z) = − iη−1 A < e − ip* η(z − d ) > q ∞


t(k1 , k2 , z) = − B < e − ip* η(z − d ) > q ∞ (z − d > 0 ) (8.75)
s(k1 , k2 , z) = −C < e − ip* η(z− d ) > q ∞

u (k1 , k2 , z) = iη−1 A < e − ip* η(z − d ) > q ∞


t(k1 , k2 , z) = B < e − ip* η(z − d ) > q ∞ (z − d < 0 ) (8.76)
s(k1 , k2 , z) = C < e − ip* η(z − d ) > q ∞

where

q ∞ = At f (8.77)

The corresponding physical-domain solutions can be found by the double Fourier


inverse transform

 u( x1, x2 , x3 )   iη−1 A
 t ( x1, x2 , x3 )  = −1
2π ∞ 
 ∫0 dθ∫  B <e
− ip* η(z − d ) > e − iη( x1 cos θ + x2 sin θ) ηq ∞ dη
(z > d )
 4π 2 0
 s( x1, x2 , x3 )   C 
 (8.78)

 u( x1, x2 , x3 )   iη−1 A
 t ( x1, x2 , x3 )  = 1
2π ∞
 B  < e − ip* η(z − d ) > e − iη( x1 cos θ + x2 sin θ) ηq ∞dη (z < d )
  4π 2 ∫0 dθ∫
0  
 s( x1, x2 , x3 )   C 
 (8.79)
8.4 Green’s Functions Using 2D Fourier Transform Method 273

These equations can be written as

 u( x1 , x2 , x3 )  −1 2π ∞
 iA 
 t ( x1 , x2 , x3 )  = 2 ∫0 dθ∫  B η < eiη [ − p* (z− d )−( x1 cos θ + x2 sin θ)] > q ∞ dη
0 
(z > d )
 s( x1 , x2 , x3 )  4π 
C η
 (8.80)

 u( x1 , x2 , x3 )  1 2π ∞  iA 
 t ( x1 , x2 , x3 )  = 2 ∫0 dθ∫  B η < eiη[ − p* (z− d )−( x1 cos θ + x2 sin θ)] > q ∞ dη (z < d )
 s( x1 , x2 , x3 )  4π 0 Cη
 
 (8.81)

Carrying out the integral with respect to η, we have


(1) For z > d,

1 2π 1
u( x1 , x2 , x3 ) =
4π 2 ∫0
dθ A <
[− p* (z − d) − ( x1 cosθ + x2 sin θ )]
> q∞

1 2π 1
t ( x1 , x2 , x3 ) = 2 ∫ dθ B < > q∞ (8.82)
4π 0 [ − p* (z − d) − ( x1 cosθ + x2 sin θ )]2
1 2π 1
s( x1 , x2 , x3 ) = 2 ∫ dθ C < > q∞
4π 0 [− p* (z − d) − ( x1 cosθ + x2 sin θ )]2

(2) For z < d,

−1 2π 1
u( x1 , x2 , x3 ) =
4π 2 ∫0
dθ A <
[− p* (z − d) − ( x1 cosθ + x2 sin θ )]
> q∞

−1 2π 1
t ( x1 , x2 , x3 ) = 2 ∫ dθ B < > q∞ (8.83)
4π 0 [ − p* (z − d) − ( x1 cosθ + x2 sin θ )]2
−1 2π 1
s( x1 , x2 , x3 ) = 2 ∫ dθ C < > q∞
4π 0 [ − p* (z − d) − ( x1 cos θ + x2 sin θ )]2

The integral interval from 0 to 2π can be reduced to the interval from 0 to π by making
use of the periodic conditions of the involved integrands. Therefore, we finally have,
in terms of the Green’s components in the I-direction at x1, x2, x3(≡z) due to a general
source in the K-direction at the point y(0, 0, d).
(1) For x3 > y3,

1 π  −1 
uIK ( x1 , x2 , x3 ) =
2π 2 ∫0  A < [ p* ( x3 − y3 ) + ( x1 cosθ + x2 sin θ )] > At  dθ
IK
1 π  −1 
tIK ( x1 , x2 , x3 ) =
2π 2 ∫0  B < [ p* (x3 − y3 ) + (x1 cosθ + x2 sin θ )]2 > At  dθ
(8.84)
IK
1 π  −1 
sIK ( x1 , x2 , x3 ) =
2π 2 ∫0 C < [ p* ( x3 − y3 ) + ( x1 cosθ + x2 sin θ )]2 > At  dθ
IK
274 Green’s Functions in an Anisotropic Magnetoelectroelastic Full-Space

(2) For x3 < y3,

1 π  1 
uIK ( x1 , x2 , x3 ) =
2π 2 ∫0  A < [ p* ( x3 − y3 ) + ( x1 cosθ + x2 sin θ )] > At  dθ
IK
1 π  1 
tIK ( x1 , x2 , x3 ) =
2π 2 ∫0  B < [ p* ( x3 − y3 ) + ( x1 cosθ + x2 sin θ )]2 > At  dθ (8.85)
IK
1 π  1 
sIK ( x1 , x2 , x3 ) =
2π 2 ∫0 C < [ p* ( x3 − y3 ) + ( x1 cosθ + x2 sin θ )]2 > At  dθ
IK

8.5 Green’s Functions in Terms of Radon Transform


Making use of the periodic conditions of the integrand over the unit sphere, Eq.
(8.14) can be rewritten as

1 π π APM (k )
uPM ( x ) =
4π 2 ∫0 dϕ ∫0 D(k )
δ ( x ⋅ k )sin θ dθ (8.86)

We now introduce the following integral variable transform

p = cos θ / sin θ (8.87)

Then, Eq. (8.86) can be changed to (Wang et al. 2006)

1 π +∞ APM ( s)
uPM ( x ) =
4π 2 ∫0
dϕ ∫
−∞ D( s)
δ( x ⋅ s)dp (8.88)

with

s = cos ϕ m + sin ϕ n + pe (8.89)

where (m,n,e) is a set of unit orthogonal base vectors, with e being used later as the
outward normal of a given boundary. To carry out the inner integral, we make use of
the following relation (Gel’fand et al. 1966) (when γ approaches zero)

1 1
= P − i sgn(γ )πδ( x) (8.90)
x + iγ x

where P stands for the principle part of the expression. Equation (8.90) can be
applied to solve δ(x) when γ approaches zero. In other words, we have

−sgn( γ )  1 
δ( x ) = Im  (8.91)
π  x + iγ 

Thus, Eq. (8.88) can be expressed alternatively as

−1 π +∞ APM ( s) sgn( γ )
uPM ( x ) =
4π 2 ∫0 dϕ∫−∞ D( s) x ⋅ s + iγ
dp (8.92)
8.6 Green’s Functions in Terms of Stroh Eigenvalues and Eigenvectors 275

Because the roots for D(s) can only appear in complex pairs, based on the Cauchy’s
residue theory, we can eventually write Eq. (8.92) as (assuming distinct roots only)

π 5 sgn(e ⋅ x )
uJM ( x ) = − Re ∫
0 ∑ g(JMK) (ϕ) s( K ) ⋅ x
dϕ (8.93)
K =1

where

s(K ) = cos ϕ m + sin ϕ n + pK e


1 AJM ( s( K ) )
g (JM
K)
(ϕ ) = (8.94)
2π D, p ( s( K ) )
5
∂D( s)
D, p ( s( K ) ) ≡
∂p
p = pK = a10 ( pK − pK ) ∏ ( p − pq )( p − pq )
q =1; q ≠ K

with a10 in the last expression being the coefficient of the 10-th power in the
polynomial of D(s).
By taking the derivatives of Eq. (8.93) with respect to x, we have

π 5 sgn(e ⋅ x )
uJM,q ( x ) = Re ∫
0 ∑ sq(K ) g(JMK) (ϕ) [ s(K ) ⋅ x]2 dϕ (8.95)
K =1

where sq( K ) indicates the q-component of the vector s( K ) .

8.6 Green’s Functions in Terms of Stroh Eigenvalues and Eigenvectors

8.6.1 General Definitions


We first define again, for a given field point x, e and k as,

e ≡ x / r = sinΘcosΦ ix + sinΘ sin Φ i y + cos Θiz


(8.96a,b)
k = m cos ϕ + n sin ϕ

with

 sin Φ  cos Θ cos Φ 


m =  −cosΦ  , n =  cos ΘsinΦ  (8.97)
 0   −sinΘ 

Equation (8.97) defines two new orthogonal unit vectors m and n in the plane with
e being its normal (forming an orthogonal coordinate system as (m × n=e), see
Figure 8.1).
We first investigate the eigenvalues p and eigenvectors a associated with the
following equation

[Q + p( R + R t ) + p2T ]a = 0 (8.98)

which is identical to Eq. (8.49) with the involved Q, R and T matrices being defined
by Eq. (8.50) but based on the two new orthogonal unit vectors m and n in Eq.
(8.97).
276 Green’s Functions in an Anisotropic Magnetoelectroelastic Full-Space

It is obvious that Eq. (8.98) is in a quadratic form of p. A convenient way to


find the eigenvalues and eigenvectors of Eq. (8.98) is to convert it to the following
equivalent linear eigenequation (as Eq. (8.52)):

N ξ = pξ (8.99)

where

ξ = (a, b)t
−1
b = [(nm) + p(nn)]a = [(mm) + p(mn))]a (8.100)
p
 −(nn) −1 (nm) (nn) −1 
N =
 ( mn)( nn) ( nm) − ( mm) − ( mn)( nn) −1 
−1

There are ten eigenvalues pJ (J = 1 to 10) and ten eigenvectors ξ of Eq. (8.99).
Because N is real, the eigenvalues pI and the corresponding eigenvectors ξI (aI, bI,
and the eigenmatrices A and B) can be ordered in the same way as in Eq. (8.60).

8.6.2 Orthogonal Relations


First, based on the definition of the matrices A and B in Eq. (8.60) and the
normalization relation Eq. (8.62), we have the following orthogonality and closure
relations (Ting 1996)

 Bt At   A A  I 0 
 B t = (8.101a)
At   B B   0 I 

 A A  B t At   I 0 
 B B   B t = (8.101b)
At   0 I 

While Eq. (8.101a) is directly the definition of the orthonormal relation among
eigenvectors, its conjugate pairs, Eq. (8.101b), is obtained from Eq. (8.101a) by
switching the order of the two matrices on the left-hand side.
One important consequence of Eq. (8.101) is that some of the matrices originated
from matrices A and B can be real. For instance, the matrix H defined as follows

H = 2iAAt (8.102)

is real, symmetric and positive definite (Bacon et al. 1979; Ting 1996).

8.6.3 Variation and Integration of Stroh Quantities in the (m,n)-Plane


and the Green’s Functions
The Stroh quantities (Stroh eigenvalues pI, eigenmatrices A and B, and their
functions) considered in this section are those under the integral for ϕ from 0 to 2π
(or from 0 to π as in Eq. (8.26)). Therefore, they might be functions of ϕ as well as the
orientation of the observation point x, that is, e.
8.6 Green’s Functions in Terms of Stroh Eigenvalues and Eigenvectors 277

Assuming that these quantities are functions of ϕ, then one can prove that the
following differential and integral relations hold (Bacon et al. 1979; Ting 1996)


pI = −(1+ pI2 ),
∂ϕ
(8.103a, b)

N = −( I + N 2 )
∂ϕ

1 π
π ∫0
P (ϕ)dϕ = i I ,
(8.104a, b)
1 π S H
∫ N (ϕ)dϕ = 
 L S 
− t
π 0

where S and L are two real matrices, defined as

S = i(2 AB t − I ), L = −2iBB t (8.105)

and P(ϕ) is the diagonal matrix of the eigenvalues pI, as a function of ϕ, that is,

P(ϕ) =< p* (ϕ) >= diag[ p1 (ϕ), p2 (ϕ), p3 (ϕ), p4 (ϕ), p5 (ϕ)] (8.106)

Furthermore, in deriving Eq. (8.104), the following diagonalization of N has been


used (Ting 1996)

  P (ϕ ) 0   Bt At 
N (ϕ ) =  A A   (8.107)
B B  0 P (ϕ )  B t At 

An immediate and important application of the integral relation (8.104b) is that the
extended Green’s function displacement matrix u (with elements uPM ) in the integral
form (8.26) can be expressed using the eigenmatrix A only as

1 π
4π 2 r ∫0
u( x ) = (nn) −1 dϕ
(8.108)
H i
= = A ⋅ At
4 πr 2 πr

Thus, for a given field point x, we can find its unit vector e (and thus m and n). With
these values, we can solve Eq. (8.99) for the eigenmatrix A, which is needed in the
extended displacement Green’s function tensor.

8.6.4 Derivatives of Extended Green’s Displacements


We still refer to Figure 8.1. For a given field point x, the unit vectors of the spherical
coordinate system can be expressed as (using m, n, and e)

m = sinΦix − cosΦi y
n = cos Θ cos Φix + cos Θ sinΦi y − sinΘ iz (8.109)
e = sinΘ cosΦi x + sinΘ sin Φi y + cos Θ iz
278 Green’s Functions in an Anisotropic Magnetoelectroelastic Full-Space

Then the gradient of any function in terms of the spherical coordinate system can
be expressed as

∂f ∂f ∂f
∇f = e +n +m (8.110)
∂r r∂Θ rsinΘ∂Φ

Or in terms of their components,

∂f ∂f ∂f ∂f
= ek + nk + mk (8.111)
∂xk ∂r r∂Θ rsinΘ∂Φ

Therefore, what one needs is just to carry out the derivatives of Eq. (8.108) with
respect to xk. For example, the first derivative of the extended Green’s functions can
be expressed as

∂uIJ 1  ∂r −1 ∂H IJ 
uIJ,k ≡ =  H IJ + r −1 (8.112)
∂xk 4 π  ∂xk ∂xk 

Making use of Eq. (8.111) and the fact that HIJ are functions of Θ and Φ only,
we have

∂uIJ 1  ∂H IJ ∂H IJ 
=  − ek H IJ + nk + mk  (8.113)
∂xk 4πr 2 ∂Θ sin Θ∂Φ 

Thus, in order to find the first derivatives of the extended Green’s displacements,
one needs to find their derivatives with respect to Θ and Φ only. Taking into
consideration Eq. (8.108), we then only need to find the derivatives of the tensor
A∙At with respect to Θ and Φ. In other words, we only need to calculate the following
two expressions:

∂A ⋅ At ∂A t ∂At
= ⋅ A + A⋅
∂Θ ∂Θ ∂Θ (8.114)
∂A ⋅ At ∂A t ∂At
= ⋅ A + A⋅
∂Φ ∂Φ ∂Φ

In order to do so, we need to first find the derivatives of the matrix N with respect
to Θ and Φ.
Noticing that

∂m ∂m
= 0, = cos Θ n + sinΘ e
∂Θ ∂Φ
∂n ∂n
= −e, = − cos Θ m (8.115)
∂Θ ∂Φ
∂e ∂e
= n, = − sin Θ m
∂Θ ∂Φ
8.6 Green’s Functions in Terms of Stroh Eigenvalues and Eigenvectors 279

then we have,

∂N  a11 a12  ∂N  b11 b12 


= = (8.116)
∂Θ a21 a22  , ∂Φ b21 b22 

where the 5×5 submatrices are defined as (with matrix (nm) being defined in Eq.
(8.27))

 ∂(nn) −1   ∂(nn) −1 
a11 =  − (nm) + (nn) −1 (em)  , a12 =  
 ∂Θ   ∂Θ 
 −(me)(nn) −1 (nm) − (mn)(nn) −1 (em) (8.117a)
a21 =  −1  , a = at
 +(mn) ∂(nn) ( nm)  22 11
 ∂Θ 

 ∂(nn)−1   ∂(nn)−1  t
b11 =  − (nm) , b12 =  , b22 = b11
 ∂Φ   ∂Φ 
−1
   b21 = {cos Θ[(nn) − (mm)] + sin Θ(en)}(nn)−1 (nm) + (mn) ∂(nn) (nm) (8.117b)
∂Φ
+ (mn)(nn)−1 {cos Θ[(nn) − (mm)] + sin Θ(ne)}
− cos Θ[(nm) + (mn)] − sin Θ[(em) + (me)]

In Eq. (8.117a, b), the involved derivatives are given as follows

∂(nn) −1 ∂(nn) ∂(nn)


= − (nn) −1 (nn) −1 , = − (ne) − (en) (8.117c)
∂Θ ∂Θ ∂Θ

∂(nn) −1 ∂(nn) ∂(nn)


= − (nn) −1 (nn) −1 , = − cosΘ[( mn) − ( nm)] (8.117d)
∂Φ ∂Φ ∂Φ

First, the Stroh eigenvalue problem (8.99) can be written as functions of the two
angles Θ and Φ (for the I-th eigenvalue)

N (Θ, Φ ) ξ I (Θ, Φ ) = pI (Θ, Φ ) ξ I (Θ, Φ ) (8.118)

where there is no summation over the repeated index I on the right-hand side. As
the convention, if the same index appears on both sides of an equation, there will
be no summation over the repeated one on either side of the equation. Also in Eq.
(8.118), ξI = (aI, bI)t.
Equation (8.118) is the normal, right eigenvalue problem. The corresponding
left eigenvalue problem is

ξtK (Θ, Φ ) JN (Θ, Φ ) = pK (Θ, Φ ) ξtK (Θ, Φ ) J (8.119)

where

J =  0I 0I  (8.120)
 
280 Green’s Functions in an Anisotropic Magnetoelectroelastic Full-Space

is a special switch matrix.


Taking the derivative of Eq. (8.119) with respect to Θ, we have

∂N ∂ξ ∂p ∂ξ
ξI + N I = I ξI + pI I (8.121)
∂Θ ∂Θ ∂Θ ∂Θ

Premultiplying by ξtK J and using (8.119), we have

∂N ∂ξ ∂p ∂ξ
ξtK J ξI + pK ξtK J I = I ξtK J ξI + pI ξtK J I (8.122)
∂Θ ∂Θ ∂Θ ∂Θ

Therefore, if I = K, we have (because ξtI J ξI = 1)

∂pI ∂N
= ξtI J ξI (8.123)
∂Θ ∂Θ

If I≠K, then, we have (because ξtI J ξK = 0 )

∂ ξI ∂N
( pI − pK )ξtK J = ξtK J ξI (8.124)
∂Θ ∂Θ

Multiplying both sides of Eq. (8.124) by ξK, and taking the summation, we have

∂N
10 ξK J
t ξI
10
∂ ξI ∂Θ ξ (8.125)
∑ ξtK J
∂Θ
ξK = ∑ K
K =1 K = 1 ( pI − pK )
K ≠I K ≠I

The left-hand side can be rewritten as

10 10
∂ ξI ∂ξ ∂ξ
∑ ξtK J ∂Θ
ξK = ∑ ξtK J I ξK − ξtI J I ξI
∂Θ ∂Θ
(8.126)
K =1 K =1
K ≠I

Noticing that ∂( ξtI J ξI ) / ∂Θ = 0, we know that the second term on the right-hand
side is zero. As for the first term on the right-hand side, it can be rewritten as

10
∂ ξI ∂ξ
∑ ξtK J ∂Θ
ξK = I
∂Θ
(8.127)
K =1

Therefore we finally have

t ∂N
10 ξK J ξI
∂ ξI ∂Θ ξ
= ∑ K (8.128)
∂Θ K = 1 ( pI − pK )
K≠I

Equations (8.123) and (8.128) are the final results that we have derived so far.
Equation (8.123) is the derivatives of the eigenvalues while Eq. (8.128) is the
derivatives of the eigenvectors.
8.7 Technical Applications of Point-Source Solutions 281

Similar results can be found for their derivatives with respect to Φ (Malen 1971;
Lavagnino 1995):

∂pI ∂N
= ξtI J ξI (8.129)
∂Φ ∂Φ

t ∂N
10 ξK J ξI
∂ ξI ∂Φ ξ
= ∑ K (8.130)
∂Φ K = 1 ( pI − pK )
K≠I

Thus, with these results, the derivatives of the eigenmatrix A (made of eigenvectors aK)
with respect to the coordinate angles Θ and Φ can be finally expressed in exact
closed form. We point out that the derivative of the eigenvalues with respect to
the coordinate angles Θ and Φ is not needed in the calculation of the derivative
of the eigenmatrix A. However, it is required if one would like to take the high-
order derivatives of the Green’s functions with respect to these physical coordinates
(Lavagnino 1995).
Remark 8.4: We summarize the derivatives of u(x) in Eq. (8.108) with respect to x.
The derivatives of any function with respect to the Cartesian coordinates xi are first
converted to those with respect to the spherical coordinates r, Θ and Φ using Eq.
(8.111) so that we arrive at Eq. (8.113). Second, because the eigenmatrix A is made
of ξI = (aI, bI)t, one only needs to find the derivatives of ξI with respect to Θ and Φ.
These are given, respectively, by Eqs. (8.128) and (8.130). Finally, in Eqs. (8.128) and
(8.130), the required derivatives of the matrix N with respect to Θ and Φ are given
by Eq. (8.116).

8.7 Technical Applications of Point-Source Solutions

8.7.1 Couple Force, Dipoles, and Moments


First, as we know that if there is an extended body force fJ in the domain V, then the
induced displacement is

uK ( x sp ) = ∫ uJK ( x sp ; xqf ) f J ( xqf )dV ( xqf ) (8.131)


V

Notice that in the general case (heterogeneous or inhomogeneous materials),


the integral is with respect to the field point of the two-point Green’s function
displacement.
For a homogenous space with body force fJ, the source and field points can be
switched and so are the indices between the extended displacement components and
extended force direction (between J and K). Thus, Eq. (8.131) can be expressed as

uK ( x ) = ∫ uJK ( x; y) f J ( y)dV ( y) (8.132)


V

For concentrated forces f (i) applied at y(i) (i = 1 to N), the induced displacements
are
282 Green’s Functions in an Anisotropic Magnetoelectroelastic Full-Space

(a) (b)
f f
δ
δ
δ
δ y
y
f f

Figure 8.2. A dipole with a net moment in (a) and a moment-free dipole in (b).

N
uK ( x ) = ∑ uJK ( x; y (i ) ) f J(i) (8.133)
i =1

Remark 8.5: Couple force and dipole “tensor.” For two concentrated forces f and −f
(same magnitude but opposite in direction) applied at y + δ and y – δ, respectively
(see Figure 8.2), the induced displacement is then

uK ( x ) = uJK ( x; y + δ) f J − uJK ( x; y − δ) f J (8.134)

Or, it can be rewritten as

uJK ( x; y + δ) − uJK ( x; y − δ)
uK ( x ) = 2 f J δ i (8.135)
2δ i

Taking the limit with δ approaching zero and letting 2fJδi = TJi be finite constants,
then Eq. (8.135) becomes

uK ( x ) = TJ i uJK,i ( x; y ) (8.136)

It is noticed that T is not a tensor because it is not a square matrix. However, for
the purely elastic case, T becomes a tensor, and it is further called a dipole tensor,
with the following physical meanings (Tji): when the force couple (or simply, a single
couple, or a couple) and their arm (or the dipole axis) are in the same direction (i.e.,
i = j), we then have a moment-free dipole (see Figure 8.2b); otherwise (i.e., i ≠ j),
there is a net moment from the couple force (see Figure 8.2a).

8.7.2 Relations among Dislocation, Faulting, and Force Moments


We first recall that the extended displacements induced by a dislocation with
component bJ applied on an internal surface ∑ with normal ni can be expressed as
(see Eq. (2.22) in Chapter 2)

uK ( x sp ) = ∫ σ iJK ( x sp ; xqf )bJ ( xqf )ni ( xqf )dΣ ( xqf ) (8.137)


Σ

It is noted that the kernel function in Eq. (8.137) is the Green’s stress with component
(iJ) at the field point xqf due to a point force in the K-th direction applied at x sp . The
8.7 Technical Applications of Point-Source Solutions 283

dislocation bJ is also regarded as the relative displacements across the inner surface
(i.e., ΔuJ). The point-force-induced stresses can be also expressed in terms of the
displacement gradients using the constitutive relations (2.5) in Chapter 2 as

σ iJK ( x sp ; xqf ) = ciJLquLK,q ( x sp ; xqf ) (8.138)

Thus, from Eq. (8.137), we have

uK ( x sp ) = ∫ ciJLquLK,q ( x sp ; xqf )bJ ( xqf )ni ( xqf )dΣ ( xqf ) (8.139)


Σ

Defining the moment density “tensor” (moment “tensor” per unit area) as

mLq ( xqf ) = ciJLq bJ ( xqf )ni ( xqf ) (8.140)

we then can express Eq. (8.139) as

uK ( x sp ) = ∫ mLq ( xqf )uLK,q ( x sp ; xqf )dΣ ( xqf ) (8.141)


Σ

Because uLK are the solutions due to the point force, uLK,q can be considered as the
solutions corresponding to the couple force (point force in the K-direction with a
force arm in the q-direction). In other words, the moment tensor and the dislocation
pair (bJni) are equivalent to each other.
It is noted that mLq is in general not a square tensor and thus, it is not symmetric.
However, for the purely elastic but anisotropic case, we have mlq = mql. In this case,
in general, nine force couples are needed to equivalently express the faulting or
dislocation in an anisotropic elastic medium. For the purely elastic case, while the nine
different pairs of the dislocation source (nibj) in Eq. (8.140) are shown in Figure 8.3,
the corresponding force couples or moment tensors are shown in Figure 8.4.
Furthermore, for the isotropic elastic case, we have

(
mij = λbk nk δ ij + µ bi n j + bj ni ) (8.142)

Let us look at a special case of the moment tensor where bi is parallel to the
dislocation plane with normal ni (i.e., the tangential displacement discontinuity
without opening). This corresponds to a slip or shear dislocation as

(
mij = µ bi n j + bj ni ) (8.143)

Equation (8.143) contains the following two examples.


Example 1. In Eq. (8.143), if we further let n = (0,0,1)t and b = (b1,0,0)t, we
then have

 0 0 µ b1 
m= 0 0 0  (8.144)
 µ b1 0 0 
284 Green’s Functions in an Anisotropic Magnetoelectroelastic Full-Space

3 3 3

(1,1) (1,2) b (1,3) b


2 2 2
nb n
n

1 1 1

3 3 3

(2,1) n (2,2) nb (2,3) b


2 2 2
n
b

1 1 1

3 3 3

(3,1) n (3,2) n (3 ,3 ) nb

2 2 2
b
b
1 1 1
Figure 8.3. The nine dislocation source pairs (nibj) in elasticity.

3 3 3

[1,1] [1,2] [1,3]

1 2 1 2 1 2

3 3 3

[2,1] [2,2] [2,3]

1 2 1 2 1 2

3 3 3

[3,1] [3,2] [3,3]

1 2 1 2 1 2

Figure 8.4. The moment components or the nine force couples (or dipoles) mlq in elasticity.
8.7 Technical Applications of Point-Source Solutions 285

(a) (b)
x3
x2
x3 T
P
b1

x1 x1
b1

Figure 8.5. (a) Schematics of a shear fault source (or a glide dislocation) in the (x1,x2)-plane
with the dislocation b1 along the x1-direction. (b) The equivalent double (force) couple,
corresponding to the moment tensor given in Eq. (8.144). In (b), the P- and T-axes are also
shown, representing the compressive and tensile quadrants.

(a) x3 (b)
x2 x3
x2
b3

x1 x1

b3
Figure 8.6. (a) An opening fault source in the (x1,x2)-plane (i.e., n = (0,0,1)t) with the dislocation
b3 along the x3-direction. (b) The equivalent couple system (three dipoles, with the third one
in x3-direction having a different magnitude), corresponding to the moment tensor given in
Eq. (8.145).

This moment density tensor consists of the force double-couple. In other words, a
slip along the fault or a sliding dislocation can be equivalently expressed by a double
force couple as shown in Figure 8.5.
Example 2. In Eq. (8.142), if n and b are in the same direction, for instance, n =
(0,0,1)t and b = (0,0,b3)t, then we have

 λ b3 0 0 
m= 0 λ b3 0  (8.145)
 0 0 (λ + 2 µ )b3 

This shows that an opening fault or climbing dislocation in the x3-direction can be
equivalently expressed by the summation of three dipoles with the weight ratio
respectively being 1, 1, and (λ + 2μ)/λ, as shown in Figure 8.6. It is pointed out that if
the three dipoles have the same magnitude, this moment tensor (8.145) will represent
an explosive source.

8.7.3 Equivalent Body Forces of Dislocations


First, from Betti’s reciprocal theorem, we have

uK ( x sp ) = ∫ [σ iJ ( xqf )uJK ( x sp ; xqf ) − σ iJ


K ( x s ; x f )u ( x f )]n ( x f )dS( x f )
p q J q i q q
S
(8.146)
+ ∫ uJK ( x sp ; xqf ) f J ( xqf )dV ( xqf )
V
286 Green’s Functions in an Anisotropic Magnetoelectroelastic Full-Space

Thus, considering the body-force only, the induced displacements are

uK ( x sp ) = ∫ uJK ( x sp ; xqf ) f J ( xqf )dV ( xqf ) (8.147)


V

Then, comparing to the expression of the dislocation-induced displacement (2.23),


we can find that the equivalent body-force of the dislocation is


fL (η) = − ∫ ciJLp bJ ( ξ)ni ( ξ) δ(η − ξ)dΣ( ξ) (8.148)
Σ ∂η p

where η is a field point in the whole problem domain V, and ξ is a point on the
dislocation surface.
Substituting Eq. (8.148) into Eq. (8.147), we have


uK ( x sp ) = − ∫ uLK ( x sp ; x p )∫ ciJLp bJ (ξ p )ni (ξ p ) δ ( x p − ξ p )dΣ (ξ p )dV ( x p ) (8.149)
V Σ ∂x p

Transferring the derivation of the delta function to that of the Green’s displacement
(i.e., by integral by parts), and further making use of the delta function properties,
we have

∂ K s
uK ( x sp ) = ∫ ciJLp bJ (ξ p )ni (ξ p ) u ( x p ; ξ p )dΣ (ξ p ) (8.150)
Σ ∂ξ p L

which is exactly Eq. (8.139).


Remark 8.6: In deriving Eq. (8.148), we have made use of the following equation
between the Green’s function and its derivative:

∂ K s ∂
uL ( x ; ξ) = − ∫ uLK ( x s ;η) δ(η− ξ)dV (η) (8.151)
∂ξ p V ∂η p

Remark 8.7: We present examples of the equivalent force (8.148). Let us consider
the purely elastic and isotropic case. If we further assume that the normal vector
n and the dislocation vector b are in the x3- and x1-directions, respectively, that is,
n = (0,0,1)t and b = (b1,0,0)t, then we have from Eq. (8.148) the equivalent forces of
this glide dislocation (or discontinuity) as (c3113=c3131=μ)


f1 (η) = − ∫ µ( ξ)b1 ( ξ)n3 ( ξ) δ(η− ξ)dΣ( ξ)
Σ ∂η3
f2 (η) = 0 (8.152)

f3 (η) = −∫ µ( ξ)b1 ( ξ)n3 ( ξ) δ(η− ξ)dΣ( ξ)
Σ ∂η1

The first expression in Eq. (8.152) is clearly equivalent to a single couple force
system in which the forces are distributed on both sides of the plane η3 = 0, with
the forces along the x1-direction (one side is positive and the other side is negative)
8.8 Numerical Examples of Dislocations 287

η3
η2

η1

Figure 8.7. Equivalence between an elastic dislocation and a single force couple as given in
the first expression of Eq. (8.152).

(Aki and Richards 1980). This is illustrated in Figure 8.7. As for the third expression
in Eq. (8.152), because the directions of n and b are exchangeable (see Eq. (8.150),
except for index J > 3), one may think of it as an equivalent system where n and b
are in the x1- and x3-directions, respectively, that is, n = (1,0,0)t and b = (0,0, b3)t. In
this case, we will have a single double-force system on both sides of the plane η1 =
0, with the forces along the x3-direction (one side is positive and the other side is
negative).

8.8 Numerical Examples of Dislocations


We consider a uniform dislocation on a unit circle located in the (x2,x3)-plane (or in
the (y,z)-plane for easy presentation) with fixed normal n = (1,0,0)t (Figure 8.8). The
uniform Burgers vectors are respectively b = b (1,0,0,0,0)t, b = b (0,1,0,0,0)t, b = b
(0,0,1,0,0)t. The 3D full-space is made of the following three materials:
Case 1: Transversely isotropic piezoelectric (TI Piezoelectric) material with its
symmetry axis along the z-axis. The material is the poled lead-zirconate-titanate
(PZT-4) ceramic from Dunn and Taya (1993), with its material property listed in
Table 2.1 of Chapter 2.
Case 2: Transversely isotropic elastic (TI Elastic) with its material property
obtained from Case 1 by just letting the coupling coefficients eijk = 0 and neglecting
the electric field.
Case 3: Isotropic elastic (ISO Elastic) with its material property being obtained
from Case 2 by using the following conditions (this gives the so-called equivalent
isotropic moduli)

( c44 + c55 + c66 )


c44 = c55 = c66 =
3

( c11 + c22 + c33 )


c11 = c22 = c33 = (8.153)
3

c12 = c13 = c23 = c11 − 2c44


288 Green’s Functions in an Anisotropic Magnetoelectroelastic Full-Space

r
θ
a
y
b

Figure 8.8. A uniform dislocation vector b of magnitude b is applied within the circle of radius
a on the (y,z)-plane. The induced fields are drawn along the circumference of a large circle
with radius r = 5b.

On the (y,z)-plane, we draw the response on the circumference of the circle r ≡


√(y2 + z2) = 5b for the following field quantities (total displacement, hydrostatic
stress, and effective stress, respectively)

u = ux2 + uy2 + uz2


σ xx + σ yy + σ zz
σh =
3
1
σe = [(σ xx − σ yy )2 + (σ yy − σ zz )2 + (σ zz − σ xx )2 ] + 3(σ xy
2 +σ 2 +σ 2 )
yz xz
2 (8.154)

Figures 8.9, 8.10, and 8.11 show, respectively, the total elastic displacement,
the hydrostatic stress, and the effective stress along the circumference of the
circle r =5b induced by a uniform Burgers vector of b = (b,0,0)t, b = (0,b,0)t, or
b = (0,0,b)t. The Burgers vector is applied within the circle of radius a = 0.5b. While
the elastic displacement is normalized by b, the stresses are normalized by a/b and
are in GPa. In each figure, we compare the variations of the same field quantity
along the circumference of the circle for the three different material cases (i.e., TI
Piezoelectric, TI elastic, and ISO elastic). The results show clearly that material
anisotropy and piezoelectric coupling could substantially affect the dislocation-
induced displacement and stress fields.

8.9 Summary and Mathematical Keys

8.9.1 Summary
In this chapter, the full-space Green’s functions (including the extended displacements
and stresses) are derived using various mathematical approaches. These include
the line-integral expression (from 0 to π) obtained through the direct 3D Fourier
transform, the explicit solutions in terms of the Stroh eigenvalues and involved
matrices, the line-integral expression (from 0 to π) obtained using the 2D (in-plane)
8.9 Summary and Mathematical Keys 289

(a) 0.03
TI piezoelectric
TI elastic
Normalized displacement u/b

ISO elastic

0.02

0.01

0
0 90 180 270 360
Degree
(b) 0.02
TI piezoelectric
TI elastic
Normalized displacement u/b

ISO elastic
0.015

0.01

0.005

0
0 90 180 270 360
Degree
(c) 0.006
TI piezoelectric
TI elastic
Normalized displacement u/b

ISO elastic

0.004

0.002

0
0 90 180 270 360
Degree
Figure 8.9. Total elastic displacement along the circumference of the circle r = 5b induced by
a uniform Burgers vector of b = (b,0,0)t in (a), b = (0,b,0)t in (b), and b = (0,0,b)t in (c), applied
within the circle of radius a = 0.5b. The elastic displacement is normalized by the magnitude
of the Burgers vector b.

Fourier transform, the line-integral expression (from 0 to π) obtained through the


Radon transform, and the explicit solutions in terms of the Stroh eigenvectors.
Because these Green’s function expressions possess different advantages and
disadvantages, they can be applied to different problems. For instance, the explicit
Green’s functions in terms of the Radon transform is very suitable for carrying out
an integral (either line or area) with respect to the source or field points, as will be
shown in Chapter 9. The Green’s functions in terms of the 2D Fourier transform
are required in deriving the corresponding half-space, bimaterial, or layered Green’s
functions.
290 Green’s Functions in an Anisotropic Magnetoelectroelastic Full-Space

(a) 0.6
TI piezoelectric
TI elastic
ISO elastic
Hydrostatic stress (GPa) 0.4

0.2

−0.2
0 90 180 270 360
Degree
(b) 0.4
TI piezoelectric
TI elastic
ISO elastic
Hydrostatic stress (GPa)

0.2

−0.2

−0.4
0 90 180 270 360
Degree
(c) 0.2
TI piezoelectric
TI elastic
ISO elastic
Hydrostatic stress (GPa)

0.1

−0.1

−0.2
0 90 180 270 360
Degree

Figure 8.10. The hydrostatic stress along the circumference of the circle r = 5b induced by a
uniform Burgers vector of b = (b,0,0)t in (a), b = (0,b,0)t in (b), and b = (0,0,b)t in (c), applied
within the circle of radius a = 0.5b. The hydrostatic stress is normalized by a/b and is in GPa.

In deriving the solutions, we have assumed in most cases that the source is at the
origin. If the source is located at point y, one needs just to replace x in the solutions
in this chapter by x−y because the full-space Green’s functions depend only upon the
relative distance between the field and source points. Furthermore, the derivatives
of these Green’s functions with respect to the source point are just opposite to those
with respect to the field point.

8.9.2 Mathematical Keys


Mathematical keys in this chapter include the 2D Fourier transform, the Radon
transform, and the Stroh formalism. The Green’s functions presented in different
8.10 Appendix A: Some Basic Mathematical Formulations 291

(a) 2
TI piezoelectric
TI elastic
ISO elastic
1.5
Effective stress (GPa)

0.5

0
0 90 180 270 360
Degree
(b) 1.2
TI piezoelectric
TI elastic
1 ISO elastic
Effective stress (GPa)

0.8

0.6

0.4

0.2
0 90 180 270 360
Degree

(c) 0.8
TI piezoelectric
TI elastic
ISO elastic
0.6
Effective stress (GPa)

0.4

0.2

0
0 90 180 270 360
Degree
Figure 8.11. The effective stress along the circumference of the circle r = 5b induced by a
uniform Burgers vector of b = (b,0,0)t in (a), b = (0,b,0)t in (b), and b = (0,0,b)t in (c), applied
within the circle of radius a = 0.5b. The effective stress is normalized by a/b and is in GPa.

forms should be mathematically equivalent and a suitable form can be selected for
attacking the given problems.

8.10 Appendix A: Some Basic Mathematical Formulations


The adjoint of a square matrix A ≡ [aij] is defined as the transpose of the matrix [Cij]
where Cij is the cofactor of the element aij. The cofactor is calculated as follows.
The minor of A’s entry aij, also known as the i,j, or (i,j), or (i,j)th minor of A, is
denoted by Mij and is defined to be the determinant of the submatrix obtained by
removing from A its i-th row and j-th column. It then follows:
292 Green’s Functions in an Anisotropic Magnetoelectroelastic Full-Space

Cij = (−1)i + j Mij (A1)

and Cij is called the cofactor of aij, also referred to as the i,j, (i,j) or (i,j)th
cofactor of A.
The Cauchy’s residue theorem: Suppose C is a positively oriented, simple closed
contour in the complex plane z. If f is analytic on and inside C except for a finite
number of singular points zk (k = 1, 2, …, n), then

n
∫ C f (z)dz = 2πi ∑ Res( f , zk ) (A2)
k =1

where Res(f,zk) denotes the residue of f at zk. For example, if f(z) is an analytic
function within C, then,
f (z)
∫ C z − z1 dz = 2πif (z1 ) (A3)

f (z)
∫ C (z − z )
1
2
d z = 2 π i f ′(z1 ) (A4)

8.11 References
Aki, K. and Richards, P. G. 1980. Quantitative Seismology: Theory and Methods. San Francisco:
W. H. Freeman and Company.
Bacon, D. J., Barnett, D. M., and Scattergood, R. O. 1979. Anisotropic continuum theory of
lattice defects. Progress in Materials Science 23: 51–262.
Dunn, M. L. and Taya, M. 1993. An analysis of piezoelectric composite materials containing
ellipsoidal inhomogeneities. Proceedings of the Royal Society of London A 443: 265–87.
Gel’fand, I. M., Graev, M. I., and Vilenkin, N. Y. 1966. Generalized Functions, Vol. 5. New York:
Academic Press.
Han, X. 2009. High-order derivatives of Green’s functions in magneto-electro-elastic materials.
International Journal of Solids and Structures 46: 3405–11.
Lavagnino, A. M. 1995. Selected Static and Dynamic Problems in Anisotropic Linear Elasticity.
PhD diss., Stanford University.
Malen, K. 1971. A unified six-dimensional treatment of elastic Green’s functions and
dislocations. Physica Status Solidi B 44: 661–72.
Pan, E. 2002. Three-dimensional Green’s functions in anisotropic magneto-electro-elastic
bimaterials. Zeitschrift für angewandte Mathematik und Physik 53: 815–38.
Pan, E. and Tonon, F. 2000. Three-dimensional Green’s functions in anisotropic piezoelectric
solids. International Journal of Solids and Structures 37: 943–58.
Pan, E. and Yuan, F. G. 2000. Three-dimensional Green’s functions in anisotropic bimaterials.
International Journal of Solids and Structures 37: 5329–51.
Ting, T. C. T. 1996. Anisotropic Elasticity: Theory and Applications. New York: Oxford
University Press.
Wang, C.Y., Denda, M., and Pan, E. 2006. Analysis of quantum-dot-induced strain and electric
fields in piezoelectric semiconductors of general anisotropy. International Journal of
Solids and Structures 43: 7593–608.
9 Green’s Functions in an
Anisotropic Magnetoelectroelastic
Bimaterial Space

9.0 Introduction
One of the advanced material structures is the laminated composite in which each
layer is assigned with different material properties in different orientations. In
so doing, the effective material properties of the composite can be significantly
improved. In such a composite structure, material property mismatch and interface
effect need to be carefully studied and thus designed. In this chapter, we present
the 3D Green’s functions due to the extended point force in a general anisotropic
MEE bimaterial space. Besides the solutions for the well-known perfect interface
condition case, solutions to various imperfect interface conditions are derived
as well. The reduced half-space cases under a variety of surface conditions are
also presented. As applications, the extended point-force Green’s functions are
utilized to solve the 3D inclusion problem, which is closely related to 3D quantum
dot (QD) semiconductor structures. Finally, recent progress on the corresponding
dislocation solutions is reviewed where the 3D dislocation loop can be in various
shapes.

9.1 Problem Description


We now consider an anisotropic and MEE bimaterial full-space where x3 > 0 and
x3 < 0 are occupied by Materials 1 and 2, respectively, with the interface at x3 = 0.
Without loss of generality, we first assume that an extended point force f = (f1, f2,
f3, −fe, −fh)t is applied in Material 1 at the source point y ≡(0,0,y3 ≡ d), with the field
point being denoted by x ≡ (x1,x2,x3 ≡ z) (Figure 9.1). In order to solve the problem,
we here artificially divide the problem domain into three regions: z > d (in Material
1), 0 ≤ z < d (in Material 1), and z < 0 (in Material 2) so that every domain is free of
any body source and therefore the balance equations (8.3) with the corresponding
material properties hold in absence of the body force term. We will need to find
the solution in each domain making use of the interface conditions at z = 0, the
source-level conditions at z = d, and the regular condition of the solution when z
approaches infinity.

293
294 Green’s Functions in an Anisotropic Magnetoelectroelastic Bimaterial Space

x3

Interface between •
Materials 1 and 2 f ≡ (f1, f2, f3, –fe, –fh)t
x2

Material1 (x3>0)
Material 2 (x3<0)

x1
Figure 9.1. A bimaterial space under extended point forces f = (f1, f2, f3, −fe, −fh)t applied in
Material 1 at source point y(0,0,y3≡d).

9.2 Solutions in Fourier Domain for Forces in Material 1


Making use of the full-space solutions in Eqs. (8.63) and (8.77), the general solutions
in the Fourier transformed domain, which satisfy the conditions (8.43) at the source
level and the regular condition as z approaches infinity, can be derived as follows
(Ting 1996; Pan and Yuan 2000).
For z > d (in Material 1):

( 1) ( 1)
u (1) (k1 , k2 , z; y) = − iη−1 A(1) 〈 e − ip* η(z − d ) 〉q ∞ − iη−1 A(1) 〈 e − ip* ηz 〉q (1)
( 1) ( 1)
t (1) (k1 , k2 , z; y) = − B(1) 〈e − ip* η(z− d ) 〉q ∞ − B(1) 〈 e− ip* ηz 〉q (1) (9.1)
s(1) (k1 , k2 , z; y) = −C (1) 〈 e − ip*( 1) η(z − d ) 〉q ∞ − C (1) 〈 e − ip*(1) ηz 〉q (1)

For 0 ≤ z < d (in Material 1):

( 1) ( 1)
u (1) (k1 , k2 , z; y) = iη−1 A(1) 〈 e − ip* η(z − d ) 〉q ∞ − iη−1 A(1) 〈e − ip* ηz 〉q (1)
( 1) ( 1)
t(1) (k1 , k2 , z; y) = B(1) 〈 e− ip* η(z− d ) 〉q∞ − B(1) 〈 e− ip* ηz 〉q (1) (9.2)
( 1) ( 1)
s(1) (k1 , k2 , z; y) = C (1) 〈 e − ip* η(z− d ) 〉q ∞ − C (1) 〈 e− ip* ηz 〉q (1)

For z < 0 (in Material 2):

( 2)
u (2) (k1 , k2 , z; y) = iη−1 A(2) 〈e − ip* ηz 〉q(2)
(2)
t(2) (k1 , k2 , z; y) = B( 2) 〈 e − ip* ηz 〉q( 2) (9.3)
(2)
s( 2) (k1 , k2 , z; y) = C ( 2) 〈 e − ip* ηz 〉q( 2)

where, in accordance with Eq. (8.77),

q ∞ = ( A(1) )t f (9.4)

The complex vectors q (1) and q( 2) in Eqs. (9.1)–(9.3) are to be determined by the
interface conditions. First, along the interface z = 0, we have the following expressions
9.2 Solutions in Fourier Domain for Forces in Material 1 295

for the displacement and traction vectors on both material sides (Material 1 in z > 0
and Material 2 in z < 0):
( 1)
u (1) (k1 , k2 , 0; y) = iη−1 A(1) 〈 eip* ηd 〉q ∞ − iη−1 A(1) q (1)
( 1)
(9.5)
t(1) (k1 , k2 , 0; y) = B(1) 〈 eip* ηd 〉q ∞ − B(1) q (1)

u ( 2) (k1 , k2 , 0; y) = iη−1 A( 2) q( 2)
(9.6)
t( 2) (k1 , k2 , 0; y) = B( 2) q( 2)

Because there will be no mixed conditions involving both the displacement and
traction, one does not need to worry about the prefactor iη−1 before the displacement.
This argument is also true for the half-space case with homogeneous boundary
conditions as will be discussed later.
Perfect interface: We first assume that across the interface, the extended displacement
and traction vectors are continuous. In other words, we have the so-called perfect
interface at z = 0

u( x1 , x2 , +0) − u( x1 , x2 , −0) = 0
(9.7)
t ( x1 , x2 , +0) − t ( x1 , x2 , −0) = 0

In the Fourier transformed domain, we have

u (+0) − u ( −0) = 0, t (+0) − t ( −0) = 0 (9.8)

Therefore the complex vectors q (1) and q( 2) are required to satisfy the following two
vector conditions (Ting 1996; Pan and Yuan 2000)
( 1)
A(1) 〈 eip* ηd 〉q ∞ − A(1) q (1) = A( 2) q( 2)
(9.9)
B(1) 〈 e ip*(1) ηd 〉q ∞ − B(1) q (1) = B( 2) q( 2)

To determine the complex vectors q (1) and q(2) in Eq. (9.9) we first rewrite it as
( 1)
A( 2) q( 2) + A(1) q (1) = A(1) 〈eip* ηd 〉q ∞
(9.10a, b)
B( 2) q( 2) + B(1) q (1) = B(1) 〈e ip*(1) ηd 〉q ∞

Equation (9.10b) can be rewritten as


( 1)
iM ( 2) ( A( 2) q( 2) ) − iM (1) ( A(1) q (1) ) = iM (1) A(1) 〈 eip* ηd 〉q ∞ (9.11)

or
( 1)
M ( 2) ( A( 2) q( 2) ) − M (1) ( A(1) q (1) ) = M (1) A(1) 〈eip* ηd 〉q ∞ (9.12)

with M (α ) being the impedance tensors defined as

M (α ) = − iB(α ) ( A(α ) )−1 (α = 1, 2) (9.13)


296 Green’s Functions in an Anisotropic Magnetoelectroelastic Bimaterial Space

Therefore, the solution of the complex vectors q (1) and q( 2) can be found by using Eqs.
(9.10a) and (9.12), similar to solving a regular linear algebraic system of equations
with two unknowns.
For the perfect interface case (interface conditions (9.7)), the solution for the
complex vectors q (1) and q( 2) has the following simple expression
( 1)
q (1) = G1 〈 eip* ηd 〉q ∞
(9.14)
q ( 2 ) = G2 〈 e ip*(1) ηd 〉q ∞

where the matrices G1 and G2 are given by

G1 = −( A(1) )−1 ( M (1) + M (2) )−1 ( M (1) − M (2) ) A(1)


(9.15)
G2 = ( A(2 ) )−1 ( M (1) + M (2) )−1 ( M (1) + M (1) ) A(1)

Remark 9.1: The first term in Eqs. (9.1) and (9.2) is the Fourier-domain Green’s
function for the anisotropic MEE full-space. Its corresponding physical-domain
solution can be found in Chapter 8 in terms of various forms. Thus, the Fourier
inverse transform needs to be carried out only for the second term of the solutions,
which we call the complementary part of bimaterial Green’s functions.

9.3 Solutions in Physical Domain for Forces in Material 1


Applying the Fourier inverse transform, the extended Green’s displacement in Eq.
(9.1) becomes (replace z by x3 and d by y3)

i
4π 2 ∫∫
( 1)
u(1) ( x1 , x2 , x3 ) = − { η−1 A(1) 〈 e − ip* η( x3 − y3 ) 〉q ∞ e − i( x1k1 + x2 k2 ) }dk1dk2
(9.16)
i
− 2 ∫∫ { η−1 A(1) 〈e − ip* ηx3 〉q (1) e − i( x1k1 + x2k2 ) }dk1dk2
( 1)

Again, the first integral in Eq. (9.16) corresponds to the extended Green’s
displacement in the full-space, which has been discussed in Chapter 8. Consequently,
the inverse transform needs to be carried out only for the second regular integral,
or the complementary part. The singularities involved in the bimaterial Green’s
function appear only in the full-space solution, which can be evaluated easily because
of its simple and explicit-form expression. Denoting the full-space Green’s function
in Material 1 by u∞ ( x1 , x2 , x3 ) and recalling the polar coordinate transform defined
in Eq. (8.47), Eq. (9.16) is then reduced to

u(1) ( x1 , x2 , x3 ) = u∞ ( x1 , x2 , x3 )
i 2π ∞
− 2  ∫ dθ∫ A(1) 〈e − ip* ηx3 〉G1 〈 eip* ηy3 〉e− iη( x1 cos θ + x2 sin θ) ( A(1) )t dη f
( 1) ( 1)

4π   0 0 
 (9.17)

Because the matrices A(1) and G1 are independent of the radial variable η, the
integral with respect to η can actually be performed analytically (i.e., Pan and Yuan
9.3 Solutions in Physical Domain for Forces in Material 1 297

2000). Assuming that x3 ≠ 0 or y3 ≠ 0, Eq. (9.17) can be reduced to a compact form.


Furthermore, noticing the following relations of the involved functions (Ting 1996)

pJ (θ + π) = − p J (θ), G1 (θ + π) = −G1 (θ), G2 (θ + π) = G2 (θ)


A(θ + π) = γ A(θ), B(θ + π) = − γ B(θ), C (θ + π) = − γ C (θ) (9.18)
(for γ = i, or γ = − i)

the regular integral over [0, 2π] can actually be reduced to the interval of [0, π].
That is

2π π
∫0 g(θ )dθ = 2 ∫ g(θ )dθ
0
(9.19)

where g(θ) stands for the integrands in Eq. (9.17) and other similar expressions.
Therefore, the regular line integral is actually over the interval [0, π].
Assuming that x3 ≠ 0 or y3 ≠ 0, the 5 × 5 Green’s displacement matrix uIK, with its
first index I for the extended displacement component at x and the second K for the
extended point-force direction at y, is found to be (in Material 1) (by further replacing
the source point by y = (y1,y2,y3) and denoting uIK for uIK as used in Chapter 8)

∞ ( x; y ) + u c ( x; y )
u IK ( x; y) = uIK IK (9.20)

1  π (1) (1) (1) t 


2π 2  ∫0
c ( x; y ) =
u IK A Gu ( A ) dθ (9.21)
 IK

and

(G1 )IJ
(Gu(1) )IJ = (9.22)
− pI(1) x3 + pJ(1) y3 − [( x1 − y1 )cos θ + ( x2 − y2 )sin θ]

In Eq. (9.20), uIK∞ ( x; y ) again denotes the Green’s displacement matrix for the

extended displacements in the full-space with Material 1 (see Eqs. (8.84) and (8.85)
in Chapter 8), and in Eqs. (9.20)–(9.22), the indices I and J take the range from 1 to
5. There is no summation over the repeated indices I and J on the right-hand side.
Similarly, the bimaterial Green’s functions for the extended stresses (traction and
in-plane stress) can be derived as follows.
In Material 1,

∞ ( x; y ) + t c ( x; y )
t IK ( x; y) = t IK IK
1  π (1) (1) (1) t 
t IK ( x; y) = 2 ∫ B Gt ( A ) dθ
c
2π  0  IK
(9.23)
sIK ( x; y) = sIK ∞ ( x; y ) + s c ( x; y )
IK
1  π
C (1)Gt(1) ( A(1) )t dθ 
2 π 2  ∫0
sIKc ( x; y ) =
 IK
298 Green’s Functions in an Anisotropic Magnetoelectroelastic Bimaterial Space

where

(G1 ) IJ
(Gt(1) )IJ = (9.24)
{− pI(1) x3 + p J(1) y3 − [( x1 − y1 )cos θ + ( x2 − y2 )sin θ]}2

In Eq. (9.23), tIK∞ ( x; y ) and s ∞ ( x; y ) are the Green’s functions for the extended
IK
stresses in the full-space with Material 1 (see Eqs. (8.84) and (8.85) in Chapter 8), and
the first index I in s ranges from 1 to 7 (see Eq. (8.55)). Again, there is no summation
over the repeated index I on the right-hand side.
In Material 2, the Green’s functions contain only the complementary parts,
which are

1  π (2) ( 2) (1) t 
2π 2  ∫0
c ( x; y ) = −
uIK A Gu ( A ) dθ
 IK
c ( x; y ) = − 1 
π
B (2)Gt(2) ( A(1) )t dθ 
2π 2  ∫0
tIK (9.25)
 IK
1 π
 C (2)G (2) ( A(1) )t dθ
2 π 2  ∫0
c ( x; y ) = −
sIK t  IK

with

(G2 )IJ
(Gu(2) )IJ = (9.26)
− pI(2) x3 + pJ(1) y3 − [( x1 − y1 )cos θ + ( x2 − y2 )sin θ)]

(G2 )IJ
(Gt(2) )IJ = (9.27)
{− pI(2) x3 + pJ(1) y3 − [( x1 − y1 )cos θ + ( x2 − y2 )sin θ]}2

There is no summation over the repeated indices I and J on the right-hand side.
Therefore, in Material 1, the bimaterial Green’s function is expressed as a sum of
the explicit full-space Green’s function and a complementary part in terms of a line
integral over [0, π]. In Material 2, the bimaterial Green’s function is expressed in
terms of a line integral over [0, π]. Concerning the complicated nature of the problem
and the final concise expression for the bimaterial Green’s function, it is concluded
that the extended Stroh formalism is truly mathematically elegant and numerically
powerful (Ting 1996). Furthermore, with regard to these physical-domain bimaterial
Green’s functions (Eqs. (9.20), (9.21), (9.23) and (9.25)), the following important
observations can be made, with some of them being similar to those made in Pan
and Yuan (2000).
Remark 9.2: For the complementary part (or the image part) of the solution
in Material 1 and the solution in Material 2, the dependence of the solutions on
the field point x and source point y appears only through matrices Gu(1), Gt(1), Gu( 2),
and Gt( 2) defined in Eqs. (9.22), (9.24), (9.26), and (9.27).
Remark 9.3: The integrals in Eqs. (9.21), (9.23), and (9.25) are regular if x3 ≠ 0 or y3 ≠
0. In other words, there will be no singularity within the integral interval. Thus these
integrals can be easily carried out by a standard numerical integral method such as
the Gaussian quadrature.
9.4 Solutions in Physical Domain with Imperfect Interface 299

Remark 9.4: If x3 ≠ 0 and y3 = 0, the bimaterial Green’s function is still mathematically


regular although some of its components may not have a direct and apparent physical
meaning when the interface is imperfect (see Dundurs and Hetenyi 1965, for the
purely elastic counterpart).
Remark 9.5: When the field and source points are both on the interface (i.e., x3 = y3 =
0), the bimaterial Green’s function is then reduced to the interfacial Green’s function.
For this special case, the line integral involved in the Green’s function expression
becomes singular and the resulting finite-part integral needs to be handled with
special cares (Pan and Yang 2003a).

9.4 Solutions in Physical Domain for Forces in Material 1 with


Imperfect Interface Conditions
The solution procedure developed in previous sections can be utilized to find the
bimaterial Green’s functions with other imperfect but homogeneous interface
conditions. The only requirement on the interface conditions is that the extended
traction and displacement cannot enter simultaneously into the same interface
condition. We choose the following two imperfect interface conditions as
illustrations.

9.4.1 Imperfect Interface Type 1


We assume that the mechanical displacement (and electric potential) and traction
(normal electric displacement) vectors across the interface are continuous and that
the magnetic potential is zero on the interface, that is, on z = 0,

u (J1) = u (J2) , t (J1) = t (J2) ( J = 1 − 4)


(9.28)
u5(1) = u5( 2) =0

It is seen that physically this interface is magnetically closed (Alshits et al. 1994) or it
is a magnetic wall, and is a very common case in electromagnetic studies (see, Papas
1988 or Volakis et al. 1998).
The interface conditions (9.28) give the following interface equations from
which the complex vectors q (1) and q( 2) are to be solved:
For J = 1–4, the first four rows in each of the following two vector equations
(also Eq. (9.9)) hold
( 1)
A(1) 〈 eip* ηd 〉q ∞ − A(1) q (1) = A( 2) q( 2)
(9.29)
B(1) 〈 e ip*(1) ηd 〉q ∞ − B(1) q (1) = B( 2) q( 2)

For J = 5, the fifth rows in each of the following two vector equations hold
( 1)
A(1) 〈 eip* ηd 〉q ∞ − A(1) q (1) = 0
(9.30)
0 = A( 2) q( 2)
300 Green’s Functions in an Anisotropic Magnetoelectroelastic Bimaterial Space

By simple additions and subtractions, Eq. (9.30) can be written equivalently as (again
for the fifth rows only)
( 1)
A(1) 〈 eip* ηd 〉q ∞ − A(1) q (1) = A( 2) q( 2)
(9.31)
A(1) 〈 e ip*(1) ηd 〉q ∞ − A(1) q (1) = − A( 2) q( 2)

Referring to Eq. (9.29) for J = 1–4, and Eq. (9.31) for J = 5, one can symbolically write
the interface conditions similar to the prefect interface case. These equations are

(1)
 (1) 〈 eip*(1) ηd 〉q ∞ − A
A  ( 2) q( 2)
 q (1) = A
(1) (9.32)
 (1) 〈 eip*(1) ηd 〉q ∞ − B
B   ( 2) q( 2)
q (1) = B

The structure of this linear system for the complex vectors q (1) and q( 2) is now
similar to that for the prefect interface case (9.9). Consequently, the solution for the
complex vectors q (1) and q( 2) should also have a similar structure as for the perfect
interface case if certain modified Stroh matrices are introduced. For this imperfect
(α ) (α)
interface case considered here, the modified Stroh matrices A and B in Eq.
(9.32) are defined as

 (α ) = A(α )
A (α = 1, 2) (9.33)

(1) ( 2)
 B11 B12 B13 B14 B15   B11 B12 B13 B14 B15 
(1)  B21 B22 B23 B24 B25  ( 2)  B21 B22 B23 B24 B25 

B =  B31 B32 B33 B34 B35  
, B =  B31 B32 B33 B34 B35 
 B41 B42 B43 B44 B45   B41 B42 B43 B44 B45 
 A51 A52 A53 A54 A55   − A51 − A52 −A53 − A54 − A55 
 (9.34)

Consequently, the solution for the complex vectors q (1) and q( 2) for this imperfect
interface model can also be expressed by Eq. (9.14) and the matrices G1 and G2 by
Eq. (9.15) but replacing the Stroh matrices A and B by the modified ones as defined
previously by Eqs. (9.33) and (9.34).

9.4.2 Imperfect Interface Type 2


This imperfect interface is relatively complicated. Here we assume that the two half-
spaces are mechanically in smooth contact at the interface, while both the electrical
and magnetic potentials are zero along the interface:

u3(1) = u3( 2) , t3(1) = t3( 2)


tα(1) = tα( 2) = 0 (α = 1, 2) (9.35)
u (J1) = u (J2) =0 ( J = 4, 5)
The interface conditions in the transformed domain can be written as follows.
For J = 1,2, the first and second rows in each of the following two vector
equations hold
9.4 Solutions in Physical Domain with Imperfect Interface 301

( 1)
B(1) 〈 eip* ηd 〉q ∞ − B(1) q (1) = 0
(9.36)
0 = B ( 2 ) q( 2 )

which can be equivalently written (by simple addition and subtraction) as


( 1)
B(1) 〈 eip* ηd 〉q ∞ − B(1) q (1) = B( 2) q( 2)
( 1)
(9.37)
B(1) 〈 eip* ηd 〉q ∞ − B(1) q (1) = −B( 2) q( 2)

For J = 3, the third rows in each of the following two vector equations hold
( 1)
A(1) 〈 eip* ηd 〉q ∞ − A(1) q (1) = A( 2) q( 2)
(9.38)
B(1) 〈 e ip*(1) ηd 〉q ∞ − B(1) q (1) = B( 2) q( 2)

For J = 4,5, the fourth and fifth rows in each of the following two vector
equations hold
( 1)
A(1) 〈 eip* ηd 〉q ∞ − A(1) q (1) = 0
(9.39)
0 = A( 2)q( 2)

which can be equivalently written as


( 1)
A(1) 〈 eip* ηd 〉q ∞ − A(1) q (1) = A( 2) q( 2)
(9.40)
A(1) 〈 e ip*(1) ηd 〉q ∞ − A(1) q (1) = − A( 2) q( 2)

Therefore, combining Eqs. (9.37), (9.38), and (9.40), the final interface conditions can
be also written as in Eq. (9.9), but with the following modified Stroh matrices:

(1) (2)
 B11 B12 B13 B14 B15   B11 B12 B13 B14 B15 
 B21 B22 B23 B24 B25   B21 B22 B23 B24 B25 
 (1)
A =  A31 A32 A33 A34 A35   ( 2)
, A =  A31 A32 A33 A34 A35 
 A41 A42 A43 A44 A45   A41 A42 A43 A44 A45 
 A51 A52 A53 A54 A55   A51 A52 A53 A54 A55 
 (9.41)

(1) ( 2)
 B11 B12 B13 B14 B15   − B11 − B12 − B13 − B14 − B15 
(1)  B21 B22 B23 B24 B25   − B21 − B22 − B23 − B24 − B25 
B =  B31 B32 B33 B34 B35  ,  ( 2)
B =  B31 B32 B33 B34 B35 
 A41 A42 A43 A44 A45   − A41 − A42 − A43 − A44 − A45 
 A51 A52 A53 A54 A55   − A51 − A52 − A53 − A54 − A55 
 (9.42)

Thus, the solution for the complex vectors q (1) and q( 2) for this imperfect interface
model can also be expressed by Eq. (9.14) with the matrices G1 and G2 being related
to the modified Stroh matrices defined by Eqs. (9.41) and (9.42).
302 Green’s Functions in an Anisotropic Magnetoelectroelastic Bimaterial Space

9.5 Special Case: Upper Half-Space under General Surface Conditions


Solutions to this half-space case can be regarded as the special case of the
bimaterial Green’s functions with the source point in Material 1. For example, the
general solutions in the Fourier transformed domain, which satisfy the conditions
at the source level and the condition at infinity, can be derived as (Pan and Yuan
2000):
For z > d:

u (k1 , k2 , z; y) = − iη−1 A〈e− ip* η(z− d ) 〉q ∞ − iη−1 A〈 e− ip*ηz 〉q


t(k1 , k2 , z; y) = − B 〈e− ip* η(z− d ) 〉q ∞ − B 〈e− ip* ηz 〉q (9.43)
s(k1 , k2 , z; y) = −C 〈e− ip* η(z− d ) 〉q ∞ − C 〈 e− ip*ηz 〉q

For 0 ≤ z< d:

u (k1 , k2 , z; y) = iη−1 A〈e− ip* η(z− d ) 〉q ∞ − iη−1 A〈e− ip*ηz 〉q


t(k1 , k2 , z; y) = B 〈e− ip* η(z− d ) 〉q ∞ − B 〈e− ip* ηz 〉q (9.44)
s(k1 , k2 , z; y) = C 〈e− ip* η(z− d ) 〉q ∞ − C 〈e− ip* ηz 〉q

where

q ∞ = At f (9.45)

and

〈 e − ip* ηz 〉 = diag[e − ip1ηz , e − ip2 ηz , e − ip3 ηz , e − ip4 ηz , e − ip5 ηz ] (9.46)

and q is an unknown complex vector. Equations (9.43) and (9.44) are just the same
as Eqs. (9.1) and (9.2) except that the superscript (1) there has been omitted here.
We can actually solve the half-space Green’s functions under very general
but homogeneous surface conditions at z = 0 as in Eq. (4.37) of Chapter 4 for the
2D case

Iu u + I t t = 0 (9.47)

with Iu and It being the 5×5 diagonal matrices satisfying Eq. (4.38).
We now need to determine q for the given general boundary conditions (9.47).
This is achieved by defining a new complex matrix K of 5×5 as

K = Iu A + I t B (9.48)

a suitable combination of the Stroh eigenmatrices A and B that are properly coupled
with the boundary conditions. It is noted that Eq. (9.48) for 3D is different than
Eq. (4.41) for 2D in such that while A and B in 2D depend only upon the material
property, A and B in 3D depend further on θ, the Fourier transformed angle.
For example, for the case in which the surface is free from extended tractions,
the matrix K has the following expression:
9.5 Special Case: Upper Half-Space under General Surface Conditions 303

 B11 B12 B13 B14 B15 


 B21 B22 B23 B24 B25 
K =  B31 B32 B33 B34 B35  (9.49)
 B41 B42 B43 B44 B45 
 B51 B52 B53 B54 B55 

If the condition on the surface is mechanically traction free, but electrically and
magnetically conducting (i.e., both the electric and magnetic potentials are zero),
the matrix K is:

 B11 B12 B13 B14 B15 


 B21 B22 B23 B24 B25 
K =  B31 B32 B33 B34 B35  (9.50)
 A41 A42 A43 A44 A45 
 A51 A52 A53 A54 A55 

We now can solve the unknown complex vector q for the half-space case. On the
boundary of the half-space z = 0, we have

u (k1 , k2 , 0; y) = iη−1 A〈eip* ηd 〉q ∞ − iη−1 Aq


(9.51)
t(k1 , k2 , 0; y) = B 〈eip* ηd 〉q ∞ − Bq

Notice that on the boundary z = 0, either the extended displacement or extended


traction components are zero (and there are no mixed boundary conditions between
the displacement and traction as required by Eq. (9.47)). Therefore, in the following
ten equations reduced from Eq. (9.51), five of them are zero.

0 = A〈eip* ηd 〉q ∞ − Aq
(9.52)
0 = B 〈eip* ηd 〉q ∞ − Bq

Premultiplying the first set by Iu and the second set by It to make them identical with
the boundary conditions, and then adding them together we have

0 = (I u A + I t B)〈 eip* ηd 〉q ∞ − (I u A + I t B)q (9.53)

Therefore, with the new matrix K, the complex vector q for all different sets of
boundary conditions (9.47) can be expressed in a simple vector equation as

q = K −1 K 〈eip* ηd 〉 At f (9.54)

Equation (9.54) is a very surprising simple result.


The physical-domain Green’s functions are similar to those in Material 1 with
the matrix G1 in Eq. (9.14) being replaced by

G1 = K −1 K (9.55)
304 Green’s Functions in an Anisotropic Magnetoelectroelastic Bimaterial Space

9.6 Bimaterial Space with Extended Point Forces in Material 2


For this case and in the Fourier transformed domain, we assume that the solutions in
the three subdomains are in the following forms.
For z > 0 (in Material 1):
( 1)
u (1) (k1 , k2 , z; y) = − iη−1 A(1) 〈 e − ip* ηz 〉q(1)
( 1)
t(1) (k1 , k2 , z; y) = − B(1) 〈 e− ip* ηz 〉q(1) (9.56)
( 1)
s(1) (k1 , k2 , z; y) = − C (1) 〈 e− ip* ηz 〉q(1)

For d < z ≤ 0 (in Material 2):


(2) (2)
u (2) (k1 , k2 , z; y) = − iη−1 A(2) 〈e − ip* η(z − d ) 〉q ∞ + iη−1 A(2) 〈e − ip* ηz 〉q( 2)
(2) (2)
t( 2) (k1 , k2 , z; y) = − B( 2) 〈e − ip* η(z − d ) 〉q ∞ + B( 2) 〈 e− ip* ηz 〉q(2) (9.57)
(2) ( 2)
s (2) (k1 , k2 , z; y) = −C (2) 〈 e− ip* η(z− d ) 〉q ∞ + C ( 2) 〈 e− ip* ηz 〉q( 2)

For z < d (in Material 2):


( 2) (2)
u (2) (k1 , k2 , z; y) = iη−1 A(2) 〈 e − ip* η(z− d ) 〉q ∞ + iη−1 A(2) 〈 e − ip* ηz 〉q( 2)
( 2) (2)
t( 2) (k1 , k2 , z; y) = B( 2) 〈 e− ip* η(z − d ) 〉q ∞ + B( 2) 〈 e − ip* ηz 〉q( 2) (9.58)
(2) (2)
s( 2) (k1 , k2 , z; y) = C ( 2) 〈 e− ip* η( z− d) 〉q ∞ + C ( 2) 〈 e − ip* ηz 〉q(2)

and

q ∞ = ( A(2) )t f (9.59)

The interface conditions at z=0 in the perfect interface case give


(2)
A(1) q (1) = A( 2) 〈eip* ηd 〉q ∞ − A( 2) q( 2)
(9.60)
B(1) q (1) = B( 2) 〈e ip*( 2 ) ηd 〉q ∞ − B ( 2 ) q ( 2 )

or,
(2)
A(1) q (1) + A( 2) q( 2) = A( 2) 〈eip* ηd 〉q ∞
(9.61)
B(1) q (1) + B(2) q(2) = B(2) 〈e ip*( 2 ) ηd 〉q ∞

Thus, for the perfect interface case, the solution for the complex vectors q (1) and q( 2)
has the following simple expression
(2)
q (1) = G1 〈 eip* ηd 〉( A(2) )t f
(9.62)
q( 2) = G2 〈 e ip*( 2 ) ηd 〉( A(2) )t f

where, the matrices G1 and G2 are given by

G1 = ( A(1) )−1 ( M (1) + M (2) )−1 ( M (2) + M (2) ) A(2)


(9.63)
G2 = ( A(2) )−1 ( M (1) + M (2) )−1 ( M (1) − M (2) ) A( 2)
9.6 Bimaterial Space with Extended Point Forces in Material 2 305

Assuming that z ≠ 0 or d ≠ 0, the 5×5 Green’s function matrix uIK, with its first index I
for the extended displacement component and the second K for the extended point-
force direction, is found to be:
In Material 1 (z > 0), the Green’s functions contain only the complementary
part, as (by further replacing the source point by y(y1,y2,y3))

1  π (1) (1) ( 2) t 
2 π 2  ∫0
c ( x; y ) =
uIK A Gu ( A ) dθ
 IK
c ( x; y ) = 1 
π
B (1)Gt(1) ( A( 2) )t dθ 
2 π 2  ∫0
tIK (9.64)
 IK
c ( x; y ) = 1 
π
C (1)Gt ( A( 2) )t dθ 
(1)
sIK
2π 2  ∫0  IK

(G1 )IJ
(Gu(1) )IJ = (9.65)
− pI(1) x3 + p J(2) y3 − [( x1 − y1 )cos θ + ( x2 − y2 )sin θ]

(G1 )IJ
(Gt(1) )IJ = (9.66)
{− pI(1) x3 + pJ(2) y3 − [( x1 − y1 )cos θ + ( x2 − y2 )sin θ]}2

There is no summation over the repeated indices I and J on the right-hand side of
Eqs. (9.65) and (9.66).
In Material 2 (z < 0), the Green’s functions are the summation of the full-space
one and the complementary part,

∞ ( x; y ) + u c ( x; y )
uIK ( x; y) = uIK IK
1  π
A(2)Gu( 2) ( A( 2) )t dθ 
2π 2  ∫0
c ( x; y ) = −
uIK
 IK
∞ ( x; y ) + t c ( x; y )
tIK ( x; y) = tIK IK
1  π (2) (2) (2) t  (9.67)
tIK ( x; y) = − 2 ∫ B Gt ( A ) dθ
c
2π  0  IK

sIK ( x; y ) = sIK ( x; y ) + sIK ( x; y )
c

c ( x; y ) = − 1 
π
sIK ∫ C (2)Gt( 2) ( A( 2) )t dθ 
2π  0
2   IK

where

(G2 )IJ
(Gu( 2) )IJ = (9.68)
− pI(2) x3 + pJ(2) y3 − [( x1 − y1 )cos θ + ( x2 − y2 )sin θ]

(G2 )IJ
(Gt( 2) )IJ = (9.69)
{− pI(2) x3 + pJ(2) y3 − [( x1 − y1 )cos θ + ( x2 − y2 )sin θ]}2

Again, there is no summation over the repeated indices I and J on the right-hand side
∞ ( x; y ) denotes the Green’s function matrix
of Eqs. (9.68) and (9.69). In Eq. (9.67), uIK
for the extended displacements in the full-space with Material 2 (Pan and Tonon
306 Green’s Functions in an Anisotropic Magnetoelectroelastic Bimaterial Space

2000), and again in Eqs. (9.65), (9.66), (9.68), (9.69) the indices I and J take the range
∞ ( x; y ) and
from 1 to 5. Also, the first index I in s ranges from 1 to 7. In Eq. (9.67), tIK
∞ ( x; y ) are the Green’s functions for the extended stresses in the full-space with
sIK
Material 2 (Pan and Tonon 2000; see also Eqs. (8.84) and (8.85) in Chapter 8).
Remark 9.6: The solutions given are for the perfect interface conditions. The Green’s
functions corresponding to different imperfect interface conditions can be derived
similarly as in the previous section for the case when the source is in Material 1.
The only change is to the Stroh matrices A and B. Furthermore, the solution for the
source in Material 2 can be obtained from that for the source in Material 1 by the
following transforms: (1) Switch the material indices between 1 and 2; (2) Replace xi
by −xi, and yi by −yi.

9.7 Special Case: Lower Half-Space under General Surface Conditions


For the half-space case in the lower half-space z < 0, the solutions in the Fourier
transformed domain can be expressed as
For d < z ≤ 0:

u (k1 , k2 , z; y) = − iη−1 A〈e− ip* η(z− d ) 〉q ∞ + iη−1 A〈 e− ip*ηz 〉q


t(k1 , k2 , z; y) = − B 〈e− ip* η(z− d ) 〉q ∞ + B 〈e− ip* ηz 〉q (9.70)
s(k1 , k2 , z; y) = −C 〈e− ip* η(z− d ) 〉q ∞ + C 〈 e− ip*ηz 〉q

For z < d:

u (k1 , k2 , z; y) = iη−1 A〈 e− ip* η(z− d ) 〉q ∞ + iη−1 A〈 e− ip*ηz 〉q


t(k1 , k2 , z; y) = B 〈e− ip* η(z− d ) 〉q ∞ + B 〈e− ip* ηz 〉q (9.71)
s(k1 , k2 , z; y) = C 〈e− ip* η(z− d ) 〉q ∞ + C 〈e− ip* ηz 〉q

and

q ∞ = ( A)t f (9.72)

For the surface z = 0 of the half-space, five of the following ten boundary
conditions hold

0 = A〈eip* ηd 〉q ∞ − Aq
(9.73)
0 = B 〈eip* ηd 〉q ∞ − Bq

Premultiplying the first set by Iu and the second set by It, and then adding them
together we have

0 = ( I u A + I t B)〈e ip* ηd 〉q ∞ − ( I u A + I t B)q (9.74)

Therefore, with the new matrix K defined in Eq. (9.48), the complex vector q for
all different sets of boundary conditions (9.47) can be expressed in a simple vector
equation as
9.8 Technical Application: QDs in Piezoelectric Semiconductors 307

q = K −1 K 〈 eip* ηd 〉 At f (9.75)

Finally, we need only to replace G2 in Eq. (9.14) by

G2 = K −1 K (9.76)

to obtain the physical-domain solutions in the lower half-space under general surface
conditions. Again, we point out that, the solution to the lower half-space domain can
be also obtained by simply replacing xi by −xi, and yi by −yi in the solution for the
upper half-space case.

9.8 Technical Application: Quantum Dots in Anisotropic Piezoelectric


Semiconductors
We recall Eqs. (4.81) and (4.82) for the extended displacements and strains due to
the extended eigenstrains within a QWR, which actually hold also for the 3D case.
In other words, we assume that there is a 3D QD inclusion of general shape within
an anisotropic MEE medium. Then the extended displacements and strains due to
the extended eigenstrain within the QD can be expressed, respectively, by Eqs. (4.81)
and (4.82).
We point out that an efficient way to carry out the surface integrals in Eqs. (4.81)
and (4.82) is to approximate the boundary of the QD inclusion by a number of
flat triangles, and carry out the integral over the triangle exactly. With increasing
triangle numbers, the approximate solution will approach the accurate one. Thus our
goal is to find the exact closed-form solution of the integrals on the right-hand sides
of Eqs. (4.81) and (4.82) over a flat triangle. This method is particularly efficient
and accurate for the surface of the inclusion made of piecewise polygonal shapes.
Figure 9.2a shows that the surface of a cubic QD is exactly represented by twelve
flat triangles while Figure 9.2b is the approximation of the surface of a spherical
inclusion by ninety-six flat triangles.
In the following sections, we present the analytical integral expressions of the
extended displacements and strains over a flat triangle of the 3D QD in different
problem domains.

Figure 9.2. The surface of a cubic inclusion represented by twelve flat triangles in (a) and the
surface of a spherical inclusion approximated by ninety-six flat triangles in (b).
308 Green’s Functions in an Anisotropic Magnetoelectroelastic Bimaterial Space

9.8.1 Analytical Integral over Flat Triangle, the Anisotropic


MEE Full-Space Case
Extending the full-space Green’s function solutions by Wang et al. (2006) to MEE
materials and substituting the results into Eq. (4.81), we can find the induced extended
displacements by a QD inclusion as follows (vector n and angle θ here are, respectively,
identical to the vector e and angle ϕ in Eqs. (8.93)–(8.95) of Chapter 8)

π
5
 sgn[n ⋅ ( x − y)] 
    uK ( y) = −ciJLm γ Lm Re ∫0
*
∑g (P)
KJ (θ)  ∫
S s( P ) ⋅ ( x − y)
ni ( x )dS( x )dθ  (9.77)
P =1  

where ni(x) again is the outward normal on the boundary S of the QD. We remark
that similar expressions for the extended displacement Green’s functions in an
MEE full-space can also be found in Chapter 8 with the detailed definitions for the
involved quantities.
Taking the derivative of Eq. (9.77), we find the corresponding extended strains
as follows.
For the elastic strains with k,r = 1,2,3:
γ *Lm ciJLm
γ kr ( y) =
2
 π 5  sgn[n ⋅ ( x − y)]  
× Re  ∫ ∑ [ sr( P ) gkJ
(P)
(θ) + sk( P ) grJ( P ) (θ)]  ∫ ( P ) ni ( x )d S( x ) d θ
 S [ s ⋅ ( x − y)]
2
 P = 1  
0

 (9.78a)

For the electric and magnetic fields with k = 1,2,3

π 5  sgn[n ⋅ ( x − y)] 
Ek ( y) = − γ *Lm ciJLm Re ∫
0 ∑  sk(P ) g4(PJ ) (θ) ∫S [ s(P ) ⋅ ( x − y)]2 ni ( x)dS( x) dθ
P =1  
 (9.78b)

π 5  sgn[n ⋅ ( x − y)] 
Hk ( y) = − γ *Lm ciJLm Re ∫
0 ∑  sk(P ) g5(PJ ) (θ) ∫S [ s(P ) ⋅ ( x − y)]2 ni ( x)dS( x) dθ
P =1  
 (9.78c)

Thus the induced field can be expressed as a line integral outside and a surface
integral over the boundary of the QD inclusion. There are two types of surface
integrals, one related to the induced extended displacements and the other to the
extended strains, expressed in the following text as

sgn[n ⋅ ( x − y)]
I iK ( y, θ) = ∫ ni ( x )dS( x )  (9.79a)
S s( K ) ⋅ ( x − y )

sgn[ n ⋅ ( x − y)]
JiK ( y, θ) = ∫ ni ( x )dS( x )  (9.79b)
S [ s( K ) ⋅ ( x −y)]2
9.8 Technical Application: QDs in Piezoelectric Semiconductors 309

(a) (b) ξ2 x3
ξ2
3
ξ1
3
2
n
q2
h q1
x (3)
x (2) x0
1 2 ξ1
l1 x0 l2
x (1) x2
l
x1
1

Figure 9.3. Geometry of the flat triangle Δ in the local coordinate (ξ1,ξ2)-plane (with corners
1,2,3) in (a), and the transformation from the global (x1,x2,x3) to local (ξ1,ξ2,ξ3) coordinates
where ξ3 is along n, the outward normal to the flat triangle in (b).

Again, if we assume that the surface S of the QD inclusion is made of an N-face


polyhedron with each being quadrangle, it can be expressed by 2N flat triangles. We
now want to integrate exactly over any flat triangle of area Δ.
For a flat triangle given in Figure 9.3a, the unit outward normal is constant. Thus
we have

sgn[n ⋅ ( x − y)]
I iK ( y, θ) = ni ∫ dΔ( x ) (9.80a)
Δ s(K ) ⋅ ( x − y)

sgn[n ⋅ ( x − y)]
JiK ( y, θ) = ni ∫ dΔ( x ) (9.80b)
Δ [ s( K ) ⋅ ( x− y)]2

We further select the following unit orthogonal local base vectors as (refer to
Figure 9.3b)

( x (2) − x (1) ) × ( x (3) − x (1) )


n=

x (2) − x (1)
q1 = ≡ (q11 , q12 , q13 )t (9.81)
x (2) − x (1)
q2 = n × q1 ≡ (q21 , q22 , q23 )t

We remark that the components for q1 and q2 are in terms of the global (x1,x2,x3)
coordinate system and that (q1,q2,n) are the same as (m,n,e) in Eq. (8.96) of Chapter 8.
The transformation between the global (xi) and local (ξi) coordinates is given by

x = x0 + ξ1q1 + ξ 2 q2 + ξ 3 n (9.82)

where x0 (x01,x02,x03) can be found as

x0 = x (1) + [( x ( 3) − x (1) ) ⋅ q1 ]q1


(9.83)
= x ( 3) − [( x ( 3) − x (1) ) ⋅ q2 ]q2
310 Green’s Functions in an Anisotropic Magnetoelectroelastic Bimaterial Space

Other triangle parameters are (refer to Figure 9.3a)

h = ( x (3) − x (1) ) ⋅ q2
l = x (2) − x (1)
(9.84)
l1 = ( x (3) − x (1) ) ⋅ q1
l2 = l − l1

In terms of the local coordinates (q1,q2,n), the global vector y can be expressed by

y = x0 + η1q1 + η2 q2 + η3 n (9.85)

where

η1 = ( y − x0 ) ⋅ q1 , η2 = ( y − x0 ) ⋅ q2 , η3 = ( y − x0 ) ⋅ n (9.86)

Recalling the expression for s(K) as in Eq. (8.94), we then have (θ ≡ φ)

s(K ) ⋅ ( x − y) = (ξ1 − η1 )cos θ + (ξ 2 − η2 )sin θ + (ξ 3 − η3 ) pK (9.87)

Then the flat triangle integrals can be reduced to

h l2 − l2 ξ 2 /h dξ 1 dξ 2
I iK ( y, θ) = − ni sgn( η3 )∫ ∫−l + l ξ / h ξ1 cos θ + ξ2 sin θ − ς K (9.88a)
0 1 1 2

h l2 − l2 ξ 2 /h d ξ 1d ξ 2
JiK ( y, θ) = − ni sgn( η3 )∫ ∫−l + l ξ / h (ξ1 cos θ + ξ 2 sin θ − ς K )2 (9.88b)
0 1 1 2

where

ς K = η1 cosθ + η2 sin θ + η3 pK (9.89)

The integrals of these two equations can be carried out exactly, with the
results being

ni sgn( η3 )h  z1K log z1K − z2 K log z2 K z1K log z1K − z3K log z3K 
I iK ( y, θ) = −  − 
cos θ  z1K − z2 K z1K − z3K 
 (9.90a)

ni sgn( η3 )h  log z1K − log z2 K log z1K − log z3K 


JiK ( y, θ) = −  −  (9.90b)
cos θ  z1K − z2 K z1K − z3K 

where

z1K = h sin θ − ς K , z2 K = l2 cos θ − ς K , z3K = − l1 cos θ − ς K (9.91)

Therefore, the inner surface integral in Eqs. (9.77) and (9.78) can be expressed in an
exact-closed form if the integral area is a flat triangle. The only numerical part in the
9.8 Technical Application: QDs in Piezoelectric Semiconductors 311

QD inclusion problem is the simple line integral over (0,π), which can be easily done
by a numerical integral approach, like the Gaussian quadrature. As we stated earlier,
this is very efficient if the QD is in a shape of polyhedron of N faces. A very typical
and well-grown QD shape is the pyramid or truncated pyramid (e.g., Bimberg et al.
1998).

9.8.2 Analytical Integral over a Flat Triangle, the Anisotropic


MEE Bimaterial Space Case
We assume that there is a 3D QD in Material 1 (x3 > 0) of the anisotropic MEE
bimaterial where the two half-spaces are in perfect contact. Because the solutions
in the bimaterial space are the superposition of the full-space solution and the
complementary parts, we need only to deal with the complementary parts when
integrating over the surface of the QD because in Section 9.8.1, we have already dealt
with the full-space part. Furthermore, in the complementary part, due to the similarity
of the expressions (Eqs. (9.22) and (9.24) versus (9.26) and (9.27)), we only need to
present the results for the flat triangle integrals in one of the half-spaces only. We take
the extended complementary displacement Green’s functions (9.21) and (9.22) as
example. Assuming that the extended eigenstrains are uniform, the integral over the
flat triangle of the QD surface for the QD inclusion-induced extended displacements
can be expressed as (for the image or the complementary part only)

1  π
uK ( y) = ciJLm γ *Lm ni ∫∆ ∫0 [ A(1)Gu
(1)
( A(1) )t ]KJ dθ d∆( x ) (9.92)
2π 2 

with Gu(1) being given by Eq. (9.22).


Exchanging the order of integration between the triangle integral and line
integral, we have

1 π (1)
uK ( y) = ciJLm γ *Lm ni
2π 2 ∫0 [ A(1) ∫Δ Gu dΔ( x)( A(1) )t ]KJ dθ (9.93)

where the inner integral over the matrix Gu(1) can be expressed as

(1) dΔ ( x )
∫Δ [Gu ]IJ dΔ( x ) = (G1 )IJ ∫
Δ − pI(1) x3 + pJ(1) y3 − [( x1 − y1 )cos θ + ( x2 − y2 )sin θ]
(9.94)

We now introduce the transformation between the global x (x1, x2, x3) and local ξ
(ξ1,ξ2,ξ3) coordinate systems associated with the flat triangle (Figure 9.3b), as in
Eq. (9.82). Then, the inner integral over the flat triangle can be expressed as (with
ξ3 = 0)

dΔ( x)
FIJ ( y, θ) ≡ ∫
∆ − pI(1) x3 + p J(1) y3 − [( x1 − y1 )cos θ + ( x2 − y2 )sin θ]
(9.95)
h l2 − l2 ξ 2 / h dξ 1 d ξ 2
=∫ ∫− l + l ξ / h
0 1 1 2 f1 (θ)ξ1 + f2 (θ)ξ 2 + f3 ( yk , θ)

where
312 Green’s Functions in an Anisotropic Magnetoelectroelastic Bimaterial Space

f1 (θ) = −(q11 cos θ + q12 sin θ + q13 pI(1) ) (9.96a)

f2 (θ) = −(q21 cos θ + q22 sin θ + q23 pI(1) ) (9.96b)

f3 ( y, θ) = −( x01 cos θ + x02 sin θ + x03 pI )


+ ( y1 cos θ + y2 sin θ + y3 pJ(1) ).
(9.96c)

The integration in Eq. (9.95) can now be carried out, and the results are

1  f1l2 + f3  f h + f3  − f1l1 + f3  f h + f3  
FIJ ( yq , θ) =  ln  2 − ln  2  (9.97)
f1  f2 − f1l2 / h  f1l2 + f3  f2 + f1l1 / h  − f1l1 + f3  

Thus, defining F being the 5×5 matrix in the preceding text, then the final result of
the extended displacements due to a flat triangle surface of the inclusion can be
expressed as

1 π
uK ( y) = ciJLm γ *Lm ni
2π2 ∫0 [ A(1)G1 F ( A(1) )t ]KJ dθ (9.98)

In summary, the extended displacements due to the flat triangle surface of the QD
can be expressed as a simple line integral after carrying out the exact closed-form
integration over the flat triangle. To find the induced extended strains, we need only
to take the derivative of Eq. (9.98) with respect to the coordinate y of the point-force
Green’s functions, which only appears in Eq. (9.96c). Thus the corresponding strains
can be also easily found. By adding the contribution from the full-space part, the
total response due to the QD can then be found for the bimaterial case.
Remark 9.7: Similar expressions can be found for the quantities in the other half-
space. For the reduced half-space case, one needs only to replace the matrices A, B,
G, as well as the eigenvalues pJ by the corresponding ones in the half-space.

9.9 Numerical Examples

9.9.1 A Pyramidal QD in a Piezoelectric Full-Space


This example is from Wang et al. (2006) where a pyramidal QD inclusion with a
uniform eigenstrain distribution (Figure 9.4) is located in a piezoelectric full-space.
For the QD, the unit nondimensional base length a = 1 and height h = 1 are specified
and the unit nondimensional extended eigenstrain components given inside the
pyramidal QD are γ ij* = 1 and Ei* = 1. The material of the piezoelectric full-space
is Gallium Arsenide (001) (cubic) with the material constants given in Chapter 2.
These properties are normalized by c0 = 10 11 N/m2, e0 = 10 C/m2 and ε0 = 10−9 C/(mV)
so that the results are dimensionless.
Figure 9.5 shows the distribution of the strain component γ33 on the vertical
plane containing VPQ that is parallel to the (x,z)-plane (a) and on the horizontal
plane containing HG (the dashed square in Figure 9.4) (b) through the centroid
9.9 Numerical Examples 313

z
V

H G Q
y

a
P
x a
Figure 9.4. Pyramidal QD under extended uniform eigenstrain fields.

(a) (b)
z 2 y 2

1.5 1.5

1 1

0.5 0.5

0 0

–0.5 –0.5

0 1 2x 0 1 2 x
Figure 9.5. Pyramidal inclusion in Gallium Arsenide (001): distribution of γ33 (a) on the vertical
plane VPQ and (b) on the horizontal plane HG of Figure 9.4. Reproduced with permission
from Wang et al. (2006): International Journal of Solids and Structures 43: 7593–608. © 2006
Elsevier.

G of the pyramid, as shown in Figure 9.4. Figure 9.6 shows the distribution of the
strain component γ12 on the vertical plane VPQ (a) and on the horizontal plane HG
(b). Finally, Figure 9.7 shows the distribution of the electric field component E2 on
the vertical plane VPQ (a) and on the horizontal plane HG (b). All these figures
demonstrate that the magnitude of the induced extended strains is larger inside the
QD than that outside. Consequently, in the QD-induced field analysis, one should
pay more attention to the induced quantities inside or on the boundary of the QD.

9.9.2 QD Inclusion in a Piezoelectric Half-Space


As the first example in this section, we assume that the half-space is made of GaAs
(001) or GaAs (111) and its surface is extended traction-free, which is also called
traction-free insulating (i.e., the tractions and the normal component of the electric
displacement are zero on the surface). We want to calculate the elastic and piezoelectric
fields on the surface of the half-space due to a buried point QD. The QD is under a
hydrostatic misfit-strain γ Ij* = γ *δ ij with γ* = 0.07 and its volume is va = 4πa3/3 where
a = 3 nm, and is located at depth d = 10 nm below the surface. Because we assume that
this is a concentrated point QD, no integration is required. The material properties
314 Green’s Functions in an Anisotropic Magnetoelectroelastic Bimaterial Space

(a) (b)
z 2 y 2

1.5 1.5

1 1

0.5 0.5

0 0

–0.5 –0.5

0 1 2 x 0 1 2 x
Figure 9.6. Pyramidal inclusion in Gallium Arsenide (001): distribution of γ12 (a) on the vertical
plane VPQ and (b) on the horizontal plane HG of Figure 9.4. Reproduced with permission
from Wang et al. (2006): International Journal of Solids and Structures 43: 7593–608. © 2006
Elsevier.

(a) (b)
z 2 y 2

1.5 1.5

1 1

0.5 0.5

0 0

–0.5 –0.5

0 1 2 0 1 2
x x
Figure 9.7. Pyramidal inclusion in Gallium Arsenide (001): distribution of E2 (a) on the vertical
plane VPQ and (b) on the horizontal plane HG of Figure 9.4. Reproduced with permission
from Wang et al. (2006): International Journal of Solids and Structures 43: 7593–608. © 2006
Elsevier.

of GaAs (001) and GaAs (111) can be found in Pan (2002a, 2002b), which are also
listed in Chapter 2. It is noted that the material properties of GaAs (111) can be also
obtained by coordinate transformation (e.g., Pan 2002a, 2002b).
Figure 9.8 shows the contours of the hydrostatic strain γkk on the surface of
GaAs (001) (a) and GaAs (111) (b) due to a point QD of volume va applied at a
distance d = 10 nm below the surface. It is observed that while the contour pattern
on the surface of GaAs (001) is in a square shape, the one on the surface of GaAs
(111) is in a triangle shape. We further mention that the contour shape of the
hydrostatic strain on the surface of an isotropic crystal due to a point QD is exactly
circular (Pan 2002b).
9.9 Numerical Examples 315

(a) (b)
15 15

0.2
10 0.6
10
0.
2 1.0 2.0 7.0
1.4 6.5
1.8 6.0

0.6
5 1.8 1.6 5 1.0 5.5
1.0

0.2
2.5 5.0
1.4
1.4

1.8

1.0

x [11–2] nm
4.5
x [100] nm

1.2
0.6

4.0
1.0

0 1.0 0 3.5
0.8 3.0

5.5 .0
4
2.5
0.2

1.4
0.6 2.5 2.0
–5 1.8 0.4 –5 1.5

0.6
1.4 0.2 1.0 1.0
0.5

0.2
0.6 1.0 0.0 0.0
–10 0.2
–0.2 –10 –0.5

–15 –15
–15 –10 –5 0 5 10 15 –15 –10 –5 0 5 10 15
y [010] nm y [–110] nm
Figure 9.8. Contours of the hydrostatic strain γkk on the surface of GaAs (001) (a) and GaAs
(111) (b) due to a point QD of volume va applied at distance d = 10nm below the surface.
Reproduced with permission from Pan (2002b): Elastic and piezoelectric fields in substrates
GaAs (001) and GaAs (111) due to a buried quantum dot. Journal of Applied Physics 91:
6379–87. © 2002, AIP Publishing LLC.

(a) (b)
15 15

1.4 –1.4

0.0
1.0

.0
0.2
0.6

–0.6
–0.2

–1

1.0
10 10
2.0 1.0
1.8
1.6 0.5
1.4 –1.0
1.8 1.2 5 0.0
5 –1
.8

0.0
1.4 .4
1.0 1.0 –0.5
–1 0.8 .0
x [11–2] nm

0.6 –2
x [100] nm

1.0 –1.0 –1.0


0.6 –0.6 0.4
0.0

0.2 –0.2 0.2 –1.5


0 –0.2 –0.0 0 –2.0
–0.6 0.2 –0.2
–1.0 0.6 –0.4
1.0 –0.6 .0 –2.5
–1.4 1.4 –2
–0.8 .0 –3.0
–1.0 –1
–5 –1.2 –5 0.0
–3.5
–1.4
–1.6 0.0 –4.0
–1.8 0.0
–4.5
–2.0
0

.8 1.8
–10
0.

–1 –2.2 –10 –5.0


.4 1. 1.4
0.2
–0.2

–1
–0.6

0.6
.0

0
1.0
–1

–15 –15
–15 –10 –5 0 5 10 15 –15 –10 –5 0 5 10 15
y [010] nm y [–110] nm
Figure 9.9. Contours of the electric potential φ on the surface of GaAs (001) (a) and GaAs
(111) (b) due to a point QD of volume va applied at distance d = 10nm below the surface.
Reproduced with permission from Pan (2002b): Elastic and piezoelectric fields in substrates
GaAs (001) and GaAs (111) due to a buried quantum dot. Journal of Applied Physics 91:
6379–87. © 2002, AIP Publishing LLC.

Figure 9.9 shows the contours of the electric potential φ on the surface of GaAs
(001) (a) and GaAs (111) (b) due to a point QD of volume va applied at a distance
d = 10 nm below the surface. It is observed that on the surface of GaAs (001), the
maximum magnitude of the electric potential is reached at the four quarters of
the surface and located at the same location in each quarter with the same value.
However, on the surface of GaAs (111), the maximum is reached at the center,
directly above the point QD. Furthermore, the maximum magnitude in GaAs (111)
is more than twice of that in GaAs (001) (Pan 2002b).
Figure 9.10 shows the contours of the vertical electric field Ez on the surface of
GaAs (001) (a) and GaAs (111) (b) due to a point QD of volume va applied at a
316 Green’s Functions in an Anisotropic Magnetoelectroelastic Bimaterial Space

(a) (b)
15 15

–0.0
0.0
0.1

–0.1
10 10
0.25 –0.0

–0.0
–0.1
0.2 .2 0.20
–0
5 –0.2
5 0.15 –0.3
–0.3
0.10

x [11–2] nm
0.1 –0.1
–0.4

0 .0
x [100] nm

0.05 –0.5
0.0
0 0.00 0 –0.6
–0.05 .6 –0.7
–0
–0.1 0.1 –0.10 –0.8
0.0

.3
–5 –0.15 –5 –0.9

–0
0.0
–0.20 –1.0
–0.25 –1.1
–10 –0.30 –10 –1.2
.2 0.2
–0
0.0
–0.1

0.1

–15 –15
–15 –10 –5 0 5 10 15 –15 –10 –5 0 5 10 15
y [010] nm y [–110] nm
Figure 9.10. Contours of the vertical electric field Ez on the surface of GaAs (001) (a) and
GaAs (111) (b) due to a point QD of volume va applied at distance d = 10nm below the
surface. Reproduced with permission from Pan (2002b): Elastic and piezoelectric fields in
substrates GaAs (001) and GaAs (111) due to a buried quantum dot. Journal of Applied
Physics 91: 6379–87. © 2002, AIP Publishing LLC.

(a) (b)
15 15
0.1 –0.1 –0.1
–0.1
10 0.2 10
0.5 0.9
0.4
0.4 0.8
.3
–0

0.3 0.7
0.3
–0

5 5
–0.1

0.6
.3

0.3
0.3

0.1

0.2
.1

0.5
.3

x [11–2] nm

–0
x [100] nm

–0

0.2
0.2

–0.3
–0.1

0.1 0.4
0.1

0.1 –0.1 0.3


0 –0.0 0 –0.1
0.2
0.2 –0.1 –0.1
–0.1 0.1
0.1

0
–0.2 0.1 .3
0.1

–0.2 0.0
–0.3 –0.1 –0.1
–5 –0.3 –5
–0.2
0.3

–0.3
–0.4 –0.3
–0.4
.3

0.2 –0.4
–0

–0.5
–10 –0.6 –10 –0.5
–0.1
–0.1
–0.3

0.1

–15 –15
–15 –10 –5 0 5 10 15 15 –10 –5 0 5 10 15
y [010] nm y [–110] nm
Figure 9.11. Contours of the vertical electric field Ez on the surface of GaAs (001) (a) and
GaAs (111) (b) due to a point QD of volume va applied at distance d = 10nm below the surface
under the traction-free and conducting surface conditions. Reproduced with permission from
Pan (2002b): Elastic and piezoelectric fields in substrates GaAs (001) and GaAs (111) due to
a buried quantum dot. Journal of Applied Physics 91: 6379–87. © 2002, AIP Publishing LLC.

distance d = 10 nm below the surface. Comparison of Figure 9.10a to 9.10b clearly


indicates the dependence of the electric field on the crystal orientation.
We emphasize that, besides dependence on the crystal orientation, the surface
conditions of the half-space could also significantly affect the QD-induced field. For
example, Figure 9.11 shows the contours of the vertical electric field Ez on the surface
of GaAs (001) (a) and GaAs (111) (b) due to a point QD of volume va applied at a
distance d = 10 nm below the surface. The boundary condition on the surface is now
traction-free and conducting (i.e., the electric potential is zero on the surface). By
comparing Figure 9.11 to Figure 9.10, the effect of electric boundary conditions on
9.9 Numerical Examples 317

(a) (b)
15 15

10 10
10.0 10 10
1.0 9.5 10 25
9.0
8.5 20 20
8.0
5 7.5 5 15
2.5 4.0
2. 7.0
5.5 5
6.5 10

1.0
20
6.0
1.0

5
x (nm)

8.5

x (nm)
5.5 10
0 5.0 0 0 0 0 0
7. 4.5 –10
0 4.0 –5
3.5 –20 –1
4.0 5.5 0
3.0 –10
2.5 2.5
–5 2.0 –5 –15
1.5 –20 –20
1.0 1.0
0.5 –25
0.0
–10
–10 –0.5 –10 –10 –30

–15 –15
–15 –10 –5 0 5 10 15 –15 –10 –5 0 5 10 15
y (nm) y (nm)
Figure 9.12. Contours of the hydrostatic strain γkk (×10−3) (a) and the electric potential φ
(×10−2V) (b) on the surface of AlN (1000) due to a point AlN QD of volume va applied at
depth d = 10nm below the surface. Reproduced with permission from Pan and Yang (2003b):
Elastic and piezoelectric fields in a substrate AlN due to a buried quantum dot. Journal of
Applied Physics 93: 2435–9. © 2003, AIP Publishing LLC.

the induced electric field becomes obvious, not only for the pattern but also for the
magnitude.
In the second example, we assume the half-space is made of piezoelectric AlN,
with AlN (0001) denoting the case in which the global z-axis is along the material
(0001)-axis, and AlN (1000) the case in which the global x-axis is along the material
(0001)-axis. Again, we model the AlN QD as a point source, with a volume va =
4πa3/3 where a = 3 nm and it is located at a depth d = 10 nm below the surface. At the
surface z = 0, the traction-free insulating condition is assumed (extended traction-
free). The misfit-strains in the AlN QD are given by γ xx * = γ * = 0.1375 and γ * =
yy zz
0.1267 in substrate AlN (0001) and γ xx * = 0.1267, γ * = γ * = 0.1375 in AlN (1000)
yy zz
(Pan and Yang 2003b). The material properties for AlN can be found from Pan and
Yang (2003b), and are also listed in Chapter 2.
While Figure 9.12a shows the contours of the hydrostatic strain γkk (×10−3) on
the surface of AlN (1000) due to the point AlN QD, Figure 9.12b plots those for the
corresponding electric potential φ (×10−2V). These results are completely different
from those on the surface of AlN (0001) due to the point AlN QD where the contours
for both hydrostatic strain and electric potential are in perfect circular shapes with a
concentration at the center (Pan and Yang 2003b).
Figure 9.13 shows the contours of the horizontal electric field Ex (×107V/m) on
the surface of AlN (0001) due to the point AlN QD, which are antisymmetric with
respect to the x-axis and symmetric with respect to the y-axis. The corresponding
contours for Ey can be simply obtained by a rotation of 90 degrees in the horizontal
(x,y)-plane due to the symmetry of material properties and the problem geometry. For
the same reason, the contours for Ez will be in perfect circular shapes. On the surface
of the AlN (1000) half-space, however, the contour shapes of the horizontal electric
fields Ex and Ey, as shown in Figures 9.14a and 9.14b are totally different from each
other and also different from those on the surface of the AlN-(0001) half-space.
318 Green’s Functions in an Anisotropic Magnetoelectroelastic Bimaterial Space

15

0 0

10
4 14
12
10
5 12
8
6

0
0
4
8
4 2
x (nm)

0 0
–4
0 0
–2
–12
–4
0

–6
–5 –8

0
–8 –10
–12
–4
–14
–10 –16

–15
–15 –10 –5 0 5 10 15
y (nm)

Figure 9.13. Contours of the horizontal electric field Ex (×107V/m) on the surface of AlN
(0001) due to a point AlN QD of volume va applied at depth d = 10nm below the surface.
Reproduced with permission from Pan and Yang (2003b): Elastic and piezoelectric fields in
a substrate AlN due to a buried quantum dot. Journal of Applied Physics 93: 2435–9. © 2003,
AIP Publishing LLC.

(a) (b)
15 15
.2

–0.2
0.4 –0
3
10 10 –0.2
4 0.4 1.8
3 1.6
2 1.4
3 0 1.2
0 1
5 0
5 –1
.4
1.0
1.6
1.0

0 0.8
–0.8

–3 –3 –1 0.6
–6 1.0
0.

–0.8
0.4

0.4
–0.2

–2
4
x (nm)

x (nm)

–9 0.4 0.2
–3 –0.2
0 –4 0 0.4
–0.2 0.0
–3 –0.8
–0.2
–5
–0.2

–6 –0.4
0.4

–1
–3 –6 1.0 1.6 .4 –0.6
–0.8

0 –7 –0.8
–5 –5
1.0

0 –8 –1.0
0 3 –1.2
–9
–1.4
–10 –1.6
–11 0.4 –0.2 –1.8
3
–10 –12 –10 –2.0
–0.2

.2
0.4

–0

–15 –15
–15 –10 –5 0 5 10 15 –15 –10 –5 0 5 10 15
y (nm) y (nm)
Figure 9.14. Contours of the horizontal electric fields (in 107V/m) Ex (a) and Ey (b) on the
surface of AlN (1000) due to a point AlN QD of volume va applied at depth d = 10nm below
the surface. Reproduced with permission from Pan and Yang (2003b): Elastic and piezoelectric
fields in a substrate AlN due to a buried quantum dot. Journal of Applied Physics 93: 2435–9.
© 2003, AIP Publishing LLC.

In summary, for the problem of an AlN QD within an AlN half-space, the


following features have been observed: (1) On the surface of the substrate AlN
(0001), the hydrostatic strain, electric potential, and vertical and horizontal electric
fields are rotationally symmetric with respect to the z-axis. However, these quantities
are not rotationally symmetric on the surface of the substrate AlN (1000). (2) A
hydrostatic strain as large as 0.01 on the surface of AlN (1000) and as large as 0.008
on the surface of AlN (0001) can be reached, both larger than that on the surface of
the substrate GaAs due to a QD with the same volume at the same depth. (3) The
electric potential on the surface of the substrate AlN (0001) is much larger than that
9.9 Numerical Examples 319

(a) (b)
x B
Surface
z
d

h 2Rt 2Rb
γ*Ij

y
A
x
Figure 9.15. Schematic of a hexagonal truncated-pyramidal AlN QD. (a) 3D view of a QD
buried in a half-space substrate. Lines A and B are parallel to z-axis and x-axis, respectively.
(b) View of the QD structure in the (x,y)-plane.

(a) (b)
10

8 00
5

0.0002
0.00

13
0.0

–0.00
17
6 0.0026
0.0012
0.0
00

3 0.0005 0.002
4 .00
5

–0 –0.0002 0.0014
–0.

2 –0.0009
003

0.0008
–0.0016
y (nm)

0 –0.0023 0.0002

0.0017
–0.003 –0.0004

–0.0013
0.0002
–2 –0.0037 –0.001
–0.0044
–4 3
–0.0016
0.00 –0.0051
0.0


00

–0.0058 –0.0022
–6
5

–0.0065 –0.0028
5
00
–8 0.0

–10
–10 –8 –6 –4 –2 0 2 4 6 8 10 –10 –8 –6 –4 –2 0 2 4 6 8 10
x (nm) x (nm)
Figure 9.16. Contour of the electric potential φ (in V) on the surface of AlN (0001) (a) and
the surface of AlN (1000) (b) due to a hexagonal truncated-pyramidal QD (Zhang 2010).

on the surface of AlN (1000), 0.8V versus 0.3V. (4) Large horizontal and vertical
electric fields, of the order 108V/m, can be induced on the surface of AlN, about two
orders of magnitude larger than that on the surface of the substrate GaAs due to a
QD with the same volume at the same depth.
In the third example, we now apply the preceding analytical solutions to investigate
the elastic and electric fields induced by QDs buried within the AlN half-space
substrate (Zhang 2010). The model is similar to that in the second example, except
that now the AlN QD has a finite size. A schematic 3D view of an AlN QD is shown
in Figure 9.15a with its cross-section in (x,y)-plane shown in Figure 9.15b. The QD size,
taken from experimental values (Arley et al. 1999), is described by the base and top
diameters of the pyramid, Rb = 8.5 nm and Rt = 4 nm; h = 4.1 nm is the QD height and
d = 6.9 nm is the distance between the top of QD and the half-space surface.
Figures 9.16a and 9.16b show, respectively, the contours of the electric potential
on the half-space surface of substrates AlN (0001) and (1000) due to the same
hexagonal truncated-pyramidal QD. It is observed that the contours of the electric
potential on the surface of AlN (0001) are rotationally symmetric about the z-axis
and exhibit a large potential well at the center, which will contribute to the lateral
carrier confinement in the dot. By contrast, the electric potential on the surface of
320 Green’s Functions in an Anisotropic Magnetoelectroelastic Bimaterial Space

(a) (b)
16

14
0.006
0.0055
12 0.005
0.004
0.0035
–0.0 0.003
10 01
0.0015 0.002
z (nm)

8 0.001
0.0035 0.004
–0.0005 0
6 01 –0.001
–0.0 –0.002
–0.0025
–0 –0.003
4 .00
–0.0045 1 –0.004
2 –0.005
–0.0065 –0.006
0
–10 –8 –6 –4 –2 0 2 4 6 8 10 –10 –8 –6 –4 –2 0 2 4 6 8 10
x (nm) x (nm)
Figure 9.17. Contours of electric potential φ (in V) in the (x,z)-plane (y = 0) in (a) AlN (0001)
and (b) AlN (1000) half-space structures. The downward trapezoid shows the vertical profile
of hexagonal truncated-pyramidal QD, with the surface at z = 0 on the bottom for easy
presentation (Zhang 2010).

Surface x B

Dx = 25 nm

Figure 9.18. Schematic of a pair of hexagonal truncated-pyramidal QDs. Dx = 25nm is the


center distance between the two QDs. The distance between their top and the half-space
surface is d = 6.9nm and the height of each QDs is h = 4.1nm (Figure 9.15a).

AlN (1000) is symmetric about the x-axis but antisymmetric about the y-axis. It can
be also observed that, similar to the point AlN QD case, the magnitude of the electric
potential on the surface of AlN (0001) is much larger than that on the surface of AlN
(1000), 0.0065V versus 0.0026V. Comparing to the corresponding solutions for the
point QD shown in the second example, the finite-size effect can be clearly observed,
especially by looking at the results on the surface of the AlN (1000) half-space.
Contours of the electric potential in the (x,z)-plane (y = 0) due to the same
hexagonal truncated-pyramidal QD are visualized in Figure 9.17. The dark areas
show regions of high electric potential and the light ones are the regions of low
electric potential. It can be seen that there is a large potential well created by the
potential difference between the pyramid base and top in AlN (0001) while large
potential wells show along the left and right sides in AlN (1000). The vertical profile
of the potential in AlN (1000) is antisymmetric about the z-axis. We can also see that
the potentials in the two structures are of the same magnitude, unlike the potential
on the surface in Figure 9.16.
Our analytical solutions can also be applied to study the interaction among
multiple QDs. For example, Figure 9.18 shows a pair of hexagonal truncated
pyramidal QDs within the AlN half-space.
9.9 Numerical Examples 321

(a) (b)
20

0
15

0
0.003 0.004
10 0.0 0.0 0.002 0.0032
01 01
0.001 0.0024
5
0

0
0.0016
0
y (nm)

0.0008

0
0 –0.001

0.001
0

0
–0.002
–0.0008
–5 –0.003
0 0 1 0 –0.0016
0.0 –0.004
–10 –0.0024
–0.005 –0.0032
–15 –0.006 –0.004

0
0
–20

–20 –15 –10 –5 0 5 10 15 20 –20 –15 –10 –5 0 5 10 15 20


x (nm) x (nm)
Figure 9.19. Contour of the electric potential φ (in V) on the surface of AlN (0001) (a) and
on the surface of AlN (1000) (b) due to a pair of hexagonal truncated-pyramidal QDs (Zhang
2010).

Surface
1.79h
d

2.155h 2.5h 3.732h

(a) z (b) z (c) z (d) z


x x x x

Figure 9.20. Geometry for (a) a cubic QD, (b) a truncated pyramid QD, (c) a pyramid QD,
and (d) a point QD. Top row is the 3D view and bottom row is the vertical (x,z)-plane view. All
these QDs have the same volume. The distance between their top and the half-space surface
is d = 2nm and the height of each QDs is h = 4nm, except for the point QD.

Figures 9.19a and 9.19b show, respectively, the contours of the electric potential
on the surface of the substrates AlN (0001) and (1000) due to the pair of QDs. It
is observed that the contour of the piezoelectric potential on the surface of AlN
(0001) is no longer rotationally symmetric in the regions directly above the QDs. It
is further observed from Figure 9.19b that a large potential well is created between
the two QDs.
In the fourth example (Zhang 2010), we apply our analytical solutions to
calculate the strain energy induced by a buried QD within the GaAs half-space
substrate. The surface of the substrate is extended traction-free. The QD is located at
322 Green’s Functions in an Anisotropic Magnetoelectroelastic Bimaterial Space

Table 9.1. Maximum Strain Energy Emax on the Surface of the Substrate GaAs
for Different QD Shapes with Different Depths (unit of energy = 118.8×1015Nm)

Depth Cubic Truncated pyramid Pyramid

d = 1nm (001) Emax 2.54 2.93 2.71


(x,y) (±1,±1) (±1,±1) (0,0)
(111) Emax 2.55 2.75 2.05
(x,y) (1,0) (1,0) (0,0)
d = 2nm (001) Emax 1.77 1.99 0.96
(x,y) (0,0) (0,0) (0,0)
(111) Emax 1.31 1.34 0.75
(x,y) (0,0) (0,0) (0,0)
d = 3nm (001) Emax 1.03 1.09 0.41
(x,y) (0,0) (0,0) (0,0)
(111) Emax 0.70 0.70 0.36
(x,y) (0,0) (0,0) (0,0)

5
4
(a) (b) (c) (d)
3
2
1
y (nm)

0
–1
–2
–3
–4
–5
5
(a) (b) (c) (d)
4
3
2
1
y (nm)

0
–1
–2
–3
–4
–5
–5 –4 –3 –2 –1 0 1 2 3 4 5 –5 –4 –3 –2 –1 0 1 2 3 4 5 –5 –4 –3 –2 –1 0 1 2 3 4 5 –5 –4 –3 –2 –1 0 1 2 3 4 5
x (nm) x (nm) x (nm) x (nm)
Figure 9.21. Normalized strain energy on the surface of the half-space substrate of GaAs
(001) (top row) and GaAs (111) (bottom row) induced by (a) a cubic, (b) a truncated pyramid,
(c) a pyramid, and (d) a point QD, where the contour values of the normalized strain energy
increase toward the center. The QD is embedded within the substrate with its top side at a
depth d = 2nm from the surface (Zhang 2010).

a depth d (= 2 nm) below the surface (Figure 9.20) and the misfit strain is hydrostatic,
that is, γ xx
* = γ * = γ * = 0.07. As for the QD shape, we assume it to be either cubic,
yy zz
pyramidal, truncated pyramidal, or point type (Grundmann et al. 1995). The QDs
have the same height h (= 4 nm, except for the point QD). To make all the QDs
(including the point QD) have the same volume, we have the base length 2.155h for
the cubic QD, upper length 1.79h and lower length 2.5h for the truncated pyramid
QD, and base length 3.732h for the pyramid QD (Figure 9.20). For the point QD
case, it is located at the middle height of the cubic QD (i.e., its vertical distance to
the surface is d+h/2). We study the effect of the QD shape and depth on the strain
energy on the surface (Pan et al. 2008).
9.9 Numerical Examples 323

(a) z (b) z (c) z


x x x

5
(a)
4 (b) (c)
3
2
1
y (nm)

0
–1
–2
–3
–4
–5
5
4 (a) (b) (c)
3
2
1
y (nm)

0
–1
–2
–3
–4
–5
–5 –4 –3 –2 –1 0 1 2 3 4 5 –5 –4 –3 –2 –1 0 1 2 3 4 5 –5 –4 –3 –2 –1 0 1 2 3 4 5
x (nm) x (nm) x (nm)

Figure 9.22. Geometry of a pyramid QD within a half-space substrate with different depth
d: the vertical (x,z)-plane view (top row), and the strain energy induced by the pyramid QD
in substrate GaAs (001) (middle row) and GaAs (111) (bottom row): (a) d = 1nm, (b) d =
2nm, and (c) d = 3nm. The contour values of the normalized strain energy increase toward the
center (Zhang 2010).

Shown in Figures 9.21a–d are, respectively, contours of the normalized strain


energy on the surface of GaAs (001) (top row) and GaAs (111) (bottom row),
induced by a buried cubic, truncated pyramidal, pyramidal, or point type QD. In this
example, the depth d = 2 nm, height h = 4 nm, and the strain energy is normalized by
118.8×10 15Nm (this normalization factor is also used for the strain energy in Table 9.1).
It is clear that different QD shapes (including the point QD) induce different strain
energy distributions on the surface of the substrate. Besides the difference of the
contour shape, the strain energy values are also different. For example, the contours
with increasing values move toward the center when the QD shape changes from
left to right (i.e., cubic, truncated pyramid, pyramid, and point type). Table 9.1 lists
the maximum strain energy values corresponding to different QD shapes at different
depths within the GaAs substrate with both (001) and (111) orientations. It is clear,
from Table 9.1, that while the induced maximum strain energy value decreases with
the depth, its value on the substrate GaAs (001) is always larger than that of the
corresponding inclined substrate GaAs (111). Furthermore, the effect of the QD
shape (cubic, truncated pyramid, pyramid, or point QD) on the strain energy is
complicated, as can be seen from Figures 9.21a–d. Also from Figures 9.21a–d, it is
observed, by comparing the top row to the bottom row, that the strain energy contour
lines of equal value all move toward the center, and that the contour shapes of the
strain energy over GaAs (001) are sharply different from those over GaAs (111).
324 Green’s Functions in an Anisotropic Magnetoelectroelastic Bimaterial Space

We now study the effect of QD depth on the strain energy distribution on the
surface. The top row of Figure 9.22 shows the depths of the pyramid QD within
the substrate where d = 1 nm, 2 nm, and 3 nm, while the strain energy distributions
on the surface of GaAs (001) and (111) are shown, respectively, in the middle and
bottom rows. The contour values of the normalized strain energy increase toward the
center. It is apparent from Figure 9.22 that as the QD moves away from the surface,
the strain energy contours approach to those due to a point QD with equal volume.

9.9.3 Triangular and Hexagonal Dislocation Loops


in Elastic Bimaterial Space
The analytical solutions developed in this chapter can also be applied to dislocations
in anisotropic bimaterials. The expressions for the dislocation-induced fields are the
integral of the point-dislocation Green’s functions over the dislocation domain. For
a dislocation loop in a flat triangle shape, the area integral over the dislocation can
be carried out analytically, similar to the one given in this chapter for the 3D QD
case. The only thing one needs to pay attention to is to express the point-dislocation
Green’s function correctly from the available point-force Green’s function using
the Betti’s reciprocal theorem. The details can be found in Chu et al. (2012) for an
anisotropic elastic bimaterial space and in Han et al. (2013) for an anisotropic MEE
bimaterial space.
We assume that there is a regular triangular dislocation with side length a lying
in the (1 1 1) slip plane, which is located in the upper half-space made of Al (x3 > 0)
(Figure 9.23). The distance from the origin of the local dislocation coordinate
system to the Al-Cu interface (x1,x2)-plane is assumed to be 2a. It is noted that the
triangular dislocation is situated symmetrically about x2 = x1. Despite this, we find
that neither the individual displacement components u1 and u2 nor the distributions
of shear stresses σ13 and σ23 are symmetric about x2 = x1. However, the distribution of
(u1+u2)/2 within the (x1,x2)-interface plane is antisymmetric and the distribution
of (u1−u2)/2 symmetric, as shown in Figure 9.24a and Figure 9.24b, respectively.

ξ2: [112]
x3: [001]

Triangular ξ3: [111] x3 > 0: Al


dislocation loop a

ξ1: [110]

x0
Interface

o x2: [010]
x1: [100]

x3 < 0: Cu

Figure 9.23. Geometry of a regular triangular dislocation with side length a in a 3D aluminum-
copper bimaterial system, where x3 > 0 is Al and x3 < 0 is Cu. Reproduced with permission from
Chu et al. (2012): Journal of the Mechanics and Physics of Solids 60: 418–31. © 2012 Elsevier.
9.9 Numerical Examples 325

(a) (b)

(u1+u2)/2b (u1–u2)/2b
0.0

02

0
00

0
3 3

00
1.0E–3 0.00

0.
06
0.
01 1.4E–3
00
.00
02
–0 8.0E–4 0.0009
2 1.2E–3
6.0E–4
0.0005

0
0
1.0E–3
1 4.0E–4
0.0 8.0E–4
2.0E–4 00

0
0 3
x2 /a

9 0.000
2

0 0 6
–0. 0.0E+0 0.00 6.0E–4
00

8 000
00
0.0

0.0 05 1

0
0 –2.0E–4 4.0E–4
0 . 0
–1
.0 04

0.0
7

–4.0E–4
–0 .00
–0 01

00

00
2.0E–4
0
.00

9
–0.00

0.0009
–0

–6.0E–4

0.0006
–2 0.0E+0
04

–8.0E–4
–2.0E–4
–3 –1.0E–3
–3 –2 –1 0 1 2 3 4 –3 –2 –1 0 1 2 3 4
x1/a x1/a

Figure 9.24. Distribution of (u1 + u2)/2 (a) and (u1 − u2)/2 (b) normalized by the magnitude
of the Burgers vector value b on the interface (x3 = 0) in the global coordinates, where the
coordinates are normalized by the side length a of the triangular dislocation. A peak value
of 0.00099b is located at point (−1.81a,0.03a,0) in (a), and that of 0.0014b is located at point
(−1.52a,1.02a,0) in (b). Reproduced with permission from Chu et al. (2012): Journal of the
Mechanics and Physics of Solids 60: 418–31. © 2012 Elsevier.

(a) (b)
σ13 + σ23 σ13 – σ23
2ηµ 0 2ηµ
–0

–0
0

.00
.00

3 1
1

0.002 1.0E–2
1.0E–2 0
0.004
–0

1 8.0E–3
2 0.00 8.0E–3
.0


0

01

–0 0
0.002 .0
04 0 6.0E–3 6.0E–3
.00 0

–0.001
2

1 4.0E–3
02

1 4
0.0050.006 0
06 4 4.0E–3
.00
–0
0.0

0.0
0.00.00 001
0.0
0 – 02 –0
2.0E–3 .
x2 /a

–0 2 2.0E–3
0 .00 01
–0 –0.0
01
0 0.0E+0 0 0.0
02

.0 0 02 0.0E+0
0.0

–0 6 0.
05 –2.0E–3 .00 00
0.0
–1 –0 4
0

–2.0E–3
4

0 4
.00

0.0 –4.0E–3 0.0


0

02
–0


0.006

–4.0E–3
–2 –6.0E–3
0.004

2
–0.004

–0.0

.00
2
0 01
0.00

–0
.0

–6.0E–3
01

–8.0E–3
–0
0

–3 –1.0E–2 –8.0E–3
–3 –2 –1 0 1 2 3 4 –2 –1 0 1 2 3 4
x1/a x1/a

Figure 9.25. Distribution of (σ13+σ23)/2 (a) and (σ13−σ23)/2 (b) normalized by ημ on the
interface (x3=0) in the normalized global coordinates, where η=b/(3a) and μ is the nominal
shear modulus of Al (Chu et al. 2012). The normalized peak value is about 0.01 and the
maximum normalized shear stresses σ13 and σ23 are equal but reach at different points.
Reproduced with permission from Chu et al. (2012): Journal of the Mechanics and Physics of
Solids 60: 418–31. © 2012 Elsevier.

Likewise, (σ13+σ23)/2 (Figure 9.25a) is antisymmetric and (σ13−σ23)/2 (Figure 9.25b) is


symmetric about x2 = x1. Peak values of these shear displacements and stresses, which
are likely regions for nucleating dislocations or voids, are offset from the line of
intersection between the dislocation glide plane and the interface plane (see figure
captions for peak values).
Remark 9.8: Dislocation solutions for materials with transverse isotropy can
be obtained from those for general anisotropic materials directly by proper
326 Green’s Functions in an Anisotropic Magnetoelectroelastic Bimaterial Space

degeneration. Yuan et al. (2013a; 2013b) and Yuan et al. (2014), however, have
recently developed a very elegant method, by expressing the potential functions in
Chapter 6 in a different form, to deduce the dislocation solutions in terms of line
integrals either for purely elastic or MEE materials with transverse isotropy.

9.10 Summary and Mathematical Keys

9.10.1 Summary
We have presented the point-force Green’s functions in general anisotropic MEE
bimaterial systems. The solutions include various imperfect interface condition
cases. The corresponding half-space solutions can be reduced from these bimaterial
Green’s functions, and again include various surface boundary conditions. Because
the solutions are derived based on the double Fourier transforms, with special
arrangements of the associated eigenvalues and eigenvectors, care must be taken
when one goes from the upper half-space solution to the lower half-space solutions,
or for the bimaterial case, goes from the case in which the source in Material 1 (z > 0)
to the case in which the source in Material 2 (z < 0).
For either the half-space case or the bimaterial space case, we also need the
corresponding full-space solution. This solution is discussed in more detail in
Chapter 8. Because our solutions are in terms of indices ranging from 1 to 5, they
include many reduced coupled or uncoupled cases as special solutions (refer to
Chapter 2). For example, one can assume that the magnetic coupling coefficients
qijk are zero so that the problem will be reduced to the anisotropic piezoelectric
case as the magnetic field will be decoupled from the rest. While we reduce the
size from 5 to 4, and therefore, the size of all the involved matrices and vectors
can be reduced from 5 to 4 (the in-plane stress dimension is reduced from 7 to 5),
care must be taken in arranging the associated eigenvalues and eigenmatrices. In
other words, in designing the program, one needs to single out the eigenvalues and
eigenmatrices corresponding to the purely magnetic case so that we are sure that
these eigensystems are totally decoupled.
We have assumed that the source and field points are both not on the interface or
surface simultaneously (i.e., either z≠0 or d≠0). Under these conditions, the involved
finite line integrals (from 0 to π) are regular and can be easily evaluated by any
common numerical integral, such as the Gaussian quadrature. For these cases, the
singularity is only in the full-space Green’s functions. By contrast, if both the source
and field points are on the interface or surface, then the involved line integrals
become singular with different orders of singularity. Reductions and numerical
calculation of these singular integrals can be found in Pan and Yang (2003a).
The corresponding point-dislocation Green’s functions can be easily found
from Eq. (2.24) deduced from Betti’s reciprocal theorem, as discussed in detail in
Chapter 2.

9.10.2 Mathematical Keys


In order to solve the bimaterial problem, we express the bimaterial solution as a
sum of the full-space solution plus a complementary or image part. The solution
9.11 References 327

is designed in such a way so that the full-space part of the solution takes care of
the jump conditions at the source level, and the image parts in both half-spaces are
introduced to satisfy the interface conditions at z = 0 and also their respective regular
conditions as the field point approaches infinity.
In carrying out the inverse Fourier transforms, the polar coordinates are introduced
so that the integration along the radial direction can be carried out exactly.
The methodology is very flexible and the bimaterial solution includes many
different imperfect but homogeneous interface cases.
Similarly, the reduced half-space solution contains many different surface
condition cases.

9.11 References
Alshits, V. I., Barnett, D. M., Darinskii, A. N., and Lothe, J. 1994. On the existence problem
for localized acoustic waves on the interface between two piezocrystals, Wave Motion
20: 233–44.
Arley, M., Rouviere, J. L., Widmann, F., Daudin, B., Feuillet, G., and Mariette, H. 1999.
Quantitative characterization of GaN quantum-dot structures in AlN by high-resolution
transmission electron microscopy. Applied Physics Letters 74: 3287–9.
Bimberg, D., Grundmann, M., and Ledentsov, N. N. 1998. Quantum Dot Heterostructures.
New York: Wiley.
Chu, H. J., Pan, E., Han, X., Wang, J., and Beyerlein, I. J. 2012. Elastic fields of dislocation
loops in three-dimensional anisotropic bimaterials. Journal of the Mechanics and Physics
of Solids 60: 418–31.
Dundurs, J. and Hetenyi, M. 1965. Transmission of force between two semi-infinite solids.
Journal of Applied Mechanics 32: 671–4.
Grundmann, M., Stier, O., and Bimberg, D. 1995. InAs/GaAs pyramidal quantum dots:
Strain distribution, optical phonons, and electronic structure. Physical Review B 52:
11969–81.
Han, X., Pan, E., and Sangghaleh, A. 2013. Fields induced by three-dimensional dislocation
loops in anisotropic magneto-electro-elastic biomaterials. Philosophical Magazine 93:
3291–313.
Pan, E. 2002a. Elastic and piezoelectric fields around a quantum dot: Fully coupled or semi-
coupled model. Journal of Applied Physics 91: 3785–96.
2002b. Elastic and piezoelectric fields in substrates GaAs (001) and GaAs (111) due to a
buried quantum dot. Journal of Applied Physics 91: 6379–87.
Pan, E. and Tonon, F. 2000. Three-dimensional Green’s functions in anisotropic piezoelectric
solids. International Journal of Solids and Structures 37: 943–58.
Pan, E. and Yang, B. 2003a. Three-dimensional interfacial Green’s functions in anisotropic
bimaterials. Applied Mathematical Modelling 27: 307–26.
2003b. Elastic and piezoelectric fields in a substrate AlN due to a buried quantum dot.
Journal of Applied Physics 93: 2435–9.
Pan, E. and Yuan, F. G. 2000. Three-dimensional Green’s functions in anisotropic bimaterials.
International Journal of Solids and Structures 37: 5329–51.
Pan, E., Zhang, Y., Chung, P.W., and Denda, M. 2008. Strain energy on the surface of an
anisotropic half-space substrate: effect of quantum-dot shape and depth. CMES-
Computer Modeling in Engineering & Sciences 41: 1–11.
Papas, C. H. 1988. Theory of Electromagnetic Wave Propagation. New York: Dover.
Ting, T. C. T. 1996. Anisotropic Elasticity. Oxford: Oxford University Press.
Volakis, J. L., Chatterjee, A., and Kempel, L. C. 1998. Finite Element Method for
Electromagnetics. New York: IEEE Press.
Wang, C. Y., Denda, M., and Pan, E. 2006. Analysis of quantum-dot-induced strain and electric
fields in piezoelectric semiconductors of general anisotropy. International Journal of
Solids and Structures 43: 7593–608.
328 Green’s Functions in an Anisotropic Magnetoelectroelastic Bimaterial Space

Yuan, J. H., Chen,W. Q., and Pan, E. N. 2013a. Line-integral representations of the displacement
and stress fields due to an arbitrary Volterra dislocation loop in a transversely isotropic
elastic full space. International Journal of Solids and Structures 50 (1): 160–75.
2014. Line-integral representations for extended displacements, stresses and interaction
energy of arbitrary dislocation loops in transversely isotropic magneto-electro-elastic
bimaterials. Applied Mathematics and Mechanics (English Edition) 35 (8): 1005–28.
Yuan, J. H., Pan, E. N., and Chen, W. Q. 2013b. Line-integral representations for the
elastic displacements, stresses and interaction energy of arbitrary dislocation loops
in transversely isotropic bimaterials. International Journal of Solids and Structures 50
(20–1): 3472–89.
Zhang, Y. 2010. Analysis of Elastic and Electric Fields in Quantum Structures by Novel
Green’s Functions and Related Boundary Element Methods. PhD diss., University of
Akron, Ohio.
Index

2D Fourier transform method, 267–74 in terms of Stroh eigenvalues and


2D governing equations, 39–41 eigenvectors, 275–81
2D potential Green’s functions, 8–9 via 2D Fourier transform method, 267–74
moments, 281–2
anisotropic elastic solids, governing equations relations among dislocation, faulting and force
and, 33–5 moments, 282–5
anisotropic magnetoelectroelastic bimaterial anisotropic piezoelectric semiconductors,
spaces, 293, 312–27 quantum dots in, 307–12
with extended point forces in material 2 anti-plane deformation, elastic isotropic full
and, 304–6 plane, 57–8
lower half-space under general surface anti-plane solutions
conditions and, 306–7 elastic isotropic bimaterial planes
pyramidal QD in piezoelectric full space, of line dislocations, 67–70
312–13 line force inside or outside circular
QD inclusion in a piezoelectric half-space, inhomogeneity, 76–80
313–24 of line forces, 65–7
quantum dots in anisotropic piezoelectric screw dislocation inside or outside circular
semiconductors, 307–12 inhomogeneity, 80–3
solutions in Fourier domain for forces in of line forces and line dislocations, 58–60
material 1 and, 294–6
solutions in physical domain for forces in BEM (boundary element method), xv, 109
material 1 and, 296–9 Betti’s reciprocal theorem
solutions in physical domain for forces in applications of, 41–6
material 1 with imperfect interface extended, 41
conditions and, 299–301 BIE (boundary integral equation), 23–4, 109
triangular and hexagonal dislocation loops in bimaterial planes
elastic bimaterial space, 324–6 elastic isotropic, 109–10
upper half-space under general surface anti-plane solutions, 65–70, 76–83
conditions and, 302–3 image force of line dislocation, 101–7
anisotropic magnetoelectroelastic full spaces, image work of line forces, 107–8
260–1, 281–7, 290–1 plane-strain solutions, 70–6, 84–101
adjoint of a square matrix, 291–2 magnetoelectroelastic, 129, 137, 138
Cauchy’s residue theorem, 292 applications in semiconductor industry,
couple force, 281–2 129–135
dipoles, 281–2 Green’s functions of line forces and line
dislocations, numerical examples of, 287–8 dislocations in, 123–8
equivalent body forces of dislocations, half-spaces and, 221
285–7 potential Green’s functions in, 9–10
Green’s functions and bimaterial spaces
in terms of line integrals, 261–5 anisotropic MEE, 293, 312–27
in terms of Radon transform, 274–5 with extended point forces in material
in terms of Stroh eigenvalues, 265–7 2, 304–6

329
330 Index

bimaterial spaces (cont.) transversely isotropic MEE bimaterial spaces


lower half-space under general surface and, 221–5
conditions and, 306–7 vertical point forces and, 193–4
quantum dots in anisotropic piezoelectric magnetic
semiconductors, 307–12 extended Boussinesq solution for, 243–4
solutions in Fourier domain for forces in transversely isotropic MEE bimaterial spaces
material 1 and, 294–6 and, 221–5
solutions in physical domain for forces in transversely isotropic MEE full spaces
material 1 and, 296–9 and, 183–7
solutions in physical domain for forces circular inhomogeneity
in material 1 with imperfect interface edge dislocations inside, 93–101
conditions and, 299–301 edge dislocations outside, 84–93
upper half-space under general surface geometry of with source in matrix, 79
conditions and, 302–3 geometry of with source inside, 78
elastic isotropic line forces inside, 76–80, 93–101
derivatives of some common functions, 172–3 line forces outside, 76–80, 84–93
dislocation solution, 166 screw dislocation inside or outside, 80–3
displacements and stresses in half-space common functions, derivatives of, 172–3
induced by point force applied on surface complex function method, 60–1
with mixed boundary conditions, 174–5 constitutive relations, fully coupled MEE
displacements and stresses in traction-free materials, 30
half-space due to point force applied on couple force, 281–2
surface, 173–4 cubic elastic solids, 39
Green’s displacements and stresses in, 158–66
mathematic keys, 172 decoupled solutions
Papkovich functions, 145–58 of horizontal point force along x-axis,
uniform loading over circular area on 194–6, 197–8
half-space surface, 166–71 piezoelectric Green’s functions, 191–3
potential Green’s functions in, 11–12 of vertical point force, 196
transversely isotropic MEE, 220–1, 257 of vertical point force and negative electric
Green’s function solutions for MEE half- charge, 193–4
space, 241–6 derivatives
horizontal point force, solutions for, 225–30 of extended Green’s displacements, 211–13,
indentation over MEE half-space, 246–56 277–81
negative electric charge, solutions for, 221–5 due to point source in x-direction, 211–12
negative magnetic charge, solutions due to point source in y-direction, 212–13
for, 221–5 of scaled Green’s function, 213–15
for other interface conditions, 237–41 due to point source in K-direction, 214
reduced, 230–7 due to point source in x-direction, 214–15
vertical point force, solutions for, 221–5 due to point source in y-direction, 215
boundary element method (BEM), xv, 109 of some common functions, elastic isotropic
boundary integral equation (BIE), 23–4, 109 bimaterial spaces, 172–3
Boussinesq, J., 165 Ding, Haojiang, xvii
Boussinesq solution, 165, 172 dipoles
extended, for electric charge, 243–4 anisotropic MEE full space and, 281–2
extended, for magnetic charge, 243–4 in elasticity, 284
extended, for vertical point force, 243–4 with net moment and moment-free, 282
extended, indentation over MEE half disk-like inclusion case, 208
space, 250–1 dislocation. See also line dislocation
finding induced displacements and stresses, anti-symmetric and symmetric distributions
166–7 and, 325
dislocation solution, elastic isotropic full and
Cauchy’s residue theorem, 292 bimaterial spaces, 166
Cerruti, V., 165 dislocation source pairs in elasticity, 284
Cerruti solution, 165 distribution of shear stresses, 325
Cerruti solutions, extended, 244–6 edge
charge inside circular inhomogeneity, 93–101
electric outside circular inhomogeneity, 84–93
extended Boussinesq solution for, 243–4 Peach-Koehler force, 102–4, 105–7
piezoelectric bimaterial space and, 230–2 equivalence between elastic dislocation and
transversely isotropic magnetoelectroelastic single force couple, 287
full spaces, 183–7 equivalent body forces of, 285–7
Index 331

geometry of regular triangular dislocation, 324 anti-plane solutions of line forces and line
hexagonal dislocation loops in elastic bimaterial dislocations, 58–60
space, 324–6 plane displacements in terms of complex
point, 41–5, 136 functions, 60–1
relations among faulting and force moments plane-strain deformation, 57–8
and, 282–5 plane-strain solutions, 62–5
schematics of shear fault source, 285 full spaces, 140
screw, 101–2 dislocation solution, 166
triangular dislocation loops in elastic bimaterial displacements and stresses in half-space
space, 324–6 induced by point force applied on
uniform, 288 surface with mixed boundary
displacements and stresses conditions, 174–5
effective stress, 291 displacements and stresses in traction-free
elastic displacement, 289 half-space due to point force applied on
extended Green’s displacements, 209–11 surface, 173–4
derivatives of, 211–13, 277–81 Green’s functions of point forces
in half-space induced by point force applied in, 140–3
on surface with mixed boundary mathematic keys, 172
conditions, 174–5 Papkovich functions, 143–5
hydrostatic stress, 290 uniform loading over circular area on half-
normalized horizontal displacement, 169, 170 space surface, 166–71
normalized stress component, 168, 169, 171 elastic materials, material properties for, 40
normalized vertical displacement, 168, 170 electric charge
in traction-free half-space due to point force extended Boussinesq solution for, 243–4
applied on surface, 173–4 negative
divergence theorem, 6–8 piezoelectric bimaterial space and, 230–2
transversely isotropic magnetoelectroelastic
edge dislocation. See also line dislocation full spaces, 183–7
inside circular inhomogeneity, 93–101 transversely isotropic MEE bimaterial spaces
outside circular inhomogeneity, 84–93 and, 221–5
Peach-Koehler force energy point of view, governing equations
in elastic isotropic bimaterial planes, 102–4 from, 50–1
interacting with circular equilibrium equations
inhomogeneity, 105–7 anisotropic MEE solids and, 30
eigenstrain field concept, 48–9 elastic displacements and, 141
eigenvalues, Stroh, 265–7, 275–81 Eshelby, John D., 46, 101
eigenvectors, Stroh, 275–81 Eshelby inclusion problem, 46–8, 198–201,
elastic bimaterial space, triangular and hexagonal 216–17
dislocation loops in, 324–6 quantum wires and, 129–135
elastic Green’s functions, 195–6 spheroidal inclusion, 216–17
elastic isotropic Eshelby inhomogeneity problem, 48–9
bimaterial planes, 109–10 Eshelby tensor, extended, 201–8
anti-plane solutions, 65–70, 76–83 extended Betti’s reciprocal theorem, 41
image force of line dislocation, 101–7 extended Boussinesq solution
image work of line forces, 107–8 for electric charge, 243–4
plane-strain solutions, 70–6, 84–101 indentation over MEE half space, 250–1
bimaterial spaces for magnetic charge, 243–4
derivatives of some common functions, 172–3 for vertical point force, 243–4
dislocation solution, 166 extended Cerruti solutions, for horizontal point
displacements and stresses in half-space force, 244–6
induced by point force applied on surface extended Eshelby tensor, 201–8
with mixed boundary conditions, 174–5 extended Green’s displacements, 209–11
displacements and stresses in traction-free derivatives of, 211–13, 277–81
half-space due to point force applied on extended line dislocations, 45–6
surface, 173–4 extended line forces, 45–6
Green’s displacements and stresses in, 158–66 extended point dislocations, 41–5
mathematic keys, 172 extended point forces
Papkovich functions, 145–58 bimaterial space under, 294
uniform loading over circular area on half- in material 2 with anisotropic MEE bimaterial
space surface, 166–71 space, 304–6
full planes, 57 relation between extended point dislocations
anti-plane deformation, 57–8 and, 41–5
332 Index

faulting moments, anisotropic MEE full space anti-plane solutions of line forces and line
and, 282–5 dislocations, 58–60
flat triangles plane displacements in terms of complex
analytical integral over, 308–12 functions, 60–1
cubic and spherical inclusions, 307 plane-strain deformation, 57–8
geometry of, 309 plane-strain solutions, 62–5
force magnetoelectroelastic, 111, 132, 138
line, 115 applications in semiconductor industry,
under general interface conditions, 126–8 129–135
in half-plane under general boundary Green’s functions of line dislocations in half-
conditions, 119–23 plane, 118–19
inside circular inhomogeneity, 76–80, 93–101 Green’s functions of line dislocations in half-
outside circular inhomogeneity, 76–80, 84–93 plane under general boundary conditions,
line force in x-, y- directions, 123–6 119–23
anti-plane solutions of, 65–7 Green’s functions of line forces, 118
in half-plane, 118 line force and line dislocation solutions in 2D,
image work of line forces, 107–8 113–17
under perfect interface conditions, 124–6 plane-strain deformation, 111–13
solutions in 2D, magnetoelectroelastic planes, full spaces
113–17 anisotropic MEE, 260–1, 281–8, 290–1
line force in z-direction, 62–3, 70–6 adjoint of a square matrix, 291–2
point force, 140–3 Cauchy’s residue theorem, 292
point force, extended Green’s functions and, 261–81
bimaterial space under, 294 elastic isotropic, 140
in material 2 with anisotropic MEE dislocation solution, 166
bimaterial space, 304–6 displacements and stresses in half-space
relation between extended point dislocations induced by point force applied on surface
and, 41–5 with mixed boundary conditions, 174–5
point force in horizontal direction displacements and stresses in traction-free
decoupled solutions of, 194–6, 197–8 half-space due to point force applied on
extended Cerruti solutions for, 244–6 surface, 173–4
Green’s displacements and stresses by, 161–6 Green’s functions of point forces in, 140–3
Papkovich functions for, 157–8 mathematic keys, 172
piezoelectric bimaterial space and, Papkovich functions, 143–5
232–4, 235–7 uniform loading over circular area on half-
solutions of along x-axis, 187–91 space surface, 166–71
transversely isotropic MEE bimaterial spaces transversely isotropic MEE, 176, 198–209
and, 225–30 decoupled solutions, 191–6, 197–8
point force in vertical direction Eshelby inclusion problem, 216–17
decoupled solutions of, 196 extended Green’s displacements, 209–13
extended Boussinesq solutions for, 243–4 general solutions in terms of potential
Green’s displacements and stresses by, 158–60 functions, 176–83
negative electric charge and, 193–4 infinite MEE space subjected to concentrated
Papkovich functions for, 156–7 sources at the point, 177
piezoelectric bimaterial space and, negative electric charge and, 183–7
230–2, 234–5 negative magnetic charge and, 183–7
solutions of, 183–7 scaled Green’s function derivatives, 213–15
transversely isotropic MEE bimaterial spaces solutions of horizontal point force along
and, 221–5 x-axis, 187–91
point force in x-direction, 150–5, 156 solutions of vertical point force, 183–7
point force in z-direction, 146–50, 155 special material coupling cases, 209
force moments, anisotropic MEE full space unit sphere with spherical angles, 200
and, 282–5
Fourier domain, anisotropic MEE bimaterial Gallium Arsenide (GaAs)
space and, 294–6 contours of electric potential on surface
Fourier transform method (2D), 267–74 of, 315
free surface, Green’s function solutions for MEE contours of hydrostatic strain on surface of, 315
half-space with, 241–2 contours of vertical electric field on surface
full planes of, 316
elastic isotropic, 57 maximum strain energy on surface of substrate
anti-plane deformation, 57–8 GaAs for different QD shapes, 322
Index 333

normalized strain energy on surface of half- under general interface conditions, 126–8
space substrate of, 322 in half-plane, 118
pyramidal inclusion in, 313, 314 in half-plane under general boundary
Gauss’ theorem, 6–8 conditions, 119–23
general anisotropic magnetoelectroelastic under perfect interface conditions, 124–6
solids, 29–32 piezoelectric, 191–3
general interface conditions, Green’s functions of point forces, 140–3
of line dislocations and line forces potential functions, 8–22
under, 126–8 in anisotropic planes or spaces, 12–13
governing equations, 29, 50 in bimaterial planes, 9–10
anisotropic elastic solids, 33–5 in bimaterial spaces, 11–12
Betti’s reciprocal theorem inhomogeneous circle in full-plane, 13–17
applications of, 41–6 inhomogeneous sphere in full-space, 17–22
extended, 41 primary on 2D and 3D, 8–9
cubic elastic solids, 39 solutions for MEE half-space, 241–6
from energy point of view, 50–1 with free surface, 241–2
Eshelby inclusion problem, 46–8 with surface electrode, 242–3
Eshelby inhomogeneity problem, 48–9 surface functions, 243–6
general anisotropic magnetoelectroelastic unified-form, 197–8
solids, 29–32 Green’s representation, Papkovich functions
in linearly elastic isotropic solid, 140–3 and, 143–5
material properties Green’s second identity, 6–8
for elastic materials, 40
of MEE reported in recent literature, 36 half-plane
for piezoelectric materials, 34 Green’s functions of line dislocations in, 119–23
for TI piezomagnetic materials, 38 Green’s functions of line forces in, 118–19
piezoelectric materials, 32–3 in half-plane under general boundary
piezomagnetic materials, 33 conditions, 119–23
transformation of MEE material properties from half-space, MEE, 221
one coordinate system to the other, 51–4 Green’s function solutions for, 241–6
transversely isotropic MEE solids, 35–6 with free surface, 241–2
transversely isotropic or isotropic elastic with surface electrode, 242–3
solids, 37–8 surface functions, 243–6
transversely isotropic piezoelectric/ indentation over, 249–56
piezomagnetic solids, 37 axisymmetric rigid indenter and, 250
two-dimensional, 39–41 loaded and unloaded stages for conical
unit relations, 54 indenter, 247
gradient relations, anisotropic MEE solids lower, anisotropic MEE bimaterial space
and, 30–2 and, 306–7
Green, George, 6–8, 22–4 uniform loading over circular area on half-
Green’s displacements and stresses, 158–66. space surface, 166–71
See also displacements and stresses upper, anisotropic MEE bimaterial space
by horizontal point force, 161–6 and, 302–3
by vertical point force, 158–60 hat transformation, 84
Green’s functions, 25 hexagonal dislocation loops, in elastic bimaterial
anisotropic MEE full space and space, 324–6
in terms of line integrals, 261–5 horizontal point forces
in terms of Radon transform, 274–5 decoupled solutions of, 194–6, 197–8
in terms of Stroh eigenvalues, 265–7 extended Cerruti solutions for, 244–6
in terms of Stroh eigenvalues and Green’s displacements and stresses by, 161–6
eigenvectors, 275–81 piezoelectric bimaterial space and, 232–4,235–7
via 2D Fourier transform method, 267–74 solutions of along x-axis, 187–91
defined, 1–6 transversely isotropic MEE bimaterial spaces
elastic, 195–6 and, 225–30
of line dislocations, 123–6
under general interface conditions, 126–8 image force of line dislocation, elastic isotropic
in half-plane, 118–19 bimaterial planes, 101–7
in half-plane under general boundary image work of line forces, 107–8
conditions, 119–23 on anti-plane line force in bimaterial planes, 108
under perfect interface conditions, 124–6 on anti-plane line force interacting with circular
of line forces, 123–6 inhomogeneity, 108
334 Index

indentation over MEE half-space, 246–56 image work of line forces, 107–8
axisymmetric rigid indenter and, 250 on anti-plane line force in bimaterial
loaded and unloaded stages for conical planes, 108
indenter, 247 on anti-plane line force interacting with
infinite circular cylinder case, 208 circular inhomogeneity, 108
infinite series summation, integral over image line inside circular inhomogeneity, 76–80, 93–101
source and, 25–7 outside circular inhomogeneity, 76–80, 84–93
inhomogeneous circle in full-plane, potential plane-strain solutions of
Green’s functions and, 13–17 elastic isotropic bimaterial planes, 70–6
inhomogeneous sphere in full space, potential elastic isotropic full planes, 62–3
Green’s functions and, 17–22 solutions in 2D, magnetoelectroelastic planes,
integral equations for potential problems, 22–3 113–17
integral over image line source, infinite series line integrals, anisotropic MEE full space
summation and, 25–7 and, 261–5
interfaces Lorentz solution, 172, 241–2
Green’s functions of line dislocations and line Love, A. E. H., v
forces under, 126–8
perfect magnetic charge
Green’s functions of line dislocations and line extended Boussinesq solution for, 243–4
forces under, 124–6 negative
with Papkovich functions, 145–55 transversely isotropic MEE bimaterial spaces
perfect-bonded, for Papkovich functions, 156–8 and, 221–5
perfectly conducting, transversely isotropic transversely isotropic MEE full spaces
MEE bimaterial spaces and, 237–8 and, 183–7
rigidly contacting and electromagnetically magnetoelectric (ME) effect, 29
insulating interface, bimaterial Green’s magnetoelectroelastic. See MEE
functions for, 238–41 (magnetoelectroelastic)
smooth, Papkovich functions and, 155–8 material properties
smoothly contacting, transversely isotropic AlN, 34, 317–321
MEE bimaterial spaces and, 237–8 BaTiO3, 34, 36, 132, 176
transversely isotropic MEE bimaterial spaces CoFe2O4, 36, 38, 132, 176
and, 237–41 for elastic materials, 40
isotropic elastic solids, 37–8 GaAs, 34
isotropic piezoelectric space, reduced TI MEE contours of electric potential on surface
bimaterial spaces and, 230–7 of, 315
contours of hydrostatic strain on surface
Kelvin’s solution, 17, 141–3 of, 315
contours of vertical electric field on surface
line dislocation, 115 of, 316
anti-plane solutions of, 67–70 maximum strain energy on surface of
extended, 45–6 substrate GaAs for different QD
Green’s functions of, 123–6 shapes, 322
under general interface conditions, 126–8 normalized strain energy on surface of half-
in half-plane, 118–19 space substrate of, 322
in half-plane under general boundary pyramidal inclusion in, 313, 314
conditions, 119–23 GaN, 34
under perfect interface conditions, 124–6 InAs, 34
inside or outside circular inhomogeneity, 80–3 InN, 34
plane-strain solutions of of MEE reported in recent literature, 36
elastic isotropic bimaterial planes, 70–6 for piezoelectric materials, 34
elastic isotropic full planes, 63–5 PZT, 36, 135
solutions in 2D, magnetoelectroelastic planes, SiC, 34
113–17 Terfenol-D, 38
line forces, 115 for TI piezomagnetic materials, 38
anti-plane solutions of, 65–7 transformation of MEE material properties from
Green’s functions of, 123–6 one coordinate system to the other, 51–4
under general interface conditions, 126–8 ME (magnetoelectric) effect, 29
in half-plane, 118 MEE (magnetoelectroelastic).
in half-plane under general boundary See also transversely isotropic (TI) MEE
conditions, 119–23 bimaterial spaces; transversely isotropic
under perfect interface conditions, 124–6 (TI) MEE full spaces
Index 335

anisotropic magnetoelectroelastic bimaterial negative electric charge


spaces, 293, 312–27 piezoelectric bimaterial space and,
with extended point forces in material 2 230–2
and, 304–6 transversely isotropic magnetoelectroelastic full
lower half-space under general surface spaces, 183–7
conditions and, 306–7 transversely isotropic MEE bimaterial spaces
quantum dots in anisotropic piezoelectric and, 221–5
semiconductors, 307–12 vertical point forces and, 193–4
solutions in Fourier domain for forces in negative magnetic charge
material 1 and, 294–6 transversely isotropic magnetoelectroelastic full
solutions in physical domain for forces in spaces, 183–7
material 1 and, 296–9 transversely isotropic MEE bimaterial spaces
solutions in physical domain for forces and, 221–5
in material 1 with imperfect interface Neumann, C., 17
conditions and, 299–301 normalized horizontal displacement,
upper half-space under general surface 169, 170
conditions and, 302–3 normalized stress component, 168,
anisotropic magnetoelectroelastic full spaces, 169, 171
260–1, 281–8, 290–1 normalized vertical displacement, 168, 170
adjoint of square matrix, 291–2
Cauchy’s residue theorem, 292 orthogonal relations, anisotropic MEE full space
Green’s functions and, 261–81 and, 276
bimaterial planes, 129, 137, 138
applications in semiconductor industry, Papkovich functions
129–135 Green’s representation and, 143–5
Green’s functions of line forces and line with perfect interface, 145–55
dislocations in, 123–8 for perfect-bonded and smooth
half-spaces and, 221 interfaces, 156–8
composites, 29 with smooth interface, 155–6
general anisotropic magnetoelectroelastic Peach-Koehler (PK) force
solids, 29–32 on edge dislocation in elastic isotropic
material properties of, 36 bimaterial planes, 102–4
physical meanings of boundary on edge dislocation interacting with circular
conditions, 121 inhomogeneity, 105–7
full planes, 111, 132, 138 on screw dislocation in elastic isotropic
applications in semiconductor industry, bimaterial planes, 101–2
129–135 perfect interfaces
Green’s functions of line dislocations in half- Green’s functions of line dislocations and line
plane, 118–19 forces under, 124–6
Green’s functions of line dislocations in half- with Papkovich functions, 145–55
plane under general boundary conditions, perfect-bonded interfaces, for Papkovich
119–23 functions, 156–8
Green’s functions of line forces in half- perfectly conducting interface, transversely
plane, 118 isotropic MEE bimaterial spaces
Green’s functions of line forces in half-plane and, 237–8
under general boundary conditions, physical domain
119–23 for forces in material 1, anisotropic MEE
line force and line dislocation solutions in 2D, bimaterial space and, 296–9
113–17 for forces in material 1 with imperfect interface
plane-strain deformation, 111–13 conditions, anisotropic MEE bimaterial
half-spaces, 221 space and, 299–301
Green’s function solutions for, 241–6 piezoelectric full space, pyramidal quantum dots
indentation over, 249–56 in, 312–13
Mindlin, Raymond D., 165 piezoelectric Green’s functions, 191–3
Mindlin solution, 172, 241–2 piezoelectric half-space, quantum dots inclusion
moments in, 313–24
anisotropic MEE full space and, 281–2 piezoelectric materials
faulting, anisotropic MEE full space and, governing equations and, 32–3
282–5 material properties for, 34
force, anisotropic MEE full space and, 282–5 piezomagnetic materials, governing equations
Muskhelishvili, Nikolos I., 57, 60 and, 33
336 Index

PK (Peach-Koehler) force moments, 281–2


on edge dislocation in elastic isotropic relations among dislocation, faulting and force
bimaterial planes, 102–4 moments, 282–5
on edge dislocation interacting with circular Poisson’s integral formula
inhomogeneity, 105–7 in 2D and 3D domains, 2–6
on screw dislocation in elastic isotropic potential Green’s functions in anisotropic plane
bimaterial planes, 101–2 or space, 15–16
plane displacements in terms of complex potential Green’s functions in bimaterial
functions, 60–1 planes, 10
plane-strain deformation potential Green’s functions, 8–22
elastic isotropic full plane, 57–8 in anisotropic planes or spaces, 12–13
full planes, magnetoelectroelastic, 111–13 in bimaterial planes, 9–10
plane-strain solutions in bimaterial spaces, 11–12
elastic isotropic bimaterial planes, 84 general solutions in terms of, 176–83
of line dislocations, 70–6 inhomogeneous circle in full-plane, 13–17
of line forces, 70–6 inhomogeneous sphere in full space, 17–22
line forces or edge dislocations inside circular 3D, 8–9
inhomogeneity, 93–101 2D, 8–9
line forces or edge dislocations outside pyramidal quantum dots
circular inhomogeneity, 84–93 under extended uniform eigenstrain fields, 313
elastic isotropic full planes in piezoelectric full space, 312–13
of line dislocations, 63–5
of line forces, 62–3 quantum dots (QD)
point dislocation, 43, 136 in anisotropic piezoelectric semiconductors,
extended, 41–5 307–12
point forces analytical integral over flat triangle, 308–12
extended contour of electric potential on the surface of
bimaterial space under, 294 AlN, 321
in material 2 with anisotropic MEE contours of electric potential on surface of
bimaterial space, 304–6 AlN, 319
relation between extended point dislocations contours of horizontal electric field on surface
and, 41–5 of AlN, 318
Green’s functions of, 140–3 contours of hydrostatic strain and electric
horizontal potential on surface of AlN, 317
decoupled solutions of, 194–6, 197–8 geometry for, 321
extended Cerruti solutions for, 244–6 inclusion, in piezoelectric half-space, 313–24
Green’s displacements and stresses by, 161–6 pyramid, geometry of, 323
Papkovich functions for, 157–8 pyramidal
piezoelectric bimaterial space and, under extended uniform eigenstrain
232–4, 235–7 fields, 313
solutions of along x-axis, 187–91 in piezoelectric full space, 312–13
transversely isotropic MEE bimaterial spaces schematic of hexagonal truncated-pyramidal
and, 225–30 AlN QD, 319
normal to interface, 146–50, 155 schematic of pair of hexagonal truncated-
parallel to interface, 150–5, 156 pyramidal QD, 320
vertical vertical profile of hexagonal truncated-
decoupled solutions of, 196 pyramidal QD, 320
extended Boussinesq solutions for, 243–4 quantum wires (QWRs), 129–131
Green’s displacements and stresses by, 158–60 in MEE half-plane, 132–3
negative electric charge and, 193–4 in piezoelectric bimaterial plane, 135
Papkovich functions for, 156–7 in piezoelectric full plane, 131–2
piezoelectric bimaterial space and, square, 132
230–2, 234–5 T-shaped, 133, 134
solutions of, 183–7
transversely isotropic MEE bimaterial spaces Radon transform, anisotropic MEE full space
and, 221–5 and, 274–5
point-source solutions, anisotropic MEE full space reduced TI MEE bimaterial spaces, 230–7
and, 281–7 isotropic piezoelectric space, 230–7
couple force, 281–2
dipoles, 281–2 scaled Green’s function derivatives, 213–15
equivalent body forces of dislocations, 285–7 screw dislocation, 101–2. See also line dislocation
Index 337

semiconductor industry, 129–135 piezoelectric Green’s functions, 191–3


singular function, 1. See also Green’s functions of vertical point force, 196
smooth interfaces of vertical point force and negative electric
Papkovich functions and, 155–8 charge, 193–4
smoothly contacting interface, transversely Eshelby inclusion problem, 198–201, 216–17
isotropic MEE bimaterial spaces spheroidal inclusion, 216–17
and, 237–8 extended Eshelby tensor, 201–8
special material coupling cases, 209 extended Green’s displacements, 209–13
spherical case, 208 general solutions in terms of potential
spheroidal inclusion, Eshelby inclusion problem, functions, 176–83
216–17 infinite MEE space subjected to concentrated
Stokes’ theorem, 7–8 sources at the point, 177
stresses. See displacements and stresses negative electric charge, 183–7
Stroh formalism, anisotropic MEE full space and, negative magnetic charge, 183–7
265–7, 275–81 scaled Green’s function derivatives, 213–15
surface electrode, Green’s function solutions for solutions of horizontal point force along x-axis,
MEE half-space with, 242–3 187–91
surface functions, Green’s function solutions for solutions of vertical point force, 183–7
MEE half-space and, 243–6 special material coupling cases, 209
unit sphere with the spherical angles, 200
3D potential Green’s functions, 8–9 transversely isotropic (TI) piezomagnetic
TI (transversely isotropic) magnetostrictive (magnetostrictive) materials, 37
materials, 37, 38 transversely isotropic (TI) solids, 35–6, 37–8
TI (transversely isotropic) solids, 35–6, 37–8 triangles, flat
TI MEE bimaterial spaces. See transversely analytical integral over, 308–12
isotropic MEE bimaterial spaces cubic and spherical inclusions, 307
TI MEE full spaces. See transversely isotropic geometry of, 309
MEE full spaces triangular dislocation loops, in elastic bimaterial
Timoshenko, Stephen P., 1 space, 324–6
transversely isotropic (TI) MEE bimaterial 2D Fourier transform method, 267–74
spaces, 220, 257 2D potential Green’s functions, 8–9
Green’s function solutions for MEE half- two-dimensional governing equations, 39–41
space, 241–6 two-point function, 2. See also Green’s
with free surface, 241–2 functions
with surface electrode, 242–3
surface functions, 243–6 unified-form Green’s functions, 197–8
horizontal point force, solutions for, 225–30 uniform loading over circular area on half-space
indentation over, 246–56 surface, 166–71
negative electric charge, solutions for, 221–5 unit relations, governing equations and, 54
negative magnetic charge, solutions for, 221–5 unit sphere with spherical angles, 200
for other interface conditions, 237–41
problem description, 220–1 vertical point forces
reduced, 230–7 decoupled solutions of, 196
isotropic piezoelectric space, 230–7 extended Boussinesq solution for, 243–4
vertical point force, solutions for, 221–5 Green’s displacements and stresses by, 158–60
transversely isotropic (TI) MEE full spaces, 176, negative electric charge and, 193–4
198–209 piezoelectric bimaterial space and, 230–2, 234–5
decoupled solutions solutions of, 183–7
of horizontal point force along x-axis, transversely isotropic MEE bimaterial spaces
194–6, 197–8 and, 221–5

You might also like