You are on page 1of 250

Sanichiro 

Yoshida

Deformation
and Fracture
of Solid-State
Materials
Field Theoretical Approach and
Engineering Applications
Deformation and Fracture of Solid-State Materials
Sanichiro Yoshida

Deformation and Fracture


of Solid-State Materials
Field Theoretical Approach and Engineering
Applications

123
Sanichiro Yoshida
Department of Chemistry and Physics
Southeastern Louisiana University
Hammond, LA, USA

ISBN 978-1-4939-2097-6 ISBN 978-1-4939-2098-3 (eBook)


DOI 10.1007/978-1-4939-2098-3
Springer New York Heidelberg Dordrecht London
Library of Congress Control Number: 2014949877

© Springer Science+Business Media New York 2015


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of
the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation,
broadcasting, reproduction on microfilms or in any other physical way, and transmission or information
storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology
now known or hereafter developed. Exempted from this legal reservation are brief excerpts in connection
with reviews or scholarly analysis or material supplied specifically for the purpose of being entered
and executed on a computer system, for exclusive use by the purchaser of the work. Duplication of
this publication or parts thereof is permitted only under the provisions of the Copyright Law of the
Publisher’s location, in its current version, and permission for use must always be obtained from Springer.
Permissions for use may be obtained through RightsLink at the Copyright Clearance Center. Violations
are liable to prosecution under the respective Copyright Law.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
While the advice and information in this book are believed to be true and accurate at the date of
publication, neither the authors nor the editors nor the publisher can accept any legal responsibility for
any errors or omissions that may be made. The publisher makes no warranty, express or implied, with
respect to the material contained herein.

Printed on acid-free paper

Springer is part of Springer Science+Business Media (www.springer.com)


Preface

This book stems from my first acquaintance with Academician Victor E. Panin
of the Soviet Academy of Sciences,1 and my daily research log that I have been
keeping since then. At the end of 1990, I had the opportunity to listen to a
lecture given by Acad. Panin at a meeting held in Tokyo, Japan. In this meeting,
a number of delegates from the Soviet Academy of Sciences gave presentations to
Japanese business people. The Acad. Panin’s lecture was on a new theory of plastic
deformation that he called physical mesomechanics. Although I understood only
20 %, or probably less, of his lecture, I was greatly fascinated by his enthusiastic
presentation and by the physical-mesomechanical view of plastic deformation. (He
was among the two or three presenters who gave the talk in English without an
interpreter.) In particular, the Maxwell-type field equations that describe plastic
deformation dynamics interested me greatly. At that time, my field of research was
laser and spectroscopy, and I was using the Maxwell equations of electrodynamics
on a daily basis. Although I did not understand the Maxwell-type field equations
of physical mesomechanics in any depth necessary to comprehend the deformation
dynamics behind them, I was able to understand that the equations described the
translational and rotational interaction of the displacement field. With my limited
knowledge of continuum mechanics, I was able to sense that material rotation and
its interaction with translational displacement is important in the plastic regime, and
that the Maxwell-type field equations represent that effect.
During the coffee break, I came to Acad. Panin to introduce myself and ask
a number of questions about his presentation. He answered each of my questions
enthusiastically. Moreover, he kindly invited me to the post-meeting banquet to be
held at the USSR embassy later that day. Of course I accepted the invitation and
attended the banquet where I was able to discuss with Acad. Panin a wide range
of topics in strength physics and material sciences. He gave me a book written in
Russian as a gift, and invited me to an international conference being held in the

1
Presently the Russian Academy of Sciences.
v
vi Preface

following summer in Tomsk, Siberia. I did not know the language at that time. I was
so interested in the book that I took Russian language courses for 2 years.
In the summer of 1991, I attended the conference in Tomsk and met a group of
scientists working in Acad. Panin’s group. The discussions I had with them were
revolutionary to me. They explained the Maxwell-type field equations, the interac-
tion between the translational and rotational displacement in plastic deformation,
and other gauge theoretical concepts in detail. To be honest, my knowledge about
gauge theories at that time was almost none. After returning home, I read a handful
of books on gauge theories and got more confused. I kept reading and learned that
the electromagnetic field is the gauge field that makes quantum mechanics locally
symmetric. This brought me to the turning point. I started to understand the concept
of gauge transformation and local symmetry. I analyzed various gauge theoretical
concepts in deformation dynamics via analogy to electrodynamics. Interestingly,
this exercise deepened my understanding on electrodynamics. I noticed a number of
different views on Faraday’s law and Ampere’s law as the interaction between the
electric and magnetic field that nature uses as a mechanism to stabilize events, e.g.,
prevent runaway increase of current. This, in turn, helped me consolidate the basic
understanding on the gauge-field nature of plastic deformation dynamics.
As I kept deepening my understanding on the physical foundation of physical
mesomechanics, I realized that the theory was much more profound than I initially
thought. It was an elegant theory capable of describing plastic deformation based
on pure physics, unlike most theories of plastic deformation that relies on phe-
nomenology or mathematical models. It indicated a number of potential engineering
applications as well. However, the work at that time was somewhat inclined toward
the mathematical aspect of the theoretical foundation with little experimental proofs.
I started to conduct experiments trying to prove various elements of the theory, such
as transverse wave characteristics of displacement field in the plastic regime. To
measure displacement field, I used an optical interferometric technique known as
the ESPI (Electronic-Speckle Pattern Interferometry). I found several interesting
phenomena that could be explained by the same physical foundation as physical
mesomechanics. Through analysis of these experimental observations, especially
with the help of analogy with electrodynamics, I conceived new ideas in the
description of deformation dynamics such as the concept of deformation charge
and its role of energy dissipation. This helped me advance the theory from the field-
theoretical description of plastic deformation dynamics to a comprehensive theory
of deformation and fracture based on the same theoretical foundation. To date, I
have continued investigating the field theoretical dynamics of deformation using the
ESPI.
As will be discussed in the following chapters, development of this theory has
not been completed. I decided to put together the knowledge and information I
gained so far as a book at this point for several reasons. First, recent experimental
observations have convinced me of the validity of the theory. Essentially, the gauge
field in deformation dynamics makes the law of linear elasticity locally symmetric.
The nonlinear dynamics in the plastic regime is formulated through the potential
associated with the gauge field. Second, these experimental observations and their
Preface vii

field theoretical interpretations demonstrate potential of engineering applications.


In particular, the use of ESPI techniques allows us to visualize the deformation field
as a full-field image, and along with field theoretical interpretations, it provides
us with various information. For example, the use of the field theoretical criteria
of plastic deformation and fracture allows us to make diagnosis of the deformation
state of a given object. Third, I would like to invite specialists of different disciplines
to this research for further development of the theory and applications. On the
theoretical side, connections with microscopic theories are very important. At this
point the theory incorporates the effect of microscopic defects that causes plastic
deformation generally into the field equations via the source terms. If a specific
form of the source term is provided by a microscopic theory, it is possible to
describe how the microscopic defect can evolve to the final fracture under a given
condition. Also, more thermodynamic argument will allow us to discuss the energy
dissipation process resulting from irreversible plastic deformation more specifically.
For applications, software development for visualization of displacement field in
objects under deformation, especially during the transitional stage from one regime
to another, e.g., from the elastic to plastic regime, will be not only an interesting
application but also helps further advancement of the theory. Numerical simulations
are also important for further tests of the theory and explore for new applications.
Lastly, I would like to share my experience of learning the Maxwell’s formalism
and the gauge theories with students. A number of electrodynamic concepts that
were unclear to me when I was in the graduate school became crystal clear through
this project. I would like to invite students and have them feel the beauty of field
theories. In my opinion, this subject is ideal to visualize the concept of local
symmetry associated with a gauge field, which is otherwise abstract and difficult
to comprehend. I tried my best to portray the complicated concept with plain terms
and analogies without going into mathematical details. It is also a unique case in
which these concepts, which are usually discussed by scientists specialized in basic
physics such as high energy or particle physics, are discussed in connection with
real world applications such as nondestructive testing of metal objects. I hope that
this book is helpful to people in any of these and related disciplines.
I tried my best to cite literature appropriately. If some papers or books do not
receive fair credit or are not cited, I apologize.
Finally, I would like to express my sincere gratitude to a number of people. First
of all, I would like to thank Acad. Victor Panin for introducing to me his beautiful
paradigm of deformation dynamics and his friendship ever since. I am grateful
to countless colleagues who always supported me during the development of this
theory, especially Professor Cesar Sciammarella for his continuous encouragement
and precious discussions. Through my learning processes of gauge theories and
continuum mechanics, I realized that I owe greatly to all the professors and teachers
from whom I received my graduate and undergraduate trainings. Without their
excellent instruction, I would have never reached the present level of understanding
on the subjects. I also thank all of my friends and students who helped me with the
experiments and computations that provided a number of supporting data.

Hammond, LA, USA Sanichiro Yoshida


Contents

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Background. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Aim, Scope, and Organization of This Book . . . . . . . . . . . . . . . . . . . . . . . . . . 5
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2 Quick Review of Theories of Elastic Deformation . . . . . . . . . . . . . . . . . . . . . . . 9
2.1 Displacement and Deformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Hooke’s Law and Poisson’s Ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.3 Principal Axis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.4 Equation of Motion and Elastic Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.4.1 One-Dimensional Longitudinal Elastic Waves . . . . . . . . . . . . . . . . 32
2.4.2 Three-Dimensional Compression Waves . . . . . . . . . . . . . . . . . . . . . . 33
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3 Quick Review of Field Theories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.1 The Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.2 Symmetry in Physics. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.3 Global and Local Transformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.4 Gauge Field. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.5 Lagrangian Formalism and Field Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.6 Electrodynamics as a Gauge Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4 Field Theory of Deformation and Fracture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.1 Gauge Theories of Deformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.2 The Big Picture of the Present Field Theory. . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.3 Field Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
4.4 Wave Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

ix
x Contents

5 Interpretations of Deformation and Fracture Phenomena


from Field Theoretical Viewpoint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
5.1 Field Equation as an Equation of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
5.2 Comprehensive Description of Deformation . . . . . . . . . . . . . . . . . . . . . . . . . . 93
5.2.1 Elastic Deformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
5.2.2 Plastic Deformation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
5.2.3 Fracture. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
5.3 Physical Meaning of Potentials. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
5.3.1 Vector Potential and Scalar Potential from Gauge . . . . . . . . . . . . 114
5.3.2 Scalar and Vector Potential from Viewpoint
of Wave Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
5.3.3 Field Forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
5.4 Physical Meaning of Charges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
5.5 Thermodynamic Consideration. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
6 Optical Interferometry and Application to Material
Characterization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
6.1 Basics of Light and Optics. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
6.1.1 Light as an Electromagnetic Wave . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
6.1.2 Interaction with Media and Photon . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
6.1.3 Laser and Coherence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
6.1.4 Imaging. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
6.2 Interference and Interferometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
6.2.1 Mathematics of Waves. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
6.2.2 Interference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
6.2.3 Michelson Interferometer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
6.2.4 Mach–Zehnder Interferometer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
6.3 Electronic Speckle-Pattern Interferometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
6.3.1 Speckles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
6.3.2 In-Plane Displacement. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
6.3.3 Out-of-Plane Displacement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
7 Experimental Observations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
7.1 Plastic Deformation Wave . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
7.1.1 Decay Characteristics. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
7.2 Vortex-Like Displacement Field. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
7.3 Observation of Charge-Like Phenomena . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
7.3.1 Charge-Like Pattern and PLC Band . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
7.3.2 Cyclic Loading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
7.3.3 Experiment with Notched Specimen. . . . . . . . . . . . . . . . . . . . . . . . . . . 197
7.3.4 Acoustic Emission . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
7.3.5 Temperature Rise Due to Plastic Deformation . . . . . . . . . . . . . . . . 201
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
Contents xi

8 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
8.1 Evaluation of Stress Concentration with Charge-Like Patterns . . . . . . . 210
8.1.1 Stabilized and Unstabilized Non-welded
Specimens, A5052-S and A5052-N. . . . . . . . . . . . . . . . . . . . . . . . . . . . 210
8.1.2 AA6063 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
8.1.3 Welded Specimens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
8.2 Plastic Deformation and Fracture Criteria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
8.2.1 Plastic Deformation Criterion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
8.2.2 Fracture Criterion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
8.2.3 ESPI Experiment on Plastic Deformation
and Fracture Criteria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
8.2.4 Interpretation of Fringe Patterns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
8.3 Evaluation of Load Hysteresis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
Chapter 1
Introduction

1.1 Background

Deformation and fracture of solid-state materials have been studied for centuries.
The subject is not only of scientific interest but also of extreme practical importance.
A great number of scholars in a wide range of disciplines developed various theories.
The history of continuum mechanics can be traced back to the Hellenic period [1].
In 1660, Robert Hooke discovered the law of elasticity, the linear relation between
tension and extension in an elastic spring. The scientists in the seventeenth and
eighteenth centuries introduced the basic concepts of strain by extending Newton’s
law of motion for a point mass to a motion law for a deformable body with a finite
volume. By the middle of the nineteenth century, Augustin-Louis Cauchy, a French
mathematician, compiled these achievements into the basic framework of three-
dimensional continuum mechanics. Cauchy’s contribution is of especial significance
because of the mathematical rigorousness of his formulation, which is contrastive to
the heuristic approach exploited by earlier scientists. This is evidenced by the fact
that his formulations are still used by engineers of the twenty-first century, including
numerous software packages for numerical simulations of solid mechanics. In the
twentieth century, continuum mechanics developed to a unifying theory combined
with the advancement in thermodynamics and rheology. Clifford Truesdell was the
major force in this development. A number of textbooks such as the one written
by Landau and Lifshitz as a volume of their theoretical physics courses [2] are
available.
Elasticity as a nonlinear problems is also studied. It was initiated by Poincare and
Lyapunov in their study of ordinary differential equations of discrete mechanics at
the end of the nineteenth century. Methods for handling nonlinear boundary-value
problems were slowly developed by a handful of mathematicians on the first half of
the twentieth century. Further information about nonlinear elasticity can be found
in [3].

© Springer Science+Business Media New York 2015 1


S. Yoshida, Deformation and Fracture of Solid-State Materials,
DOI 10.1007/978-1-4939-2098-3__1
2 1 Introduction

In the area of plastic deformation, there are several mathematical descriptions


[4]. One is nonlinear deformation theory. In this theory, the stress is expressed as
a nonlinear function of the strain, as opposed to a linear function of the strain as
in the case of the linear elastic theory. This nonlinearity can also be viewed as
that the stress–strain relation is locally linear (infinitesimally linear at each value of
strain) on the stress–strain curve but the stiffness (the ratio of the stress to the strain)
reduces as the strain increases beyond the linear limit. Although this approach is
accurate as long as the stiffness is known as a function of strain, it cannot account
for irreversibility of the deformation. Flow plasticity theory [5] assumes that the
total strain can be decomposed into elastic and plastic parts. The plastic strain is
determined from the linear elastic relation to the stress of the material. The plastic
part is determined from a flow rule and a hardening model.
Another important theory of plasticity is dislocation theory. In 1934, Egon
Orowan [6], Michael Polanyi [7], and Geoffrey Ingram Taylor [8], approximately
at the same time, published papers to explain plastic deformation in terms of
dislocations. Experiments show that plastic deformation results from slip on specific
crystallographic planes in response to shear stress along the planes. However,
observed shear strength is orders of magnitude lower than theory. This discrepancy
indicates that slip is caused by a mechanism where lattice defects referred to as
dislocations move along the plane. When the dislocations complete the movement
at the end of the plane, slip occurs on the entire slip plane.
Fracture mechanics was developed during World War I by an aeronautical
engineer, Alan Arnold Griffith [9], to explain the fact that fracture of brittle materials
occur under two orders of magnitude lower stress than the theoretically predicted
value. He developed a thermodynamic theory that states that the growth of a crack
requires the creation of two new surfaces and hence an increase in the surface
energy. At the same time, it reduces the elastic energy stored in the material.
The difference of them (surface energy–elastic energy) can be interpreted as the
free energy. As the crack area increases, the surface energy increases linearly and
the elastic energy decreases quadratically, hence the free energy has a maximum.
Griffith postulated that the crack would increase spontaneously leading to fracture
beyond this maximum point because the free energy would decrease monotonically.
Accordingly, he defined the corresponding crack length as the critical crack length.
Griffith’s theory explains the fracture behavior of brittle materials well. However,
the actual energy necessary to fracture a ductile material is orders of magnitude
higher than the corresponding surface energy. During World War II, a group under
George Rankine Irwin [10] realized that in ductile materials the plastic zone
developing at the crack tip increases in size with the applied load, dissipating
the energy as heat. Hence, an energy dissipative term has to be added in the
energy balance. With this modification, the theory explains experiment for ductile
materials.
A number of researchers apply above-mentioned theories to engineering prob-
lems as well. Timoshenko [11], for example, describes the use of elasticity for a
number of practical applications.
1.1 Background 3

Sensor technology for defect detections has been advanced tremendously. A


number of techniques based on various principles such as ultrasonic/optical imag-
ing, eddy current, and other nondestructive technology have been developed and
used in the fields. Recent advancement allows us to detect micro-cracks in a very
early stage. On the material development side, a number of new materials with
additional designed properties such as reinforcement and anti-corrosion have been
developed.
In spite of all these developments, catastrophic accidents still occur. Aircraft parts
fail after passing the pre-flight inspection, structures such as bridges and stages
collapse totally unexpectedly. In many cases, the cause is unknown. Apparently,
the problems at least partly come from the limitation of the theory that the
inspection procedure is based on. The problem is not necessarily in incompleteness
of individual theories relevant to a given problem. Rather, it is the lack of connection
between the regimes of deformation. As discussed above, most of the currently
available theories are applicable selectively to a certain regime of deformation. They
are accurate in describing the dynamics in the corresponding regime. However,
in reality, the mechanical state cannot be characterized by a single regime. Even
in a specimen about to fracture, deformation in some parts are still in the elastic
regime. Flow plasticity theory may appear to be able to handle elastic and plastic
deformation simultaneously via the elastic and plastic parts of the strain. However,
this is a parameterized model, and the elasticity and plasticity are differentiated with
parameters, not physical laws. The use of these theories does not allow us to model
the transition from one regime to another. In the real world, on the other hand,
accurate description of the transition is extremely important.
The scale level is another issue. Micro and nano-technology is an emergent
field in various engineering disciplines. Most theories of solid mechanics were
developed for macroscopic objects. Mechanical properties of a material at the micro
and nanoscopic levels can be substantially different from those at the macroscopic
level. This imposes limitations in the applicability of existing theories to micro and
nanoscopic objects. In particular, those theories based on experimentally evaluated
parameters may need substantial modifications for applications to micro and nano-
scale systems. The issue is not necessarily limited to micro and nano-technologies.
Macroscopic fracture begins at the atomic level. The process starts with an atomistic
defect, grows to a macroscopic crack, and eventually the fracture of the entire object.
It is important that the theory can describe the transitions from one scale level to the
next on the same physical basis. If the crack generation can be predicted in an earlier
stage, the inspection technology will drastically advance.
The above-addressed issue clearly indicates the necessity of a theory capable of
describing all stages of deformation on the same physical basis, independent of the
scale level. It is obvious that such a theory must be based on a fundamental level of
physics. In this regard, the gauge theoretical approach that Panin et al. employed to
formulate dynamics of plastic deformation is promising. Details of their approach
can be found elsewhere [12–15]. In short, their approach is as follows: they
describe deformation with a transformation of GL.3; R/ group (three-dimensional
general linear group over real numbers [16]), and request local symmetry [17] in
4 1 Introduction

the transformation. In other words, they allow that the transformation matrix is
coordinate dependent and request that the dynamics be expressed in the same form
before and after the transformation. This requires replacement of usual derivatives
with covariant derivatives, or equivalently, introduction of a gauge. They find an
appropriate gauge and Lagrangian associated with it. Based on the principle of least
action, they derive field equations for each group element. After summation over
the group index, the field equations take a form analogous to Maxwell equations
of electrodynamics. The solution to the field equations represents transverse wave
characteristics of the displacement field in the plastic regime.
The present theory is based on the Panin’s formalism. When summed over the
group index (after contraction over the index representing the group elements) the
GL.3; R/ transformation matrix becomes the deformation tensor widely used in
linear elastic theory. Thus, the Panin’s approach can be interpreted as requesting
local symmetry in linear elasticity, which indicates that this formulation should
reduce to the conventional continuum mechanics. Subsequent analyses [18–20]
indicate that the transverse wave characteristics in the displacement field in the
plastic regime are driven by the shear restoring force of the material represented
by the shear modulus and that the longitudinal effect in the plastic regime is not
elastic force proportional to displacement but rather an energy dissipative effect.
Further, it has been found that fracture occurs when the material loses both its shear
resiting force mechanism and the longitudinal energy dissipating mechanism, hence
it loses all mechanisms to convert the mechanical work done by the external agent
to another form of energy. These altogether indicate the possibility of describing all
the stages, from the elastic through fracture, based on the same field equations.
One quite interesting feature of the field equations is that they are analogous to
Maxwell’s equations of electrodynamics. As will be explained in various sections
throughout this book, the similarity of the present theory with electrodynamics is
not limited to the mathematical resemblance. There are a number of similarities in
the physical behaviors between the displacement field and electromagnetic field. In
fact, the physical meaning of various behaviors of the displacement field can be
interpreted based on the analogy with electrodynamics, and these interpretations
have led to further understanding of the deformation dynamics. A number of
experimental observations have also been interpreted based on the analogy, and that
has facilitated the theoretical development.
It should be noted that the present theory does not refer to a cause of deformation,
in the same sense as an equation of motion does not refer to the specifics of the
external force. The present field equations describe the relation between transla-
tional and rotational modes of displacement. The cause of irreversible deformation
is incorporated into the field equations through the source terms. It is possible to
integrate the present theory with a microscopic theory, e.g., a dislocation theory, to
deal with the cause of deformation. It is an important future subject.
1.2 Aim, Scope, and Organization of This Book 5

1.2 Aim, Scope, and Organization of This Book

The primary aim of this book is to introduce the field theoretical approach to
deformation and fracture. The theoretical foundation is described and supporting
experiments are discussed. It should be emphasized that the present theory is still
developing. This book is not to present the completed form of the theory; rather it is
to invite researchers to consider the viewpoints of the present theory and hopefully
apply the approach to their own purpose.
The materials that this book tries to cover are quite interdisciplinary. It is likely
that most engineers are unfamiliar with the concept of local symmetry and gauge
transformation. Continuum mechanics is not a subject that scientists deal with on a
regular basis. It is my intention to describe the big picture of the approach, rather
than going into details of the content in each discipline. Those who are interested
in more detailed information are encouraged to read books or other resource of the
subject field.
It is also my intention to invite people of various disciplines, engineers, scientists,
and technicians. This is because exchanges of opinions among researchers in
different disciplines are most important for further development of this approach.
For this reason, much efforts have been made to explain the concepts of each
discipline in such a way that people in other disciplines can digest them as easily as
possible. Special attention is paid so that the reader does not need prior knowledge
except for basic physics and engineering. Basic knowledge on electrodynamics,
gaseous electronics, solid mechanics, and quantum mechanics will be helpful, but
not prerequisite. For this purpose, whenever seems necessary, extra explanations
are added. In some occasions, the mathematical derivation may appear lengthy and
perhaps redundant.
The objectives of each chapter are as follows. It is intended that each chapter is,
to some extent, self-contained so that the reader can use them as a reference. For
instance, those who are interested in applying an optical interferometric technique
to mechanical analysis, Chap. 6 will be useful. In this chapter, basic concepts of light
and optics which may not appear directly related to the interferometric techniques
are described. The purpose of the provision of these materials is to facilitate the
interferometric experiments. Conducting an optical interferometric experiment with
and without these knowledges make a huge difference in the experimental efficiency
and the quality of the results. The interferometric fringe-contrast will be much better,
for example, if we use an optical interferometer with the proper understanding of
the coherence in laser light, as opposed to simply follow the procedures to get
outputs from the interferometer. Other chapters are written with the same general
philosophy.
Chapter 2 reviews continuum mechanics. Basic physical concept of elasticity and
mathematical description such as the strain tensor, stress tensor, and constitutive
equations are discussed. Some of the continuum-mechanical concepts and mathe-
matical expressions are used in the present field approach. The goal of this chapter is
to facilitate the description of the mathematical procedure used to derive the present
6 1 Introduction

field equations in Chap. 5. Those who are familiar with continuum mechanics can
skip this chapter.
Chapter 3 discusses the gauge transformation and various underlying concepts
such as symmetry in physics, covariant derivatives, local symmetry, and gauge
potential. These concepts are not easy to digest for those who are not familiar
with the field. An effort has been made to discuss the complex concept as easily as
possible. The goal of this chapter is to describe the big picture of the concept without
going into mathematical details. Those who are familiar with gauge transformation
can skip this chapter.
Chapters 4 and 5 discuss the present field theory in detail. Chapter 4 focuses
on the formalism of the theory. The concept of gauge transformation discussed in
Chap. 3 is applied to the displacement field of a solid-state medium under plastic
deformation and the resultant field equations are discussed. Chapter 5 discusses
the physical meaning of the field equations and various concepts derived from
the field equations. One of the field equations is interpreted as the equation of
motion that governs the dynamics of a unit volume in the object under deformation.
Wave dynamics of the displacement field as solutions to the equation of motion
are discussed. The energy dissipative nature of plastic deformation is argued via
the concept of deformation charge, which is analogous to the electric charge. The
physical meanings of the charge and its interaction with the displacement field
are discussed. Through these discussions, the field equations are argued as the
governing equations of deformation for all stages; the elastic, plastic, and fracturing
stage. The form of the term representing the longitudinal force in the above-
mentioned equation of motion differentiates one stage from another.
Chapter 6 explains optics and optical interferometry that are used in the support-
ing experiments discussed in Chaps. 7 and 8. Interestingly, some of the behaviors
of light as an electromagnetic wave are analogous to the deformation wave. This is
not surprising because the field equations of the deformation field are analogous to
Maxwell equations of electrodynamics. Various similarities between the light and
deformation fields are discussed. It will help us digest the field theoretical dynamics
of deformation discussed in Chaps. 4 and 5 from a different angle.
Chapters 7 and 8 present experimental results that support the present field
theory. Chapter 7 focuses on various field theoretical concepts such as decaying,
transverse displacement-waves in the plastic regime and the deformation charge
and its behavior that causes energy dissipation. Chapter 8 discusses engineering
applications of the present theory including diagnosis of the current deformation
regime (elastic, plastic, or fracturing regime) and evaluation of load hysteresis for a
given object.
References 7

References

1. Dugas, R.: A History of Mechanics. Editions du Griffon, Neuchatel (1955)


2. Landau, L.D., Lifshitz, E.M.: Theory of Elasticity. Course of Theoretical Physics, vol. 7, 3rd
edn. Butterworth-Heinemann, Oxford (1986)
3. Antman, S.S.: Nonlinear Problems of Elasticity. Applied Mathematical Sciences, vol. 107.
Springer, New York/Budapest (1995)
4. Hill, R.: The Mathematical Theory of Plasticity. Oxford University Press, Oxford (1998)
5. Lubliner, J.: Plasticity Theory. Courier Dover, New York (2008)
6. Orowan, E.: Z. Phys. 89, 605, 614, 634 (1934)
7. Polanyi, M.: Z. Phys. 89, 660 (1934)
8. Taylor, G.I.: Proc. R. Soc. A145, 362 (1934)
9. Griffith, A.A.: Philos. Trans. A 221, 163–198 (1920)
10. Irwin, G.R.: “Fracture Dynamics,” Fracturing of Metals. American Society for Metals,
Cleveland (1948)
11. Timoshenko, S.P., Goodier, J.N.: Theory of Elasticity. McGraw-Hill, New York (1951)
12. Panin, V.E., Grinaev, Yu.V., Egorushkin, V.E., Buchbinder, I.L., Kul’kov, S.N.: Sov. Phys. J.
30, 24–38 (1987)
13. Panin, V.E.: Wave nature of plastic deformation. Sov. Phys. J. 33(2), 99–110 (1990)
14. Danilov, V.I., Zuev, L.B., Panin, V.E.: Wave nature of plastic deformation of solids. In: Panin,
V.E. (ed.) Physical Mesomechanics and Computer-aided Design of Materials, vol.1, p. 241.
Nauka, Novosibirsk (1995) (Russian)
15. Panin, V.E.: Physical fundamentals of mesomechanics of plastic deformation and fracture of
solids. In: Panin, V.E. (ed.) Physical Mesomechanics of Heterogeneous Media and Computer-
Aided Design of Materials. Cambridge International Science Publishing, Cambridge (1998)
16. Ibragimov, N.H.: Transformation Groups and Lie Algebras. World Scientific, Singapore (2013)
17. Elliott, J.P., Dawber, P.G.: Symmetry in Physics, vol. 1. Macmillan, London (1984)
18. Yoshida, S., Siahaan, B., Pardede, M.H., Sijabat, N., Simangunsong, H., Simbolon, T.,
Kusnowo, A.: Phys. Lett. A 251, 54–60 (1999)
19. Yoshida, S.: Phys. Mesomech. 11, 140–146 (2008)
20. Yoshida, S.: Scale-independent approach to deformation and fracture of solid-state materials.
J. Strain Anal. 46, 380–388 (2011)
Chapter 2
Quick Review of Theories of Elastic Deformation

This chapter describes conventional approaches to elastic deformation known as


continuum mechanics or theory of elasticity. The goal of this chapter is to discuss
mathematical descriptions of kinematic and dynamics that these conventional
theories use. Some of them are basis of the present field theory. It is not my
intention to cover a wide area of the subject of elastic deformation. Rather, it
is to prepare for the mathematical procedures developed in later chapters where
we derive various equations of the present theory and interpret their physical
meanings via comparison with conventional approaches. For this purpose, some
of the concepts are viewed from different angles than conventional theories of
elasticity. For complete description of elastic deformation, the reader is encouraged
to refer to other books [1–6].

2.1 Displacement and Deformation

Consider point P1 located at a coordinate point .x1 ; y1 ; z1 / in an object. We express


!
this point with a position vector OP1 as illustrated in Fig. 2.1. Similarly, another
point in the same object P2 at .x2 ; y2 ; z2 / can be expressed with position vector
!
OP2 . Apparently, the coordinates .x; y; z/ identify position of different points in the
same object, and the coordinate origin can be viewed as a reference point affixed to
the object. Such coordinates are referred to as Lagrangian coordinates.
Now consider in Fig. 2.2 that point P1 is displaced to another point P10 . To express
the displacement, we introduce a displacement vector . As a three-dimensional
vector,  has three components that can be expressed with a coordinate system,
.x ; y ; z /. While both .x; y; z/ and .x ; y ; z / have the dimension of length and
are three-dimensional vectors, the meaning of these two vectors are different from
each other. The set of components .x; y; z/ identify the position in a given object

© Springer Science+Business Media New York 2015 9


S. Yoshida, Deformation and Fracture of Solid-State Materials,
DOI 10.1007/978-1-4939-2098-3__2
10 2 Quick Review of Theories of Elastic Deformation

!
Fig. 2.1 Position vector OP1 z
representing a point in an
object P1 (x,y,z)

OP1
y

Fig. 2.2 Point P1 and point P2' dx


P2 displaced for different
amounts
x(r + dr)
P2 dr'
dr
P1' dr' = dr + dx
P1 x(r)

relative to the origin .x; y; z/ D .0; 0; 0/. On the other hand, .x ; y ; z / does not
identify a point in the object. Instead, it represents the change in the position of a
point in the object as a result of some physical event such as exertion of a force by an
external agent. Thus, their reference point is not affixed to the object but rather to the
frame of analysis. This type of coordinates is referred to as Eulerian coordinates.1
Deformation is defined as the situation where different points in the same
object are displaced differently. Note that as will be discussed in detail below, the
coordinate point after displacement is the addition of the initial position vector
and the displacement vector. As such, the position vector after the displacement
is expressed in the Eulerian coordinates. That is why some books [5] state that the
point before deformation is in Lagrangian coordinates and that after is in Eulerian
coordinates.
Figure 2.2 illustrates the situation where points P1 and P2 are displaced
differently as the object is deformed. Here the two points are considered to be
!
separated by an infinitesimal distance dr.

1
Consider you are still at .x; y; z/ on the earth (the object). Since you are not moving on the
earth, the value of .x; y; z/ does not change. However, since you are moving with the earth, the
displacement vector representing your motion with reference to the sun keeps changing.
2.1 Displacement and Deformation 11

! ! !
drDOP2  OP1 : (2.1)

By the deformation, the two points are displaced by  1 and  2 .

! !
OP10 D OP1 C 1
! !
OP20 D OP2 C 2 (2.2)

Thus, the distance between the two points after the deformation is expressed as
! ! ! !
dr 0 DOP20  OP10 Ddr C. 2   1 /: (2.3)

The two displacement vectors can be viewed as two values of a common displace-
ment function .x; y; z/ at points .x1 ; y1 ; z1 / and .x2 ; y2 ; z2 /. Further, as the distance
between points P1 and P2 is infinitesimally small, we can put

 2   1 D .x2 ; y2 ; z2 /  .x1 ; y1 ; z1 / D .x1 C dx; y1 C dy; z1 C d z/  .x1 ; y1 ; z1 /


!
D d  .x; y; z/: (2.4)

Therefore, the change in the infinitesimal distance due to the deformation can be put
as follows:
! !
dr 0 Ddr Cd .x; y; z/: (2.5)

Here, for the purpose of generalization, the suffix is dropped in the rightmost term.
Since each component of the displacement vector is a function of .x; y; z/.

@i @i @i


d i D dx C dy C d z; i D x; y; z: (2.6)
@x @y @z

In matrix notation, d  can be put


0 1
d x
d i D @d y A
d z
0 @ @ @x 1
x x
(2.7)
B @x
B @ @y @z C 0dx 1
B y @y @y C
C@ A
DB C dy
B @x @y @z C
@ @z @z @z A d z
@x @y @z
12 2 Quick Review of Theories of Elastic Deformation

Hence, Eq. (2.5) becomes


0 @ @x @x 1
x
0 0 1 0 1 B @x dx C
dy C dz
B @ @y @z C
@y C
dx dx
@dy 0 A D @dy A C B C
y @y
B dx C dy C d zC
0 B @x @y @z C
dz dz @ @z @z @z A
dx C dy C dz
@x @y @z
0 0 @ @ @ 11 (2.8)
x x x
0
B 100 1 B @x @y @z CC 0 dx 1
B B @ @ @ CC
B B y y CC
D B@0 1 0A C B CC @ dy A
y
B B @x @y @z CC
@ 001 @ @z @z @z AA d z:
@x @y @z

More concisely,
 
i 0 @i
.dx / D ıij C j dx j  Uij dx j : (2.9)
@x

@i @i
Here ıij is the Kronecker’s delta. The matrices j and Uij D ıij C j are called
@x @x
the displacement gradient tensor and the deformation gradient tensor, respectively.
It is convenient to divide the displacement gradient tensor into the symmetric and
asymmetric terms.
0 1 0    1
@x @x @x @x @x @y 1 @z @x
B @x
1
C C
B @y @z C B
C B  @x 
2
@y @x 2  @x @z CC
B @y @y @y C B 1 @x @ @ 1 @ @ y C
B CDB C
y y z
C C
B @x @y @z C B @z C
B C B 2  @y @x   @y  2 @y C
@ @z @z @z A @ 1 @x @z 1 @z @y @z A
C C
@x @y @z 2 @z @x 2 @y @z @z
0    1
1 @y @x 1 @x @z
B 0   
B   2 @x @y 2 @z @x C C
B 1 @y @x 1 @z @y C
CBB  0   C
B 2 @x @y    2 @y @z C C
@ 1 @x @z 1 @z @y A
   0
2 @z @x 2 @y @z
(2.10)

As Fig. 2.3 illustrates, the symmetric part represents strain and the asymmetric part
rotation. The former is referred to as the strain tensor and the latter as the rotation
tensor. They can be concisely expressed as follows:
2.1 Displacement and Deformation 13

Fig. 2.3 Strain and rotation


part of deformation tensor

Normal strain +

Rotation

Shear strain

@j @i
ij D i
C j; (2.11)
@x @x

@j @i
!ij D i
 j; (2.12)
@x @x
The strain tensor can further be divided into the normal strain and shear strain terms.

ij D n C sh : (2.13)

Here,
0 1
@x
B @x 0 0 C
B C
B @ y C
n D B 0 0 C (2.14)
B @y C
@ @z A
0 0 ;
@z
0   1
@x @y @z @x
B 0 C C
B  @y @x  @x @z C
C
1B @x @y @z @y C
sh D B C 0 C C (2.15)
2B
B  @y @x    @y @z CC
@ @x @z @z @y A
C C 0
@z @x @y @z

For simplicity, let’s define the line element vector  for before and 0 for after the
deformation.
14 2 Quick Review of Theories of Elastic Deformation

0 1
dx
 D @dy A (2.16)
dz
0 01
dx
0 D @dy 0 A (2.17)
d z0

With these expressions, the deformation gradient tensor and the deformation as
a transformation become

Uij D ıij C n C sh C !ij (2.18)


0 D U  (2.19)

It is worth exploring the physical meaning of each term of the deformation


gradient tensor. Consider two-dimensional deformation in the x-y plane in Fig. 2.4
where the x and y components of an infinitesimal line element vector dx and dy
are transformed to dx 0 and dy 0 , respectively, by deformation. Since translational
displacement is not of our interest, the tail of the line element vector after the
deformation is shifted to that of the before deformation. For simplicity, we consider
a two-dimensional case here but the same argument holds for three dimensions.
From Fig. 2.4,
! ! !
dx 0 D dx C AA0
! ! !
dy 0 D dy C BB 0 (2.20)

∂xy B'
dy
∂y
B B"
∂xx
dy
∂y

dy' A" A'


dy

dx'
∂xy
dx
∂x
Fig. 2.4 Physical meaning of O
dx A ∂xx
each term of deformation dx
gradient tensor ∂x
2.1 Displacement and Deformation 15

Considering the spatial dependence of the displacement vector  D x xO C y y,


O

!
   
0 @x @y
AA D dx xO C dx yO
@x @x
!
   
@x @y
BB 0 D dy xO C dy yO (2.21)
@y @y

So,
!

  
0 @x @y
dx D dx xO C dx xO C dx yO (2.22)
@x @x
!
   
0 @x @y
dy D dy yO C dy xO C dy yO (2.23)
@y @y

With Eqs. (2.16) and (2.17) substituted into Eq. (2.19), and Fig. 2.4 along with
Eqs. (2.22) and (2.23), the meaning of each term of Uij in Eq. (2.18) can be
interpreted as follows.
The first term, the unit matrix ıij , represents the undeformed part of the line
element, i.e., the first terms of Eqs. (2.22) and (2.23). The second term n represents
the normal strain expressed by the second term of Eq. (2.22) and the third term of
Eq. (2.23).
The geometric meaning of the third term sh representing the shear strain
!
becomes clear from the following discussion. Consider the scalar product of dx 0
!
and dy 0 .

! !          
@x @y @x @y
dx 0  dy 0 D dx C dx xO C dx yO  dy xO C dy C dy yO
 @x  @x @y @y
@y @x
 C dxdy (2.24)
@x @y

where
 the
 second-order
  terms of the derivatives of the displacement
@x @x
etc: are neglected. By definition, this scalar product can be
@x @y
written as follows:
ˇ ! ˇˇ ! ˇ
! ! ˇ ˇˇ ˇ
dx  dy D ˇˇdx 0 ˇˇ ˇˇdy 0 ˇˇ cos  0
0 0

s s
       
@x 2 @y 2 @x 2 @y 2
D 1C C C 1C cos  0 dxdy
@x @x @y @y

 cos  0 dxdy; (2.25)


16 2 Quick Review of Theories of Elastic Deformation

! !
@
where  0 is the angle between dx 0 and dy 0 , and @x
; y << 1 (small deformation
@x @y
or the length change x ; y is much smaller than dx; dy) is used. Equating the
right-hand side of Eqs. (2.24) and (2.25), we obtain
   
@y @x 
C D cos  0 D sin  0   0 (2.26)
@x @y 2 2

Here the small angle approximation is used for 2  0 . Equation (2.26) indicates that
  ! ! ! !
@y
@x
C @x
@y
is the change in the angle between dx and dy to dx 0 and dy 0 caused
by the deformation.
The geometrical meaning of the third term !ij in Eq. (2.18) can be understood in
! !
a similar fashion. Consider the vector product of dx and dx 0

! !
    
@x @y @y 2
dx  dx 0 D dx xO  dx C dx xO C dx yO D dx zO (2.27)
@x @x @x

By definition

! !
dx  dx 0 D dx dx 0 sin x zO  dx 2 sin x zO (2.28)
! !
where x is the angle between dx  dx 0 . From Eqs. (2.27) and (2.28),

@y
D sin x  x : (2.29)
@x
Similarly, we can easily find that

@x
D sin y  y : (2.30)
@y

From Eqs. (2.29) and (2.30), we find that

@y @x
 D x  y ; (2.31)
@x @y
!
i.e., the rotation represents the difference between the angle of rotation of dx
!
and dy.
2.2 Hooke’s Law and Poisson’s Ratio 17

2.2 Hooke’s Law and Poisson’s Ratio

The argument of the preceding section does not yet refer to dynamics because
the concept of force has not been introduced. The underlying force law in elastic
deformation is essentially Hooke’s law, which states that elastic media exert resistive
force proportional to the displacement from the equilibrium position in response to
an external force. Here the equilibrium position is the position that a part of the
medium takes when there is no external force acting on that part. Different parts of
a given elastic object have their own equilibrium positions that are different from
one another. The simplest way to visualize this type of force is a series of point
masses connected with springs, as illustrated in Fig. 2.5. In this context, the i th
mass’ equilibrium is established when neither of the springs directly connected to
this mass is stretched or compressed. Note that the force exerted by other stretched
springs are not external force on the i th mass.
The force proportional to the displacement from the equilibrium position can
be evaluated by differentiating the potential energy Up .x/ with respect to the
space coordinate. Here x is the distance from the equilibrium position. By Taylor-
expanding the potential energy with respect to x, we can express the situation as
follows:
1
U.x/ D U.0/ C U 0 .0/x C U 00 .0/x 2 C    (2.32)
2

F .x/ D U 0 .0/ C U 00 .0/x C    (2.33)

ith

ith

Displacement of ith mass

ith

stretched unstretched stretched

Fig. 2.5 Spring mass system to represent elasticity. All springs are unstretched (top). All springs
are stretched (middle). Springs connected to the i th mass are unstretched and all other springs are
stretched (bottom)
18 2 Quick Review of Theories of Elastic Deformation

Fig. 2.6 Normal stress acting l Dl


on x surface

Ax
f

At the equilibrium x D 0, the force must be zero because the atom is at the bottom
of the potential well. Thus, U 0 .0/ D 0. From this viewpoint, the spring force is
interpreted as the case when we take up to the second-order term of the potential
function. By putting U 00 .0/ D k 2 where k is the stiffness or the spring constant of
the spring, we can rewrite Eq. (2.33) as follows:

F .x/ D kx (2.34)

To discuss deformation, the total external force has much less meaning than force
per unit area. The force exerted on a unit area is called the stress. The concept of
stress is most conveniently explained through consideration of a cube in a medium
as shown in Fig. 2.6. In a one-dimensional case where the force is in the positive
x-direction, Eq. (2.34) leads to3

xx D Exx (2.35)

Here xx is the normal stress, xx is the normal strain, and E is the Young’s
modulus. Here Young’s modulus represents the stiffness, as the spring constant does
in Eq. (2.34). Since xx is acting on a plane of unit area, its dimension is “force per
area,” or in the SI unit N=m2 . Also, as clear from Fig. 2.6, the force associated with
xx is differential force, or the difference in the normal force acting on the left
surface of the cube and the right surface. Each of these forces is the spring force at
the corresponding surface. If the normal force acting on the two surfaces is the same,
the cube would not be stretched or compressed; it would simply be accelerated as
a rigid body. Accordingly, the quantity multiplied to the stiffness on the right-hand
side must be “stretch per length,” or strain. The unit of the stiffness E is therefore
.N=m2 /=.m=m/ D N=m2 .

f
xx D (2.36)
Ax

2
A negative sign is used on the right-hand side to emphasize that the force is centripetal, or opposite
to the displacement from the equilibrium.
3
See Sect. 2.4.1 for further descriptions about the Young’s modulus in the context of dynamics.
2.2 Hooke’s Law and Poisson’s Ratio 19

Note that the first subscript in xx denotes the plane defined by the direction of
the outward unit vector normal to the plane and the second subscript denotes the
direction of the force acting on the plane.
It is useful to consider the relation between the spring constant and the Young’s
modulus. Consider an elastic material (spring) of length l and its cross-sectional
area Ax being stretched by force f . Denoting the stretch of this material with l,
we can relate f and l with the spring constant k.4

f D k  l (2.37)

Here the direction of the stretch is x. Then the normal strain xx is

l
xx D (2.38)
l
Substituting Eqs. (2.36) and (2.38) into Eq. (2.37), we obtain

xx A D .kl/xx (2.39)

Comparing Eqs. (2.35) and (2.39), we find

A
kDE (2.40)
l

Notice that the spring constant k depends on the size of the object, whereas the
stiffness (Young’s modulus) E is a material constant. Below, p we will find that
the phase velocity of one-dimensional longitudinal wave is E=, where  is the
density of the material. This indicates that the longitudinal wave’s velocity, e.g,
sound velocity, is uniquely determined by the material. This is contrastivepto the
angular resonant frequency of an object of mass m and spring constant k is k=m;
it depends on the object’s size (mass).
In the above argument, the differential force is normal to the surface. What if
the differential force is parallel to the surface; for example, the force acting on the
top surface of the cube in Fig. 2.6 parallel to the surface is different from the force
acting on the bottom surface parallel to the other force on the top surface? The cube
will experience shear strain as shown in Fig. 2.3. The stress corresponding to the
transversely differential force is referred to as the shear stress and the constant of
proportionality in its relation to the shear strain is referred to as the shear modulus G.

yx D Gyx (2.41)

4
To avoid complexity, only the absolute value of the force is considered here.
20 2 Quick Review of Theories of Elastic Deformation

Fig. 2.7 Stress vectors y

sz
B sn
sx

o A x
C

z
sy

Here, as is the case of the normal stress xx , the first subscript y indicates the plane
that the force is acting on and the second subscript x indicates the direction of the
force.
So far, the direction of the differential force has been one-dimensional. To extend
this to three dimensions, we need to treat the stress as a three-dimensional vector.
Consider a stress on a given plane ABC in Fig. 2.7. From the force equilibrium on
the infinitesimally small tetrahedron OABC .5

 n dSn D  x dSx C  y dSy C  z dSz (2.42)

Here  k ; k D n; x; y; z is the stress vector acting on plane k and dSk is the area
of plane k. Considering that the ratio dsi =dsn ; i D x; y; z is the direction cosine
of a normal vector of triangular area ABC with axis i , Eq. (2.42) can be put in the
following form:

 n D nx  x C ny  y C nz  z (2.43)

where ni ; i D x; y; z is the direction cosine of the normal vector to the axis xi .


Now the stress vector acting on each of the x, y, z-plane can be expressed in terms
of the unit vector xO i .

 x D xx xO C xy yO C xz zO (2.44)


 y D yx xO C yy yO C yz zO (2.45)
 D zx xO C zy yO C zz zO
z
(2.46)

The coefficients of the unit vectors can be put in the form of a tensor referred to as
the stress tensor as follows:

5
The effect of body force such as gravity is omitted as it does not have a substantial effect in the
argument in this chapter.
2.2 Hooke’s Law and Poisson’s Ratio 21

0 1
xx xy xz
Œij
D @yx yy yz A (2.47)
zx zy zz

From Eq. (2.43) and Eqs. (2.44)–(2.46),


 
 n D nx xx C ny yx C nz zx xO C nx xy C ny yy C nz zy yO

C nx xz C ny yz C nz zz zO
 . ni ix /xO C . ni iy /yO C . ni iz /Oz (2.48)

Here row i represents the stress vector  i introduced in Eq. (2.43), and column j
represents the j -component of each stress vector  i . In the last line of Eq. (2.48),
the summation over index i is omitted for simplicity. So, for example, yz is the
z component of stress vector  y , which is acting on plane y. This convention is
consistent with that for Eq. (2.36).
Now consider relations among stress tensor components. Considering force
equilibrium for a given volume enclosed by surface S , we obtain the following
equation.
Z Z Z Z

 dS D
n
nx xi C ny yi C nz zi dS D 0 (2.49)
S S

where n is the unit vector normal to the surface S ( n is the component of 


normal to the surface dS) and Eq. (2.48) is used in going through the second equal
sign. Using Green theorem, Eq. (2.49) can be converted to an integration over the
volume as
Z Z Z  
@xi @yi @zi
C C dV D 0 (2.50)
V @x @y @z

In order for Eq. (2.50) to hold in any volume, it follows that

@xi @yi @zi


C C D0 (2.51)
@x @y @z

or for each component

@xx @yx @zx


C C D0
@x @y @z
@xy @yy @zy
C C D0
@x @y @z
@xz @yz @zz
C C D0
@x @y @z
22 2 Quick Review of Theories of Elastic Deformation

Next considering rotational equilibrium, we obtain


Z Z
r   n dS D 0 (2.52)
S

Here r D x xO C y yO C zOz. With the use of Eq. (2.48) and Green theorem, this leads to
Z Z Z      
@xz @yz @zz @xy @yy @zy 
xO y C C z C C C yz  zy
V @x @y @z @x @y @z
CyO Π 
C zO Π 
(2.53)

From Eq. (2.52), the first part of the angle bracket Œ


is zero. This leads to the
following relation between ij and its diagonal counterpart j i as

xy D yx ; yz D zy ; zx D xz (2.54)

This indicates that the stress tensor Eq. (2.47) is symmetric.


0 1
xx xy zx
Œij
D @xy yy yz A (2.55)
zx yz zz

With the stress tensor (2.55), the one-dimensional constitutive equation (2.35) can
be extended to three-dimensions as follows:

ij D Cijkl kl (2.56)

Here ij is the .i; j / component of strain tensor defined by Eq. (2.11). In accordance
with the above convention, ij denotes the j component of the stress vector acting
on plane i . Similarly, kl denotes the .k; l/ component of the strain tensor. Each
combination of k and l represents the degree of freedom in deformation; e.g.,
.k; l/ D .x; y/ represents the shear deformation in a plane parallel to the x-y plane
as defined by Eq. (2.57). The coefficient Cijkl represents the response of the material
to the external force represented by ij for each degree of freedom.
 
1 @y @x
xy D C (2.57)
2 @x @y

With the use of symmetry of ij and ij , eq. (2.56) can be expressed in a matrix
form as follows:
2.2 Hooke’s Law and Poisson’s Ratio 23

Fig. 2.8 Meanings of C12


e yy l Dl
and C21
Dl

l
e xx
s xx
s yy
y
z x

0 1 0 10 1
xx C11 C12 C13 C14 C15 C16 xx
B C BC   B
C26 C Byy C
C
B yy C B 21 C22 C23 C
B C B CB C
B zz C BC31        C B zz C
B CDB CB C (2.58)
Bxy C BC41        C Bxy C
B C B CB C
@ yz A @C51        A @ yz A
zx C61     C66 zx

The matrix Ckl is called the stiffness tensor. Consider the meanings of C12 and
C21 . If we write only the relevant term in xx and yy in Eq. (2.58), we obtain the
following expressions.

xx D C12 yy (2.59)


yy D C21 xx (2.60)

As Fig. 2.8 illustrates, Eq. (2.59) describes how much the volume expands normally
along y-axis when the volume is subject to the normal stress applied on the x-plane.
Similarly, Eq. (2.60) describes how much the volume expands normally along x-axis
when the volume is subject to the normal stress applied on the y-plane. Apparently,
the volume does not know which axis is x or y, meaning that if we switch the x-axis
with the y-axis, the physical situation is unchanged. It is obvious that Ckl D Clk ,
or the stiffness matrix is symmetric. We can repeat the same argument for C13 and
C31    to find there are C.6; 2/ D 15 (the number of 2-combinations from a set
of 6-elements) redundancies. This reduces the number of independent Ckl from the
total number of 6  6 D 36 to 21.
In the case that the material’s response is symmetric, e.g., symmetric about the
x-y plane, for example, we can further reduce the number of the elements in the
stiffness matrix. Consider a new coordinate system x 0 y 0 z0 that is symmetric with
xyz system about the x-y plane; x D x 0 , y D y 0 and z D z0 . With this coordinate
transformation, the sign of the z-component of a given vector is flipped, but the signs
of the x and y components are unchanged. Therefore, xz D x 0 z0 , yz D y 0 z0
and the sign of the other stress tensor components is unchanged. Similarly, yz D
y 0 z0 , zx D z0 x 0 and the other strain tensor components are unchanged. Now
consider the normal stress on the x D x 0 plane with the two coordinate systems.
24 2 Quick Review of Theories of Elastic Deformation

Table 2.1 Coordinate x y z


transformation
O
.x/ O
.y/ .Oz/
x0 x 0 x x 0 y x 0 z
.xO0 / .xO0  x/
O .xO0  y/
O .xO0  zO/
y0 y 0 x y 0 y y 0 z
.yO0 / .yO0  x/
O .yO0  y/
O .yO0  zO/
z0 z0 x z0 y z0 z
.zO0 / .zO0  x/
O .zO0  y/
O .zO0  zO/

x 0 x 0 D C11 x 0 x 0 C C12 y 0 y 0 C C13 z0 z0 C C14 x 0 y 0 C C15 y 0 z0 C C16 z0 x 0


D C11 x 0 x 0 C C12 y 0 y 0 C C13 z0 z0 C C14 x 0 y 0  C15 y 0 z0  C16 z0 x 0 (2.61)
xx D C11 xx C C12 yy C C13 zz C C14 xy C C15 yz C C16 zx (2.62)

As mentioned above, this coordinate transformation does not change the normal
stress on the x-plane; x 0 x 0 D xx . It follows that C15 D C16 D 0. Repeating the
same argument for the symmetry about the yz and zx-plane, we can simplify the
stiffness matrix as follows:
0 1
C11 C12 C13 0 0 0
BC C C C
B 12 22 23 0 0 0 C
B C
BC C C 0 0 0 C
ŒCkl
D B 13 23 33 C (2.63)
B 0 0 0 C44 0 0 C
B C
@ 0 0 0 0 C55 0 A
0 0 0 0 0 C66

Next, consider that the material is isotropic. In this case, the constitutive relation
should remain the same if we rotate the material around an axis. First examine how
the stress vector is transformed by coordinate transformation from O  xyz to O 
x 0 y 0 z0 . Referring to Eq. (2.43),
0
 x D x 0 x  x C x 0 y  y C x 0 z  z (2.64)

where x 0 x etc are the direction cosine (see Table 2.1). Substituting Eqs. (2.44)–
(2.46) into the right-hand side of Eq. (2.64), we obtain
0  
 x D x 0 x xx C x 0 y yx C x 0 z zx xO C x 0 x xy C x 0 y yy C x 0 z zy yO

C x 0 x xz C x 0 y yz C x 0 z zz zO
 x 0 i ix xO C x 0 i iy yO C x 0 i iz zO (2.65)

Now by definition, x 0 x 0 is the xO0 component of  x .


0

0 
x 0 x 0 D  x  x 0 x xO C x 0 y yO C x 0 z zO
2.2 Hooke’s Law and Poisson’s Ratio 25

Table 2.2 Coordinate x y z


transformation 0
x cos  sin  0
y0  sin  cos  0
z0 0 0 1

D x 0 x x 0 i ix C x 0 y x 0 i iy C x 0 z x 0 i iz


 x 0 j x 0 i ij (2.66)

Applying the same procedure as Eq. (2.66) to the other components, we find the
general expression to relate the stress tensor components before and after the
coordinate transformation as follows.6

i 0 j 0 D i 0 i j 0 j ij (2.67)

Now transform the coordinate system around the z-axis and make analysis on the
constitutive relation. Referring to Table 2.2,

x 0 x 0 D x 0 x x 0 x xx C x 0 y x 0 x yx C x 0 x x 0 y xy C x 0 y x 0 y yy


D xx cos2  C yx sin  cos  C xy cos  sin  C yy sin2 
D xx cos2  C yy sin2  C 2xy cos  sin 
y 0 y 0 D xx sin2  C yy cos2   2xy cos  sin 
z0 z0 D zz
 
x 0 y 0 D yy  xx cos  sin  C xy cos2   sin2 
y 0 z0 D yz cos   zx sin 
z0 x 0 D yz sin  C zx cos  (2.68)

Similarly, the strain tensor components are expressed with the new coordinate
system as follows:

x 0 x 0 D x 0 x x 0 x xx C x 0 y x 0 y yy C x 0 y x 0 x yx C x 0 x x 0 y xy


D xx cos2  C yy sin2  C 2xy cos  sin 
y 0 y 0 D xx sin2  C yy cos2   2xy cos  sin 
z0 z0 D zz
 
x 0 y 0 D yy  xx cos  sin  C xy cos2   sin2 

6
In Eq. (2.67) the order of the coefficients x 0 j x 0 i in Eq. (2.66) is switched for better visibility.
26 2 Quick Review of Theories of Elastic Deformation

y 0 z0 D yz cos   zx sin 


z0 x 0 D yz sin  C zx cos  (2.69)

Consider expressing x 0 x 0 in two different ways. First, express the stress tensor
component on the right-hand side using the strain-tensor components ij and the
stiffness tensor components in Eq. (2.63).

x 0 x 0 D xx cos2  C yy sin2  C 2xy cos  sin 



D C11 xx C C12 yy C C13 zz cos2 

C C12 xx C C22 yy C C23 zz sin2 
C2C44 xy cos  sin 

D xx C11 cos2  C C12 sin2 

Cyy C12 cos2  C C22 sin2 

Czz C13 cos2  C C23 sin2 
Cxy .2C44 / cos  sin  (2.70)

Next, express x 0 x 0 using i 0 j 0 and the stiffness tensor components in Eq. (2.63), and
convert the strain component expression after the coordinate transformation with
those of before the transformation, ij .

x 0 x 0 D C11 x 0 x 0 C C12 y 0 y 0 C C13 z0 z0



D C11 xx cos2  C yy sin2  C 2xy cos  sin 

CC12 xx sin2  C yy cos2   2xy cos  sin 
CC13 zz

D xx C11 cos2  C C12 sin2 

Cyy C11 sin2  C C12 cos2 
Czz C13
Cxy 2 .C11  C12 / cos  sin  (2.71)

Compare Eqs. (2.70) and (2.71) for the coefficients of the same strain tensor matrix,
ij . In order for x 0 x 0 expressed in these two ways to be the same for any angle ,
the following equalities are necessary for the coefficients of yy , zz , and xy .
From yy

C12 cos2  C C22 sin2  D C12 cos2  C C11 sin2  (2.72)


2.2 Hooke’s Law and Poisson’s Ratio 27

From zz

C13 cos2  C C23 sin2  D C13 (2.73)

And from xy

2C44 cos  sin  D 2 .C11  C12 / cos  sin  (2.74)

From Eqs. (2.73)–(2.74), it follows that

C11 D C22 (2.75)


C13 D C23 (2.76)
C44 D C11  C12 (2.77)

Repeating the same procedure for coordinate rotations about the x and y axes, we
find the following conditions for the case of isotropic materials.

C11 D C22 D C33 ; C12 D C13 D C23 ; C44 D C55 D C66 D C11  C12 (2.78)

Since C11 D C22 D C33 are related to C44 D C55 D C66 in the last
expression, Eq. (2.78) indicates that the stiffness tensor of isotropic materials has
two independent components as follows:

C12 D C13 D C23  (2.79)

C44 D C55 D C66  2 (2.80)

and are called as Lamé’s first and second constants, respectively. All these
arguments simplify the stiffness tensor for isotropic materials into the form with the
use of the two Lamé’s constants, and allows us to express the constitutive relation
as follows:
0 1 0 10 1
xx C 2 0 0 0 xx
B C B C 2 C B
0 0 0 C Byy C
B yy C B C
B C B CB C
B zz C B C 2 0 0 0 C B zz C
B CDB CB C (2.81)
Bxy C B 0 0 0 2 0 0 C Bxy C
B C B CB C
@ yz A @ 0 0 0 0 2 0 A @ yz A
zx 0 0 0 0 0 2 zx

Here the physical meaning of C44 is the stiffness that connect the shear stress xy
and shear strain xy [Eq. (2.57)] as
 
C44 @y @x
xy D C44 xy D C (2.82)
2 @x @y
28 2 Quick Review of Theories of Elastic Deformation

Fig. 2.9 Schematic


illustration of Poisson ratio

e xx

e zz
e yy

From Eq. (2.41), the shear stress can be related to the shear strain with the shear
modulus (G).
 
@y @x
xy D G C (2.83)
@x @y

From Eqs. (2.82) and (2.83), we find C44 D 2G.


Next relate the Lamé’s constants to Young’s modulus and Poisson ratio. Consider
a volume is subject to one-dimensional force in Fig. 2.9. As the volume is stretched
along the x-axis, the material is usually compressed in the orthogonal direction. For
the isotropic case, the compressions in the y and z direction are the same. The ratio
is referred to as the Poisson’s ratio.
ˇ ˇ
ˇ t ˇ
D ˇˇ ˇˇ (2.84)
n

Here n is the normal strain in the direction of the applied, one-dimensional stress,
and t is the transverse strain. In the case of Fig. 2.9, n D xx and t D yy D zz .
Putting xx D n , yy D zz D xy D yz D zx D 0, xx D n , yy D zz D t ,
xy D yz D zx D 0 in Eq. (2.81), we obtain

n D . C 2 /n C 2 t (2.85)
0 D n C 2. C /t (2.86)

From Eq. (2.86),


ˇ ˇ
ˇ t ˇ
D ˇˇ ˇˇ D (2.87)
n 2. C /

Eliminating t from Eqs. (2.85) and (2.86),

.3 C 2 /
n D n (2.88)
. C /

Therefore,
2.3 Principal Axis 29

Table 2.3 Various moduli


. ; / .E; G/ .E; / .G; /
.3 C 2 /
E E E 2.1 C /G
. C
E
G G G
2.1 C /
E  2G

2. C / 2G
G.E  2G/ E 2 G

3G  E .1 C /.1  2 / .1  2 /

n .3 C 2 /
ED D (2.89)
n . C /

Table 2.3 shows relation among different moduli.

2.3 Principal Axis

Strain tensor Eq. (2.11) is symmetric and can be diagonalized by a rotational


transformation of the coordinate system [7]. Once diagonalized, all the shear
components of the resultant strain tensor become zero; ij D 0; i ¤ j .
This means that we can always find a coordinate system with which a given
strain can be expressed in terms of normal strains only. This is understandable
because when an infinitesimally small volume at a point is deformed, the cor-
responding displacement vector has three-translational degrees of freedom, and
stretch/compression is the differential displacement along the axes of these degrees
of freedom. By aligning the coordinate axes to these three direction, we can always
express the total strain as a combination of the three normal strains.
Similarly, stress tensor Eq. (2.55) can be diagonalized, and this can be understood
as follows. Whatever coordinate system you may chose, the stress tensor vector
has three components. In other words, the stress tensor can be expressed as the
summation of the three components. The three axes of the coordinate system
corresponding to the diagonal tensor are called the principal axes.
If we apply a diagonalized strain tensor to the constitutive equation (2.81), it
is clear that the stress tensor corresponding to the left-hand side is also diagonal
because all shear stress terms ij ; i ¤ j become zero. This is not surprising
because Hooke’s law states that stretch or compression is parallel to the external
force causing it. Non-diagonal terms in the stiffness tensor such as C12 that connects
a normal stress xx and normal strain yy in an orthogonal direction is not a direct
consequence of Hooke’s law. It is not associated with a elastic modulus but instead
due to the Poisson’s effect represented by the Poisson ratio . In other words, linear
elastic deformation is an orientation preserving mapping [2].
30 2 Quick Review of Theories of Elastic Deformation

The procedure to make a diagonal tensor from a square tensor is known as


diagonalization. Take a moment to review the procedure and discuss the physical
meaning. Consider diagonalization of a strain tensor expressed in the symmetric
form as follows:
0 10 1 0 10 1
xx xy zx dx1 dx2 dx3 1 0 0 dx1 dx2 dx3
@xy yy yz A @ dy1 dy2 dy3 A D @ 0 2 0 A @ dy1 dy2 dy3 A (2.90)
zx yz zz d z1 d z2 d z3 0 0 3 d z1 d z2 d z3

Rewrite Eq. (2.90) in the following form


20 1 0 13 0 1
xx xy zx 1 0 0 dx1 dx2 dx3
4@xy yy yz A  @ 0 2 0 A5 @ dy1 dy2 dy3 A D 0 (2.91)
zx yz zz 0 0 3 d z1 d z2 d z3

and consider the first column of Eq. (2.91).


0 10 1
xx  1 xy zx dx1
@ xy yy  2 yz A @ dy1 A D 0 (2.92)
zx yz zz  3 d z1

The physical meaning of Eq. (2.92) is as follows. If we multiply a strain matrix


to the infinitesimal line element vector expressed in the x1 y1 z1 coordinate system,
the resultant differential displacement vector associated with the strain has only a
normal strain component; d xx ¤ 0, d xy D d zx D 0. In order for this condition
to be true for a nonzero line element vector, it is necessary that Eq. (2.92) is not
solvable as a set of linear equations. This condition is expressed as the determinant
of the matrix is zero. That is,
0 1
xx  1 xy zx
det @ xy yy  2 yz A D 0 (2.93)
zx yz zz  3

This leads to the following characteristic equation, i ; i D 1; 2; 3 is one of the three


roots.

3  I1 2 C I2  I3 D 0 (2.94)

Here I1 , I2 , and I3 are known as the invariants defined as follows:

I1 D xx C yy C zz (2.95)


I2 D xx yy C yy zz C zz xx  2
xy  2
yz  2
zx (2.96)

I3 D xx yy zz  xx yz


2
 yy zx
2
 zz xy
2
C 2xy yz zx (2.97)
2.3 Principal Axis 31

Examine an example of diagonalization using a two-dimensional case. Equa-


tion (2.69) expresses how shear strain tensor components are transformed under
coordinate rotation around the z- axis by angle . For xy the transformation is
 
x 0 y 0 D yy  xx cos  sin  C xy cos2   sin2  (2.98)

What we need is to find the angle  that makes the value of Eq. (2.98) zero. Dividing
the right-hand sides of this equation by cos2 , we obtain the following quadratic
equation of tan .
 
yy  xx tan  C xy 1  tan2  D 0 (2.99)

Using the identity

2 tan 
tan 2 D (2.100)
1  tan2 
Thus, substituting Eq. (2.100) into Eq. (2.99) we find,

2xy
tan 2 D (2.101)
xx  yy

When the angle of the coordinate rotation satisfies the condition (2.101), the strain
tensor has only the normal components.
With the principal axis, the invariants take simple forms. Denoting the normal
strains with the principal axis xx D x , : : :,

I1 D x C y C z (2.102)
I2 D x y C y z C z x (2.103)
I3 D x y z (2.104)

For a given total strain, the values of I1 ; I2 ; and I3 are determined by Eqs. (2.102–
2.104) no matter what coordinate system may be used. The characteristic equa-
tion (2.94) provides us with the normal strains along the principal axes from strain
tensor components based on a given coordinate system. From this viewpoint, it can
be said that these invariants connect different frames (coordinate systems). Naively
speaking, this is similar with the deformation charge that connects different frames
each of which individually represents the local linear elasticity. The deformation
charge and related issues will be discussed in Chaps. 4 and 5.
32 2 Quick Review of Theories of Elastic Deformation

2.4 Equation of Motion and Elastic Waves

2.4.1 One-Dimensional Longitudinal Elastic Waves

Consider a block in an elastic medium in Fig. 2.10. Assume the materials at the
front and rear ends of this block are both stretched. This situation can be modeled
as a point mass at x D x connected to two springs one at x D x and the other at
x D x C dx. If the spring constant is k, the equation of motion is

d 2 d .x/ d .x C dx/


m 2
D k dx C k dx (2.105)
dt dx dx

Here  is the displacement. The first term on the right-hand side represents the force
in the negative x direction exerted by the rear spring. Being stretched, this spring
tries to return to its natural length. Consequently, it pulls the point mass to the left.
The second term is the force exerted by the front spring, which pulls the point mass
to the right as it returns to its natural length. Using the second derivative of , we
can write the spatial derivative in the second term as follows:
 
d .x C dx/ d d .x/ d .x/ d 2 .x/
D .x/ C dx D C dx (2.106)
dx dx dx dx dx 2

So, Eq. (2.105) can be put as follows:


 
d 2 d .x/ d .x/ d 2 .x/ d 2 .x/ 2
m D k  C C dx dx D k dx (2.107)
dt 2 dx dx dx 2 dx 2

In the case of actual elastic medium, replace kdx (the total force at x) with EdS
and m with dSdx. Here E is the Young’s modulus, dS is the cross-sectional area
and  is the density.

d 2 d 2 .x/
dSdx 2
D EdSdx (2.108)
dt dx 2

dx
f (x) = −k dx
dx

Fig. 2.10 Block in an elastic


medium. Net force is dx (x + dx)
differential force from the f (x + dx) = k dx
front and back ends dx
2.4 Equation of Motion and Elastic Waves 33

and finally,

d 2 d 2 .x/
 2
DE (2.109)
dt dx 2
Equation (2.109) is known as the equation of longitudinal elastic wave.

2.4.2 Three-Dimensional Compression Waves

Consider Fig. 2.11. The x component of the net external force acting on a cubic unit
volume can be found by differentiating the x components of normal stress on the
x-plane with respect to x, the x components shear stress on y-plane with respect
to y, the x components shear stress on z-plane with respect to z, and add the three
together. The y and z components of the net external force can be found in the same
fashion. Thus, the equation of motion can be expressed as follows:

@2 x @xx @yx @zx


 D C C (2.110)
@t 2 @x @y @z
@2 y @xy @yy @zy
 D C C (2.111)
@t 2 @x @y @z
@2 z @xz @yz @zz
 D C C (2.112)
@t 2 @x @y @z

Substituting each stress tensor component of the constitutive equation (2.81) into
the right-hand side of Eq. (2.110), we obtain the following equation.

@2 x @
@xy @zx
 2
D . C 2 /xx C yy C zz C 2 C 2
@t @x @y @z
 
@ @x @y @z
D . C 2 / C C
@x @x @y @z
   
@ @y @x @ @x @z
C C C C
@y @x @y @z @z @x

s yx (y + dy)
y
s zx (z)

z x s xx (x + dx)
s xx (x)

s yx (y)
Fig. 2.11 Net force on a
s zx (z + dz)
cube of unit volume
34 2 Quick Review of Theories of Elastic Deformation

   
@2 x @2 x @2 x @ @x @y @z
D C C C . C / C C
@x 2 @y 2 @z2 @x @x @y @z
D r 2 x C . C /r   (2.113)

Using the same procedure for Eqs. (2.111) and (2.112), we obtain the following set
of equations.

@2 x
 D r 2 x C . C /r   (2.114)
@t 2
@2 y
 D r 2 y C . C /r   (2.115)
@t 2
@2 z
 D r 2 z C . C /r   (2.116)
@t 2
By differentiating Eq. (2.114) by x, Eq. (2.115) by y, Eq. (2.116) by z and adding
the resultant equations, we obtain the following wave equation.

@2
 .r  / D . C 2 /r 2 .r  / (2.117)
@t 2

Equation (2.117) represents that .r / propagates as a wave with the phase velocity
p
. C 2 /=. This wave is called the three-dimensional compression wave. Here
the physical meaning of .r  / is the volume change.
By differentiating Eq. (2.115) by z. Equation (2.116) by y and subtracting the
former from the latter, and repeating this pattern for other combinations, we obtain
the equation of rotation wave as follows:

@2 !yz
 D r 2 !yz (2.118)
@t 2
@2 !zx
 D r 2 !zx (2.119)
@t 2
@2 !xy
 D r 2 !xy (2.120)
@t 2
p
Note that the rotational wave propagates with the phase velocity =.

References

1. Spencer, A.J.M.: Continuum Mechanics. Longman, London/New York (1980)


2. Marsden, J.E., Hughes, T.J.R.: Mathematical Foundations of Elasticity. Prentice-Hall, Engle-
wood Cliffs (1983)
References 35

3. Landau, L.D., Lifshitz, E.M.: Theory of Elasticity. Course of Theoretical Physics, vol. 7, 3rd
edn. Butterworth-Heinemann, Oxford (1986)
4. Antman, S.S.: Nonlinear Problems of Elasticity. Applied Mathematical Sciences, vol. 107.
Springer, New York/Budapest (1995)
5. Sciammarella, C.A., Sciammarella, F.M.: Experimental Mechanics of Solids. Wiley, London
(2012)
6. Kunio, T.: Kotai Rikigaku no Kiso (Foundation of Solid Mechanics), Baihu-kan, Tokyo (1977)
[in Japanese]
7. Boas, M.L.: Mathematical Methods in the Physical Sciences, 3rd edn. Wiley, London (2006)
(see p. 148)
Chapter 3
Quick Review of Field Theories

In this chapter, we consider the concept of “the field.” We tend to use the word
“field” or “field equations” casually. However, in a physical theory the field has a
specific meaning associated with the dynamics that the theory describes. Generally
speaking, the field can be classified into two groups; material fields and gauge
fields [1, 2]. The former is the field of physical quantities such as energy and
momentum. The gauge field has a little different meaning; it is a compensation
field to satisfy certain symmetry. A good example is the electromagnetic field as a
gauge field to make the Shrödinger equation locally symmetric [3]; in other words,
it is necessary to introduce the electromagnetic field to make a phase transformation
of a charged particle locally symmetric, or allow the phase transformation to be
coordinate dependent, as opposed to all the charged particles in the universe phase-
transform in the same fashion (globally symmetric).
The concept of “symmetry” [4] is also worth considering carefully. In physics,
symmetry means that a physical law does not depend on the choice of the coordinate
system. As an example, consider that you make a pendulum with a string of a certain
length and a bob, and measure the period. Suppose that you make this measurement
on the earth and your measured period is 10 s. If you repeat the same measurement
on the moon, the period will be longer in proportion to the square root of the ratio
of the gravitational field strength on the earth to that on the moon. However, the
relationship between the length of the string and the period, i.e., the underlying
physical law, is the same. The symmetry regarding a physical law is referred to as
invariance.
As easily imagined, symmetry is always associated with a certain operation. On
our daily life, for example, line symmetry means that a given shape appears exactly
the same when it is flipped over the line of symmetry and point symmetry means
that a shape appears the same when it is rotated around the point of symmetry
by a certain angle. Thus we say that the former is associated with an operation
of reflection (inversion) and the latter is associated with an operation of rotation.
Similarly, invariance in physical theory is associated with a certain transformation.

© Springer Science+Business Media New York 2015 37


S. Yoshida, Deformation and Fracture of Solid-State Materials,
DOI 10.1007/978-1-4939-2098-3__3
38 3 Quick Review of Field Theories

The above example of a pendulum indicates that Newton’s second law is invariant
under the transformation of a coordinate system fixed on the earth to that on the
moon.
Such symmetry can be divided into global symmetry and local symmetry. A
global symmetry is a symmetry that holds when all points in the space-time
experience the same transformation. A local symmetry is a symmetry that holds
when transformation varies from point to point. Generally, a globally symmetric
theory is not necessarily locally symmetric because local symmetry is more strict
requirement than global symmetry. While allowing a given physical system to
experience different transformation from point to point, local symmetry requires that
the underlying physical law is invariant. However, if a certain force is introduced, it
becomes possible to regain the symmetry while allowing the transformation varies
from point to point. In other words, we can introduce a force field just to compensate
the effect of the coordinate dependence of the transformation so that the dynamics
can be formulated with a common physical law anywhere and anytime. This leads
to the description of a new dynamics associated with the force. The field associated
with this force is called the gauge field, sometimes compensation field or connection
field.
The goal of this chapter is to consider these concepts without going into
mathematical details. For further details, the reader is encouraged to read books,
e.g., [4] (symmetry in physics); [1–3, 5] (gauge field and transformation); and [6]
(general relativity).

3.1 The Field

In physics, a field is defined as a quantity that has a value at each point of space-
time. If the quantity is a scalar, the field is called a scalar field and if it is a vector,
a vector field. A tensor field can be defined in the same fashion. A temperature
distribution in a three-dimensional space is an example of a three-dimensional scalar
field. As mentioned at the beginning of this chapter, a field is associated with force
and underlying dynamics. From this viewpoint, the real significance of the fact that
temperature is a function of space coordinates is that it leads to flow dynamics.
As an example, when the air of a certain region has a temperature distribution, it
is more important to analyze the air flow resulting from the temperature field than
the temperature distribution itself. When a hurricane comes to southern states of
the USA, the meteorologist keeps mentioning the water temperature of the Gulf of
Mexico. This is because the temperature gradient determines the hurricane wind
power as warm air goes up and cooler air goes down. It is easily imagined that this
is related to a certain force pattern.
Dimension-wise, force is first-order spatial differentiation of energy. If a force
field is related to potential energy in the following form, it represents energy
conservative dynamics.
3.1 The Field 39

F D r (3.1)

Proof. Take curl of F and use the identity r  r D 0 ;

r F D0 (3.2)

From the Stokes’s theorem and Eq. (3.2),


I Z Z
F  ds D .r  F /  d A D 0 (3.3)

Equation (3.3) indicates that the force field is conservative because the net work
along a closed loop is zero. t
u
From Eq. (3.1), it follows that the potential energy difference between point a
and b can be found by the following integral operation.
Z b
D F  ds (3.4)
a

Note that whereas Eq. (3.4) indicates that point b is higher than point a in potential,
an object placed at point b does not necessarily pass point a when it is released. It
goes in the direction that maximizes the absolute value of d =ds. In other words,
the field force acts on an object in the direction in which the potential changes most
steeply. From this viewpoint, the differential form Eq. (3.1) is a stronger statement
than the integral form Eq. (3.4).
The situation that an object under the influence of a field force moves in the
direction of maximum potential change is unchanged when the object cannot move
freely or under a certain constraint. In this case, the object moves along the slope of
the constraint or along the gradient of the constraint. Let’s consider a gravitational
potential field as an example of a conservative field. Imagine in Fig. 3.1 that you
form a flat surface and a non-flat surface using a sheet of cloth, and gently place

y
f y
f

Fig. 3.1 Flat and non-flat


surfaces formed by a sheet of x
cloth x
40 3 Quick Review of Field Theories

a little ping-pong ball at a certain location of the surface. It is assumed that the
magnitude of the gravitational field is constant and its direction is downward. The
ping-pong ball cannot go through the cloth, so the surface acts as the constraint. It
is easily imagined that if you place the ping-pong ball on the flat surface, it will stay
at the same location forever. This is because there is no way for the ping-pong ball
to change the gravitational potential energy on the flat surface; the constraint in this
case does not allow the ping-pong ball to change the gravitational potential energy.
If you place the ping-pong ball on the non-flat surface, it will move in the direction
where the slope is the steepest, because that is the direction where the gravitational
potential energy changes the most. Once the ball starts to move, it will keep moving
until it hits the bottom of a local well. There the ball may go back and forth around
the deepest point of the well. If there is no energy dissipating mechanism such as
friction, it will be oscillating forever. More realistically, the ball will eventually stop
oscillating because the energy dissipating mechanism uses up the potential energy.
Then the ball will stay at the bottom of the well forever for the same reason as the
flat surface case.
This dynamics can be argued in two different ways. First, we can claim that the
ball moves towards a place where the gravitational potential is the lowest. At each
moment, it keeps seeking the direction where the gravitational potential changes
most steeply under the given constraint. The friction and other nonconservative
resisting force are handled as the corresponding decrease in the total mechanical
energy. Alternatively, we can find the net external force acting on the ball at each
moment. The gravitational force is downward at any point on the surface, the normal
force is perpendicular to the surface and outward, and the friction is opposite to the
velocity of the object. The vector of the normal force and the friction varies as the
object moves around on the surface. The tangential component of the net external
force determines the motion of the ball at each moment. Call the first the potential-
based approach and the second the net-force-based approach.
The present field theory, as is the case of most field theories, treats the dynamics
using the potential-based approach. In the above example, the direction of the
steepest surface at a given point is the direction in which the gravitational potential
changes most greatly. Thus, the task of finding the conservative force can be
replaced by the task of finding the direction of the steepest slope at a give
point. Mathematically, this can be argued as follows. Say the magnitude of the
gravitational field strength is g (constant) and its direction is negative z. Then the
gravitational potential energy .z/ for mass m is .z/ D mgz. Because of the
constraint, the object cannot move along the z-axis. Instead, it can move along the
surface. Let s be the variable that represents the path of the object. Then the change
in the gravitational potential energy over the path is given as follows

d @ d z
D (3.5)
ds @z ds
3.1 The Field 41

Here @ @z
D mg is determined by the gravitational field independent of the surface.
Therefore, the change in the gravitational potential energy over the path d ds
is
dz
maximized when the slope ds is maximized. The object selects the path that
maximizes the slope ddsz .
To discuss the above dynamics more quantitatively, consider two points P and Q
on a half-pipe-like surface as shown in Fig. 3.2. The axis of the full-pipe (cylinder)
is oriented along the y-axis. At point P , the slope along the y-axis is null and that
along the x-axis is negative. You can easily imagine that if an object is released at
point P , it will slide along the surface and the y-component of its velocity vector
is zero. On the other hand, the gradient at point Q is null both along the x-axis and
y-axis. An object will stay at this point, if you place it very gently. However, once it
is displaced either direction along the x-axis, it will start moving along the surface.
As point P gets farther away from point Q along the x-direction, the slope along
the surface gets steeper. As a beginner skier, you do not want to come to such a
point.
In accordance with the above argument, this dynamics can be considered in
a cross-sectional plane normal to the y-axis and containing point P of the half-
pipe (Fig. 3.3). Let s be the variable to represent the path of the object along the
circumference. We are interested in the change in the gravitational potential energy
along the surface, d =ds. Since the gravitational potential energy is .z/ D mgz
and the infinitesimal change in z and s can be related as d z D ds sin  with the
use of angle  (Fig. 3.3), the explicit form of d =ds in this case can be found as
follows

d @ d z
D
ds @z ds
D mg sin  (3.6)

According to Eq. (3.1), this represents the conservative force acting on the object
under the given constraint.

d
fs D  D mg sin  (3.7)
ds
Of course this force agrees with the one found through the analysis of the normal
force, vertical gravitational force, as indicated in Fig. 3.3.
In the above analysis, we have the additional constraint that the object moves
along the circumference to prove the consistency between the potential-based
approach and the net-force-based approach. Next, extend the idea by allowing the
ball to move any direction on the surface, and find out the direction the ball follows
at a given point. This task boils down to maximize the change d =ds in Eq. (3.5).
To generalize the argument, assume that the ball can take a path in any direction in
xyz-coordinates. Under this condition, the partial derivative @ is replaced with the
gradient operator, and the @ =@s with the directional derivative [7] as follows
42 3 Quick Review of Field Theories

Fig. 3.2 Ping-pong ball on a f


half-pipe

y
P

Fig. 3.3 Physical meaning of z


potential gradient
q N

f = mgz
ds
dz q
q ds sinq = −dz
fs = mg sinq
mg

df ∂f dz
− =− = (−mg) (−sinq)
ds ∂z ds

@ @ @
r D xO C yO C zO (3.8)
@x @y @z
d
D r  u (3.9)
ds
Here, u is a unit vector in the given direction that we are interested in analyzing the
potential change. To find the direction of the object motion at a given point on the
surface, all we do is to find u that maximizes the directional derivative r  u. It
is easily proven that in the case of Fig. 3.2, a unit vector u along the circumference
maximizes the directional derivative as follows. Since the potential is proportional
to z, the potential change is maximized when the path is selected to maximize d z=ds

dz @z dx @z dy
D C (3.10)
ds @x ds @y ds

In Eq. (3.10), dy=ds D 0 because the y-axis is parallel to the ridge. So, naturally,
when the path is perpendicular to the y-axis, the slope d z=ds is maximized.
3.2 Symmetry in Physics 43

3.2 Symmetry in Physics

The concept of symmetry is always associated with a certain operation and an


unchanged feature. On our daily life, line symmetry means that a given shape
appears exactly the same when it is flipped over the line of symmetry and point
symmetry means that a shape appears the same when it is rotated around the point
of symmetry by a certain angle. For example, an equilateral triangle is symmetric
about the three axis shown in Fig. 3.4. We say that line symmetry is associated
with an operation of reflection (inversion) and point symmetry is associated with
an operation of rotation.
In physics, symmetry also means that a physical feature remains unchanged
after a certain operation. It is said that the feature is invariant under the operation.
This concept is not merely applied to geometric feature such as a certain crystal
structure symmetric around some axes. It is extended to more abstract concepts
such as invariance associated with coordinate transformation. Consider how a vector
behaves under coordinate transformation in Fig. 3.5. When the x–y coordinate
system is transformed to the x 0 –y 0 coordinate system, clearly the components of
the vector change. However, the magnitude of the vector or the summation of the
square of all the components is unchanged. We say that the length of a vector is
invariant under a coordinate transformation. Here we consider a two-dimensional
case but the same argument holds for three dimensions.
The symmetry associated with the above coordinate transformation has a phys-
ical meaning. Consider that the vector represents the velocity of an object. It is
obvious that the speed (the magnitude of velocity) does not depend on how you
set up a coordinate system to analyze the motion of the object, while the direction
of the motion depends on the coordinate system. Consider the object moves to the
North and your x-axis (the first axis) is directed to the East and your y-axis (the
second axis) is to the North. Apparently, the velocity vector’s x-component (first
component) is zero. However, if another coordinate system whose x-axis (the first
axis) is directed to the North, this time the vector’s first component is the same as

Fig. 3.4 Symmetry of shapes


44 3 Quick Review of Field Theories

Fig. 3.5 Symmetry of vector y


under coordinate
transformation y'

x'

Fig. 3.6 Stretch of a spring y


and invariance of Hooke’s law
y'
dy
x'
dx'
dy'

dx x

the vector’s magnitude itself. The speed of the object is the same when measured
by either coordinate system. Later this concept leads to the invariance of norm of
a matrix.
The concept of symmetry in physics has more depth. It is further extended to
invariance in physical theories associated with transformation. In this context, it is
worth while considering a vector and its component before and after transformation.
Referring to Fig. 3.6, consider that we analyze a spring and mass system in two
coordinate systems .x; y/ and .x 0 ; y 0 /. Since the dynamics (the physical law) is
determined by the mass and spring neither of which depends on the choice of
a coordinate system, we know that the equation of motion and its solution is
the same in both coordinate systems. In other words, although the component of
displacement vector differs depending on the choice of the coordinate system, the
equation of motion remains the same because the force vector and the displacement
vector experience the same change in their components under the coordinate
transformation; while the displacement vector component changes from dx to dx 0 ,
the force vector component changes from fx to fx0 in the exactly the same fashion
as the displacement vector, and therefore, the relationship between the displacement
and the force remains the same. We can see that Hooke’s law is invariant under the
coordinate transformation.
The more interesting question is whether or not the differentials obey the same
coordinate transformation as the vector; i.e., if we relate the components of the
displacement vector before and after transformation, .x ; y / and .x0 ; y0 / with a
transformation matrix M , are their differentials .d x ; d y / and .d x0 ; d y0 / related
in the same fashion? If the answer is “Yes,” then we can definitely describe the
3.2 Symmetry in Physics 45

physics before and after the coordinate transformation using the same equation.1
Let’s examine this strictly.
!   
x0 M11 M12 x
D (3.11)
y0 M21 M22 y

The differential expressed in the before transformation coordinates can be given as


follows
0 1 0 1
    @x @x @x @x  
B @x dx C dy

d x D
d x
DB @y C B
C D B @x @y C dx
C
(3.12)
y d y @ @ y @ y A @ @ y @ y A dy
dx C dy
@x @y @x @y

Now the question is if the following equation is true.


!   
d x0 M11 M12 d x
D (3.13)
d y0 M21 M22 d y

And the answer is “Yes.”


Proof.
0 1
! @x0 @x0
d x0 B @x dx C @y C
dy C
DB
@ @y0 @y0 A
d y0
dx C dy
@x @y
0 1
@  @ 
B @x M 
11 x C M 
12 y dx C M 
11 x C M 
12 y dy C
DB
@ @ 
@y
@ 
C
A
M21 x C M22 y dx C M21 x C M22 y dy
@x @y
0 1
@x @y @x @y  
B M C M M C M
@y C
11 12 11 12
DB @x @x @y C dx (3.14)
@ @x @y @x @y A dy
M21 C M22 M21 C M22
@x @x @y @y

1
These differentials are necessary when we compute d x =dt etc.
46 3 Quick Review of Field Theories

Fig. 3.7 Vector under y


transformation

h'
h

0 1
  @x @x  
M11 M12 B C
B @x @y C dx
D
M21 M22 @ @y @y A dy
@x @y
  
M11 M12 d x
D
M21 M22 d y

t
u
This argument indicates that the differential operation is invariant under coordi-
nate transformation. Next, we consider the case where the vector itself experiences
a linear transformation. Laws of dynamics in general are expressed by differential
equations. If the differential operation is invariant under the linear transformation,
laws of dynamics are invariant as well; i.e., after the transformation the dynamics
can be expressed by the same differential equation as before the transformation,
regardless of the choice of the coordinate system. Let’s consider this problem
more quantitatively using linear transformation U indicated in Fig. 3.7. Let U be
transformation matrix that transforms vector  to 0 .

0 D U  (3.15)

Derivatives of 0 can be obtained as

@0 @ @
i
D i .U / D U i (3.16)
@x @x @x
Here x i denotes a coordinate variable, and the transformation U is assumed to be
coordinate independent (@U=@x i D 0). Under this condition, comparison of the
leftmost and rightmost sides of Eq. (3.16) indicate that the derivatives of the vector
are transformed in the same fashion as the vector. Since the transformation matrix
does not depend on the coordinate variable, this type of transformation is referred
to as a global transformation. When the transformation matrix depends on the
3.3 Global and Local Transformation 47

coordinate variables, the derivative of the transformation matrix becomes nonzero.


Consequently, the derivatives of the vector are not transformed in the same way as
the vector. This type of transformation is referred to as a local transformation and is
the main topic of the next section.

3.3 Global and Local Transformation

As we discussed at the end of the preceding section, when a linear transformation


is coordinate independent, the differential of a vector is transformed in the same
fashion as the vector. Consequently, the governing equation of the dynamics is
invariant under the transformation. When the transformation is coordinate depen-
dent, Eq. (3.16) must be rewritten as follows

@0 @U @
i
D iCU i (3.17)
@x @x @x
Before discussing mathematical details about local transformation and associated
dynamics, let’s briefly consider the physical meaning of global transformation.
Essentially, global transformation means that if a vector experiences a transfor-
mation at a certain point of a space, all other vectors in the space is transformed
in the same fashion. As will be discussed in more detail in the next section
as an example, if the wave function of a charged particle experiences a phase
transformation and the transformation is global, all other charged particles in this
universe are transformed in the same way; this is totally unrealistic. It is more
natural to allow different charged particles to experience different transformations.
Then according to Eq. (3.17), we encounter the situation where the differentials of a
vector are not transformed in the same way as the vector. This in turn indicates that
the governing equation (in the case of charged particles, the Shrödinger equation)
cannot be written in the same way after the transformation. We know that this
is not the case. Then how can we solve the problem? Is it possible to let the
transformation be local and assume that the Shrödinger equation is still valid? The
short answer is that quantum dynamics alone is not sufficient to solve the problem.
There must be another field dynamics that makes the phase transformation local and
still Shrödinger equation valid. This type of the field is called the compensation or
gauge field. In this particular case, the gauge field is nothing but the electromagnetic
field, and the associated dynamics is electrodynamics. When a charged particle
experiences a locally defined phase transformation, all the other charged particles
know this transformation and the associated change in this charged particle through
electrodynamics. In other words, as one charged particle changes its phase by being
displaced, all the other charged particles also change their own trajectory by means
of the electric and magnetic field force. More will be discussed about this particular
case in Sect. 3.6.
Now let’s go back to Eq. (3.17) and make more quantitative consideration about
the gauge field. The first step is to modify the differentiation operation so that the
48 3 Quick Review of Field Theories

derivative is transformed in the same way as the vector. Formulaically, we can make
the differential to take the same form as vector by replacing the usual derivative in
Eq. (3.16) by a covariant derivative Di defined as follows

@
Di D  i  @i  i (3.18)
@x i
The additional term i is referred to as the gauge term. Note that Di is associated
with the differentiation with respect to the coordinate variable x i , and that for the
differentiation with respect to each coordinate variable a gauge is defined.
With the use of the covariant derivative, the left-hand side of Eq. (3.17) can be
written as

Di0 0 D Di0 .U / (3.19)

Here the prime on D indicates that the differentiation operation is applied to the
vector after the transformation, and the right-hand side of Eq. (3.19) explicitly
expresses that Di0 0 is the differential of the vector after the transformation. Equating
this to the transformation of the derivative, U.Di /, we obtain the following
equation

Di0 .U / D U .Di / (3.20)

Equation (3.20) indicates that with this differential operation Di , the differential of
the transformed is the same as the transformed of the differential. In other words,
this differential operation can transform the differential in the same fashion as the
vector.
The next step is to find out the necessary condition on i in Eq. (3.18) so
that Eq. (3.20) holds. Substitute Eq. (3.18) to Eq. (3.20), we obtain the following
equation
 0 
@i U  C U @0i   i0 .U / D U .@i /  U .i / (3.21)

As Eq. (3.14) indicates, @0i D @i . Therefore, the second term on the left-hand side
cancels the first term on the right-hand side, and Eq. (3.21) leads to the following
condition

i0 .U / D U .i / C .@i U /  (3.22)

Here the left-hand side of Eq. (3.22) represents the situation “after” the transforma-
tion and the right-hand side represents that of “before.” Viewing i0 U , U i , and
@i U as operators on a vector, we can put Eq. (3.22) as follows

i0 U D U i C @i U (3.23)
3.4 Gauge Field 49

Further, by multiplying U 1 from the right, we obtain

@U 1
i0 D U i U 1 C U (3.24)
@x i
Equation (3.24) is the transformation that the gauge term i needs to obey as the
vector  is transformed by U . The combination of the transformation on the vector
and the gauge is called the gauge transformation

0 D U  (3.25)
@U 1
i0 D U i U 1 C U (3.26)
@x i

3.4 Gauge Field

The set of Eqs. (3.25) and (3.26) has the following mathematical and physical
meanings. When vector  experiences linear transformation as Eq. (3.25) where the
transformation matrix U depends on the coordinate variables, the transformation of
the vector’s derivatives cannot be expressed in the same form as the vector because
of the extra term @U =@x i  in Eq. (3.17). This means that under this condition the
usual differential operation is not invariant under the transformation. However, if we
replace the usual partial derivative with a covariant derivative by adding the gauge
term i [Eq. (3.18)] and if the gauge term is transformed by Eq. (3.26), we can
formulaically recover the same form of transformation for the vector’s derivatives
as the vector [Eq. (3.20)].
The content of the preceding paragraph can be viewed as follows. With the
use of the covariant derivative Di instead of the usual partial derivative @i , the
differential operation can be defined in the same way after the transformation as
before the transformation. In other words, the equation to describe the dynamics
that governs vector  can be expressed with the same differential equation after
the transformation. In this situation, we say that the law of dynamics is invariant
under the transformation. Since the transformation U is defined locally, we call this
symmetry (invariance) local symmetry (invariance). The quantity we introduce to
regain local symmetry, , is called the gauge. As we will see below, the gauge
represents a potential and the actual dynamics is described by the differential
equation that regains the local symmetry plus the potential. The transformations
(3.25) and (3.26) are called the gauge transformations.
The local symmetry gained in the above operation makes the dynamics of 
invariant in each individual local frame; it does not refer to inter-local frame
dynamics. To describe the dynamics in the global level, we need additional
50 3 Quick Review of Field Theories

information. This information is hidden inside the gauge term that we introduce
as the additional term in the differentiation operator. From the view point of the
global dynamics, this effect can be viewed as a certain constraint for each individual
local dynamics imposed via inter-local frame interaction. From this viewpoint, it is
natural to interpret the gauge term as a potential. The force field resulting from this
potential is called the gauge field. The force field describes the inter local-frame
dynamics. From the viewpoint that this field connects local frames, mathematicians
call it a connection field. Also, it can be interpreted that we need to introduce the
gauge field as a penalty to make the local dynamics formulaically invariant. From
this stand point this field is also called a compensation field.
As the gauge term operates on vector , the potential associated with the gauge
term is essentially a vector quantity. Calculation of the total differential for the i th
component of vector  will clarify this situation
     
@i @i @i
Di D  x i dx C  y i dy C  z i d z  di  Ai
@x @y @z
(3.27)
Here the suffix i denotes the i th component of vector , and Ai is the corresponding
component of the vector potential.
Once the vector potential is obtained, the gauge field can be obtained from a
quantity called field stress tensor F . Essentially, the field stress tensor represents
the interaction of the vector with the vector potential. Being originating from
the gauge, which is part of the differential operation as Eq. (3.18) indicates, the
field stress tensor can be interpreted as an additional effect that the vector 
experiences upon differentiation. To discuss this effect, first consider the meaning
of differentiation in Fig. 3.8. Differentiation with respect to the coordinate variable
x is basically measurement of the change in ’s components over an infinitesimal
change along the x-axis. If this operation is taken inside the local frame, the
differentiation is the usual partial differentiation @=@x. However, if the infinitesimal
change represented by dx involves multiple local frames, it becomes unclear

h (x + dx)
h (x)

y + dy
y

Fig. 3.8 Meaning of


differentiation x x + dx x
3.4 Gauge Field 51

Fig. 3.9 Two vector along a


geodesic

P’

whether the change in the vector component associated with the operation is due to
an actual change in the vector or to a change in the orientation of the local frame. As
a well-known example, consider a vector perpendicular to a geodesic [6] in Fig. 3.9.
The geodesic is the shortest line between two points that lies in a given surface.
Consider a vector is perpendicular to a gigantic sphere near the north pole. The
vector represents the reaction to the sphere’s gravitational force on a mass. Since
the gravitational force is directed to its center of the sphere, this vector is outward.
Another vector is found somewhere between the north pole and the equator. This
vector represents the reaction of another mass placed there to the net external force
acting on it. This mass is identical to the one placed in the other place near the
north pole. On this mass, however, an external force tangential to the surface of the
sphere is applied in addition to the gravitational force by the sphere. Consequently,
the vector is slightly tilted from the perpendicular direction to the sphere’s surface.
Consider that you are trying to find out the external forces on the above
two masses through analysis of the vector representing the reaction. For this
purpose, you are comparing the two vectors by placing the first vector’s tail on
the second vector’s tail without changing the orientations of the vectors. Obviously,
the comparison shows that the vectors’ directions are different and the difference
contains two sources. The first source is due to the additional external force on
the second mass. The other source is due to the fact that the two masses are
placed on different locations on the geodesic, and consequently, the directions of the
gravitational force are different between the two masses when the vectors are placed
on top of each other. Numerically, the changes in the vector due to the two sources
are indistinguishable. Physically, only the former is meaningful for the purpose of
finding the non-gravitational, external force. Thus, in order to analyze the external
forces, it becomes necessary to remove the difference in the vectors’ directions due
to the curvature of the surface (the shape of the path) from the total difference.
52 3 Quick Review of Field Theories

The same problem always occurs when we perform differentiation along a


geodesic. By moving on the geodesic, we find the direction of the vector at the
initial point and that at the final point are different from each other. But with a
simple differentiation, we cannot tell the difference is due to physics or the shape of
the path we take to perform the differentiation. In order to extract the pure physics
part of the difference, we need to remove the effect due to the shape of the path. If
we have local frames at both points of this analysis in such a way that the local z-axis
is normal to the surface, it is necessary for us to know the relative orientation of the
two local z-axis. This is the meaning of the covariant derivative; the removal of the
change associated with the change in the frame is represented by the gauge term.
The same argument holds for the differentiation with respect to each coordinate axis.
The next level of the argument is to figure out a net effect of the gauge on the
vector in the dynamics. To this end, we can differentiate along the x-axis first and
then y-axis, counterclockwise, to move from .x; y/ to .x C dx; y C dy/, and then
compare the result with the case when we conduct the differentiation clockwise.
Figure 3.10 illustrates this clockwise and counterclockwise differential operation
schematically. If the results of the two differentiations are different, the gauge field
has a net effect on the vector’s dynamics. The field stress tensor F is the indicator
of this effect. This concept will be revisited in the next chapter with the gauge field
that makes linear elastic theory locally symmetric as a specific field stress tensor (see
Sect. 4.3). As will be discussed later in this chapter (Sect. 3.5), the field stress tensor
itself is not invariant under the transformation, but the trace of its inner product is
invariant. This constitutes the Lagrangian of the gauge field


F D D D  dx dx (3.28)

The component of the field stress tensor F is the gauge field variable that
describes the dynamics resulting from the request of local invariance [1, 2] in
the original dynamics that governs vector . In the case of electrodynamics, F
is the electromagnetic field that makes quantum mechanics locally invariant, and
vector  describes the quantum mechanical state of the charged particle. With the
existence of the electromagnetic field, all the charged particles are connected. If one
particle moves, it changes the electromagnetic field felt by all the other particles.
Alternatively, we can think of the situation as all the charged particles are under the

dxm
h(xm + dxm, xn + dxn )

Fig. 3.10 The difference


between the clockwise and
counterclockwise dxn xn
differentiation of  with dxn
respect to point .x ; x / and
.x C dx ; x C dx /
xm
represents the interaction with
the gauge field h (xm, xn ) dxm
3.5 Lagrangian Formalism and Field Equation 53

influence of the electromagnetic potential, and movement of any particle changes


the potential.
The potential associated with the gauge field can be intuitively understood by
considering an experiment of a pendulum on a stormy day. You are measuring the
period of a pendulum in a swirling wind pattern. As the wind pushes the bob, the
period based on the measurement is off the theoretical value. The original dynamics
described by the equation of motion of a pendulum is hidden. You need to remove
the effect of the wind to find the original dynamics.

3.5 Lagrangian Formalism and Field Equation

Lagrangian is an abstract concept and somewhat difficult to digest. However, it is


extremely important to understand field theories. Before discussing the Lagrangian
of the gauge field, let’s quickly review the Lagrangian formalism and related topics
in general.
In one word, Lagrangian can be characterized as a quantity that describes the
dynamic state of an entity at the most fundamental level. It is a function of
position and velocity. It is naturally understood that the position and velocity are
the minimum set to describe the motion of a particle. In the Lagrangian formalism,
the momentum and force are defined as the derivative of the Lagrangian with respect
to the velocity and position

@L
pi D (3.29)
@xP i
@L
fi D i (3.30)
@x

Here, p i and f i are the i th components of the momentum and force vector,
respectively. Lagrangian is given as the difference of the kinetic energy and potential
energy of the particle

1 2
LD mv  (3.31)
2
Here m is the mass of the particle, v is the velocity of the particle, and is the
potential energy. Substitution of Eq. (3.31) into Eqs. (3.29) and (3.30) leads to

p i D mvi (3.32)
@
fi D (3.33)
@x i
Note that Eq. (3.33) is identical to Eq. (3.1).
54 3 Quick Review of Field Theories

Fig. 3.11 Action considered q


under particle moving from
q1 to q2
q2

Possible paths q(t)

q1

t1 t2 time

In description of dynamics, Lagrangian is based on the least action principle


(Hamilton’s principle), which states:
“When a dynamical system moves from one point to another within a time interval, the
actual path is the one that minimizes the integral of the difference between the kinetic and
potential energy.”

Here action S is defined as such time integral


Z
SD Ldt (3.34)

Consider Fig. 3.11 where q is a generalized coordinate and a particle located at


position q1 at time t1 is displaced to q2 at t2 . The question becomes what the actual
path that this particle takes in moving from q1 to q2 is. According to Hamilton’s
principle, the answer is the path that minimizes the variation of action; ıS D 0.
First consider a case where Lagrangian is a function of the generalized coordinate q
P L.q; q/
and its time derivative q; P
Z t2 Z t2  
@L @L
ıS D ıLdt D ıq C ı qP dt
t1 t1 @q @qP
Z t2    t2
@L d @L @L
D ıq  ıq dt C ıq (3.35)
t1 @q dt @qP @qP t1

Here the second term of the integrand is integrated in par in going through the last
equal sign. By definition, the last term is null (because at time t1 and t2 the particle
is at a single point or ıq.t1 / D ıq.t2 / D 0). Since ıS is zero for any ıq, we can
equate Eq. (3.35) to zero and divide the resultant equation by ıq. This leads to the
famous Euler Lagrangian equation of motion.
3.5 Lagrangian Formalism and Field Equation 55

@L d @L
 D0 (3.36)
@q dt @qP

It is intuitive to use Eqs. (3.29) and (3.30) in Eq. (3.36)

d dv
f  p Df m D0 (3.37)
dt dt
Here, it is assumed that the mass m is independent of time. This is the Newton’s
equation of motion for an object of mass m and acceleration a D d v=dt .
In the case that the Lagrangian is a function of the spatial derivative, we define
Lagrangian density L as
Z
LD L.q; q 0 ; q/dx
P (3.38)

Here for simplicity, we consider a one-dimensional case.

ıS D 0 (3.39)
@L @L @L
ıL D ıq C 0 ıq 0 C ı qP (3.40)
@q @q @qP
Z Z Z Z Z  
@L @L 0 @L
ıS D ıLdt D ı Ldxdt D ıq C 0 ıq C ı qP dxdt (3.41)
@q @q @qP

Here Z  x 2 Z x2 Z x2
x2 @L 0 @L @ @L @ @L
ıq dx D ıq  ıqdx D  ıqdx (3.42)
x1 @q 0 @q 0 x1 x1 @x @q 0
x1 @x @q 0
Z t2  t2 Z t2 Z t2
@L @L @ @L @ @L
P D
ı qdt ıq  ıqdt D  ıqdt
t1 @qP @qP t1 t1 @t @qP t1 @x @qP

(3.43)

Therefore,
Z t2 Z x2 Z t2 Z x2 Z t2 Z x2
@L @L 0 @L
ıS D ıqdxdt C ıq dxdt C P
ı qdxdt
t1 x1 @q t1 x1 @q 0 t1 x1 @qP
Z t2 Z x2 Z x2 Z t2 Z x2 Z t2
@L @ @L @ @L
D ıqdxdt  0
ıqdxdt  ıqdt dx
t1 x1 @q x1 t1 @x @q x1 t1 @t @qP
Z t2 Z x2  
@L @ @L @ @L
D   ıqdxdt D 0 (3.44)
t1 x1 @q @x @q 0 @t @qP
56 3 Quick Review of Field Theories

This leads to the following form of Euler–Lagrangian equation of motion

@L @ @L @ @L
 0
 D0 (3.45)
@q @x @q @t @qP

Equation (3.45) can be interpreted in the following way. Consider you move an
object under a potential by carrying it from one position to another. The object
changes its status through the first term @L @q
. In response, nature compensates this
change by adjusting other forms of energy so that the resultant change in the action is
minimum. If the object’s energy (Lagrangian) is dependent on the spatial derivative
with respect to x (like elastic energy of a volume proportional to the differential
displacement between front and rear surface of the volume along the x-axis),
 @ @L the
the second term @x @q 0
changes accordingly. Similarly, if the Lagrangian depends
on the velocity (the kinetic energy), the third term changes. In this fashion, we can
add any number of compensation terms as the Lagrangian has dependence on them,
@2 @L
e.g., @x 2 @q 00 .
Generally,

@L @ @L @ @L @2 @L @n @L @n @L
 0
 C 2 00   C.1/n n .n/ C  C.1/n n .n/ D 0
@q @x @q @t @qP @x @q @x @q @t @q
(3.46)
Here @q .n/ denotes either spatial or temporal derivative of nth order depending on
the context.
The above equation (3.46) can easily be extended to three dimensions by
replacing the spatial differentiation with the gradient operator as follows

@L @L @ @L
r   D0 (3.47)
@q @.rq/ @t @qP

In the case where the Lagrangian depends on higher-order derivatives such as the
second order time derivative, the above equation has the corresponding term as

@L @2 @L @L @ @L
C 2 r   D0 (3.48)
@q @t @qR @.rq/ @t @qP

Now extend the above idea to a gauge field using the transformation U defined by
Eq. (3.25). Our first task is to see if the product F F defined above by Eq. (3.28)
is invariant under the transformation U


F D D D  dx dx (3.28)
3.5 Lagrangian Formalism and Field Equation 57

Here the invariance of F F simply means that the value of this product remains
the same after the transformation as before; i.e.,
0
F F D F F 0 (3.49)

From Eq. (3.20), the covariant derivative after the transformation can be put in the
following form

D 0 D UDU 1 (3.50)

0
Using this expression, we can write F in the following form
h i  
0
F D D 0 ; D 0 dx dx D D 0 D 0  D 0 D 0 dx dx

D UD U 1 UD U 1  UD U 1 UD U 1 dx dx

D UD D U 1  UD D U 1 dx dx


D U D ; D U 1 dx dx
D UF U 1 dx dx (3.51)

0
Apparently, F ¤ F and the field stress tensor itself is not invariant under
transformation U . However, from the mathematical identity t r .AB/ D t r .BA/
(Commutativity under trace), we can easily find that the inner product of the field
stress tensor is invariant
h i 

0
t r F F 0 D t r UF U 1 UF U 1 D t r U F F U 1

 
D t r F F U 1 U D t r F F (3.52)

Equation (3.52) indicates that the inner product of the field stress tensor in the
form of Eq. (3.28) is invariant under transformation U , and can be part of the
Lagrangian. In Chap. 4, we apply this concept to the vector potential associated
with the gauge that makes the linear elastic theory locally symmetric, and construct
Lagrangian that leads to the field equations that the present theory is based on. In
fact, it will be shown that the term F F constitutes the Lagrangian of a free
particle. It will be discussed that for full description of the deformation field, the
interaction term of the Lagrangian is necessary (see Sect. 4.3)

L / F F .for free charges/ (3.53)


58 3 Quick Review of Field Theories

3.6 Electrodynamics as a Gauge Theory

The dynamics of a charged particle q under potential V is given by Shrödinger


equation [8, 9] in the following form
 
„2 2 @
.i r/ C qV ‰.r; t / D i „ ‰.r; t / (3.54)
2m @t

Here „ is the Plank constant, m is the mass of the charged particle whose electric
charge is q, and ‰ is the wave function of the charged particle. On the left-hand
side of Eq. (3.54), the first term represents the kinetic energy and the second term
the potential energy. In the kinetic energy term, i „r is the quantum mechanical
momentum pO

pO D i „r (3.55)

The solution to Shrödinger equation (3.54) is the wave function of the charged
particle. Consider that the charged particle is under phase transformation

‰ 0 D ‰e iq (3.56)

Here ‰ 0 denotes the wave function after the transformation. A wave function
represents the probability of a particle as a function of time and space. A phase
transformation on a particle alters the probability to find the particle; in other words,
it changes the location of the particle at a given time. If all charged particles are
subject to the same phase transformation, it means that all the charged particles
change their locations at the same time in the same fashion. This is absolutely
unrealistic. It is natural to assume that the phase transformation is coordinate
dependent, or  is a function of time and space.
Now consider if the wave function satisfies Shrödinger equation (3.54) after
a coordinate dependent phase transformation. Since Eq. (3.54) contains spatial
and temporal differentials, it is necessary that differentials of ‰ obey the same
transformation. Differentiate ‰ 0 allowing coordinate dependence of 

r 0 D r.‰e iq / D .r‰/e iq C ‰.i q/re iq (3.57)

This is apparently different from the transformation of the differential, .r‰/e iq . In
other words, transformation of the differential is not the same as the differential of
the transformed. This means that the Shrödinger equation in the form of Eq. (3.54)
is not invariant under the coordinate dependent phase transformation. It is necessary
to replace the usual derivatives with covariant derivatives. Use the following form
of covariant derivatives and examine if they suffice

D D r  i qA (3.58)
3.6 Electrodynamics as a Gauge Theory 59

@
D0 D C i qV (3.59)
@t

Since the use of covariant derivative (3.58) alters i r ! .i /.r  i qA/ D i r 
qA, the spatial differentiation after the transformation becomes
 
i r  qA 0 ‰ 0 D i r  qA 0 ‰e iq
D i.r‰/e iq C .i /‰.i q/.r/e iq  qA 0 ‰e iq

D i r C qr  qA 0 ‰e iq (3.60)

Equation (3.60) represents the spatial differentiation of the transformed wave


function, i.e., it prescribes the differential operation after the transformation. On
the other hand, transformation of the spatial differentiation is

.i r  qA/ ‰ 0 D .i r  qA/ ‰e iq (3.61)

From comparison of Eqs. (3.60) and (3.61), it follows that in order for the spatial
differential to be invariant under the transformation, A has to obey the following
transformation. Conversely, if A obeys the following transformation, the spatial
differential of the wave function is transformed in the same fashion as the wave
function

A 0 D A C r (3.62)

Similarly, the temporal differentiations is invariant before and after if the following
condition is satisfied

@
V0 DV  (3.63)
@t
Equations (3.62) and (3.63) indicate that if these conditions are held, the differential
operation can be defined the same before and after the phase transformation, and
thereby Shrödinger equation is invariant under the phase transformation.
Now what is the physical meaning of the additional terms i qA and i qV ? In
accordance with the general discussion of gauge terms and gauge transformation
presented above, these terms connect phase transformations occurring at different
coordinate points. When a charged particle undergoes phase transformation at a
certain point, other charged particles are connected with this phase transformation
via these terms. The situation can be interpreted as all charged particles are in a
potential, and when one charge changes its location via a phase transformation,
all other charges are informed of the movement of this charge via a change in
the potential. Since a change in a potential energy is field force, this change
in potential should be addressed in terms of a field force. Indeed, as will be
discussed later in Chap. 5 [see Eq. (5.90)], by replacing the usual derivative with
60 3 Quick Review of Field Theories

the covariant derivatives in the expression of momentum (3.55), we can derive


the famous Lorenz force. All charged particles in the universe know one another’s
movement via the Lorenz force, or electromagnetic force. From this viewpoint, we
say that electrodynamics makes quantum mechanics (Shrödinger equation) locally
symmetric. In this context, the electromagnetic field is a gauge field. Griffiths [5]
discusses the Lagrangian that leads to field equations of this dynamics known as
Maxwell’s equations of electrodynamics.

References

1. Chaichian, M., Nelipa, N.F.: Introduction to Gauge Field Theories. Springer, Berlin (1984)
2. Frampton, P.H.: Gauge invariance. In: Gauge Field Theories, Chap. 1. The Benjamin/Cummings
Publishing Company, Menlo Park (1987)
3. Aitchson, I.J.R., Hey, A.J.G.: Gauge Theories in Particle Physics. IOP Publishing, Bris-
tol/Philadelphia (1989)
4. Elliott, J.P., Dawber, P.G.: Symmetry in Physics, vol. 1. Oxford University Press, Oxford (1985)
5. Griffiths, D.J.: Introduction to Electrodynamics, 3rd edn. Prentice Hall, Upper Saddle River
(1999)
6. Kenyon, I.R.: General Relativity. Oxford University Press, Oxford/New York (1990)
7. Boas, M.L.: Mathematical Methods in the Physical Sciences, 3rd edn., p. 290. Wiley, New York
(2006)
8. Griffiths, D.J.: Introduction to Electrodynamics, 3rd edn. Prentice Hall, Upper Saddle River
(1999)
9. Schiff, L.I.: Quantum Mechanics, 3rd edn. McGraw-Hill Kogakusha, Tokyo (1968)
Chapter 4
Field Theory of Deformation and Fracture

The goal of this chapter is to apply the field theoretical idea discussed in the
preceding chapter to deformation of solid-state materials. Using the formalism
discussed in Chap. 3, we derive the field equations that govern the displacement
field of objects under deformation. Section 4.1 briefly describes existing filed-
theoretical approaches to deformation including field theories applied to dislocation.
Some of the ideas used in these theories are basis of the present field theory. In
Sect. 4.2 we consider the big picture of the theoretical basis that the present theory
is based on. The underlying concept that leads to the formalism that describes
nonlinear dynamics of plastic deformation as interrelationship among local frames
is explained qualitatively. Here the local frame is a logical unit each of which
describes its own linear elastic deformation. From this viewpoint, it is referred
to as the deformation structural element. In Sect. 4.3 the concept explained in
Sect. 4.2 is applied to coordinate-dependent transformation of linear elasticity. The
Lagrangian formalism is applied to the interrelationship among local frames. The
actual Lagrangian is found and field equations are derived. In Sect. 4.4 transverse
waves are found as solutions to the field equations. Various phenomena observed
in the wave characteristics are discussed. All these derivations will be made
formulaically. Their physical meanings will be considered in detail in Chap. 5.

4.1 Gauge Theories of Deformation

Field theoretical approaches to deformation have a long history. Since the middle
of the last century, researchers have known that moving dislocations behave as
if they are a relativistic particle. Their velocities approach that of sound waves
in relativistic fashion [1–5]. In addition, a number of authors discuss analogy of
dislocation dynamics with Maxwell’s electrodynamics [3, 6–11].

© Springer Science+Business Media New York 2015 61


S. Yoshida, Deformation and Fracture of Solid-State Materials,
DOI 10.1007/978-1-4939-2098-3__4
62 4 Field Theory of Deformation and Fracture

The similarities to general relativity and electrodynamics indicate the possibility


of formulating the dynamics of dislocations based on a gauge theory. In fact
several authors [12–16] derive gauge field equations that represent dislocation
dynamics. Egorushkin [14] derives wave equations of plasticity based on his gauge
theoretical approach [13] and explains irreversibility of plastic deformation based
on thermodynamics (entropy production). More recently, Lazar [17] derives a closed
system of field equations in a very elegant, quasi-Maxwellian form as equations of
motion for dislocations.
Let us take a brief look at the Egorushkin’s formulation as an example.
Egorushkin [13–15] formulates dislocations and disclinations in the following way.
In general, deformation can be expressed as linear transformation of line element
vector R as R 0 D AR C a, where R and R 0 denote the line element vectors before
and after the transformation, and A and a represent rotation and translation of the
line element vector, respectively. When defects exist, there is a jump in displacement
on the left and right side of the defect. This jump can be expressed with a pair of
linear transformations of this form, each applied to the left and right side of the jump

R 0l D Al R C al (4.1)
R 0r D Ar R C ar (4.2)

Allowing coordinate dependence in the transformation and requesting local sym-


metry, Egorushkin has derived field equations. Details about the derivation of the
field equations can be found in [13–15]. Without considering the effect of thermal
expansion as a source of the stress, the Egorushkin’s field equations can be put as
follows

c12
r J D  p (4.3)
S1

r J D  (4.4)
@t
1 @J 1
r ˛ D 2 C  (4.5)
c1 @t S1
r ˛ D 0 (4.6)

where J is the disclination current [13] (flow of disclinations), ˛ is the dislocation


density, c1 is the phase velocity, S1 is the coupling constant, p is the momentum
density (flux) of the surrounding elastic material, and  is the stress.
Consider the physical meanings of the above field equations. Equation (4.3) can
be viewed as an equation of continuity, which states that the source of disclination
current J is the momentum flux p. Dimensional analysis on Eq. (4.4) indicates that
r  J is in .1=m2 /=s D 1=.m2 s/ because the dislocation density ˛ is in 1=m2
on the right-hand side. Hence current J is in 1=.m s/. This dimension of J can be
interpreted as the momentum flux density p in .kg  m=s/=m2 normalized to the
unit mass; .kg  m=s/=m2 =kg D 1=.m s/. Thus, current J can be identified as a
4.1 Gauge Theories of Deformation 63

quantity representing the flow of disclinations generated by a flux of momentum of


the surrounding material. Equation (4.5) indicates that a stress of the surrounding
elastic field or a temporal change of the disclination flow generates dislocations
of density ˛. Here r  ˛ on the left-hand side indicates that the direction of the
dislocation density vector is perpendicular to the stress vector. This can be naively
understood as follows. Imagine a shear stress is applied to an array of atoms. If
this stress generates a positive dislocation on one side, it will generate a negative
dislocation on the other side. This situation that dislocations have opposite signs
on the opposite sides of a stress vector is cylindrically symmetric around the stress
vector; hence the resultant dislocation density vector is vortex-like as represented
by r  ˛. The fact that dislocations can also be generated by the temporal change
of J can be understood by remembering that J is generated by a momentum flux
p; the temporal change of momentum flux is a stress. Equation (4.4) states that the
time evolution of the dislocation density ˛ is due to the curl of J , not the divergence
of it as Egorushkin points out (p. 318 of [14]). Together with Eq. (4.5), this indicates
the wave-like nature of J and ˛. As Egorushkin points outs in [14], this wave-like
behavior of the defect fields is similar to the wave-nature of an electromagnetic
field where J corresponds to the electric field, ˛ to the magnetic field, and  to
the external (conduction) current. Equation (4.5) can be viewed as corresponding to
the electromagnetic case where both a temporal change of the electric field and
a conduction current can be a source of the magnetic field whereas the former
contributes to the conservative wave nature of the field and the latter to the energy-
dissipative nature of the field (see, for example, Chap. 7, 7.3.5 Maxwell’s Equations
in Matter, of Griffiths [18]). From this viewpoint, the temporal change of J is the
energy-conservative source and the stress  is the energy-dissipative source of the
dislocation density.
Thus, according to the field equations (4.3) and (4.5), the momentum and stress
of the surrounding material are the sources of the defects. This is consistent with the
field equations of dislocation derived by Lazar [17] in which the momentum and the
force stress are the sources of the dislocation excitations.
Panin et al. [19, 20] have applied a similar gauge theoretical approach to
plastic deformation using GL.3; R/ Lie transformation group. Here the linear
transformation represents the distortion of the medium under deformation where
each group element of GL.3; R/ corresponds to a degree of freedom in distortion
such as stretch along the x-axis and shear in the xy-plane. By allowing coordinate
dependence for the transformation and requesting local symmetry, they have derived
the following field equations

1 a0
r  S a  f abc .A a  S c / D  J (4.7)
l2
a
1 @S 1 a
.r  R a /  f abc ŒA b  R c
D C J (4.8)
cl2 @t l2
@R a
.r  S a /  f abc ŒA b  S c
D (4.9)
@t
r  R a  f abc .A a  R c / D 0 (4.10)
64 4 Field Theory of Deformation and Fracture

Here S a is the temporal change (the time derivative) of the distortion tensor
component a D 1; 2; 3; : : : ; 9, A b is the gradient of the distortion tensor component
reflecting the gauge field, R is the gradient of the bend-torsion tensor component,
f abc are the structural constants of the GL.3; R/ group, l is the dimensional
parameter, cl is the phase velocity of the gauge field, and J a0 and J a are defined
as follows
˛ Pˇ
J a0 D g ij ak
i k  j (4.11)

J a D g ij ak ˛ ˇ
i k .D  /j C˛ˇ (4.12)

where  is the density of the material, ak


i is the generator of the GL.3; R/ group,
 is the local frame, and D D @  ak i B a
is the covariant derivative.
After summation over the group indexes (contracting with respective to internal
coordinates), they rewrite the field equations in the following form [19–22]

r  V D g ij ˛i P j˛ (4.13)
!
r V D (4.14)
@t
1 @V
r ! D 2 C g ij ˛i .D j˛ / (4.15)
cl @t
r ! D 0 (4.16)

where V is the velocity and ! is the angle of rotation.1


The idea behind the present field theory of deformation and fracture originates
from the field equations (4.13)–(4.16). Through consideration on the physical
meanings of these equations and other equations and concept derived from these
equations, the author notices that the dynamics can be formulated directly from
the linear elastic theory by requesting its local symmetry without dealing with the
complexity of GL.3; R/. The rest of this chapter will be used to discuss the present
field theory.

4.2 The Big Picture of the Present Field Theory

The present theory is based on two postulates. The first postulate is that a solid
medium of any deformation status locally obeys the law of linear elasticity. In other
words, in any stage of deformation, the elastic, plastic and fracturing stage, we can

1
It seems that a negative sign is missing either on the right-hand side of Eq. (4.14) or on the entire
right-hand side of Eq. (4.15). Without one of the negative signs, the above equations do not yield
the wave equation that the authors of [20, 22] claim.
4.2 The Big Picture of the Present Field Theory 65

always find local regions where the law of linear elasticity is applicable [23–25]. We
call such a local region as the deformation structural element. The second postulate
is that as long as the medium remains as a continuum, it is possible to logically
connect the deformation structural elements by introducing a connection field and
that the introduction of the connection field leads to a new dynamics that describes
deformation of objects comprising multiple deformation structural elements.
Let’s discuss the first postulate first. The statement that a solid medium locally
obeys the law of linear elasticity means that within each deformation structural
element, we can define a local frame that acts as principal axes to describe the local
linear elasticity. The applicability of linear elasticity to local frames is justified by
the fact that we can always find a local region where the inter-atomic potential can
be approximated by a quadratic function regardless of the stage of deformation. This
indicates that if we go down to the atomistic level, the displacement of an atom from
its equilibrium position can be described by elastic force.
The applicability of linear elasticity to a local region becomes invalid when the
local region has defects such as dislocations. In that case, the force law does not
take the form of linear elasticity if applied to the entire region. Let’s consider this
situation with Fig. 4.1, which schematically illustrates the deformation of a local
region containing a defect under a tensile load. When a solid medium is linear-
elastically deformed, we can always find principal axes and express the deformation
with a set of normal deformation. As this figure indicates, the deformation of this
entire local region cannot be expressed with normal deformation only, because the
four blocks around the defect undergo rotations of mutually different directions
as the material is weaker against the tensile load along the central line where the
defect exists. Note that, however, each of the four blocks individually behaves linear-
elastic deformation. In other words, in this case the entire local region is viewed as
consisting of four deformation structural elements.
All these observations indicate that when dislocations or other defects exist, it
becomes necessary to introduce multiple deformation structural elements. Since
each deformation structural element has its own principal axis, their linear elasticity

Fig. 4.1 Local containing a


defect at the center
66 4 Field Theory of Deformation and Fracture

cannot be expressed with a common (global) coordinate system. In order to describe


the deformation of the entire local region with the global coordinates, it is necessary
to introduce a new dynamics. This dynamics must be consistent with the linear
elasticity of each deformation structural element that form the entire local region,
and at the same time, must “connect” the linear elasticity of each individual
deformation structural element. How can we formulate the new dynamics? This
is where the gauge (connection) field kicks in as the main player for the second
postulate. This dynamics is closely related to the differential operation, as will be
discussed in the following paragraphs.
As we observed in Chap. 2, the law of elasticity is based on the force proportional
to the local stretch, or differential displacement. In each individual deformation
structural element, this differentiation-based dynamics is valid with the use of
the local frame. From the global viewpoint, because each deformation structural
element has its own orientation as Fig. 4.1 illustrates, the usual differential operation
cannot be used for the reason described in Chap. 3 in conjunction with the geodesics
(see Fig. 3.9). However, as we discussed in Chap. 3, we can solve this problem by
replacing the usual derivatives with covariant derivatives, or adding an appropriate
gauge term to the usual derivative. This gauge and the associated potential is the
basis of the formulation of the new dynamics that connect local linear elasticity.
By applying the Lagrangian formalism, we can derive the prescription of the
dynamics, or the field equations that govern nonlinear deformation including plastic
deformation. We say that this gauge recovers the local symmetry of the linear
elasticity when the transformation matrix is coordinate dependent.
The situation is analogous to the local symmetry of quantum mechanics; when
we allow coordinate-dependent phase-transformations for charged particles, it
becomes necessary to introduce the electromagnetic field so that the Shrödinger
equations is valid at the global level [26]. From the viewpoint of mechanics, the
dynamics has an extra force associated with the interaction of charged particles with
the electromagnetic field. This force law is described by the field equations known
as Maxwell’s equations. We say that it is necessary to introduce electrodynamics to
regain local symmetry in quantum mechanics. From the perspective of electrody-
namics, the electric charges act as sources of the electric and magnetic fields. In the
field theoretical sense, it can be said that the electric charges are charge of symmetry
to make quantum dynamics locally symmetric. Naturally, the electric charges as
sources are common to the state of before the phase transformation and that of after.
This can be interpreted as the reason why the electric charge is a conserved quantity.
In the displacement field of solid-state materials under deformation, the same
argument as above holds. To make the law of elasticity locally symmetric, a gauge
field must be introduced. By remembering that the law of elasticity is essentially an
orientation preserving mapping [27] as we discussed in Chap. 2, we notice that it
is natural to characterize the gauge is related to rotation of deformation structural
element. Indeed as we discuss later in this chapter, the vector potential associated
with this gauge is closely related to the rotation of deformation structural elements.
It is also possible to identify charges of symmetry as quantities analogous to the
electric charge and current. In addition, it is possible to argue the dynamics in
4.3 Field Equations 67

the plastic regime based on charges referred to as the plastic deformation charges.
Interestingly, fracture can be considered as a phenomenon analogous to electric
breakdown [28] of dielectric media where the electric charges become the main
energy carrier as the discharge changes from the glow regime to the arc regime [29].
From this viewpoint, it is natural to interpret that the plastic deformation charge is
responsible for the energy dissipation in the plastic regime and plays an important
role in the transition from the later plastic regime to the fracturing regime.
As argued at the beginning of this section, the existence of dislocations and other
defects necessitates the introduction of multiple deformation structural elements
because the elastic force law does not apply to the entire medium globally. Indeed,
as we will see later in this and next sections, dislocations and defects act as source
terms of the field equations. In other words, dislocations and other defects create the
situation where the gauge field must be introduced to the global dynamics, and hence
they act as the sources for the field variables. From this viewpoint, a perfectly elastic
medium is defect-less so that the elasticity can be described by one local frame. In
the case of a perfectly elastic medium, the entire medium can be represented by one
deformation structural element and thereby the local frame is the global frame.

4.3 Field Equations

In Chap. 2, we discussed elastic deformation as a transformation represented by the


deformation gradient tensor. Recall that using the principal axis, it is possible to put
the formulation in the form where the shear component is null. View Eq. (2.9) as
transformation of dx to dx 0 ;

0 D U  (4.17)
U D I Cˇ (4.18)

Here  is the line element vector whose component is infinitesimal line element
dx i ; i D 1; 2; 3, I is the unit matrix, and ˇ is the displacement gradient tensor
defined in Eq. (2.9)
 
0i @i
dx D ıij C j dx j  Uij dx j (2.9)
@x
0 @ @x @x 1
x
B @x @y @z C
B @ @y C
@i B y @y C
ˇD j DB C (4.19)
@x B @x @y @z C
@ @z @z @z A
@x @y @z
68 4 Field Theory of Deformation and Fracture

In Chap. 3, we discussed that in order to regain local symmetry in the dynamics


represented by a transformation matrix when it is coordinate dependent, it is
necessary that the differential of the vector transform in the same fashion as the
vector, i.e., the differential of the transformed is the same as the transformed of the
differential

Di0 .U / D U .Di / (3.20)

This requires the usual differentiations be replaced with covariant derivatives (see
Sect. 3.3)

@
Di D  i  @i  i (3.18)
@x i
The additional term i called the gauge term compensates the effect that the
differentials are not transformed in the same way as the vector due to the coordinate
dependence of the transformation matrix. It was shown that for the total differential
of the i th component of vector , the effect of the gauge term is conveniently
expressed via the potential Ai as follows
     
@i @i @i
Di D  x i dx C  y i dy C  z i d z  di  Ai (3.27)
@x @y @z

Equation (3.27) prescribes the differentiation of line vector component i so that


the vector’s differentials are transformed in the same fashion as the vector when the
transformation matrix is coordinate-dependent.
Now extend the idea to make the elastic force law invariant under the above
transformation when ˇ (the displacement gradient tensor) is coordinate dependent
[24, 25]. In the theory of elasticity, the elastic force is proportional to the stretch
or the differential displacement d i . Therefore, for the theory to be invariant, the
differential displacement vector must be transformed in the same fashion as the
displacement vector itself. Otherwise, after the transformation the elastic force law
cannot be written in the same form as before the transformation. This means that
after the transformation, U.D/ must represent the differential of the transformed
displacement, i.e.

D 0 .U / D U.D/ (4.20)

as is the case of Eq. (3.20). Thus, essentially, we can repeat the same procedure as
Sects. 3.3 and 3.4, and derive the covariant derivative for displacement vector as
follows
     
@i @i @i
Di D  x i dx C  y i dy C  z i d z  d i  Ai
@x @y @z
(4.21)
4.3 Field Equations 69

Before continuing the discussion further, it is interesting to note that the condition
that the transformation matrix U is coordinate independent means that differen-
tiation of Eq. (4.18) with respect to a coordinate variable is zero. As the unit
matrix I is a constant, this means that ˇ is coordinate independent or constant.
Since the components of ˇ is the first order derivatives of displacement, the
coordinate independence of U in turn means that the coordinate dependence of the
displacement is as high as the first order, i.e., the deformation is linear.
In elastic deformation, the rotation matrix represents rigid body rotation of the
material, which does not involve length change. In Eq. (4.21), the actual change
in the length of displacement vector is all in d. Thus, A can be interpreted as
representing a displacement vector that rotates the deformation structural element so
that the differential displacement vector contains only the change due to physically
true deformation, not to the geometrical effect (see Fig. 3.9 and explanation
about the differentiation along the geodesic). Figure 4.2 illustrates this situation
schematically.
From the viewpoint of the potential that makes the dynamics of deformation
locally symmetric, A can be interpreted as follows. From the global point of
view, a given object experiencing plastic deformation consists of a number of
deformation structural elements. Within each deformation structural element, the
local deformation can be characterized as linear elastic deformation represented
by orientation preserving mapping [27]. Because of the geometric effect, each
individual deformation structural element has its own preferred orientation, the
orientation of the principal axes along which the spatial differentiation of the
displacement has normal components only (the shear strain components are all
null). Therefore, in order to perform the operation of differentiation to describe the
dynamics globally, it is necessary to remove the geometric effect. In other words, it
is necessary to align all deformation structural elements so that all vectors can be
differentiated with one operation. The displacement vectorA serves as the aligning
mechanism for this purpose, as Fig. 4.3 illustrates schematically. It is important to
note that this geometrical effect is due to the nonlinear nature of the deformation at
the global level. From the analysis of A, we can formulate the dynamics that makes
the deformation nonlinear from the global point of view. More will be discussed on
the physical meaning of A in the next chapter (Sect. 5.3).

dx

Gx

Fig. 4.2 Meaning of gauge


in deformation
70 4 Field Theory of Deformation and Fracture

B’

B’
A’
B
A’
A
Potential A

B’
B

A

B A

Fig. 4.3 Local elastic deformation is represented by each individual orientation-preserving map-
ping that describes differential displacement of the local region. In order to deal with all the local
deformations simultaneously, it is necessary to align orientation of all local deformation structural
elements so that each of them is still a orientation preserving mapping and differentiation can
be handled globally. The vector potential A represents the displacement that rotates all local
deformation structural elements so that they are aligned for the common differentiation with the
global coordinate variables

Splitting the distortion matrix into the symmetric part (strain matrix) and
asymmetric part (rotation matrix), as we discussed in Chap. 2, the differential
displacement can be written in the following form
0 1 0 10 1
dx 0 !z !y dx
D D ˇ @dy A  @ !z 0 !x A @dy A
dz !y !x 0 dz
0 1 (4.22)
!z dy C !y d z
D d   @ !z dx  !x d z A
!y dx C !x dy

Comparison of Eqs. (4.21) and (4.22) indicates that the three components of A can
be identified as

Ai D !j dx k  !k dx j ; i; j; k D x; y; z (4.23)

Equation (4.23) shows that as expected, A represents displacement associated with


rotation.
Now consider the interaction with the field via the vector potential. As we
discussed in Chap. 3, the interaction with the potential reflecting the gauge field can
4.3 Field Equations 71

Fig. 4.4 Gauge field strength B


for nonlinear deformation

Dxs
xn

xm
A

be evaluated through comparison of clockwise and counterclockwise differentiation


along a closed path. (See Fig. 3.10 and explanation about the field stress tensor.)
Apply this operation to this particular case and examine the interaction between
 and A. Figure 4.4 illustrates an infinitesimal differential displacement D s
(difference between the displacement of two neighboring points). Consider moving
from point A to point B along two paths under the influence of the potential; the first
path is to move along the x axis and then the x axis (clockwise), and the second
path is counterclockwise. Dropping the second-order differentials, the clockwise
case is

D .D s dx / dx D @ @ s dx dx  @ . s dx / dx C   s dx dx
(4.24)
Here from the definition,  s dx can be interpreted A , and so

1
D .D s dx / dx D @ @ s dx dx  @ A dx C A A (4.25)
s

The counterclockwise case can be expressed with the vector potential in the same
fashion. Thus the difference between the clockwise and counterclockwise case is
  1

D .D s dx / dx D D s dx dx D @ A dx  @ A dx C A ; A
s
(4.26)
In the infinitesimal limit, dx D dx D ds, and division of the above equation by
ds leads to

 1

D ; D s ds D @ A  @ A C A ; A  F (4.27)
s ds

The quantity F appearing on the rightmost hand side of Eq. (4.27) is the field
stress tensor that we discussed in Chap. 3. In the present case, F represents the
interaction of the deformation dynamics with the gauge field via the associated
potential. As we observed in Chap. 3, F itself is not invariant under the gauge
transformation, but its inner product F F is invariant [25]. Naively speaking,
72 4 Field Theory of Deformation and Fracture

the invariance of the inner product of F can be considered as an extension of the


invariance of the vector’s magnitude under coordinate transformation (see Fig. 3.5).
As we discussed in Chap. 3, the inner product of F is closely related to the
Lagrangian that represents the dynamic associated with the gauge field. Below, we
consider the Lagrangian associated with vector potential A. It is convenient to use
the relativistic form of metric tensor g for the calculation of the inner product

g D g
0 1
1 0 0 0
B0 1 0 0C (4.28)
DB@0 0
C
1 0A
0 0 0 1

Thus we define the coordinate and potential with the relativistic, four-vector
notations as2

x D x0; x1; x2; x3

x D x 0 ; x 1 ; x 2 ; x 3 (4.29)

where

x 0 D ct (4.30)

and
!
0 1 2 3  

A D ;A ;A ;A D A0 ; A1 ; A2 ; A3
c
  !
0 0
A D ; A1 ; A2 ; A3 D .A0 ; A1 ; A2 ; A3 / D ; A1 ; A2 ; A3 (4.31)
c c

 m
Here c is the phase velocity. 0 D A0 c m is the scalar potential associated
s
with the longitudinal elasticity, whose physical meaning will be discussed in detail
in the next chapter (Sect. 5.3).

2
So far we did not pay attention whether the index to specify the elements of a vector is a
subscript or superscript. In the four-vector notation, the index in the subscript or superscript form
is specifically defined as described here.
4.3 Field Equations 73

The derivatives can also be written with the relativistic, four-vector notation as
   
@ 1 @ @ @ @
@ D ; r D ;  ;  ; 
@x 0 c @t @x 1 @x 2 @x 3
   
@ 1 @ @ @ @
@ D ;r D ; ; ; (4.32)
@x 0 c @t @x 1 @x 2 @x 3

Thus the field stress tensor components with temporal and spatial terms are
!
@Ai @A0 1 @Ai 1 @ 0 1 @Ai @ 0 vi
F0i D @0 Ai  @i A0 D  D   D C D
@x 0 @x i c @t c @x i c @t @x i c

@Ai @A0 1 @Ai 1 @ 0 vi


F 0i D @0 Ai  @i A0 D C D C D (4.33)
@x 0 @x i c @t c @x i c

Similarly, the field stress tensor component with spatial terms only are

@Aj @Ai
Fij D @i Aj  @j Ai D  C D ! k
@x i @x j
@Aj @Ai
F ij D @i Aj  @j Ai D  i
C j D ! k (4.34)
@x @x
Thus,
0 1
0 v1 =c v2 =c v3 =c
Bv1 =c 0 ! 3 ! 2 C
F DB
@v =c !
C (4.35)
2 3
0 ! 1 A
v3 =c ! 2 ! 1 0

and
0 1
0 v1 =c v2 =c v3 =c
Bv1 =c 0 ! 3 ! 2 C
F DB
@v2 =c ! 3 0 ! 1 A
C (4.36)
v3 =c ! 2 ! 1 0

Consequently,

F F D F01 F 01 C F10 F 10 C F02 F 02 C F20 F 20 C    C F12 F 12 C F21 F 21 C   


 1 2 
.v / .v2 /2 .v3 /2

D 2 C C C 2 .! 1 /2 C .! 2 /2 C .! 3 /2
c c c
 2 
v
D 2 2  ! 2 (4.37)
c
74 4 Field Theory of Deformation and Fracture

Here vi and ! i are the rate of displacement (velocity) and rotation, respectively,
associated with the displacement vector A, or the additional dynamics the material
obeys because of the coordinate dependence of the transformation matrix represent-
ing the deformation gradient tensor.
As we discussed in the preceding chapter [Eq. (3.53)], we can construct
Lagrangian of a free particle using this F .3 Lagrangian has the form of
kinetic energy minus potential energy. From Eq. (4.37), we know that F F
is dimensionless. Noting that the second term of Eq. (4.37) has the form of the
square of rotation, we find that multiplication of the shear
 modulus of the medium
to the entire equation makes this quantity to be in N m2 D J m3 or have a
dimension of Lagrangian density. Thus, we can identify that the Lagrangian density
as follows
 2 
G 1 v
L D  F F D G  ! 2
(4.38)
4 2 c2

Here G is the shear modulus in .N m2 /. Noting that the phase velocity of a shear
wave is related to the shear modulus and the density  of the medium as
s
G
cD (4.39)


Equation (4.38) can be further rewritten as follows

v2 G! 2
LD  (4.40)
2 2
As expected, Eq. (4.40) has the form of “kinetic energy density” minus “potential
energy density.”
This is the Lagrangian of the gauge field itself, corresponding to the Lagrangian
consisting of electric and magnetic energy of the electromagnetic field as the gauge
field to make quantum dynamics locally symmetric. To discuss the dynamics due to
this gauge field, it is necessary to consider the Lagrangian of interaction. Taking the
first order term only, we can include this interaction term in the expression of the
Lagrangian density as follows

G v2 G! 2 G
LD F F C Gj A D  C j 0 A0 C Gj i Ai (4.41)
4 2 2 c
Here A is the four vector defined by Eq. (4.31). j is the four vector representing the
charge of symmetry, or the quantity that prescribes the way the gauge field interacts

3
Here the word particle means a local material represented by a global coordinate point.
4.3 Field Equations 75

with the solid-state material obeying the local dynamics of linear elasticity. From
the viewpoint of global dynamics, j can be viewed as the source term
 
j0 1 2 3
j D ;j ;j ;j (4.42)
c

Remember that A has the dimension of length. Considering that L, G and c are in
(J/m3 ), (N/m2 ) and (m/s), we note that j 0 and j i are in 1/s and 1/m, respectively.
Here the suffix 0 and i denote the temporal and spatial components. The physical
meaning of the four vector j will be discussed in detail in the next chapter
(Sect. 5.4).
With Lagrangian density (4.41) substituted to the Euler–Lagrangian equation of
motion in the form of Eq. (3.45) discussed in Chap. 3, we can derive field equations.
As this process is somewhat complicated mathematically, we take some time and
consider it step by step. The Euler–Lagrangian equation of motion associated with
A can be given as follows

@L @L
@  D0 (4.43)
@.@ A / @A

Here the first term of Eq. (4.43) represents the variation of the action associated with
F F term of the Lagrangian density (4.38) and the second term represents the
Lagrangian density associated with the interaction term. Note that the interaction
term of (4.38) does not have explicit dependence on @ A . First consider the partial
derivatives of the first term of Eq. (4.43)

F F D F .@ A  @ A / D F @ A  F @ A D F @ A  .F /@ A
D F @ A C F @ A D F @ A C F @ A D 2F @ A
D 2.@ A  @ A /@ A (4.44)

Here F D F is used at the end of the first line, and and are swapped in
the second line.
Differentiate the rightmost-hand side of Eq. (4.44) with respect to @ A . This is
a differentiating of the product of 2.@ A  @ A / and @ A both of which depends
on @ A . First consider differentiating (@ A  @ A ) part. Obviously,

@
Œ2.@ A  @ A /
@ A D 2@ A (4.45)
@ A

Since the indices and are cycled, at some point becomes and becomes .
Therefore, the differentiation yields the following term as well

@
Œ2.@ A  @ A /
@ A D 2@ A (4.46)
@ A
76 4 Field Theory of Deformation and Fracture

Thus, the differentiation of .@ A  @ A / part results in

2@ A  2@ A D 2 .@ A  @ A / (4.47)

The other term @ A also contains @ A as the indices cycle. To visualize this,
rewrite the rightmost-hand side of Eq. (4.44) as follows

2.@ A  @ A /@ A D 2.@ A  @ A /.@˛ g ˛ /.Aˇ g ˇ /

D 2.@ g ˛ A g ˇ  @ g ˇ A g ˛ /@˛ Aˇ

D 2.@˛ Aˇ  @ˇ A˛ /@˛ Aˇ D 2.@ A  @ A /@ A (4.48)

Here in the last step ˛ and ˇ are replaced with and , respectively. Thus, through
the differentiation with respect to @ A , this part also yields

2 .@ A  @ A / (4.49)

Consequently, the first term of Eq. (4.43) becomes


"
#
G @ F F G
@ D @ Œ2.@ A  @ A / C 2.@ A  @ A /

4 @.@ A / 4
D G@ .@ A  @ A / D G@ F (4.50)

The second term of Eq. (4.43) is rather straightforward. From Eqs. (4.41) and
(4.42),
@L @ G 0 G
D j A0 D j 0 (4.51)
@A0 @A0 c c
@L @  i
D Gj Ai D Gj i (4.52)
@Ai @Ai
So, Eq. (4.43) becomes
1 0
 @ F 0 D j (4.53)
c
@ F i D j i (4.54)

Here Eq. (4.53) is for D 0, and (4.54) is for D 1, 2, or 3. Consider the explicit
form of the left-hand side of Eqs. (4.53) and (4.54) using Eqs. (4.32) and (4.36).
4.3 Field Equations 77

D 0 case (time component)

1 1
@ .F 0 / D .@1 F 01 C @2 F 02 C @3 F 03 / D  .@1 v1 C @2 v2 C @3 v3 / D  r  v
c c
(4.55)
D i; i D 1; 2; 3 case (space component, i D 1 as an example)
  v 
1
@ .F / D  @0 F 10 C @2 F 12 C @3 F 13 D  @0 C @2 .! 3 / C @3 ! 2
c
 
1 @v
D 2 C .r  !/1 (4.56)
c @t 1

Finally, from Eqs. (4.55) and (4.56), we find the field equations (4.53 ) and (4.54)
in the following form. Note that from Eqs. (4.53) and (4.54), the indices have been
lowered; j D .j 0 =c; j 1 ; j 2 ; j 3 /; j D .j 0 =c; j 1 ; j 2 ; j 3 /.

r  v D j0 (4.57)
 
1 @v
C .r  !/ D j (4.58)
c2 @t

From Eq. (4.34), we find that

@! k @ @Aj @ @Ai
D i
 (4.59)
@t @t @x @t @x j
Changing the order of the temporal and spatial partial differentiation on the right-
hand side of Eq. (4.59) and using Eq. (4.33), we find that

@! k @ @Aj @ @Ai @vj @vi


D i  j D i  j
@t @x @t @x @t @x @x

Thus, we find the following relationship between v and !

@!
r vD (4.60)
@t
Further, noting that ! is curl of A in Eq. (4.34) and the mathematical identity
r  .r  A/ D 0 (here A denotes an arbitrary vector), we can put the field
equations (4.58) in the same form as Maxwell equations are usually expressed with

r  v D j0 (4.61)
@!
r v D (4.62)
@t
78 4 Field Theory of Deformation and Fracture

1 @v
r ! D  j (4.63)
c 2 @t
r ! D 0 (4.64)

Here j0 and j are the source terms and the physical meaning of each individual
equation of Eqs. (4.61)–(4.64) will be discussed in detail in the next chapter
(Sect. 5.4). For comparison, Maxwell equations are shown below

1
r E D e (4.65)
0
@B
r E D  (4.66)
@t
@E
r  B D  0 0 C 0 je (4.67)
@t
r B D 0 (4.68)

Here E is the electric field, 0 is the electric permittivity in free space, e is the
electric charge density, B is the magnetic field, 0 is the magnetic permeability in
free space, and j e is the electric current density.

4.4 Wave Solutions

The field equations (4.61)–(4.64) are analogous to Maxwell equations, as seen


above. Indeed, as the case of electrodynamics, they yield wave equations. Before
getting into mathematical details, it is worth considering the mechanism that an
electromagnetic field generates the wave characteristics. Consider in Fig. 4.5 that a
magnetic field at a certain point in the space increases over time for some reason.
According to Faraday’s law, this induces an electric field in the surrounding area via
r  E as illustrated by the solid loopy line in Fig. 4.5. This induced electric field, in
turn, induces magnetic fields in accordance with Eq. (4.67). In Fig. 4.5, the induced
magnetic field is illustrated with dashed lines. At the point in the space where the
initial magnetic field increases over time, the induced magnetic field is downward, or
in the direction of decreasing the initial change of the magnetic field. Similarly, if the
initial magnetic field decreases over time, the same mechanism induces a secondary
magnetic field so that the decrease in the initial magnetic field is diminished. If we
start with an increasing/decreasing initial electric field, the same mechanism induces
secondary electric field to decrease/increase the initial electric field. In other words,
the change in the initial field is always opposed by this mechanism, as if a spring
opposes motion of a mass attached to it whether the mass is pulled or pushed from
the equilibrium position by an external agent.
4.4 Wave Solutions 79

Fig. 4.5 Temporal increase ∂Bind


in magnetic field is ∂B < 0
> 0 ∂t
suppressed by Faraday’s ∂t
mechanism

∇×E <0
Bind

Fig. 4.6 Faraday’s mechanism generates secondary magnetic fields at all points. The initial change
in the magnetic field induces the electric field (solid lines) at all radii. This electric field in turn
induces the secondary magnetic field at all points in the space (dashed lines)

This “spring-like” effect is the electromagnetic version of nature’s opposing


mechanism known as Lenz’s law in the broad sense; “Nature reacts to a change by
inducing an effect that opposes the initial change.” Note that this nature’s negative
feedback mechanism comes from the negative sign appearing on the right-hand side
of Eq. (4.66). Without this negative sign, the direction of the induced electric field
would be opposite to the one seen in Fig. 4.5 leading to a catastrophic result; once a
magnetic field increases over time, it keeps increasing forever.
The above-mentioned Lenz’s law not only opposes the change in the initial
field, but it propagates the effect through the surrounding space. This is because
the mechanism involves curl of the other field in the process to compensate the
temporal change of the initial field. As illustrated in Fig. 4.6 schematically, when
the initial magnetic field changes over time, the electric field is induced in a plane
perpendicular to the magnetic field in a concentric-circular fashion at all radii. This
generates secondary magnetic fields at all points in the space. As we observe below,
this causes oscillations at all points with a certain time delay, generating a sinusoidal
transverse wave.
The wave characteristics are easily seen by decoupling E and B from each other.
Apply the curl of Eq. (4.66) and differentiate Eq. (4.67) with respect to time

@B @
r  r  E D r  D r B (4.69)
@t @t
@ @2 E @je
r  B D  0 0 2 C 0 (4.70)
@t @t @t
80 4 Field Theory of Deformation and Fracture

Equating the left-hand side of Eq. (4.69) and the right-hand side of Eq. (4.70), using
the mathematical identity r  r  E D r.r  E /  r 2 E , and assuming no charge
or r  E D 0 and hence je D 0, we obtain the following wave equation for E

@2 E
r 2 E D  0 0 (4.71)
@t 2
Similarly, by differentiating Eq. (4.66) with respective to time, applying the curl of
Eq. (4.67) and using the same assumption and mathematical identity, we obtain the
wave equation of the same type as Eq. (4.71) for B

@2 B
r 2 B D  0 0 (4.72)
@t 2
Wave equations in the form of Eqs. (4.71) and (4.72) can be put

1 @2
r2 D (4.73)
cph @t 2

where is the vector quantity travels as a wave and cph is the phase velocity. Thus,
the quantity 0 0 can be identified as the reciprocal of the square of the phase
velocity of the electronic magnetic wave ce0 known as the speed of light in free
space

1
ce0 D p (4.74)
 0 0

As known well, in a linear medium characterized by electric permittivity  and


magnetic permeability the phase velocity becomes

1
ce D p (4.75)


Thus with the form of Eq. (4.75), we can write the wave equations for the two field
variables E and B as

1 @2 E
r 2E D (4.76)
ce2 @t 2
1 @2 B
r 2B D (4.77)
ce2 @t 2

Equations (4.76) and (4.77) have the form of “secondary temporal derivative is a
constant times secondary spatial derivative.” We know that in general a sinusoidal
function of the following form satisfies this relationship. Thus we find that the
4.4 Wave Solutions 81

general solution to the wave equations (4.76) and (4.77) can be written in the
following form

E D E0 sin.ce t C r/ (4.78)
B D B0 sin.ce t C r/ (4.79)

The general solutions (4.78) and (4.79) represent a monochromatic, longitudinal,


or transverse wave. The fact that an electromagnetic wave is always a transverse
wave, as known well, does not come from the wave equations (4.76) or (4.77). It
comes from additional constraints on E and B imposed by Maxwell equations.4
Notice that Eqs. (4.76) and (4.77) have been derived from Eqs. (4.66) and (4.67),
only two of the Maxwell equations. It is Eq. (4.65) under the condition of no charge
or its right-hand side is zero that forces the electromagnetic wave to be always
a transverse wave. More generally speaking, whereas every solution to the field
equations must obey the wave equation, the converse is not true. As we will see
below, this has a significant meaning in the deformation counterpart.
Whether a given wave is a longitudinal or transverse wave is usually determined
by whether the direction of the oscillation is parallel or perpendicular to the direction
of the wave propagation. From the viewpoint that a wave carries energy, it is also
possible to argue that if a wave is longitudinal/transverse, it carries the energy
parallel/perpendicular to the oscillating field variable. This leads to the well-known
fact that an electromagnetic wave carries the electromagnetic energy perpendicular
to both the electric and magnetic field. This energy flow is known as Poynting vector.
An important consequence is that the electromagnetic wave characteristic does
not possess a mechanism to carry energy in the direction of the electric or magnetic
field whereas the electric force acting on an electric charge is parallel to the
electric field. In fact, the energy associated with electric force does not contribute
to propagation of the electromagnetic energy through space. Rather, it dissipates
the electromagnetic energy through the work of displacing the charge by exerting
a force on it. As much as the kinetic energy that the charge gains in the process,
the electromagnetic field loses the energy stored in the space. Normally, the kinetic
energy gained by the charge is converted to heat as the medium exert resisting force
to the charge so that it does not get accelerated. The resultant flow of charges at
the constant velocity is called the conduction current whereas the constant velocity
is called the drift velocity of the charges. This phenomenon is known as Ohmic
loss. Note that this energy dissipation is determined by the conductivity of the
medium that the electromagnetic wave travels through; i.e., it is determined by the
medium’s property not the field itself. In the case of the magnetic force, since it
is perpendicular to the motion of the charge, it does not do any work. From the
viewpoint of potential, we can argue that by moving in the direction of the field,
the charge loses the electric potential, or the conduction current flows to lower

4
See p. 378 of [18].
82 4 Field Theory of Deformation and Fracture

the electric potential in the same sense as water flows to a lower place losing the
gravitational potential.
The above situation can be argued as follows as well: when an electric energy
is carried transversely by an electromagnetic wave, the associated electromagnetic
wave is potentially non-energy-dissipative, i.e., the energy stays in the form of elec-
tromagnetic field energy without being transferred to other form of energy such as
thermal energy. However, once the mechanism changes so that the electromagnetic
field carries the energy longitudinally via electric force, the electromagnetic energy
gets dissipated via Ohmic loss, i.e., it is not recoverable as electromagnetic energy.
An interesting and good example of such switching energy carrying mechanism
from transverse to longitudinal is the transition from a glow discharge to an arc
discharge. Often, this transition leads to electric breakdown of the medium [28],
such as lightning. There is an interesting theoretical similarity between electric
breakdown of gas media and fracture of solid-state media [29], as will be discussed
in detail in the next chapter.
The above argument is a subtle issue of the charge conservation law that states
that the electric charges are never generated or lost. What if the charge dissipating
the energy disappears? Where would the energy go? The charge conservation is
essential for the energy conservation law. The charge in fact moves in the form
of current conserving the mass and energy. From the gauge theoretical viewpoint,
the subject is closely related to the Noether’s theorem which states “if Lagrangian
is invariant under a continuous transformation, a current of symmetry and a
corresponding conservation law necessarily exist.” In this case, the electric charge
is the charge of symmetry that makes the phase transformation of charged particles
locally symmetric.5
The charge conservation can be argued from Maxwell equations as well. Apply-
ing the divergence to Eq. (4.67) and using Eq. (4.65) and mathematical identity
r  r  ! D 0, we obtain the so-called equation of continuity

@e
D r  je (4.80)
@t
Equation (4.80) explicitly states that the change in the number of charges in a
unit volume is exactly the same as the net current flowing through a surface that
encloses the unit volume. If the total number of the charge in the unit volume
decreases/increases over time, the net flow is outward/inward.
The similarity between deformation dynamics and electrodynamics is not limited
to the form of the field equation and the resultant wave characteristics of the field.
A number of phenomena observed experimentally in deformation and fracture of
solid-state materials can be explained based on the analogy to electrodynamics
[30, 31]. While field theoretical interpretations of various deformation phenomena

5
See for example Chap. 4 of [26].
4.4 Wave Solutions 83

Fig. 4.7 Lenz’s mechanism x(x,y)


in deformation

x(x,y)

are subjects of the next chapter, it will be useful to discuss the analogy quickly
referring to the above paragraphs about electrodynamics.
The deformation-dynamics version of the nature’s negative feedback mechanism
represented by Lenz’s law can be explained rather straightforward. As known well,
solid-state materials generally possesses elasticity or the ability to exert spring force.
Here the spring force is defined as the force whose direction is opposite to the
displacement from the equilibrium position and magnitude is proportional to the
displacement. Microscopically, this force comes from the inter-atomic force that
can be approximated by force of this kind. When all atoms in a given material are
considered to be points masses, in other words if we ignore structures in the material,
this force is simply associated with translational dynamics. However, as soon as we
consider a structure of a finite volume in the material, dynamics of the rotational
degrees of freedom kicks in. As Fig. 4.7 illustrates the situation schematically, a
deformation structural element is rotated by surrounding deformation structural
elements that it shares a border. For instance, when the left deformation structural
element shown in Fig. 4.7 is rotated counterclockwise by external force (torque),
the central deformation structural element rotates clockwise due to the force exerted
by the former along the boundary. Because of the elasticity, this force is elastic
force making the central deformation structural element go back to the neutral
(equilibrium) rotational position. In other words, the material tries to go back to
the neutral situation through this rotational and translational interaction. Since the
displacement at the boundary is caused by shear strain, the force is shear force
associated with the shear modulus. By denoting the translational and rotational
positions with  and !, respectively, this mechanism can be expressed in a fashion
analogous to the magnetic induction. From the perspective of translational dynamics
at the boundary, this can be viewed as follows. Due to the differential rotation of the
neighboring deformation structural elements, r !, the local region at the boundary
exerts restoring (elastic) force. According to Newton’s second law, this force is
proportional to the acceleration, or the secondary temporal derivative of . Noting
that velocity v is the first temporal derivative of , we can express this dynamics as

@v
 Gr  ! D  (4.81)
@t
84 4 Field Theory of Deformation and Fracture

Fig. 4.8 Shear force due to


differential rotational
displacement

Fig. 4.9 Transverse wave


characteristics of !

where G is the shear modulus and  is the density. Figure 4.8 illustrates schemati-
cally the shear force due to differential rotation.
Importantly, as soon as the central deformation structural element rotates clock-
wise, the right deformation structural element starts to rotate counterclockwise.
Thus the rotational motion propagates through a chain of deformation structural
elements, accompanying resistive translational force at the boundaries of each
neighboring pair. The torque exerted on elements alternates in going successive
deformation structural elements whereas the force at the boundaries also alternates.
At certain point of time, the resistive force at the boundary equals the force exerted
by the neighboring element. Due to the inertia at that moment, the deformation
structural elements keep rotating in the same direction, generating resistive force
in the opposite direction until the deformation structural elements reach their
rotation limits. Being spring force due to the material’s elasticity, the force is
always centripetal generating negative acceleration. In other words, the dynamic
is transmitted through the material as a wave. If we set up a rotation vector in each
deformation structural element as

!Dr  (4.82)

in the direction normal to the plane of the element, the dynamics can be viewed as
a transverse wave of !. Figure 4.9 illustrates the situation schematically.
4.4 Wave Solutions 85

For more quantitative argument, consider eliminating the translational displace-


ments from Eqs. (4.81) and (4.82). Noting that v is time derivative of , we can
rewrite Eq. (4.81) into the following form

@2 
 Gr  ! D  (4.83)
@t 2

Taking curl and using the mathematical identity r  r  ! D rr  !  r 2 !


along with r  ! D 0, [Eq. (4.64)], we can rewrite the left-hand side of Eq. (4.83) as
follows

 Gr  r  ! D rr  ! C Gr 2 ! D Gr 2 ! (4.84)

Similarly, taking curl of both-hand sides and using r   D ! on the right-hand


side, we can rewrite Eq. (4.83) as follows

@2 !
 Gr  r  ! D  (4.85)
@t 2
Equating the right-hand sides of Eqs. (4.84) and (4.85), we obtain the following
equation

@2 !
Gr 2 ! D  (4.86)
@t 2
p
Equation (4.86) is an equation of wave traveling at the phase velocity .G=/. In
continuum mechanics, the solution to this equation is known as the rotational wave
or deformation wave. Its relation to the wave dynamics of plastic deformation will
be discussed in the next chapter.
By eliminating !, it is also possible to derive a wave equation that governs v.
The solution to the resultant wave equation yields the well-known shear wave.
However, the process is a bit more complicated than the case of rotational wave, and
will be discussed in the next chapter after exploring the physical meanings of the
underlying, dynamics (Sect. 5.2). Note that both !-wave and v-wave are transverse
waves. The former corresponds to the magnetic wave and the latter the electric
wave in electrodynamics. The wave generation mechanism through the rotational
and translational interaction corresponds to the magnetic induction. The !-wave
and v-wave resulting from the above mechanism are transverse wave because the
resistive force at the boundary is based on shear force, not normal force. This may
sound a matter of course but in fact it has subtleness in differentiating elastic and
plastic deformation both of which the field equations (4.61)–(4.64) yield as solutions
under different physical conditions. More details about this will be discussed in the
next chapter.
The similarity to electrodynamics can also be discussed to explain energy
dissipation. In fact, the discussion based on the analogy to electrodynamics in this
86 4 Field Theory of Deformation and Fracture

context is crucially important to define plasticity as an energy dissipating process.


This is a topic of the next chapter but the essence should be discussed here in
the context of wave decay. In short, the energy dissipation in a material under
deformation can be argued via a flux of momentum density. In order to make
arguments along this line, it is convenient to define the deformation charge in
analogy to the electric charge. The right-hand side of Eq. (4.61), j0 corresponds
to the electric charge density. From the field theoretical point of view, both
charges can be considered as a charge of symmetry; so this correspondence is not
surprising. In continuum mechanics, divergence of displacement r is known as
volume expansion [32]. From this viewpoint, the left-hand side of Eq. (4.61) can
be interpreted as the rate of volume expansion of a unit volume. It is possible
to view the change in the volume expansion over a unit time as the expanded or
compressed volume flows out from the coordinate point associated with the unit
volume. From this viewpoint, the charge density j0 can be viewed as the flux of
volume expansion. Applying the divergence to Eq. (4.63) and using Eq. (4.61), as
we did for the electromagnetic case, we obtain an equation of continuity governing
the deformation charge and its current j

1 @j0
Dr j (4.87)
c 2 @t
The equation of continuity (4.87) can be discussed in conjunction with the momen-
tum of a unit volume. In general, the phase velocity of a mechanical wave has
the form of square root of “elastic modulus/density” where the elastic modulus
represents the elasticity that causes the transfer of the oscillatory motion from
one point to the next of the medium and the density represents the inertia of the
oscillatory motion. It is important to note that the phase velocity, not the frequency
or wavelength, is intrinsic to a given medium. In other words, each medium has
characteristic difficulty in oscillation (the inertia or the density) and characteristic
strength of restoring force (such as the Young’s modulus) per unit volume. Denoting
the elastic modulus with  and the density , we can rewrite Eq. (4.87) as follows

@j0 @
 D .j0 / D r  .j / (4.88)
@t @t
With the help of Eq. (4.61), the quantity j0 can be put as follows:

 j0 D .r  v/ (4.89)

Dimensional analysis on the right-hand side of Eq. (4.89) indicates that the quantity
j0 is in (kg/m3  1=s/ D .kg=s  1=m3 /, and that it represents the rate of mass flow
coming out of the unit volume. Since each particle has mass and velocity, this causes
momentum flow-out from the unit volume. Substitution of Eq. (4.89) into Eq. (4.88)
leads to

@
 .r  v/ D r  .j / (4.90)
@t
4.4 Wave Solutions 87

The left-hand side of Eq. (4.90) represents the temporal change in the momentum
flow-out from the volume, and the equation can be viewed in two ways. The first way
is that it describes the force exerted by the neighboring material on a unit volume and
the resultant change in the momentum flow-out rate. According to Newton’s second
law, external force on an object causes temporal change in momentum of the object.
From this viewpoint, .j / on the right-hand side of Eq. (4.90) can be interpreted
as the external force acting on the unit volume at the surface. The right-hand side
represents the differential force that causes the momentum change. If r  .j / > 0,
the differential force is outward. This causes the volume of the elastic material to
be expanded. As long as the outward force acts on the surface, the volume keeps
expanding. Consequently, the rate of the momentum flow-out decreases. With the
negative sign, the left-hand side of Eq. (4.90) represents the decrease in the rate.
In this view, .j / having a unit of .kg  m=s2 /=m3 can be identified as the force
density. If the force is removed, the volume stops expanding and goes back to the
initial volume. The initial momentum flow-out rate is resumed.
In the second view, j on the right-hand side of Eq. (4.90) represents an
irreversible flow of j0 D .r  v/ through the boundary of the volume.6 This
situation is better represented if we rewrite Eq. (4.90) using .r  v/ D j0 as
follows:
@
.j0 / D r  .j / (4.91)
@t
j D W d .j0 / (4.92)

With the negative sign, the right-hand side of Eq. (4.91) indicates that vector
.j / is inward to the volume where the left-hand side indicates that this inward
.j / contributes to the increase of the quantity .j0 / over time. Equation (4.92)
explicitly represents this flow of .j0 / with Wd being the drift velocity. Since the
flow is irreversible, .j0 / can be viewed as “free” particles. The right-hand side
of Eq. (4.92) is in (m/s) (kg/m3 ) (1/s), or the change in the momentum per unit
volume over unit time. It represents an irreversible change in the total momentum
of the volume due to the flow of the free particles through the boundary. From this
viewpoint, j can be identified as the flux of momentum density passing through
the boundary surface of the unit volume. The situation corresponds to plastic flow in
which local concentrated strain (a plastic front) moves in a certain direction causing
permanent (irreversible) deformation behind. In this context, the right-hand side of
Eq. (4.91) can be interpreted as differential, energy-dissipative force acting on the
unit volume. See Sect. 5.2.2 for more detailed discussion about the irreversible flow
of .j0 / D .r  v/.
In the electrodynamic analogy, the first view corresponds to polarization of bound
electrons due to an external electric field, and the second view corresponds to a

6
The flow is irreversible because unlike the first view there is no mechanism to reverse the temporal
change represented by the left-hand side.
88 4 Field Theory of Deformation and Fracture

Fig. 4.10 Poynting vector of ∂xx ∂xx


deformation dynamics S = v × t = −vG (xˆ × z)
ˆ = vG ŷ
∂y ∂y

xx (yP)
∂xx
P v = vxˆ t = Gw = −G ẑ
∂y
y
∂x
x w = − x ẑ
z ∂y
Q
xx (yQ)

conduction current. Based on the analogy, we call j0 the deformation charge. Here
j0 corresponds to the positive electric charge density. This difference in the sign
convention from electrodynamics does not have a physical significance; it comes
from the way we define the charge of symmetry in Eq. (4.41). More will be discussed
on the charge and related issues in the following chapters.
In electrodynamics, the Poynting vector Se is defined as
1
Se D E B (4.93)
0
Based on the similarities discussed above (and also in the next chapter), the
deformation-dynamics version of Poynting vector can be deduced as follows:

S D Gv  ! (4.94)

Figure 4.10 illustrates this Poynting vector schematically. Consider point P where
the shear stress is G.d!z =dy/  dy D G! (N/m2 ). In one second, the displacement
of point P is Px D vx (m/s). Thus the power is G!vx (W/(m2 )). This is the power
per unit area that propagates in the direction indicated in Fig. 4.10.

References

1. Frank, F.C.: On the equations of motion of crystal dislocations. Proc. Phys. Soc. (Lond.) A 62,
131–134 (1949)
2. Eshelby, J.D.: Proc. Phys. Soc. A 62, 307 (1949)
3. Eshelby, J.D.: Supersonic dislocations and dislocations in dispersive media. Proc. Phys. Soc.
B 69, 1013–1019 (1956)
4. Leibfried, G., Dietzeh, H.D.: Z. Phys. 126, 790 (1949)
5. Weertman, J.: Uniformly moving transonic and supersonic dislocations. J. Appl. Phys. 38,
5293–5301 (1967)
6. Kröner, E.: Die inneren Spannungen und der Inkompatibilitätstensor in der Elastizitätstheorie.
Zeitschrift für angewandte Physik 7, 249–257 (1955)
References 89

7. Kröner, E.: Continuum theory of defects. In: Balian, R., et al. Physics of Defects (Les Houches,
Session 35), pp. 215–315. North-Holland, Amsterdam (1981)
8. Holländer, E.F.: The basic equations of the dynamics of the continuous distribution of
dislocations I,II,III. Czech. J. Phys. B 10, 409–418 (1960); 10, 479–487 (1960); 10, 551–560
(1960)
9. Bovet, D.: The continuous theory of dislocations in elastostatics and elastodynamics. Sol.
Mech. Arch. 4, 31–96 (1979)
10. Mistura, L.: Crystal dislocations and Dirac monopoles. Int. J. Eng. Sci. 33, 2149–2159 (1995)
11. Seeger, A.: Z. Natwf. 8a, 47 (1955); Handbuch der Physik, VII/l, p. 622. Springer, Berlin
(1953)
12. Kadić, A., Edelen, D.G.B.: Historical Remarks. In: A Gauge Theory of Dislocations and
Disclinations. Lecture Notes in Physics, vol. 174, Chap 1. Springer, New York/Tokyo (1983)
13. Egorushkin, V.E.: Gauge dynamic theory of defects in nonuniformly deformed media with a
structure, interface behavior. Sov. Phys. J. 33, 135–149 (1990)
14. Egorushkin, V.E.: Dynamics of plastic deformation: waves of localized plastic deformation in
solids. Rus. Phys. 35, 316–334 (1992)
15. Egorushkin, V.E.: Gauge theory in mechanics of continuum media. In: Panin, V.E. (ed.)
Physical Mesomechanics and Computer-Aided Design of Materials, vol. 1, p. 241. Nauka,
Novosibirsk (1995, in Russian)
16. Edelen, D.G.B.: A correct, globally defined solution of the screw dislocation problem in the
gauge theory of defects. Int. J. Eng. Sci. 34, 81–86 (1996)
17. Lazar, M.: On the fundamentals of the three-dimensional translation gauge theory of disloca-
tions. Math. Mech. Solids 16, 253–264 (2011)
18. Griffiths, D.J.: Introduction to Electrodynamics, 3rd edn. Prentice Hall, Upper Saddle River
(1999)
19. Panin, V.E., Grinaev, Y.V., Egorushkin, V.E., Buchbinder, I.L., Kul’kov, S.N.: Sov. Phys. J. 30,
24–38 (1987)
20. Panin, V.E.: Wave nature of plastic deformation. Sov. Phys. J. 33(2), 99–110 (1990)
21. Danilov, V.I., Zuev, L.B., Panin, V.E.: Wave nature of plastic deformation of solids. In: Panin,
V.E. (ed.) Physical Mesomechanics and Computer-Aided Design of Materials, vol. 1, p. 241.
Nauka, Novosibirsk (1995, in Russian)
22. Panin, V.E.: Physical fundamentals of mesomechanics of plastic deformation and fracture of
solids. In Panin, V.E. (ed.) Physical Mesomechanics of Heterogeneous Media and Computer-
Aided Design of Materials. Cambridge International Science Publishing, Cambridge (1998)
23. Yoshida, S.: Field theoretical approach to dynamics of plastic deformation and fracture. AIP
Conf. Proc. 1186, 108–119 (2009)
24. Yoshida, S.: Physical meaning of physical-mesomechanical formulation of deformation and
fracture. AIP Conf. Proc. 1301, 146–155 (2010)
25. Yoshida, S.: Scale-independent approach to deformation and fracture of solid-state materials.
J. Strain Anal. 46, 380–388 (2011)
26. Aitchson, I.J.R., Hey, A.J.G.: Gauge Theories in Particle Physics. IOP Publishing, Bris-
tol/Philadelphia (1989)
27. See for example, Marsden, J.E., Hughes, T.J.R.: A point of departure. In: Mathematical
Foundations of Elasticity. Prentice-Hall, Englewood Cliffs (1983)
28. Kaiser, K.L.: Electromagnetic Compatibility Handbook. CRC Press, Boca Raton (2004)
29. Yoshida, S.: Consideration on fracture of solid-state materials. Phys. Lett. A 270, 320–325
(2000)
30. Yoshida, S.: Interpretation of mesomechanical behaviors of plastic deformation based on
analogy to Maxwell electromagnetic theory. Phys. Mesomech. 4(3), 29–34 (2001)
31. Yoshida, S.: Physical mesomechanics as a field theory. Phys. Mesomech. 8(5), 15–20 (2005)
32. Landau, L.D., Lifshitz, E.M.: Theory of Elasticity. Course of Theoretical Physics, 3rd edn.,
vol. 7. Butterworth-Heinemann, Oxford (1986)
Chapter 5
Interpretations of Deformation and Fracture
Phenomena from Field Theoretical Viewpoint

The goal of this chapter is to discuss various concepts evolving from the field
equations derived in the preceding chapter. In Sect. 5.1, one of the field equations
is interpreted as the equation of motion that governs the wave-like dynamics of
a unit volume in solid-state objects under deformation. The physical meaning of
each term of the field equation is discussed. This equation of motion has two
force terms, which are interpreted as the transverse effect (shear force) and the
longitudinal effect (normal force), respectively, acting on the unit volume. This
will be followed by Sect. 5.2 where all stages of deformation, from the elastic
through fracturing stage, are comprehensively formulated based on the same field
equations. It will be shown that the dynamics of different stages are characterized
by the difference in the form of the longitudinal effect term in the field equation.
The argument will show that the longitudinal effect in the elastic regime represents
the elastic force whose magnitude is proportional to the displacement and that the
longitudinal effect in the plastic regime represents the momentum loss that the
unit volume is subject to. The mechanism of this momentum loss is explained
through the introduction of the deformation charge and its flow (current) analogous
to the electric charge and current. Sections 5.3 and 5.4 will focus on the gauge
theoretical aspect of the present formalism. In Sect. 5.3, the physical meaning of
the potential associated with the local symmetry of the linear elasticity is discussed.
The physical meanings of the deformation charge and current will be discussed in
conjunction with the energy dissipative nature of plastic deformation. The analogy
of the deformation charge and current to the electric charge and current will be
discussed based on the underlying physics. Behaviors of the deformation charge
and current are discussed based on the concept of field force and Lenz’s law. In
Sect. 5.4, the physical meaning of the deformation charge will be discussed from
the viewpoint of symmetry charge associated with the local symmetry. Section 5.5
is to briefly discuss the irreversibility of plastic deformation from the viewpoint of
thermodynamics. A short discussion on the connection of the present formalism to
the generation and propagation of dislocations will also be made.

© Springer Science+Business Media New York 2015 91


S. Yoshida, Deformation and Fracture of Solid-State Materials,
DOI 10.1007/978-1-4939-2098-3__5
92 5 Interpretations of Deformation and Fracture Phenomena from Field Theoretical. . .

5.1 Field Equation as an Equation of Motion

Rearranging the terms, we can put the field equation (4.63) in the following form
[1, 2].

1 @v
D r  !  j (5.1)
c 2 @t
As discussed in the preceding chapter [see Eq. (4.81)], the first term on the right-
hand side of Eq. (5.1) represents the shear effect on a unit volume by neighboring
volumes.
p Thus, it is natural to identify the phase velocity on the left-hand side as
G= [see Eq. (4.39)]. We can then rewrite Eq. (5.1) as follows:

@v
 D Gr  !  Gj (5.2)
@t
Here  is the density and G is the shear modulus. The left-hand side of Eq. (5.2)
is the change in the momentum of the unit volume over time. Thus, Eq. (5.2) can
be viewed as an equation of motion. According to Newton’s second law, the right-
hand side can be interpreted as representing the external force acting on the unit
volume. Since the unit volume is in a continuum medium, this force must be exerted
by surrounding materials. Of the two terms on the right-hand side, the first term is
identified as shear force exerted by the neighboring unit volumes that sandwich this
unit volume (Fig. 4.8). It is a transverse effect.
The second term, on the other hand, represents a longitudinal effect. In fact, this
term plays a very important role in the present formalism from the stand point of
comprehensive description of deformation dynamics of all stages. In accordance
with Newton’s second law, the net external force acting on a volume can be viewed
as the change in the momentum of the volume per unit time. When the unit volume
is flowing, as is the case of a given unit volume of an object under deformation,
the temporal change can be considered as a flux. This position is more clearly
understood via the equation of continuity (4.87).

1 @j0
Dr j (4.87)
c 2 @t
With the same substitution of c as above, and the use of Eq. (4.89), Eq. (4.87) can
be put in the following form:

@j0 @.r  v/ @
 D D .r  v/ D r  .Gj / (5.3)
@t @t @t

So, Gj can be viewed as a flow (flux) of r  .v/, or divergence of momentum


density. Notice that Gj is a special case of j discussed at the end of the preceding
chapter, where being a pure plastic case the elastic modulus  is the shear modulus
5.2 Comprehensive Description of Deformation 93

G. With the negative sign in front, the right-hand side of Eq. (5.3) represents the net
flow of momentum flux into the volume. The left-hand side represents the resultant
increase of the deformation charge j0 D .r  v/. With W d being the velocity of
the flow, we can put Gj in the following form:

Gj D W d r  .v/ (5.4)

In this context, the velocity W d represents the flow of the spatial dependence of the
velocity field, not a flow of a material. From this standpoint, it can be interpreted
as a wave velocity in the broad sense. With an appropriate form of j , the equation
of motion (5.2) can represent a given stages of deformation. Below, we explore the
explicit form of j under various conditions.

5.2 Comprehensive Description of Deformation

5.2.1 Elastic Deformation

We derived the field equations (4.61)–(4.64) by allowing coordinate dependence


of the transformation matrix that represents locally linear elasticity. It is natural
to assume that these field equations are intrinsically able to represent elastic
deformation at the global level as well. The elastic deformation is essentially
a longitudinal effect; it is based on Hooke’s law which states that an elastic
object is stretched or compressed in the direction of the applied force where the
magnitude of the stretch or compression is proportional to the force. Note that the
deformation perpendicular to the applied force is not due to the effect represented
by Hooke’s law, but to the Poisson’s effect which simply describes the amount of the
perpendicular deformation based on an empirically known constant. The argument
in the preceding section shows us the longitudinal force is represented by the second
term on the right-hand side of Eq. (4.63), j . From the gauge theoretical viewpoint,
this term represents how the Lagrangian interacts with the displacement field. Thus,
it should be possible to argue the elastic limit in conjunction with j .

Compression Wave

The equation of compressive elastic wave can be derived from the equation of
motion (5.2). Take divergence of Eq. (5.2).

@.r  v/
 D G.r  j / (5.5)
@t
Here the first term on the right-hand side of Eq. (5.2) vanishes because the
mathematical identity states that r  .r  !/ D 0 for any vector !. Replacing v
94 5 Interpretations of Deformation and Fracture Phenomena from Field Theoretical. . .

Fig. 5.1 Differential volume


(∇ ⋅ x )
expansion

(∇ ⋅ x )

(∇ ⋅ x )

with displacement  as,

@
vD (5.6)
@t
Equation (5.5) becomes as follows:

@2
 .r  / D G.r  j / (5.7)
@t 2

In continuum mechanics [3–6], the quantity .r  / is known as the volume


expansion (Fig. 5.1). The elastic compressive wave is a wave of such volume
expansion. Thus it is possible to interpret the temporal derivative @=@t as being
associated with the elastic compressive wave’s propagation. With this interpretation,
the right-hand side of Eq. (5.7) should represent the spatial variation of .r/. Below
we discuss this interpretation.
Consider first replacing the temporal derivative appearing on the left-hand side
of Eq. (5.7) with spatial derivative. Essentially, a wave is propagation of oscillation
through space, where the temporal periodicity is the period T and the spatial
periodicity is the wavelength . The phase velocity vph is the ratio between them
vph D =T , which can be viewed as the conversion factor from the temporal to
spatial periodicity. In general, a traveling wave function .t; r/ can be put in the
following form [7].

.t; r/ D .!t  k  r/ D .!t  k.˛x C ˇy C  z// (5.8)

where ! D 2=T is the angular frequency, k is the propagation vector, r is the


position vector representing the coordinate points, and ˛, ˇ, and  are the directional
cosines of the propagation vector. The temporal derivative of Eq. (5.8) is

@ 0
D! (5.9)
@t
5.2 Comprehensive Description of Deformation 95

Here 0 .s/ D d =ds with s being an independent variable for the function . The
spatial derivative in this context is the directional derivative along the propagation
of the wave. Referring to Eq. (3.9), we obtain
 
@ @ @
.r /  kO D xO C yO C zO  .˛ xO C ˇ yO C  zO/
@x @y @z
0 0
D .k˛ xO  kˇ yO  k zO/  .˛ xO C ˇ yO C  zO/ D k (5.10)

Here,
@ 0
D .k˛/
@x
@ 0
D .kˇ/
@y
@ 0
D .k / (5.11)
@z

are used in going through the second equal sign, and

˛2 C ˇ2 C  2 D 1 (5.12)

is used in going through the third equal sign. The condition (5.12) comes from the
definition of ˛ xO C ˇ yO C  zO as the unit vector in the direction of the propagation
O1
vector k.
From Eqs. (5.9) and (5.10),

1@ 1
0
D D  .r /  kO (5.13)
! @t k
Therefore, in wave dynamics, the temporal differentiation can be converted to the
spatial differentiation as follows:

@ !
D  .r /  kO D c.r /  kO D .r /  c kO D .r /  c (5.14)
@t k
where c is a vector representation of the wave traveling with the phase velocity c.

1
The fact that the magnitude of a vector having directional cosines as its components is unity
can be proved as follows. Put u D ˛ xO C ˇ xO C  x. O Suppose the angle between u and the x-
axis is ˛ . Then the x-component of u is ux D u  xO D juj jxj O cos ˛ D juj cos ˛ D juj ˛.
Similarly, uy D juj cos ˇ D juj ˇ; uz D juj cos  D juj . So, juj2 D .ux /2 C .uy /2 C .uz /2 D
.˛ 2 C ˇ 2 C  2 / juj2 . It follows that .˛ 2 C ˇ 2 C  2 / D 1, i.e., u is a unit vector parallel to k or
u D k.O
96 5 Interpretations of Deformation and Fracture Phenomena from Field Theoretical. . .

Substituting .r  / for , the temporal derivative of Eq. (5.7) becomes

@.r  /
D r.r  /  c (5.15)
@t

Similarly, considering r 2 ,
   
@ @ @ @ @ @
.r 2
/D xO C yO C zO  xO C yO C zO
@x @y @z @x @y @z
 
@ @ @ 
D xO C yO C zO  k˛ 0 xO  kˇ 0 yO  k 0
zO
@x @y @z
 0 
@ @ 0 @ 0
D k ˛ Cˇ C
@x @y @z
 00 00

D k k˛  kˇ  k 00 D k 2 00 (5.16)

and

@2 00
D !2 (5.17)
@t 2
we obtain

@2 !2 2
D .r / D c 2 .r 2 / (5.18)
@t 2 k2
0
Here, Eq. (5.11) is applied to in going through the last equal sign of Eq. (5.16).
Again, substituting .r  / for ,

@2 .r  /
D c 2 .r 2 .r  // D c 2 r  r.r  / (5.19)
@t 2
Replacing the temporal derivative in the left-hand side of Eq. (5.7) with the
rightmost-hand side of Eq. (5.19), we obtain

c 2 r  r.r  / D r  .Gj / (5.20)

Comparing both-hand sides of this equation, we find that

Gj D c 2 r.r  / (5.21)


5.2 Comprehensive Description of Deformation 97

The elastic compressive wave velocity is known to be [3]


s
C 2G
cD (5.22)


where is the Lamé’s constant and G is the shear modulus. Substituting Eq. (5.22)
for c in Eq. (5.21), we obtain

Gj D . C 2G/r.r  / (5.23)

Equation (5.23) is the longitudinal force on the unit volume causing an elastic
compressive wave. The force is due to the gradient of the volume expansion, i.e.,
the differential volume expansion at the boundary of the unit volume as Fig. 5.1
illustrates.
From Eq. (5.23), we can identify the current j for three-dimensional elastic case
to be in the following form:

. C 2G/
j D r.r  / (5.24)
G
Substituting this expression of Gj into Eq. (5.7), we obtain

@2
 .r  / D r  Œ. C 2G/r.r  /
(5.25)
@t 2
which leads to

@2 . C 2G/
.r  / D r 2 .r  / (5.26)
@t 2 

This is a more usual form known as the equation of compression wave in continuum
mechanics [3, 8].
Alternatively, by substituting Eq. (5.23) to Eq. (5.2), this wave equation can be
given in the following form:

@v
 D Gr  ! C . C 2G/r.r  / (5.27)
@t
This form explicitly indicates that the external force acting on the unit volume has
shear force portion (the first term on the right-hand side) and longitudinal force
associated with the gradient of volume compression.
98 5 Interpretations of Deformation and Fracture Phenomena from Field Theoretical. . .

Further, using ! D r  , r  r D rr    r 2  and replacing v on the right-


hand side with  using Eq. (5.6), we can eliminate the rotation ! from Eq. (5.27) as
follows:

@2 
 D Gr 2  C . C G/r.r  / (5.28)
@t 2

Rotation (Deformation) Wave

The equation of motion for the wave known as rotation (deformation) wave in
continuum mechanics can be derived from Eq. (5.27). Consider curl of both-hand
sides of Eq. (5.27). Using mathematical identities r  .r  !/ D r.r  !/  r 2 !
and r  rf D 0 (f is a scalar function), and r  ! D 0 [Eq. (4.64)], we obtain

@.r  v/
 D Gr 2 ! (5.29)
@t
Again, replacing v with  and using ! D r  , this becomes

@2 !
 D Gr 2 ! (5.30)
@t 2
Equation (5.30) is known as the equation of rotational (deformation) wave in
continuum mechanics.

One-Dimensional Elastic Wave

The dynamics of one-dimensional elastic wave, also known as the longitudinal


elastic wave propagating through a bar of elastic medium, is most conveniently
discussed via wave equations derived from field equations (4.62) and (4.63). Taking
curl of Eq. (4.63) and temporal derivative of Eq. (4.62), we can eliminate ! from
these equations.

@.r  !/
r  .r  v/ D (5.31)
@t
@.r  !/ 1 @2 v @j
D 2 2  (5.32)
@t c @t @t

With the use of r  .r  v/ D r.r  v/  r 2 v, this leads to

1 @2 v @j
D r 2v   r.r  v/ (5.33)
c 2 @t 2 @t
5.2 Comprehensive Description of Deformation 99

Integrating both-hand sides of Eq. (5.33) with respect to time, we obtain the
following wave equation for .

1 @2 
D r 2   j  r.r  / (5.34)
c 2 @t 2
In continuum mechanics, the wave equation of a longitudinal elastic wave is given
as follows:

@2 
 D Er 2  (5.35)
@t 2
Here E is the Young’s modulus. Equation (5.34) reduces to Eq. (5.35) if the current
j and the phase velocity c satisfy the following conditions.

j D r.r  / (5.36)

E
c2 D (5.37)


Equation (5.36) can be interpreted as a one-dimensional version of volume expan-


sion shown in Fig. 5.1. Equation (5.37) is the phase velocity of a longitudinal
elastic wave traveling through an elastic medium of Young’s modulus E. As a
one-dimensional wave, the displacement field associated with Eq. (5.35) represents
linear elasticity. Therefore, in this case, there is no need to regain local symmetry
with the introduction of vector potential A. Thus, the coordinate dependence of
rotation, r  !, becomes irrelevant to the dynamics; hence it is reasonable to put
r  ! D 0. Substituting this condition along with the expression of j given by
Eq. (5.36), the equation of motion in this case can be identified from Eq. (5.2) as
follows:

@v
 D Er.r  / (5.38)
@t

In the context of one-dimensional longitudinal wave, r.r  / can be interpreted


as @2 =@x 2 , where x denotes the coordinate variables of the axis along which the
longitudinal wave propagates. Substitution of @2 =@x 2 for r.r  / lets us write
Eq. (5.38) in the form of one-dimensional wave equation.

@v @2 
 DE 2 (5.39)
@t @x
The right-hand side of Eq. (5.39) represents the normal stress per unit length or force
density in the unit of ŒN=m3
. In the present formalism, it can be viewed as a force
exerted by the displacement field.
100 5 Interpretations of Deformation and Fracture Phenomena from Field Theoretical. . .

Equations (5.24) and (5.36) indicate that current j for the elastic case is given
in the form of gradient. Mathematically, r  rf D 0 (f is a scalar function).
This means that the current j does not possess rotational nature. It is interesting
to explore how this fact affects the wave characteristics of the rotational mode
of displacement ! discussed above. For this purpose, let us derive the wave
equation governing ! directly from the field equations. Taking temporal derivative
of Eq. (4.62) and curl of Eq. (4.63), eliminate .

@ @2 !
r v D (5.40)
@t @t 2
1 @.r  v/
r  .r  !/ D  2 r j (5.41)
c @t
This leads to the following wave equation of !.

1 @2 !
D r 2!  r  j (5.42)
c 2 @t 2
Since j is proportional to gradient of r   [Eqs. (5.24) and (5.36)], r  j D 0.
Thus,

1 @2 !
D r 2! (5.43)
c 2 @t 2
p
With the use of shear wave velocity G= [Eq. (4.39)], Eq. (5.43) reduces to
Eq. (5.30) derived above.
In summary, in the elastic case, the longitudinal effect j represents force
proportional to the gradient of r  .2 Being proportional to the gradient of a scalar
function, by nature j is non-rotational.

5.2.2 Plastic Deformation

As briefly discussed in the preceding chapter, in the plastic regime the longitudinal
effect represented by the first field equation [Eq. (4.61)] is not elastic or restoring
force. Rather, it is momentum loss associated with a flow of a charge. The loss of
longitudinal restoring force causes r  v appearing in Eq. (4.61) not to contribute to

2
Here r   represents the volume expansion of the infinitesimal region at the coordinate point
where one end-surface of the unit volume is located. Thus, its gradient represents the differential
volume expansion of the unit volume. Note that the force associated with the volume expansion is
due to the elastic force proportional to the displacement; when a volume is expanded, the elastic
force on the end surfaces pushes back the volume to oppose the expansion by exerting resiting
force proportional to the displacement from the equilibrium.
5.2 Comprehensive Description of Deformation 101

Effective Force
vx (resistive)
Gj

vx
>0
Wd x

Fig. 5.2 A one-dimensional positive charge flows in the positive x direction. Particles behind the
passage of charge (hatched region) lose kinetic energy as the velocity decreases. The material loses
mechanical energy. This dynamics can be viewed as energy dissipative resiting force acting on the
charge in the negative x direction

the wave dynamics. The transverse effect represented by Gr  ! [Eq. (5.2)] is still
effective. Consequently, the field equation yields a transverse wave, for the same
reason as an electromagnetic wave is always a transverse wave as already discussed
(see Sect. 4.4). The momentum loss associated with a flow of a charge causes the
displacement wave to decay. Thus, in the pure plastic regime, the deformation wave
is a decaying transverse wave. Below, we consider this situation quantitatively.
Under the condition where charges flow with drift velocity W d [Eq. (5.4)] as the
shear force Gr  ! acts on the unit volume, we can rewrite the equation of motion
(5.2) by replacing Gj with W d .r  v/.

@v
 D Gr  !  W d .r  v/ (5.44)
@t
When a charge flows, the material loses mechanical energy stored in the
displacement field. Consider the situation in Fig. 5.2 where a one-dimensional
positive charge @vx =@x > 0 drifts in the positive x direction. As the positive slope
of velocity’s x dependence drifts forward in the positive x direction, the particles
behind the passage of the charge lose velocity as compared with before the passage.
Consequently, the corresponding block of the material loses the total momentum.
The same argument holds for a negative charge @vx =@x < 0 when it drifts against
the velocity vx or in the negative x direction.
According to Newton’s second law, a momentum change over time is equivalent
to an external force acting on the object. From this viewpoint, W d .r  v/ on the
right-hand side of Eq. (5.44) can be interpreted as energy dissipative force acting
on the charge contained in a unit volume of the material (force density). This is
contrastive to the elastic case where Gj represents restoring force. While both
forces act on a volume in the same direction, one is energy conservative force with
102 5 Interpretations of Deformation and Fracture Phenomena from Field Theoretical. . .

its magnitude proportional to the displacement (spring force) and the other is energy
dissipative force whose magnitude is proportional to the velocity (velocity damping
force).
Naively speaking, the mechanism that the particles behind the passage of the
charge lose velocity can be ascribed to the generation of defects. These defects
weaken the stiffness at the tailing edge of a positive charge. Consequently, the
particles in the material behind the charge lose their velocity (left behind because the
material ahead cannot pull them strongly enough to maintain the original velocity
due to the weakened stiffness). This picture qualitatively agrees with the general
explanation of the Lüders band [9]. In the case of a negative charge, the same
argument applies to the leading edge of the charge. It is important to note that
the drift velocity of a charge is different from the particles’s velocity. As will be
discussed below, the drift velocity depends on the material and its condition as well
as the degree of plastic deformation (how close to fracture). In fact, the ratio of the
charge’s drift velocity to the particle velocity is an important factor that determines
the degree of plasticity as will be discussed shortly under “Charge flow and decay
of displacement wave.”

Charge in Elastic Regime

Note that having a charge in a volume does not necessarily mean that the material
is being deformed plastically. The mechanical (elastic) field loses energy when
a charge drifts in the way described above. Even in the pure elastic regime,
the velocity field can have a charge-like pattern, and the field does not lose
mechanical energy. Consider a one-dimensional compression wave, as an example.
The stretched portion of the wave has a pattern of @vx =@x > 0, and the pattern
can travel in the positive x direction. However, as a compression wave, there is
always a compressed portion in the material next to the stretched portion. While the
stretched portion loses the net momentum via particles’ flow-out rate r  v > 0,
the compressed portion gains momentum via particle flow-in r  v < 0. Overall,
the field does not lose mechanical energy. This argument indicates that the elastic
case should be discussed as a special case where the charge-like pattern drifts at the
phase velocity of the compression wave. Below, we consider this argument more
quantitatively.
The situation is most conveniently argued via analysis of Gj as a flow of charge
.r  v/ in comparison with the corresponding elastic longitudinal force. In the
plastic regime, the longitudinal effect is

Gj D W d .r  v/ (5.45)

Using v D @=@t and the conversion from the temporal derivative to spatial
derivative for .r  / wave given by Eq. (5.15), we can rewrite Eq. (5.45) as follows:
5.2 Comprehensive Description of Deformation 103

@.r  /
Gj D W d  D W d  .r.r  //  c D W d c .r.r  //  cO (5.46)
@t

where cO is the unit vector in the direction of c. Compare Eq. (5.46) with the
previously derived expression of longitudinal elastic force (5.21).

Gj D celas
2
r.r  / (5.47)

Here the subscript “elas” is used to emphasize that Eq. (5.21) represents the elastic
longitudinal force. The scalar-product part of the rightmost-hand side of Eq. (5.46)
is the projection of .r.r  // vector onto the cO axis (the cO component of the vector).
Thus, the vector represented by the entire rightmost-hand side of Eq. (5.46) is anti-
parallel to W d vector in direction, and Wd c times the cO component of .r.r  //
in magnitude. On the other hand, the vector represented by the right-hand side of
Eq. (5.47) is anti-parallel to .r.r  // in direction, and jr.r  /j times celas2
in
magnitude. These two vectors are identical under the conditions: (a) .r.r  // is in
the same direction as c so that its cO component is equal to its magnitude itself, and
(b) the magnitude of W d vector and c vector are equal to celas . Condition (a) can be
understood as follows. When Eq. (5.46) represents elastic deformation, the elastic
force is proportional to .r.r  // and the wave is a longitudinal wave. Therefore,
the direction of the elastic wave is the same as the elastic force.
These arguments indicate that the elastic wave is a special case where (r  v/
travels in the direction of the elastic longitudinal wave at the drift velocity Wd equal
to the phase velocity of the elastic wave.

W d D c D c elas (5.48)

In this context, the charge is the compressed/stretched pattern of the displacement,


and the longitudinal wave is a compression wave. The elastic phase velocity celas is
the rate of energy flow (the longitudinal wave phase velocity, not Poynting vector).

Charge Flow and Decay of Displacement Wave

Experiments [10–12] based on optical interferometry indicate that under certain


conditions the displacement field of a tensile-loaded specimen can show a spatial
pattern that can be interpreted as a developed, one-dimensional charge. Figure 5.3
illustrates such a pattern schematically. Here, the charge is said to be one-
dimensional because the velocity has spatial dependence in only one direction; in
Fig. 5.3, the xs -axis represents the coordinate axis that velocity v is dependent on,
and xp - and z-axes are orthogonal to the xs -axis; i.e., @=@xs ¤ 0 and @=@xp D
@=@z D 0. Under this condition, r  v D @vs =@xs . The word “developed” is used to
mean that this pattern of spatial dependence spans across the width of the specimen.
This type of patterns (the charge-like patterns) are observed to drift along the xs
axis with their shapes (the pattern of @vs =@xs ) practically unchanged. Figure 5.4
104 5 Interpretations of Deformation and Fracture Phenomena from Field Theoretical. . .

Fig. 5.3 Developed


charge-like pattern vp

vs
z

shows an example of such a one-dimensional charge drifting, where the top images
are interferometric patterns representing contours of constant velocity,3 the middle
picture shows that the charge drifts to the right with the head velocity vr and tail
velocity vl , and the bottom picture illustrate the situation where the particles lose
momentum as the charge drifts. More will be discussed about developed charges
and their drifting characteristics in the following sections.
Another series of similar tensile experiments [13, 14] indicate that the dis-
placement field shows decaying transverse wave characteristics and that the wave
characteristics disappear as a developed charge appears. One of these experiments
[13] shows a slowly decaying transverse wave that decays (loses the oscillatory
behavior) when a developed charge appears in a very late stage of deformation.4
This experimental result leads to a hypothesis that undeveloped charges exist in an
early stage of plastic deformation causing the mechanical energy to dissipate. With
the development of deformation, the charge grows in scale and eventually its spatial
dependence spans across the width of the specimen. This hypothesis is consistent
with the claim by Panin et al. that the scale of plastic deformation develops toward
the fracture [17].
Based on the above hypothesis, the decaying wave dynamics in the plastic regime
can be argued by analogy with electrodynamics. Figure 5.2 and the associated
discussion indicate that when a positive charge .r  v/ > 0 flows in the direction of
the particles’ velocity v, the mechanical energy of the system is dissipated. Gauge
theoretical consideration on the field force indicates that a positive charge (r v/ > 0
experience force in the same direction as v (see Sect. 5.3.3 below). This allows us
to put

W d D 0 v (5.49)

3
More precisely, the dark stripes in the top images are contours of the velocity component that
the interferometer is sensitive to. In this particular case, the interferometer is sensitive to the
vertical component (the component perpendicular to the longer side of the specimen in Fig. 5.4).
Therefore, strictly speaking, the contours represent the constant value of the vertical component
of the velocity. However, a number of similar experiments [10, 15, 16] using an interferometer
with two-dimensional sensitivity indicate that both in-plane displacement components show stripes
exactly at the same locations. This strongly indicates that these stripes represent the contours of
the same absolute value of the velocity vector, not its vertical component only.
4
About the beginning of the final 20 % of the life span of the specimen under the constant pulling
rate.
5.2 Comprehensive Description of Deformation 105

x x+Dx
Dx

vl vh vl vh

x x+Dx

vh vh

vl vl

x x

Fig. 5.4 One-dimensional, developed charge

where 0 > 0 is a constant of proportionality. As we discussed in the preceding


chapter [see near Eq. (4.92)] and earlier in this chapter [Fig. 5.2 and texts near
Eq. (5.44)], the passage of charge causes energy dissipation associated with the
generation of defects. From this viewpoint, the magnitude of W d is determined
by how fast the material is yielded or how easily defects are generated. Thus, it is
considered that the value of 0 is related to the dislocation velocity; the faster the
dislocation flows, the greater Wd as compared with v. As known well, the dislocation
velocity varies over many orders of magnitude depending on the material and its
conditions such as the temperature [18]. So, it is reasonable to assume that 0 is a
material-dependent constant.
Using Eq. (5.49), we can put W d .r  v/ D 0 v.r  v/, and rewrite Eq. (5.44)
as follows:

@v
 D Gr  !  0 .r  v/v
@t
D Gr  !  c v (5.50)

Here c defined by Eq. (5.51) corresponds to the conductivity of electrodynamics.

c D 0 .r  v/ (5.51)

The physical meaning of c can be interpreted as follows. v.r  v/ is the total


momentum carried by the unit volume per second as its particles flow with the
average velocity v. If 0 D 1, the drift velocity Wd is equal to vs ; the term
c v D 0 .r  v/v appearing on the right-hand side of Eq. (5.50) represents this
106 5 Interpretations of Deformation and Fracture Phenomena from Field Theoretical. . .

total momentum carried by the mass  drifting with the drift velocity v.5 When
the value of 0 is higher than 1, some particles are left behind the pattern r  v
as it drifts with velocity o v which is higher than the particle velocity v. The
hatched area in the bottom right picture of Fig. 5.4 illustrates those “left-behind”
particles schematically. The higher the value of 0 , the more particles are left
behind because the hatched area increases, meaning that the more momentum hence
the kinetic energy is lost. 0 represents the degree of this momentum loss, and
c v D 0 .r  v/v represents the actual momentum loss in Eq. (5.50). For the
same 0 and v, the higher the value of r  v (steeper the slope of d v=dx in the
one-dimensional case shown in Fig. 5.4), the greater the c . In the case of a one-
dimensional charge, this means that the narrower the charge, the steeper the slope
d v=dx and therefore the greater the 0 or the more energy dissipative. Experiments
[10, 11] indicate that developed, one-dimensional charges tend to narrow in their
width in the final stage of plastic deformation transitional to the fracturing stage,
indicating that towards the end of deformation the material becomes more energy
dissipative. The stress concentration causing the necking (strain concentration) in
the last stage of tensile loading can be understood as an example of such elevated
energy dissipation. It is interesting to note that this process is analogous to the
situation where in a late stage of electric breakdown of gases the conductivity is
raised as the free charges (usually heavy ions) are localized [19]. Normally, the
elevated local free-charge density leads to an arc discharge, which in turn leads to
the electric breakdown of the gas [20].
Elimination of ! from Eq. (5.50) with the use of field Eq. (4.62) leads to the
following wave equation that governs v.

@2 v @v
 2
 Gr 2 v C c D Gr.r  v/ (5.52)
@t @t
The optical-interferometric image-pattern representing a developed, one-
dimensional charge (e.g., the top image in Fig. 5.4) normally consists of equidistant,
parallel fringes, indicating that r  v D d v=dx is approximately a constant [10–12].
With the assumption that r  v drifts without changing its shape (r  v is constant
or r.r  v)=0), its spatial dependence can be assumed to be null and the right-hand
side of Eq. (5.52) can be put zero.

@2 v @v
 2
 Gr 2 v C c D0 (5.53)
@t @t
In the case of a one-dimensional charge, the general solution to Eq. (5.53) can be
put in the following form:

5
In the one-dimensional picture, v.r  v/ D vs .@vs =@xs /. Here, vs .@vs =@xs / can be viewed as
d vs =dt D .@vs =@xs /.dxs =dt / D .@vs =@xs /vs . In this view, the velocity vs D dxs =dt can be
interpreted as the drift velocity of .@vs =@xs /.
5.2 Comprehensive Description of Deformation 107

vp D e j.!f tkxs / (5.54)

Here vp is the amplitude of the transverse wave propagating in the positive xs


direction, as defined in Fig. 5.3, j is the imaginary unit j 2 D 1, and !f is a
complex angular frequency defined as

!f D !0 C j !Q (5.55)

Substitution of Eq. (5.54) into Eq. (5.53) leads to the following characteristic
equation

Gk 2  .!02  !Q 2 /  c !Q  j!0 .2!Q  c / D 0 (5.56)

Equating the real and imaginary parts of the above to zero, respectively, the
following conditions are obtained:
c
!Q D (5.57)
2
s
G 2 2
!0 D k  c2 (5.58)
 4

Thus the general solution in this case can be put in the following form (ignoring the
initial phase).
 1=2 !
 2c t G 2 2
vp D e cos k  c2 t  kxs (5.59)
 4

Similarly, elimination of v leads to the following wave equation for !.

@2 !
 C Gr 2 ! D 0 (5.60)
@t 2
The non-decaying nature of Eq. (5.60) indicates that in the plastic regime, the shear
force is still restoring; consequently, there is no mechanism to decay the rotational
dynamics directly. In the electrodynamic analogy, this is equivalent to the fact that
the magnetic force does not do any work on the charged particles because the force
is perpendicular to the motion of the particles. Note that the !-wave also decays via
field equation (4.62) where v decays according to the above argument.
108 5 Interpretations of Deformation and Fracture Phenomena from Field Theoretical. . .

c
The term e  2 t on the right-hand side of Eq. (5.59) indicates that the displace-
ment wave vp decays with the time constant c D 2=c . From Eq. (5.51),

2 2
c D D (5.61)
c 0 .r  v/

where 0 is a material dependent constant as discussed above. Equation (5.61)


indicates that the greater .r  v/, the shorter the time constant. A greater value of
.r  v/ means a higher strain concentration. Thus, this can be interpreted as that the
higher the initial strain concentration, the faster the displacement wave decays. As
we will see in the section of fracture below, .r  v/ is a figure of merit that indicates
how close a given material under plastic deformation is to its fracture.

Developed Charge

One-dimensional charge-like patterns of the displacement field like the one shown
in Fig. 5.4 are observed in experiments based on a full-field optical interferometric
technique known as the Electronic Speckle-Pattern Interferometry (ESPI) [21].
Details about the ESPI and observations of charge-like patterns will be discussed
in Chaps. 6 and 7, respectively. In short, optical interferometric patterns are formed
in the following fashion. The specimen is illuminated with a pair of interferometric
laser beams sensitive to a certain direction, e.g., the x, y, or z axis of an xyz
coordinate system. A pair of images is taken with a digital camera before and
after the deformation of interest occurs. By subtracting the image captured before
the deformation from the one captured after the deformation, the displacement
occurring during the deformation can be visualized as the so-called fringe pattern.
A fringe pattern consists of bright and dark stripes (fringes). The dark fringes
represent the contour along which the displacement component the interferometer
is sensitive to is an integer multiple of a unit value determined by the configu-
ration of the interferometer.6 Fringe patterns like Fig. 5.4 are formed by in-plane
sensitive interferometer, meaning that the pair of the interferometric laser beams
has sensitivity to the x or y direction in the xy plane. Normally, a fringe pattern
referred to as the developed charge in this book consists of several, approximately-
equidistant, parallel, straight dark-fringes running across the width of the specimen.
If the interferometer is sensitive to the x-component of the displacement, x , as
an example, its x dependence @x =@x can be considered as part of r   D
@x =@xC@y =@yC@z =@z. By dividing the displacement by the duration dt between
the times when the two images are captured, r   can be converted to r  v.
Experiments with an interferometer having two-dimensional sensitivity (sensitive to
both the x and y directions in the xy plane) indicate that the developed charge-like

6
In the case of an interferometer with visible optical frequency, the dark fringes appear for an
increment of the corresponding displacement of typically a few hundreds nm.
5.2 Comprehensive Description of Deformation 109

pattern resulting from the x-sensitive and y-sensitive interferometer are identical to
each other in terms of the location and orientation of the parallel fringes. This means
that the displacement vector, and hence the velocity vector, have spatial dependence
only normal to the interferometric fringes, or on xs only in Fig. 5.3.7
Referring to Fig. 5.3, analyze the displacement field of a developed charge
more quantitatively [22]. Using the coordinate system xp , xs and z, and based on
the assumption that the displacement components have only xs dependence, i.e.,
@=@xp D @=@z D 0, we can express the velocity and rotation components as follows:

@vp @vs @vz @vs


r v D C C D (5.62)
@xp @xs @z @xs
@s @p @p
!z D  D (5.63)
@xp @xs @xs
@p @z
!s D  D0 (5.64)
@z @xp
@z @s @z
!p D  D (5.65)
@xs @z @xs
@!z @!s @2 p
.r  !/p D  D 2 (5.66)
@xs @z @xs
@!p @!z
.r  !/s D  D0 (5.67)
@z @xp

From Eq. (5.44), the equation of motion can be put for the p and s component
separately as follows:

@vp @2 p @vs
 D G 2  Wdp  (5.68)
@t @xs @xs
@vs @vs
 D Wds  (5.69)
@t @xs

Here Wdp and Wds are the p and s components of the drift velocity. The first term on
the right-hand side of Eq. (5.68) represents the p-component of the shear force and
the second term the p-component of the energy dissipative force. Equation (5.69)
indicates that the s-component of the shear force is null, meaning that when a
developed charge is formed the material loses restoring force normal to the parallel
fringes representing the charge.

7
The two-dimensional interferometric results do not guaranty that the displacement vector does
not depend on z. However, since we are making two-dimensional analysis in the xy plane, the z
dependence is not significant.
110 5 Interpretations of Deformation and Fracture Phenomena from Field Theoretical. . .

The argument that the material loses shear force as a developed charge is formed
along with the experimental observations [11, 23, 24] that those developed charges
are formed instantaneously indicates that the stress drops known as the serrations
can be explained as a consequence of the loss of shear resistance associated with the
formation of a developed charge. The serrations [25] are a zigzag pattern of stress–
strain characteristics that accompanies the appearance of a shear-band referred to as
the Portevin-Le Chatelier (PLC) band [26]. In each cycle of the zigzag pattern, an
abrupt stress drop is followed by a slower stress recovery. In the present context, the
pattern of parallel-fringes represents the PLC band and the material loses the shear
resistance normal to the PLC band. Further, an experiment [23] indicates that the
formation of a developed charge-like pattern is accompanied by acoustic emission
originating the location where the pattern appears. This observation indicates that
when a developed charge is formed a fracture of small scale occurs and that causes
the loss of the shear resistance. It is possible that the stress drop and the formation
of a developed charge are observed coincidentally when a dislocation flow reaches
one side of the specimen.8
Figure 5.5 shows a series of developed-charge like fringe patterns traveling from
the lower end of a specimen to the other end along with the stress recorded at the
same time as the fringe patterns. Notice that during the period when the charge-
like pattern completes the travel, the stress remain at the same level except for the
small ripples considered to be serrations. Towards the end of the travel, those ripples
are not as conspicuous as the beginning. It is not clear at this time whether the
stress drops become less conspicuous towards the end because the stress drop is
smaller when the corresponding developed charge is formed or the same charge
glides without being accompanied by a stress drop.
Equation (5.61) indicates that the decay time constant of displacement wave
depends on (r  v); the greater (r  v), the shorter the time constant. In the one-
dimensional case as in Fig. 5.4, this means that the greater the slope @vs =@xs , the
faster the wave decays, i.e., the smaller the region in which the developed charge
is concentrated, the faster the wave decays. Later in Chap. 7, we will consider
the decay constant of a transverse displacement wave, based on (r  v) for two
cases as examples; the first case is where a transverse wave is observed to decay
until a developed charge appears. There is no visible, developed charge until the
wave decays. In this case, the charge @vs =@xs can be assumed distributed over
the entire specimen, and therefore the density (r  v) is low. The second case
is where developed charges appear intermittently. Every time a developed charge
appears, the wave decays. In this case the developed charge is observed in a narrow
region, typically a few mm or less in width. In both cases, the decay constant of
the displacement wave estimated based on experimentally determined 0 shows
semiquantitative agreement with experiment. For more details about this discussion,
see Sect. 7.1.

8
As will be discussed in detail in Chap. 7, a number of experiments indicate that the developed
charge represents the PLC band.
5.2 Comprehensive Description of Deformation 111

77 100 130 163 190 218 250 280 336

350
300
77 336
250
stress (MPa)

200
150
100
50
0
0 0.01 0.02 0.03 0.04 0.05 0.06
strain

Fig. 5.5 Developed, one-dimensional charge-like pattern traveling from the bottom to top of
specimen and stress

Transition from Elastic to Plastic Regime

In the preceding section, we discussed the equivalence of the drift velocity of


the deformation charge, W d , and the phase velocity of the elastic wave, c elas . In the
plastic regime, the mechanical energy is dissipated when a deformation charge
drifts. In the elastic regime, an elastic wave carries strain energy. The equivalence
of W d and c elas indicates that the transition from the elastic regime to plastic
regime can be argued based on a change in the form of energy flow. In the elastic
regime, the mechanical energy is in the form of spring energy. When the dynamics
is longitudinal-wave like, the energy is carried from one point to another in the
form of compressive/stretching energy. At one moment the material has compressive
elastic energy at a certain point of the specimen; a half cycle later, the same point
will have stretching energy. When the dynamics is transverse-wave like, the energy
is carried in the form of torsional energy. These are conservative energy. When
the deformation develops to the plastic regime and the energy dissipative effect
represented by Gj D Wd .r  v/ dominates over the energy conservative shear
restoring force G.r !), the material loses the mechanism of carrying conservative
112 5 Interpretations of Deformation and Fracture Phenomena from Field Theoretical. . .

energy in the form of a wave, and the energy provided by the external agent starts
to be dissipated via the flow of .r  v/. This process is naturally understood as
the reduction in the shear modulus due to the increase in the dislocation density or
other defects. As much as the elastic energy decreases with the reduction of the shear
modulus, the system loses energy via the energy dissipative force Gj D W d .rv/.
It is interesting to note that the above-mentioned transition from elastic deforma-
tion to plastic deformation is similar to the transition from a glow discharge to an
arc discharge in gaseous electronics. When a discharge is in the glow regime, the
electric energy provided by the external circuit is deposited to excited states of
the gas species (atoms or molecules) via collisions of the electrons accelerated by
the external electric power with the atoms or molecules [27]. If the excited state is an
electronic state, the bound electron orbiting the nucleus of the atom changes its orbit
to one with a higher energy level. The electron is still bound to the original atom.
Some of the collisions result in ionization. In this case, the electrons are released
from the bound state to be free electrons. Ionization also generates ions, which are
also free charges. These free electrons drift under the influence of the electric field
due to the external circuit, e.g., the voltage applied to a fluorescent light bulb. The
resultant charge flow is referred to as the conduction current.
The transition from elastic to plastic deformation can be argued in a similar
fashion. The irreversible deformation resulting from the generation of defects such
as dislocations corresponds to ionization. The process of ionization is irreversible
in the sense that free charges generated by ionization are not bound by the original
nucleus.9 The quantity .r  v/ corresponds to free charges, and Gj D W d .r  v/
to the conduction current. From Eqs. (5.49) and (5.51),

W d .r  v/ D c v (5.70)

The existence of free charges .r  v/ makes the material conductive, and the degree
of conductance is represented by the conductivity c . The energy dissipation due to
Gj D W d .r  v/ corresponds to the Ohmic loss in the electric case represented
by the conduction current in the form of the product of the electric conductivity e
and the electric field E ; je D e E . The higher the value of .r  v/, the more the
medium becomes energy dissipative.

9
These free charges can be recombined with an atom or molecule though the process known as
recombination. In this context, however, the effect of ionization is that it raises the conductivity.
From this standpoint, unless the recombination rate exceeds the ionization rate, which is unlikely
in the normal circumstance, the ionization can be considered as an irreversible process.
5.2 Comprehensive Description of Deformation 113

5.2.3 Fracture

Transition from a late stage of plastic deformation to fracture can be argued based on
the plastic deformation charge .r v/. Moreover, the analogy of plastic deformation
charge to the electric free charge discussed in the preceding section seems to work
here as well. In this case, the transition from the plastic, deformation stage to the
pre-fracture stage is analogous to the transition from the glow discharge to the arc
discharge in gaseous electronics, and the transition from the pre-fracture to the
fracture stage is analogous to the transition from the arc discharge to the electric
breakdown of a gas medium. Before starting to argue fracture, take a moment and
consider the gaseous electronics case. As the external energy (voltage) is increased,
so is the rate of ionization, creating more free charges. In the early stage of the
increase in ionization, the free charges are still uniformly distributed in the gas,
contributing to the atomic excitation via collisions with lower state atoms. With
the increase of free charges, the gas medium becomes more conductive. Once the
level of free charges resulting from ionization reaches a certain level, the free
charges tend to be distributed unevenly. This raises the local conductivity at a certain
location of the gas medium much higher than the rest of the medium. Once the
conductivity is raised in a local region, the conduction current flowing through
that region increases. Consequently, the more energy from the external circuit is
input to the local region, further increasing the local conductivity. This triggers the
positive feedback mechanism where the conduction current through the local region
increases in a catastrophic manner, leading to the final arc discharge short-circuiting
the gas, or the phenomenon known as the electric breakdown of the gas medium.
The deformation version of the transition is as follows. After Gj D W d .r  v/
becomes the dominant mechanism of energy flow, the charge density .r  v/
tends to keep increasing at a certain location. The reason is as follows. The
increase in W d .r  v/ as the material’s resiting mechanism is accompanied by the
corresponding decrease in G.r !/ as the other resisting mechanism. The decrease
in G.r  !/ is caused by reduction in the shear modulus G as the increase in
.r  v/ at a specific location enhances non-uniformity in deformation. The increase
in r  v can be interpreted as increase in strain. Similar to the electric case where
the localization of free-charge concentration is enhanced by the positive feedback
mechanism in a late stage of glow discharge, r  v becomes more localized in a
late stage of plastic deformation or the pre-fracture stage. This is the well-known
phenomenon referred to as strain localization. Eventually, r  v becomes infinity.
r  v > 0 represents the rate of particles flowing out of a volume. So if it becomes
infinite, the volume becomes empty. This is interpreted as the fracture as the final
stage of plastic deformation.
In this process, the external agent, the tensile machine applying a load to the
specimen for example, continues providing energy. The material dissipates this
energy via Gj D W d .r  v/. In other words, W d .r  v/ has a finite value. If
.r  v/ approaching infinity, the drift velocity W d approaches zero. This leads to
the hypothesis that the plastic deformation charge must stop flowing as the material
114 5 Interpretations of Deformation and Fracture Phenomena from Field Theoretical. . .

fractures. A number of ESPI-based experiments [10, 11, 13, 24] indicate that the
developed, one-dimensional charge-like pattern stops moving prior to the fracture
of the specimen, supporting this hypothesis.
From the viewpoint of wave dynamics, the transition to fracture can be argued
as follows.
p As the shear modulus G decreases, the phase velocity of the shear
wave G= decreases. The phase velocity is the product of the frequency and
the wavelength, where the latter is determined by the geometrical factor such as
the length of the object. Thus, in most cases, the wavelength remains the same
until the fracture. So, at the point when the shear modulus decreases to zero, so
does the phase velocity and the frequency. In an oscillatory system, zero frequency
means the displacement is one-way. In the present case, this corresponds to the
situation where the displacement associated with the shear deformation becomes
monotonic. As the shear deformation continues in such a monotonic fashion, the
object eventually breaks along the line of the monotonic displacement.
There are experimental observations that support the above argument regarding
transitions from the elastic to plastic, and the plastic to fracture stage in connection
with the wave decay and charge’s drift velocity. These will be discussed later in
Chap. 7.

5.3 Physical Meaning of Potentials

5.3.1 Vector Potential and Scalar Potential from Gauge

Equations (4.33) and (4.34) can be put in the following form:

@A
vD C r 0 (5.71)
@t
! D r A (5.72)

From the original meaning of gauge as a compensation term of differentials and the
fact that vector potential A is the spatial part of it indicates that A is the differential
displacement at the boundary of neighboring deformation structural elements. The
relationship between the gauge and the vector potential is not straightforward, but
it can be understood as follows. We describe deformation as transformation that a
local line element vector  experiences to become 0 [Eqs. (4.17) and (4.18)].

0 D U  D .I C ˇ/ (5.73)

When we calculate the total differential of a displacement vector, because of the


spatial dependence of the transformation matrix [the displacement gradient tensor,
Eq. (4.19)]
5.3 Physical Meaning of Potentials 115

0 1
@x @x @x
B @x @y @z C
B C
B C
@i B @y @y @y C
ˇD DB
B @x
C (4.19)
@xj B @y @z C C
B C
@ @z @z @z A
@x @y @z

It is obvious that the differentiation of a vector is not transformed in the same way
as the vector.

U.d / D d.U / D .d U // C U.d / (5.74)

Equation (5.74) indicates that when the transformation matrix is dependent on the
spatial coordinates, the physics law defined by Eq. (5.73) cannot be written in the
same fashion before and after the transformation. It becomes necessary to replace
the usual partial derivative with a covariant derivative by adding the gauge term.

@
Di D D @i  i ; i D x; y; z (3.18)
@x i
Thus the total differential of the displacement becomes,
     
@i @i @i
Di D  x i dx C  y i dy C  z i d z  d i  Ai (4.21)
@x @y @z

Here i is the gauge associated with differentiation with respect to the coordinate
axis xi , and Ai is the corresponding component of the vector potential. The above
mathematical operation indicates that
When the displacement tensor hence the deformation tensor depends on the coordinate
points, the deformation is nonlinear because the first-order derivative is not constant.
Therefore, the law of linear elastic deformation does not apply. However, by replacing the
usual derivatives with covariant derivatives by adding the above gauge terms, it is possible
to use the law of linear elasticity to describe the nonlinear situation. From the physical
viewpoint, the gauge term generates the additional displacement represented by the vector
potential A. In other words, we can cram the portion of the displacement that the linear
elastic law cannot handle into the vector potential. Figure 5.6 illustrates such a displacement
schematically. As much as the deformation structural element rotates as a rigid body in
response to external force, the differential displacement does not contribute to the elastic
force prescribed by the law of linear elasticity.

Note that this vector potential is due to the coordinate dependence of the transfor-
mation matrix, different from the usual vector potential in elasticity that leads to
shear elastic wave [28]. The usual vector potential is due to the way external force
is applied to the material. This vector potential causes rotation even if the external
force exerts no torque. In other words, part of the material rotates due to a self-
induced cause such as non-uniformity of the elastic-modulus induced by generation
of defects.
116 5 Interpretations of Deformation and Fracture Phenomena from Field Theoretical. . .

Ax

z wy
dz
y x

xj xj
?
xi xi

Fig. 5.6 Schematic illustration of vector potential A as a displacement vector representing rigid
body rotation of a deformation structural element. By removing rotation, all deformation structural
elements can be differentiated on a common basis

The other potential 0 can be argued as the temporal version of the gauge term.
As is the case of the spatial version, the temporal component of the gauge is
introduced as the additional term to the usual temporal derivative. This temporal
part of gauge concept may appear less intuitive than the spatial part, but it
can be understood as follows. Linear elasticity states force is proportional to
spatial differentiation of displacement. According to Newton’s second law, force
is proportional to the temporal derivative of velocity. Therefore, it can be said that
linear elasticity connects the spatial differentiation of displacement and temporal
differentiation of velocity. If we modify the spatial differentiation by adding the
gauge term to the spatial differentiation to make linear elasticity locally symmetric,
we need to modify the temporal part as well so that spatial-temporal connection is
preserved. When we express the gauge effect by potential, the spatial effect is the
vector potential and the temporal effect is the scalar potential. If we express the
gauge effect by field variables, v and !, these spatial and temporal potential effects
are connected on the right-hand side of Eq. (5.71). It is interesting to note that in
Eq. (5.71), they are connected in the form of temporal differentiation of the spatial
potential and spatial differentiation of the temporal potential.
The above-mentioned connection between the temporal and spatial part of
potential can be argued in connection with the gauge invariance in the deformation
field. Consider that the vector potential A is changed by gradient of function  as

A ! A 0 D A C r (5.75)

With the mathematical identity r  r D 0, this change does not affect ! in


the field equations. As for the other field variable v, Eq. (5.71) indicates that if the
potential 0 changes as
5.3 Physical Meaning of Potentials 117

0 @
0 ! 0 D 0  (5.76)
@t

v remains the same as well. Apparently, if 0 experiences the change (5.76), neither
v or ! is affected as long as A changes according to Eq. (5.75). In other words,
the above pair of changes in the spatial and temporal part of the potential does not
affect the field equations, meaning that the dynamics is unaffected. The connection
between the temporal and spatial potential has such a synergetic nature. Also note
that the fact that the changes in the potentials do not affect the dynamics indicates
that the potentials are not physical quantities but rather artificial parameters.
It is worth while considering the spatial and temporal differentiation and their
connection from the viewpoint of dynamics. The equation of motion under linear
elasticity is in the form of “(mass)  (temporal differentiation of velocity) D
(stiffness)  (spatial differentiation of displacement).” Here the mass describes the
inertia or the difficulty in changing the velocity over time whereas the stiffness
describes the difficulty in changing the displacement over space. The two changes
are connected by the equation of motion through the force. The greater the mass or
the greater the stiffness, the greater the force. Importantly, both mass and stiffness
are intrinsic to the material. In other words, once we pick a material, we do not have
control over the relation between the temporal and spatial differentiation. Nature
determines how the material behaves based on the relation. For example, the natural
frequency of a p spring-mass system is solely dependent on the spring constant k and
the mass m as .k=m/. It does not depend on how the mass is oscillated.
Similarly, when oscillation is transferred from one part to another in an elastic
medium, the phase velocity is intrinsic to the elastic medium in the form of
the square root of “stiffness/density.” The phase velocity does not depend on the
frequency or the wavelength of the medium. Equation (5.44) states that when we
allow coordinate-dependence of the transformation describing linear elasticity, the
connection of the spatial and temporal differentiation in the material is represented
by the phase velocity in the form of the square root of “the shear modulus/density.”
The spatial and temporal terms of the gauge portion j0 and j are also connected by
a phase velocity. As Eq. (4.41) indicates, the temporal and spatial gauge potential
terms of the Lagrangian interact with the material field via j0 =c and j i . Here c can
be viewed as the phase velocity of these charges in the sense that it connects the
temporal and spatial gauge effects. In Eq. (5.45), j is connected with j0 D .r  v/
via the drift velocity W d as Gj D W d .r  v/, and in Eq. (5.48) we observe
that in the elastic limit the drift velocity reduces to the elastic phase velocity. It is
not coincident that Wd reduces to the phase velocity. It reflects that phase-velocity
nature of the connection between j0 and j .
Thus, from the gauge theoretical point of view, the scalar and vector potentials are
understood in terms of the temporal and spatial differentiation. The relation between
temporal and spatial effects is more intuitively understood in conjunction with
wave dynamics. In this context, it is convenient to express the temporal and spatial
components of the derivatives and potentials comprehensively using the relativistic,
four-vector notation. Here the relativistic notation is useful not because we deal
118 5 Interpretations of Deformation and Fracture Phenomena from Field Theoretical. . .

with relativistically high velocity but we want to connect the temporal and spatial
components of the related quantities. Wave dynamics deals with both components
comprehensively. We discuss this as a new section next.

5.3.2 Scalar and Vector Potential from Viewpoint of Wave


Dynamics

In the context of wave dynamics, the connection between the temporal and spatial
dependence can be understood as follows. When boundary conditions are provided,
the above-argued phase velocity connects the temporal and spatial periodicities
as “phase velocity = wavelength/period.” The temporal boundary condition can be
given as the driving frequency of the external force (input disturbance) and the
spatial boundary condition can be determined by the dimension (size) of the actual
object. The point here is that as a property intrinsic to the material, the phase
velocity is unchanged if the driving frequency or the object size is changed. When
the input disturbance is at a certain frequency fi n , the wavelength of the wave
traveling through the medium is determined as cph =fi n . If the input disturbance is
an impulse (i.e., containing all frequency components or having unlimited choices
of temporal periodicity) from one end of the object having a length of L, the
dominant wave is such that the wavelength of the fundamental wave is either 2L
or L depending on the constraint on the other end, and the frequency is determined
as “frequency = phase velocity/wavelength.” This is how the temporal and spatial
variations are connected with each other, and why we can argue the temporal and
spatial periodicity interchangeably through the phase velocity.
The above concept can be conveniently described with the relativistic four vector
notation, which explicitly expresses the time component in parallel to the space
component via the phase velocity. Before discussing potentials 0 and A, take a
moment and express an elastic wave equation with the four vector notation. Recall
that elastic wave equations can be put in general in the following form:

@2 
2
D r2 D cph
2
r2 (5.77)
@t 

Here is the wave function,  is the density of the material,  denotes the elastic
modulus, and the ratio = is the square of the phase velocity cph . Table 5.1 lists
some typical cases. Recall the four-vector derivatives Eq. (4.32) discussed in the
preceding chapter.
   
@ 1 @ @ @ @
@ D ; r D ;  ;  ; 
@x 0 c @t @x 1 @x 2 @x 3
   
@ 1 @ @ @ @
@ D ; r D ; ; ; (4.32)
@x 0 c @t @x 1 @x 2 @x 3
5.3 Physical Meaning of Potentials 119

Table 5.1 Some typical K Wave type


elastic waves
E  Longitudinal wave
G  Shear wave
C 2G r  Compression wave
G r  Rotation wave

With Eq. (4.32), wave equation (5.77) can be put in the following form:

 1 @ 2
.@ / @ D 2  r2 D0 (5.78)
cph @t 2

The four vector notation Eq. (5.78) literally represents the above-discussed inter-
changeability of temporal and spatial differentiation with the factor of phase velocity
2
via cph appearing in the denominator of the temporal derivative.
Now extend the idea to the potential due to the gauge terms. Equation (4.31)
expresses the scalar and vector potential in the four vector notation.
 
0 1 2 3 
A D ;A ;A ;A D A0 ; A1 ; A2 ; A3 (4.31)
c

The phase velocity appearing in the denominator of the scalar potential represents
how fast the effect of the additional displacement A propagates through the medium.
In the context of electrodynamics, this is nothing but the delay vector potential;
the electromagnetic oscillation travels at the speed of light. A charged particle
notices the fact that another charged particle at a far distance is oscillating by
receiving the news that travels at the speed of light. In the present context, the
signal associated with the coordinate dependence of the transformation matrix (the
news that the dynamics has an extra element because of the coordinate dependence
of the transformation matrix) propagates at this phase velocity. As the potential
representing the temporal version of the gauge effect, the scalar potential has the
same factor of phase velocity in the four vector notation as the time component of
the four-vector derivatives.
Since the vector potential A has a unit of length (m), the scalar potential 0 is
in .m  m=s/. Notice that if we multiply the charge q whose volume density is the
charge density .r  v/ in .kg=.m3  s// [and therefore q in (kg/s)], the resultant
quantity has a unit of energy; ..kg  m=s2 /  m/ D .N  m/ D .J/. If we multiply
the same quantity q in (kg/s) to the other potential A in (m), the resultant quantity
has a unit of momentum; .kg  m=s/. In this view, we can interpret the scalar and
vector potential in connection with potential energy and kinetic energy (momentum)
caused by the quantity q in (kg/s). From the field theoretical viewpoint, this quantity
can be interpreted as the source of the additional dynamics due to the coordinate
dependence of the transformation matrix, or the charge. This leads to the concept
of the field force associated with the potential and kinetic energy. In the four vector
120 5 Interpretations of Deformation and Fracture Phenomena from Field Theoretical. . .

notation, the scalar and vector potential multiplied by the charge constitutes a four
momentum. We will discuss the field force in the next section.
 0 

q ; A1 ; A2 ; A3 D p 0 ; p 1 ; p 2 ; p 3 (5.79)
c

5.3.3 Field Forces

Field forces10 are most conveniently discussed with Hamiltonian. The above-
mentioned force associated with the scalar potential 0 may not sound like a field
force, if the field force is defined as the non-contact force. However, from the
viewpoint that this force is proportional to the field variable v, as will be discussed
soon, it can be treated as a field force. There is no truly a contact force at the
atomistic level as no material entity has contacts with another entity.
Based on the similarity to electrodynamics, we start the argument by putting the
Hamiltonian in the same form as that of charged particles, and see if we can discuss
the resultant force and related physics reasonably. To this end, let’s quickly review
Hamiltonian of a charged particle. The Hamiltonian of an electric charge q in an
electromagnetic field represented by scalar potential and vector potential A can
be given by the following equation [29, 30]:

1
H D .p  qe A/2 C qe (5.80)
2m
With generalized coordinates qi and momentum pi , Hamilton’s equations of motion
are [31]
@H
qPi D (5.81)
@pi
@H
pPi D  (5.82)
@qi
From the first Hamiltonian equation of motion Eq. (5.81), we obtain
d xi 1
D vi D .pi  qe Ai / (5.83)
dt m
which leads to

10
The content of this section does not introduce a new concept to the present theory. It is a different
view of the dynamics discussed in preceding sections in terms of the potential associated with
the gauge and energy dissipation due to plastic deformation. It helps us understand the effect of
plastic deformation as the synergetic process between the translational and rotational mode of
displacement.
5.3 Physical Meaning of Potentials 121

pi D mvi C qe Ai (5.84)

Here vi is the velocity (the i th component) of the charge. Equation (5.84) indicates
that the momentum of a charged particle has the electromagnetic momentum in
addition to the product of its mass and velocity. By differentiating Eq. (5.84) with
respect to time,

dpi d vi dAi
Dm C qe
dt dt dt
 
d vi @Ai @Ai dx1 @Ai dx2 @Ai dx3
Dm C qe C C C
dt @t @x1 dt @x2 dt @x3 dt
 
d vi @Ai @Ai @Ai @Ai
Dm C qe C v1 C v2 C v3 (5.85)
dt @t @x1 @x2 @x3

On the other hand, from the other Hamiltonian’s equation of motion Eq. (5.82), we
obtain

dpi qe @A @
D .p  qe A/   qe (5.86)
dt m @xi @xi

From Eq. (5.84),

p  qe A
Dv (5.87)
m
Substituting Eqs. (5.87), (5.86) becomes

dpi @A @
D qe v   qe (5.88)
dt @xi @xi

Equating the right-hand side of Eqs. (5.85) and (5.88), we obtain


 
d vi @Ai @Ai @Ai @Ai @A @
m C qe C v1 C v2 C v3 D qe  v  qe (5.89)
dt @t @x1 @x2 @x3 @xi @xi

Here the order of the inner product on the right-hand side is switched for better
@A
visibility. Noting that @x i
has the term @Ai
@xi
and canceling the corresponding term on
the left-hand side, we can rewrite this equation in the following form:
 
d vi @Ai @ @Ai @Ai @Ai @A
m D qe  qe  qe vi C vj C vk C qe v
dt @t @xi @xi @xj @xk @xi
     
@Ai @ @Aj @Ai @Ak @Ai
D qe C C qe vj  C qe vk 
@t @xi @xi @xj @xi @xk
122 5 Interpretations of Deformation and Fracture Phenomena from Field Theoretical. . .

 
@Ai @
D qe C C qe vj .r  A/k  qe vk .r  A/j
@t @xi
 
@Ai @
D qe C C qe .v  .r  A//i
@t @xi
D qe Ei C qe .v  B/i (5.90)

With the vector notation, this becomes

dv
m D qe E C qe v  B (5.91)
dt
Equation (5.91) is the famous Lorenz force acting on the charge qe [29, 32].
Now consider the case of deformation dynamics. The Hamiltonian is given as
follows:
1
H D .p C qm A/2  qm 0 (5.92)
2m

where qm is the deformation charge whose density is .j0 / D .r  v/ and the
mass is m, i.e., qm D m.r  v/, p is the momentum of qm , and A and 0 are the
potentials of the deformation field. As previously discussed, .j0 / corresponds to
the electric positive charge [see the two paragraphs following Eq. (4.92) in Chap. 4].
Note that the sign in front of qm in Eq. (5.92) is opposite to that in front of qe in
Eq. (5.80). This difference comes from the sign on the right-hand side of Eq. (5.71)
from the electromagnetic version as shown below.

@A
vD C r 0 (5.71)
@t
@A e
E D  r e (5.93)
@t
This difference in the sign between the deformation field and electromagnetic field
is not physically significant. It simply comes from the difference in the way the
potentials are defined. In the electromagnetic case, the scalar potential is defined
as such that the electric field is opposite to the slope (gradient) of the potential;
in a potential field, a positive charge moves down the slope to get accelerated. To
the contrary, according to Eq. (5.71) the potential 0 for the deformation charge
is defined such that the v filed is the same sign as the gradient of the potential;
in a potential field, a positive charge qm moves up the slope. Consequently, the
material containing the positive charge loses its momentum. This situation is clearer
if we derive an equation corresponding to Eq. (5.84) as follows. From the first
Hamiltonian equation of motion Eq. (5.81),

d xi 1
D wi D .pi C qm Ai / (5.94)
dt m
5.3 Physical Meaning of Potentials 123

and therefore,

pi D mwi  qm Ai (5.95)

Here wi is the i th component of the velocity of the charge qm . Equation (5.95)


indicates that the total momentum of the piece of the material containing the positive
charge qm is its mass times velocity minus the momentum loss due to the upslope
motion of the charge in the potential. By differentiating Eq. (5.95) with respect to
time,

dpi d wi dAi
Dm  qm
dt dt dt
 
d wi @Ai @Ai @Ai @Ai
Dm  qm C w1 C w2 C w3
dt @t @x1 @x2 @x3
 
d wi @Ai @Ai @Ai @Ai
Dm  qm C wi C wj C wk (5.96)
dt @t @xi @xj @xk

From the other Hamiltonian’s equation of motion Eq. (5.82),

dpi qm @A @
D .p C qm A/  C qm (5.97)
dt m @xi @xi

Substituting .p C qm A/ in the form of Eq. (5.95)

dpi @A @
D qm w  C qm
dt @xi @xi
 
@A1 @A2 @A3 @
D qm w1 C w2 C w3 C qm
@xi @xi @xi @xi
 
@Ai @Aj @Ak @
D qm wi C wj C wk C qm (5.98)
@xi @xi @xi @xi

Equating the right-hand side of Eqs. (5.96) and (5.98), we obtain


 
d wi @Ai @Ai @Ai @Ai
m  qm C wi C wj C wk
dt @t @xi @xj @xk
 
@Ai @Aj @Ak @
D qm wi C wj C wk C qm (5.99)
@xi @xi @xi @xi

Eliminating common terms from both-hand sides and rearranging the orders of the
terms, this equation becomes
124 5 Interpretations of Deformation and Fracture Phenomena from Field Theoretical. . .

      
d wi @Ai @ @Ai @Aj @Ai @Ak
m D qm C C qm wj  C wk 
dt @t @xi @xj @xi @xk @xi
 
@Ai @ 
D qm C C qm wj .r  A/k C wk .r  A/j
@t @xi
 
@Ai @
D qm C  qm .w  .r  A//i
@t @xi
D qm vi  qm .w  !/i (5.100)

Thus, the force on charge qm , F q is identified as

F q D qm v  qm w  ! (5.101)

where v is the velocity of a point in the object under deformation, ! is the


material rotation around the point, and w is the velocity of charge qm D m.r  v/.
Equation (5.101) indicates that when the v field has divergence, the field exerts force
on that diverging part. As is the case of electrodynamics, the field exerts two kind
of forces; the first on the charge itself and the second on moving charges.

F 1 D qm v D m.r  v/v (5.102)


F 2 D qm .w  !/ D m.r  v/w  ! (5.103)

The first force F 1 D qm v corresponds to the electric force. Being proportional


to velocity v, this force is essentially energy dissipative (a velocity-damping force).
The only exception is when this term happens to be proportional to the displacement
in the elastic limit as has been discussed earlier in this chapter. In this case, this
force becomes centripetal. The second force F 2 D qm w  ! corresponds to the
magnetic force. As is the case of magnetic force, this force does not do work because
it is perpendicular to the flow of qm .
Let’s consider the physical meaning of these forces from the gauge theoretical
point of view. These forces originate from the potential associated with the
introduction of the gauge term in the covariant derivatives. The gauge terms are
necessary to make the transformation matrix representing linear elasticity locally
symmetric. In other words, when we try to describe curvilinear dynamics with the
linear elastic theory, the points of the object under deformation interact with the
gauge field. F q is the force representation of this interaction, and the deformation
charge is the medium that connects the material and the gauge field. Thus, it can be
said that F q is the force that describes the interconnection among local deformation
structural elements within which linear elasticity holds. The permanent deformation
in the plastic regime is due to the energy dissipative nature of the first force F 1
acting on the deformation charge.
5.3 Physical Meaning of Potentials 125

From the material’s perspective, the deformation charges behave like force as
well. According to Eq. (4.41), the potential energy associated with the gauge field is

G 0
U D j A0 C Gj i Ai (5.104)
c
If we evaluate field force Fm by the usual prescription of gradient of potential
energy, rU , using the potential energy in the form of Eq. (5.104) and the four
potential
 
0
A D ; A1 ; A2 ; A3 D .A0 ; A1 ; A2 ; A3 / (4.31) (5.105)
c

as the independent variable, we obtain


 
j0 1 2 3
Fm D G ;j ;j ;j D Gj (5.106)
c

The spatial components of this vector G.j 1 ; j 2 ; j 3 / is the force density vector
discussed in the preceding chapter. It is important to note that G.j 1 ; j 2 ; j 3 / is the
energy dissipative flow of charge qm and that therefore the force density acting on
the material is fundamentally energy dissipative.
The interrelation between j 0 and .j 1 ; j 2 ; j 3 / has a subtle and interesting aspect.
It represents Lenz’s law, or the nature’s opposing mechanism to suppress a change
caused by an external agent. This effect is conveniently understood via analogy
to Faraday’s law in electrodynamics. So, first consider the electrodynamic case.
Consider in Fig. 5.7 a ring conductor and an external magnetic field going into
the plane of the ring perpendicularly. Suppose that this external magnetic field
increases with time. According to Faraday’s law, an electric field is induced along
the conductor in the counterclockwise direction. Consequently, free charges in the
conductor flows in the counterclockwise direction. This induced current generates
a magnetic field perpendicular to the plane of the ring in the direction coming out
of the plane so that the increase in the external magnetic field into the plane is

Iind Fmag

Fig. 5.7 Faraday’s law on an


increasing external magnetic
and a ring conductor. The
solid circle represents the
external magnetic field
increasing into the plane of
the ring. The dashed circles
represent the induced
magnetic field coming out of
the plane
126 5 Interpretations of Deformation and Fracture Phenomena from Field Theoretical. . .

F2 = −qmW × w

j = Wd r (∇ ⋅ n)

Fig. 5.8 Lenz’s law on a deformation structural element experiencing increasing clockwise
rotation due to external force. The solid circle represents the rotation due to the external force. The
dashed circles represent the induced rotation expressed by the field equation (4.63). The second
field force F2 pushes the boundary inward so that the torque due to the external force is reduced

suppressed. On the other hand, the induced current feels inward magnetic force
exerted by the external magnetic field. This magnetic force decreases the ring
diameter so that the flux (the effect of the increasing magnetic field) is reduced.
Thus, the Lenz’s law can be argued in two ways; first, the induced current opposes
the increase of the external magnetic field through the induction of the opposing
magnetic field. Second, the magnetic force on the induced current reduces the area
enclosed by the conductor so that the flux of the increasing external magnetic field
is reduced.
The deformation dynamics version of the above effect can be explained as
follows. Consider a deformation structural element experiences an increase in
clockwise rotation due to an external agent in Fig. 5.8. According to the field
equation (4.62), the velocity along the boundary is clockwise. Consequently,
positive free charges qm existing near the boundary of the deformation structural
element flow in the clockwise direction. According to the field equation (4.63),
this flow of charges induces rotation (r  !) in the direction that suppresses
the initial clockwise rotation due to the external agent. This flow of charges feel
vector F2 according to Eq. (5.103) toward the center of the deformation structural
element. Making the radius shorter, this force reduces the torque due to the external
agent,11 effectively diminishing the increase in the clockwise rotation. Note that
this is the energy-dissipative version of Lenz’s law for deformation dynamics, as
opposed to the energy-conservative Lenz’s law based on the restoring rotational
force dues to shear force discussed in the preceding chapter (see Fig. 4.7 and
associated explanation).

11
Torque is the vector product of the arm length l and force F ; tau D l  bmf . The decrease of
the radius corresponds to a decrease in l .
5.4 Physical Meaning of Charges 127

5.4 Physical Meaning of Charges

In the description of the gauge theoretical dynamics of deformation discussed


in the preceding section, the deformation charge was addressed as somewhat
abstract quantity associated with the compensation effect of the dynamics. The field
equations indicate that the deformation charge is closely related to the temporal part
of the charge of symmetry associated with the local symmetry of linear elasticity.
In this section, we explore the physical meaning of the charges of symmetry in
conjunction with the compensation effect of the deformation dynamics.
The deformation charge .j0 / and its flow Gj D W d .r  v/ D W d .j0 /
[see Eq. (5.45)] are introduced to the dynamics as the source terms to the field
equations as

r  v D j0 (5.107)
@v
Gr  ! D   Gj (5.108)
@t
p
Here Eq. (5.108) is derived by substituting the phase velocity c D G= into
the original field equation (4.63), and Gj has been identified as the density of
the energy dissipative, longitudinal force acting on a unit volume of a material
under plastic deformation. The quantities j 0 and j are originally introduced as
the temporal and spatial parts of the charge of symmetry associated with the local
symmetry of linear elasticity. The underlying physical meaning of all these are
understood as follows. We started off the theory with the postulate that even in
the plastic or fracturing regime of deformation, locally the material has elasticity.
This justifies to describe the local deformation with the law of linear elasticity. We
refer to those regions where linear elasticity is locally applicable as the deformation
structural element. To describe the nonlinear behavior of deformation in the plastic
and fracturing regimes, we allow coordinate dependence on the transformation that
describes linear elastic deformation, and request that the transformation is locally
symmetric. This necessitates us to replace the usual derivatives with covariant
derivatives, meaning that the differentials are transformed in the same fashion as the
vector itself. The gauge is the additional term in the covariant derivatives, and its
function is to connect the deformation structural elements so that the dynamics can
be described on the global level. From the dynamics point of view, this connection is
equivalent to introduce a potential field that accounts for the dynamics that the local
linear elastic theory cannot cover. From this standpoint, the potential field is referred
to as a compensation field to preserve the local symmetry of linear elasticity. v and
! are the vector representation (more precisely the vector field variables) of the
potential field. Based on this, we derive the Lagrangian density (4.41). The actual
interaction of this compensation field and the material under plastic deformation is
the field force discussed in the preceding section. The deformation charge is the
physical entity that the field force acts on. From the gauge theoretical viewpoint, the
charge of symmetry can be interpreted as the Noether’s current discussed in Chap. 4.
128 5 Interpretations of Deformation and Fracture Phenomena from Field Theoretical. . .

The field force acts on the plastically deformed material of mass m as m0 .rv/v,
as discussed in the preceding section. Here the parameter 0 measures the degree
of the plasticity. Per unit volume, this can be put as Gj D v0 .r  v/. In other
words, when the deformation of a material is developed and part of the material
starts to experience irreversible deformation, it loses mechanical energy via the
energy-dissipative force density Gj acting on the material as the first field force
F1 . Under the influence of F1 , the deformation charge flows. If the charge flows
in a displacement field where rotation ! changes over time, the second field force
acts on the charge flow so that the temporal change of ! is reduced. This is based
on the beauty of nature, the opposing mechanism that prevents further deformation.
This self-organized mechanism can be understood as follows. When deformation
develops to the level where m.r  v/ appears, the displacement field of the material
becomes rotation-like because those regions where the charge m.r  v/ is generated
are weaker than the other regions. Consequently, even if the external load does not
provide torque, the corresponding deformation structural element tends to rotate.
Then by the mechanism explained with Fig. 5.8, nature reduces the further damage
of the material due to the external load. Below, let’s look at the charges more closely.
Equation (4.41) states

G v2 G! 2 G
LD F F C Gj A D  C j 0 A0 C Gj i Ai (4.41)
4 2 2 c
Thus,

c @L
j0 D (5.109)
G @A0
1 @L
ji D (5.110)
G @Ai

Equations (5.109) and (5.110) explicitly represent that j 0 and j i constitute the
temporal and spatial components of the charge in the four vector notation.
 
j0 1 2 3
j D

;j ;j ;j (4.42)
c

In this context, the phase velocity c is the velocity of disturbance. Continuity


Eq. (4.87) says

1 @j0
Dr j (4.87)
c 2 @t
Using the conversion rule from the temporal to spatial differentiation in wave
dynamics, Eq. (5.14),
5.4 Physical Meaning of Charges 129

@
D .r /  c (5.14)
@t
1 @j0 1 1
D  2 .rj0 /  c D r.j0 /  cO (5.111)
c 2 @t c c
This leads to
1
r.j0 /  cO D r  j (5.112)
c
The left-hand side of Eq. (5.112) is the directional derivative in the direction of c.
 
1 1 @j 0 cx @j 0 cy @j 0 cz
r.j0 /  cO D C C (5.113)
c c @x c @y c @z c

where cx , cy , and cz are the x, y, and z components of the phase velocity vector c.
If we put

j0 j 0  cx cy cz 
j D cO D  xO C yO C zO (5.114)
c c c c c
the right-hand side of Eq. (5.112) becomes as follows:
   0 
@ @ @ j cx cy cz 
r j D xO C yO C zO   xO C yO C zO
@x @y @z c c c c
 0 
1 @j cx @j 0 cy @j 0 cz
D C C (5.115)
c @x c @y c @z c

The rightmost-hand side of Eq. (5.115) is identical to the right-hand side of


Eq. (5.113). This means that the expression of j (5.114) satisfies Eq. (5.112). Rear-
ranging Eq. (5.114) into the form that explicitly expresses the temporal component
of the four vector j and noting cO  c D c, we obtain the following relationship
between the temporal and spatial part of j .

j 0 D j  c (5.116)

Therefore,

j0 c
D j  D j  cO (5.117)
c c
Equation (5.117) indicates that the first component of j in the four vector notation
(4.42) is the unit vector of the spatial part of j in the direction of c. Remember
that j 0 and j i respectively originate from the temporal and spatial part of the
gauge used as the additional term in the temporal and spatial covariant derivative.
130 5 Interpretations of Deformation and Fracture Phenomena from Field Theoretical. . .

Equations (5.116) and (5.117) indicate that j 0 =c and j i represents the same gauge
in different perspective; the former in the time-domain differentiation and the latter
in the space-domain differentiation, and the two are connected with the phase
velocity c.

5.5 Thermodynamic Consideration

The discussions we have made so far in the preceding chapter and this chapter
are based on mechanics. In these discussions, we ascribe the irreversibility of
plasticity to the deformation charge and its flow. In the field equations, the effects
of charge and its flow are incorporated into the deformation dynamics as the source
terms. They literally provide the source of plastic deformation-like behaviors of
the displacement field such as decaying wave characteristics and strain localization.
However, they do not refer to the original cause of these effects. In addition, the
field equations do not describe the energy conversion from the form of mechanical
energy of the material to other forms such as thermal and sound energy resulting
from the energy-dissipative process. Intuitively, we know that the generation and
propagation of dislocations and other defects cause plasticity. We also know that the
irreversibility of plastic deformation should be accounted for by thermodynamics.
For further development of the theory, considerations based on thermodynamics and
connections to dislocation theories are essential. These are physically fundamental
and immensive subjects, and are basically beyond the scope of this book. However,
since they are so important, a brief discussion is provided in this section.
According to the second law of thermodynamics, the entropy change dS of a
system undergoing any infinitesimal, reversible process is given by dQ=T , where
dQ is the heat supplied to the system and T is the absolute temperature of the
system. In a material being deformed, part of the material exerts force on the
neighboring part causing a stress. As an example, consider a part of a material being
stretched by neighboring parts in Fig. 5.9. The increase in volume of this part causes
a positive change in entropy dS > 0, because the degree to which the probability
that the system is spread out over different possible microstates increases. Here
a microstate specifies all molecular details about the system including the position
and velocity of every molecule. If this deformation process is reversible, this positive
change in entropy must be compensated by one way or another. If the neighboring
part exerting the volume expansion of this part is being compressed (as is the case of
a compressive elastic wave), the neighboring part should experience a negative
change in its entropy. In this fashion, it is possible that the total net entropy change in
the entire material is zero, and the deformation is totally reversible. Here, the entropy
must flow from one part to others in the form of heat dQ. If all the parts of the
material are being stretched and the deformation is still reversible, the entire system
must absorb the corresponding heat from outside. If the process is adiabatic, the
temperature of the stretched part must decrease. If all the parts are stretched, the
5.5 Thermodynamic Consideration 131

dQ

Fig. 5.9 Heat flow in material being deformed elastically. As the left part is stretched, the right
part is compressed. The entropy increases in the left part and decreases in the right part

temperature of the entire material decreases. This phenomenon is known as thermo-


elastic effect. Later in Chap. 7, we will discuss a case where the temperature of
a metal specimen under uniform tensile deformation decreases initially when the
deformation is in the linear elastic regime, and the temperature goes up when the
deformation enters the plastic regime. If the deformation is irreversible, the entropy
of the stretched part must be increased without exchange of heat, or with an entropy
increase greater than the heat flow, dS > dQ=T . Panin and Egorushkin [33] discuss
a thermodynamic approach to a description of fragmentation of solids.
This thermodynamical argument of plastic deformation is closely related to the
generation and propagation of defects. Egorushkin [34] expresses this situation by
the following expressions.

dS
 D r  I s C s (5.118)
dt
.rT /2  vh  
s D 2
 2 ..˛  vh /  rT / C  (5.119)
T T T
Here,  is the density of the material, I s is the entropy flow and s is the entropy
production,  is the thermal conductivity, T is the temperature, ˛ is the linear defect
density,  is the elastic stress, vh D cjQ d where cQ being the plastic deformation
front velocity and j d the flux of linear defect (the plastic front flows, as the defects
flows, in the opposite direction under very high stress associated with the work
needed to push the defects to go through the material (V.E. Egorushkin, 2008, private
communication), and vh represents the work associated with the defect flow due
to hydrostatic tension. It is interesting to note that the right-hand side of entropy
production expression (5.119) contains elastic stress  , indicating the interaction
between the elastic stress field and irreversibility in plasticity via entropy.
From the viewpoint that the source of plasticity is related to micro-defects
such as dislocations, the energy-dissipative force Gj should be somehow
related to force on those micro-defects. Naively speaking, the proportionality of
Gj D W d .r  v/ D c v [see Eq. (5.70)] is related to the fact that frictional force
acts on mobile dislocations [18].
When the temperature gradient is zero, the condition of irreversibility becomes
vh  
 >0 (5.120)
T
132 5 Interpretations of Deformation and Fracture Phenomena from Field Theoretical. . .

The entropy of metals is associated with lattice vibration [35]. An increase in


entropy thus is associated with some sort of acoustic phenomenon. This is consistent
with the above-mentioned observation that the formation of developed, charge-like
patterns, which represents irreversible deformation, is accompanied by acoustic
emission [23].

References

1. Yoshida, S.: Phys. Mesomech. 11, 140–146 (2008)


2. Yoshida, S.: Scale-independent approach to deformation and fracture of solid-state materials.
J. Strain Anal. 46, 380–388 (2011)
3. Landau, L.D., Lifshitz, E.M.: Theory of Elasticity, Course of Theoretical Physics, vol. 7, 3rd
edn. Butterworth-Heinemann, Oxford (1986)
4. Marsden, J.E., Hughes, T.J.R.: A point of departure. In: Mathematical Foundations of
Elasticity. Prentice-Hall, Englewood Cliffs (1983) (see for example)
5. Spencer, A.J.M.: Continuum Mechanics. Longman, London/New York (1980)
6. Sciammarella, C.A., Sciammarella, F.M.: Experimental Mechanics of Solids. Wiley, London
(2012)
7. Hecht, E.: Optics, 4th edn. Addison-Wesley, Reading (2001)
8. Graff, K.F.: Wave Motion in Elastic Solids. Oxford University Press, London (1975)
9. Lüders, W.: Dinglers Polytech. J. 155, 18–22 (1860)
10. Yoshida, S., Suprapedi, R., Widiastuti, R., Pardede, M., Hutagalung, S., Marpaung, J.S.,
Muhardy, A.F., Kusnowo, A.: Direct observation of developed plastic deformation and its
application to nondestructive testing. Jpn. J. Appl. Phys. 35, L854–L857 (1996)
11. Yoshida, S.M., Muhamad, I., Widiastuti, R., Kusnowo, A.: Optical interferometric technique
for deformation analysis. Opt. Exp. 2, 516–530 (1998)
12. Yoshida, S., Ishii, H., Ichinose, K., Gomi, K., Taniuchi, K.: J. Appl. Mech. 72, 792–794 (2005)
13. Yoshida, S., Siahaan, B., Pardede, M.H., Sijabat, N., Simangunsong, H., Simbolon, T.,
Kusnowo, A.: Phys. Lett. A 251, 54–60 (1999)
14. Yoshida, S.: Interpretation of mesomechanical behaviors of plastic deformation based on
analogy to Maxwell electromagnetic theory. Phys. Mesomech. 4(3), 29–34 (2001)
15. Nakamura, T., Sasaki, T., Yoshida, S.: Analysis of Portevin-Le Chatelier effect of Al-Mg alloy
by electronic speckle pattern interferometry, advancement of optical methods in experimental
mechanics. In: Conference Proceedings of the Society for Experimental Mechanics Series, vol.
3, pp. 109–117 (2014)
16. Yoshida, S., Rourks, R.L., Mita, T., Ichinose, K.: Physical mesomechanical Criteria of plastic
deformation and fracture. Phys. Mesomech. 12(5–6), 249–253 (2009)
17. Panin, V.E., Zuev, L.B., Makarov, P.V., Yegorushkin, V.E., Gorbatenko, V.V., Danilov, V.I.:
Method and apparatus for nondestructive testing of the mechanical behavior of objects under
loading utilizing wave theory of plastic deformation, US patent, 5,508,801 (1996)
18. Suzuki, T., Takeuchi, S., Yoshinaga, H.: Dislocation Dynamics and Plasticity. Springer,
Berlin/New York (1989)
19. Yoshida, S.: Consideration on fracture of solid-state materials. Phys. Lett. A 270, 320–325
(2000)
20. Raizer, Y.P., Allen, J.E., Kisin, V.I.: Gas Discharge Physics. Springer, Heidelberg (1997)
21. Meinlschmidt, P., Hinsch, K.D., Sirohi, R.S.: Selected Papers on Electronic Speckle Pattern
Interferometry: Principles and Practice. SPIE Press, Bellingham, Washington, Milestone
Series, vol. 132 (1996)
22. Yoshida, S.: Physical meaning of physical-mesomechanical formulation of deformation and
fracture. In: AIP Conference Proceedings, vol. 1301, pp. 146–155 (2010)
References 133

23. Yoshida, S.: Optical interferometric study on deformation and fracture based on physical
mesomechanics. J. Phys. Meso. Mech. 2(4), 5–12 (1999) [in English and Russian]
24. Yoshida, S., Toyooka, S.: Field theoretical interpretation of dynamics of plastic deformation
-Portevin-Le Chatelie effect and propagation of shear band. J. Phys. Condens. Matter 13,
6741–6757 (2000)
25. Courtney, T.H.: Mechanical Behavior of Materials, 2nd edn. Waveland Press, Long Grove
(2005)
26. Mertens, F., Franklin, S.V., Marder, M.: Phys. Rev. Lett. 78, 4502 (1997)
27. Cherrington, B.E.: Gaseous Electronics and Gas Lasers. Pergamon Press, Oxford (1979)
28. Harris, J.G.: Linear Elastic Waves. Cambridge University Press, Cambridge/New York (2001)
29. Aitchson, I.J.R., Hey, A.J.G.: Gauge Theories in Particle Physics. IOP Publishing, Bris-
tol/Philadelphia (1989)
30. Schiff, L.I.: Quantum Mechanics, 3rd edn. McGraw-Hill Kogakusha, Tokyo (1968)
31. Thornton, S.T., Marion, J.B.: Classical Dynamics, 5th edn. Brooks/Cole-Thomson Learning,
Belmont (2004)
32. Griffiths, D.J.: Introduction to Electrodynamics, 3rd edn. Prentice Hall, Upper Saddle River
(1999)
33. Panin, V.E., Egorushkin, V.E.: Physical Mesomechanics and nonequilibrium thermodynamics
as a methodological basis for nanomaterial science. Phys. Mesomech. 13(5–6), 204–220 (2009)
34. Egorushkin, V.E.: Dynamics of plastic deformation: waves of localized plastic deformation in
solids. Russ. Phys. 35, 316–334 (1992)
35. Kittel, C.: Introduction to Solid State Physics. Wiley, New York (1956)
Chapter 6
Optical Interferometry and Application
to Material Characterization

The main goal of this chapter is to provide basic information regarding an optical
interferometric technique known as the Electronic Speckle-Pattern Interferometry
(ESPI). In Chaps. 7 and 8, we will discuss a number of experimental observations
that support the present field theoretical approach along with possible applications
of the present theory to engineering problems. Most of these experiments are based
on ESPI. For those who are not familiar with ESPI or optical interferometry in
general, but interested in using those techniques for their own experiments, some
basic concepts of optics are reviewed as well. While a number of commercial
system of optical interferometers are available, often basic understanding of optics
is essential for proper use of those commercial systems and correct interpretation
of data provided by the systems. In Sect. 6.1 basic information of light and optics
is discussed. Various properties of light as an electromagnetic wave, its interaction
with materials along with basics of optics and optical devices including laser are
discussed. Incidentally, some parts of this section are scientifically analogous to
the present field theory. It is expected that these parts help us digest the content of
Chaps. 4 and 5 from a different perspective. Section 6.2 describes practical aspects
of interferometry, including mathematical expressions useful for analysis of optical
interferometry and typical optical interferometer and their working principles.
Section 6.3 focuses on speckles and the principle of ESPI. Note that the intention of
this chapter is to provide supporting materials for Chaps. 7 and 8. The information
provided in each section is limited to that directly related to the contents of Chaps. 7
and 8. For more detailed information, the reader is encouraged to refer to specialized
books on optics [1, 2], electrodynamics [2–4], optical interferometry [5–7], and
quantum mechanics [4, 8, 9]. Some of these books are cited in the text below
whenever the relevance is high for a specific topic.

© Springer Science+Business Media New York 2015 135


S. Yoshida, Deformation and Fracture of Solid-State Materials,
DOI 10.1007/978-1-4939-2098-3__6
136 6 Optical Interferometry and Application to Material Characterization

6.1 Basics of Light and Optics

“Is light a wave or particle” is one of the most frequently asked questions regarding
light. In a sense, the question is similar to asking whether a water molecule in the
ocean is a wave or particle. Both water molecules and photons behave like a wave,
carrying energy from one point to another. However, there are crucial differences
between the photon and the water molecule. Water waves travel only through water
but light waves can travel in vacuum. Light wave as an electromagnetic wave carries
the information regarding a charged-particle’s motion to other charged-particles [4].
Water waves carry simply water molecules energy, although it can be huge energy
like a gigantic Tsunami.
The photon as a particle has additional subtleness. It becomes meaningful only
in the context of interaction with matters such as absorption by an atom. When the
frequency of light incident to an atom is the same or close to the resonant frequency
of a radiative transition of the atom, the optical energy is transferred to the atom as
the photon energy h (h is the Plank constant and is the frequency.) Thus, without
the existence of an atom or molecule, even if we know the frequency of the light
and hence the value of h , the photon energy is almost meaningless. It is not that a
number of particles each having energy of h are flowing in vacuum like a number
of tennis balls having the same kinetic energy are moving.
While light carries optical energy both in vacuum and through a medium, the two
cases should be discussed fundamentally differently. Below, we consider the case
when light is transmitted through vacuum first, and then the other case considering
the interaction of matters with light.

6.1.1 Light as an Electromagnetic Wave

An electromagnetic wave is given as a solution to the Maxwell equations [1–3].


When it travels in vacuum where there is no charge, the right-hand side of the
first equation, the source term, can be put to zero. With the lack of charges, the
conduction current is also zero on the right-hand side of the third equation as well.
Accordingly, the Maxwell’s equations take the following forms:

r E D 0 (6.1)
@B
r E D  (6.2)
@t
@E
r  B D  0 0 (6.3)
@t
r B D 0 (6.4)
6.1 Basics of Light and Optics 137

where E is the electric field, B is the magnetic field, 0 is the electric permittivity in
vacuum (free space), and 0 is the magnetic permeability in vacuum. Taking the curl
of Eq. (6.2) and substituting the resultant expression into time derivative of Eq. (6.3)
(like we did in Chap. 4), we obtain the following differential equation:

@2 E
r 2 E D  0 0 (6.5)
@t 2
p
Equation (6.5) is a wave equation with phase velocity of 1= 0 0 . The general
solution to Eq. (6.5) has the following form:
 
t
E D E0 sin p ˙kr (6.6)
 0 0

Here k is the propagation vector and r D x xO C y yO C zOz. The argument of the sine
function is called the phase of the wave. The first term of the phase represents how
the phase changes as the time elapses and the second term represents the spatial
variation in the phase in the direction of the propagation vector. The term s 
k  r is the component of the space coordinate in the direction of the propagation.
Notice that when the sign between the temporal and spatial parts of the phase term is
positive, the point of the same phase shifts towards the negative direction of s D kr
with the elapse of time because it is necessary to reduce the value of s to keep the
phase the same when t increases. In other words, the wave propagates in the negative
direction of the s-axis. Similarly, if the sign between the temporal and spatial part
of the phase term is negative, the wave propagates in the positive direction of the
s-axis.
The oscillatory behavior of an electromagnetic wave comes from the Faraday’s
law of magnetic induction, Eq. (4.66). Faraday’s law represents the Lenz’s law
in electrodynamics based on the synergetic interaction between the electric and
magnetic fields. In Chap. 4, we analyzed Faraday’s law and how the oscillatory
behavior of the electromagnetic field propagates based on the field equations (6.2)
and (6.3) (see Figs. 4.5 and 4.6). Here we consider the effect through analysis of LC
oscillation in Fig. 6.1.
Suppose the capacitor is fully charged, and you close the switch. The positive
charges on the positive electrode will flow through the inductor. The voltage stored
in the capacitor at a given time is related to the charge at that time as

switch

q(t) d
v(t) = v(t) = L i(t)
C dt

Fig. 6.1 LC circuit


138 6 Optical Interferometry and Application to Material Characterization

q.t /
vC .t / D (6.7)
C
Here vC is the voltage across the capacitor, C is the capacitance, and q is the charge
stored in the capacitor. The voltage drop across the inductor changes in proportion
to the temporal change in the current flowing through.

d i.t /
vL .t / D L (6.8)
dt

Here L is the inductance and i.t / is the current flowing through the inductor.
Equations (6.7) and (6.8) lead to the following differential equation:

q.t / d2
 L 2 q.t / D 0 (6.9)
C dt
The general solution to Eq. (6.9) has the following form:
 
1
q.t / D A sin p t (6.10)
LC
p p
Here the oscillation frequency is 1=.2 LC /; the greater the value of LC , the
slower the oscillation. This makes sense because large capacitance means it takes a
long time to charge up the capacitor and large inductance means that the inductive
reactance is high making the temporal change of the current slow (and therefore
makes the oscillation period long). If there is a resistor in the circuit, the motion of
the charge decays.
The LC circuit model can be used to explain the propagation of light as well.
Consider a light wave that travels through a medium whose electric permittivity is
 and magnetic permeability is . To apply the Maxwell equations to this case, we
need to replace 0 and 0 on the right-hand side of Eq. (6.3) with  and . It follows
p
that the phase velocity of the light passing through the medium is c D 1=  .
The propagation of the light wave at this phase velocity can be explained by
viewing the space filled with this medium as comprising effective capacitors and
inductors. See Fig. 6.2 where the LC circuit shown in Fig. 6.1 is the source of an
electromagnetic wave. Consider a moment when the conduction current flowing
through the inductor increases. This induces a magnetic field in the space around the
inductor. As the conduction current increases, so does the induced magnetic field.
In a small area labeled “Area A,” the increase in the induced magnetic field induces
a looped displacement current proportional to @B=@t as shown in the figure. This
displacement current induces a magnetic field crossing through Area A upward,
opposing the increase in the magnetic field induced by the conduction current
(Faraday’s law). This magnetic field induced by the displacement current in turn
causes a downward increase in the magnetic field at the next point, at the cross-
sectional area labeled “Area B.” In this fashion, by the mechanism of Faraday’s law,
6.1 Basics of Light and Optics 139

Fig. 6.2 E and B field


change travels by Faraday’s
law
B

Area A Area B

B

t

the field change is transferred to the next point of the space one after another. Wave
equation (6.5) represents this effect; the effect is transferred at the phase velocity of
p
1=  .

Electric and Magnetic Energy

The wave characteristics of electromagnetic fields can also be argued from the
viewpoint of electromagnetic energy stored in the space. The electromagnetic
energy density in a vacuum is given by

0 2 1 2
uD E C B (6.11)
2 2 0

where 0 and 0 are the electric permittivity and magnetic permeability in free
space.
It is worth deriving the above equation from the energy stored in an LC circuit
like the one shown in Fig. 6.1. Considering the change in the energy stored in a
capacitor by adding an infinitesimal amount of charge dq and using Eq. (6.7), we
can express the energy after charging an initially empty capacitor for time t as
Z t Z t
q.t / 1 2 1
vC .t /dq D dq D q D CV 2 (6.12)
0 0 C 2C 2

Here q is the charge on the capacitor and vc is the voltage across the capacitor, and
the relationship q D C vC is used in going through the rightmost equal sign.
Similarly, considering that the power into an inductor is vL i and integrating for
time the energy change in an inductor is1
Z t Z t
di 1
vL .t /i.t /dt D L idt D Li 2 (6.13)
0 0 dt 2

1
Integrate in part or consider the time derivative of .1=2/Li.t /2 .
140 6 Optical Interferometry and Application to Material Characterization

Here VL is the voltage across the solenoid and L is the inductance of the solenoid.
Thus the total of the energy stored in the capacitor UC and the energy stored in
the solenoid UL at the moment when the voltage across the capacitor is vC and the
current flowing through the solenoid is iL is given as

1 2 1
UC C UL D C vC C LiL2 (6.14)
2 2
Assume that the circuit consists of a parallel plate capacitor and a solenoid of N
turns. The capacitance C is given as

A
C D (6.15)
d

Here d is the distance between the parallel plates and A is the common area of
the plates. So, assuming that the electric field E is uniform and therefore given as
E D vC =d , the stored energy is

1 2 1 A 1  vC 2 1
C vC D  v2C D  dA D E 2 .Ad / (6.16)
2 2 d 2 d 2

Since Ad is the volume, Eq. (6.16) indicates that the energy per unit volume is given
by E 2 =2. The first term of Eq. (6.11) can be interpreted as a special case where the
capacitor happens to be the free space. Similarly, the inductance L is given as

N 2A
LD (6.17)
l

Here l is the length of the solenoid and A is the cross-sectional area of the solenoid.
Since the magnetic field is given as B D N i= l , and therefore the current i can be
expressed as i D Bl=. N / the stored energy becomes
 2
1 2 1 N 2A l 1 2
Li D B D B .Al/ (6.18)
2 2 l N 2

Again, since Al is the volume, this indicates that the energy stored per unit volume
is given by B 2 =.2 /, which corresponds to the second term of Eq. (6.11).
From these arguments and Fig. 6.2, we can view the propagation of an electro-
magnetic wave as consecutive transfer of the capacitive and inductive energy density
to the neighboring space via Faraday’s law. At a given point of space, the form of
energy alternates between the capacitive electric energy and the inductive magnetic
energy. At a moment when the space is fully charged, the capacitive energy is at
its maximum. At the same time, since the temporal change of the stored charge
(current) is zero, the inductive energy is zero at this moment. On the other hand, at
a moment when the space’s stored energy is zero, the temporal change of the stored
charge (current) is at the maximum. So, the inductive energy is at the maximum.
6.1 Basics of Light and Optics 141

Therefore, the total electromagnetic energy density umax at a given point is given as
follows:
0 2 1 2
umax D E D B (6.19)
2 peak 2 0 peak

where Epeak and Bpeak denote the peak values of the electric and magnetic fields.
p
Since this energy travels at the speed of c D 1= 0 0 , the energy flux per unit
cross-sectional area is

c0 E 2
I D (6.20)
2

This quantity is called the intensity of the light wave, and has the unit of (W/m2 ).
The energy flow can be conveniently represented by Poynting vector S .

1
S D E B (6.21)

It is worth examining Eq. (6.21) based on Faraday’s law referring to Fig. 6.2. For
simplicity, consider the case where the magnetic field is oriented along the y-axis,
the electric field is oriented along the x-axis, and the magnetic field is traveling in
the positive z-direction (Fig. 6.3). Further, consider a moment when the magnetic
field is decreasing over time. Under this condition, Eq. (4.66) becomes
 
@By @Ex @Ez @Ex
D  .r  E /y D   D (6.22)
@t @z @x @z

From the condition that the magnetic wave propagates in the positive z-direction,
we can put its wave characteristics in the following form:

By D Bpeak sin.!t  kz/ (6.23)

From the condition that the magnetic field decreases over time and Eq. (6.22),

@By @ @Ex
D Bpeak sin.!t  kz/ D Bpeak ! cos.!t  kz/ D  (6.24)
@t @t @z

So,
Z
!
Ex D Bpeak ! cos.!t  kz/d z D Bpeak sin.!t  kz/ D Epeak sin.!t  kz/
k
(6.25)
From these arguments, the electric wave becomes as shown in Fig. 6.3. Note that the
direction of the wave is E  B. Since !=k is the phase velocity of the wave, from
142 6 Optical Interferometry and Application to Material Characterization

Fig. 6.3 Poynting vector y

z
x

By Ex

Eq. (6.25) it is known that the ratio of the amplitude of the electric field to that of
the magnetic field is the phase velocity or the speed of light c.

Ex D cBy (6.26)

Based on the analogy to electrodynamics, the electric and magnetic fields corre-
spond to v and ! of the deformation field. The field equation (4.62) and the equation
of rotational wave (5.30) lead to the deformation-field version of Eq. (6.26) to be
s
G
vx D !y (6.27)


It is interesting to note that Eq. (6.27) shows the fact that as the rotational wave
propagates, the v wave propagates in the same direction at the same velocity being
the wave of the differential displacement causing the rotation. This pictures the
deformation-field version of Poynting vector [see (4.93), (4.94) and Fig. 4.10].

6.1.2 Interaction with Media and Photon

As discussed above, light propagates more slowly in media than in air or vacuum
because the permittivity and permeability is higher than in vacuum. From the fact
that the color does not change in a transparent medium, we know that the frequency
does not change in media. Since the phase velocity is the product of the frequency
and wavelength, this indicates that light changes the wavelength as much as it
changes the velocity in a medium. The ratio of the speed of light in a medium to
that in vacuum is known as the index of refraction n.

c0 0
nD D (6.28)
c
6.1 Basics of Light and Optics 143

Here c is the speed of light, is the wavelength, and the suffix 0 indicates
vacuum. Using the electric permittivity and magnetic permeability in the medium
and vacuum, we can express the index of refraction as follows:
r
c0  p
nD D D  r r (6.29)
c  0 0

Here r is the relative permittivity


p and r is the relative permeability. In most media,
r  1, and therefore n   r .

Response of Dielectric Media

Let us consider the underlying physics that determines the phase velocity of light.
Dielectricity and conductivity are the two most important properties of a medium
when a light wave propagates through it. Dielectricity is the property that dielectric
media possess, in which pairs of positive and negative charges line up under the
influence of an external electric field. The situation in the dielectric material in a
parallel plate capacitor is a good example as shown in Fig. 6.4. Conductivity is
the property of conductors, in which free charges move under the influence of an
external electric field. In a dielectric medium, the positive and negative charges in
each charge pair are bound with each other. Therefore, when an external electric
field is applied, they stop moving after being shifted in mutually opposite directions.
This situation can be argued as that the binding force due to the electric field between
the paired charges resists the electric force due to the external field. The pair of
unlike charges is called the dipole. The external field does more work to separate
the unlike charges in a dipole farther if the charge is greater or the distance between

Fig. 6.4 Dielectricity in a


parallel plate capacitor
144 6 Optical Interferometry and Application to Material Characterization

the pair is greater. Thus, the degree of this effect can be measured by the product of
the charge and the separation known as the dipole moment p.

p D qd (6.30)

Here q is the charge of the dipole and d is the vector whose magnitude is the distance
between the unlike charges in the dipole and the direction is from the negative to
positive charge. In a unit volume of a given dielectric medium, there are a number
of dipoles. To consider the overall effect per unit volume, it is necessary to sum all
of the individual dipole moments. The resultant quantity is called the polarization
vector P.
It is important to consider the polarization in connection with the electric field
and the electric permittivity of free space (in vacuum). Apply Gauss’s law (4.65) to
a unit volume that has both bound charges and free charges.

 0 r  E D  b C f (6.31)

Here b and f denote the bound and free charges in the unit volume of interest.
By considering the electric potential due to a dipole, it can be easily shown that the
polarization vector is related to the bound charge density as follows [Apply Gauss’s
law to a thin closed volume just to cover the top electrode of the capacitor shown
in Fig. 6.4 considering that the electric field is null above and on the sides of the
electrode and inward to the closed volume below the electrode, see Chap. 4 of [3]].

r  P D b (6.32)

Equations (6.31) and (6.32) lead to

r  .0 E C P/ D f (6.33)

The quantity inside the parenthesis on the left-hand side of Eq. (6.33) is known as
the electric displacement vector D.

D D 0 E C P (6.34)

In the case of linear media, the electric displacement vector is proportional to the
electric field with the electric permittivity as the constant of proportionality.

D D E (6.35)

From Eqs. (6.34) and (6.35), it is clear that in a linear medium the polarization vector
is proportional to the electric field.

P D D  0 E D E  0 E (6.36)
6.1 Basics of Light and Optics 145

We can put the above equation as

P D 0 e E (6.37)

The quantity e represents how susceptible the polarization is to the electric field
and called the electric susceptibility of the medium. Using the electric susceptibility,
we can express the electric displacement vector as follows:

D D E D 0 E C 0 e E (6.38)

Comparison of Eqs. (6.35) and (6.38) leads to the following relationship between
the electric permittivity in the medium and that in free space.

 D 0 .1 C e / (6.39)

Further, by using the relative permittivity r  =0 and referring to Eq. (6.29), we
can relate the index of refraction and the electric susceptibility.

1
p
nŠ r D .1 C e / 2 (6.40)

In usual gases e << 1 . Thus the above expression is further simplified as

1
n Š 1 C e (6.41)
2
Equation (6.41) indicates that the index of refraction follows the frequency depen-
dence of susceptibility. This is quite important to understand various phenomena
associated with light transmission in media including absorption, phase change, and
dispersion.
The frequency dependence of the electric susceptibility can be discussed though
consideration of response of a dipole to a light wave as an alternating electric field.2
Consider the force acting on the electron of a dipole when a light wave passes
through the dielectric medium. The equation of motion can be expressed as follows:

d 2x dx
m D kx  m C qE (6.42)
dt 2 dt
Here x is the spatial coordinate representing the displacement of the electron from
its equilibrium position, m is the mass of the electron, k is the spring constant of
the binding force,  is the damping coefficient associated with the conductivity of
the medium, and E is the electric field of the light wave propagating through the

2
See Chap. 8 of [2] or Chap. 9 of [3].
146 6 Optical Interferometry and Application to Material Characterization

medium. The approximation of the binding force as a spring force is justified by


that the potential of the bound electron is well approximated by a quadratic function
of x; i.e., the force as the first order derivative of the potential is proportional to x.
Lettingpthe natural (angular) frequency of the oscillation due to the spring force be
!0 D k=m and using the exponential notation for the electric field of the light
wave as E D E0 e i!t , Eq. (6.42) can be put in the following form of a second order
linear differential equation.

d 2x dx q
C C !0 2 x D E0 e i!t (6.43)
dt 2 dt m
The particular solution to Eq. (6.43) can be put in the following form:

xQ D xQ 0 e i!t (6.44)

Substituting Eq. (6.44) into Eq. (6.43), we obtain


 2 q
!  i ! C !0 2 xQ 0 e i!t D Ee i!t (6.45)
m
Hence,

qE 1
xQ 0 D (6.46)
m .!0 2  ! 2 /  i !

leading to a complex dipole moment of

q2 1
Q / D q xQ 0 e i!t D
p.t / D q x.t E0 e i!t (6.47)
m .!0 2  ! 2 /  i !

Assuming that there are N molecules in the unit volume and each molecule has fj
electrons whose resonance is at !j and damping coefficient is j , we can put the
polarization in the following form:

q2N X fj
PD   E (6.48)
m j ! 2  ! 2  i !
j j

From Eq. (6.37), the electric susceptibility becomes

P q2N X fj
e D D   (6.49)
0 E 0 m j ! 2  ! 2  i !
j j
6.1 Basics of Light and Optics 147

Then from Eq. (6.41), the index of refraction takes the following form:

1 q2N X fj
n Š 1 C e D 1 C   (6.50)
2 20 m j ! 2  ! 2  i !
j j

Apparently, the index of refraction is a complex number. The physical meaning of a


complex refractive index becomes clear through consideration of the phase velocity
c D !=k D c0 =n as follows:
! ! ! !
kD D n D nR C i nI (6.51)
c c0 c0 c0
Here nR and nI denote the real and imaginary parts of the index of refraction (6.50)

q2N X fj !j 2  ! 2
nR D 1 C  (6.52)
20 m j !j 2  ! 2 2 C .!j /2

q2N X fj j !
nI D  (6.53)
20 m j !j 2  ! 2 2 C .!j /2

So, the imaginary part of the propagation constant becomes

! ! q2N X fj j ! q2N !2 X fj j
nI D  2 D  2
c0 c0 20 m !j 2  ! 2 C .!j /2 20 mc0 !j 2  ! 2 C .!j /2
j j

(6.54)

Thus, the exponential term of the light wave, becomes

1 n nR ! nR ˛
i!.t z/ i!.t z/ i!.t z/  nI z i!.t z/  z
i.!tkz/ c0 D e c0 e c0 c0 e 2
e De c De De
(6.55)
Here ˛ is the absorption coefficient in power (square of amplitude).

! q2N !2 X fj j
˛D2 nI D  (6.56)
c0 0 mc0 j !j  ! 2  .!j /2
2

Figure 6.5 illustrates the index of refraction and absorption coefficient as a function
of the frequency of the incident light wave near one of the resonances corresponding
to a certain fj in Eqs. (6.50) and (6.56). Notice that as the frequency increases from
the lower end toward the resonance, the index of refraction increases. This means
that within the spectrum width of the incident light, the higher frequency (shorter
wavelength) travels more slowly than the lower frequency (longer wavelength). This
phenomenon is known as the dispersion of light.
148 6 Optical Interferometry and Application to Material Characterization

1
α
n-1
0.8
absorption and refractive index (au)

0.6

0.4

0.2

−0.2

−0.4
3 3.5 4 4.5 5 5.5 6 6.5 7
optical frequency (au)

Fig. 6.5 Absorption and index of refraction near a resonance

The plots of Fig. 6.5 can be viewed as a transfer function of energy from the
optical field to the atomic system. Depending on the frequency of the incident light
relative to the resonant frequency of the dipole of the medium, the amplitude (the
dashed-line) and phase (solid-line) of the dipole oscillation is altered according to
Fig. 6.5. When the frequency of the incident light matches the resonant frequency,
total energy transfer can occur from the optical field to the atomic system of the
medium. Consequently, the atomic system changes its energy state. In this situation,
the optical field behaves as if it is a particle. This particle is known as the photon.
We will make a quick review of the photon in the next section.

Photon

The photon is known as quantization of electromagnetic field. What does it really


mean? My interpretation is as follows. The concept of photon becomes meaningful
only through interaction with a medium. When a dielectric medium is exposed to
the alternating electric field of a light wave, the dipoles are oscillated by the electric
force; the electric energy is transferred to the dipoles’ kinetic energy. If the medium
is conductive, there is a loss, meaning that part of the electromagnetic energy in
the incident light is lost as the dipole oscillates. Consequently, the transmitted
(re-emitted by the atomic system) light has lower energy than the incident light.
We call this phenomenon absorption. As indicated in Fig. 6.5, absorption occurs at
6.1 Basics of Light and Optics 149

or near the resonance. If the incident light is white, only the frequency component
near the resonant frequency of the dielectric medium is absorbed. The rest of the
spectrum is transmitted through. That is why leaves are green in sunlight. In this
view, the dielectric medium provides a transfer function of incident to transmitted
light’s electromagnetic energy. It is easily imagined that this reemission can cause
phase delay, as a transfer function of a mechanical or electrical system shift the
phase of the input signal. This determines the phase velocity of light inside the
medium. Notice that the index of refraction Eq. (6.41) has the electric susceptibility
e as part of the expression. As Eq. (6.37) indicates, e represents how strongly the
dipole responds to the electric field of the incident light.
The quantization of the energy originates from the atomic system of the dielectric
medium. Quantum mechanics describes particles via their wave functions of
probability. In other words, according to quantum mechanics, the location of an
electron cannot be specified in the form of a definite position at a definite time.
Instead, quantum mechanics provides us the probability to find the electron as a
function of time and space. Consequently, the physical quantities of the electron
such as its energy are provided as an expectation value, and the energy state of
the atom is computed as an eigen state identified by a discrete number called the
quantum number. Thus, the energies of different atomic states are not continuous,
but separated by a finite increment. As such, the electromagnetic energy absorbed
by such an atomic system is a discrete quantity.
This discrete nature of quantum states can be intuitively understood as follows.
Bound electrons of an atom orbit around the nucleus. The projection of the orbiting
motion onto a linear axis is harmonic oscillation-like; if the orbit is circular at
a constant speed, the projection to an axis in the plane of the circle is purely
harmonic (e.g., the projection to the x-axis when the circle is in an x–y plane).
If the projected motion is not purely harmonic, it can be expanded to a series of
harmonic oscillations (Fourier theorem). So, we can model the oscillatory motion
of an electron with a spring-mass system under harmonic oscillation without losing
the generality. Imagine you are watching a harmonic oscillation of a mass attached
to a spring from a side. If you were asked “what is the position of the mass?” How
would you answer the question? You cannot answer like “the mass is x .cm/ away
from the origin” because the mass keeps moving. We could definitely say that the
mass is between the two turning (end) points of the oscillation. Probably the best you
could do is to provide the probability to find the mass between the turning points
quantitatively, or the expectation value of the location of the mass. Alternatively,
you could take a number of still photos of the mass on the same film superposing
the multiple exposures and show it to the questioner; the resultant picture will show
more images of the mass near the two turning points than the middle. If you take
statistically sufficient number of shots, the image will be symmetric around the
center of the oscillation. The statistics can be argued in terms of the velocity (i.e.,
the momentum) of the mass. In this case, the velocity is expected to be highest at the
center of the oscillation. In addition, once we know the probability of the location of
the mass, we can calculate the spring potential energy as well as the mass’s kinetic
energy from the velocity. So we can express the energy of the system in the same
150 6 Optical Interferometry and Application to Material Characterization

Fig. 6.6 Concept of laser


resonator

fashion. The famous Heisenberg’s uncertainty principle can be naively understood


along the same line of argument.

6.1.3 Laser and Coherence

The word laser stands for Light Amplified Stimulated Emission of Radiation. As
this word indicates, amplification of stimulated emission is essential to obtain laser
light. The coherence of a laser beam, meaning that the phase difference of the light
wave between two points apart by the same amount in either in time or space
is constant,3 is generated in an optical resonator. Figure 6.6 illustrates an optical
resonator schematically. For generation of coherent light with an optical resonator,
essentially two conditions must be established. First, the total number of photons
emitted from the laser medium must exceed the total number of photons absorbed
by the medium. This condition is referred to as the population inversion. Second,
the light wave must be well defined and completely controlled inside the optical
resonator so that the stimulated emission occurs in a coherent fashion.

Population Inversion and Optical Amplification

When an atom4 at the ground state absorbs a photon by the mechanism discussed
in the preceding section, the atom is excited to a higher-energy state (the excited
state). An excited state created in this fashion does not last forever. After the period
called the radiative life time, the excited state gives up energy to a photon and goes
back to the ground state. This process is called the spontaneous emission [2, 10].
If a photon of the same energy as the energy difference between the excited and
the ground states is incident to an atom at the excited state, the atom goes back to
the ground state emitting a photon having the same energy as the incident photon.
Consequently, the number of the photons of this energy is increased by one. This
process is called the stimulated emission [2, 10]. Here the state resulting from

3
These are called the temporal and spatial coherence, respectively. As a wave, the spatial coherence
is equivalent to the temporal coherence. [Recall the conversion between the temporal and spatial
differentiations in wave dynamics expressed by Eq. (5.14).]
4
The same argument holds for a molecule.
6.1 Basics of Light and Optics 151

stimulated emission does not necessarily have to be the ground state. It can be any
lower state as far as radiative transition is allowed [11] and the energy gap is the
same as the incident photon. In thermal equilibrium, atoms at a higher-energy state
are less populated than the lower states. Therefore, the absorption always exceeds
emission. If the number of the atoms at the excited state is greater than the lower
state, the emission exceeds the absorption. This situation is called the population
inversion. In this situation, the number of the photon of the same energy as the
initial photon keeps increasing as the light wave passes through the population-
inverted medium. In other words, the light intensity grows as the light wave travels.
We say that the light is amplified or the medium has positive laser gain.
To establish a population inversion, the laser active medium must be excited by
an external energy source such as an electrical, optical, and chemical energy source.
The excitation by an external source is referred to as pumping. Depending on the
method of pumping, the laser is referred to as an electrically pumped laser, optically
pumped laser, etc. The stimulated emission converts part of the pumped energy to
optical energy.

Optical Resonator and Phase Condition

Consider a population-inverted atomic system (called the gain medium) is placed


in a space between a pair of mirrors forming an optical resonator as shown in
Fig. 6.6. Here the block near the center of the space is the gain medium. As the
light wave goes back and forth between the mirrors, it passes through the gain
medium, and consequently, its intensity is amplified in each trip. This increase in the
optical intensity is referred to as the laser gain. Normally, there is some mechanism
of optical loss such as absorption or scattering in the resonator. If the laser gain
overcomes the loss for each round trip of the light, the net gain is positive. By
making one of the mirrors partially transmissive, we can establish a steady-sate
situation where the net laser gain is balanced by the transmission through the mirror.
In this situation, the transmitted light can be used as coherent light that the resonator
outputs. The partially transmissive mirror is called the output mirror of the laser
resonator.
To make the pair of the mirrors resonate with the optical field, the light
wave at the output mirror after N th round trip must be in-phase with the
N  1th; N  2th : : :1st round trips. This condition is established by making the
resonator length equal to an integer multiples of a half the wavelength; a round-
trip optical path is an integer multiple of the wavelength. The shape of the inner
surface of the resonator mirror defines the shape of the wavefront at the location
of the mirror. Normally, the transverse size of the gain medium is smaller than the
distance between the resonator mirrors. Consequently, the light wave expands with
a spherical wavefront as it travels toward the mirrors. To confine such a wave, the
inner surface of both or one of the mirrors is shaped like part of a sphere, as indicated
152 6 Optical Interferometry and Application to Material Characterization

in Fig. 6.6. In this way, the mirrors focus the light wave back to the gain medium on
reflection, and thereby confine the optical field stably in the intra-resonator space.5
In a given optical resonator, the optical field resonance occurs in certain modes
called the resonant modes. This is similar to the situation where you generate a
standing wave with a string grabbing its two ends and oscillating it using something
like a guitar pick. If the string is L in length, the wavelength of the standing wave can
be an integer multiple of L=2. If the integer is N , the standing wave has N C 1 anti-
nodes; e.g., if N D 0, there is one anti-node. In an optical resonator, these integers
correspond to longitudinal mode numbers.6 Being a wave with a two-dimensional
wavefront, the light wave inside an optical resonator has transverse modes called the
TEM (Transverse Electro-Magnetic) modes as well. With two integers that specify
the numbers of anti-nodes in the two transverse directions, the transverse modes are
denoted TEM 00; TEM10; : : :. Say the coordinates system is xyz with x and y in
the transverse directions and z along the optical path, TEM10 means there are two
anti-nodes along x-axis and one along y-axis.

Gaussian Beam and Propagation

For many applications where the optical phase is important such as optical
interferometry, a laser device designed to generate the output light beam in the
TEM00 mode is used. Often this mode is called the Gaussian mode as the transverse
distribution of the electric field (the amplitude), hence the intensity as well, are in
the form of Gaussian functions. The advantage of the Gaussian mode is that the
amplitude and phase are given by simple mathematical expressions and therefore the
analysis is simple. The amplitude of a Gaussian beam can be expressed as follows:
0 1
1 ik
i.kz /r 2 @ C A
w0 2
w .z/ 2R.z/
E.x; y; z/ D e (6.57)
w.z/

where E is the electric field amplitude, k D 2= is the wave number for
wavelength , is the on-axis longitudinal phase delay called the Gouy phase, w.z/
is the beam radius called the beam spot size where the amplitude is 1=e of the peak
on the axis at a given z, w0 is the minimum value of w.z/ at the z location called the
beam waist, and R.z/ is the radius of curvature of the wavefront. The z-dependence
of the beam spot size and radius of curvature are given as follows when the beam

5
This type of resonator is called the stable resonator.
6
Normally, the distance between the resonator mirrors is much longer than the wavelength of the
light, and therefore the longitudinal mode number is large; e.g., if the wavelength is 1 m and the
resonator length is 50 cm, the mode number is 50  102 =.1  106 =2/ D 0:5=57 D 1  106 .
6.1 Basics of Light and Optics 153

Fig. 6.7 Gaussian beam

waist is located at z D 0. A Gaussian beam is uniquely defined if a combination of


the beam waist location and the minimum spot size w0 is specified.
s
 
z 2
w.z/ D w0 1 C (6.58)
zR
 2 !
zR
R.z/ D z 1 C (6.59)
z

Here zR is the parameter called the Rayleigh length defined as zR D .  w20 /= .


Rayleigh length indicates how quickly a Gaussian beam diverges. R is the radius of
the circle whose circumference is the contour of the phase, which is a function of
z, the axial distance from the beam waist. A small value of R means that the phase
value varies quickly as you deviate from the beam axis in a transverse direction
at z where R is evaluated. A Gaussian beam keeps diverging in accordance with
Eqs. (6.58) and (6.59) after coming out from the resonator. When we use a Gaussian
beam for an application, we can calculate the beam size and radius of curvature at
a given axial distance with Eqs. (6.58) and (6.59). A Gaussian beam can be easily
converted to another Gaussian beam with a lens, as a lens changes the radius of
curvature.
Figure 6.7 illustrates a sample Gaussian beam around its beam waist, where the
horizontal axis z is the beam axis, the dashed lines indicate the beam spot size at
each z, and the solid lines are the wavefronts (the curvature represented by R) at
several z’s. Note that to generate this Gaussian beam, we can place at any point on
the z axis a mirror whose inner surface is identical to the wavefront at the location.

6.1.4 Imaging

Imaging devices including human eyes consist of an optical sensor that detects
optical intensity and a mechanism such as a converging lens to focus multiple light
rays coming from the object to the sensor. The role of the focusing mechanism is
to map a single point on the object to a single point on the detector’s sensing plane
(the image plane). Figure 6.8 illustrates this mechanism. Here lo is the distance from
the object to the lens, li is the distance from the lens to the image plane, and f is
the focal length of the lens. Three rays coming from the tip of the object shown in
154 6 Optical Interferometry and Application to Material Characterization

Fig. 6.8 Thin lens equation lo

ho

f
hi

li

1 1 1 li hi
+ = M= =
li lo f lo ho

this figure are converging at the same single spot on the film plane. This is the point
where the image of the tip of the object is formed. In a camera, the image plane is
where a film should be placed. Imagine you displace the film towards the lens (the
arrow enclosed by dashed lines in Fig. 6.8). The upper ray coming from the tip of
the object will hit the image plane closer to the axis than the lower ray. This means
that the rays coming from the tip of the object do not converge at the same spot on
the film, and therefore the image will be blurry. It is clear that there is only one point
on the axis where a clear image can be formed. This point can be found from the
following relation among lo , li and f , known as the thin lens equation7

1 1 1
D C (6.60)
f lo li

The other points we should discuss with Fig. 6.8 are the image is upside down
and its size is not the same as the object. The former point is always true, including
our eyes. We do not see things upside down because our brain turns the image back
to the right-side-up orientation. The size of the image is related to the size of the
object by the following equation.

li hi
D M (6.61)
lo ho

Here the ratio M is called the magnification of the imaging system.

7
This equation is an approximation of a more general equation applicable to thick lenses. However,
this equation will be applicable to most cases in our laboratory.
6.2 Interference and Interferometry 155

6.2 Interference and Interferometry

6.2.1 Mathematics of Waves

The simplest form of wave is a monochromatic harmonic wave. According to


Fourier’s theorem, a given wave can be expressed as a linear combination of a
harmonic wave and its higher-order harmonics. From this viewpoint, the discussion
we make here for a monochromatic harmonic wave is is applicable to a given wave.
Consider a sine wave traveling in the positive x-direction with amplitude A,
frequency f , and wavelength . It is customary to use the angular frequency
and wave number instead of the frequency and wavelength. These parameters are
defined as
2
! D 2f D (6.62)
T
2
kD (6.63)

Here T is the period of the oscillation. With the use of ! and k the sine wave takes
the following form:

D A sin.!t  kx/ (6.64)

The argument of the sine function is the phase, as defined in Eq. (6.65).

 D !t  kx (6.65)

Essentially, interferometry is a metrology that uses a relative phase or a phase


change to measure the physical quantity of interest for a given application. Solid
understanding of the concept of the phase is important. Let’s take some time here to
consolidate the understanding of the phase and related concepts.
In harmonic waves, the phase term consists of the time and space factors. The
time factor, or the first term on the right-hand side of Eq. (6.65), measures the change
in the value of the sine function due to a change in time in the unit of the temporal
periodicity T . Equation (6.62) converts this change in time in (s) to (rad); one period
in (s) corresponds to 2 in (rad). As you may notice, the symbol ! used to denote
the angular frequency is often used for rotational velocity. In fact, there is a close
connection between circular motion with a constant speed and oscillatory motion
of wave dynamics. Consider in Fig. 6.9 a point P under a constant-speed circular
motion. The radius of the circle is A and point P moves on the circle counter-
clockwise at the rotational velocity !. Figure 6.9 represents the moment when the
angle made by point P is t .t / from the initial position on the horizontal line. Here
the subscript t indicates that t .t / is the time factor of the phase expression (6.65).
The sine curve shown on the right to the circle is the trace of the height of point P
156 6 Optical Interferometry and Application to Material Characterization

Q
P
qt (t)
qt (t) qt (t + Dt)

Fig. 6.9 Circular motion and phase

Fig. 6.10 Spatial periodicity


and phase velocity of a x3 x2 x1
traveling wave
t+dt t

x1 x2 x3 x

from the horizontal line as it moves on the circumference. It is seen that the angle
t .t / corresponds to the phase of the sine curve. Point Q is another point on the
circumference under the same circular motion as point P . Here, point Q is ahead
of point P by a certain angle. On the sine curve, this difference can be viewed as a
phase difference corresponding to the time difference of t D .Q  P /=!.
Now consider a number of disks placed parallel in Fig. 6.10. Each disk is
connected with the neighboring ones. The disks are spinning at a constant angular
velocity as is the case of Fig. 6.9. The direction of the spinning motion is counter-
clockwise as viewed from the left. The waveforms drawn to the right of the disks
plot the vertical positions of the reference points, denoted by a star or circle on the
side of each disk. The coordinate points on the axis indicate the distance of each disk
measured from the rightmost disk. (x goes leftward on the disks and rightward on
the axis of the waveform.) The solid line represents the vertical position of the stars
and the dashed line represents those of the circles. The respective waveforms can be
viewed as a snapshot of the vertical positions of the corresponding reference point.
Analyze the behavior of the solid and dashed lines. Pay attention to the point
where the dashed and solid lines cross each other for the first time (the right of x3 ).
On the left of this point, the dashed line is above the solid line. In the region between
this point and the point where the dashed and solid lines cross each other for the
second time (in going along the x-axis to the right), the dashed line is below the solid
6.2 Interference and Interferometry 157

line. Since the disk spins counterclockwise as viewed from the left, the reference
point denoted by the circle is ahead of the star. This means that at a certain time
later, say dt .s/ later from the current time t .s/, the star will come to the point
in each disk where the circle is currently at. In the plot on the right of Fig. 6.10,
the solid line follows the dashed line in time. In other words, the spatial periodicity
shifts to the left as much as shown in Fig. 6.10 in dt . Such a shift per unit time is
called the phase velocity. Say the shift of the spatial periodicity is dx (the horizontal
distance between the peak of the dashed and solid waveforms is dx). Then the phase
velocity is found to be dx=dt. This correspondence between the space and time
domain shifts connected by the phase velocity is the basis of the conversion between
the temporal and spatial differentiations in wave dynamics expressed by Eq. (5.14).
Note that this spatial shift corresponds to the vertical shift of the reference point for
each disk, as the vertical arrow indicated at x D x1 -x3 in the figure. The vertical
shift is the change in the value of the oscillatory quantity that travels as a wave. In
the case of an electromagnetic wave, this vertical change represents the electric field
strength that varies as a function of time.

Complex Notation

The value of a sinusoidal function in the form of Eq. (6.64) is absolutely determined
if the amplitude and phase are given.8 This means that an electromagnetic wave in
the form of Eq. (6.64) can be fully characterized by a combination of the amplitude
and phase at a given point in space and time. On the other hand, we know a very
handy way to handle waves in this form; express it with a complex number. Denoting
the amplitude and phase of a wave A and , we can express a wave a as follows:

a D Ae i D Ae i.!tCkr/ (6.66)

where ! is the angular frequency, k is the propagation vector of the wave, and r is
the spatial coordinate variables r D x xO C y yO C zOz. The scalar product appearing in
the rightmost-hand side of Eq. (6.66) is the phase associated with the space variable
as explained in Chap. 5 [see Eq. (5.8)]. Figure 6.11 illustrates the complex notation
of an amplitude for a case where the wave is propagating in the positive z-direction.
With this notation, we can evaluate the intensity I of electromagnetic waves as the
magnitude of the complex number.

aa D Ae i A e i D A2 e 0 D A2 (6.67)

8
More strictly speaking, the initial phase 0 must be included in the expression as A sin.!t kx C
0 / to generalize the argument. However, in the present context, setting 0 D 0 does not change
the main point of the argument. To avoid complexity set the initial phase to 0 in the discussion
here.
158 6 Optical Interferometry and Application to Material Characterization

Fig. 6.11 Complex notation Im


of light field

A
f = wt − kz
Re

Ay z

Ax

Fig. 6.12 Light with rotating polarization

When the electric field has two spatial components, say the x- and
y-components, and each propagates at its own phase velocity, Eq. (6.66) must
be put in the following form:

a D ax xO C ay yO D Ax e i.!tCkx r/ xO C Ay e i.!tCky r/ yO (6.68)

where the subscripts x and y denote the respective components. Note that the
difference in the phase velocity is reflected in the propagation k as the frequency is
the same for the x and y components. As the two components propagate at different
speed, the orientation of the total amplitude vector a varies as the wave propagates.
For example, if the difference in the phase velocity creates the phase situation at a
certain point where the x component of the vector is zero [e.g., Ax cos.=2/
and the
y component is nonzero, the total vector a is oriented along the y-axis. A moment
later, the x component will have a nonzero value and the orientation of the total
vector is not lined up with the y-axis any longer. In this way, the total vector rotates
around the axis. The light wave whose amplitude stays in the same orientation is
called linearly polarized light, and the light wave whose orientation varies with its
propagation is called unpolarized light. Figure 6.12 illustrates the situation where
an unpolarized light propagates rotating its polarization.
6.2 Interference and Interferometry 159

a b
Im y
f = wt − kz f = wt − kz
A

Ay

Re Ax x
Ax Ay

Fig. 6.13 Circular polarization. (a) The phase difference of the x and y components of the
transmitted light complex plane rotating polarization; (b) the x and y components of the vector
representing the transmitted light

The phenomenon of rotating amplitude is observed when light passes through a


birefringent medium [1, 2]. The birefringence is the property where the refractive
index depends on the axis affixed to the material. A number of crystals have this
property, and the axis along which the phase velocity is faster than the other axis is
called the fast axis, and the other the slow axis. Consider a linearly polarized light
is incident to a birefringent crystal in such a way that the incident polarization is
45ı from both the fast and slow axes. If the crystal thickness (the distance that the
light travels within the crystal) is such that the overall phase difference between the
components propagating along the two axes is =2, the transmitted light will have
mutually the same amplitude with the phase difference of =2. The transmitted
light keeps this phase difference (as long as the medium after the transmission is
not birefringent). Consequently, the total amplitude vector will keep rotating around
the optical path. This type of light is called circularly polarized light. Figure 6.13a
illustrates in the complex plane the situation where the x component of a light wave
transmitted through a birefringent crystal is advanced by =2 in phase relative to
the y component. The amplitude of the total vector is the algebraic addition of the
real parts of the two components. Figure 6.13b illustrates the total vector of the same
moment as Fig. 6.13a in the xy plane of the actual space. As the x and y components
rotate counterclockwise in the complex plane at the same speed corresponding to the
phase velocity in air, the actual electric field vector rotates counterclockwise in the
xy plane. In the xy plane, the x and y components of the vector oscillate along the
respective axis with the phase difference of =2; when the x component is at its
positive peak, the y component is zero, : : :. A birefringent medium whose thickness
creates the overall phase difference of =2 is called the quarter-wave plate (because
=2 is a quarter of a period).
Now consider a linearly polarized light is incident to the same type of birefringent
crystal whose thickness is such that the overall phase difference between the light
160 6 Optical Interferometry and Application to Material Characterization

a Incident light b Transmitted light


s s

A A

a As As a

Af f Af f

Fig. 6.14 Half-wave plate. (a) The components of linearly polarized light vector incident to the
half-wave plate. The fast axis aligned to the x-axis and the slow axis is to the y-axis. Being linearly
polarized light, its fast-axis and slow-axis components are in-phase; (b) the components of light
vector transmitted through the half-wave plate at the moment when the slow-axis component is the
same phase as the incident light shown in (a). Since the fast axis is ahead of the slow axis by  in
phase, its component is flipped as compared with the incident light

components propagating along the fast and slow axes is  (in other words twice
as thick as a quarter-wave plate). Suppose the fast axis is aligned with the x-axis
and the slow axis is to the y-axis, and the incident light’s polarization is at angle
˛ to the slow axis as Fig. 6.14a illustrates. At the end of the travel through the
crystal, the component along the fast axis is ahead of the component along the slow
axis. Therefore, at the moment when the slow-axis component is the same phase
as the incident light shown in Fig. 6.14a, the fast-axis component is opposite to
the incident light (the same magnitude and opposite sign). Consequently, the total
vector is rotated by 2˛ from the incident light. Since the phase shift is  (rad),
the polarization of the transmitted light stays as shown in Fig. 6.14b; it is still linear
polarization. In other words, the birefringent crystal rotates a linearly polarized light
by angle 2˛. The situation is the same for any value of angle ˛. A birefringent
medium whose thickness creates the overall phase difference of  is called the half-
wave plate. A half-wave plate is widely used to turn a linearly polarized light for a
given angle by the mechanism illustrated in Fig. 6.14.
Birefringent media are used for various other useful applications such as phase
modulation and amplitude modulations. For further details, see [1, 2].
6.2 Interference and Interferometry 161

Fig. 6.15 Superposition of i(wt − kl1) i(wt − kl2)


two light waves A1e + A2e
P

i(wt − kl1) i(ωt − kl2)


A1e A2e

6.2.2 Interference

With the complex notation, we can discuss interference quite conveniently. Consider
two light waves are being superposed at a certain point of the space. Call the two
fields a1 and a2 . Assume that the two light waves have the same polarization.

a1 D A1 e i1 D A1 e i.!tkl1 / (6.69)


a2 D A2 e i2 D A2 e i.!tkl2 / (6.70)

At the point where the distance from the light source is l1 (m) for light wave a1 and
l2 (m) for a2 , the total field is the summation of the amplitude of the two waves as
follows:

a1 C a2 D A1 e i.!tkl1 / C A2 e i.!tkl2 / (6.71)

Therefore, the total intensity is

.a1 C a2 /.a1 C a2 / D a1 a1 C a2 a2 C a1 a2 C a1 a2

D A21 CA22 CA1 A2 e i.!tkl1 / e i.!tkl2 / CA1 A2 e i.!tk11 / e i.!tkl2 /

D A21 C A22 C A1 A2 e ik.l2 l1 / C A1 A2 e ik.l2 l1 /

D A21 C A22 C 2 cosŒk.l2  l1 /


(6.72)

Figure 6.15 illustrates the above situation where the two light waves are being
superposed with each other at point P . Consider that you vary l2 keeping l1 the
same by moving the origin of the second wave away from the point P . As the
difference ıl D l2  l1 changes over multiple wavelengths, the phase difference
ı D k.l2  l1 / D 2 ıl sweeps through periods of 2. This causes the total
162 6 Optical Interferometry and Application to Material Characterization

5
A1=A2
4.5
A1=0.8*A2
4

3.5
total intensity (au)

2.5

1.5

0.5

0
0 1 2 3 4 5 6 7 8 9
Δ φ (rad)

Fig. 6.16 Fringe visibility for two cases. Case one: the amplitude of the two light waves to be
superposed is the same. Case two: the amplitude of one of the light waves is 20 % lower than the
other

intensity to change ranging from the minimum corresponding to ı D .2N C 1/


(cos ı D 1) and the maximum corresponding to ı D 2N  (cos ı D 1) where
N is an integer. By superposing two light waves in this fashion, it is possible to
observe the relative phase difference associated with the relative optical path length
as an intensity variation. This mechanism of converting path length to intensity is the
essence of interferometry. Figure 6.16 plots the intensity variation as a function of
relative phase difference. The solid line is the case when the two light waves have the
same amplitude. The dashed line is the case when one of the light wave’s amplitude
is 20 % lower than the other. Notice that the former has higher maximum and lower
minimum than the latter. From the viewpoint of converting a phase difference to
intensity difference, the former is more favorable because the intensity differentiates
the phase difference more efficiently. To measure this efficiency, the fringe visibility
s is defined as follows [Eq. (6.73)]. For better visibility, it is important to make the
two laser beams’ intensity as close as possible to each other in actual interferometer
setups.

Imax  Imin
s (6.73)
Imax C Imin

Here Imax is the maximum intensity and Imin is the minimum intensity.
6.2 Interference and Interferometry 163

2
E1 + E2 = 2A2 {1 + cos (f1 − f2)} = 2A2 (1 + cos Δf)

end-mirror 1

4.5
4
3.5

intensity (au)
end-mirror 2 3
2.5
2
1.5
1
0.5
0
0 5 10 15
phase difference Δφ (rad)

Fig. 6.17 Photo detector signal varies from 0 to 4A2 as optical path differences go through fringes

6.2.3 Michelson Interferometer

The Michelson interferometer is one of the most widely used interferometers.


Figure 6.17 illustrates the principle of Michelson type interferometers. The laser
beam is split by a beam splitter into two paths, called the interferometric arms.
The two beams are reflected by the total reflector placed at the end of each arm.
The returning beams are recombined at the beam splitter and the interference can
be observed behind the beam splitter. In accordance with Eq. (6.72), the intensity
behind the beam splitter is at the maximum when the round-trip optical path lengths
along the two arms are the same or different by an integer multiple of 2 in
phase. Note that the laser beam experiences round trip in each arm, the condition
of the phase difference 2 behind the beam splitter corresponds to the arm length
difference of a half wavelength. The condition that the optical intensity is maximized
behind the beam splitter is referred to as the constructive interference or bright fringe
condition. Similarly, the optical intensity behind the beam splitter is at the minimum
when the round-trip optical path lengths are different by .2N C 1/. This condition,
referred to as the destructive interference or the dark fringe condition, is usually
more useful than the constructive interference condition because the intensity varies
around the minimum value approximately linearly in proportion to the arm length
difference. In accordance with Fig. 6.16, the minimum intensity becomes precisely
zero when the beam splitter splits the original intensity exactly 50 % for each arm.
When one of the end reflectors is slightly tilted, the optical path difference within
the width of the beam is observed as parallel fringes (Fig. 6.18).
164 6 Optical Interferometry and Application to Material Characterization

Fig. 6.18 The optical path y end-mirror 1


difference is observed as the
dark and bright fringes x

end-mirror 2

i(wt−kz1) if1
E1 (t, z1) = Ae = Ae

Vout
0 ≤ Vout ≤ 4A2

i(wt−kz2) if2
E2 (t, z2) = Ae = Ae

Fig. 6.19 Mach–Zehnder interferometer

Fig. 6.20 Mach–Zehnder i(wt−kz1) if1


interferometer. The optical E1 (t, z1) = Ae = Ae
path difference is observed as
the dark and bright fringes

Vout

i(wt−kz2) if2
E2 (t, z2) = Ae = Ae

6.2.4 Mach–Zehnder Interferometer

Figure 6.19 shows another common interferometer called the Mach–Zehnder


interferometer. The first beam splitter splits the incident laser beam and the second
beam splitter recombines the laser beams from the two arms. The two corner mirrors
fold the respective beams so that they are superposed at the second beam splitter.
Unlike the case of a Michelson interferometer, the laser beams do not make a
round trip in each arm. Therefore, the phase difference observed behind the second
beam splitter directly corresponds to the optical path difference. Like the case of
Michelson interferometer, it is possible to form fringes by tilting one of the corner
mirrors (Fig. 6.20).
6.3 Electronic Speckle-Pattern Interferometry 165

lz lf
1.22 ≅ 1.22
D D
= 1.22 lF

Fig. 6.21 Speckles (left) and the mechanism of formation. A number of rays fall into the
same pixel interfering with each other. In a given pixel, the resultant optical field is coherent
superposition of all rays falling into it. This determines the speckle’s phase

6.3 Electronic Speckle-Pattern Interferometry

6.3.1 Speckles

When a coherent beam is applied to a rough surface, multiple light rays reflected at
various points around a point on the object come to the same spot on the image plane
causing interference. Depending on the relative phases of these multiple waves, the
resultant intensity can vary from zero to any number. Consequently, a number of
dots of various intensities are formed on the image of the surface (Fig. 6.21). These
dots are referred to as the speckles. If the image is formed on the image plane, there
is point-to-point correspondence between the object plane and the image plane, as
indicated in Fig. 6.8. This means that the rays forming each speckle on the image
plane must originate from a local area representing the point on the object surface
that corresponds to the point on the image plane where the speckle is formed. Each
speckle is different from one another in the intensity because the surface condition of
the local area on the object is different from point to point. Therefore, each speckle
can be used to identify the corresponding local area on the object surface.
Resulting from superposition of coherent light rays, each speckle is characterized
by a certain amplitude and phase. Figure 6.22 illustrates the situation, where each
pair of Ai and i denotes the amplitude and phase of the complex number that
represents a ray being superposed with other rays to form a speckle. It is seen
that no matter how many rays may be added, the resultant complex number can
be expressed by a single amplitude and phase. This complex number represents the
sum of the distance that each ray forming the same speckle travels to reach the
point on the image plane. Therefore, a change in the complex number of a speckle
indicates the change in the distance between the local point on the object surface
and the corresponding point on the image plane. Hence, it can be used as a tool
to measure the displacement of the local point on the object surface. Note that the
phase of each speckle is a random value. It does not vary continuously from one
point to next on the object plane.
166 6 Optical Interferometry and Application to Material Characterization

Im
AN, fN A3, f3
A2, f2

A1, f1

Re
Ai, fi

Fig. 6.22 Phase of a speckle as superposition of optical paths of a number of light rays that form
the speckle

Now consider two speckle fields are formed by the same object on the same
image plane by a set of two light beams originating from the same laser source.
Since the two speckle fields are formed by the same laser source, speckles from the
respective fields are correlated. Consequently, a speckle in the first field A1 e i1 and
a speckle from the second field A2 e i2 are added algebraically as A1 e i1 C A2 e i2 ,
if they overlap each other. This results in a combination of an amplitude and phase
as A3 e i3 by the same mechanism as Fig. 6.22. If we can create a situation where
a displacement of an object point causes the phase of a speckle in the first speckle
field to decrease and the phase of a speckle in the second field that overlaps the other
speckle in the first field to increase, the relative phase between the two speckles,
3 , changes in proportion to the displacement. If the displacement is uniform on
the object surface, the relative phase of all those speckle pairs will change for the
same amount. If the displacement is not uniform, e.g., the object is stretched, the
relative phase change will vary from point to point on the image plane, depending
on the amount of the displacement of the corresponding point on the object surface.
If the relative phase change is an integer multiple of 2, the intensity will not
change. Therefore, if we subtract the superposition of the two speckle fields of
“before the deformation” from that of “after the deformation,” the intensity of those
area where the pair of speckles that satisfy the condition of “relative phase change is
an integer multiple of 2” becomes zero. This is the basic principle of the Electronic
Speckle-Pattern Interferometry or ESPI. Below we will discuss this principle more
quantitatively for in-plane displacement and out-of-plane displacement.
6.3 Electronic Speckle-Pattern Interferometry 167

Before going to the quantitative description of ESPI, let’s discuss the speckle
size. The speckle size is important from the viewpoint of applications because
normally the sensor is a digital device consisting of an array of pixels. The spatial
resolution of the phase information contained in speckles is determined by the size
of the speckles relative to the pixel. If the pixel size is much greater than the average
speckle size, each pixel cannot resolve the phase difference among the speckles.
The speckle radius can be approximated by the first minimum of Airy disk [1]
which is given by the following equation.

sin A  1:22 (6.74)
D
where A is the angle at which the first minimum occurs, measured from the
direction of incoming light, is the wavelength, and D is the aperture size. Being a
small angle, sin A  tan A . Thus the actual size s D z tan A can be approximated
by z sin A where z is the distance between the image plane and aperture of the lens.
In most cases, the distance between the lens and the image plane is approximately
equal to the focal length. In that case, the speckle size is given by the following
expression.

z
s D z tan   z sin   1:22 D 1:22 F (6.75)
D

where F  f =D is the F -number of the imaging lens.

6.3.2 In-Plane Displacement

Based on the argument regarding the phase of speckles, consider the optical
arrangement shown in Fig. 6.23. A laser beam is split into two beams, one of which
is reflected by the left mirror and the other by the right mirror in the figure. The
solid and dashed lines represent two rays within the width of the beam. The two
beams are striking the same point on the object surface, forming speckles on the
image plane, respectively. For the sake of explanation, let’s assume that the speckle
due to the left beam is a horizontally longer ellipse and that due to the right beam is
a vertically longer ellipse. Imagine that the point being struck by the two beams is
displaced by ı to the right. This causes the optical path from the source for the left
beam to be longer by ı sin ˛, and for the right beam to be shorter by ı sin ˛. Here ˛
is the angle of incidence common to the left and right beam. Let =2 be the phase
change corresponding to the optical path length change of ı sin ˛.

ı sin ˛
D 2 (6.76)
2

Let the phase of the speckle due to the left beam be left and that due to the
right beam be right . Letting the two beams have the same intensity of I , the
168 6 Optical Interferometry and Application to Material Characterization

δ
f
2
a

left right
f f before

d 2 2 iqleft iqright −iqleft −iqright
(e +e ) (e +e ) ⇒ cos (qleft − qright)
= +

f f f f
= + iqleft + iqright − −iqleft + −iq −
(e 2 +e 2 ) (e 2 + e right 2 ) ⇒ cos (qleft − qright + f)

after

Fig. 6.23 Principle of phase change detection due to in-plane displacement of a pair of speckle
fields

intensity of the total optical field due to the superposed speckles before and after
the displacement can be put as follows.
Before:
 
I e ileft C e iright e ileft C e iright D 2I C 2I cos.left  right / (6.77)

After:
  
I e i.left C 2 / C e i.right  2 / e i.left C 2 / C e i.right  2 / D 2I C 2I cos.left  right C /
(6.78)

Consequently, if the intensity of “before” is subtracted from “after” electronically,


the result becomes as follows:

2.left  right / C
2I cos.left  right C /  2I cos.left  right / D 4I sin sin
2 2
(6.79)

In Eq. (6.79), is proportional to the displacement via Eq. (6.76), and the phase
we are interested in finding. left and right are the phase of a speckle resulting
from superposition of a number of rays reflected off the object surface as indicated
by Fig. 6.22, and unknown. Therefore, we cannot find by solving Eq. (6.79).
However, the second term on the right-hand side becomes zero when =2 is equal
to an integral multiple of  no matter what left and right may be. From Eq. (6.76),
6.3 Electronic Speckle-Pattern Interferometry 169

Acoustic transducer

Object

Beam
expander
Reference surface

Laser

Fig. 6.24 Typical out-of-plane sensitive ESPI setup

the displacement that satisfies the condition of equal to an integral multiple of 


can be found as follows:

ı sin ˛
2 D m; m W integer (6.80)

which leads to


ıDm D mı0 (6.81)
2 sin ˛
When the object experiences deformation, the displacement is not uniform over
the object surface. In this case, those regions where the displacement is an integer
multiple of ı0 defined by Eq. (6.81) become black in the whole view of the electronic
image representing the subtraction of the before image from the after image.
These black regions are called “dark fringes” of a subtracted image, and present
the contour of displacement that satisfies condition (6.81). ı0 can be viewed as
representing the displacement corresponding to the case when m D 1 in Eq. (6.81).
From this viewpoint, ı0 can be interpreted as the minimum displacement we can
find by analysis of fringes. When the angle of incidence is 45ı and the wavelength
is 632.8 nm (helium neon laser), as an example, ı0 is equal to 447.5 nm.

6.3.3 Out-of-Plane Displacement

Out-of-plane deformation can be analyzed based on the same principle as the


in-plane case discussed above except for the interferometer setting. Figure 6.24
170 6 Optical Interferometry and Application to Material Characterization

illustrates a typical interferometer sensitive to out-of-plane displacement along with


a sample fringe pattern obtained with a vibrating metal plate. The interferometer
is basically a Michelson interferometer, where one end mirror is replaced with an
object to be analyzed (specimen) and the other end mirror is with a reference plate.
The specimen is oscillated with an acoustic transducer from the rear side. Behind
the beam splitter, the speckle field due to the reflection from the specimen and the
speckle field due to the reflection from the reference plate are superposed on the
image plane of a digital camera.

References

1. Hecht, E.: Optics, 4th edn. Addison-Wesley, Reading (2001)


2. Yariv, A.: Introduction to Optical Electronics. Holt Rinehart and Winston, New York (1971)
3. Griffiths, D.J.: Introduction to Electrodynamics, 3rd edn. Prentice Hall, Upper Saddle River
(1999)
4. Aitchson, I.J.R., Hey, A.J.G.: See Chapters 2 and 4 of Gauge Theories in Particle Physics. IOP
Publishing, Bristol/Philadelphia (1989)
5. Meinlschmidt, P., Hinsch, K.D., Sirohi, R.S. (eds.): Selected Papers on Electronic Speckle
Pattern Interferometry: Principles and Practice. SPIE Milestone Series, vol. 132. SPIE Optical
Engineering Press, Bellingham (1996)
6. Sirohi, R.S. (ed.): Speckle Metrology. Marcel Dekker, New York (1993)
7. Sciammarella, C.A., Sciammarella, F.M.: Experimental Mechanics of Solids. Wiley, Hoboken
(2012)
8. Griffiths, D.J.: Introduction to Quantum Mechanics, 2nd edn. Pearson Prentice Hall, Upper
Saddle River (2005)
9. Schiff, L.I.: Quantum Mechanics, 3rd edn. McGraw-Hill Kogakusha, Tokyo (1968)
10. Yardley, J.T.: Introduction to Molecular Energy Transfer. Academic, New York (1980)
11. Herzberg, G.: Molecular Spectra and Molecular Structure. Spectra of Diatomic Molecules, vol.
1, 2nd edn. Krieger Publishing Company, Malabar (1989)
Chapter 7
Experimental Observations

In this chapter, we will discuss experimental observations from various viewpoints


of the present field theory. Section 7.1 presents plastic deformation waves observed
experimentally in the velocity field of thin-plate specimens under tensile loads. The
oscillatory and decaying characteristics of the observed waves are argued based on
the field equations presented in Chap. 4 and their physical interpretations discussed
in Chap. 5. Section 7.2 discusses the vortex nature of the displacement field. The
boundary of a pair of experimentally observed vortexes in the displacement field
is interpreted as representing the shear force of the plastic regime in the form
of differential rotation r  ! as discussed in the earlier chapters. Section 7.3
focuses on the deformation charge and its behaviors in connection with dynamics of
plastic deformation and fracture. Optical interferometric patterns are interpreted as
representing the deformation charge under various loading and specimen conditions,
and the experimental observations are argued from the viewpoints of the present
field theory and conventional theories. Energy dissipation of plastic deformation
associated with the movement of plastic deformation charges is discussed and
its connections with acoustic emission and heat release are argued based on
experimental observations.

7.1 Plastic Deformation Wave

In Chap. 5, we discussed that in the plastic regime materials still possess recovery
force that leads to wave characteristics in the displacement field. What differentiates
the dynamics in the plastic regime from that in the elastic regime is that the longitu-
dinal effect is represented by energy dissipative force proportional to differential
velocity r  v, as opposed to elastic restoring force proportional to differential
displacement r  . The wave characteristics in the plastic regime is caused by
differential rotation that exerts shear force. Therefore, the displacement wave in

© Springer Science+Business Media New York 2015 171


S. Yoshida, Deformation and Fracture of Solid-State Materials,
DOI 10.1007/978-1-4939-2098-3__7
172 7 Experimental Observations

Tensile load

P1 Mirror

P2 α
α Beam splitter
P3 Beam 1

Beam 2
Mirror Mirror CCD camera

Laser

Computer TV monitor

Fig. 7.1 One-dimensional in-plane sensitive ESPI. P1 –P3 are monitoring points where displace-
ment is evaluated

the pure plastic regime is a transverse wave. The energy dissipation occurs when the
deformation charge .r  v/ flows under the influence of the velocity field v.
Yoshida et al. [1] reports an experimental study in which decaying, transverse
wave characteristics are observed in the velocity field u (the rate of displace-
ment component perpendicular to the tensile axis) in a thin-plate specimen being
deformed under a constant tensile speed. Figure 7.1 illustrates the experimental
arrangement used in this study. The optical setup is a standard, dual-beam ESPI
configuration [2–4] sensitive to horizontal, in-plane displacement. The specimen is
an aluminum-alloy (A6063) plate of 200 mm long (100 mm in effective length),
25 mm wide, and 1 mm thick. The grain size of this material is approximately
150 m. A tensile load is applied at a constant head speed ranging from 0:1 to
1 mm/min. At each time step, a specklegram (a full-field image of the specimen
formed by the superposition of the speckle fields of the two laser beams) is taken
with a CCD (charge coupled device) camera. Each specklegram is subtracted from
the specklegram obtained at a previous time step. The interval between the pair
of specklegrams for the subtraction is determined such that the number of fringes
formed by the subtraction is appropriate.
Once a fringe pattern is formed for each time step, the displacement on each dark
stripe (fringe) of the fringe pattern (uN ) is evaluated from the fringe order N and
the fringe separation calibrated to the actual displacement [ı0 in Eq. (6.81)]. The
values of u in between fringes are evaluated by interpolation. Figure 7.2 illustrates
a typical fringe pattern and the situation where the value of u is evaluated at grid
7.1 Plastic Deformation Wave 173

Fig. 7.2 Typical fringe


pattern formed with the
subtraction method

points by interpolation. In this fashion, u at the three reference points along the
central, vertical axis of the specimen (Fig. 7.1) are found at each time step.
Figure 7.3 plots the oscillatory behavior of u at the three reference points P1 –P3
indicated in Fig. 7.1 as a function of time. The loading characteristic recorded by the
tensile machine and the cumulative displacement of reference point P2 (called the
x-wave) are also plotted. The tensile machine’s cross-head speed for this experiment
is 0.1 mm/minD 1:67 m/s. The oscillation in u begins slightly before the loading
characteristic reaches the yield point (at 85 % of proof stress). This is consistent
with the field theoretical argument discussed in Chaps. 4 and 5 that the transverse
wave characteristics (in this case the wave characteristics in the velocity component
perpendicular to the tensile axis) begin as the deformation enters the plastic regime.
When an initially linear elastic material enters the plastic regime the entire specimen
breaks into local regions within each of which the dynamics can still be expressed
by linear elasticity. From the viewpoint of the global coordinates, this corresponds
to that the rotation is no more coordinate-independent, and therefore r  ! ¤ 0.
From the dynamical viewpoint, this is when the shear force represented by Gr 
! [Eqs. (4.81), (5.2)] becomes effective, which causes the wave characteristics in
the velocity-field to rise. The x-wave indicates that point P3 keeps shifting in one
direction until the load reaches the critical point corresponding to 85 % of proof
stress. When the stress reaches this critical point, a small scale fracture occurs so
that the material loses its shear stability. This causes the dynamics to enter the plastic
regime described by the field equations (4.61)–(4.64). Accordingly, the transverse
displacement-wave rises. The ESPI setup sensitive to the horizontal (x) component
of the dynamics detects it as the excitation of u-wave. The x-wave oscillates around
the new equilibrium established after the small scale fracture occurs, as seen in
Fig. 7.3.
A careful examination of the u-waves in Fig. 7.3 around the first peaks indicates
that there are phase delays among the three points. The peak at P3 appears first on
the time axis followed by the one at P2 , and the peak at P1 is behind the two. The
phase delays become less prominent around the second peaks, and there is no visible
phase delay around the third peaks. The dashed lines indicate the phase delays at
the respective peaks. This trend of decrease in the phase delay can be interpreted
as a reduction in the phase velocity. As discussed in Chap. 4 [see Eq. (4.86) p and
the paragraph that follows Eq. (4.86)], the phase velocity is equal to G=. It
174 7 Experimental Observations

0.6
u (mm/min), x (mm) and load (au)
0.4
failure
0.2

0
0 10 20 30
−0.2

−0.4

−0.6
Time (min)

u at P1 u at P2 u at P3

load x at P2 equilibrium

Fig. 7.3 Transverse displacement wave rising as the deformation enters the plastic regime

Fig. 7.4 Peak position as a 15


reference point
location (cm)

function of time
10

5
y = –11.01x + 163.9
0
13.5 14 14.5 15
time at peak (min)

is possible to interpret that the reduction in the phase velocity is caused by the
reduction in the restoring force (the sear modulus G) that the material experiences
as the plastic deformation develops.
From the phase delay observed around the first peaks, the phase velocity of the u-
wave at that time can be estimated. Figure 7.4 plots the times when the u-waves that
the five reference points (P1  P3 plus points between P1 and P2 , and between P2
and P3 ) reach their respective peaks. From the slope of this plot, the phase velocity
is estimated to be 110 mm/min.
Table 7.1 lists experimentally estimated phase velocities of plastic deformation
waves in aluminum alloys. The top three rows are the phase velocities of the
u-wave evaluated with the same method as in Fig. 7.3. The material and dimensions
of the specimen are the same as the case of Fig. 7.3 for all these three cases but the
7.1 Plastic Deformation Wave 175

Table 7.1 Plastic deformation waves observed under various tensile speeds
Tensile Phase
speed velocity
(mm/min) (mm/min) Period (min) Wavelength (mm) Material L (mm) PDW Ref.
0.1 110 5.4 572 A6063 161 u-Wave [1]
0.35 83 3.6 299 A6063 84.4 u-Wave [1]
1.0 118 3.5 413 A6063 117 u-Wave [1]
0.1 6.6 2.9 19.4 A85 50 Shear-wave [5]

tensile speeds are various. Notice that the observed phase velocities are similar to
one another, independent of the tensile speed. This is consistent with the argument
that
p the phase velocity is determined by the shear modulus and the density as
G=, independent of the tensile speed or strain rate. Also listed in this table is the
period of oscillation estimated from the time plot of u like Fig. 7.3. From the phase
velocity and the period, the wavelength can be calculated as shown in the table.
It is interesting to note that the calculated wavelength is approximately a quarter
of the effective length L of the specimen (the gauge length or the length of the
part between the tensile machines’ grips). This indicates that the specimen swings
horizontally pivoting around a point near one of the grips where the displacement is
minimum and the portion near the other grip swings most largely.
The fourth row of Table 7.1 summarizes the results of a similar experiment
conducted by Danilov et al. [5], in which the authors evaluate the shear strain field of
a tensile loaded aluminum alloy specimen using a speckle photographic technique
to measure the in-plane displacement field. The specimen size of 50  10  1:5 mm
is comparable to the other experiments listed in Table 7.1. They evaluate the shear
strain by numerically differentiating the displacement data obtained with the speckle
photographic experiment. The period of the shear wave is calculated from the phase
velocity and the wavelength of the shear wave reported in [5].
Since the period of a shear wave is the same as the displacement wave or the
velocity wave, it is possible to estimate the phase velocity of the velocity-wave
from the shear-wave’s period. Assuming that the specimen swings transversely in
the same fashion as above, i.e., the part near one of the tensile machine’s grips is
stationary and the part near the other grip swings most largely, the wavelength of the
velocity wave can be estimated to be 50 mm  4 D 200 mm (the specimen length
is a quarter of the wavelength). From this wavelength and the period of 2.9 min,
the phase velocity can be estimated to be 200 mm  2:9 min D 68:9 mm=min.
This value is on the same order of magnitude as the other three cases shown in
Table 7.1. The observation that the phase velocity of the velocity-wave observed in
aluminum alloys A6063 and A85 is similar to each other is reasonable because the
phase velocity is determined by the elastic modulus and the material density and
both of these quantities are similar in A6063 and A85.
Figure 7.5 plots a case where the u-waves observed at the three reference points
initially show an oscillatory behavior with a phase delay, but unlike Fig. 7.3, they
lose the oscillatory behavior before reaching the second peak and start to decay
176 7 Experimental Observations

0.12
failure
0.08

P1
u (mm/min)

0.04
P2
0
P3
−0.04

−0.08
0 5 10 15
time (min)

Fig. 7.5 Over-damping like u-wave

exponentially. It is interpreted that in this case the plasticity develops so rapidly


that after the first half period of oscillation, the wave shows an over-damping-like
behavior.
Figure 7.6 shows another case where developed charge-like patterns appear
repetitively starting at a relatively early stage of plastic regime. The charge-like
pattern initially appears at various locations on the specimen randomly in an
intermittent fashion. Around the beginning of the last one-third of the plastic regime
(around 25 min on Fig. 7.6), it starts to drift continuously from the top to the bottom
of the specimen. After the completion of the second round of this continuous top-
to-bottom drifting pattern, it becomes stationary near the vertical center of the
specimen, and stays there until the specimen fracture at that location about three
min later. The u-wave shows some correlation to this movement of the charge-like
pattern. It rises prior to the yield point, as is the case of Fig. 7.3. However, when the
charge-like pattern appears for the first time, the wave characteristic is interrupted.
When the charge-like pattern appears intermittently, the displacement component u
shows a somewhat oscillatory behavior. This can be interpreted as follows: as the
transversal displacement (x or the displacement component parallel to u) increases
from the neutral point (corresponding to the position prior to the yield point), the
stress reaches the critical point where the stress is concentrated at a certain point
of the specimen and the stress energy is released by the formation of the same
type of small-scale fracture as discussed in Fig. 7.3 (see the paragraph explaining
this figure). This causes the wave to decay, and the abrupt change in the strain
distribution resulting from the small-scale fracture is observed as the charge-like
pattern. Once the stress is released completely, the transverse displacement resumes
causing the u-wave to reappear. In other words, in the situation where the charge-like
pattern appears intermittently, the decaying wave characteristics in the transverse
displacement observed in Fig. 7.3 repeat as the charge-like pattern appears and
7.1 Plastic Deformation Wave 177

0.25
u at P1
0.15
u (mm/min) u at P2
0.05
u at P3
−0.05
Load
−0.15 (au)
−0.25
0 10 20 30 40
14
Charge location (cm)

12
10
8
6
4
2
0
0 10 20 30 40
Time (min)

Fig. 7.6 Discontinuous u-wave observed with the charge-like pattern

disappears. As this process repeats, the degree of plasticity increases, raising the
value of c or making the material more conductive in terms of the electrodynamic
analogy, as discussed in Chap. 5. This causes the amplitude of the intermittent
u-wave to decrease; in the last stage where the charge-like pattern drifts continu-
ously, the amplitude of the intermittent u-wave decays exponentially.

7.1.1 Decay Characteristics

Above, we discussed three types of plastic deformation waves; (a) Oscillatory-


decaying case, Fig. 7.3, (b) Over-damping case, Fig. 7.5, and (c) Intermittent case,
Fig. 7.6. In the case of (a, b), a developed charge-like pattern does not appear
till a late stage of deformation where the wave decays. In the case of (c), a
developed charge-like pattern appears intermittently, and the wave decays on each
appearance of the developed charge-like pattern. Apparently, the decay of the
wave characteristics and the appearance of the developed charge-like pattern are
correlated.
The correlation between the decay of the wave characteristics and the appearance
of the charge-like pattern can be discussed semi-quantitatively. In Chap. 5, we
discussed the decay of displacement wave in conjunction with the magnitude of
178 7 Experimental Observations

the charge density that causes the wave energy to dissipate. With the assumption
that r.r  v/ D 0 and that .r  v/ drifts without changing its shape, we derived an
expression of the velocity field in the following form:
c  1=2 !
 t G 2 c2
vp D e 2 cos k  2 t  kxs (5.59)
 4

Here vp is the amplitude of the transverse wave propagating in the positive


xs direction, as defined in Fig. 5.3. Experimentally, vp can be evaluated as the
temporal derivative of the xp component of the displacement field, or the component
perpendicular to the propagation of the transverse wave. Note that the velocity u
discussed in Figs. 7.2, 7.3, 7.4, 7.5 and 7.6 is the x-component of the velocity where
the x-axis is perpendicular to the tensile axis. In the ESPI setup shown in Fig. 7.1,
the x-axis is horizontal. c is defined by Eq. (5.51).

c D 0 .r  v/ (5.51)

Here, 0 is a material-dependent constant as discussed in Chap. 5 [see Eq. (5.49)


and the following paragraph]. Equation (5.59) indicates that under the condition
that the initial charge density .r  v/ is low, the spatiotemporal characteristic of
v (the rate of displacement in m/s) is a simple, exponentially decaying sinusoidal
wave. Equation (5.61) relates the decay time constant c and the charge .r  v/.

2 2
c D D (5.61)
c 0 .r  v/

In the context of the u-waves discussed in Figs. 7.3, 7.4, 7.5 and 7.6, the charge
r  v D @u=@x C @v=@y. Here, the y-axis is parallel to the tensile axis, or the vertical
axis in Fig. 7.1. In the present type of tensile load, the displacement component
perpendicular to the tensile axis is much smaller than that parallel to the tensile
axis. Thus, we can approximate the charge as @u=@x C @v=@y  @v=@y D d v=dy.
So, Eq. (5.61) can be rewritten as follows:

2
c D (7.1)
0 .d v=dy/

Equation (7.1) indicates that if 0 and .d v=dy/ are known, we can estimate the
decay time c . Below, we estimate the decay time for case (a)–(c).
0 is the ratio of the charge’s drift velocity Wd to the average velocity of the
particles inside the charge vs ; Wd D 0 vs . For the component parallel to the tensile
axis, Wdy D 0 vy . 0 is a material constant, as discussed in Chap. 5. In the case of
a developed one-dimensional charge, the average velocity v is on the order of the
tensile machine’s cross-head speed. So, from measurement of Wdy and the cross-
head speed, we can find 0 .
7.1 Plastic Deformation Wave 179

The value of .d v=dy/ can be estimated from the cross-head speed and the width
of .d v=dy/, or how long it spans along the y-axis. If d v=dy is the only normal
strain, the particle velocity at the tail of d v=dy is 0 and the particle velocity at the
head of it is the same as the tensile machine’s cross-head speed.1 If the specimen
experiences normal strain outside of .d v=dy/, the tail velocity is higher than 0 and
the head velocity is lower than the cross-head speed. However, in most cases, the
number of fringes outside .d v=dy/ is much lower than that inside .d v=dy/. So, it
is reasonable to estimate .d v=dy/ as the cross-head speed divided by the width of
.d v=dy/ along the tensile (y-) axis.
Experiments indicate that for typical aluminum alloy specimens under tensile
loading, the drift velocity Wdy is on the order of 10 cm/min for the tensile machine’s
cross-head speed of 0.35 mm/min (Fig. 7.12). Assuming that the particles’ average
velocity is on the order of the tensile’machines cross-head speed, this leads to 0
on the order of 10 (mm/min)/0.35 (mm/min) = 285. Toyooka et al. [6] have made
similar experiments on an aluminum alloy using the cross-head speed of 3:9 m/s
and found that the drift velocity of a developed charge-like pattern reduced from
1.8 mm/s at the first appearance in the initial stage of the plastic regime on the
loading curve to 0.5 mm/s in the middle stage of the plastic deformation or a halfway
point of the life-span of the specimen, and that the drift velocity does not decrease
any more. This leads to 0 ranging from 128–461, or an average of 294. So, it is
a good assumption that 0 for aluminum alloys to be approximately 300. For the
estimation of r  v D d v=dy, we can argue as follows:
First consider case (a). Figure 7.7 plots the oscillation at P2 to examine the
decay behavior closely. The dashed lines are exponential fits to the peaks of the
oscillation. It is seen that the wave decays exponentially. Also shown in Fig. 7.7 is
the appearance of the charge-like fringe patterns (Fig. 5.4). Notice that the charge-
like pattern starts to appear when the u-wave is about to lose the oscillatory
character; when the first charge-like pattern appears, the oscillatory behavior of u
is in the phase to swing upward in Fig. 7.7 but immediately after this point of time it
shows a sharp change in the direction followed by a monotonic upward trend. This
monotonic behavior of u can be interpreted as the final fracturing stage. The load
drops sharply in a synchronized fashion as the monotonic behavior of u.
Since a developed charge-like pattern does not appear till the wave decays, we
can assume that the charge d v=dy that causes the decay of the wave is uniformly
distributed over the entire specimen. Assuming that the particle’s velocity of d v=dy
is equal to zero at the tail and equal to the cross-head speed at the head, and using
the specimen’s effective length of 100 mm, the value of d v=dy can be estimated as
d v=dy D 1:67 (m/s)/100 (mm) = 1.67  105 .1=s/. Thus, c D 2=.300  1:67 
105 / D 400 s D 6:7 min. This decay time constant agrees with the experimental
value seen in Fig. 7.7.

1
An experiment indicates that when a developed, charge-like pattern is formed in a tensile load, the
material on one side of the pattern is translationally stationary (experiencing rigid-body rotation
only) and the material on the other side shifts at the same rate as the cross-head. See Fig. 7.19 and
discussion under Sect. 7.3.1 below.
180 7 Experimental Observations

15
u, load and location of charge (au)
10

0
0 5 10 15 20 25 30

−5 u at P2
load
charge
−10
trend
Time (min)

Fig. 7.7 u-Wave at point P2 decays exponentially

In case (b) observed in Fig. 7.5, the decay is exponential-like but faster than case
(a). From the fact that a developed charge-like pattern does not appear initially, as is
the case of Fig. 7.3, the faster decay can be explained as the value of d v=dy is higher
than case (a) but not as high as case (c) that causes the appearance of a developed
charge-like pattern. The time constant read from Fig. 7.5 is approximately 2.5 min.
This value is shorter than the case (a)’s time constant of 6.7 min, but of the same
order of magnitude.
The decay behavior of case (c) seems to be characterized by two time constants;
the short time constant associated with the fast decay due to the formation of a
developed charge, and the long decay representing the decrease of the oscillation
amplitude of the discontinuous wave. These time constants can be argued in the
same way with the use of different widths for d v=dy. The width of the developed
charge-like pattern that makes the wave discontinuous is typically of the order of
1 mm. Assuming that one developed charge appears at a time, we can set d v (the
differential velocity between the head and tail end of the charge-like pattern) equal
to the tensile-machine’s cross-head speed. This means that the value of d v=dy is
two orders of magnitude higher than case (a) where the same d v is distributed
over 100 mm (as opposed to 1 mm). Since 0 is the same as case (a), the time
constant is two orders of magnitude shorter than case (a); 400 s=100 D 4 s. This
time constant shows an order-of-magnitude agreement with the rapid decay of the
wave; in Fig. 7.6, the discontinuous wave resumes rising two or three times in 10 s,
i.e., the rise time is between 10=3 D 3:3 s and 10=2 D 5 s.
The slower decay of the discontinuous wave can be argued in the following way.
Figure 7.8 plots the u-wave at point P2 . The dashed lines are exponential fits like the
one for case (a) in Fig. 7.7. Since this decay represents the decrease in the amplitude
7.2 Vortex-Like Displacement Field 181

Shear-band location (au)


0.4
Exponential fit
0.2 Shear-band
u, mm/min

0.0

−0.2

−0.4
0 10 20 30 40
Time, min

Fig. 7.8 Discontinuous u-wave observed with the charge-like pattern; direct comparison

of the discontinuous wave (not the decay of the wave due to the formation of a
developed charge), we can assume that the charge is distributed over the entire
specimen, as is the case of Fig. 7.7. Thus, it is expected that the decay time constant
is similar to that of Fig. 7.7, 400 s. The exponential fit in Fig. 7.8 indicates a time
constant of approximately 10 min D 600 s, which is of the same order of magnitude
as case (a).
From experiments on developed charges discussed later under Sect. 7.3.1, 0 for
carbon steels is estimated to be an order of magnitude lower than aluminum alloys.
For the cross-head speed of 2:5 m/s (Fig. 7.15), the drift velocity is observed to
be 1  104 m/s and therefore 0 D 1  104 =2:5  106 D 40. Similarly, with
the cross-head speed of 25 m/s (Fig. 7.18), 0 D 4  104 =25  106 D 16.
At this time no plastic deformation wave data are available for carbon steel, and
these values cannot be verified by experiment. However, it is interesting to note
that aluminum alloys in general generate developed charge-like patterns more easily
than carbon steel. Especially, aluminum alloys with a few percent of magnesium in
the chemical composition form developed charge-like patterns conspicuously [7].
Pure aluminum usually does not form a developed charge-like pattern until the pre-
fracturing stage [8]. It is believed that the formation of shear bands is initiated at
the microscopic level resulting from interaction between dislocations and obstacles
such as solute atoms in solid solution or precipitates [9,10]. It is likely that the value
of 0 is determined by those microscopic interactions of defect with surrounding
atoms [11].

7.2 Vortex-Like Displacement Field

In Yoshida et al. [12], the rotational displacement is observed with the use of a
two-dimensional, in-plane sensitive ESPI arrangement illustrated in Fig. 7.9. Of
particular interest is the appearance of a pair of developed vortexes having mutually
182 7 Experimental Observations

Tensile load

Mirror
α
α Beam splitter
Beam 1
Beam 2
Mirror CCD camera
Mirror

Laser

Computer TV monitor

Fig. 7.9 Two-dimensional in-plane sensitive ESPI

opposite rotational direction in the acceleration field, and its connection with the
charge-like pattern. The left two images of Fig. 7.10a, b, are the two-dimensional
displacement fields .u; v/ at time steps 31 and 32 obtained from the same fringe
analysis explained above (see Fig. 7.2 and its explanation). The third image (c)
is the difference of (a, b). Since (a, b) are obtained by subtracting the pair of
interferogram at the respective time steps, the resultant vector fields represent
velocity (displacement over the time difference used to form the subtraction fringe
pattern). Thus, the vector field (c) represents the acceleration. The mutually opposite
vortexes observed in (c) therefore indicate that a pair of mutually opposite torques
(the clockwise torque in the upper region of the specimen and the counterclockwise
torque in the lower region), creating concentrated shear stress along the boundary.
As Fig. 7.10d indicates, a charge-like pattern appears immediately afterward. The
specimen in fact fractures along the boundary of the vortexes.
The above observation can be explained based on the field equations as follows.
The acceleration field seen in Fig. 7.10 can be interpreted as the first term on the
right-hand side of field Eq. (4.63).

1 @v
r ! D  j (4.63)
c 2 @t
According to Eq. (4.63), the acceleration is caused by the shear force represented
by Gr  ! appearing on the left-hand side. The clockwise and counterclockwise
rotational pattern shown in (c) can be interpreted as r  !. The negative sign on the
7.3 Observation of Charge-Like Phenomena 183

Fig. 7.10 Vortex-like displacement field observed in a metal-plate specimen under tensile loading

Fig. 7.11 Curl of ! causing


shear force along the
boundary of vortexes. Above
the boundary where the arrow
indicates the acceleration,
rotation ! is into the page
and below the arrow ! is out
of the page. The right
drawing illustrates the
direction of r  ! ∇×w

right-hand side of Eq. 4.63 indicates that the actual force (torque) is in the opposite
direction to r  ! indicated by the arrow in Fig. 7.11.

7.3 Observation of Charge-Like Phenomena

The charge-like pattern is first observed unexpectedly in a study of the displacement


field pattern in an aluminum alloy thin-plate specimen under monotonic tensile
loading [13]. Figure 7.12 shows several charge-like patterns observed in the study
of [13], along with its location on the specimen. The charge-like patterns are
observed in ESPI fringe patterns formed with the same subtraction method as
184 7 Experimental Observations

a b c d

e
500
Charge location

c
(pixel)

250

0
700 1100 1500 1900
Time (s)

Fig. 7.12 First observation of charge-like pattern. Cross-head speed = 0:35 mm=min D
5:83 m=s. The charge-like pattern drifts at approximately 10 cm=min D 1:67 mm=s

Fig. 7.2. The charge-like pattern starts to appear when the load reaches the yield
point. In an early stage, sometimes multiple patterns are observed (Fig. 7.12a).
The charge-like patterns at this stage are dynamic on the specimen. As the
deformation progresses, the pattern tends to move more slowly, and eventually
becomes completely stationary at a certain location of the specimen. The specimen
breaks along the stationary pattern.
Subsequent studies on this charge-like pattern indicate the following features.
(a) The pattern appears at the same time at the same location in the fringe patterns
formed by ESPI sensitive to in-plane displacement parallel and perpendicular to the
tensile axis. (b) In monotonic tensile experiment on a specimen free of initial stress
concentration, the pattern begins to appear near the yield stress, and is dynamic
along the specimen until it becomes stationary as mentioned above. (c) In monotonic
tensile experiment on a specimen with initial stress concentration (such as a notch,
through holes, and welding), the pattern tends to be stationary from the beginning.
(d) The dynamic pattern is accompanied by a stress drop and a stress recovery
immediately after the formation of the pattern. Consequently, the loading curve
shows the zigzag pattern known as the serration. (e) The pattern appears under cyclic
loading as well. (f) If the load is applied until fracture of the specimen, the fracture
line always overlaps the stationary pattern.
These features indicate similarity of the charge-like pattern with the phe-
nomenon known as the stretcher-strain, Lüders band (front), shear band or Portevin
Le-Chatelier (PLC) band. Below, studies on this similarity are discussed. The
relation to stress concentration will be discussed in the next chapter.
7.3 Observation of Charge-Like Phenomena 185

Lamp Upper Mirror


crosshead

Beam Helium neon laser


expander

CCD Mirror
camera
Video Beam
camera Glass Beam splitter
plate Mirror expander

Lower
crosshead Tensile/Cyclic load

Fig. 7.13 Simultaneous observation of charge-like pattern and Lüders band

7.3.1 Charge-Like Pattern and PLC Band

In Yoshida et al. [14], simultaneous observation of Lüders bands and charge-like


patterns is conducted. With the use of the setup shown in Fig. 7.13, the Lüders
bands are observed with the naked eye from the rear side of the specimen, while
the charge-like patterns are observed with the one-dimensional ESPI configuration
sensitive to in-plane displacement. The material used in this study is commercially
available structural steel SS400. Prior to the machining, the material is annealed at
900 ı C. The grain size is measured to be 15 m. After the annealing, the material
is machined to be two types of specimens; the first type is a bone shaped specimen
with a parallel part of 50 mm long, 10 mm wide, and 5 mm thick, and the second
type is a rectangle of 80 mm long (the grip-to-grip length), 15 mm wide, and 3 mm
thick. For the visualization of the Lüders band to the naked eye, the rear surface
of the specimen is polished with an abrasive paper of grit number 2000 [15]. The
front surface is first polished with an abrasive paper of grit number 100 and then
painted white for visualization of the charge-like pattern with a high contrast. The
rear surface of the specimen is illuminated with a fluorescent lamp, and the image
is recorded with a video recorder.
The ESPI setup used for the front surface of the specimen is a typical dual-beam
configuration [4]. Specklegrams of the specimen are taken with a CCD (Charged
Coupled Device) camera at a frame rate of 1 frame/s, and the image data are stored
into computer memory at the same rate. The tensile load is applied at constant cross-
head speeds. The applied load and the stroke of the dynamic grip is recorded at a
rate of 10 sample/s. The locations of the Lüders band and the charge-like pattern
are read from the images with rulers attached to the lower grip of the test machine
on both sides. The interferometric fringe patterns are formed by subtracting the
specklegram taken at one time step from the specklegram taken several time steps
before, which corresponded to the total elongation of 10–30 m (depending on the
cross-head speed).
186 7 Experimental Observations

Fig. 7.14 Observed Lüders band and charge-like pattern

Fig. 7.15 Locations of 10


Location from lower grip (cm)

charge-like patterns and 9


Lüders band observed
8
simultaneously. The
cross-head speed is 2:5 m/s 7
6
5
4 Lüders
3
Charge-like pattern
2
1 theory
0
0 100 200 300 400
Elapsed time (s)

Figure 7.14 shows the Lüders bands and charge-like patterns observed in the first
type of the specimen at the cross-head speed of 2:5 m/s. In this particular run, a pair
of Lüders band and charge-like pattern appear near the lower end of the specimen
and propagate upward. Except that the Lüders band and the corresponding charge-
like pattern are slanting in mutually opposite orientations because they are observed
from the opposite sides of the specimen, they propagate in a very similar fashion.
Figure 7.15 plots the locations of the centers of the Lüders band and charge-like
pattern measured from a reference point near the lower grip as a function of time.
The solid lines are the best fits to the data points. From the slopes of these lines, the
propagation velocities of the Lüders band and the charge-like pattern are found to
be 1:11  104 m=s and 1:13  104 m=s, respectively. Considering that the error
7.3 Observation of Charge-Like Phenomena 187

350

300 charge-like pattern


B
250
Stress (MPa)

200 300
charge Lüders
Lüders 250
150
200
A
100
150

50 100
−4.34E-18 0.002 0.004 0.006
0
0 0.01 0.02 0.03 0.04 0.05 0.06
Strain

Fig. 7.16 The stress–strain characteristic in the plastic regime recorded at the same time as
Fig. 7.14. The expanded view indicates the stress values when the Lüders front and charge-like
pattern begin to appear

associated with the reading of the locations of the Lüders line and charge-like pattern
is estimated to be ˙2 %, it is fair to say that the Lüders band and the charge-like
pattern propagate at the same velocity.
Figure 7.16 shows the stress–strain characteristic in the plastic regime along with
its expanded view near the yield point. This characteristic is recorded at the same
time as Fig. 7.14. The extents where the Lüders bands and charge-like pattern appear
are marked, respectively. From the observed yield elongation (region between points
A and B) and the specimen length of 50 mm, the yield strain L can be estimated to
be 0.0197. Using this yield strain, the cross-head speed Vc D 2:5106 m=s and the
relationship VL D VC =L derived by Sylwestrowicz and Hall [16], the theoretical
velocity of the Lüders front can be estimated to be VL D 1:27  104 m=s. This
value shows a good agreement with the measurement, as the dashed line in Fig. 7.15
indicates.
While the theoretical velocity of the Lüders band based on the yield strain agrees
with experiment, Fig. 7.16 indicates that the Lüders front or charge-like pattern does
not extend the whole span of the yield elongation. A possible explanation of this
observation is that while the plastic deformation front begins to propagate at the
lower end of the specimen at the yield stress and completes the propagation at the
upper end of the specimen at the end of the yield elongation, the initial and final
parts of the propagation are not observed as the Lüders band or charge-like pattern.
The fact that the charge-like pattern appears earlier and disappears later than the
Lüders band indicates that the former has higher sensitivity as an indicator of the
plastic deformation front. The expanded view in Fig. 7.16 shows that the charge-like
188 7 Experimental Observations

114 124 126 128 137 138 149 161 173 181 184 198

Fig. 7.17 Lüders front (upper) and charge-like pattern (lower) observed simultaneously at the
cross-head speed of 25 m/s. The Lüders’ front taken at time step 184 is not shown because of
poor quality

patten begins to appear at 8 % lower stress than the Lüders band.2 This observation
is consistent with the previous finding by Funamoto [17] for the same material that
the charge-like pattern begins to appear near a circular hole at a stress value 11 %
lower than the corresponding stretcher-strain.
At the cross-head speed of 25 m/s, a similar relationship between the Lüders
bands and charge-like patterns is observed. Figure 7.17 shows the Lüders bands
and charge-like patterns observed in the second type of specimen (the rectangular
specimen) at this cross-head speed. In this run, two pairs of Lüders band and charge-
like pattern begin to appear at the upper and lower end of the specimen, respectively,
within a time lag less than 30 s. At this cross-head speed, the Lüders bands show less
sharp edges, as is normally the case. The two pairs of Lüders band and charge-like
pattern propagate along the length of the specimen in mutually opposite directions.
Figure 7.18 shows the locations of the pairs as a function of time. Like Fig. 7.15, the
Lüders band and charge-like pattern of the same pair propagate at the same velocity,
which is reasonably close to the theoretical Lüders band velocity (dashed lines)
estimated in the same fashion as Fig. 7.15. (The lower Lüders band and charge-like
pattern show somewhat higher velocity than the theoretical value, and its reason
is not clear.) Unlike Fig. 7.15, however, both the Lüders band and the charge-like

2
Note that in Fig. 7.18 the charge-like patterns appear to begin later than the Lüders bands. This is
due to the fact that the view area of the CCD camera to take the specklegrams is smaller than the
view area of the video camera to take the Lüders bands and that the locations where the charge-like
patterns begin to appear are out of the view area of the CCD camera.
7.3 Observation of Charge-Like Phenomena 189

10 upper Lüders

Locations from lower grip (cm)


9 lower Lüders
8 upper charge
7 lower charge
6
5
4
3
2
1
0
50 100 150 200 250
Elapsed time (s)
Fig. 7.18 Locations of Lüders front and charge-like pattern as a function of time. The cross-head
speed is 25 m/s

pattern fluctuate around the fitted straight lines, indicating that when the Lüders
band forms a less sharp edge, it does not propagate at a constant velocity but
somewhat fluctuates around a mean value. The charge-like pattern observed under
this condition follows a similar, fluctuating trend, indicating that it still represents
the same phenomenon as the Lüders band. When tested at the cross-head speed of
2:5 m/s, the second type of specimen shows a straight trend similar to Fig. 7.15.
Based on this result, it is speculated that the fluctuating trend observed in Fig. 7.18
is not caused by the difference in the shape of the specimen but the difference in the
cross-head speed.
The fringe patterns observed in the regions on the specimen divided by a
charge-like pattern typically consist of approximately equidistant and horizontally
parallel fringes (see Figs. 7.14, 7.17, and 7.19). A system of perfectly equidistant
and horizontally parallel fringes observed in a horizontally sensitive ESPI setup
represents a rigid body rotation of the object [18]. This observation, therefore,
indicates that the deformation is concentrated at a Lüders band in such a way that
the regions divided by the band experience rigid-body like rotations [12]; i.e., there
is a stress concentration at the Lüders band so that the material on either or both
sides of the band experiences rigid-body like rotation.
Figure 7.19 shows fringe patterns formed by subtracting specklegrams obtained
at several time-steps from a common specklegram (frame #180). The number shown
underneath each fringe pattern denotes the number of the frames involved in the
subtraction. The fringe systems seen in these patterns have the following features.
The region above the charge-like pattern (region 1) shows nearly the equidistant,
horizontally parallel system representing a rigid-body like rotation. The region
inside the charge-like pattern (region 2) shows a dense fringes. These fringes are
190 7 Experimental Observations

Fig. 7.19 Charge-like pattern based on same before image

parallel to each other along the orientation of the developed, charge-like pattern
(along the xp axis in Fig. 5.3), and can be interpreted as representing a concentrated
strain normal to the charge-like pattern (normal to the xp axis in Fig. 5.3). The region
below the charge-like pattern (region 3) does not show a clear fringe structure. As
the number of the time-step increment increases in going from (a–f), the number of
fringes in region 1 increases rather constantly. The number of fringes in region 2
also increases, at a rate much higher than region 1. In going from (a–c), the fringe
density in region 2 becomes four or five times higher, while the number of fringes
in region 1 increases by one. The fringe pattern in region 3 remains more or less
the same in (a) through (f). These features can be interpreted as follows. When
the Lüders band is formed, the material in the vicinity of the band experiences a
localized deformation. As the time goes by, the part of the specimen held by the
dynamic grip (region 3) is displaced in the vertical direction as a rigid body (without
substantial deformation), forming no clear fringe structure. This motion sustains the
localized deformation in region 2. While this happens, the localized deformation
moves up by some mechanism as the other side of the material (region 1) rotates as
nearly a rigid body at an approximately constant rate.
The study in [14] leads to the following conclusions. (a) It confirms that the
charge-like pattern represents the same phenomenon as the Lüders band. It is
observed that the charge-like pattern appears at about 10 % lower stress than the
Lüders band, indicating that it has higher sensitivity than the Lüders’ band as an
indicator of the plastic deformation front. (b) Comparison of Figs. 7.15 and 7.18
indicates that the charge-like patterns drift faster when the tensile machine’s cross-
head speed is higher. This is consistent with the observation made by Mertens
et al. [19] that in their monotonic tensile experiment on an aluminum alloy the
propagation speed of the plastic deformation front varies in proportion to the loading
speed.
The mechanism that makes the drift velocity of charge-like patterns increase
with the cross-head speed has not been fully understood. In Chap. 5, based on the
7.3 Observation of Charge-Like Phenomena 191

energy-dissipation argument, we derived the expression that states that the drift
velocity W d of a positive charge is proportional to the particle velocity v via a mate-
rial constant 0 [W d D 0 v Eq. (5.49)]. With the assumption that the average veloc-
ity of the particle in the charge-like pattern is proportional to the cross-head speed,
this explains that a positive charge moves in the same direction as the cross-head.
In Fig. 7.18 the lower charge-like pattern moves in the same direction as the cross-
head. However, the charge-like pattern in Fig. 7.15 and the upper charge-like pattern
in Fig. 7.18 move in the opposite direction to the cross-head. A possible explanation
for those charge-like patterns moving opposite to the cross-head’s motion is that
they represent a negative charge d v=dy < 0 (y is the tensile direction and v is the
y-component of the velocity); as a small-scale fracture occurs3 on the formation
of the charge-like pattern, the fracture is concentrated on the side of the charge-
like pattern closer to the dynamics cross-head so that the rest of the material inside
the charge-like pattern recoils toward the static grip of the tensile machine making
a pattern of v.=dy < 0 (i.e., compression or the material closer to the small-
scale fracture recoils the most). The ESPI setup used in this experiment is not
sensitive to the direction of the displacement and therefore whether a given parallel
fringe pattern represents compression or stretch is unknown. It is also possible
that the movement of the charge-like pattern is governed by a completely different
mechanism and a positive charge can move opposite to the particle’s velocity.
However, if a one-dimensional, positive charge moves opposite to the particle’s
velocity, as indicated in Fig. 5.4, it causes the material to gain momentum. It is
unlikely to occur as a spontaneous process. Further study on the mechanism of the
movement of charge-movement is definitely necessary.

7.3.2 Cyclic Loading

Similar relationship between the Lüders-band like strain concentration (called the
shear-band)4 and the charge-like interferometric band-pattern has been observed
under cyclic loading with the use of the experimental setup shown in Fig. 7.13 [20].
The specimen used in this study is a 90 mm long (effective length), 15 mm wide,
and 3 mm thick rectangular carbon-steel (SS400) plate. The cyclic load is applied
with a load ratio5 R D 0 at frequency of 1 Hz. For the purpose of expediting the
deformation in the elastic regime, a monotonic load is applied up to 1,000 kgw.
Then the cyclic load with R D 0 is initiated at 1 Hz. After four cycles, the load is
increased by 4 kgw. In this pattern, the cyclic load is increased until the specimen

3
As we discussed in Fig. 7.3.
4
The strain concentration observed with the naked eye under the cyclic loading appears exactly the
same as the tensile loading case. However, since the terminology Lüders band is usually used to
mean strain concentration under tensile stresses, the pattern is referred to as the shear-band.
5
R is defined as the ratio of the minimum stress min to the maximum stress max as R D min =max .
In the present case, R D 0 means that no compressive stress is applied to the specimen as the
applied tensile stress is increased.
192 7 Experimental Observations

Crosshead position (r.u) 2

1.8

1.6

1.4

1.2

1
804 806 808 810 812 814
time (s)

Fig. 7.20 Cyclic loading pattern. Plus symbol indicates the cross-head position where a speckle-
gram is taken

fractures. When the load setting is increased after each set of four cycles (called the
set), there is a slight time lag before the load is actually applied to the specimen.
Therefore, the interval between the last cycle of a given set and the first cycle of the
following set is slightly longer than 1 s, whereas the interval between consecutive
cycles within the same set is exactly 1 s. On the other hand, the CCD camera takes
the image at the exact frame rate of 1 frame/s. Consequently, in each frame the
relative position of the cross-head in its sinusoidal pattern varies slightly. Figure 7.20
illustrates this situation for a time window in which pairs of the charge-like patterns
and shear-bands are observed. In the figure, the C mark denotes those positions of
the cross-head at which the CCD camera takes images.
Figure 7.21 shows the stress–strain characteristic resulting from the cyclic
loading. When the maximum (peak) stress reaches 330 N/mm2 , the stress–strain
characteristic begins to show a plateau similar to the yield plateau commonly
observed under tensile loading. This is when a pair of shear-bands appears on the
rear side of the specimen; one near the top end of the specimen and the other near
the bottom end. On appearing at the respective locations, the two bands begin to
propagate in mutually opposite directions towards the center of the specimen (see
the lower half of Fig. 7.22). On the front side of the specimen, a pair of charge-like
patterns appears somewhat closer to the center, and begin to propagate towards the
center of the sample at the same velocity as the shear bands (see the upper half of
Fig. 7.22).
Figure 7.22 shows the propagations of the charge-like patterns and the shear-
bands toward the center of the specimen. The images are arranged in such a way
that vertically aligned pairs of the charge-like patterns and the shear-bands have
similar distances between the respective upper and lower components. (Compare
the distances between the arrowheads inserted in the vertically aligned images.
Note that the charge-like pattern images and shear-band images are presented at
7.3 Observation of Charge-Like Phenomena 193

plateau
450
400
350
stress (N/mm2)

300
250
200
150
100
50
0
0 5 10 15 20 25 30
strain (%)

Fig. 7.21 Stress–strain characteristic under cyclic loading

Fig. 7.22
Shear-bands/charge-like
patterns pairs showing similar
distances. The charge-like
patterns appear at the
boundary that separates the
fringe pattern into different
systems. The number shown
below each pair indicates the
elapsed time when the
charge-like patterns (the left
number) and the shear bands
(the right number) appear.
Note that the charge-like
patterns pair always precede
the shear-band pair

838/878 s 847/886 s 859/894 s 860/900 s

different magnifications.) The numbers at the bottom denote the elapsed time from
the beginning of the loading. It is clearly seen that the charge-like patterns and shear-
bands are formed in the same orientation and propagate at the same rate, indicating
that they represent the same physical event. The charge-like pattern pair continues
194 7 Experimental Observations

10
upper shear-band
9 lower shear-band
8 upper charge
7 lower charge
locations (cm)

6
5
4
3
2
1
0
800 820 840 860 880 900 920
time (s)

Fig. 7.23 Drifts of the charge-like patterns and the shear-bands observed under cyclic loading

propagating until they merge with each other near the center of the specimen.
Similarly, the shear-band pair continues propagating until they merge with each
other at the same location as the charge-like pattern pair merge. During this period,
the strain-stress characteristic shows the yield-plateau like trend. In Fig. 7.21 the
beginning and the end of the propagation of the shear-bands are marked with vertical
lines.
As seen in Fig. 7.22, the fringes between the charge-like patterns are essentially
equidistant and horizontally parallel. Equidistant, horizontally parallel fringes
formed by an ESPI interferometer sensitive to horizontal displacement represent
rigid-body rotation around an axis normal to the plane [18]. It is interpreted that
the regions between the charge-like patterns experience rigid-body rotations in
such a way that the regions separated by a charge-like pattern rotate in opposite
directions (if one rotates clockwise the other counterclockwise). This interpretation
is consistent with the previous one that a charge runs along the boundary of two
regions rotating in opposite directions so that the boundary has high shear stress
(see Fig. 7.19 and the associated discussion).
Figure 7.23 shows the locations of the charge-like patterns and the shear-bands
as a function of the elapsed time. The top two data sets having a negative slope are
the traces of the charge-like pattern and the shear-band that begin to appear near
the top end of the specimen propagating downward. The other two data sets are the
charge-like pattern and the shear-band that begin to appear near the bottom of the
specimen propagating upward. The straight lines are the best fit to the respective
data sets. The slopes of these lines indicate the mean propagation velocity.
7.3 Observation of Charge-Like Phenomena 195

6
crosshead

displacement of crosshead and


5 upper shear-band

shear-band (cm)
4

0
800 820 840 860 880 900 920
time (s)

Fig. 7.24 Positions of cross-head and shear band as a function of time

Two arguments can be made in Fig. 7.23. First, the charge-like patterns propagate
ahead of the corresponding shear-bands. On the horizontal scale, it is observed that
the charge-like pattern appears approximately 40 s earlier than the corresponding
shear-band at a given location. From the recorded stress–strain diagram (Fig. 7.21),
the time lag of 40 s is equivalent to an increase in the mean stress (the mean value
of the maximum and minimum stress) of 1.5 N/mm2 . Since the mean stress in
this period is about 170 N/mm2 , this means that the charge-like pattern appears at
approximately 1:5=170 D 0:9 % lower stress than the corresponding shear-band.
The fact that the charge-like pattern appears at a lower stress than the shear-band is
consistent with the above analysis under the monotonic tensile loading case in which
the same material is used. However, in the monotonic tensile case, the charge-like
pattern appears at about 10 % lower stress than the Lüders band. It is unclear at
this point why the difference in the onset stress between the charge-like pattern and
shear-band in the cyclic loading case (0:9 %) is an order of magnitude smaller than
the monotonic loading case (10 %).
The second argument associated with Fig. 7.23 is that the data points fluctuate on
the time scale. In the case of the monotonic loading (Figs. 7.15 and 7.18), the trend
of the equivalent data points is straight (except for the slight fluctuation observed in
the fast tensile speed case of Fig. 7.18). As mentioned above, the trend is consistent
with the observation by Mertens et al. [19] that the propagation speed of the plastic
deformation front increases with the loading speed. It is possible that the observed
temporal fluctuation of the shear-band is in proportion to the cyclic displacement
field on the specimen. Figure 7.24 plots the displacement of the upper shear-band
relative to its initial position and the displacement of the cross-head relative to
its position at the moment the shear-band appears. Here the cross-head position
means the position of the dynamic grip of the cyclic test machine, and therefore
its displacement represents the dynamic, total displacement field on the specimen.
196 7 Experimental Observations

0.5 crosshead
0.4 upper shear-band
Shear-band location and
crosshead position (r.u.)
0.3
0.2
0.1
0
−0.1
−0.2
−0.3
−0.4
−0.5
800 810 820 830 840 850 860
time (s)
Fig. 7.25 Deviation of positions of cross-head and shear band from the mean values as a function
of time

Clearly, the shear-band follows the cross-head, i.e., the motion of the shear-band is
proportional to the cyclic displacement field on the specimen. To examine this effect
more precisely, Fig. 7.25 compares the deviation of the upper shear-band position
from the fitted straight line and the cross-head position relative to its mean value on
the same time scale (the deviation from the center of oscillation, see Fig. 7.20). Clear
correlation is observed, indicating that the shear-band follows the cross-head on a
moment-by-moment basis. The lower shear-band shows similar correlation with the
cross-head, although it is somewhat lower than the upper one.
The above observations indicate that the charge-like pattern formed under cyclic
loading behaves similarly to those formed under tensile loading. The charge-like
pattern begins to appear at a lower stress level than the shear-band, indicating that
it has higher sensitivity than the shear-band as a sensor of the stress concentration,
as the charge-like pattern is more sensitive than Lüders band under tensile loading.
Dynamic behaviors of the charge-like pattern under cyclic loading are similar to
the shear-band, indicating that both represent the same physical event. Moreover,
the fact that both move in coherent with the movement of the test machine’s
cross-head indicates that the way the stress concentration represented by the shear-
band is formed under cyclic loading is similar to that under tensile loading.
Under tensile loading, it is widely accepted that the Lüders band or PLC band
represents the plastic deformation front. It is also widely accepted that low-level
cyclic loads induce fatigue in the material. These observations altogether indicate
the fundamental similarity between plastic deformation and fatigue as a physical
process. Hasegawa et al. [21] have conducted an ESPI based research to study the
effect of fatigue on plastic deformation. They apply various fatigue loads to the same
7.3 Observation of Charge-Like Phenomena 197

a b c
Crack
I

III
IV

Fig. 7.26 Charge-like pattern divides plastic and elastic zones. (a) ESPI fringe pattern represent-
ing near-crack displacement field, (b) I–III elastic zones, IV plastic zone, and (c) another case
where the boundary of the plastic zone is more prominent

type of specimen prior to a monotonic tensile load and found that the fatigue load
promotes local plastic deformation. Description of the fatigue process based on the
field theory is definitely an important future subject.

7.3.3 Experiment with Notched Specimen

The charge-like pattern shows an interesting behavior when fast tensile loads are
applied to notched specimens [22]. Figure 7.26 shows typical fringes observed when
a fast tensile load (the cross-head speed of the order of cm/s) is applied to a notched
carbon-steel specimen (S50C) of approximately a 70 mm  70 mm square with
thickness of 13 mm. A bright ring-like pattern is observed at the boundary of fringe
systems that apparently have different pattern; one is an essentially equidistant
vertically parallel system and the other is a system consisting of curved fringes.
As will be discussed in the next chapter, in accordance with the present field theory,
the former is interpreted as representing elastic deformation and the latter plastic
deformation. Based on this interpretation, regions I–III: are elastic zones, and region
IV is a plastic zone. It is interesting to note that the situation that the charge-like
pattern divides the elastic and plastic zone is the same as the situation that the
Lüders band is located at the plastic front, separating the plastically deformed and
still elastic regions.
Figure 7.27 shows the change in the diameter of the ring-shaped charge like
pattern observed in the same experiment as Fig. 7.26. Here the horizontal axis
is the elapsed time from the reference time when the fringe pattern Fig. 7.28(1)
is observed. The ring diameter is seen to increase for the first 15 s or so, stops
increasing at that point, and remains the same until the specimen fractures.
Figure 7.28 are the fringe patterns at four different stages labeled (1–4) in Fig. 7.27
where (3) is when the ring diameter stops increasing. The phenomenon that the
ring diameter stops increasing can be interpreted as the charge-like pattern stops
198 7 Experimental Observations

Diameter Size Vs. Time


180
(3)
160
Diameter Size (px)
(2)
140
(4)
120

100

80
(1) ImgA009600-ImgA009640 : 0 sec

60
0 5 10 15 20 25 30 35 40
Time (seconds)

Fig. 7.27 Change in diameter of ring charge-like pattern

moving. From this viewpoint, it corresponds to the situation where the band-shaped
charge-like pattern stops moving in Fig. 7.12 before fracture. Note that the specimen
fractures soon after the ring-shaped pattern stops moving. This is consistent with
the field theoretical fracture criterion discussed in Chap. 8 that when a charge-like
pattern stops moving the material loses the mechanism to dissipate energy and
therefore it fractures.
It is worth discussing the ring-shaped charge-like pattern from the viewpoint
of conventional mechanics. For this purpose, a finite element analysis based on
continuum mechanics has been conducted. Figure 7.29 shows the Von-Mises stress
distribution (top left), the rate of the vertical component of the displacement (top
right), and the temporal trend of the diameter of the maximum Von-Mises stress
(bottom right) [23]. The bottom left image is the experimental (ESPI) fringe
pattern that the finite element analysis models. As the top left figure indicates,
the charge-like ring overlaps the maximum Von-Mises stress (the isosurface of the
highest Von-Mises stress). This is consistent with the fact that the Lüders band,
which overlaps the linear charge-like pattern, represents the plastic deformation
front where the tensile stress causes that particular location to yield. The top-right
fringe pattern resulting from the finite element analysis qualitatively agrees with
the experimental fringe pattern shown at the bottom left. The computed temporal
trend of the maximum Von-Mises stress shown in the bottom-right figure shows the
saturating behavior around the same time as the experiment. Since the finite element
analysis uses conventional elasto-plastic model, there is no way to simulate the
effect after the maximum Von-Mises stress stops increasing. Of course it is possible
to manipulate the constitutive relationship to match up with the experimental trend
of the ring diameter, but since such a model is not based on a deformation theory, it is
not meaningful to do so. Thus the computation is terminated as shown in this figure.
7.3 Observation of Charge-Like Phenomena 199

(1) (2)

(3) (4)

Fig. 7.28 Ring charge-like pattern observed in a specimen with a notch under fast tensile loading

Further theoretical development regarding the dynamics of the stress concentration


represented by the charge-like pattern is necessary.

7.3.4 Acoustic Emission

It is accepted that the Lüders band is a type of the shear band known as the PLC
band. It is known that the formation of a PLC band is accompanied by an abrupt
stress drop known as the serration. The coincidence of the charge-like pattern and
the Lüders band or shear band indicates the likeliness that the formation of a charge-
like pattern is accompanied by a stress drop. Moreover, it is considered that such
an abrupt stress drop is accompanied by a partial breakage of the material, which
indicates the possibility that the formation of a charge-like pattern is somehow
correlated to acoustic emission. In [24], simultaneous observation of charge-like
patterns and acoustic emissions is attempted with the use of the same type of the
ESPI setup as Fig. 7.1. The specimen used in this study is an aluminum alloy
(A6063) thin plate of 100 mm in effective length, 25 mm in width, and 1 mm in
thickness. A pair of identical acoustic sensors are attached to the two ends of the
specimen to which a tensile load is applied at a constant rate of 1 mm/min. The
location of acoustic emission is estimated from the difference in the time at which
the respective sensors detect the signal. An in-plane sensitive ESPI is arranged in
front of the specimen to take images to find the formation of charge-like patterns.
200 7 Experimental Observations

Isosurface von Mises stress Isosurface y-velocity

0.06

0.04

0.02

0
0 0.02 0.04 0.06 0 0.02 0.04 0.06

Diameter Size Vs. Time, COMSOL


30
ESPI
COMSOL 200 MPa yield
25 COMSOL 350 MPa yield
Diameter size (mm)

20

15

10

0
85 90 95 100 105 110 115 120 125 130 135
Time (seconds)

Fig. 7.29 Finite element modeling of displacement around a notch. In the bottom-right figure, the
plots label “ESPI” is the same as Fig. 7.27

Figure 7.30 shows the locations of the acoustic emission estimated in this way
on the same time axis as the locations of the charge-like patterns observed by the
ESPI. This particular specimen shows charge-like patterns at three locations at three
different times. The patterns are slant band-like patterns running at a certain angle to
the tensile axis. The vertical error bars in Fig. 7.30 indicate the locations of the left
end and the right end of each slant pattern. The data acquisition rate of the acoustic
emission is lower than that of the charge-like pattern. Therefore, for each data point
of acoustic emission, multiple charge-like patterns can be detected. The horizontal
error bars for the acoustic emission data points indicate the uncertainty due to the
fact that the data acquisition time for the acoustic emission is longer than the ESPI.
Within these vertical and horizontal margins of error, the formations of the charge-
like pattern and acoustic emission are coincident. This strongly indicates that when
a charge-like pattern is generated, a certain breakage of atomic bonding takes place,
emitting an acoustic wave.
7.3 Observation of Charge-Like Phenomena 201

500

400
Location (pixel)

Charge-like pattern
300
Acoustic emission
200

100

0
0 10 20 30 40 50
Data #

Fig. 7.30 Simultaneous observation of charge-like pattern and acoustic emission

7.3.5 Temperature Rise Due to Plastic Deformation

In Chap. 5, we derived the following equations:

Gj D Wd .r  v/ (5.45)
W d D 0 v (5.49)
c D 0 .r  v/ (5.51)

Equation (5.45) states that the longitudinal effect in the plastic regime is the flow
of the charge .r  v/. Equation (5.49) indicates that the drift velocity W d is
proportional to the particle’s velocity v and the constant of the proportionality 0
rates the degree of energy dissipative nature of the material; greater the value of
0 > 1, the more energy dissipative the material is. Finally, under the condition that
r.r  v/ D 0 and the pattern of .r  v/ unchanged, the degree of energy dissipation
can be expressed with c . From these equations, the flow of charge .r  v/ can be
put in the following form:

Gj D W d .r  v/ D 0 v.r  v/ D c v (7.2)

Being proportional to velocity v, Gj D c v can be interpreted as a velocity


damping force. Indeed, in the wave equation governing v, c is the coefficient of the
damping term. The quantity c can be interpreted as corresponding to the electric
conductivity.

@2 v @v
 2
 r 2 v C c D0 (5.53)
@t @t

From the above argument, it is expected that the flow of charge .r  v/ generates
heat as the mechanical energy is dissipated. In [25] temperature rise associated with
202 7 Experimental Observations

load
Upper
crosshead Mirror
Beam expander Laser

thermistor 2
thermistor 1 Mirror

CCD
Beam splitter
Beam expander

Mirror
Lower crosshead

Fig. 7.31 Experimental arrangement to study temperature rise

a flow of charge is investigated with the use of an ESPI setup sensitive to horizontal,
in-plane displacement (Fig. 7.31). The specimens tested are 0.2 mm-thick aluminum
and brass plates of approximately 20 mm wide and 100 mm long. The specimen
has a notch on a side near the vertical center so that fracture necessarily occurs
there. Two thermistors are attached; one near the notch and the other near the top
of the specimen by the grip of the tensile machine. A CCD (charge coupled device)
camera captures the whole image of the specimen at a preset constant interval, and
the image data are sent to a computer where image subtraction is made for fringe
formation. Temperature data collected by the thermistors are sent to the computer
via a data acquisition board along with the load and elongation data collected by the
tensile machine. The tensile machine is operated at a constant cross-head speed of
20 m/s.
Figure 7.32 shows typical fringe patterns obtained before and after the crack
starts to grow from the notch. The horizontal fringes observed above and below
the horizontal line passing though the notch indicate that the specimen experiences
bodily rotation (clockwise/counterclockwise above/below the notch). A pair of
linear charge-like patterns forming a v-shaped pattern sharing the vertex [circled
in (b)] begin to appear when the crack tip begins to grow. Interestingly, a point-like
pattern appears near the vertex, slightly away from the crack-tip. This point-like
pattern leads the crack growth as the linear charge patterns drift ahead of it at a
constant velocity (c). A consideration based on the present field theory indicates
that this situation satisfies the pre-fracture condition. This observation will be
discussed in the next chapter in conjunction with the field theoretical criteria of
plastic deformation and fracture.
Figure 7.33 compares snapshots of fringe patterns of the brass and aluminum
specimens. The high intensity regions highlighted with circles are where the fringe
density is much higher than the surrounding area indicating strain concentration.
7.3 Observation of Charge-Like Phenomena 203

Fig. 7.32 Charge-like pattern leading a crack-tip. Typical fringe patters before (a) and after
(b, c) the crack begins to grow. The wedge-shaped fringe pattern circled in (b) is considered to be
a deformation charge. Four images in (c) show a similar charge observed in a different experiment
indicating that it propagates to the right as the crack grows

Fig. 7.33 Spot-like charge in


brass and line-like charge in
aluminum

The strain concentration is spot-like in the brass specimen and line-like in the alu-
minum specimen. Figure 7.34 plots the temperatures measured with the thermistors
and the load measured with the tensile machine as a function of the elapsed time.
The arrows in Fig. 7.34 indicate when these patterns start appearing. In the case of
brass, the spot-like pattern moves along with the tip of the crack at a few mm ahead
of the crack tip. In the case of aluminum, the pattern extended as the load increases
until it reaches the maximum point and the specimen fractures. This difference
204 7 Experimental Observations

Fig. 7.34 Variation of temperature and load as a function of elapsed time

indicates the fact that brass is more ductile than aluminum. The strain concentration
can be interpreted as the current of deformation charge, analogous to the electric
conduction current. Thus, it is expected to accompany a temperature rise.
The temporal variations in temperature and load are recorded as the tensile load
is applied at a constant cross-head speed of 20 m/s. Figure 7.34 shows typical data
for the brass specimen (top) and the aluminum specimen (bottom). The temperature
data measured near the notch appears more oscillatory than those measured near
the upper grip of the tensile machine. This is because the thermistor used near the
notch has a higher sensitivity than the one used near the top of the specimen. A total
of 25 specimens are tested. Thorough analysis of the data obtained with all these
specimens results in the following findings.
1. In all 25 specimens tested, commonly to the brass and aluminum, a crack starts to
grow from the notch when the load starts to decrease after reaching the maximum
value.
2. The load-decrease after the maximum is more gradual in brass than aluminum,
indicating that brass is more ductile than aluminum.
3. In all 25 specimens tested, commonly to the brass and aluminum, the temperature
near the notch is higher than near the grip, indicating that the notch region is
closer to the heat source than the top region of the specimen.
4. In most cases, a temperature decrease is observed at the beginning of loading.
7.3 Observation of Charge-Like Phenomena 205

5. In all the aluminum specimens except for one case, the temperature starts to
rise when the crack starts to grow. In the only exceptional case, the temperature
rise is prior to the initiation of the crack growth, i.e., the temperature rises while
the load is still rising, as is the case of the brass data in Fig. 7.34.
6. In all brass specimens, on the other hand, the beginning of the temperature rise
is prior to the initiation of the crack growth, as seen in Fig. 7.34.
The above findings regarding the temperature changes can be interpreted as
follows. The initial temperature decrease [finding (4)] is apparently due to the
phenomenon known as the thermo-elastic effect [26]. As discussed in Sect. 5.5,
the volume expansion of the specimen accompanies an increase in entropy. Con-
sequently, in accordance with the second law of thermodynamics, the system
absorbs heat from outside, or if the environment is adiabatic, the temperature of
the system decreases. In the present case, the process is practically adiabatic and the
temperature of the specimen is decreased.
There are two possible causes of the observed temperature rise. First, the
temperature rise can be a consequence of heat release associated with the crack
formation and propagation. When a crack is formed and propagates, the difference
between the material’s internal energy and the energy needed to form and propagate
the crack is dissipated as heat. This is a subject of fracture mechanics, and details
are discussed in a number of references [27–29]. Second, it can be due to energy
dissipation associated with the flow of charge .r  v/. As discussed earlier in this
chapter and in Chap. 5, when a deformation charge flows the material loses the
momentum irreversibly. This effect was referred to as the energy dissipative (or
damping) force in the earlier sections. Under the practically adiabatic environment,
this causes a temperature rise of the specimen.
The temperature rise observed after the crack initiation can be understood as
dissipation of the first type. At the beginning of deformation when the restoring
force represented by the shear force Gr  ! appearing on the right-hand side
of the equation of motion in the plastic regime [see Eqs. (5.2), (5.44) or (5.50)] is
dominant, the material stores the elastic energy (the spring potential energy) as the
internal energy. When the crack starts to grow, the material becomes discontinuous
and loses its capability of storing this internal energy. Consequently, it releases the
energy as heat. When the sample is brittle, like the aluminum case in Fig. 7.34, these
energy storing phase and heat releasing phase are sharply divided in time. That is
why the onset of the heat release coincides with the onset of load decrease in the
aluminum data in Fig. 7.34.
The temperature rise prior to the onset of load decrease can be understood as the
second type. Naturally, the effect of energy dissipative force is more prominent if
the material is more ductile. The above observation indicates that the temperature
rise prior to the load decrease is much more prominent in the brass specimen than
the aluminum specimen [findings (5) and (6)]. This can be explained as the fact
that brass is more ductile than aluminum in consistence with finding (2). When the
energy dissipation due to the damping force becomes substantial, the material can
release heat while the restoring force is still effective; thus, the load increase due
206 7 Experimental Observations

23.6

23.4
Brass
23.2
Temp (ºC)

23.0

22.8

22.6
time
22.4
-50 0 50 100 150 200 250 300 350
Load (kgf)

23.0
Aluminum

22.8
Temp (ºC)

22.6

22.4

time
22.2
-20 0 20 40 60 80 100 120
Load (kgf)

Fig. 7.35 Load versus temperature rise

to the restoring force and heat release due to the damping force can coexist. This
is consistent with previous observations in various materials that when charge-like
patterns drift in the specimens the oscillation in the displacement field decays (the
oscillation is damped).
Figure 7.35 shows the same data as Fig. 7.34 in the form of the load change versus
the temperature change. In the brass data (top), the load increases along the lower
path reaching the maximum value of 320 kgf, and returns to zero along the upper
path. The gradual decrease in temperature in the lower path indicates the above
finding (4), the initial temperature decrease. The slight temperature rise in the lower
path near the maximum load indicates the above finding (6), the onset of temperature
rise prior to the crack initiation. In the aluminum data (bottom), the forward path
and return path are almost overlapping each other, while the return path is much
smoother than the forward path indicating the fact that the load decrease is much
faster than brass. There is no temperature rise near the maximum load in the forward
path, indicating that the onset of the temperature rise is at the same time as the
onset of load decrease. The smoothness observed in the return path indicates the
References 207

Fig. 7.36 Load versus temperature rise

fact that the load-decrease/temperature-rise is faster in aluminum. The temperature


difference at the zero load point between the forward path and return path indicates
the total heat release. It is clear that the brass specimen releases much more heat
than the aluminum specimen.
The original energy source of the released heat is the work done by the tensile
machine. The work done by the tensile machine can be computed numerically
through integration of the product of the load and elongation with respect to time. In
Fig. 7.36, the total temperature rise is plotted as a function of the tensile machine’s
work for both brass and aluminum. The density and specific heat of the brass and
aluminum are taken into account. The relationship between the tensile machine’s
work and the temperature rise is reasonably linear, supporting the above interpreta-
tion.

References

1. Yoshida, S., Siahaan, B., Pardede, M.H., Sijabat, N., Simangunsong, H., Simbolon, T.,
Kusnowo, A.: Phys. Lett. A 251, 54–60 (1999)
2. Sciammarella, C.A., Sciammarella, F.M.: Experimental Mechanics of Solids. Wiley, Hoboken
(2012)
3. Meinlschmidt, P., Hinsch, K.D., Sirohi, R.S. (eds.): Selected Papers on Electronic Speckle
Pattern Interferometry: Principles and Practice, SPIE Press, Bellingham, Washington, vol. 132
(1996)
4. Sirohi, R.S. (ed.): Speckle Metrology. Marcel Dekker, New York (1993)
5. Danilov, V.I., Zuyev, L.B., Mnikh, N.M., Paninand, V.Y., Shershova, L.V.: Phys. Met. Metall.
71, 187 (1991)
6. Toyooka, S., Widiastuti, R., Qingchuan, Z., Kato, H.: Dynamic observation of localized
strain pulsation generated in the plastic deformation process by electronic speckle pattern
interferometry Japan. J. Appl. Phys. 40, 310–313 (2001)
208 7 Experimental Observations

7. Nakamura, T., Sasaki, T., Yoshida, S.: Analysis of Portevin-Le Chatelier effect of Al-Mg alloy
by electronic speckle pattern interferometry, advancement of optical methods in experimental
mechanics. In: Conference Proceedings of the Society for Experimental Mechanics Series, vol.
3, pp. 109–117 (2014)
8. Sasaki, T., Yoshida, S.: Revealing load hysteresis based on electronic speckle pattern interfer-
ometry and physical mesomechanics. Phys. Mesomech. 15(1–2), 47–57 (2012)
9. Bird, J.E., Newman, K.E., Narasimhan, K., Carlson J.M.: Heterogeneous initiation and growth
of sample-scale shear bands during necking of Al–Mg sheet. Acta Metall. 35, 2971–2982
(1987)
10. Korbel, A., Embury, J.D., Hatherly, M., Martin, P.L., Erbsloh, H.W.: Microstructural aspects
of strain localization in Al–Mg alloys. Acta Metall. 34, 1999–2009 (1986)
11. Guo, Z.X. (ed.): The Deformation and Processing of Structural Materials. Woodhead Publish-
ing, Cambridge (2005)
12. Yoshida, S., Muhamad, I., Pardede, M., Widiastuti, R., Siahaan, M.B., Kusnowo, A.: Optical
interferometry applied to analyze deformation and fracture of aluminum alloys. Theor. Appl.
Fracture Mech. 27, 85–98 (1997)
13. Yoshida, S., Widiastuti, S.R., Pardede, M., Hutagalung, S., Marpaung, J.S., Muhardy, A.F.,
Kusnowo, A.: Direct observation of developed plastic deformation and its application to
nondestructive testing. Jpn. J. Appl. Phys. 35, L854–L857 (1996)
14. Yoshida, S., Ishii, H., Ichinose, K., Gomi, K., Taniuchi, K.: J. Appl. Mech. 72, 792–794 (2005)
15. Ichinose, K., Kosaka, Y., Fukuda, K., Taniuchi, K.: Detection of the failure zone caused by
cyclic loading. J. Intell. Mater. Syst. Struct. 10, 214–220 (1999)
16. Sylwestrowicz, W., Hall, E.O.: The deformation and aging of mild steel. Proc. Phys. Soc. Lond.
Sect. B 64, 495–502 (1951)
17. Funamoto, Y.: The observation of yielding region of the circular hole vicinity under the
gradually increase cyclic loading and application of ESPI in the failure region of low carbon
steel, p. 102. Master’s thesis. Tokyo Denki University, Tokyo (2002, in Japanese)
18. Yoshida, S., Suprapedi, Widiastuti, R., Astuti, E.T., Kusnowo, A.: Phase evaluation for
electronic speckle-pattern interferometry deformation analysis. Opt. Lett. 20, 755–757 (1995)
19. Mertens, F., Franklin, S.V., Marder, M.: Phys. Rev. Lett. 78, 4502 (1997)
20. Yoshida, S., Ishii, H., Ichinose, K., Gomi, K., Taniuchi, K.: Observation of optical interfero-
metric band structure representing plastic deformation front under cyclic loading. J. Jpn. Appl.
Phys. 43, 5451–5454 (2004)
21. Hasegawa, S., Sasaki, T., Yoshida, S., Hebert, S.L.: Analysis of fatigue of metals by electronic
speckle pattern interferometry, to be presented at Society for Experimental Mechanics, 2014
Annual Meeting (2014)
22. Okiyama, T., Ichinose, K., Yoshida, S.: 17th Annual Meeting of the Japan Society of
Mechanical Engineers Kanto Branch, March 18–19 (2011)
23. Liu, A.: Mechanics and Mechanisms of Fracture, an Introdcution. ASM International, Materi-
als Park (2005)
24. Yoshida, S.: Optical interferometric study on deformation and fracture based on physical
mesomechanics. J. Phys. Meso. Mech. 2(4), 5–12 (1999, in English and Russian)
25. Yoshida, S., Gaffney, G.A., Schneider, C.W., Rourks, R.L.: Field theoretical approach to
deformation and fracture. In: Proceedings of SEM XI International Congress, June 2–5,
Orlando, pp. 144–149 (2008)
26. Pieczyska, E.: Thermoelastic effect in austenitic steel referred to its hardening. J. Theor. Appl.
Mech. 2(37), 349–368 (1999)
27. Anderson, T.L.: Fracture Mechanics: Fundamentals and Applications, 3rd edn. CRC Press,
New York (2004)
28. Kobayashi, H.: Fracture Mechanics. Kyoritu-Shuppan, Tokyo (1993)
29. Tetelman, A.S., McEvily, J.A., Jr.: Fracture of Structural Materials. Wiley, New York (1966)
Chapter 8
Applications

In this chapter, we will explore potential applications of the present field theory in
conjunction with the Electronic Speckle-Pattern Interferometric (ESPI) techniques
discussed in the previous chapters. Section 8.1 discusses the use of the developed
charge-like pattern (discussed in Chaps. 5 and 7) for prediction of failure of metal
specimens under tensile loads. It has been observed in a number of experiments
that the behavior of the developed charge-like pattern is strongly related to the
level of stress concentration. When the stress concentration is at a low level, the
pattern normally moves along the specimen. As the level of stress concentration
increases, the pattern tends to be stationary, and the specimen always fails at the
location where the pattern becomes stationary. Several cases of specimen and its
initial stress conditions including welded specimens will be discussed. Section 8.2
describes the plastic deformation and fracture criteria derived from the present
theory, and demonstrates an application of them to an ESPI-based tensile analysis.
The transition from the elastic to plastic regime and that from the plastic to fracture
regime are diagnosed from the specific features of ESPI fringe pattern that represent
the plastic deformation and fracture criteria. The results of diagnosis are found
consistent with the loading characteristics. In Sect. 8.3, the same concept as Sect. 8.2
is applied to fatigue. Metal specimens are tensile-preloaded to various stress levels,
and released from the load. Subsequently, these specimens are tensile-reloaded at
a minimum stress level (in the linear elastic regime substantially lower than the
yield stress) and their response is monitored with ESPI as a two-dimensional,
full-field fringe patterns. The observed fringe patterns are analyzed based on the
plastic deformation and fracture criteria. Results of these analyses indicate that
this technique can be used to reveal the load hysteresis of a given specimen, the
hysteresis that cannot be revealed from analysis of the loading characteristic.

© Springer Science+Business Media New York 2015 209


S. Yoshida, Deformation and Fracture of Solid-State Materials,
DOI 10.1007/978-1-4939-2098-3__8
210 8 Applications

8.1 Evaluation of Stress Concentration with Charge-Like


Patterns

In the preceding chapter, various properties of charge-like patterns are discussed.


Some of the properties strongly indicate that the formation of a charge-like pattern
is related to stress concentration. In [1] a series of experiments are conducted
to explore the relationship between behaviors of the charge-like pattern and
initial stress concentration with the use of an in-plane displacement sensitive
ESPI arrangement. The specimens used in this study are plates of 150 mm in
effective length, 25 mm wide, and 2 mm thick. The material tested are aluminum
alloys A5052H112, A5052H32, and AA6063. The first two of these are standard
materials with specification guaranteed by the manufactures. They have the same
chemical composition and both are work-hardened after the casting process. The
only difference between them is that A5052H32 is stabilized after the work-
hardening treatment while A5052H112 is not. AA6063 is a low grade material
whose specification is not guaranteed except that its composition is known. Some of
the A5052H32 specimens are welded. Table 8.1 summarizes the conditions of the
specimens.

8.1.1 Stabilized and Unstabilized Non-welded Specimens,


A5052-S and A5052-N

These are non-welded specimens and therefore considered to be free of initial


stress concentration. Figures 8.1 and 8.2 show the temporal behavior of the charge-
like pattern and the corresponding stress–strain characteristics for these specimens.
The temporal behavior of the charge-like pattern is expressed by the location of
the charge-like pattern in the unit of the pixel number of the image taken by the
CCD camera. The stress–strain characteristics are expressed by the tensile load in
kN . Both are plotted as a function of specimen’s total elongation in mm. Points
A–C indicated in Fig. 8.1 denote, respectively, the point when the first charge-like

Table 8.1 Specimens


Specimen ID Material Stabilization Welding condition
A5052-S A5052H32 Stabilized Not welded
A5052-N A5052H112 Unstabilized Not welded
AA6063 AA6063 Unstabilized Not welded
A5052-SW A5052H32 Stabilized Shallow welded, vweld D 6 m=min
A5052-DW A5052H32 Stabilized Deep welded, vweld D 2 m=min
A5052-BW A5052H32 Stabilized Butt welded, vweld D 2 m=min
A5052-GC A5052H32 Stabilized Graphite coated, vweld D 2:5 m=min
8.1 Evaluation of Stress Concentration with Charge-Like Patterns 211

Charge location (pixel)


5.0
B C 500
Load (kN) 4.0
A 400
3.0
300
2.0 F
200
1.0
100
0
0 10 20
Elongation (mm)

Fig. 8.1 Charge-like pattern observed in A5052H work-hardened after casting followed by
stabilization, non-welded specimen

Stress drop

Charge location (pixel)


5.0
500
4.0 1st pattern
400
Load (kN)

3.0 300
F
2.0 200
1.0 100
0
0 5
Elongation (mm)

Fig. 8.2 Charge-like pattern observed in A5052H work-hardened after casting, unstabilized, non-
welded specimen

pattern appears, the charge-like pattern begins to decelerate, and the charge-like
pattern becomes completely stationary. Point F denotes the vertical location where
the fracture occurs. The temporal behavior of the charge-like patterns show clear
difference between the two specimens. The charge-like pattern observed in the
stabilized specimen is initially dynamic and becomes stationary prior to the fracture.
On the other hand, the charge-like pattern observed in the unstabilized specimen
is stationary from the beginning. Note in Fig. 8.1 that the moment when the first
charge-like pattern appears coincides with the initiation of the zigzag character
of the stress–strain curve. The fact that the stress shows the pattern of decreasing
and increasing reciprocally indicates that the stress localization associated with
the formation of one charge-like pattern fades out before the stress localization
associated with the formation of the next charge-like pattern is generated. It is
likely that each appearance of the charge-like pattern corresponds to different stress
concentration. In the case of Fig. 8.2, on the other hand, the stress does not show
clear zigzag characteristics, indicating that in this case a single stress concentration
is responsible for the fracture.
212 8 Applications

The appearance of charge-like pattern and the zigzag character of the stress–
strain curve are correlated in the following fashion. Each of the abrupt stress drop is
accompanied by the formation of a charge-like pattern. Once a charge-like pattern
is formed, the stress drops abruptly and recovers to approximately the same level
as before the abrupt drop. The next The next formation of a charge-like pattern is
accompanied by a similar combination of abrupt stress drop and recovery. In this
fashion, the stress drop/recovery and the formation of charge-like patterns repeat.
The detail pattern of the stress drop/recovery and the feature of the corresponding
charge-like pattern depend on a number of factors such as the grain size. Nakamura
et al. [2] made a thorough investigation on the dependence on the grain size.
While the temporal behaviors of the charge-like patterns are completely different,
the stabilized and unstabilized specimens have important properties in common.
First and most importantly, both specimens fracture at the location where the charge-
like pattern becomes stationary. (Note that the pixel number where the charge-like
pattern becomes stationary in Fig. 8.1 is the same as where the fracture is seen
in the object image shown right to the dynamic characteristics.) In Fig. 8.2, the
fracture occurs at the location where the charge-like pattern becomes stationary,
though an image of the fractured object is not shown. Second, the first charge-like
pattern appears as soon as the object enters the plastic regime in the stress–strain
characteristics. Third, once the charge-like pattern becomes completely stationary
(point C in Fig. 8.1), the stress begins to decrease.
These observations can be interpreted as follows. First, the fact that the specimen
fractures at the location of a stationary charge-like pattern proves the hypothesis
discussed in Chap. 5 (Sect. 5.2.3) that “the plastic deformation charge must stop
flowing as the material fractures; i.e., as long as the external agent applies a load,
the energy dissipation in response to it .r  v/W d is finite. If Wd D 0 under this
condition, the rate of the volume expansion r  v must be infinite.” Second, the fact
that the first charge-like pattern appears after the material enters the plastic regime
indicates that in both cases the specimen does not have an initial stress concentration
that causes local plastic deformation in a particular location before the entire
specimen yields. In other words, it is only after the specimen shows global plasticity
that the part of the specimen reaches the deformation-level in which the stress
concentration is high enough to cause a strain concentration observed as a charge-
like pattern. The third observation that the stress begins to decrease as soon as the
charge-like pattern becomes stationary indicates that when the above condition of
fracture that the volume expansion rate approaches infinity with Wd D 0, it takes
some time for the .r  v/ to “empty” the volume.
Based on this interpretation, the behavior of the respective specimens can
be explained as follows. Since the two specimens are the same in the material
compositions and the post casting hardening treatment, and the only difference
is that A5052-S is stabilized after the hardening treatment, the difference in
the temporal behavior of the charge-like pattern is considered to be due to the
stabilization treatment. In the case of A5052-N, when the first charge-like pattern
appears around the yield point the strain localization is already on the level of
a stationary charge-like pattern due to the post casting hardening treatment not
followed by stabilization treatment. At the location of the stationary charge-like
8.1 Evaluation of Stress Concentration with Charge-Like Patterns 213

pattern, there is an initial stress concentration that is higher in degree than other
locations but not high enough to form a charge-like pattern. In the case of A5052-S,
on the other hand, the stabilization treatment homogenizes stress concentration so
that there is no particular region that has higher stress concentration than the other
parts and this causes the charge-like pattern to be dynamic until a later stage when
it becomes stationary. Thus, from the dynamic characteristics of the charge-like
pattern, it is possible to diagnose that A5052-N has a low level stress concentration
from the beginning, although it is not welded.

8.1.2 AA6063

This is a specimen of a low-grade aluminum alloy. The peculiarity of the behavior


of the charge-like pattern observed in this specimen is that the 15 specimens
prepared from the same base plate vary in a completely random fashion. In the
case of A5052H base specimens, the charge-like patterns observed in the stabilized
specimens show the dynamic behavior in the early stage of plastic deformation
whereas those observed in the unstabilized specimens are stationary from the
beginning. In contrast, the charge-like patterns observed in some of the AA6063
specimens are dynamic at the beginning and become stationary in a later stage,
whereas the charge-like patterns observed in the others are stationary from the
first appearance. This is probably because as a low grade material, the base plate
lacks homogeneity in the mechanical property; some specimens happen to be from
relatively more homogeneous part and others are from less homogeneous part of the
base plate.
Interestingly, this diversity in the behavior of the charge-like pattern seems to
provides us with an insight into the relationship between the charge-like pattern and
the plasticity. In Table 8.2, the behaviors of the charge-like patterns observed in all
the fifteen AA6063 specimens prepared from the same base plate are summarized.
The charge-like pattern’s behavior is discussed in terms of when the first pattern
appears, and when the pattern becomes stationary. These parameters are compared
with the maximum load (corresponding to the ultimate strength) observed in the
loading characteristics, along with when the loading characteristics show the yield
point, maximum load, and fracture expressed in the unit of frame number of the
ESPI image. All the specimens are tested under the same tensile speed, and therefore
the frame number is proportional to the total elongation of the specimen.
From Table 8.2 the following features are found.
(i) When the charge-like pattern is dynamic, it begins to appear near the yield
point, as is the case of A5052-S.
(ii) Once a charge-like pattern becomes stationary, soon or later the stress begins
to decrease. (Compare the frame number when the first stationary charge-like
pattern appears and that when the load reaches the maximum. They are close
to each other whether or not the charge-like pattern is dynamic initially.) This
is also true for the A5052-S and A5052-N cases.
214 8 Applications

Table 8.2 Behaviors of charge-like patterns observed in AA6063 specimens


1st CLP Yield point Stationary CLP Max load point Fracture point Max load Ductility
(frame #) (frame #) (frame #) (frame #) (frame #) (kN) factor
50 60 180 175 195 4.48 30.1
44 44 150 150 171 4.08 31.1
143 101 143 151 161 6.14 9.8
98 49 98 105 116 6.07 11.0
138 63 138 142 156 6.62 14.0
122 81 122 131 146 5.80 11.2
50 50 157 156 184 5.80 23.1
107 61 107 111 126 6.90 9.42
15 16 95 124 146 4.00 32.5
72 28 72 93 106 5.00 15.6
94 29 94 100 116 6.00 14.5
121 39 121 123 141 7.00 14.6
96 45 96 109 126 7.00 11.6
100 30 100 103 121 6.30 14.4
31 30 31 31 104 3.70 20.0

(iii) When the charge-like pattern is dynamic or stationary from the beginning,
as compared with the case when a charge-like appears at a later stage and
stationary from the first appearance, the maximum load tends to be lower
and the plastic regime (the distance between the yield point and fracture
point on the loading characteristics) tends to be longer. In other words, the
specimen is more ductile if the charge-like pattern appears in the early stage,
regardless dynamic or stationary. For the purpose of quantifying these tendency
connecting to the ductility, a figure of merit called the ductility factor is
defined as (yield point–fracture point)/maximum load. Here the yield point
and fracture point are in the unit of data number, and the maximum load is in
kN. The numerator indicates how long the plastic regime in the total life of the
specimen and the reciprocal of the denominator indicates how easy to elongate
the specimen; overall this figure of merit indicates how ductile the specimen
is.
Figure 8.3 shows the ductility factor and the maximum load as a function of the
frame number when the first charge-like pattern appears. It is clearly seen that the
earlier the charge-like pattern begins to appear, the lower is the maximum load and
the higher is the ductility factor. This is understandable because earlier appearance
of a charge-like pattern means a faster growth rate of defects, hence faster evolution
of plastic deformation.
8.1 Evaluation of Stress Concentration with Charge-Like Patterns 215

Fig. 8.3 Ductile factor of low-grade aluminum alloy specimens

500
Charge location (pixel)

8.0
1st pattern
Load (kN)

4.0 250

0 0
0 5 10 15
Elongation (mm)

Fig. 8.4 Charge-like pattern observed in A5052H work-hardened, stabilized, shallow-welded


specimen

8.1.3 Welded Specimens

Figures 8.4, 8.5, 8.6 and 8.7 show the temporal behaviors of the charge-like patterns
and stress–strain characteristics observed in the welded specimens together with
an image of the charge-like pattern observed in the final stage. The material of
216 8 Applications

500

Charge location (pixel)


8.0

Load (kN) 250


4.0

1st pattern
0 0
0 5
Elongation (mm)

Fig. 8.5 Charge-like pattern observed in A5052H work-hardened, stabilized, deep-welded speci-
men

Charge location (pixel)


500

5.0 1st pattern


Load (kN)

250
2.5

0 0
0 5
Elongation (mm)

Fig. 8.6 Charge-like pattern observed in A5052H work-hardened, stabilized, butt-welded speci-
men
Charge location (pixel)

500
5.0
Load (kN)

250
2.5

1st pattern
0 0
0 5
Elongation (mm)

Fig. 8.7 Charge-like pattern observed in A5052H work-hardened, stabilized, graphite coated and
butt-welded specimen
8.1 Evaluation of Stress Concentration with Charge-Like Patterns 217

Fig. 8.8 Void observed in shallow-welded, deep-welded, butt-welded and graphite-coated speci-
mens

these specimens is the same as A5052-S. When a specimen is welded, voids


formed in the heat-affected zone around the weld cause stress concentration. Defect
densities for these specimens are shown in Fig. 8.8. Below we discuss the temporal
behaviors of the charge-like pattern observed in the respective specimens. Here the
welding method for the shallow-welded (A5052-SW) and deep-welded (A5052-
DW) specimens is the bead-on-plate (the welding laser is applied on the specimen
without any joint), and that for the butt-welded (A5052-BW) and the graphite-
coated (A5052-GC) specimens is butt-welding. The graphite-coated specimen is
coated by graphite prior to the welding.
The charge-like pattern observed in the shallow-welded (A5052-SW) in which
the void density is very low shows basically the same temporal behavior as the
non-welded case (A5052-S). The charge-like pattern begins to appear past the
yield point and it is dynamic until a very late stage. This indicates that the initial
stress concentration is negligibly small. Note that, however, the specimen eventually
fractures along the weld-line. From this standpoint, the initial stress concentration
caused by the weld influences the fracture, while its degree is not higher enough
than the other part of the material to cause the charge-like pattern to be stationary at
the beginning.
When the specimen is deep-welded (A5052-DW), the first charge-like pattern
appears prior to the yield point. This indicates that there is local plastic deformation
that forms a charge-like pattern near the weld while other parts are still elastically
deformed. The temporal behavior of the charge-like pattern is basically stationary
but occasionally dynamic. This is a sort of intermediate situation between the cases
of the shallow-welded specimen and the butt-welded specimen. It is interesting that
unlike the non-welded and shallow-welded cases, the stress–strain characteristic
does not show a zigzag pattern. This indicates that in the case of the deep-welded
specimen, the stress is relaxed through enhancement of the strain occurring at the
location of the existing charge-like pattern rather than forming a strain concentration
at a new location accompanying an abrupt stress drop. Note that as soon as the
charge-like pattern finally becomes stationary, the stress decreases as is the non-
welded case.
In the case of the butt-welding (A5052-BW), the charge-like pattern is com-
pletely stationary from the beginning at the weld where the specimen fractures
eventually. The first charge-like pattern is observed in the elastic region of the
stress–strain characteristics. It can be said that the initial stress concentration of
this specimen is strong enough to cause intensive strain localization at the weld and
218 8 Applications

Table 8.3 Summary of observation in stabilized A5052 specimens


Void 1st CLP CLP temporal Yield load Max load Stationary CLP Fracture
Weld type density regime behavior (kN) (kN) location location
No weld No Plastic Dynamic 6.5 9.2 296/274 296/274
Shallow No Plastic Dynamic 4.0 6.5 236/273 236/273
Deep Very low Elastic Dynamic 6.8 7.6 264 264
Butt Low Elastic Stationary 3.6 4.5 240 240
Graphite High Elastic Stationary – 5.0 252 252

maintain it at the same location until the material fractures. As is the case of the
deep-welded specimen, after the appearance of the first charge-like pattern the other
part of the specimen is elastically deformed. This indicates that the level of the local
plastic deformation of the butt-welded specimen is such that the associated strain
localization is intense enough to make the charge-like pattern stationary but since it
is localized the other part of the material is still elastically deformed.
When the specimen is coated by graphite, it is well known that the void density
is increased. In this case, the charge-like pattern begins to appear as soon as the
tensile load is applied. The charge-like pattern is stationary from the beginning
and the specimen fractures before the stress–strain curve reaches the yield point.
The characteristics of the local plasticity is basically similar to the case of the butt-
welded specimen, but since the associated strain localization is more intense the first
charge-like pattern appears earlier and the material fractures there while the other
part is still being elastically deformed. In other words, the level of local plastic
deformation is higher than the butt-welded specimen and it can cause the specimen
to fracture while the other part is still being deformed elastically.
Table 8.3 summarizes the results of the stabilized A5052 specimens. From these
observations, the following things can be said: (1) all the specimens eventually
fracture where the charge-like pattern becomes stationary; (2) the charge-like pattern
becomes stationary at the weld; (3) as the void density, hence the degree of initial
stress concentration, increases, the charge-like pattern (i) tends to begin appearing
prior to the yield point and (ii) tends to be stationary from the beginning.
In (3), (i) indicates the existence of local plastic deformation and (ii) indicates
that the level of associated strain localization is intense from the beginning. In terms
of the level of the initial stress concentration, (i) can appear at a lower level. In
other words, when the degree of the initial stress concentration increases, the first
symptom is the initiation of local plastic deformation that is revealed as the first
charge-like pattern that appears in the elastic regime. As the level of the initial stress
concentration increases, the local plastic deformation can cause intensive strain
localization causing the charge-like pattern to be stationary from the beginning. In
this fashion, it is possible to diagnose the level of the initial stress concentration by
noting when the first charge-like pattern appears and whether it is stationary.
8.2 Plastic Deformation and Fracture Criteria 219

8.2 Plastic Deformation and Fracture Criteria

8.2.1 Plastic Deformation Criterion

As discussed earlier in various occasions in this book, the present theory formulates
plastic deformation as dynamics associated with the gauge field that is necessary
to recover local symmetry in the theory of linear elasticity. Mathematically, the
deformation dynamics is expressed by the transformation matrix U defined by
Eq. (4.18) and the distortion tensor defined by Eq. (4.19) where beta has coordinate
dependence.

0 D U  D .I C ˇ/ (5.73)
0 1
@x @x @x
B @x @y @z C
B C
B C
@i B @y @y @y C
ˇD DBB @x @y @z C
C (4.19)
@xj B C
B C
@ @z @z @z A
@x @y @z

The elements of matrix (4.19) are first order derivatives of the displacements.
Therefore, the coordinate dependence of this matrix means that the differentiation
of the first order derivatives of the displacement with respect to the coordinate
variables is nonzero. In other words, the displacements are second or higher
order function of the coordinate variables, or they are nonlinear. In the context
of deformation mechanics, the situation can be expressed in the following way:
as the deformation of an initially linear elastic material develops, the population
of dislocations and other defects is increased. This prevents the material from
obeying the law of linear elasticity as a whole, whereas locally it can still being
elastically deformed. Consequently, in the global coordinate, the displacements
becomes nonlinear. Consider this condition in terms of the field equations (4.61)–
(4.64).

r  v D j0 (4.61)
@!
r v D (4.62)
@t
1 @v
r ! D  2 j (4.63)
c @t
r ! D 0 (4.64/
220 8 Applications

Fig. 8.9 Schematic


illustration of r  ! ¤ 0 (∇×ω)y>0

ωz>0

(∇×ω)x>0 (∇×ω)x>0

ωz<0 ωz>0

(∇×ω)y>0

The term r  ! on the left-hand side of Eq. (4.63) is a second-order derivative of


the displacement (spatial differentiation of omega, which is a first-order derivative of
displacement). Therefore, the above-mentioned condition of nonlinear displacement
can be interpreted as this term becomes nonzero.
This condition r  ! ¤ 0 is a sufficient condition that the transformation
matrix is coordinate-dependent, or deformation cannot be expressed by the global
coordinates as linear elasticity. It is not a condition that the deformation is plastic.
The deformation becomes plastic when the charge drifts under the influence of
the velocity field causing energy dissipation. This condition can be conveniently
expressed as the longitudinal effect Gj represents an energy dissipative flow of
charge in the form of Eq. (5.45) as opposed to elastic force in the form of Eq. (5.23)
[or Eq. (5.36) for a one-dimensional case]. Thus, it is reasonable to hypothesize the
following conditions as a criterion for the initiation of plastic deformation in an
initially linear elastic material [3].

r ! ¤ 0 (8.1)
Gj D Wd .r  v/ (8.2)

where Gj is the longitudinal force and Wd is the drift velocity of the charge .r v/.
Note that the first condition (8.1) is a condition of nonlinear deformation, not
plastic deformation from the theoretical viewpoint. However, in reality, it is likely
that the material is partially plastically deformed when r  ! is nonzero because,
as schematically illustrated by Fig. 8.9, it is likely that defects exist along the
boundaries of neighing blocks when they experience differential ! [4]. As will
be discussed below, experiment indicates that the first condition is sufficient to
diagnose that an initially linear elastic material enters the plastic regime under
tensile loading.
8.2 Plastic Deformation and Fracture Criteria 221

8.2.2 Fracture Criterion

As we discussed in Sect. 5.2, the term r! appearing on the left-hand side of condi-
tion (8.1) represents the shear force acting on the unit volume (see Fig. 4.8). Fracture
occurs when this shear force and the mechanism to dissipate the mechanical-energy
input from the external agent in the form of the flow of charge .r  v/ stop
functioning. Now under this condition, if the external agent keeps exerting force, the
material needs to dissipate it so that the total energy is conserved. The condition of
the nonzero energy to be dissipative can be expressed as Gj ¤ 0, and the conditions
that the shear force and the energy dissipation mechanism to stop functioning are,
respectively, r! D 0 and W d D 0. In order for this condition to be true, it follows
that r  v ! 1 or the particle flow-out rate from the unit volume becomes infinite
(see Sect. 5.2.3). Thus, the fracture criterion can be expressed by the following
conditions [3]:

r ! D 0 (8.3)
Gj D Wd .r  v/ ¤ 0 (8.4)
Wd D 0 (8.5)

Further, the transitional stage from plastic deformation to fracture can be


interpreted as the situation where the material has lost the shear force mechanism but
still possesses the energy dissipation mechanism. From this viewpoint, the critical
fracture criterion can be expressed as follows:

r ! D 0 (8.6)
Gj D Wd .r  v/ ¤ 0 (8.7)
Wd  1 (8.8)

8.2.3 ESPI Experiment on Plastic Deformation


and Fracture Criteria

Experiments have been conducted to test the above criteria (8.1)–(8.8) with the use
of Electronic Speckle Pattern Interferometry (ESPI). In [5], an ESPI setup is used to
form fringe patterns representing the two-dimensional displacement field resulting
from a tensile load, and the status of deformation is diagnosed based on the criteria.
Figure 8.10 illustrates the experimental arrangement used in this study. Two in-
plane sensitive ESPI setups are arranged on the front and rear sides of the specimen
attached to a tensile machine. The ESPI on the front side is sensitive to horizontal
displacement of the specimen and the one on the rear side is sensitive to vertical
displacement. Mirrors are arranged to image both sides of the specimen on the same
222 8 Applications

FM load Horizontally sensitive ESPI


BE
Laser
BS IM FM BE

CCD
BE
FM BS
BE FM
Vertically sensitive ESPI FM

Fig. 8.10 Experimental setup. FM folding mirror, BS beam splitter, BE beam expander, IM
imaging mirror. Multiple imaging mirrors were used to make the object distances for the front and
rear surfaces of the specimen the same, but only one is shown in this figure to avoid complexity

CCD camera so that the speckle-images associated with the horizontal displacement
and vertical displacement can be captured simultaneously. The specimen is a tin
plate of 20 mm wide, 100 mm long, and 0.4 mm thick. One side of the specimen is
curved so that the width of the vertical center is narrowest (15 mm) and thereby
fracture is always initiated on this side at the vertical center (see upper part of
Fig. 8.11). The speckle-images are taken at 30 frames/s as the tensile machine
applies the load at a constant pulling rate (cross-head speed) of 4 m/s.
By subtracting the speckle-image taken at a certain time step from the one taken
at another time step, a fringe pattern representing the displacement experienced by
the specimen between the two time steps can be obtained. The top half of Fig. 8.11
shows fringe patterns obtained at several time steps in the same tensile experiment.
As these images indicate, each fringe pattern shows a number of dark fringes. Each
dark fringe represents the contour of displacement whose phase change is an even
integer multiple of .

8.2.4 Interpretation of Fringe Patterns

By approximating the dark fringes observed in Fig. 8.11 with second order polyno-
mials, we can express the corresponding displacement as follows:

x .x; y/ D a2 x 2 C a1 x C b2 y 2 C b1 y D mx0 (8.9)


y .x; y/ D c2 x C c1 x C d2 y C d1 y D ny0 ;
2 2
(8.10)

where x and y denote the horizontal and vertical displacement, respectively, m


and n are integers, and x0 and y0 represent the unit displacement corresponding to
the phase change of D 2. It is reasonable to eliminate the term proportional to
8.2 Plastic Deformation and Fracture Criteria 223

a b c d e

41 H 402 H 2422 H 2741 H 3175 H

42 V 403 V 2423 V 2742 V 3176 V

1600
1400 (e)
1200 (b)
(c) (d)
1000
load (N)

800
600 (a)
400
200
0
0 1000 2000 3000 4000
frame #

Fig. 8.11 Horizontally (top row) and vertically (bottom row) sensitive fringes

xy because the curved fringes are symmetric about the x or y axes (Fig. 8.11). With
these polynomial approximations, different stages of deformation can be expressed
as follows.

Elastic Condition

Based on the argument made in conjunction with the plastic deformation criterion,
the elasticity is characterized by the condition that the distortion tensor components
are independent of the coordinates, i.e., the first order derivatives of displacement
(@x =@x; @x =@y, etc.) are constant. In Eqs. (8.9) and (8.10), this means that the
coefficients for the second order terms are all zero, or the dark fringes are equally-
spaced, straight lines. If this condition is true, the deformation is elastic.
224 8 Applications

Table 8.4 Possible Case 1 2 3 4 5 6 7 8


combinations of coefficient of
x .x; y/ to satisfy condition a2 ¤0 ¤0 ¤0 ¤0 D0 D0 D0 D0
(8.11) a1 ¤0 ¤0 D0 D0 ¤0 D0 ¤0 D0
b1 ¤0 D0 ¤0 D0 D0 ¤0 ¤0 D0

Plastic Condition

If the above condition of elasticity is not true, the deformation is plastic. (As
mentioned above, although the first condition (8.1) is a condition of nonlinear
deformation, not plastic deformation from the theoretical viewpoint, it is likely that
the material is partially plastically deformed when r  ! is nonzero.) It follows that
if the dark fringes are curved or not equally spaced, the deformation is plastic.

Critical Fracture Condition

In the xy plane, the critical fracture conditions (8.6) and (8.7) can be expressed as
follows:

@!z @2 y @2 x
.r  !/x D D  D 2b2 y D 0; i:e:;b2 D 0 (8.11)
@y @x@y @y 2
@!z @2 y @2 x
.r  !/y D  D . 2  / D 2c2 x D 0; i:e:;c2 D 0 (8.12)
@x @x @x@y
@x @y
Gj D . C /W d D .2a2 x C a1 C 2d2 y C d1 /W d ¤ 0 (8.13)
@x @y

where Eq. (8.2) is used in Eq. (8.13). Equation (8.11) indicates that under the critical
fracture condition the horizontally sensitive fringes x .x; y/ does not have second
order dependence on y while it can have second order dependence on x. Similarly,
Eq. (8.12) indicates that the vertically sensitive fringes y .x; y/ does not have
second order dependence on x but can have second order dependence on y. Table 8.4
lists all the possible combinations for the coefficients of the horizontally sensitive
fringes to satisfy condition Eq. (8.11), and Table 8.5 shows the shape and expression
of the dark fringes for each combination. Table 8.6 shows the corresponding cases
for the vertically sensitive fringes.
The fringe shapes observed in Tables 8.5 and 8.6 can be conveniently sum-
marized as follows. If horizontally/vertically sensitive fringes lose their horizon-
tal/vertical curvature, it means that the critical fracture condition is satisfied.
Equivalently, if horizontally/vertically sensitive fringes show a horizontal/vertical
curvature, the material is in the plastic regime but does not meet the critical fracture
condition.
With the above arguments in mind we now analyze the fringe patterns seen in
Fig. 8.11. The images in the upper and lower rows are, respectively, the horizontally
8.2 Plastic Deformation and Fracture Criteria 225

Table 8.5 Equation and shape of horizontally sensitive fringes


x .x; y/ under critical fracture condition
Case Equation and shape of fringes
1 a2 x 2 C a1 x C b1 y D mu0 , vertical parabolas
2 a2 x 2 C a1 x D mu0 , compressed, vertical straight lines
3 a2 x 2 C b1 y D mu0 , vertical parabolas
4 a2 x 2 D mu0 , compressed, vertical straight lines
5 a1 x D mu0 , equally-spaced, vertical straight lines
6 b1 y D mu0 , equally-spaced, horizontal straight lines
7 a1 x C b1 y D mu0 , equally-spaced, slant straight lines
8 Tirvial

Table 8.6 Equation and shape of vertically sensitive fringes


y .x; y/ under critical fracture condition
Case Equation and shape of fringes
10 d2 y 2 C d1 y C c1 x D nv0 , horizontal parabolas
20 d2 y 2 C d1 y D nv0 , compressed, horizontal straight lines
30 d2 y 2 C c1 x D nv0 , horizontal parabolas
40 d2 y 2 D nv0 , compressed, horizontal straight lines
50 d1 y D nv0 , equally-spaced, horizontal straight lines
60 c1 x D nv0 , equally-spaced, vertical straight lines
70 d1 y C c1 x D nv0 , equally-spaced, slant straight lines
80 Tirvial

and vertically sensitive fringes. The pair of the images in the same column are
the fringes formed from the speckle-images taken at the same time steps. The
lower part of Fig. 8.11 is the loading characteristics of the specimen. The labels
(a–e) indicate the points on the loading curve when the fringe patterns (a–e)
in Fig. 8.11 are formed. The following observations support the above argument
regarding the relationships between the fringe shapes and the stage of deformation.
Fringes in Fig. 8.11a represent elastic deformation. Fringes are all straight lines and
equally spaced. The loading curve in Fig. 8.11 supports this observation indicating
that the stress–strain relationship is in the linear range when these fringes are
formed. Fringes in Fig. 8.11b correspond to the point where the linear stress–
strain relationship is about to finish in the loading curve. Note that near the
vertical center of the images, the horizontally/vertically sensitive fringes show
horizontal/vertical parabolic shapes. In accordance with the above argument, this
indicates that the material is being deformed plastically but has not reached the
critical fracturing stage. The same tendency (the horizontally/vertically sensitive
fringes show horizontal/vertical parabolas) continues till point (c).
At point (d) in Fig. 8.11, the horizontally/vertically sensitive fringes lose the
horizontal/vertical curvatures, becoming straight lines. These satisfy the critical
fracture conditions (8.6) and (8.7). Indeed, short after this point, the loading curve
226 8 Applications

begins to decrease supporting this observation. It is interesting to note that the


critical fracture condition is satisfied in the horizontally and vertically sensitive
fringes at the same time.
Fringes in Fig. 8.11e are typically observed in the last stage of deformation
where the load decreases monotonically. Normally, the horizontally and vertically
sensitive fringes show similar, slant straight lines that move with the crack tip until
the specimen completely fractures, as shown in Fig. 7.32; the fringe pattern can be
classified as case 7 of Table 8.5. Apparently, @x =@x ¤ 0 , and condition (8.13) is
satisfied.

8.3 Evaluation of Load Hysteresis

Another potential field of application is diagnosis of load hysteresis. By applying


a minimum load to a given object and using the plastic deformation and fracture
criteria discussed in the preceding section, it is possible to diagnose in what stage of
deformation the object currently stands. With the use of an ESPI setup similar to the
one discussed in the preceding section, the response of the object to the minimum
external load can be visualized as interferometric fringe patterns, and from the shape
of the fringes the current deformation status can be diagnosed. This is a new field
of application and little experimental data is available at this time. In this section
preliminary results obtained in previously conducted experiment [6,7] are discussed.
The experimental arrangement is similar to the one discussed in the preceding
section (Fig. 8.10) except that the interferometer is one-dimensional, being sensitive
to in-plane displacement parallel to the tensile axis. The specimen used in this study
is a 100 mm long, 20 mm wide, and 0.5 mm thick rectangular aluminum plate with
shallow notches at the center of the 100 mm sides. A number of specimens with
these dimensions are prepared for the following experiments. For a base-line data,
first, a tensile load is applied to a specimen at a constant head speed of 20 m/s
until it breaks. The solid line labeled “full load” in Fig. 8.12 shows the loading
characteristics of this experiment. Next, four groups of pre-loaded specimens are
prepared at the following four different stress levels; 85 kgf (Specimen A), 100 kgf
(Specimen B), 106 kgf (Specimen C), and 109 kgf (Specimen D). Finally, all these
preloaded specimens are reloaded up to a stress level of 80 kgf five times. The four
preload levels are indicated in Fig. 8.12 along with the reloading characteristics
of the respective specimens. Note that the maximum reload level of 80 kgf is
within the linear range of the loading characteristics when the specimen is pulled
until fracture (the full loading characteristics). In the right plot of Fig. 8.12, the
reloading characteristics are shifted on the time scale so that the initial rise of
the reload overlaps the full loading curve. Notice that all the preloaded specimens
show reloading characteristics perfectly overlapping each other and the full loading
characteristics. This indicates that it is impossible to reveal the loading hysteresis
by examining the reloading characteristics.
To facilitate the discussion regarding the fringe patterns in connection with the
displacement field of the examined specimen, let’s first consider the general pattern
8.3 Evaluation of Load Hysteresis 227

140 140
120 120
100 100
load (kgf)

load (kgf)
80 80
60 60
40 40
20 20
0 0
0 30 60 90 120 0 30 60 90 120
time (s) time (s)

specimen A preload A specimen A specimen B


specimen B preload B specimen C specimen D
specimen C preload C full load
specimen D preload D
full load

Fig. 8.12 Loading characteristics of full load and reloads. The reload characteristics are shifted
on the right figure so that rising slopes can be compared with the full loading characteristics

a b

Fig. 8.13 Fringe patterns representing uniform stretch (a) and rotation (b). The dashed lines in
the right picture denote the horizontal components of the displacement under pure rotation, which
a horizontally sensitive interferometer is sensitive to

of fringes. Refer to Fig. 8.13 and consider the fringes observed with an ESPI setup
sensitive to horizontal displacement. When the specimen experiences horizontal
stretch to the right, for example, the right end of the specimen is displaced the
most largely, the left end is displaced the least and the rest parts are displaced in
proportion to the distance from the left end. Consequently, the fringe pattern consists
of vertical straight lines. If the deformation is uniform, the vertical lines are equally
spaced. On the other hand, when the specimen experiences pure rotation, the fringe
pattern observed in the same ESPI setup consists of horizontal straight lines. If the
deformation is a combination of these, which is usually the case, the resultant fringe
pattern is a mixture of these patterns.
228 8 Applications

x Stage 1 Stage 2 Stage 3 Stage 4 Stage 5

Fig. 8.14 Representative fringe-pattern observed in each stage. The arrows indicate rotation of
the material

a b

Preload A

Preload B

Preload C

Preload D

Reload 3 Reload 4 Reload 5 Reload 3 Reload 4 Reload 5

Fig. 8.15 Fringe patterns observed with reload level of 10 kgf (a) and 50 kgf (b)

When reloaded, the preloaded specimens generally show the following trends. As
the applied load is increased from 0 to 80 kgf, the fringe-patterns show changes that
can be classified into five stages. Figure 8.14 shows representative fringe patterns
observed in each stage. The actual stress level at which the fringe-pattern changes
from one stage to the next depends on the preloading condition and the number of
reloading, as indicated by Figs. 8.15, 8.16 and 8.17.
These transitions in the stage of fringe patterns can be explained in terms of
the recovery mechanism in plasticity in conjunction with the fracture criterion
represented by Eqs. (8.3) and (8.4). As the deformation develops, the material loses
the recoverability mechanism [the plastic shear recovery force G.r  !/], and
instead, the energy dissipation current [Gj D Wd .r  v//] dominates. In a two-
dimensional picture in an xy-plane, the plastic recovery force and the current can be
expressed in terms of the in-plane displacement and rotation around an axis normal
to the plane. Under the present experimental condition, since the interferometer has
the sensitivity to the horizontal (x) component of the in-plane displacement in the
xy-plane, the plastic recovery force and the energy dissipation current are detected
in the fringe pattern only in the spatial variation of the horizontal displacement as
shown below.
 
@!z @!z
G.r  !/ D G xO  yO
@y @x
 2   2  
@ y @2 x @ y @2 x
DG  xO   yO
@x@y @y 2 @x 2 @x@y
 2 
@ x @2 x
D G  2 xO C yO (8.14)
@y @x@y
90 90
Reload 1 Reload 2
80 80
70 70
60 stage 5 60 stage 5
50 50 stage 4
stage 4
40 40
stage 3 stage 3

Load L (kgf)
30

Load L (kgf)
30
stage 2 stage 2
20 20
stage 1 stage 1
10 10
8.3 Evaluation of Load Hysteresis

0 0
A B C D A B C D
Sample Sample

90 90
Reload 3 Reload 4
80 80
70 70
60 stage 5 60 stage 5
50 stage 4 50 stage 4
40 40
stage 3 stage 3

Load L (kgf)
Load L (kgf)
30 30
stage 2 stage 2
20 20
stage 1 stage 1
10 10
0 0
A B C D A B C D
Sample Sample

Fig. 8.16 Change in fringe-patterns as reload number increases


229
Preload A Preload B
230

90 90
Sample A Sample B
80 80
70 70
60 stage 5 60 stage 5
50 50 stage 4
stage 4
40 40
stage 3 stage 3
30 30

Load L (kgf)
Load L (kgf)
stage 2 stage 2
20 20
stage 1 stage 1
10 10
0 0
1 2 3 4 5 1 2 3 4 5
Reload time N Reload time N

Preload C Preload D
90 90
Sample C Sample D
80 80
70 70
60 stage 5 60 stage 5
50 stage 4 50 stage 4
40 40
stage 3 stage 3
30 30

Load L (kgf)
Load L (kgf)
stage 2 stage 2
20 20
stage 1 stage 1
10 10
0 0
1 2 3 4 5 1 2 3 4 5
Reload time N Reload time N

Fig. 8.17 Change in fringe-patterns as preload level increases for each reload
8 Applications
8.3 Evaluation of Load Hysteresis 231

   
@x @y @x
Gj D .r  v/W d D C Wd D@ Wd (8.15)
@x @y @x

Here !z is the z-component of the rotation, x and y are the x and y-components
of the displacement, and xO and yO are the unit vectors. Note that in Eq. (8.14) the
x and y-components of ! and the terms containing y are dropped because the
interferometer does not have sensitivity to these components. Similarly, the second
term in the parenthesis is dropped in going through the last equal sign in Eq. (8.15).
With these expressions, the fringe pattern of each stage can be explained as below.
Stage 1
In this stage, the fringe patterns are approximately vertically parallel, indicating
that the deformation is basically elastic. If the deformation is purely elastic,
the fringes representing displacement x observed in the horizontally sensitive
interferometer ought to be equidistant, vertical straight lines. As the left-most
image in Fig. 8.14 shows, the fringe patterns seen in Stage 1 are normally not
completely parallel indicating that the deformation in this stage is not purely
elastic. This is consistent with the fact that the loading curve is not completely
linear in the initial rising part.
Stage 2
In this stage, the fringe patterns consist of horizontal wavy lines, indicating that
.r !/x D @!z =@y ¤ 0 and .r !/y D @!z =@x ¤ 0. Referring to Eq. (8.14),
this represents the situation where neither of the x- and y-components of the
plastic recovery force is zero. Thus, it is interpreted that in this stage, the material
has the recovery mechanism both in parallel and perpendicular to the tensile load.
Stage 3
In this stage, the fringe patterns consist of horizontal straight lines. This pattern
can be interpreted as .r  !/x D @!z =@y ¤ 0 and .r  !/y D @!z =@x D 0,
i.e., the material has lost the recovery force in the direction perpendicular to the
tensile axis but still possesses its component parallel to the tensile load.
Stage 4
In this stage, the horizontal lines observed in Stage 3 begin to rise. As the load
increases, they keep rising until they are mostly vertical in Stage 5. This transition
can be interpreted as the situation where the material are losing the recovery force
parallel to the tensile load as well, .r  !/x D @!z =@y D 0, and the energy
dissipation mechanism represented by Eq. (8.15) becomes dominant.
Stage 5
In this stage, the fringes stop rotating and become mostly vertical, indicating that
the recovery force associated with r  ! is not effective [.r  !/x D @!z =@y D
0 and .r  !/y  D @!z =@x D 0]. The fringe pattern can be interpreted as
representing the @x =@x C @y =@y term in Eq. (8.15).
While Specimens A–D commonly show the transitions from Stage 1 through
Stage 5, the load level at which the fringe pattern changes from one stage to the next
depends on the preload and reload conditions. Figure 8.15a shows typical fringe
patterns observed in the third through fifth reloads in Specimen A–D at the reload
232 8 Applications

level of 10 kgf. Here the rows represent the preload conditions (A–D from the top
to bottom), and the columns represent the reload (third–fifth reload from the left
to right). The fringe patterns seen in the top two rows (in Specimens A and B) are
typical wavy fringes observed in Stage 2. The fringes seen in the two right columns
in the third row are straight fringes observed in Stage 3. The fringes seen in the
bottom row are the rising fringes observed in Stage 4. Notice that in Specimen C
(third row), the horizontal fringes are becoming more straight as the number of
reload increases, indicating that the transition from Stage 2 to Stage 3 takes place as
the reloading repeats. From this viewpoint, the leftmost pattern in the third row is
an intermediate pattern between the “Stage 2 wavy” and “Stage 3 straight” fringes.
Similarly, Fig. 8.15b shows typical fringes observed at the higher reload level of
50 kgf. The row and column arrangements are the same as Fig. 8.15a. The fringe
patterns seen in this figure are basically the rising fringes in Stage 4. Notice that
in all rows the fringes become more upright as going from the left to right. This
indicates that as the reloading repeats, the specimen experiences transition toward
Stage 5.
Generally speaking, Fig. 8.15 indicates that the higher the preloading level or the
number of reloading, the specimen tends to experience the transitions from the initial
to final stages at lower reload levels. Figures 8.16 and 8.17 present this trend more
explicitly, where the former sorts out the data in terms of the preload conditions and
the latter sorts out in terms of the number of reloading. The dashed line indicates the
variation in the transition from Stage 2 to Stage 3, i.e., the load level at which the
material loses the ability to exert the recovery force in the direction perpendicular to
the applied load. The dotted line indicates the variation in the transition from Stage
3 to Stage 4, i.e., the load level at which the material loses the recoverability in the
direction parallel to the tensile axis as well. The following observations are found.
Observation 1
As the reloading is repeated, generally speaking, the transitions from one stage
to the next take place at lower reload levels.
Observation 2
The above trend is more prominent in Specimens A and B than Specimens C and
D. In particular, the reload level that causes the transition from Stage 2 to Stage
3 reduces remarkably from the first to second reloading in Specimen A and B.
Observation 3
In all reloads, Specimen A and B show similar trends and Specimens C and D
show similar trends, respectively. However, trends observed in the two groups
are different from each other.
Observation 4
Specimens C and D barely show Stage 2 fringes, indicating that when this
specimen is preloaded to the level of 106 kgf or higher, the critical fracture
criterion is almost satisfied so that the material barely exerts the recovery force
from the beginning of reloads.
References 233

Observation 5
In the first reload, Specimens A and B show Stage 2 pattern till the end of reload
(80 kgf). In the second reload and after, they show Stage 3 pattern. This strongly
indicates that the Specimen A and B experience fatigue in the first reload, and
that the effect caused by the mechanism of fatigue is similar to the effect caused
by the tensile load.
Notice that whereas the above observations and the features observed in the
fringe patterns enable us to diagnose the preload level, Fig. 8.12 indicates that the
specimens preloaded to different levels show the same reloading characteristics.
This indicates that for the purpose of revealing loading hysteresis, it is necessary to
analyze the spatial distribution of the displacement as seen in the fringe patterns.
The loading characteristic, being representing the total displacement, is unable to
reveal the loading hysteresis.

References

1. Muchiar, S.Y., Muhamad, I., Widiastuti, R., Kusnowo, A.: Optical interferometric technique for
deformation analysis. Opt. Exp. 2, 516–530 (1998)
2. Nakamura, T., Sasaki, T., Yoshida, S.: Analysis of Portevin-Le Chatelier Eect of Al–Mg alloy
by electronic speckle pattern interferometry, advancement of optical methods in experimental
mechanics. In: Conference Proceedings of the Society for Experimental Mechanics Series,
vol. 3, pp. 109–117 (2014)
3. Yoshida, S.: Consideration on fracture of solid-state materials. Phys. Lett. A 270, 320–325
(2000)
4. Yoshida, S.: Phys. Mesomech. 11, 140–146 (2008)
5. Yoshida, S., Rourks, R.L., Mita, T., Ichinose, K.: Physical mesomechanical Criteria of plastic
deformation and fracture. Phys. Mesomech. 12(5–6), 249–253 (2009)
6. Yoshida, S., Gaffney, G.A., Yoshida, K.: Revealing load hysteresis based on physical-
mesomechanical deformation and fracture criteria. Phys. Mesomech. 13(5–6), 337–343 (2010)
7. Sasaki, T., Yoshida, S.: Revealing load hysteresis based on electronic speckle pattern interfer-
ometry and physical mesomechanics. Phys. Mesomech. 15, 47–57 (2012)
Index

A propagation, 94–95
AA6063, low-grade aluminum alloy spatial differentiation, 95–96
charge-like patterns, 213, 214 Conductivity
ductile factor, 215 and dielectricity, 143
Acoustic emission electrodynamics, 105
developed charge-like pattern, energy dissipation, 81
110 free charges, existence, 112
ESPI, 199, 200 Constitutive relation
horizontal error bars, 200 coordinate transformation, 25–26
PLC band, 199 Lamé’s constants, 27
Action Continuum mechanics
definition, 54 displacement, 86
Lagrangian, 54, 75 elasticity, 9
element analysis, 198
equation of motion, 98
C field approach, 5
Cauchy, A.-L., 1 Hellenic period, 1
Charge-like-pattern longitudinal elastic, 99
acoustic emission, 200 Panin’s approach, 4
cyclic loading, 191–197 rotational/deformation waves, 85
drifts, 194 volume expansion, 94
features, 184 Conversion from temporal derivative to spatial
notched specimen, 197–199 derivative, 94, 102–103
and PLC band, 185–191 Covariant derivatives
temperature rise, plastic deformation, deformation structural elements, 127
201–207 differential operation, 49
Coherence displacement vector, 68
description, 150 gauge, 4
population inversion and optical local frames, 52
amplification, 150–151 Lorenz force, 60
Compressive elastic waves partial, 115
continuum mechanics, 97–98 symmetry in physics, 6
differential volume expansion, 94 temporal and spatial, 129
equation of motion, 93 transformation, 56
neighboring part, 130 vector, 47, 48

© Springer Science+Business Media New York 2015 235


S. Yoshida, Deformation and Fracture of Solid-State Materials,
DOI 10.1007/978-1-4939-2098-3
236 Index

Cyclic load serrations, 110


drifts, charge-like patterns, 195 transverse wave characteristics, 104
load ratio, 191 Dielectricity, 143
monotonic loading, 195 Directional cosines
positions, function of time, 195, 196 description, 95
shear-bands/charge-like patterns, 193 propagation vector, 94
sinusoidal pattern, 192 Directional derivative
stress concentration, 196 along propagation of wave, 95
stress–strain characteristics, 192, description, 41–42, 129
193 Dislocation theory, 2, 4, 130
Displacement
displacement gradient tensor, 12, 67, 68,
D 114
Decay characteristics, plastic deformation Kronecker’s delta, 12
wave, 177–181 Lagrangian coordinates, 9
Decay time constant matrix notation, displacement vector,
of displacement waves, 110 11–12
u-Wave at point P2 decays exponentially, position vector, 9, 10
179, 180 Distortion gradient tensor, 63, 64, 223
Deformation charge Drift velocity
drift velocity, 111 deformation charge, 111
energy dissipation, 172 description, 81
field force, 128 developed charge-like pattern, 179
linear elasticity, 127 elastic longitudinal wave, 103
material fractures, 113–114 plastic deformation, 102
plastic regime, 66–67 Ductile factor, 214, 215
self-organized mechanism, 128–130 Dynamic charge-like pattern
Deformation gradient tensor AA6063, 213
description, 11–12 A5052-S, 213, 217
elastic deformation (see Elastic maximum load, 214
deformation) stabilized specimen, 211, 213
physical meaning, terms, 14 welded specimens, 215, 216
Deformation structural elements
clockwise rotation, 126
counterclockwise, 84 E
differential displacement, 114 Egorushkin’s formulation, 62
displacement vector, 69 Elastic deformation
elastic force law, 67 compression wave (see Compressive elastic
gauge, 127 waves)
linear elasticity, 66 and displacement, 9–17
local region, 65 equation of motion and elastic waves,
rotation, 66 32–34
shear strain, 83 Hooke’s law and Poisson’s ratio (see
torque, 128 Hooke’s law)
Developed charge longitudinal effect, 93
discontinuous u-wave, charge-like one-dimensional elastic wave, 98–100
pattern, 176, 177 principal axis, 29–31
displacement field, 109 rotation (deformation) wave, 98
ESPI, 108 Elastic wave
fringe pattern, 108 compressive, 93, 130
one-dimensional, 105, 110, 111 direction, 103
pattern, 104 equation of longitudinal, 33
p-component, 109 one-dimensional, 32–33, 98–100
plastic deformation, 104 shear, 115
Index 237

strain energy, 111 geometrical effect, 69


typical cases, 118–119 kinetic and potential energy density, 74
Electric permittivity, 78, 80, 137–139, 143–145 Lagrangian formalism, 66, 74
Electronic speckle-pattern interferometry local elastic deformation, 69, 70
(ESPI) longitudinal effect, 100
developed charge, 108 Maxwell equations, 77
fatigue effect, 196 metric tensor, 72
fringe patterns, 184 momentum and stress, 63
in-plane displacement, 168–169 motion, 92–93
out-of-plane displacement, 169–170 one-dimensional elastic wave, 98
plastic deformation and fracture, 221–222 oscillatory behavior, 137
two-dimensional in-plane sensitive, phase velocity, 72
181–182 rightmost-hand side, 75
typical dual-beam configuration, 185 shear wave, 74
Entropy solid-state medium, 6
heat flow, material being deformed spatial partial differentiation, 77
elastically, 131 stress tensor component, 73
lattice vibration, 132 time and space component, 77
in plasticity, 131 transformation matrix, 68
Equation of continuity transverse waves, 61
deformation charge, 86 vector potential, 70
description, 82 velocity and rotation, 74
longitudinal effect, 92 Field forces
source of disclination current, 62 charged particle momentum, 121
Equation of motion curvilinear dynamics with linear elastic
longitudinal effect, 92 theory, 124
neighboring unit volumes, 92 deformation charge, 122–123, 127
one-dimensional longitudinal elastic waves, Faraday’s law, 125
32–33 Hamiltonian, 120
three-dimensional compression waves, and Lenz’s law, 91, 126
33–34 potential and kinetic energy, 119
wave velocity, 93 Field stress tensor, 50, 73
ESPI. See Electronic speckle-pattern Field theory
interferometry (ESPI) conservative force, 41
Eulerian coordinates, 10 directional derivative, 42
Euler–Lagrangian equation of motion, 54, 56, elastic, plastic and fracturing stages, 64–65
75 electromagnetics, 37
field equations, 67–78
flat and non-flat surfaces, 39
F gauge theory, 49–53, 58–64
Faraday’s law, 78, 125, 137–141 global and local transformation, 47–49
Field equations gravitational potential energy, 40
deformation-field version, 142 half-pipe-like surface, 41, 42
dislocation dynamics, 62 integral operation, 39
displacement vector, 68 Lagrangian formalism and field equation,
elastic force law invariant, 68 53–57
electric permittivity, 78 linear elasticity, 65
electrodynamics, 4 line symmetry, 37
energy conversion, 130 local frames, 66
Euler–Lagrangian equation, 75 orientation preserving mapping, 66
Faraday’s mechanism, 71, 79 physical law, 37
four-vector notation, 73 plastic deformation charge, 67
gauge field strength, 70, 71 potential gradient, 41, 42
gauge term compensates, 68 quantum mechanics, 66
238 Index

Field theory (cont.) wave function, 57


solid medium, 65 wave-like behavior, 63
Stokes’s theorem, 39 Gauge transformation, 6, 49, 59
structural elements, 65 Gaussian beam
symmetry in physics, 43–47 amplitude, 152
temperature distribution, 38 Gaussian mode, 152
wave solutions, 78–88 Rayleigh length, 153
Flow plasticity theory, 2, 3 Geodesics, 51, 52
Four vector notation Global transformation, 47–49
derivatives and potentials, 117 Griffith, A.A., 2, 63
elastic wave equation, 118 Griffith’s theory, 2
phase velocity, 119
scalar and vector potential, 119
temporal and spatial differentiation, 119, H
128 Hamilton’s equation, field forces, 54, 120
Fracture criterion Hooke’s law
description, 221 coordinate transformation, 24–26, 44
ESPI experiment, 221–222 isotropic materials, 27
Fracture deformation. See also Field theory i th mass’ equilibrium, 17, 18
description, 113 Lamé’s constants, 28
energy flow mechanism, 113 longitudinal effect, 93
phase velocity, 114 moduli, 29
Fracture mechanics, 2, 205 one-dimensional case, 18
phase velocity, one-dimensional
longitudinal wave, 19
G poisson’s ratio (see Poisson ratio)
Gauge field shear modulus, 19
compensation, 50 shear stress and strain, 19
covariant derivative, 52 spring mass system, 17
differentiation, 50 stiffness matrix, 23
external forces, 51 stiffness tensor, 23, 26
geodesics, 50, 51 stress tensor, 18, 21–23, 26
gravitational force, 51 stretch/compression, 29
Lagrangian, 52 vectors, stress, 19
local frame, 50 Young’s modulus, 19
pendulum, 53
stress tensor, 50
transformation, 49 I
vectors derivatives, 49 Index of refraction
Gauge theory description, 142–143
charged particle, 59 and electric susceptibility, 145
dimensional analysis, 62 near resonance, 148–149
dislocation dynamics, 62 real and imaginary parts, 147
Egorushkin’s formulation, 62 Initial stress concentration, 184, 210, 212, 217,
linear elastic theory, 64 218
local symmetry, 63 In-plane displacement
Lorenz/electromagnetic force, 60 “dark fringes”, 169
phase transformation, 58 superposed speckles, 168
phase velocity, 64 Interference and interferometry
Shrödinger equation, 58 circular motion and phase, 155, 156
spatial differentiation, 59 complex notation, 157–161
stress vector, 63 electromagnetic wave, 157
temporal differentiations, 59 fringe visibility, 162
velocities approach, 61 harmonic waves, 155
Index 239

Mach–Zehnder interferometer, 164–165 Local frame


Michelson interferometer, 163–164 connection field, 50
monochromatic harmonic wave, 155 differentiation-based dynamics, 66
spatial periodicity and phase velocity, 156, elastic medium, 67
157 partial differentiation, 50
superposition, light waves, 160–161 plastic deformation, 61
Irwin, G.R., 2 solid medium, 65
vector component, 51
Local symmetry
L actual dynamics, 49
Lagrangian coordinates, 9, 10 coordinate dependence, 63
Lagrangian density, 55, 74, 75, 127 covariant derivatives, 6
Lagrangian formalism electrodynamics, 66
covariant derivative, 57 global symmetry, 38
dynamical system, 54 linear elasticity, 4
Euler–Lagrangian equation, 54 local frame, 49
gradient operator, 56 plastic deformation, 219
Hamilton’s principle, 54 potential field, 127
kinetic energy and potential energy, 53 quantum mechanics, 66
least action principle, 54 transformation, 3
linear elastic theory, 57 vector potential, 99
local frames, 61 Local transformation, 47–49
Newton’s equation of motion, 56 Lüders band
nonlinear deformation, 66 and charge-like pattern, 190
one-dimensional case, 55 at cross-head speed, 188
spatial derivative, 55 locations, 185, 186
stress tensor, 57 naked eye visualization, 185
velocity, 56 observation, 185, 186
Lame’s constants, 28, 97 propagation velocity, 186, 187
Landau, L.D., 1 stress concentration, 189
LASER. See Light amplified stimulated tensile loading, 196
emission of radiation (LASER)
Least action principle. See Hamilton’s
equation, field forces M
Lenz’s law, 79, 83, 91, 125, 126, 137 Mach–Zehnder interferometer, 164–165
Lifshitz, E.M., 1 Magnetic permeability, 78, 80, 137–139, 143
Light amplified stimulated emission of Matrix diagonalization, 29–30
radiation (LASER) Metric tensor, 72
Gaussian beam and propagation, 152–153 Michelson interferometer
optical resonator and phase condition, beam splitter, 163
151–152 optical path difference, 163
population inversion and optical principle, 163
amplification, 150–151 Micro and nano-technology, 3
resonator, 150
Light, electromagnetic wave, 136–142
Load hysteresis N
experimental arrangement, 226 Noether’s theorem, 82
full load and reloads, loading Nonlinear deformation theory, 2
characteristics, 227 Non-welded specimens. See Stabilized and
horizontal line fringe patterns, 231 unstabilized non-welded specimens
preload level and fringe patterns, 230 Notched specimen
reload number and fringe patterns, 229 diameter change, ring charge-like pattern,
transitions, fringe patterns stage, 228 197, 198
vertically parallel fringe patterns, 231 fast tensile loads, 197, 198
240 Index

Notched specimen (cont.) oscillation, 117


finite element modeling, 198 rotational/deformation waves, 85
ring-shaped charge like pattern, 197–198 scalar potential, 72
Von-Mises stress distribution, 199–200 shear-wave’s period, 175
temporal and spatial periodicity, 118
wave travel, 95
O Young’s modulus, 99
Ohmic loss, 82, 112 Photon
One-dimensional elastic wave absorption, 148
continuum mechanics, 99 discrete nature, quantum states, 149
longitudinal elastic wave propagation, 98 quantization of energy, 149
normal stress per unit length/force density, Plastic deformation
99–100 charge flow and displacement wave,
One-dimensional longitudinal elastic waves, 103–108
32–33 coordinate dependence, matrix, 219
Optical interferometry developed charge, 108–111
dielectric media, 143–148 elastic regime charge, 102–103
ESPI (see Electronic speckle-pattern generation of defects, 102
interferometry (ESPI)) linear elastic material, 219
imaging, 153–154 longitudinal restoring force, 100
interference, 161–163 nonlinear displacement, 219
laser and coherence, 150–153 one-dimensional positive charge flows, 101
light as electromagnetic wave, 136–142 schematic illustration, r! ¤ 0, 220
Mach–Zehnder interferometer, 164–165 transition from elastic to, 111–112
Michelson interferometer, 163–164 Plastic deformation wave
photon (see Photon) decay characteristics, 177–181
waves, 155–161 discontinuous u-wave, 176–177
Optical resonator estimated phase velocities, 174–175
laser gain, 151 fringe pattern, subtraction method, 172,
optical field resonance, 151–152 173
output mirror, laser resonator, 151 one-dimensional in-plane sensitive ESPI,
TEM, 152 172
Orientation preserving mapping, 29, 66, 69, 70 oscillatory behavior, 173, 174
Out-of-plane displacement, 169–170 over-damping, u-wave, 176
Over damping, 176, 177 peak position, function of time, 174
shear strain field, tensile loaded aluminum
alloy, 175
P shear wave, 175
Panin, V.E., 3, 4, 63, 104, 131, transverse displacement wave, 173, 174
Phase transformation, 37, 47, 58, 59, 66, 82 PLC band. See Portevin Le-Chatelier (PLC)
Phase velocity band
aluminum alloys, 174 Poisson ratio
charge-like pattern drifts, 102 description, 28–29
crystals, 159 and Hooke’s law (see Hooke’s law)
denominator, 119 Polarization vector, 144
electric permittivity, 80 Population inversion
energy flow, 103 definition, 150
frequency and wavelength, 114 laser active medium, 151
gauge field, 64 spontaneous emission, 150
magnetic field, 142 stimulated emission, 150
mechanical, 86 Portevin Le-Chatelier (PLC) band
one-dimensional longitudinal wave, 19 and charge-like pattern, 185–191
Index 241

serrations, 110, 199 mechanism of formation, 165


stress drop, 199 phase and amplitude, 165
tensile loading, 196 radius, 167
Poynting vector, 81, 88, 103, 141, 142 size, 167
Preload specimens, 226, 228, 230–232 Stabilized and unstabilized non-welded
Propagation vector specimens
description, 94 A5052-N, 212
directional derivative, 95 A5052-S, 212
plasticity, 130, 131 plastic deformation, 212
stress drop/recovery, 212
stress-strain characteristics, 210
Q temporal behavior, charge-like patterns,
Quantum dynamics, 47, 66, 74 210
Stationary charge-like pattern
AA6063 specimens, 213, 214
R A5052 specimen, 218
Reload specimens, 228–233 butt-welding specimen, 217
Rotation graphite coated specimen, 218
deformation structural element, 66, 84 strain localization, 218
deformation tensor, 13 Stokes’s theorem, 39
displacement gradient tensor, 12 Strain tensor
elastic deformation, rotation matrix, 69 components, 22, 23, 25
equation, rotation wave, 34 description, 12
point symmetry, 43 diagonalization, 29–30
(deformation) wave, 85, 98 divisions, 12–13
matrix, 26
shear, 31
S Stress concentration, charge-like patterns
Scalar potential, 72, 114–120, 122 AA6063, low-grade aluminum alloy,
Sensor technology, 3, 196, 199 213–215
Shear band ESPI experiment, 221–222
charge-like pattern, 196 fracture criterion, 221
cyclic loading, 192, 195 fringe patterns interpretation, 222–226
drifts, 194 load hysteresis, 226–233
formation, 181 plastic deformation criterion, 219–220
as function of time, 194, 195 specimen conditions, 209, 210
Lüders band (see Lüders band) stabilized and unstabilized non-welded
propagations, 192–194 specimens, 210–213
temporal fluctuation, 195 welded specimens, 215–218
upper shear-band position, 196 Stress drop
Shear modulus A5052H, 210, 211
density, 92 charge-like pattern formation, 212
dislocation density, 112 PLC band formation, 199
longitudinal effect, 4 zigzag pattern, 110
phase velocity, 114 Stress tensor
shear strain, 19 components, 21, 25, 73
shear wave, 74 description, 20
tensile speed/ strain rate, 175 diagonalization, 30
Shrödinger equation, 37, 47, 58, 59, 66 gauge field variable, 52
Speckles. See also Electronic speckle-pattern vector, 29
interferometry (ESPI) Symmetry in physics
description, 165 coordinate transformation, 43, 44
242 Index

Symmetry in physics (cont.) V


differential operation, 46 Vector potential
displacement vector, 44 counterclockwise case, 71
equilateral triangle, 43 deformation structural elements, 66
gauge transformation, 6 field forces, 120–126
global transformation, 46 gauge field, 50
Hooke’s law, 44 linear elastic theory, 57
shapes, 43 regain local symmetry, 99
vector, 46 scalar, 114–118
velocity, 43 scalar potential from gauge, 114–118
wave dynamics, 118–120
Void density, 217, 218
T Volume expansion
Temperature rise, plastic deformation differential, 94
charge-like pattern, crack-tip, 202, 203 gradient force, 97
crack formation, 205 neighboring part, 130
density and specific heat, brass and Von-Mises stress, 198
aluminum, 207 Vortex-like displacement field
energy dissipative force, 205 acceleration field, 182, 183
ESPI setup, 202, shear force, 183
fringe patterns, brass and aluminum, tensile loading, 183
202–203 two-dimensional in-plane sensitive ESPI,
initial temperature, 205 181, 182
load vs. temperature rise, 206, 207
longitudinal effect, plastic regime, 201
measurement, thermistors, 203, 204 W
temporal variations, 204 Wave plate
Tensile load half-wave plate, 160
Lüders band, 196 quarter-wave plate, 159
monotonic, 195 Wave solutions
notched specimens, 197, 198 characteristics, 79
one-dimensional in-plane sensitive ESPI, continuum mechanics, 85
172 deformation structural element, 84
stress concentration, 106 elastic modulus/density, 86
stress–strain characteristics, 210 electrodynamics, 85
two-dimensional in-plane sensitive ESPI, electromagnetic wave, 81
182 equation of continuity, 82
vortex-like displacement field, 183 external electric field, 87
Thermodynamics Faraday’s mechanism, 79,
decaying wave characteristics and strain free particles, 87
localization, 130 kinetic energy, 81
heat flow, material being deformed Lenz’s mechanism, 78, 83
elastically, 130, 131 longitudinal/transverse, 81
metals entropy, lattice vibration, 132 mass and velocity, 86
micro-defects, plasticity, 131–132 momentum, 86
plastic deformation, 131 Newton’s second law, 83
solids fragmentation, 131 Noether’s theorem, 82
theory, 2 Ohmic loss, 82
Thin lens equation, 154 phase velocity, 80
Three-dimensional compression waves, 33–34 Poynting vector, 88
Timoshenko, S.P., 2 rotational and translational interaction, 83
Index 243

secondary temporal derivation, 80 Welded specimens


shear force, 84 butt-welding (A5052-BW), 217
spring-like effect, 79 charge-like patterns, A5052H, 215–216
translational displacement, 85 deep-welded (A5052-DW), 217
transverse, 84 local plastic deformation, 218
wave equation, 80 shallow-welded (A5052-SW), 217
wave generation mechanism, 85 stabilized A5052 specimens, 218

You might also like